You are on page 1of 646

URANIUM DIOXIDE:

PROPERTIES AND
NUCLEAR APPLICATIONS

Edited by
J. BELLE

Naval Reactors, Division of Reactor Development

United States Atomic Energy Commission

For sale by the Superintendent of Documents, U.S. Government Printing Office


Washington 25, D.C. - Price $2.50,paper cover
Chapter I
URANIUM DIOXIDE AND ITS APPLICATION TO
NUCLEAR POWER REACTORS
I. H. MANDIL AND R. G. Scott

1.1 INTRODUCTION

A reactor fuel material and the fuel element shapes into which it
can be fabricated constitute the fundamental building blocks around
which a nuclear reactor is designed. Thus far the capabilities of the
nuclear physicist and the thermal and hydraulic designers have, in
final analysis, been limited by ability of the metallurgist to provide
suitable reactor materials. One fuel material which in very recent
years has gained prominence as the cornerstone of the power reactor
industry is uranium dioxide.
The purpose of this chapter is to identify the factors which are of
primary importance in the consideration of uranium dioxide as a
nuclear power reactor fuel. No attempt has been made to provide
an exhaustive appraisal of all possible fuel systems and applications
nor to present a comprehensive historical review of UO, as a fuel
material. Rather, specific examples have been chosen to typify salient
features of each major UO, fuel system and to delineate those con
ditions under which fuel systems might be employed to advantage.
To date, the major experience has been with bulk UO, fuel elements
in pressurized water reactors; accordingly, special attention has been
given to that application.
Following a brief history of the use of UO2, the more important
fuel material and fuel element requirements for power reactor ap
plications are reviewed. A discussion follows concerning those charac
teristics of UO2 pertinent to its application in nuclear fuel elements.
A brief analysis is made of the more significant UO, reactor fuel
systems; in a final section, specific fuel element configurations are
discussed.

1.2 HISTORICAL BACKGROUND

Prior to 1940, commercial use of uranium dioxide was virtually


nonexistent. Small quantities were used to provide a rich opalescent
1.
2 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

yellow in fine glassware or to create black, orange, and yellow glazes


for chinaware and porcelain. In addition, limited amounts of this
oxide found application as a catalyst in a few complex organic proc
esses. The small quantities required for these uses were produced by
essentially laboratory techniques. Little was known about the
physical properties of UO2.
Efforts to study the phenomenon of nuclear fission created new
interest in uranium dioxide as a source of uranium metal. Actually,
UO, itself found early application in nuclear research. This initial
use, however, was predicated more upon availability than upon recog
potentialities As early July
its

nition of

as
as
useful fuel. 1941,

a
when research efforts were still directed toward attempting

to
demon
strate experimentally the feasibility controlled nuclear fission, com

of
by
pacted UO2 powder was used the assembly

of
Fermi the first

in
8-foot graphite

an
of
exponential lattice. This assembly, consisting
cubic structure containing natural UO2, was built
of Co

at
tons
7

lumbia University for the determination nuclear constants. Later,

of
pressed, natural UO, and UAOs formed
40

approximately
of

tons

a
major part the fuel complement the first zero power nuclear
of

of of

reactor. Built under the west stands Stagg Field the University

at
Chicago, this reactor initiated the first self-sustaining nuclear chain
of

on

1942, answering decisively the question

of
2,

reaction December the


feasibility
of

controlled nuclear reaction.


a

During the decade that followed, development


of

controlled nuclear
fission created rapidly expanding demand for uranium dioxide.
a

Techniques were devised and facilities were installed for commercial


production tonnage quantities UO, from natural ores for re
of

of

uranium metal. By July 1942, production high purity


of
to

duction
UO, had increased from relatively few pounds per year rate
to
a

ton per day. With the growth nuclear technology for


of

of

over
1

power reactor application, evaluation UO,


of

fuel material was


as
a

begun. For example, beginning mid-1955, bulk UO2, the form


in
in

Zircaloy-2 clad fuel rods containing sintered, natural UO, pellets,


of

was developed for the Shippingport Pressurized Water Reactor. De


this rod-type UO, fuel element was completed by
on

velopment work
late 1956. This application was the first use power
of

bulk UO2
in
a

utilizing enriched UO.


in an

reactor. Stainless-steel clad fuel plates


dispersion stainless steel also found application the first Army
in

Package Power Reactor which began power operation April 1957.


in

power systems are discussed further


in

These and other reactor


Sect. 1.9.
expand
of

Production and dioxide has continued


to

use uranium
until today UO, widely compound
in

the most used uranium the


is

nuclear power reactor field.


URANIUM DIoxIDE APPLICATION To NUCLEAR Power REACTORS 3

1.3 NONNUCLEAR USES OF URANIUM DIOXIDE

Although the objective of this chapter is to discuss the applications


of UO, to power reactors, it is of interest to review briefly the current
nonnuclear applications of this material. From the mid-1920's until
1942, when all existing uranium facilities were diverted to the pro
duction of uranium metal for military uses, the total consumption of
various uranium oxides by industry was only 100 to 200 tons per year.
As previously noted, the major use of these oxides was as coloring
agents in glasses and ceramic glazes. Industry developed satisfactory
substitutes for uranium oxides when these materials were removed
from the market in 1942. Although uranium oxides containing 0.36
percent or less of the U* isotope became available in substantial
quantities during mid-1958, their annual commercial nonnuclear use
amounts to a mere 1 to 2 tons per year. It is estimated that the cost
of this depleted uranium would have to be cut in half to become
competitive with the substitute materials now produced.
To develop alternate applications for the substantial quantities of
depleted uranium now available, the U.S. Bureau of Mines, in co
operation with the U.S. Atomic Energy Commission, has several re
search programs in progress. The more promising of these include
the use of uranium or uranium oxides for the following applications:
(1) catalysis in a variety of chemical processes, (2) production of
high density media for mineral concentration, (3) development of
ferrous and nonferrous alloys for special uses, and (4) replacement
of zinc and magnesium in cathodic protection systems.

1.4 GENERAL CONSIDERATIONS OF FUEL MATERIALS AND


FUEL ELEMENTS

Before examining the specific applications of UO, as a reactor fuel


material it is worthwhile to consider some of the requirements in
herent in the development of fuel elements suitable for power reactor
use. While this discussion is directed primarily at pressurized water
applications, similar considerations arise in the development of a fuel
element for other reactor types.

1.4.1 Types of Fuel

Uranium-base materials can be divided into the two main categories


of highly enriched and high uranium content fuels.

(a) Highly Enriched Fuels


This type of fuel contains on the order of 5 to 95 weight percent
of uranium highly enriched in the U* isotope. The remaining fuel
**4-ca ra -- -
4 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

material is composed of a nonfissile diluent; this material serves


a number of purposes which include improving corrosion resistance,
increasing mechanical strength, and decreasing thermal gradients in
the fuel to facilitate heat transfer.
Highly enriched fuels generally find application in reactor systems
where space limitations make high power densities desirable. Also,
highly enriched fuels have been successfully used in seed and blanket
type reactors. For this reactor concept the highly enriched seed fuel
is normally arranged to form an annulus within a region of natural
uranium fuel called the blanket. During operation the neutrons leak
from the highly enriched seed to cause fission in the natural uranium
blanket. In this type of core about 50 percent or more of the total
power is generated in the natural uranium blanket.

(b) High Uranium. Content Fuels

The major portion of this type of fuel material consists of natural


uranium. In some cases, small quantities of enriched U* (0 to 5
weight percent) are added for reasons of nuclear physics. This latter
type of fuel composition is often referred to as a slightly enriched fuel.
Uranium dioxide for high uranium content applications is generally
used in bulk form, such as pellets in the shape of right circular cylin
ders. A group of pellets is clad with a tube and end caps, for example,
to form the finished fuel element. Highly enriched UO, fuel, on the
other hand, generally takes the form of a dispersion in a metal matrix
or may in bulk form in combination with other metallic oxides
be used
such as ZrO2 or Tho. Highly enriched UO, has also been used as
a dispersant in a ceramic matrix, such as BeO, and in a graphite
matrix.

1.4.2 Fuel Element Performance Requirements

Whether a fuel element is developed for highly enriched or high


uranium content (natural or slightly enriched) application, to be of
significant value it should meet the following performance require
lments:

(a) Dimensional Stability

The fuel material should be capable of withstanding high fuel burn


up without having appreciable swelling or significant deterioration of
its mechanical properties. Failure to meet these requirements could
result in severe limitations of the power capabilities and useful life
of the nuclear reactor.
URANIUM DIOXIDE APPLICATION TO NUCLEAR POWER REACTORS 5

(b) Fission Product Retention

The fuel material should not release an appreciable amount of


uranium or fission products to the coolant system in the event of a fuel
element cladding failure. This requirement is based on the desirability
of permitting prolonged reactor operation after a fuel element becomes
defective, thus, avoiding an early forced shutdown of the plant to
remove the defective fuel element. The fuel material should be re
sistant to corrosion by the - coolant media both before and after exten
sive irradiation.

(c) Thermal Performance

Continued high thermal performance of a fuel element both under


steady state and transient reactor conditions should be maintained
for the useful life of the reactor. Any rapid deterioration of the
thermal capabilities of the fuel element caused, for example, by a
loss in thermal conductivity with prolonged irradiation will clearly
limit the useful life of the fuel element.

(d) Fabricability

A useful fuel element should be readily fabricable in quantity by


production techniques at reasonable cost. It should be recognized,
however, that the least expensive fuel element may not necessarily be
the most desirable. For example, fuel elements in which radial or
axial variation in uranium loading can be readily accomplished, or in
which burnable nuclear poisons can be incorporated, may cost more to
fabricate; however, in most instances the increased fabrication costs
are offset by the greater useful fuel element life.

(e) Inspectability

The fuel element should be of a design that readily lends itself to


adequate nondestructivetesting to establish that it does not contain
harmful defects. At the current stage of experience in reactor
development, the importance of providing high integrity fuel ele
ments cannot be overemphasized.

(f) Chemical Processing

Because it is necessary to reprocess chemically spent fuel elements


to reclaim residual uranium, consideration should be given to incorpo
rating design features which simplify the reprocessing operation.

The relative importance of these characteristics varies, of course,


with each specific power application. The degree to which these
6 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

requirements are met is a direct measure of the usefulness of a power


reactor fuel element.

1.4.3 Fuel Element Design

In recent years UO, fuel elements have been successfully fabricated


in various shapes, such as solid rods, tubular elements, and flat plates,
with lengths varying from approximately 10 to 100 inches. Alumi
num, stainless steel, and Zircaloy have been used to clad UO2 fuel
for operation in a variety of coolants, including high-temperature
pressurized water, liquid metals, and organic compounds. Virtually
the full spectrum of enrichment in the U* isotope from natural to
fully enriched UO, is represented in fuel element assemblies in opera
tion or under development.
It should be noted, however, that while the adequacy of UO2 as a
basic fuel material has been demonstrated, one cannot conclude that
the numerous UO, fuel element designs under consideration have been
proved. Actually, only
a very small number of power reactors with
UO, fuel elements have, as yet, been operated for extended periods
under actual service conditions. In this regard, the Pressurized
Water Reactor (PWR) at Shippingport, Pa., (see Fig. 1.1) thus far
(July 1, 1961)has performed successfully for approximately 13,000
effective full power hours (EFPH) with Zircaloy-2 clad natural UO.
fuel rods for an equivalent maximum fuel depletion of 20,000 MWD/
Ton UO2. This operation, extending over a period of more than 3%
years, has resulted, as of July 1, 1961, in an average burnup in the nat
ural UO, fuel of approximately 4,500 MWD/Ton UO, the highest
natural UO, fuel depletion achieved so far in any operating power
reactor. The Army Package Power Reactor (APPR) at Fort Belvoir,
Va., with fuel elements consisting of highly enriched UO, dispersed
in a stainless steel matrix clad with stainless steel, has operated suc
cessfully for a calculated peak fuel burnup of approximately 18 × 10°
ſissions/cc.
The operating characteristics of UO, fuel elements are strongly
dependent upon a complex interrelationship of numerous variables,
such as geometrical shape, cladding and matrix material compositions,
fabrication techniques, heat flux, central temperature, and other
important reactor design parameters. For this reason a useful fuel
element cannot be satisfactorily developed without specific reference
to a particular reactor application to provide the required balance
among the often conflicting requirements of the metallurgist, the
thermal designer, and the physicist. For example, the metallurgist
and thermal designer may recommend a tubular UO2 fuel element
clad internally and externally with stainless steel in preference to a
rod of similar cross-sectional area because of superior thermal per
URANIUM DIOXIDE APPLICATION TO NUCLEAR POWER REACTORS 7

FMI PIPING

contROL. Rod
DRive MECHANISM

ti-
FM CONDUIT

PRESSURE
contRol. Rod VESSEL HEAD
sHROUD Assembly

-
H.
HOLD-DOwn
BARREL
control. Rod
sHart Assembly

OLTLET NOZZLE

CORE support
SPRING

controL. Rod CORE CAGE


-> SEMBLY

SEED CLUSTER
PREssure
v--SE-
THERMAL SHIELDS

BLAnxet Assembly

INLET NOzzl_E

FIGURE 1.1. Major Components of the PWR Shippingport Reactor Assembly [1].
8 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FIGURE 1.2. Ruptured UO, Fuel Rod [2].

formance. Their analysis may further show that, although the tubular
element is more costly to manufacture, its use could result in a net
reduction in fabrication cost by decreasing the number of required
fuel elements. On the other hand, the physicist would likely demon
strate the over-all superiority of rods for the application on the basis
that the increased fuel loading, required as a result of the greater
amount of stainless steel cladding in the tubular element, more than
offsets its anticipated advantages.
While significantly more information is available today than existed
when serious consideration was first given to UO, as a fuel material,
considerably more information must be accumulated before UO, fuel
element designs can be optimized. In the absence of the detailed
information which can only be obtained from evaluation of numerous
high burnup fuel elements, conservative estimates must be used for
such factors as thermal conductivity, fission gas release rates, and
dimensional stability athigh fuel burnup. Since these factors
markedly affect fuel element thermal performance, there is a great
incentive to be able to predict them accurately in order to improve
power output, increase core life, or reduce cost. Despite this incentive,
indiscriminate reductions in design margins cannot be permitted.
Fuel elements can fail, releasing their contained fission products if
subjected to excessive service requirements (see Fig. 1.2 and Chap. 9).
Because of the complexity of the interrelated physics, thermal, and
metallurgical factors which affect the performance of a power reactor
fuel element, much of the final evaluation work in the development
of a fuel element must be accomplished through long term in-pile
and out-of-pile tests under simulated operating environments. More
over, as the useful lifetime of UO, fuel elements increases, so has the
proof-test time required for their evaluation been extended. Thus,
the development of a reliable long-life fuel element is a long and
expensive undertaking.
While there exists a broad area for additional improvement, the
operating experience obtained thus far indicates that, with proper
URANIUM DIOXIDE APPLICATION TO NUCLEAR POWER REACTORS 9

Selection of design parameters and care in manufacture and inspection,


reliable UO, fuel elements can be fabricated for nuclear power reactor
uSe.

1.5 CHARACTERISTICS OF UO, FOR POWER REACTOR


APPLICATION

A list of some of the more important power reactors in operation


or in various stages of design and construction which use UO, fuel
elements in one form or another is given in Table 1.1. Examination
of the fuel element types described in Table 1.1 shows that three im
portant material configurations have been used or are currently
being evaluated: (1) UO, in bulk form, such as sintered pellets, (2)
UO, dispersed in a metal matrix, and (3) UO, in combination with
other ceramic materials.
The key to the existing widespread interest in UO, as a nuclear
fuel material lies in its attractive physical properties which make it
well suited for reactor applications. Notable among these advan
tageous characteristics are exceptional irradiation stability, high
neutron utilization (low radiation capture cross section), stability in
high-temperature water and other coolant media, and relative ease of
manufacture and chemical reprocessing after irradiation. These de
sirable properties, however, are tempered by low thermal conductivity,
poor thermal shock resistance, and the present lack of a suitable
technique for bonding uranium dioxide directly to metallic cladding.
Moreover, where high uranium atom density is required, UO, provides
only 51 percent as many uranium atoms per unit volume as does
uranium metal (see Appendix A).

1.6 APPLICATION OF BULK URANIUM DIOXIDE

1.6.1 Introduction

Bulk UO2, that is solid UO, in geometrical shapes such as cylin


drical pellets, solid rods, tubes, or plates, is generally used for ap
plications requiring a maximum concentration of UO2 per unit volume
of fuel element. Natural or slightly enriched (1 to 5 percent U*)
UO, in fuel elements of this type.
is employed Conversely, highly
enriched UO, is not used in bulk form because of the high power
densities and excessively large thermal gradients that would result
from such an application.
The thermal conductivity of UO, is considerably lower (one-fifth
to one-tenth) than that of metallic fuels and even lower than most
ceramic materials. Thus, careful control of fuel element thickness
TABLE 1.1—NUCLEAR POWER REACTORS UTILIZING URANIUM DIOXIDE

Maxi- Cladding Basic fuel element


Reactor inunn description
Begin thermal Mod- heat Fuel, enrich

9%
Reactor designation Core type Reactor designer power output Coolant erator flux ment Uns
operation (Mw) Btuſ Thick- Dimensions,
hr- ness, Metal Type inches
ftax103 inches
x
x

||
||
Army Power Package PWR------- Alco Products---- Apr. 1957--| 10----- H10---| H2O.--- 224 Fully enriched 0.005 304SS,------- Plates---------| 0.03 2.70
Reactor (Fort Bel- UO2 disper- 23
voir, Va.) sion in stain
less steel
||

Belgian Thermal PWR_______ Westinghouse----| Dec. 1960-- 40----- H2O---| H2O.-- 430 4%,sintered 0.020 348SS------- Rods---------- 0.34OD x56
Reactor BR-3 bulk UO2
(Mol, Belgium)
x

Carolinas Virginia Pressurized-- Westinghouse----| June 1962-- 60----- D2O---| D2O.--- 230 2%,sintered 0.025 Zr-2--------- Rods---------- 0.50OD
(Parr, S.C.) bulk UOz 104(axially
zoned into
four seg
ments)
x

Dresden (Morris, Ill.)- BW R------- General Electric-- Jan. 1960--| 626----| H2O---| H2O.---
|
2771.5%,sintered 0.030 Zr-2--------- Rods---------- 0.49OD 52
bulk UO2
i

||

Elk River (Elk BWR- Allis-Chalmers--- June 1961--|73----- H2O---| H10--- 310 Fully enriched 0.020 304SS,------- Rods---------- 0.440D x64
River, Minn.) UOz-Tho

-
sintered solid
solutions
x

||

Experimental Gas Gas Cooled...l-------- ----------- Heli-


||Graph--------- 2.5%,sintered 0.020 304SS,------- Rods (fuel 1.10OD
Cooled Reactor unn ite tubular pel- tubular 25%
(Oak Ridge, Tenn.) lets pellets)
x

|
||
||

|
Halden (Halden S.E., Boiling -------------------|------------ 20----- D20---| D20--- 300 Natural sin- 0.070 Aluminum Rods----------| 1.00OD
Norway) heavy tered UO2 alloy 94
water
x

Humbolt Bay PWR----___ General Electric-- June 1962.--|193----| H2O---| H2O--- 340 2.1%,sintered 0.030 Zr-2---------| Rods---------- 0.56OD
(Eureka, Calif.) bulk UO: 78

-- -- -- - -
()
x
in
Rods. 0.32

&
Indian Point (Bu Pwn. habcock Wil Sept. 1961. hizo hao 450 Fully enriched 0. trºo ss
cox UOz-94% 16.5
chanan, N.Y.)
Thor sin
tored solid
solutions
x

Kahl (Kahl, W. bw R. General Electric Feb. 1961. H10... H10... 270 2.6%,sintered 0.034 Zr-4------- Rods (Axi 0.56 OD
Germany) bulk UO: ally zoned
into two
compart
ments
x

(44,000 H2O... H2O.-- 5% sintered Zirconium Rods---------- 0.32OID 56


SHP2) bulk UO: alloy
x

&
Babcock Wil H2O... H2O.--- 4% sintered 304SS-------- Rods---------- 0.32OD 96
cox bulk UO:
x

Nuclear Power Dem Canadian Gen D-O--- D20--- Natural sin Zr-2--------- Rods (rods 1.0OD 19.5
2

onstration Reactor eral Electric tered bulk spaced by


Co. UO: Zircaloy
fins)
x

.
Pathfinder (Sioux BWR with Allis-Chalmers. June 1962.-- H2O H2O.-- Boiler-1.7% UOz 0.030 0.45OD 78

S.
Falls, Dak.) nuclear
superheat
x

Superheater 0.006 1.0 OID 78


fully enriched
UO2 disper
sion in stain
less steel
x

0.47OD

1
Shippingport Core Westinghouse---- Dec. 1957. 231---- H2O.-- H2O.--- 375 Blanket-sintered 0.030
(Shippingport, Pa.) (Blan natural UOz 10%
ket)
x
x

--

2
Shippingport Core Westinghouse---- Reactor H10 Seed Sintered ZrO2 0.020 Plates--------- 0.076 3.5
(Shippingport, Pa.) corn UOn 93% 96
plete
April
1962
x
x

Blan Sintered natural 0.020 Zr-4--------- Plates.-------- 0.140 3.5


ket UO: 96
390
-:
TABLE 1.1—NUCLEAR POWER REACTORS UTILIZING URANIUM DIOXIDE—Continued

Maxi- Cladding Basic fuel element


Reactor milulin description
Begin thermal Mod- heat Fuel, enrich

9%
Reactor designation Core type Reactor designer power output Coolant erator flux ment U235
operation (Mw) Btu/ Thick- Dimensions,
hr- ness Metal Type inches
ft2x103 inches
x

|
---
--
--

|--
Societá Elettronu- BWR______. General Electric. Nov. 1962- 507____ H2O H2O---| 285- 1.8%sintered 0.020 SS. Rods---------- 0.49OD
cleare Nazionale- bulk UO2 104
Senn (Punza Fiame,
Italy)
x
x

-||
||
--

|
Vallecitos (Plesanton, BWR------- General Electric--| Nov. 1957 30-----| H2O---| H2O---|300. ---| Fully enriched 0.005 SS- ------| Plate----------| 0.025 2.9 -
Calif.) UO2 disper- 37
sion in stain
less Steel
|-|

..
-
-
x

Vordnezh (Vordnezh PWR ------|------------------- 1961 760----| H2O---| H2O---| 440----| 1.5% sintered 0.024 Zirconium Rods----------| 0.40OD
Region, USSR) UO2 102
||
x

_
__

. -||

___
___
Yankee (Rowe, PWR Westinghouse.---| Dec. 1960 392----| H2O---| H2O---| 450----| 3.4% sintered 0.015 348SS____ Rods---------- 0.34OD 90
Mass.) UO2
URANIUM DIOxIDE APPLICATION TO NUCLEAR POWER REACTORS 13

is required to prevent excessive temperature and possible melting.


The relatively high melting point of UO, (2,760° C) compensates
only partially for this low thermal conductivity.
Bulk UO, is clad to prevent fission product contamination of the
reactor coolant media as well as to retain the fuel particles formed
by the thermal fracture of the UO, during operation. Aluminum,
stainless steel, and Zircaloy have been successfully employed as clad
ding materials for UO2. In addition, bulk UO, has been coated with
ceramic films. Such coated UO, fuel materials, however, have not
found application in power reactors primarily because in operation
this type of fuel element exhibits thermal fracture due to the poor
thermal shock resistance of UO2.
Bulk UO, fuel has been successfully fabricated by a number of
different manufacturing processes including the following: (1) hot
pressing or cold compaction followed by sintering at an elevated
temperature, (2) vibratory compaction of loose powders followed by
swaging of the clad element, (3) slip casting and sintering, and (4)
ceramic extrusion and sintering.

1.6.2 Reactor Performance

The resistance of bulk UO, to irradiation damage is exceptional.


Since this topic is discussed in detail in Chap. 9, only brief reference
to representative results will be made here.

(a) Dimensional Stability

Shippingport Reactor fuel development work has shown that


Zircaloy-2 clad high density (approximately 96 percent of theoreti
cal) bulk UO, in the form of 0.040-inch-thick, 14-inch-wide UO,
platelets exhibits essentially no dimensional change after a burnup
of 25,000 MWD/Ton of UO, with a central temperature and heat
flux of approximately 1,200° F (650° C) and 600,000 Btu/hr-ft",
respectively. Continued irradiation under these conditions yields an
approximate volume increase of 5 percent in 0.040-inch-thick fuel
after a burnup of 60,000 MWD/Ton of UO,. This latter exposure
represents destruction of approximately one out of every 12 uranium
atoms.

(b) Fission Gas Retention

Irradiation tests indicate that the release of fission gas from high
density UO, with fuel burnup as large as 25,000 MWD/Ton of UO,
is on the order of only 0.5 to 1 percent of the total fission gas produced.
Test results also indicate that an increase in fuel burnup causes
greater gas release. For example, at 60,000 MWD/Ton of UO, burnup
URANIUM DIOxIDE APPLICATION TO NUCLEAR POWER REACTORS 17

It is apparent that fuel element design can be modified to alter the


effects of the factors listed above, but such modifications generally
create side effects which must also be evaluated. As an overly simpli
fied example, consider the following: Fuel rod length can be increased
to lessen the effect of end cap flux peaking and coolant pressure drop
at the entrance and exit of each fuel element bundle. Such a modi
fication might likewise reduce fabrication costs. These advantages,
however, must be balanced against the increased tendency of a long
rod thermally to distort, thereby adversely affecting coolant flow to
produce hot spots. The use of long rods can also result in fuel ratchet
ing or contact between rods because of excessive distortion. These
effects are all potential causes of cladding failure.

(b) Differential Thermal Earpansion

When a UO,
fuel element is incorporated into a reactor design it
becomes restrained by other structures required for such purposes as
to direct coolant flow, to provide structural support, to accommodate
control element motion, and to allow for core instrumentation. As
a result of different thermal conditions and material compositions,
the fuel element is subjected to stresses in addition to those developed
by its role as a fuel element alone.

(c) Reactor Transients

The poor thermal conductivity of UO, not only affects steady state
conditions but also the transient behavior of a reactor. In fact, in
many instances the poor thermal conductivity of UO, can cause the
transient behavior of the core to become an overriding consideration
in the design of the reactor. For example, under a loss of coolant flow
accident, the heat capacity and poor thermal conductivity of UO,
slows down the rate of heat release from the fuel elements to such
an extent that it no longer matches the decay rate of the coolant flow.

1.7. APPLICATION OF UO, DISPERSIONS IN A METAL


MATRIX

1.7. I Introduction

The concept of dispersing UO, particles throughout a metal struc


ture has long been recognized as a practical means to improve the
utilization of UO, as a fuel material. The objective of this technique
is to localize the irradiation damage within small controllable zones
throughout a structural body or matrix that is itself relatively un
damaged by fission fragments or by some other irradiation effect. In
18 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

this manner the favorable irradiation resistance of bulk UO2 can


pro

its
be exploited while compensation for undesirable properties

is
vided through the improved thermal conductivity, shock resistance,
and mechanical strength As will

be
of
the matrix material. indicated
below, the specific design UO2 dispersion fuel compromise

of

is
a
a
of

as
numerous factors which includes physics and thermal well

as
metallurgical considerations. With regard the current status

to

of
UO, dispersion fuels should recognized that, although
on

be
work

it
considerable effort has been applied the in-pile and out-of-pile

to
UO, dispersions metallic matrices, the potential
of

of
evaluation

in
this important fuel element concept has just begun probed. This

be
to
situation primarily results from two causes: the large number

of
independent variables available for evaluation and the extended time
(approximately
to

years) required obtain performance

to
data
2
1

high burnup. So far, the limited oper


on

specific fuel dispersions to


ating experience with UO, dispersion fuels that have been irradiated
high fuel burnup has been encouraging. For example, the fuel
to

Army Package Reactor, consisting

of
elements for the first Power

a
percent dispersion highly UO, type 302B
20

of

volume enriched

in
satisfactorily

an
steel, had, June operated
of

to
as

stainless 1960, aver.


age fuel burnup approximately

an
of

10° fissions/cc and estimated


×
8

peak burnup
of

10° fissions/cc, with


18

central fuel element tem


×

perature approximately 580° Considerably more work, how


of

F.

ever, required with the UO, dispersion fuel concept evaluate more to
is

its

thoroughly performance and permit reduced fabrication costs


to
be
its

its
before limitations can adequately understood and capabilities
fully exploited. -

The paragraphs that follow discuss the parameters which signifi


cantly affect the performance UO, dispersions, summarize the ad.
of

vantages and shortcomings this fuel concept, and review repre


of

sentative irradiation results and fabrication processes for specific UO.


dispersion fuel materials.

1.7.2 Significant Parameters for UO Dispersions

brief analysis those factors which most significantly affect the


of
A

performance UO, dispersions metal matrix will provide


of

in

bet.
a
a

ter understanding the merits and potential limitations


of

of

this fuel
concept. Three such matrix material, the
of

factors are the choice


a

arrangement dispersed UO,


of

microstructural the within the matrix.


and the density and composition the dispersed fuel.
of

(a) Matriæ Material

based upon physical and


of

Selection suitable matrix material


is
a

properties, such high-temperature strength, ductility,


as

chemical
URANIUM DIOXIDE APPLICATION TO NUCLEAR POWER REACTORS 19

thermal conductivity, thermal expansivity, and corrosion resistance,


that are necessary for a specific application. Other considerations in
clude the rateof reaction of interdiffusion between the UO2 and the
matrix material at the temperatures encountered during fabrication
and reactor operation. Neutron poison effects, fabricability, and to
tal

cost are likewise prime considerations

of
matrix

in
the selection

a
material for specific reactor applications. At present, such materials
austenitic and ferritic stainless steels, aluminum, and iron have
as

found useful application matrix materials for UO, dispersions.


as

(b) Microstructural Arrangement

dispersion fuel depends large measure upon the


of

The success

in
a

continuous, undamaged metal matrix surrounding small


of

presence
a

local islands of fuel even after extensive irradiation. Since fission


fragments cause structural deterioration the matrix within zone
of

a
by

the recoil range fission products from the fuel (on the
of

defined
microns), sufficient separation must established be
10

be
of

order
tween the dispersed fuel particles preclude overlapping the dam

of
to

matrix regions. Otherwise, the required network undamaged


of
aged
will For this reason, the UO, particle
be

matrix not maintained.


UO, within the fuel
as

size and shape,


of

the volume fraction


as

well
body, greatly influence the performance dispersion fuel.
of
a

UO, particle size


on

the performance dispersion


of

in in a of

The effect
a
by

illustrated qualitatively fuel system


be

of

fuel can consideration


containing the particle
of

fixed volume fraction


an A

UO2. reduction
a

the dispersed fuel will result


of

of

increase the area


in

size
UO, adjacent the metal matrix, since the ratio
of
to

the total amount


fuel and matrix present remains constant. Hence,
of

in

decrease
a

fission fragment
an

of

particle size causes increase the total amount


in

the matrix. Should the particle size become too small,


to

damage
irradiation damage
to
be

the metal matrix can


to

concentrated the
point that structural integrity lost. Similarly, there exists upper
an
is

UO. Should individual UO, particles become too large, fuel


to

limit
fabrication adversely affected that nonhomogeneous dispersion
in
is

obtained and the dispersed fuel tends exhibit the undesirable


to
is

properties bulk UO, fuel.


of

The shape the dispersed UO, particles importance.


of

of

also
is

Sharp fuel corners can cause local regions high stress concentration
of

decreased integrity
of

the metal matrix.


in

which can result


(c)

Density and Composition

UO,
of be
of

The significance density within the metal matrix can


by

recognized the fission gas release properties


of

consideration
UO. mentioned previously.
57.4789 0–61—3
20 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

If dimensional stability of the fuel element is to be maintained, the


relatively small quantities of fission gas released from the UO, fuel
during irradiation must be retained without local distortion of the
matrix. Since for a given volume of released fission gas the total gas
pressure available to cause matrix deformation decreases as the volume
present for gas accumulation increases, it is logical to consider means
of providing additional gas expansion volume. In evaluating the best
method for retaining fission gases, however, two competing phenomena
must be considered. On the one hand, since present fabrication tech
niques result in essentially intimate contact between the fuel and
metal matrix, additional gas expansion volume can be most easily
provided through the use of lower density UO2 dispersions. Decreas
ing the density of UO2, however, has the inherent disadvantage of
resulting in a larger fractional release from the UO, lattice of the total
fission gas produced. Hence, selection of the proper UO2 density for
a specific application can be critical and for the present, at least, must
be made largely through experimentation, since it does not lend itself
to accurate calculation.

1.7.3 Fabrication of UO, Dispersions

The fabrication of UO, dispersions in a metallic matrix requires


the use of powder metallurgy techniques. Properly prepared metal
powders are blended with UO, particles, compacted and sintered or,
in some cases, extruded to form dispersion fuel fillers. These fuel
fillers are subsequently clad by a suitable metallurgical bonding
process, such as roll bonding or inert gas pressure bonding.
The cost of fabricating UO, dispersion fuel elements is generally
greater than that required for the manufacture of bulk UO, fuel ele
ments, primarily because powder metallurgy methods must be used.
Fabrication costs are further increased in those instances in which
matrix materials which react with the UO, are used. Aluminum, for
example, requires special fabrication techniques to minimize inter
action with UO2.
If proper care is not exercised in the fabrication of a dispersion fuel
filler, the UO, particles will break up during manufacture. Such frac
ture of the dispersed UO, can result in sharp corners with their ac
companying stress concentration factors or cause stringering of the
fuel, thereby creating corridors of weakness between fuel particles.
Stringered fuel not only can destroy the strength of a dispersion
matrix but can also provide interconnected avenues for corrosion at
tack in the event of a cladding failure.
The physical properties of UO, dispersion fuel compacts are
strongly dependent upon the fabrication processes used in their manu
facture. For example, U.O. particle size range, distribution, density,
URANIUM DIOXIDE APPLICATION TO NUCLEAR POWER REACTORS 21

matrix strength, and ductility are among the properties affected by


the specific fabrication techniques employed.

1.7.4 Reactor Performance

(a) Fission Gas Retention and Dimensional Stability

Although the dimensional stability and fission gas retention char


acteristics of UO, are quite good, dispersion fuels offer promise of
extending the potential fuel depletion limits of bulk UO, by restrain
ing the highly depleted UO, in a strong, ductile, metal matrix. The
fact that such restraint is potentially possible is evidenced by actual
irradiation results obtained on Zircaloy-clad bulk UO, fuel plates
which showed greater fuel swelling after large fuel burnup with
13-inch-wide fuel compartments than with 14-inch-wide compart
ments. Calculations indicate that the cladding provided a greater
resistance to fuel motion in the case of the 14-inch compartment.
This topic is discussed in Chap. 9.

(b) Corrosion. Resistance

As previously noted, UO, exposed to high temperature water after


depletion of approximately 60,000 MWD/Ton of UO, (18× 10”
a fuel

fissions/cc) exhibits marked swelling (about 20 volume percent).


Similar swelling can be avoided through the use of a properly designed
dispersion fuel element. Since a corrosion resistant metal matrix com
pletely surrounds each fuel particle, permitting only limited contact
between fuel and coolant in the event of a cladding failure, the use
of metallurgically bonded cladding serves to minimize further the
effects of any loss in corrosion resistance experienced by the fuel phase
after extensive irradiation by restricting the volume of fuel that can
be exposed to coolant through a cladding defect.

Metallurgically Bonded Cladding


(c)

enhancing thermal conductivity and minimizing the


to
In

addition
fuel under clad defect conditions, the integral bond attain
of

corrosion
able between dispersion fuel component and cladding eliminates the
by

waterlogging As
or

possibility ratchetting.
of

fuel-element failure
previously, UO, fuel elements are susceptible
to

discussed clad bulk


such failure mechanisms because of the lack of such bond between
a

cladding and fuel.

(d) Thermal Conductivity

The thermal conductivity U(), dispersion dependent upon


of

is
a

UO,
of

the relative thermal conductivities the and matrix materials


18 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

this manner the favorable irradiation resistance of bulk UO2 can


pro

its
be exploited while compensation for undesirable properties

is
vided through the improved thermal conductivity, shock resistance,
and mechanical strength As will

be
of
the matrix material. indicated
below, the specific design UO2 dispersion fuel compromise

of

is
a
a
of

as
numerous factors which includes physics and thermal well

as
metallurgical considerations. With regard the current status

to

of
UO, dispersion fuels should recognized that, although
on

be
work

it
considerable effort has been applied the in-pile and out-of-pile

to
UO, dispersions metallic matrices, the potential
of

of
evaluation

in
this important fuel element concept has just begun probed. This

be
to
situation primarily results from two causes: the large number

of
independent variables available for evaluation and the extended time
(approximately
to

years) required obtain performance

to
data
2
1

high burnup. So far, the limited oper


on

specific fuel dispersions to


ating experience with UO, dispersion fuels that have been irradiated
high fuel burnup has been encouraging. For example, the fuel
to

Army Package Reactor, consisting

of
elements for the first Power

a
percent dispersion highly UO, type 302B
20

of

volume enriched

in
satisfactorily

an
steel, had, June operated
of

to
as

stainless 1960, aver.


age fuel burnup approximately

an
of

10° fissions/cc and estimated


×
8

peak burnup
of

10° fissions/cc, with


18

central fuel element tem


×

perature approximately 580° Considerably more work, how


of

F.

ever, required with the UO, dispersion fuel concept evaluate more to
is

its

thoroughly performance and permit reduced fabrication costs


to
be
its

its
before limitations can adequately understood and capabilities
fully exploited. -

The paragraphs that follow discuss the parameters which signifi


cantly affect the performance UO, dispersions, summarize the ad.
of

vantages and shortcomings this fuel concept, and review repre


of

sentative irradiation results and fabrication processes for specific UO.


dispersion fuel materials.

1.7.2 Significant Parameters for UO Dispersions

brief analysis those factors which most significantly affect the


of
A

performance UO, dispersions metal matrix will provide


of

in

bet.
a
a

ter understanding the merits and potential limitations


of

of

this fuel
concept. Three such matrix material, the
of

factors are the choice


a

arrangement dispersed UO,


of

microstructural the within the matrix.


and the density and composition the dispersed fuel.
of

(a) Matriæ Material

based upon physical and


of

Selection suitable matrix material


is
a

properties, such high-temperature strength, ductility,


as

chemical
20 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

If dimensional stability of the fuel element is to be maintained, the


relatively small quantities of fission gas released from the UO, fuel
during irradiation must be retained without local distortion of the
matrix. Since for a given volume of released fission gas the total gas
pressure available to cause matrix deformation decreases as the volume
present for gas accumulation increases, it is logical to consider means
of providing additional gas expansion volume. In evaluating the best
method for retaining fission gases, however, two competing phenomena
must be considered. On the one hand, since present fabrication tech
niques result in essentially intimate contact between the fuel and
metal matrix, additional gas expansion volume can be most easily
provided through the use of lower density UO2 dispersions. Decreas
ing the density of UO2, however, has the inherent disadvantage of
resulting in a larger fractional release from the UO, lattice of the total
fission gas produced. Hence, selection of the proper UO2 density for
a specific application can be critical and for the present, at least, must
be made largely through experimentation, since it does not lend itself
to accurate calculation.

1.7.3 Fabrication of UO, Dispersions

The fabrication of UO, dispersions in a metallic matrix requires


the use of powder metallurgy techniques. Properly prepared metal
powders are blended with UO, particles, compacted and sintered or,
in some cases, extruded to form dispersion fuel fillers. These fuel
fillers are subsequently clad by a suitable metallurgical bonding
process, such as roll bonding or inert gas pressure bonding.
The cost of fabricating UO, dispersion fuel elements is generally
greater than that required for the manufacture of bulk UO, fuel ele
ments, primarily because powder metallurgy methods must be used.
Fabrication costs are further increased in those instances in which
matrix materials which react with the UO, are used. Aluminum, for
example, requires special fabrication techniques to minimize inter
action with UO2.
If proper care is not exercised in the fabrication of a dispersion fuel
filler, the UO, particles will break up during manufacture. Such frac
ture of the dispersed UO, can result in sharp corners with their ac
companying stress concentration factors or cause stringering of the
fuel, thereby creating corridors of weakness between fuel particles.
Stringered fuel not only can destroy the strength of a dispersion
matrix but can also provide interconnected avenues for corrosion at
tack in the event of a cladding failure.
The physical properties of UO, dispersion fuel compacts are
strongly dependent upon the fabrication processes used in their manu
facture. For example, U.O. particle size range, distribution, density,
URANIUM DIOXIDE APPLICATION TO NUCLEAR POWER REACTORS 21

matrix strength, and ductility are among the properties affected by


the specific fabrication techniques employed.

1.7.4 Reactor Performance

(a) Fission Gas Retention and Dimensional Stability

Although the dimensional stability and fission gas retention char


acteristics of UO, are quite good, dispersion fuels offer promise of
extending the potential fuel depletion limits of bulk UO, by restrain
ing the highly depleted UO, in a strong, ductile, metal matrix. The
fact that such restraint is potentially possible is evidenced by actual
irradiation results obtained on Zircaloy-clad bulk UO, fuel plates
which showed greater fuel swelling after large fuel burnup with
13-inch-wide fuel compartments than with 14-inch-wide compart
ments. Calculations indicate that the cladding provided a greater
resistance to fuel motion in the case of the 14-inch compartment.
This topic is discussed in Chap. 9.

(b) Corrosion. Resistance

As previously noted, UO, exposed to high temperature water after


depletion of approximately 60,000 MWD/Ton of UO, (18× 10”
a fuel

fissions/cc) exhibits marked swelling (about 20 volume percent).


Similar swelling can be avoided through the use of a properly designed
dispersion fuel element. Since a corrosion resistant metal matrix com
pletely surrounds each fuel particle, permitting only limited contact
between fuel and coolant in the event of a cladding failure, the use
of metallurgically bonded cladding serves to minimize further the
effects of any loss in corrosion resistance experienced by the fuel phase
after extensive irradiation by restricting the volume of fuel that can
be exposed to coolant through a cladding defect.

Metallurgically Bonded Cladding


(c)

enhancing thermal conductivity and minimizing the


to
In

addition
fuel under clad defect conditions, the integral bond attain
of

corrosion
able between dispersion fuel component and cladding eliminates the
by

waterlogging As
or

possibility ratchetting.
of

fuel-element failure
previously, UO, fuel elements are susceptible
to

discussed clad bulk


such failure mechanisms because of the lack of such bond between
a

cladding and fuel.

(d) Thermal Conductivity

The thermal conductivity U(), dispersion dependent upon


of

is
a

UO,
of

the relative thermal conductivities the and matrix materials


22 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS |

as well as upon the volume fraction of UO, present and the density
of the fabricated fuel body. Since the thermal conductivity of the
common metal matrix material is considerably better than that of
UO2, the combined thermal conductivity of a UO2 dispersion is sub
stantially better than that of bulk UO. In addition, since UO2 dis
persions can be metallurgically bonded to fuel element cladding,
further improvement in thermal conductivity can be made. These
two effects, the metallurgically bonded cladding and the reduced
thermal resistance of the dispersion fuel filler, can account for as much
as a factor of eight- to tenfold increase in effective thermal conduc
tivity over that attainable in unbonded bulk UO, fuel elements.

1.7.5 Limitations

merits of UO, dispersion fuel materials it should


In considering the
be recognized that this fuel system does possess inherent limitations.
First, UO, dispersions are not suitable for applications requiring high
uranium content fuels. Thus, reactor designs predicated on the use
of slightly enriched or natural uranium dioxide fuels do not lend them
selves to the use of dispersion type fuel elements. Second, for the
present, at least, the fabrication techniques required for dispersion fuel
manufacture are generally more complex and expensive than those
employed in producing other types of UO, fuel elements. Finally,
the dispersed local regions of high fuel content cause high concen
trations of fission damage and localized zones of large internal stresses.
Should the combined effects of localized stresses and fission fragment
damage to the matrix cause cracking or other structural deterioration
of the matrix, the objective of the dispersion fuel element would be
defeated.

1.8 APPLICATION OF URANIUM DIOXIDE IN COMBINATION


WITH OTHER CERAMIC MATERIALS

As previously noted, ceramic additives are used with enriched


uranium dioxide to produce fuel material for high temperature and
high burnup applications. The primary purpose of such filler
materials is to provide a diluent to improve fission gas retention, to
reduce thermal gradients, and in some instances to increase mechanical
strength, thermal conductivity, or thermal shock resistance. PuC2 and
ThC), have also been added to enriched UO, for the utilization of
plutonium as a fuel and thorium as a fertile material. All these
materials should have low nuclear absorption cross sections to avoid
excessive parasitic capture of neutrons.
When UO, is fabricated with ceramic additives two types of fuel
systems result: solid solutions and dispersions.
URANIUM DIOXIDE APPLICATION TO NUCLEAR POWER REACTORS 23

1.8.1 Solid Solution

ZrO2—UO, Th(), UO, CeO2–UO, are typical of this type of fuel


system. The system ZrO2–25 to 30 weight percent UO, is under con
sideration as a fuel material for the seed portion of the second Ship
pingport reactor core. Continuing irradiation test results on Zircaloy
clad, 0.036-inch-thick plate samples of this material operating at heat
fluxes on the order of 800,000 Btu/hr-ft 4 have shown less than 0.5
percent increase in thickness after 25X10° fissions/cc.
A special type of solid solution
system is one with more than one
phase. The ternary ZrO2–CaO-UO2, for some compositions, is repre
sentative of this type of fuel material. This fuel appears to have
better irradiation resistance than the Al2O3–UO, and BeO-UO, ma
terials, but is not as good as UO2. For example, tests performed on
samples of 0.040-inch-thick ZrO2–13 weight percent CaO-17 weight
percent UO, indicate good dimensional stability to burnups of
12× 10° fissions/cc where swelling approximately twice that of UO,
is noted.

1.8.2 Dispersed UO, Phase in an Oxide Matrix

BeO-UO, and Al2O3–UO2 are fuel compositions representative of


this system. This type of fuel exhibits some of the properties of UO,
dispersions in a metallic matrix. With regard to irradiation stability,
several samples of Al2O3–21 weight percent UO2 show a saturated
swelling of approximately 20 volume percent after less than 0.1 per
cent burnup of the total atoms present. While BeO-UO2 is somewhat
more resistant to irradiation effects than Al2O3–UO2 it too shows
substantial swelling. Test results indicate that the size of UO2 par
ticles significantly affects the magnitude of fuel swelling in a BeO
matrix. For example, a fuel plate containing UO2 particles, am
average 10 microns in diameter, increased more than 25 percent in
thickness after 12×10%" fissions/cc burnup. Under similar conditions
fuel with 150-micron UO, particles showed approximately one-third
less swelling.

1.9 SPECIFIC UO, FUEL ELEMENT CONFIGURATIONS

Thus far, the discussion has been primarily directed toward the
general properties of UO, as a fuel material and the broad require
ments of a useful reactor fuel element. To illustrate more clearly
the nature of UO, power reactor fuel elements, five examples have
been chosen from the group of reactors listed in Table 1.1. These fuel
elements are representative of five basically different power reactor
24 URANIUM DIOxIDE:

-
-ſ
J!---
PROPERTIES AND NUCLEAR APPLICATIONS

-
END CAP

u02 FUEL
PELLET

26. PELLETS
|
se
CLADDING

Figure 1.3. PWR Core 1 Blanket Element [1].


URANIUM DIOXIDE APPLICATION TO NUCLEAR POWER REACTORS 25

CORNER CHAMFERED
AND ONE ROD OMITTED
TO GIVE CLEARANCE
FOR SAMPLE TUBE

TuSE SHEET

FUFL Roos (I2O)

—-º -

FIGURE 1.4. PWR Core 1 Bundle [1].

systems. Detailed descriptions of the reactors listed below as well


1.1

others enumerated in Table are provided Refs,


to
in

6.

as the
3

1.9.1 The Shippingport Reactor—PWR


(a) Core
1

The rod-type fuel element used the blanket region


of

Core for
in

1
the

of

Shippingport reactor was the forerunner the type that,


at

Present, the most widely used UO, fuel geometry for power reactors.
is

1.3

Figs.
As

illustrated and 1.4, the basic fuel element consists


in

high density (93 percent theoretical) natural UO, fuel


95
to
26
of
26 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Ş
-
witw.
A

COOuant
flowPassage
SAMPLING
RAxt

RAKE
unit

TopBundue
spacer
(INTEGRAL
with
º SAMPuno
RaxE)

-shell Assembly

viewa
witwa
\lº||

FIGURE 1.5. PWR Core Blanket Assembly [1].


1

pellets, 0.357 inch OD and 0.350 inch long, assembled into Zir.
a

caloy-2 tube approximately inches long, 0.420 inch OD, with


10

0.028-inch nominal wall thickness. The tube sealed with Zircaloy-2


is
by

end caps automatic Heliarc welding. fuel bundle consists


of
A

120 fuel rods welded into suitably drilled Zircaloy-2 end plates which
space and support the fuel elements square lattice. Seven such
in
a

bundles are placed Zircaloy tube shell complete the final blanket
to
in
a

fuel assembly.
As shown Fig. 1.5, the fuel assembly fitted with sampling
in

is

detect and locate during reactor op


to

tube connection that used


is

eration any fuel assembly that should develop fuel element cladding
a

failure, thereby releasing fission products Figure


1.5
to

the coolant.
URANIUM DIOXIDE APPLICATION TO NUCLEAR POWER REACTORs 27

UPPERcover PLATE

Lower cover PLATE

FUEL RECEPTAcLE PLATE

FIGURE 1.6. Schematic Representation of Cladding and Fuel Components for the
PWR Core 2 Blanket Fuel Element.

also depicts a removable flow orifice that can be replaced to accom


modate reactor power shifts due to plutonium buildup in the natural
uranium fuel elements.
A discussion of the numerous problems encountered in the course
of developing the Shippingport blanket fuel element and the means
used to obtain their solutions are given in Ref. 1.

(b) Core 2

The blanket fuel elements for Shippingport Core 2 will be Zircaloy


clad UO, fuel plates. Similarly, it is expected that the seed fuel
elements will be Zircaloy-clad ZrO2–UO, fuel wafers in plate
geometry. It is anticipated that these plate-type ceramic fuel ele
ments will yield high thermal performance for an extended fuel
burnup.
The blanket fuel elements will contain 0.100-inch-thick, 0.250-inch
wide, 6-inch-long fuel compartments in a grid array as shown in
Fig. 1.6. The fuel compartments are surrounded by 0.030-inch-wide
ribs and clad with 0.020-inch-thick Zircaloy. Finished fuel elements
will be approximately 41% inches wide and 8 feet long.
Seed fuel elements will consist of 0.036-inch-thick, 0.250-inch-wide,
28 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

6-inch-long fuel compartments arranged in a manner similar to that


of the blanket fuel element. Since the compartment fuel plate lends
itself readily to selective fuel loading, the seed elements will contain
zones of different fuel enrichment to improve core thermal per
formance. In addition, a burnable nuclear poison material will be
placed in selected compartments to improve core performance and
to extend its useful life.

1.9.2 Army Package Power Reactor—APPR 1

Fuel plates are composed of a dispersion of enriched UO, in stain


less steel clad with 304L stainless steel. A burnable nuclear poison
in the form of B,C is likewise contained in the fuel matrix. As indi
cated by Fig. 1.7, 18 of these fuel plates and two stainless-steel end
plates are brazed to spacing combs to form a subassembly. The over
all plate dimensions are 0.030 inch thick, 2.78 inches wide, 23.0 inches
long, with a 0.13-inch coolant space between plates. Similar fuel
assemblies are used to form part of the moving reactor control
elements.

FIGURE 1.7. Fuel Assembly, APPR 1 [6].


URANIUM DIOxIDE APPLICATION TO NUCLEAR Power REACTORs 29

1.9.3 Nuclear Power Demonstration Reactor—NPD–2

This reactor site is located approximately 16 miles up the Ottawa


River from the Chalk River Atomic Energy Project in Canada. Its
fuel elements consist of Zircaloy-2 clad high-density, sintered natural
UO2 dioxide pellets. These pellets are ground to

fit
0.025-inch

a
thick wall, Zircaloy—2 tube with uniform clearance. End caps are
automatically seal welded argon atmosphere. Each fuel rod

an
in

is
long As illustrated Fig. 1.8,

in

in
19.5 inches and inch diameter.
1

14incHES
l

NCHEs
3
|

FIGURE 1.8. Two NPD-2 Fuel Clusters [3].

two Zircaloy wires are wound about each element 19.5-inch pitch.
on
to a

The fuel elements are arranged groups form completed


of
in

nine
spiral promote mixing
of

fuel assemblies. The wire


is as to

intended
is

coolant radially throughout the fuel assembly well provide


to
an as as

spacing individual elements. Heavy water primary


of

used
a

coolant, and each fuel assembly contained within individual


is

pressure tube the reactor structure. This arrangement allows re


in

mote refueling operation.


of

fuel assemblies with the reactor


in

1.9.4 The Experimental Gas Cooled Reactor—EGCR

This reactor, currently the design stage, provide prototype


in

to
is

power producer experimental facility utilizing recirculating


an

and
30 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONs

I3.625 IN.

FIGURE 1.9. EGCR Fuel Element [5].

helium as a coolant. The high temperature capabilities of UO, fuel


elements are expected to be utilized in this reactor where coolant is
to enter at 510° F F. The EGCR, to be erected
and exit at 1,050°
on the Clinch River at Oak Ridge, Tenn., is scheduled for completion
by October 1962.
It is planned to use fuel elements of the type shown in Fig. 1.9
for this reactor. Cored UO, tubular pellets 0.70 inch OD, 0.32 inch
ID, and 0.75 inch long are stacked in a 251%-inch-long, 0.020-inch

H
thick, 304 stainless-steel tube. A 346-inch-thick MgO insulating spacer
is placed between the 0.015-inch-thick stainless end caps and the fuel.
The central void is used to eliminate center melting and to provide
a reservoir for fission gas retention. The UO, fuel will be enriched
to approximately 2.5 percent U*. Seven fuel elements are arranged
29.397
in overal-L
LENGTH
29Odoin stacked
LENGTH

oxideFuel.PELLETs

centeRingPLugs space
Rs
*-edoin
4.cº.In

-
--- -

TYPE3o4staural-Ess-s---
sta-LessTues
5.o.o.o.
in Magnesium
ox-DE
PELLETs
supportsuite
ve
GRAPHiteGRADEAGot

FIGURE 1.10. EGCR Fuel Assembly [5].


URANIUM DIoxIDE APPLICATION To NUCLEAR Power REACTORs 31

within a stainless-steel tube surrounded by a graphite moderator sleeve


as shown in Fig. 1.10. Six of these fuel assemblies are loaded into
a fuel channel within the reactor.
Because of the high cladding temperature and large thermal
gradients in this application, the effects of thermal stress are more
pronounced than in lower temperature applications. To minimize
the tendency of thermal gradients to produce fuel element bow, spacer
elements have been located at the midpoint of each fuel element. A
larger than normal fission gas retention volume has been provided
to reduce the effect of fission gas pressure on the cladding at elevated
temperatures.

1.10 SUMMARY

Duringthe past few years, UO, has gained wide acceptance as a


nuclear power reactor fuel material. This situation is in large part
due to its unusually good resistance to irradiation damage. In par
ticular, its use in bulk form has been demonstrated in the successful
operation of the Shippingport Reactor. As discussed in this chapter,
UO, also indicates potential as a power reactor fuel in a variety of
other material configurations, such as, for example, dispersions of
UO2 in a metal matrix and combinations of UO, with such metal
oxides as ZrO2, Th(O2, and BeO.
While UO, by no means provides a universal solution to all reac
tor solid fuel material requirements, and actually there has been rela
tively little power reactor operating experience with this fuel, the
considerable experimental evaluation of this material in- and out-of
pile has yielded results which provide strong incentive to exploit
further the potentialities of this fuel material.

REFERENCES
1. U.S. Atomic Energy Commission, “The Shippingport Pressurized Water Reac
tor,” Addison-Wesley Publishing Company, Inc., Reading, Mass., 1958.
2. J. D. Eiche NBERG, P. W. FRANK, T. J. KIsIEL, B. LUsTMAN, and K. H. VogFL,
“Effects of Irradiation on Bulk Uranium Dioxide” in “Fuel Elements Con
ference, Paris,” TID–7546, Mar. 1958, pp. 616–717.
3. “Nuclear Power Plants, Part 1,” “Proceedings of the Second United Nations
International Conference on the Peaceful Uses of Atomic Energy,” Vol. 8,
United Nations, Geneva, 1958.
32 URANIUM Dioxide: PROPERTIES AND NUCLEAR APPLICATIONS

4. “Nuclear Power Plants, Part 2,” “Proceedings of the Second United Nations
International Conference on the Peaceful Uses of Atomic Energy,” Vol. 9,
United Nations, Geneva, 1958.
5. “Gas Cooled Reactors,” Monograph No. 7, Journal of the Franklin Institute,
May 1960.
6. “Power Reactors,” U.S. Atomic Energy Commission, Technical Information
Service, May 1958.
Chapter 2

PREPARATION OF URANIUM DIOXIDE


J. A. FELLows, Editor

2.1 INTRODUCTION

Two main types of large-scale industrial methods have been used to


prepare uranium dioxide. By far the largest quantities of UO, have
been prepared by denitration of uranyl nitrate hexahydrate,
UO2(NO2)2·6H2O, to UO, followed by hydrogen reduction. Al
though the largest percentage of this oxide type manufactured in the
United States has been used for further conversion to UF, and then to
either uranium metal or the hexafluoride, large quantities have been
used directly in fuel element fabrication for nuclear power reactors."
The second most important industrial preparation method for UO2
is one suitable for fuel enriched in U*
content. This oxide is pre
pared from the hexafluoride through an ammonium diuranate inter
mediate and is commonly referred to as ADU oxide. A variety of
other methods, including variations of the two main types, has been
studied on an experimental scale, but none has been put into large
scale plant practice in the United States.
The following topics are discussed in this chapter:”
1. A short historical survey of the earlier experience with production
of UO2.
2. A of the present knowledge of the preparation of UO,
discussion
from uranyl nitrate and the presently available techniques for the
reduction of UOs to a grade of UO, suitable for use as a ceramic
fuel material.
3. A description of the commercial reduction of UO, from ammonium
diuranate prepared from uranium hexafluoride of assorted en
richments.
4. An outline of the preparation of UO, by steam and water oxidation
of uranium metal.

* This oxide source, referred to in this book as MCW (Mallinckrodt Chemical Works)
oxide. was used for the blanket portion of the Shippingport PWR.
* The chemistry of uranium ore processing is outside the scope of this book. References
on this subject are listed at the end of this chapter.
33
PREPARATION OF URANIUM DIOXIDE 35

was boiled off and the solution evaporated at a boiling point of 118°
to 120° C to obtain molten UNH, which was subsequently fed to a
stainless-steel gas-fired pot for denitration [2]. The contents of the
pot were agitated with a single-blade stirrer and heated for some
7 hours to form a UO,
powder containing less than 1.5 percent com
bined nitrate and water and less than 0.01 percent U3Os. Approxi
mately 25 percent of the product was in the form of a hard cake on
the pot wall, with the remainder a 30-mesh powder.
In nearly all cases, the final stage of preparation of UO, had been
the simple hydrogen reduction of UO, at elevated temperature. In
the initial plant process at St. Louis, the UO, was charged into
stainless steel trays about 1% inches deep [3]. Eight trays at a time
were placed in a stainless-steel box and heated in an electric furnace.
The box was flushed with CO, and hydrogen was admitted (cracked
ammonia was soon substituted because of the shortage of hydrogen
gas). After heating for 6 hours at 800°C, the box was withdrawn for
cooling with the hydrogen gas still flowing. A water spray was used
to hasten the final cooling, and CO, was again used as a purge. (The
prolonged cooling to room temperature was necessary to avoid re
version to UAOs at temperatures in air above 100° C.) This method
consistently produced material containing better than 97 percent UO,.
The high temperature of this reduction created a very stable form
of UO2, a type later found to be desirable for ceramic fuel material
use. Not until much later, however, was it realized that whenever
the preparation of UF, is intended, this high reduction temperature is
a distinct disadvantage in terms of the resultant reactivity of the
UO, with HF (see Chap. 3).
The production of UO, continued for a number of years on a batch
basis until the termination of World War II
permitted time and effort
for the engineering of improved plant handling. Since that date,
the production techniques have been converted to continuous processes,
and many improvements in equipment and process control have been
achieved. There have also been opportunities to investigate alternate
chemical procedures for the preparation of oxide powders of special
characteristics.

2.3 HYDROGEN REDUCTION OF UO, OBTAINED FROM


URANYL NITRATE
W. G. Weber

2.3.I Manufacture of UO,

(a) Concentration of Uranyl Mitrate Solutions

The process for manufacturing UO, from uranyl nitrate solutions


consists of two major steps: (1) The uranyl nitrate solution is con
57.4789 O–61–4
PREPARATION OF URANIUM DIOXIDE 35

was boiled off and the solution evaporated at a boiling point of 118°
to 120° C to obtain molten UNH, which was subsequently fed to a
stainless-steel gas-fired pot for denitration [2]. The contents of the
pot were agitated with a single-blade stirrer and heated for some
7 hours to form a UO,
powder containing less than 1.5 percent com
bined nitrate and water and less than 0.01 percent U3Os. Approxi
mately 25 percent of the product was in the form of a hard cake on
the pot wall, with the remainder a 30-mesh powder.
In nearly all cases, the final stage of preparation of UO, had been
the simple hydrogen reduction of UO, at elevated temperature. In
the initial plant process at St. Louis, the UO, was charged into
stainless steel trays about 1% inches deep [3]. Eight trays at a time
were placed in a stainless-steel box and heated in an electric furnace.
The box was flushed with CO, and hydrogen was admitted (cracked
ammonia was soon substituted because of the shortage of hydrogen
gas). After heating for 6 hours at 800°C, the box was withdrawn for
cooling with the hydrogen gas still flowing. A water spray was used
to hasten the final cooling, and CO, was again used as a purge. (The
prolonged cooling to room temperature was necessary to avoid re
version to UAOs at temperatures in air above 100° C.) This method
consistently produced material containing better than 97 percent UO,.
The high temperature of this reduction created a very stable form
of UO2, a type later found to be desirable for ceramic fuel material
use. Not until much later, however, was it realized that whenever
the preparation of UF, is intended, this high reduction temperature is
a distinct disadvantage in terms of the resultant reactivity of the
UO, with HF (see Chap. 3).
The production of UO, continued for a number of years on a batch
basis until the termination of World War II
permitted time and effort
for the engineering of improved plant handling. Since that date,
the production techniques have been converted to continuous processes,
and many improvements in equipment and process control have been
achieved. There have also been opportunities to investigate alternate
chemical procedures for the preparation of oxide powders of special
characteristics.

2.3 HYDROGEN REDUCTION OF UO, OBTAINED FROM


URANYL NITRATE
W. G. Weber

2.3.I Manufacture of UO,

(a) Concentration of Uranyl Mitrate Solutions

The process for manufacturing UO, from uranyl nitrate solutions


consists of two major steps: (1) The uranyl nitrate solution is con
57.4789 O–61–4
36 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

centrated by evaporation to the approximate composition of uranyl


nitrate hexahydrate, and (2) the UNH is heated further, causing addi
tional dehydration and then denitration to yield UO, [4, 5]. The
equation for this reaction is:
heat
2UO,(No.),6H,0° $12H,ot +4No. 1 +o, t +2UO, Eq. (2:1)

The method utilized for concentration of the uranyl nitrate solutions


is dependent upon the initial concentration of liquor as received from
the preceding purification steps. This concentration can vary from
90 grams to as high as 280 grams of uranium per liter and from 0.00
to 0.05N HNOs content.
Boiling points of these liquors vary from 120° C (1,200 grams of
uranium per liter) to 143° C (1,450 grams per liter). Concentration
is done both with boildown tanks and with vertical tube evaporator
systems. The degree to which the solutions are concentrated varies
between plants, depending upon the subsequent method to be utilized
in denitration to UOs. Because the freezing points for these concen
trated liquors can vary from 58° to 116°C, the entire system must
be steam-traced (or jacketed and insulated) to prevent freezing, and
all

permit ade

as
sloped and fitted
be

in

to
lines must such manner
a

quate draining and flushing with water. Generally, AISI type 304L
stainless steel used throughout the system, although type 309Cb
is

has proved successful for the evaporator tubes.


Vapors from the various concentrating units are handled

in
several
ways. The condensed liquor from vapors containing entrained uranyl
appropriate point
an

nitrate recycled the process. Condensates


in
to
is

containing enough nitric acid make recovery economical are pumped


to

no

or
concentrating units. Condensates that contain nitric
to

uranium
acid are discarded.
Four the sites currently manufacturing UOs carry out the de
of

nitration step batch reactors, while fifth uses continuously


in

a
a

stirred, trough-type denitrator. Development work also proceeding


is
on

type.
of

another continuous denitrator fluidized-bed


a

(b) Batch Denitration

The major portion commercial production has been done gas


of of

in

heated, agitated pots the type shown Fig. 2.1 [3]. Two sizes
in

are currently used. the two (capacity, gallons)


45

50
of

The smaller
to

dish-bottom pot with 30-inch ID,


and 18-inch vertical side wall.
is
a

The pot generally made type 304L


of

and 30-inch dish radius.


or
is
a

347 stainless steel. Each pot equipped with vertical gearhead


is

driven agitator having


an

output speed rpm.


36
of
by

The agitator, fabricated from 4-inch type 304L stainless-steel


14
PREPARATION OF URANIUM DIOXIDE 37

FIGURE 21. Denitration Pots [3].

the pot with


fit

bar stock, is designed to


of

the contour clearance


a
of

design necessary large build


to

about inch. Such prevent


¥9

is
a

up crusted UO, between the agitator and the pot wall. The pots
of

months, generally burning out


on
18
of

life from
to

have the
6
a

bottoms, which are subjected the highest temperature.


to

Normally, 7%-hp motor the agitator drives.


on

possible
a to
is

use
it

However, most plants,


of

small amount sulfuric acid added


in

is

the concentrated uranyl nitrate solution after


to

has been drained


it

into the pots. This increases the caking tendencies, that con
so

siderably more power the agitator and necessitates the


on

required
is
of

use 15-hp drive motor. -


a

on

The small pots operate 5-hour cycle. The concen


nominal
a

trated uranyl nitrate solutions are stored heated tanks, and the
in

liquors are pumped through circulating liquor header above the


a

pots (55 gallons are periodically withdrawn from the header pot
as
a

charge). Heating ceramic-lined furnace with four


in

carried out
is

gas burners near the bottom the pot. The entire heating operation
of

strict time-cycle basis. During the first minutes, only


on

25
is

done
a

liquor
on
of

two the four burners are turned concen


to

increase the
tration without causing excessive boiling. To prevent splashing, the
agitator not used during the boildown period. Then, all four burn
is

ers are turned on. Heating proceeds for the next


55

minutes until
the material has been concentrated point that approaches
to

such
it
a
38 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

a doughy consistency, and the agitator must be used to prevent solidifi


cation into a concrete mass. During the following 2-hour period of
heating in this manner, there is a heavy evolution of NO2 fumes as
the UO, (NOA), is decomposed to UOs. Then two of the burners are
shut off, and 20 minutes later all heating is stopped. Cooling is con
tinued for another hour until the powder is cool enough to be unloaded
by a pneumatic system. Agitation is continued through the cooling
cycle to aid in cooling, to prevent a burned product, and also to ensure
the formation of a granular UO, free of lumps which can be easily
discharged through the pneumatic tubes. Unloading and reloading
require another 20 minutes.
The second type of pot used in UO, production (Fig. 2.1) has a
design similar to that of the 50-gallon pot. The larger pot has a 66
inch ID, a 42-inch vertical side wall, and a dished bottom. A 40-hp
gearhead drive is used to rotate an agitator which has about the same
configuration and clearance as those employed with the small pot.
The large size pot operates on a nominal 8-hour cycle. As in the
case with the small pots, concentrated uranyl nitrate solution (275
gallons) is drained into the pot from a circulating liquor header.
Heating is done by means of three concentric rings of small radiant
gas burners located about 22 inches below the pot bottom. Opera
tion of these pots differs somewhat from that of the small pots in that
the burners are controlled by the temperature of the combustion gases.
Initially, this temperature is adjusted to about 620° C for 2 hours,
then reduced to 510°C for 4 more hours. At this point, heating is
stopped, and the powder is cooled for an additional 11% hours with
the agitator still turning.
The operating conditions for each type of pot have been developed
empirically to produce a satisfactory quality of UOa. The denitra
tion cycle of the larger pots is longer and, hence, at a slower rate. The
products from the two pot sizes appear to differ in reactivity and also
in the larger proportion of the orthorhombic crystalline structure
from the smaller pots. Difference in type of trioxide and in reac
tivity affects the reduction to dioxide and the final properties of the
dioxide powder. Denitration at excessive temperatures can cause the
formation of UAOs on the pot wall, since at temperatures above ca.
590° C this oxide is the more stable form. Conversely, denitration
performed at too low a temperature results in a product containing
excessive water and nitrate.
Both types of pots have close-fitting, half-circle removable lids.
Thus, an opening is provided for the unloading of the UO, powder
by means of a pneumatic handling system. Figure 2.2 shows the
manner in which the unloading operation is performed on a small
pot. The U(), powder, normally containing some lumpy material
PREPARATION OF URANIUM DIOXIDE 39

FIGURE 2.2. View of a Denitration Pot Being Unloaded Pneumatically.

formed around the agitator shaft and on the pot walls, is drawn into
a pneumatic tube and fed into a collection hopper. The powder then
passes through a hammer mill and a continuous sampler and is finally
fed into a packaging hopper for loading into a suitable container.

(c) Agitated-Trough Continuous Denitration

The most commonly used methods for continuous denitration of


uranyl nitrate hexahydrate involve the addition of the uranyl nitrate
liquor to a hot layer of UO, powder, either in a horizontal stirred bed
or in a fluidized bed. As yet, these methods have not been widely
used. However, they may become more important, since they can be
made to produce particle shapes that differ from those of the batch
methods. They also offer several advantages over the pot type of
operation, e.g., reduced radiation exposure during pot unloading, re
duced uranium dust hazard, elimination of foaming or caking or both
in the pots, and a more uniform loading of the acid recovery system.
The agitated trough type of continuous denitration (Fig. 2.3) was
first extensively investigated on a pilot plant basis in 1953; in 1956
a plant was started in operation at Hanford [6, 7]. This process
utilizes a round-bottom trough, 26 inches wide and 12 feet long, made
40 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

AGITATOR DRIVE

UNH FEED LINEs—

U03 Powder BED


OFF-GAS LINE

-
Powder DISCHARGE

Powder HOPPER

AGITATOR PADDLES AND DIRECTION


OF ROTATION

|- ELECTRICAL HEATER ELEMENTS

FIGURE 2:3. Continuous Denitrator for Conversion of Uranyl Nitrate Hexahy


drate to Uranium Trioxide [8]. (From Charles D. Harrington and Archie E.
Ruehle, “Uranium Production Technology,” D. Wan Nostrand Co., Inc., Prince
ton, N.J., 1959.)

of 1-inch-thick type 348 stainless steel. An 11-foot section of the


trough is heated in an electric resistance furnace, the elements of
which are controlled by thermocouples inserted at intervals in the
trough shell. An operating temperature of 510° to 540° C is normally
used to prevent excessive metal temperatures which can contribute to
shell distortion.The agitator is of the nonconveying T-bar type and
consists of an 8-inch diameter type 348 stainless-steel shaft with 12
arms. Spacing between blades provides for the placement of feed in
jector pipes and thermocouples. The design and operation of the
agitator were found to be critical. The material in the trough must
not be conveyed at an excessive rate, the agitator must be adjustable
so that close clearances can be maintainedbetween the agitator blades
and the trough wall, and the agitator must have enough strength to
maintain these small clearances under operating conditions. A clear
ance of no more than 1% inch must be maintained because the tendency
of cake to form on the trough wall has a pronounced effect on heat
transfer through the reactor wall and can radically affect the general
operation of the reactor. Conversely, the presence of sulfate in the
feed tends to improve heat transfer.
Heat transfer can also be affected by agitator speed. Plant tests
have shown that the peripheral speed of the agitator must be such as
1.4

to produce 1 to times gravitational acceleration order achieve


to
in
14 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

a release of 8 to 10 percent of the total fission gas produced has been


measured. It is postulated that this remarkable ability of UO, to
retain fission gases without marked structural effects results from
accommodation of the gas atoms in the UO, lattice. Fission gases
appear to be released from the UO2 body primarily by a diffusion
mechanism. Thus, the amount of exposed UO, surface, fuel operating
temperature, and time at temperature are important variables in de
termining fission gas release.
Experimental observations indicate that fission gas release is sub
stantially increased above a temperature of approximately 2,000° F.
As the fuel approaches the melting point of UO, essentially

all
of
the
fission gases are released. Lower UO, fuel density likewise known

is
For example,
to

increase fission gas release from irradiated fuel.


under similar exposure conditions Shippingport type fuel rod con

a
taining percent dense fuel was observed
94

to

10
have times the fission
gas releases rod containing percent dense fuel. 97
of
a

(c) Stability High Temperature Water


to

Prolonged irradiation tests have been performed


which demonstrate
the continued stability UO. water after rela
high temperature
of

in

tively high fuel burnup. Zircaloy-2 clad UO, fuel rods that have
been intentionally defected with the cladding
in
0.0005-inch hole
a

have operated satisfactorily with swelling under irradiation for


no

a
period year fuel burnup approximately 12,000
of

of
to
in

excess
a
1

UO. At significantly higher fuel burnup (60,000


of

MWD/Ton
MWD/Ton), Zircaloy-2 clad fuel plates containing 0.040-inch-thick
14-inch-wide UO, fuel platelets have shown volume increase ofap
a

proximately percent when exposed high temperature water (ap


20

to

proximately fuel surface). Recent in-pile and out-of-pile


at

700°
F

that high burnup UO, can lose stability


its

data indicate, however,


high temperature water.
in

(d) Thermal Fracture

Some breakup bulk UO, during reactor operation


of

as

observed
is

Fortunately, however, in-pile thermal


of

result thermal shock.


a

cycling experiments have confirmed that the limited thermal shock


UO, prevent pulverization UO, bodies
of

of
to

resistance sufficient
by is

during operation process


of

continual thermal fracture. There


a

evidence that reporteddecreases UO, thermal conductivity


in in
is

irradiation effects may large part more correctly


be
to

attributed
assigned cladding during
of
to

lack contact between fuel and


a

operation.
42 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONs

Figure 2.4. Unground UO, as Taken from a UO, Pot; Approximately X*


Reduction Factor, 1/3.
ARATION OF URANIUM DIOXIDE 43

FIGURE 2,5. [7].


-

44 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS -

FrgurE 2.6. Layered Formation of UO, Particles Made in a Continuous Deni.


trator; X125 [7].
PREPARATION OF URANIUM DIOXIDE 45

TABLE 2.2—PARTICLE SIZE DATA: UO, MADE IN AGITATED TROUGH [6]

Screen size, mesh

8 16 40 60 80 100 200 3.25

Maximum percent
passing----------- 100. 0 ||100. 0 || 99.3 93.8 67.9 58.6 || 10.7 5. 5
Average percent
passing----------- 99.9 || 99.7 | 82. 7 || 54. 7 || 26. 0 | 16.9 3. 8 0. 5
Minimum percent
passing----------- 99. 2 || 98. 1 || 16. 0 2. 6 0. 7 0.3 0.0 0.0

(d) Fluid-Bed Denitration

As a means for contacting finely divided liquids or solids with a gas,


the fluid-bed process offers several distinct advantages over other
more established methods. These advantages include the fact that

is,
there is little limitation on the size of fluidized-bed reactors. It
therefore, possible capital equipment
in
to

effect substantial reductions


by

of

one large fluid-bed replace several small


to

costs the use reactor


units utilized for another method. Additional savings

in
lowered
maintenance costs and production downtime may
be

obtained, since
there are no moving mechanical parts fluidized reactors. An equally
in

important advantage from the standpoint efficient operation and


of

quality control
be

the fact that uniform temperature can achieved,


is

simplified. application the fluid-bed prin


of

and heat removal The


is

ciple production logical,


be
to

appears
of

UOs
to

the since excellent


heat transfer between metal and fluidized solids can be obtained this
in

type
of

reactor.
The development fluidized-bed process for application
of

to

the
a

by

uranyl the Argonne


of

denitration nitrate was undertaken


in

1953
National Laboratory [9]. Initially, experiments were conducted
in
a

bench-scale, 3-inch diameter pipe containing three fluidized beds, each


supported by perforated plate.
of of

later version the denitrator


A
a

Argonne demonstrate the practicability


at

the process
to

used
is

Fig. 2.7 [8,10]. The body


of

of

shown this denitrator consists


in

to a

6-inch diameter stainless-steel pipe. sintered metal plate serves


A

support the bed UO, and also ensure adequate distribution


of

of
to

the
fluidizing air across the entire cross section
In
of

to

the tube. order


achieve adequate distribution and contact with the rapidly moving
particles UO, the reactor denitrator, the uranyl nitrate solution
of

in

by

sprayed into the tube spray nozzle. Dust entrainment re


is

is
a
by

the tube through


of

passing the exit gases out the top


of

duced
sintered metal filters. The UO, produced
of

the form small


of in
is

spherical particles having


an

average diameter
to

150 200 microns.


The UO, regular intervals from the bottom
at

of

withdrawn the
is
46 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

PRESSURE RELIEF off GAs


vaLVE
BED-LEveL TELEGAGE

PURGE AIR —-–Q |


|

__*
– DISENGAGING SECTION

Uos CHARGE
HoPPER
-
.
P.
B
PTFºtºs,
2 BANKS-4
ºf
PoRous stalNLESS STEEL
#55.
FILTERS EACH

ºD
| T-Divider-DisengagiNG

- E
SECTION

PURGE AIR
D | |T-THERMocouple Tube
3-way PLUG
VALVE VSZ -—Dise NGAGING SECTION
Sº º
PRESSURE TAP
PLUG WALVE

-
FLUIDI2ED
LU UO-3. BED
º-CALROD HEATERS

AIR

SPRAY NOZZLE
C
FEED

SPRAY NOZZLE

AIR- -—
PRESSURE TAP

GAs CHAMBER

-THERMocoupl_E TUBES
POROUS STANLESS

-—
STEEL DISK PLUG VALVE

PRODUCT RECEIVER

PLUG VALVE-- -—PLUG valve

-—SAMPLE collectOR

PRODUCT RESERVOIR

FIGURE 2,7. Six-Inch Pilot Plant Fluidized-Bed Denitrator [8]. (From


Charles D. Harrington and Archie E. Ruehle, “Uranium Production Tech.
nology,” D. Wan Nostrand Co., Inc., Princeton, N.J., 1959.)
PREPARATION OF URANIUM DIOXIDE 47

reactor through a product reservoir which can be replaced when full


without interrupting operations. This particular unit is capable of
operating at a rate of 100 pounds of UO, per square foot per hour.
Using this design experience, Mallinckrodt Chemical Works built a
Somewhat larger, 4-foot-high, 10-inch diameter unit of type 347 stain
less steel for further study in the pilot plant. Figure 2.8 shows a
schematic diagram of this denitrator [8, 11]. The unit differs from
the one at Argonne in several respects:
1. Heat transfer has been increased by the use of eighteen 4-inch
diameter, 2-kw electric heaters which protrude into the reaction bed
through the bottom perforated air distributor plate.
2. The fluidizing air is preheated in a gas-fired furnace.

3. The product is removed from the denitrator by means of a 1%

inch diameter pipe located 33 inches from the bottom of the bed.
4. A cyclone separator is used to separate dust from the off gases,
with the solids returning into the powder feed hopper and, thence,
back into the reaction bed. This unit has been satisfactorily oper
ated at rates of 150 pounds of UO, per hour per square foot.
Operation of the reactor proceeds as follows:
1. Air is first passed through the spray nozzles.
2. The reactor is charged with approximately 200 pounds of UOs
from the powder feed hopper.
. The flow of preheated fluidizing air is started.
. The bed is brought up to temperature with the electrical heaters.
:. Water is fed through the feed nozzles and, after a short time, the
change-over is made to uranyl nitrate solution which has been pre
viously adjusted to the desired concentration in the holding tanks.
6. The feed rate is adjusted to the desired value, and the system is
balanced for optimum performance.
Table 2.3 presents a comparison of some of the more pertinent physi
cal and chemical properties of the various types of commercially pro
duced UOs. It is readily apparent that the material produced by the
two continuous methods has a markedly lower number of fine particles.
This condition is desirable from one standpoint, for the presence of
very small particles is a potential health hazard. However, whenever
fine particles are absent the chemical reactivity of the product is fre
quently low. But the reactivity may be improved by grinding the
material.

2.3.2 Reduction of UO, with Hydrogen

(a) Laboratory Kinetic Studies

As would the method of producing the UO, has a


be expected,
marked effect on its rate of reaction with hydrogen and on the nature
of the rate law followed. Figure 2.9 shows examples of the wide
48 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

*—
TO NITRIC ACID

Hi
UNH LINE REcovery


s
STEAM tº

UNH
- HOLD TANK LOADING
OFF-GAs
LINE
(
Powder
FEED
HoPPER

*/
RECYCLE
LINE

º i.
ſº
s CLEAN OUT
PORT-- Rotary
[CD VALve

U-TUBE REFEED
MANOMETER LINE

siſ

THERMO
COUPLE
| WELL
ExTERNAL
HEATERs (42)

|||| –8 -—
(4)

METER X ºnozzles PRODUCT


INTERNAL LINE
HEATERS (18)

NOZZLES
~
UNH PUMP Ø * Roſa
AIR

ATOMIZING
AIR -- 7 METER
DISTRIBUTING
`
PLATE Rotary
VALve

E FLUID AIR
"TPREHEATER
To
v.ACUUM
2
li /

AIR
BURNER ROTAMETER OBLAST
--
GATE

alsº Yºgas

FLUID ZING
AIR

FIGURE 2,8. Ten-Inch Pilot Fluidized-Bed Denitrator [8]. (From Charles D.


Harrington and Archie Ruehle, “Uranium Production Technology,” D. Van
E.

Nostrand Co., Inc., Princeton, N.J., 1959.)


PREPARATION OF URANIUM DIOXIDE 49
roo T I - I
|
I I I

90
I-TYPE UOs PRODUCT OF

III
CONTINUOUS DENITRATOR
2-TYPE UO3 FROM UNH
-

III
80
(REFINERY A)
3-TYPE UO3 FROM UNH

III


70 (REFINERY B)
iſ
3 #

4-AMORPHous Uos
60
-
IT
f

-
§ o

50
it

->
C.
-
3 #

40
2

- -
30
9

-
3

20
4.

lo -
I
I

I
l

IO 2O 40 5O 60 8O

9
o

TIME IN MINUTES

Hydrogen Reduction Rate Curves for Warious Uranium


at

FIGURE 2,9.
C

475°
Trioxide Preparations Harrington and Archie
D.

[8]. (From Charles

E.
Ruehle, “Uranium Production Technology,” Wan Nostrand Co., Inc., Prince
D.

ton, N.J., 1959.)

TABLE 2.3–PROPERTIES OF COMMERCIAL UO,

Batch UOs Continuous UOs

Small pots Large pots Agitated trough Fluid bed”

Percent UO,---------- 98.3 97.8 99.4


|-
-
-
-
-
-
-
-
-
-

Percent U20s---------- 0.36 0.96 |------------


0.
1

H.O- 1.07
0.

15
1.

Percent
0
|-

0. --
-
-
-
-
-
-
-
-
-
-
-

-
-
-
-
-

NO. ----------|------------ 0.2


0.

58
5

Percent
Fe, ppm-------------- 13 <50 <50 <50
Cr, ppm-------------- -------- - - - -
4
5
2

Yi, ppm-------------- <5 <10


1

ppm --------------- 0–100 250 |1, 500–2, 000 0–800


S.

4,

Bulk density
4.
3.

2.

3
2
5
7
-
-
-
-

-
-

-
-
-
-

SCREEN ANALysis

Percent plus 40 mesh--- 0.5 (+20) 1–1.5 (wet) 15 16


||

(unground) (unground)
Percent minus 325
,

mesh-- <1
2

50 75 (wet)
-

-
-
-
-
-
-
-
-
-
-
-
-

(unground) (unground)

"10-inch diameter pilot plant unit.


50 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I T I I i I
|OO

80

60

| 40 ºr

2O

I I
o 4o 8o I2O 14O 2OO 240 280

TIME IN MINUTES

FIGURE 2,10. Dependence of Reduction Rate of Type Orange Oxide on Tem III
perature [8]. (From Charles D. Harrington and Archie E. Ruehle, “Uranium
Production Technology,” D. Van Nostrand Co., Inc., Princeton, N.J., 1959.)

variation in reduction rates obtained with Type UO, * (orange. III


colored orthorhombic crystals) produced by various means. Figure
2.10 presents the results of some thermobalance studies and shows the
rate of reduction of UOs in a large excess of hydrogen. An apparent
activation energy of 35.1+0.6 kcal/mole can be obtained from a plot
of log k vs 1/T for the major portion of the curves (10 to 90 percent
completion) which follows a first order rate law [12]. Rather close
agreement of 34.6 kcal/mole has been obtained on UOs produced in
commercial equipment at Hanford [13]. Hydrated UO, reduces at
appreciably faster rates than the anhydrous Type product III
ob.

tained from UO, (NOA), .6H2O [13, 14]. The enhanced reactivity
by

not destroyed dehydration the UO, hydrate prior


of

to

reduction.
is

Comparisons UO, produced from UO2(NO2), and ammonium


of

diuranate (ADU) show the ADU UO, more closely aggregated


be
to

[15]. The ADU UO, consists large, closely packed aggregates


of

having mean diameter over 100 microns (diameter range


to
of

220
a

microns), while the UNH UO, has


of

mean diameter 20 microns


a

range
of 45

microns). Surface area measurements made


to

(diameter
1

UO, particle size and reduction temperature


to

determine the effect

UOs, Y—UOs, has frequently as Type III


to

been referred oxide


of

The most stable form


*
PREPARATION OF URANIUM DIOXIDE 51

upon the particle size of the resulting UO, show that reduction at
700° C causes a decrease in the mean particle diameter by a factor of
2, while annealing at 900° C in hydrogen causes an increase in aver
age particle diameter.
Anderson, et al., found that particle size increased when reduction
was performed above 650° C [16]. These same investigators found
that the density and particle size of the UO, depends upon the
density and particle size of the UO, from which it has been made
and upon the reduction temperature. They also found greater particle
growth during reduction when using amorphous UO, having a small
particle size. However, the crystallite size of the UO, apparently
depends only on the temperature of the reduction (see Chap. 3) [16].
Kuhlman [17] found that ammonia is less effective than either.
hydrogen or a 3-to-1 molar mixture of hydrogen and nitrogen in
reducing UOs. This situation exists in spite of favorable thermo
dynamics. The reduction rate is slightly reduced by the presence of
nitrogen diluent [18]. The use of water as a diluent significantly
as a
reduces the reduction rate, a 10-mole-percent dilution providing a
reduction rate only one-third of that obtained with pure hydrogen
[19, 20). A marked decrease in reduction rate has also been noted
with the presence of minute amounts of hydrogen fluoride [21, 22].
This might be attributed to the formation of a protective film of
uranyl fluoride on the surface of the UOs, since it is known that
UO.F, is reduced by hydrogen at a much lower rate than UO, [23].
An increase in the reduction rate of UOs can be achieved by incorpo
rating a sulfate group in the UO, through addition of sulfuric acid,
sulfamic acid, sulfur, etc., to the uranyl nitrate solution before the
denitration step [24, 25]. A minimum of 150 ppm is required to
induce an appreciable increase in reduction rate. No satisfactory ex
planation of the effect has been demonstrated, although it is known
that the diffraction pattern of sulfated UO, is diffuse, in contrast to
the sharp pattern obtained for unsulfated UOs.

(b) Batch Reduction

The first large-scale production of UO, from UO, was performed


in reactors of the type shown in Fig. 2.11. Essentially, this con
sists of a stainless-steel box about 13 inches high by 17 inches wide
by 47 inches long into which eight stainless-steel trays are placed on
a rack. One end of the box is fitted with a bolted end plate through
which the trays are loaded and unloaded. Gas enters through a pipe
which extends through the end plate to the rear of the box, thereby
permitting the gas to pass completely through the box and leave by a
pipe which extends just inside of the end plate.
Loading of the incoming UO, into the reactor trays is done manually

57.4789 O–61–5
52 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONs

FIGURE 2,11. Batch UO, Reactor [8]. (From Charles D. Harrington and
Archie E. Ruehle, “Uranium Production Technology,” D. Wan Nostrand Co.,
Inc., Princeton, N.J., 1959.)

under a hood. The UO, is scooped out by hand and spread evenly
on a tray. When the reactor box is filled with eight trays, the end is
bolted in place, and the dissociated ammonia supply line and a CO.
purge line are connected. The reactor is placed in an electric furnace.
As a safety precaution, to prevent possible hydrogen explosions, a
3-minute purge with CO2 is given the box to remove all residual air
before the dissociated ammonia is admitted to the reactor. The disso
ciated ammonia is made by passing anhydrous ammonia through a
standard ammonia cracking unit at 870° C.
The mixture of nitrogen and hydrogen is measured through rotam
eters and is fed to the reactor boxes by means of long rubber hoses
hanging from the ceiling. This affords much flexibility in switching
the boxes. The gas flows over the UO, for 5 hours with the furnace
temperature at 815° C. The box is then removed from the furnace and
placed in a cooling area with the dissociated ammonia supply line still
connected. After air cooling for 31% hours, it is placed under a water
spray for 5 hours. All cooling is done with the dissociated ammonia
flowing through the box to prevent the UO, from being oxidized to
PREPARATION OF URANIUM DIOXIDE 53

U.Os. The flow of dissociated ammonia is stopped, and the box is


purged with CO2 for 3 minutes. The trays are dumped manually (un
der a hood) into a special 30-gallon drum with baffles. The drum is
rolled for 15 minutes, and the UO, is dumped into a hopper from which
packaging is done.
The exhaust gas carrying the excess dissociated ammonia is burned
immediately as it leaves the reactor. The burned gas then passes
through a reverse-jet, high-efficiency bag dust collector to prevent
contamination of the surrounding atmosphere.
This type of reactor requires 1.5 times the stoichiometric
about
hydrogen needed for the reaction. It
has a production rate of about
40 pounds of UO, per hour and uses about 4 pounds of dissociated am
monia per hour.

(c) Stirred-Bed Reduction

Development work on a continuous stirred-bed reduction process for


converting UO, to UO, was begun by Mallinckrodt Chemical Works
in 1949. Pilot plant scale experiments showed that the product was
comparable to that obtained from batch processes and that the opera
tion was cheaper and safer. In 1951 the first manufacturing plant to
utilize this process was built at St. Louis, Mo.; since then three more
plants have been built at various sites by the AEC [26]. Figure 2.12
shows a schematic diagram of a stirred-bed continuous reactor. Basic
ally, this operation involves passing hydrogen gas over UOs that is
being conveyed through a heated tube by means of a screw type mech
an ISrn.

(1) EQUIPMENT. The UO, is received in 30-gallon drums (820


pounds of UO,) or in hoppers (about 5,000 pounds of UOs). In order
to minimize the dust problem, the dumping from the 30-gallon drums
into a large hopper is done in a completely enclosed system. The mate
rial in the hopper is fed by a screw conveyor through a cone-type
crusher into a 6-inch Hapman disk-in-tube type of conveyor. The
Hapman conveyor is essentially a chain (made of magnesium disks
linked together) which is drawn slowly through a stainless-steel tube.
The hopper and screw feeder are necessary to prevent choking of the
crusher when a drum is dumped. The conveyor moves the UOs to
the top of the building and distributes it to large storage weigh hop
its

pers. Each reactor has own individual weigh hopper, containing


UO,
In

about 8,000 pounds. cases where the shipped 5,000-pound


in
is

hoppers, the hopper hoisted directly the top the building where
of
to
is

platform, hopper
on

of

scale and the bottom the connected


is

is

set
to it

feed UO, From this point


to

to

the mechanism used the reactor.


on, the equipment UO, storage.
of

the same for both methods


is
54

NBAO

JO1Ow38 TTBHS
URANIUM

LH9IT 18Oddſſs
0BBdS-ETgvlaeva
Svº 13.Tuno HSºmd
dº Tº Sv930v. 1 13TNI BA||80

©||H

XXXXXXXXXX&
3O^Od 13.TunO

M
830AWOd1BTNI

QN3 0NIww38
||| {&
·

ºhnĐINH ‘ZI' IboſdáL snomuſquo;O poſſ-po-1111S uoņoupoſ uoqoboſ [8]


DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
PREPARATION OF URANIUM DIOXIDE 55

R188oNFLIGHT PUSHERBLADE

FIGURE 2,13. Stirred-Bed Agitator Conveyor [8]. (From Charles D. Harring


ton and Archie E. Ruehle, “Uranium Production Technology,” D. Van Nos
trand Co., Inc., Princeton, N.J., 1959.)

A screw feeder on the bottom of the weigh hopper feeds into a small
seal hopper, thus, preventing the reacting gases from leaking back into
the main feed hopper. A paddle type level indicator in the small seal
hopper maintains the seal and actuates the screw feeder on the bottom
of the weigh hopper to maintain a constant level. The UOs is fed out
of the seal hopper and into the main reactor by means of a variable
speed screw feeder.
The reactor consists of a 16-inch diameter horizontal tube, approxi
mately 22 feet long, made of 14-inch type 309 stainless steel. Inside
the tube a specially designed rotating agitator-conveyor, made in a
form similar to that of a reel-type lawn mower (Fig. 2.13), rotates
slowly at constant speed, imparting a slight forward motion to the
powder. The conveyor, also of type 309 stainless steel, possesses
exceptional strength at high operation temperatures.
The outside diameter of the rotor is 14 inches. A 6-inch diameter
pipe shaft supported on bearings at each end of the reactor acts as
the main support and prevents the conveyor from touching the inner
surface of the tube. Experience has shown that if the agitator is
allowed to come in contact with the tube over much of its length,
sufficient galling will occur between the agitator blades and the tube
to create small metallic particles which can cause excessive contamina
tion of the product.
The product is fed out of the reactor into a horizontal cooling screw
about 6 inches in diameter and 10 feet long and from there into a
blending and packaging system. The cooling screw is partially im
mersed in a trough containing cooling water, and cooling water is
also circulated through the hollow shaft of the screw conveyor. The
agitator conveyor is normally rotated at four rpm. A tachometer
is attached to the rotor drive in each system. Recording and indi
cating instruments are used on the hydraulic pressure system, and
recording wattmeters are used on the electric motors of the agitator
drives. Slight caking in the reactor causes no undue increase in drive
torque. The small screw feeders and rotary valves of the system are
driven by variable-speed hydraulic transmissions which are individ
ually powered by electric motors.
For about 18 feet of its length the reactor tube is encased in an
electric resistance furnace (rated at 38 kw). The furnace itself is
56 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS l

divideu into four heating zones, with heating elements located along
the sides of the furnace. A control thermocouple, inserted through
the side of the furnace, is situated midway in each temperature con
trol zone, so that its junction rests beneath the center of the reactor
tube, just outside the shell. This arrangement permits variation of
the temperature between zones. An additional recording thermo
couple is located in each zone.
Dissociated ammonia gas (from a standard 2,000 cfm dissociator)
is automatically metered by a rotameter-type flow controller into the
reactor and passes countercurrent to the motion of the UOs bed.
The exit gas (hydrogen, nitrogen, and water formed in the reaction)
is burned immediately after leaving the reactor. A hood and a dust
collector connection are located over the top of the flame, and the
burned gases are drawn through a reverse-jet, bag-type dust collector.
Enough additional air is drawn into the vent line after the burner to
ensure that the temperature of the gases entering the dust collector
is not above 90° C (so that the bags do not ignite) and also to ensure
that in the event of a flame failure (this is always a possibility, even
though an automatic sparking device is used) the concentration of
hydrogen is diluted well below the explosive limit. A thermocouple
is placed in the exhaust flame in order to sense a flame failure and
automatically to shut off the flow of dissociated ammonia to the re
actor. Ammonia flow is also automatically stopped and the reactor
purged with nitrogen in the event of power failure. Failure of the
dust collector for any reason causes the exhaust gases to be automati
cally diverted to the atmosphere through a stack.
Vents to dust collectors are provided on all storage and feed hop
pers, and a dust hood encloses the end of the reactor. The hood is
used only when it becomes necessary to remove the agitator from the
tube. The dust hazard involved in removing agitators is minimized
by pulling the agitator into a large black iron coffin that is placed
immediately adjacent to the end of the reactor tube. The coffin is then
closed and can be removed to another area for cleaning and repair
without scattering dust throughout the work area.
Since temperatures as high as 1,000° C are not uncommon at certain
spots in the reactor, the fabrication materials must have good creep
strength characteristics at high temperatures. They must be resistant
to attack by hydrogen, hydrogen sulfide, and nitric acid vapors as well
as being resistantto the abrasive action of the powders passing
through. Illium R and types 309 and 316 stainless steels have proved
successful for the reactor tube. The agitator ribbons may be made of
Illium R, Hastelloy C, or type 316 stainless steel, while the pipe shaft
used to support the ribbons is made of Inconel or type 309 or 316
stainless steel. Asbestos is generally used for gasketing and packing.
PREPARATION OF URANIUM DIOXIDE 57

(2) OPERATION. Since this reaction is highly temperature depen


dent, the instruments used to record and control temperature must be
checked frequently to ensure their accuracy. Excessively high indi
vidual particle temperature can cause the particle to become sintered
its

or
fused with other particles. The possibility

of
on surface this
occurring becomes readily apparent when one considers that the high
dissipated during the reaction time. The

be
of

heat reaction must


reactive UO, naturally causes higher rate
of

extremely

of
use an heat

a
evolution, and correspondingly higher individual particle tempera
a

ture may result. Conversely, insufficient heat

in
the reactor can cause
incomplete conversion UO. Ordinarily, the reactors are operated
to

on
control temperature all four zones; however, because
of

650°
of at

of C
a

the reactivity the UO, feed, this may

be
varied
in

differences
widely. the reactor bed temperature very difficult, since
of

Control

is
the thermocouples are all located outside the reactor tube. Tempera
high
as

have been recorded when thermocouples were


as

tures 1,000°
C

placed on the rotating shaft the agitator near the point


of

of
UOs
entry into the reactor. also interesting in

in
to
It

note that one


is

high were reached when operating


as

stance temperatures
as

1,150°
C

with highly reactive UO, even though the reactor had been controlled
by

This high metallurgical


of C.

temperature
at

650° was determined


a

analysis pieces the agitator.


of

Uranium dioxide produced the continuous reduction reactor for


in

UF, nominal operating


at

produced
be
C. to

immediate conversion can


a

temperature However, UO, produced for storage and


of

650°
This
C.
at

subsequent use must necessary because


be

made 815°
is

UO, produced below this temperature tends reoxidize upon ex


to

posure air, even The production rate


to to

temperature.
at

room
is
by

approximately 500 pounds per hour the capacity


of

limited the
cooling screw. powder
of

in

Retention time the reduction reactor


hours, with bed depth speed
of

inches. Variation
to

to

in
of is

11%
2

2
a

the agitator has only slight effect


on
of

rotation retention time


is a

However, the re
of

in

the manner which constructed.


it

because
by

tention time can easily more pushers.


be

of

decreased the addition


common practice use about 1.5 times the stoichiometric
It

to
is

of

dissociated ammonia required for the reaction order


in

to

amount
provide safety factor against inaccuracies the dissociated am
in
a

UO,
or

against surge entering the reduction


of

monia flow rate


a

reactor.
Occasionally, difficulty the dust collection system,
in

encountered
is

by

buildup
or

back pressure
of

created material the reactor


in
is
a

In

the exhaust gas exit line.


of

these cases, the flow


or

tube dust
in

automatically shut off, and nitrogen


of

cracked ammonia auto


is

is

matically flushed into the reactor. Recording instruments and alarms


indicate such stoppages
of

the cracked ammonia flow.


58 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

(d) Vibrating Tray Process

Basically, the vibrating tray process involves the use of an elongated.


enclosed tray positioned in a furnace and so equipped that it can be
vibrated periodically. Vibration of the tray imparts motion to

the
particles the tray and moves them slowly from one end

on

to
the other.
Coincidentally, either hydrogen gas

or
dissociated ammonia passed

is
through the tray

of
the UOs

to

in to
cause the conversion UO2.

In
vestigation this type
of early 1948

of
reactor was first undertaken
by

Union Carbide Nuclear Company the Oak Ridge Gaseous Diffu.

at
an

sion Plant develop continuous process for converting

to
effort
in

a
UOs UF, [27]. The method proved successful, and full scale
to

a
production plant currently operation Oak Ridge [28].

at
in
is

(1) EquipMENT. drawing vibrating tray system

of
A

is
shown

to a
Fig. 2.14. The unit similar commercial, mechanically oscil.
in

is

lated conveyors. However, because the vibration carried out

is

at
high temperature, considerable strengthening required.

is
The UO, received 5-ton hoppers which are hoisted and attached
in
of is

top hoppers. (A system also provided for han


to

the the feed

is
dling material 30-gallon drums; automatic unloading and drum
in

washing and drying facilities are incorporated into the system.) The
UO, fed from the storage hopper through seal hopper by means
is

a
of

screw conveyor and into the reactor. The reactor 15-foot

or is
a

a
long, 2-foot wide trough (made type 34.
of

12-inch thick Inconel


stainless steel) with 6-inch high roof which slightly arched pro

to
is
a

vide additional strength. The tray supported heavy 4-inch diam


on
is
or

eter pipes type 347 stainless steel which extend through


of

Inconel
the furnace wall the outside, where they are welded
in

to

slots
to
are a
heavy pipe frame
on

the furnace. The pipe frames


of

each side
springs
to

as

fastened rocker arms and torsion bars (these elements act


of

the load from the drive unit) which are connected


to to

remove most
on

heavy concrete inertia block. The block supported heavy


is
a

springs separate prevent transmis


to to

to

order from the floor and


in

it

by

the building. The tray


of

of

vibration
an

sion vibrated means


is
by

eccentric shaft (driven variable speed 10-hp motor) connected


to
a

the framework. The angle which the tray vibrated controlled


at

is

is
by

adjusting the angle


It

which the rocker arms are set.


at

neces.
be is
all

pipework, such
as

sary that gas and powder connections, made


through flexible bellows assemblies.
In

spite
elaborate precautions some vibration
of

transmitted
is

to

the electric furnace which surrounds the tray. This necessitates the
heavy 14-inch diameter heating elements. To improve tempera
of

use
ture control the reaction bed, the furnace divided into three heat
in

is

can

ing zones, each containing top and bottom heating element that
a

controlled independently.
be
GAs Porous
ouTLET CARBON

I
FILTER
GAS

|
suPPORT ARMs PIPE FRAME
T
Powder
INLET ARMs OVEN

|
wn
VARIABLE-SPEED DRIVE

EccENTRIC

*
POWDER
- OUTLET

…] \º. HEAD - ADJUSTING CLAMP

TORSION BARs
2

A
-
-

. P.
E.

FIGURE 2,14. Typical Vibrating Tray UO, Reactor [8]. (From Charles D. Harrington and Archie Ruehle,

D.
“Uranium Production Technology,” Van Nostrand Co., Inc., Princeton, N.J., 1959.)
§
60 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

In the Oak Ridge process most of the hydrogen is supplied as a by


product of the fluorine generators used for UF, production. To sup
plement this and to allow uninterrupted operation when the fluorine
cells are not being used, a 15-cell electrolytic hydrogen generator is
provided which can deliver 4.6-pound moles per hour. A pump re
cycles the hydrogen through a trombone pipe coil (heated by an elec
tric furnace) and into the reactor. Water produced by the reduction
reaction is condensed by recycling the exit gases through a water
jacketed pipe heat exchanger, the water being eliminated with the ex
cess water supplied for the pump. The concentration of hydrogen in
the recycle stream is checked by a continuous gas analyzer. When the
concentration becomes too low, the spent gas is automatically vented
from the system, and pure hydrogen is automatically admitted. If
dissociated ammonia is used, the exit gases are burned. Appropriate
safety precautions, such as those described in previous sections, are
observed.
The UO, product is discharged into a 1,500-pound capacity seal hop
per placed on scales. A screw feeder then discharges the powder for
packaging.
(2) OPERATION. Control of temperature in each zone of the reactor
is achieved with a thermocouple located at the center of the zone, just
below and to the outside of the middle of the tray. Another indicating
thermocouple is located at the center of each heating zone, just above
the top zone of the tray. With this arrangement, it is difficult to as
certain the exact relationship between reaction bed temperature and
control temperature; but it has been determined that the bed depth
does have a pronounced effect on bed temperature. As shown in Fig.
2.15, increased bed depths cause increased bed temperatures for a
given tray (control) temperature. A control temperature of 815° C
is normally used. Essentially, no reaction occurs until the bed tem
perature reaches 540° C. Once this temperature is reached, the reac
tion proceeds rapidly, with a sharp rise in bed temperature.
The powder velocity through the reactor can be varied to give re
tention times from 1% to 11% hours (dependent upon the type of UO,
being processed). This is done by varying the frequency and time of
oscillation. Normally, a frequency of 750 cycles per minute is applied
for 17 seconds once every 14 minutes. The reactor assembly is de
signed to operate at the resonant frequency of 750 cycles per minute
in order to decrease wear on the tray supports and the drive mech
anism. An increase in the drive stroke above /s inch can cause pro
nounced wear on the flexible connections. Studies have shown that a
14-inch stroke at 750 cycles per minute will move the powder at about
0.13 foot per second. With a given set of vibrating conditions, the
bed depth is directly proportional to the feed rate. Therefore, the
retention time is not affected by changes in feed rate.
PREPARATION OF URANIUM DIOXIDE 61

I I i I I

i8OOH O TRAY TEMP. - Iloo" F


2.
O TRAY TEMP.s 1200°F
...”
17OOH- 26 -
e
29
I6OO 2^
* 2^
-
2^
u
15OOH-
O
-
# _2^
:
a 1400 H.
O
92° -
> O 2^ O
* 2^
• 'O O
-
1300 H.
Q.”
O Ole
O
• 9.6
1200 |- —
2^ O
*
| | OO l I l l |
o O.5 I 1.5 2 2.5 3

BED DEPTH IN IN CHES

FIGURE 2.15. Bed Temperature as Influenced by Bed Depth during Vibrating


Tray Reduction of UO, [8]. (From Charles D. Harrington and Archie E.
Ruehle, “Uranium Production Technology,” D. Van Nostrand Co., Inc., Prince
ton, N.J., 1959.)

(e) Fluid-Bed Process

Successful production of UF, from UO, utilizing fluid-bed reactors


has been experimentally demonstrated by Union Carbide Nuclear Co.
at Oak Ridge and by Mallinckrodt Chemical Works at St. Louis [29].
These sites have utilized equipment in which the UO, is reduced to
UO2 in one fluid-bed reactor which, in turn, is connected to a second
fluid-bed reactor which hydrofluorinates the UO, to UF. Since the
fluid-bed process has proved successful in this application, it appears
logical to assume that it may prove applicable to the production of
UO2 for fuel elements. As yet, this has not been tried, but it is antici
pated that it will be investigated by Mallinckrodt in due course. It,
therefore, seems appropriate to include a brief description of facili
ties currently available.
A diagram of the fluid-bed reactor in use at St. Louis is shown in
Fig. 2.16. The unit consists essentially of two vertical type 347
stainless-steel cylinders (14-inch ID by 72 inches high) having cone
bottoms. The two cylinders are located in close proximity and are
connected by a 3-inch diameter feed overflow line and 1-inch diameter
gas pressure equalizer line. A sintered stainless-steel plate across
GASES TO BURNER GAS
AND DUST COLLECTOR EQUALIZER LINE DUST FROM uos TRANSFER HOPPER
DUST COLLECTOR

E-3 SCALE

d
AIR To VIBRATOR

º c
Lu u FEEDER
an co scREW
ou -

c
ScREW FEEDER FEED HOPPER
2 z
NG

~~~ FEED OVERFLOW LINE

NITROGEN
_
GAS INLET

PRODUCT OVERFLOW LINE

SEAL HoPPER

U02 PRODUCT

FIGURE 2,16. Fluidized-Bed Reactor for Producing UO.


PREPARATION OF URANIUM DIOXIDE 63

the bottom of the reactors serves to support the UO, and also ensures
adequate distribution of the fluidizing gas across the entire cross sec
tion of the tube.

Heat is required to initiate the reaction but, since the reaction is


highly exothermic, cooling is required once the reaction has started.
Each reactoris heated externally with 12 Calrod electric heaters
which are so connected that six of the heaters can be manually con
trolled while the other six are instrument controlled. The units may
also be cooled by means of a low-pressure steam coil welded to the
outside of the shell. The reactors are enclosed in a removable, double
walled insulating can, and a 2-inch thickness of Kaylo block insula
tion is placed over the outside.
A 3-inch diameter screw feeder with a variable speed drive is used
to feed the UO, out of the transfer hopper and into the No. 1 bed at
a point 14 inches above the porous plate. Material from No. 1 bed
overflows through the 3-inch line into the No. 2 bed, and the product
from No. 2 bed overflows into a separator hopper and, thence, to a
seal hopper.
The cracked ammonia is metered through rotameters. Sufficient gas
flow to ensure complete fluidization is assured by interlocking the UO,
screw feeder with the cracked ammonia flow controller. The off-gas
passes through a cyclone to remove any UO, dust which is recycled
to the UO, separator hopper.
The cleaned hydrogen gas is burned and
then exhausted to the atmosphere through a bag-type dust collector.
Temperature control is provided in the burned gas line to ensure that
the bags are not burned.
The No. 1 within the range of 605° to 665°C
bed unit is maintained
by means a recording-controlling thermocouple extending through
of
the reactor wall to the center of the bed above the point where the
UO, feed is introduced and approximately at the half-height of the
fluidized section. Two additional thermocouples, one located near the
top and the other near the bottom of the fluidized bed, provide the
means of appraising the degree of fluidization that exists. Since
active agitation should promote rapid heat transfer and, thus, tem
perature uniformity within the bed, a minimum differential between
these thermocouple readings is interpreted as reflecting a highly
fluidized condition. To guard against possible equipment damage in
case of a high-temperature surge from the exothermic reaction, a
thermocouple in contact with the outer skin of the reactor wall
operates a power cut-off to function at a preselected temperature limit.
The central control thermocouple governs both the Calrod heaters
and the steam cooling coils. Relatively smooth operation has been
obtained with an overlapping control. As heat is generated by the
exothermic reaction and the input from the Calrod heaters is being
64 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

decreased, the overlap cuts in the delivery of steam to the cooling


coils at a temperature 11° C below the final cut-out temperature for
the Calrods. This type of operation provides for approximately 95
percent conversion of UO, to UO, before transfer of the powder to the
No. 2 bed.
The No. 2 bed unit normally operates between 510° and 580° C,
since the heat liberated from the remaining 5 percent of the conversion,
plus that supplied by the Calrods, is insufficient to maintain the tem
perature setting of 620° C. There has, however, been no evidence of
an adverse effect on product quality.
Sustained operation of these reactors has been experienced for over
a year. Since the UO, feed to the reactor has a wide range of particle
size and density, it has been necessary to vary the fluidization veloci
ties from 0.8 fps to 1.2 fps. Operating rates of 900 pounds of UO.
per hour with 98 percent conversion to UO, has been the normal
experience.

2.4 THERMAL DECOMPOSITION AND REDUCTION OF AMMO


NIUM DIURANATE
G. W. Tompkin

2.4.1 Introduction

The use of enriched UO, in nuclear power reactors in the United


States has stimulatedtonnage production of uranium dioxide, using
uranium hexafluoride from the gaseous diffusion process as the starting
compound. Uranium dioxide, which can be processed without further
activating treatment into high-density UO, sintered shapes, can be
prepared through the ammonium diuranate flow scheme. The two
major steps of the process are (1) hydrolysis of UF, and precipitation
of ammonium diuranate (ADU) and (2) thermal decomposition and
reduction of ammonium diuranate to yield uranium dioxide. Uranium
dioxide with similar ceramic properties can also be prepared by the
thermal decomposition and reduction of ADU precipitated under cer
tain conditions, starting with uranyl nitrate solutions rather than
uranium hexafluoride.

2.4.2 Hydrolysis of UF, and Precipitation of ADU

Uranium hexafluoride, normally handled in metal cylinders, is


heated above the vaporizing temperature of 62°C and metered into the
reaction vessel with ammonia and water (Fig. 2.17). Various types
of scrubbers or contactors can be used to mix the ammoniacal liquor
with the incoming UF, to obtain complete hydrolysis. Type 304L
stainless steel is satisfactory for the reaction vessel if ammonia is pres
it in excess of the required to react with the hydrofluoric
almount
65

|
PREPARATION OF URANIUM DIOXIDE

PUMP
vENT STEAM NH3

H2O
| J. FILTER
NH3
UFe -
- rºoºoºss
Er
SLURRY ADU DRYING

|*
AT r REDUCTION
CAKE | 175°C 800°C
-
2UF6º – 14NH33.+ 7 H2O=
2 (NH4)2 U207 “2 H2
2-2 2=
DISCARD
(NH4)2U207 - 12NH4F 2UO2 + 2NH3 + 3H2O

PRODUCT K HAMMER
MILL

FIGURE 217. Flow Sheet for UO, Production from UF, by Ammonium Diuranate
Route [3].

acid formed by hydrolysis. The hydrolysis and precipitation reac


tions are shown separately below:
UFe+2H2O->UO.F.--4HF Eq. (2.2)
2UO.F,4-2NH,OH+H.O— (NH,), U.O., +4HF Eq. (2.3)
HF--NH.OH->NHLF--H2O Eq. (2.4)
Anhydrous UFe, by itself or diluted with nitrogen, can be handled
in steam-traced lines of copper, mild steel, or stainless steel without
attack on the lines or contamination of the product. Glass rotameters,
heated to prevent the condensation of UFe, can be used for metering.
Leaks which may develop during operation are quickly detected by
the characteristic white cloud of HF and UO.F, which forms upon
contact of UFs with moist air.
The ADU crystal is the precursor of the general structure of the
final UO2. Since a semiamorphous uranium dioxide is to be prepared,
it is desirable to precipitate fine, highly imperfect ADU crystals
(see Chap. 3). Such an ADU crystal is prepared by the rapid precipi
tation of ADU under conditions of low uranium solubility. In the
system under discussion, however, the strong solubilizing effect of the
fluoride ion which builds up as the reaction progresses must be taken
into account. If precipitation is to be carried out under conditions
of controlled solubility of uranium in such a manner as to achieve
high ceramic activity “ in the uranium dioxide and at the same time
realize some degree of filterability of the ADU slurry, careful pro
gramming of the antisolubility ions, ammonium and fluoride, must be
maintained. If the excess NHA present is too great relative to the
fluoride concentration, a slurry which is difficult to precipitate will

* The term “ceramic activity” has come into general use in referring to a UO2 preparation
which sinters readily to high density. Such material is then called an active oxide.
66 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

-
I I i I I I I I T I I I I I
GRAMS FLUORIDE / LITER
--- 37-39
18-25
-1
ź * * * * ** 10- 12 -
u

<
z -
<
or
P
co -
-E

-co
E
-
-:
>
<.
u.
-
o
>
!- N. -
– *~
g
S -T - - -
a
vo *--
l | l l | I
50 60 70

NH3 /LITER
GRAMS OF ExCESS

FIGURE 2.18. Solubility of Ammonium Diuranate as Influenced by Excess Am


monia Employed during the Hydrolysis of UFe.

result. If
the excess ammonia concentration is too low, a coarse,
easily filtered ADU will have been prepared, but subsequent processing
will yield a uranium dioxide product which has poor ceramic activity.
It natural to choose a course in which the error lies on the
seems
high ammonia side. However, this alternative tends to produce pyro
phoric uranium dioxide after thermal decomposition and reduction.
This tendency toward pyrophoricity is discussed in Sect. 2.4.3.
The rapidly changing condition of uranium solubility during the
hydrolysis and precipitation of UF is indicated in Fig. 2.18, which
relates the amount of solubilized uranium to the excess ammonia
concentration for three fluoride concentrations covering practical con
ditions for preparation of ceramically active UO, by hydrolysis of
UFe into ammoniacal ADU slurries.

2.4.3 Decomposition and Reduction of Ammonium Diuranate

The conversion of ADU to black uranium oxide (UAOs) and sub


sequent reduction to uranium dioxide can be done as two independent
steps or can be combined into a one-step process involving the de
composition and reduction of ADU in the presence of hydrogen in a
retort. The latter process is the one most commonly used. The con
ditions for decomposition of ADU and reduction to UO, are dictated
PREPARATION OF URANIUM DIOXIDE 67

by the requirement that the highest reactivity possible be


ceramic
retained in the UO, product. The conversion of ADU to UO, is
further complicated by occluded fluoride in the range of 1 to 4 per
cent in the ADU.
Ammonium diuranate containing occluded fluoride can be reduced
in closed retorts of Inconel, Hastelloy C, or Hastelloy X at tem
peratures of 760° to 870° C. Steam is passed over the reacting bed
during the thermal decomposition to pyrohydrolyze and remove
fluoride, which is carried from the retort as HF and ammonium
fluoride. The time required for fluoride removal in this temperature
range is strongly dependent on steam flow and bed depth. The treat
ment can take from 20 minutes to a few hours. Reduction, which in
part is being done concurrently with decomposition and pyrohydroly
sis, is continued after the steam treatment has been stopped. Bed
depth in unstirred trays is ordinarily limited to an inch or two.
Deeper beds of ADU tend to result in the preparation of less active
UO2.
A tendency of UO, to air oxidize at room temperature is associated
with uranium dioxide which is ceramically active. This characteristic
dictates that the retort and UO, be thoroughly cooled before opening.
Pyrophoric UO, can be prepared by 760°C pyrohydrolysis and re
duction of ADU prepared by rapid precipitation under conditions
of low uranium solubility.
Short of being pyrophoric, ceramically active UO2 will, upon expo
sure to air, partially oxidize by the absorption of oxygen. The cubic
lattice is retained, however, in contrast to the case of pyrophoric oxide
which actually burns in air to form orthorhombic U.Os. Uranium
dioxide, which upon exposure to 25° C air oxidizes to the extent
indicated by the formula UO, will ordinarily
os,

os. be

more active ceram


ically than UO,
in to an

oxide which air oxidizes Both samples are


to

UO, less during


or

presumed
of
to

have been reduced the extent


or

production the retort.

2.4.4 Important Process Variables

The most important process variables are associated with the pre
cipitation ADU from the uranyl fluoride
or
of

solution from the


hydrolysis precipitation
of

and uranium hexafluoride. The most


the production uranium dioxide, starting with
of

notable variable
in

DU precipitated from fluoride solution, the extremely high


A

is
a

complexing and solubilizing action the fluoride ion present


of

the
in

ammonium diuranate slurry. One practical concentration which


at

precipitate ADU fluoride per liter. At this


to

grams
25

of

about
is

grams per
an

level,
of

fluoride
to
15

excess ammonia concentration


5

57.4789 O-61–6
68 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

liter will result in the preparation of ADU from which ceramically


active UO, can be prepared.
The preparation of ADU from
solutions containing higher fluoride
ion concentrations will
require greater amounts of excess ammonia to
precipitate an active type ADU.
Active UO, can also be prepared from uranyl nitrate solutions by
proceeding through the ammonium diuranate flow scheme. Because
of the absence of fluoride in the system, corrosion problems are less
severe. The weaker complexing tendency of the nitrate ion contrasted
with the strong solubilizing effect of fluoride ion permits the use of
markedly higher uranium concentrations in the starting solutions.
The conversion of ADU to UO2 can be done under conditions
similar to those described above, except that stainless-steel reactors
can be used. A particularly active ADU can be prepared by rapidly
precipitating uranyl nitrate solutions containing about 100 grams of
uranium per liter with concentrated ammonium hydroxide at a pH
of 6.7.

2.5 EXPERIMENTAL METHODS OF UO, PREPARATION


J. C. Clayton
2.5.1 Water and Steam Oxidation of Uranium

Finely divideduranium metal will react with boiling water to yield


a tetravalent state that has been variously interpreted as either UO,
or uranous hydroxide [30]. The reaction represents a fire and ex
plosion hazard, since the evolution of hydrogen is sufficiently turbu
lent to carry metal fines to the bath surface where vigorous combustion
of both uranium and hydrogen results. Aeration of the slurry achieves
further oxidation, producing a hexavalent compound that approaches
UOs. Boiling water will also oxidize massive uranium metal (but
at a slower rate) with a similar uncertainty (in the light of present
knowledge) as to the exact nature of the end product [31].
Several methods of preparing UO, by steam oxidation of uranium
metal and uranium hydride have been reported [31–34]. Uranium
metal foil is readily oxidized with high-pressure steam (2,200 psig,
343° C, 6 to 72 hr.) to produce a slightly oxygen-deficient UO,
(UO, sº—UO, ao)." Attempts to detect the presence of unreacted
uranium or UH, have been unsuccessful. The as-prepared oxide tends
to retain the physical shape of the original uranium metal foil, but
readily crumbles into a fine, dull-black powder.
Oxidation of uranium with flowing steam at pressures between 3.5
and 7.5 psig and temperatures between 150° and 450° C produces
*
* See Chap. 6 for a of stoichiometry.
discussion
PREPARATION OF URANIUM DIOXIDE 69

a powder having an O/U ratio greater than 2.00 (2.01 to 2.20). It


has been found that the heat of oxidation is so great that under certain
conditions the reaction runs away, and the temperature of the uranium
rises to as much as 250° C above the steam temperature. The larger
the sample size, the lower the steam temperature needed to cause this
action. The oxide is similar to the high-pressure material in appear
ance. When the oxidation rate is controlled by limiting the steam
flow rate and temperature, a highly lustrous crystalline powder is
obtained.

Uranium metal readily reacts with flowing superheated steam at one


atmosphere to form UO, U.O.s, or some intermediate mixture, de
pending upon the temperature. Wathen reports that up to 225°C
theproduct is mainly UO, ; at higher temperatures, the product is
mainly UAOs [35].
Uranium metal reacts with hydrogen at 150° to 250° C to form
uranium hydride. The UHA is decomposed by heating in vacuo or in
an inert atmosphere at 300° to 450° C, and the resulting uranium
powder is steam oxidized to UO, [32, 34].

2.5.2 Precipitation Methods

(a) Introduction

A variety of precipitation procedures for preparing UO, has been


studied both as laboratory and large-scale plant processes. The neces
sity of controlling such factors as pH, concentration, temperature,
rate of stirring, digestion time, and even the geometry of the reaction
vessels increases the relative costs of the precipitation methods. In
addition, many of the precipitates have low settling rates, are diffi
cult to filter, and must be thoroughly washed. However, in spite of
these handicaps, certain precipitation processes appear to be econom
ically feasible. Many of these methods yield a UO, with a high
density, small particle size, and excellent sinterability. Moreover,
any large-scale method of producing UO, enriched in must be U*
made from a UF, Uranium hexafluoride can be hydrolyzed
source.
to give a convenient starting solution for many precipitation reactions.

th) / O, Prepared from Uranium Peror/de

The production of UO2 from uranium peroxide is done on a pilot


plant and commercial scale in Europe |36–43]. Several patents have
been issued for the preparation of UO, -2H2O, and factors affecting
precipitation have been studied The composi
its

detail [44–50].
of in
of

tion this compound subject some debate the literature.


in
is
a

Freshly precipitated air-dried uranium peroxide has been reported


of

water per atom


of

contain from two over four molecules


to

to
70 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

uranium [31, 51, 52]. However, Leininger, et al., found that widely
varied conditions of precipitation and drying have no appreciable
effect on the composition of UO, .2H2O [53]. They attributed the
results of other workers to insufficient drying. Unless the uranium
peroxide powders are pulverized and spread in thin layers, only the
surface dries, leaving a moist center. There seems to be agreement,
however, that the precipitation of uranium peroxide from acidic
uranyl solutions by hydrogen peroxide addition is a quantitative reac
tion. Under proper conditions the UO,-2H2O formed will settle
rapidly, filter well, and dry to a friable powder. It is concluded that
a hydrated peroxyuranic acid structure (H.UOs: H2O) best represents
a compound of the stoichiometry UO,-2H2O [54]. This structure is
supported by infrared absorption data and reactions with a strong base
of the uranium peroxide compound.
Precipitation is nearly quantitative over the pH range 1.0 to 3.5
[38, 40, 47–50, 55]. Below a pH of one the precipitate becomes soluble
and of uncertain composition. The recommended optimum pH value
is around 2.5 [47–50, 56]. Since NH,OH is often used to adjust the
pH, considerable uranium is held in solution as an ammonium per
uranate complex when the pH is above 3.5. The concentration of the
solutions used is not particularly critical [40, 46, 47, 55]. Quantita
tiveness is assured with uranium solutions between 0.1 and 0.5 molar or
higher; however, potentiometric titrations in more dilute solutions
indicate that the mechanism of the reaction is dependent on the hydro
gen ion and uranium concentrations [48]. The volume of the starting
solution and the rate of H2O, addition have no effect on the reaction if
there is adequate mixing. The amount of H2O, used is not important
as long as it is in excess. Reversing the order of addition by adding
a uranyl salt to an H2O, solution still gives a precipitate of composi
tion UO, 2H2O, although of a different crystalline form [56]. The
physical appearance of the precipitate appears to be a function of the
anions present [53]. The reaction itself is adversely affected by a
number of ions whose effects must be minimized for a successful pre
cipitation [39,49, 50, 55].
No optimum value was found for the reaction temperature in the
range from 0° to 60°C [47]. Vigorous stirring increases the settling
rate, reduces the filtration time, and eliminates the need for washing
the UO,-2H2O [53]. Some digestion of the precipitate is necessary
[47]. The drying conditions affect the composition only in extreme
cases, for instance, when large lumps are dried for short intervals or
when the temperature is high enough to cause thermal decomposi
tion [53].
Data on the thermal behavior of UO, 2H2O are not in agreement
[31, 53, 57, 58]. The exact temperatures at which uranium peroxide de
PREPARATION OF URANIUM DIOXIDE 71

composes into UO, and UAOs are in dispute. This may be due to
the factthat the initial rate of thermal decomposition is affected by the
precipitation conditions [53]. In addition, because of the slow rate of
conversion, reproducible values for compositions between UO,-2H2O
and UO, can be obtained on heating for short intervals, but these
values do not correspond to equilibrium conditions [53]. Uranium
peroxide can be converted into UO, either by heating in hydrogen or
by first pyrolyzing the precipitate in air to UO, or U.O, and then re
ducing with hydrogen, carbon monoxide, ammonia, cracked ammonia,
or ethyl alcohol [16, 31, 34, 37,43,46, 59–61]. When the UO,-2H.O is
calcined at 200° to 400°C, UO, is obtained; at higher temperatures,
decomposition of the UO, to U.O,
The temperatures of reduc
occurs.
tion to UO, vary C. Uranium dioxide powders
from 500° to 1,000°
prepared by reduction at the lower temperatures are pyrophoric.
They can be stabilized, however, by using longer reduction times or
higher reduction temperatures, or both, or by maintaining the UO,
under an inert atmosphere.

I'(), Prepared from Ammonium


(e)

Džuranate

Enriched UO, manufactured commercially from ammonium an


is

UF,
by

prepared hydrolysis
of
diuranate intermediate the with

a
dilute ammonia solution (see Sect. 2.4.2). However, other precipitants
urea, ammonium carbonate, gaseous ammonia, and hydrazine
as

such
appear feasible [15, 34, 43, 62–68]. Precipitation from urea and am
monium carbonate solution, for example, yields larger
of

material
a

crystallite and particle size and smaller specific surface which sig
is

by by

nificantly more crystalline and free filtering than that obtained


other reagents [34, 63, 65, 66]. Ammonium diuranate prepared
precipitation with gaseous ammonia and ammonium hydroxide usu
is

ally poorly crystallized powder with low density and large sur
a

face area.

claimed that stoichiometric ammonium diuranate, (NHA). U.O.,


It
is

precipitated and isolated from


be

an

cannot aqueous system [48, 63,


uranium and excess NH,OH concen
67,

70]. Variations, such


as

69,
trations, order reagents, pH, and type anion, re
of

of

of

addition
amorphous precipitate with
an

an

portedly result consistently


in

NH, [63]. However, Tridot, using con


of

mole ratio
U

to

about 0.5
1
:

ductometric and potentiometric titration techniques, found indications


(NH,).U.O,
[71]. Also, true (NH,).U.O., formed
of of
by of

the existence
the reaction the uranyl nitrate hydrates with liquid ammonia,
X-ray diffraction pattern the compound precipitated
as

had the same


from aqueous uranyl nitrate solution with NH, OH [72]. Since the
composition uncertain, the precipitate will
be

in
to

referred
is

exact
this discussion as ammonium uranate.
72 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Many of the factors affecting the precipitation of ammonium uranate


have been investigated [15, 34, 62, 63, 68, 73, 75]. Precipitation is
complete before the pH of the liquor is seven, but a higher pH is rec
ommended for batch processes [62, 68,74]. Data on the concentration
of the solutions used are not in complete Lister and agreement.
Gillies found that the concentration of uranium (6 to 130 g U/l has
little effect on the character of the precipitate produced [15]. How
ever, they reported that in other British studies the efficiency of
precipitation varied with the initial uranium concentration. Watson
concluded that in the batch precipitation of ammonium uranate with
NH,OH, low uranium concentration (below 300 g/l, high or excess
NH, OH concentration, increased temperature, and rapid precipitation
all contribute to the development of a fine precipitate of small particle
size [62, 68, 74]. When ammonium uranate is precipitated continu
ously with NH,OH, only control of pH is necessary, since as the pH
of precipitation increases the filtration rate decreases rapidly. Mod
erate stirring and some aging of the ammonium uranate results in a
more uniform crystalline product [15, 63]. The settling rate of the
precipitate is most rapid around 60°C, but decreases markedly as the
rate of addition of the ammonium hydroxide is increased [15]. The
drying conditions are not critical [15,62]. However, when the uranyl
salt solution is added to the NH,OH, the precipitate formed is gelat
inous, difficult to dry, and, when dry, is very hard and has a high
bulk density [15]. When ammonium uranate is precipitated with
properties are essentially independent precipi
its

gaseous ammonia,
physical properties of

of
In

tation conditions [62, 68]. general, the


ammonium uranate vary drastically with the precipitant used [15, 34,
43,62–67].
pyrolyzed UO, and U3Os. The exact
be

Ammonium uranate can


to

by

temperature which decomposition occurs the precipi


at

affected
is

tant and precipitation conditions [31, 58, 75]. The UO, and UAOs
so

UO, with hydrogen


or
be

carbon monoxide
to

obtained can reduced


directly
be

[16, 31, 32, 34, 43, 59–61, 64, 65]. Ammonium uranate can
by

or

converted into UO2 heating hydrogen carbon monoxide


in

[16, 32–34, 62, 66, 68, 73, 74, 76, 77]. The maximum temperature
of

Uranium di
C.

reduction has been varied between 500° and 1,000°


by

oxide powders prepared the lower temperatures tend


at

reduction
pyrophoric.The depth sample
be

of

the
in
to

the containers used


in

important 77]. The tem


an

operation [62,
be

reduction can variable


perature ignition the higher uranium
of

of

to

the ammonium uranate


critical factor [15, 16, 34]. Increasing
be
to

oxide has been found


a

temperature decreased the specific


to

the calcination from 500° 800°


C

by

both the higher parent oxide and the resulting UO,


of

surface
a
of

factor ten several cases.


in
PREPARATION OF URANIUM DIOXIDE 73

A method has been developed in which ammonium uranate is directly


converted into UO, [78]. Reduction is achieved by the hydrogen
produced by the dissociation of the ammonia released by pyrolysis of
the ammonium uranate to UAOs. The temperature of reaction must
be above450° C for appreciable ammonia dissociation and UAO,
reduction to occur. The optimum reduction temperature was found to
be about 700° C.

A hydrometallurgical reduction technique has been devised for pro


ducing UO2 from crude ammonium uranate of unspecified particle
size, thereby eliminating rigid preliminary control of purity and
particle size of the starting material [79, 80]. Ammonium diuranate
is treated with carbon dioxide, hydrogen, a platinum catalyst, and
anthraquinone under pressure (900 to 1,000 psi) to form a uranyl
carbonate intermediate which is converted to UAOs. In the presence
of the catalyst, anthraquinone is reduced to hydranthraquinone which
reacts with U.O. to form U.O., and finally UO. Stopping the reduc
tion at intermediate stages gives uranium oxide of any O/U ratio
desired. By varying carbon dioxide pressure during reduction, the
surface area of the oxide product can be controlled. Uranium trioxide
can also be used as a source material for this process.

(d) CO, Prepared from Uranyl Oaxalates

Uranyl oxalate hydrates (UO.C.O., H2O, UO.C.O.3H2O,


UO.C.O., xH2O) can be conveniently prepared by the addition of
oxalic acid to a uranyl salt solution [34, 81–83]. The solid phase is
the trihydrate in HNO, of a strength up to 33 percent, while in more
concentrated acid the solid phase is anhydrous uranyl oxalate [81].
Some limited investigations have been carried out on the preparation
of UO2 from uranyl oxalate [31, 34, 82, 84, 85]. Uranyl oxalate can
be converted into UO, by straight thermal decomposition in vacuo,
by direct reduction with hydrogen or carbon monoxide, or by cal
cination to a higher oxide followed by reduction with hydrogen or
carbon monoxide. A distinctly acicular type of UO, can be obtained
by the latter two procedures [34].

2.5.3 Reduction of Higher Uranium Oxides

(a) (ranium. Dioſcide Prepared from U,0,

Since it is extremely difficult to prepare pure anhydrous UO, and


since the X-ray structure and reactivity of UO, are dependent upon
its

previous history, some studies have been made preparing UO.


on
by

from U.O. obtained either pyrolysis uranyl nitrate


of

the direct
hexahydrate via UO, [15, 16, 31, 34, 86–88].
In
or

these studies the


uranyl nitrate hexahydrate and UO, were ignited air
to

UAOs
in

at
74 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

800° to 850° C and then reduced to UO, at temperatures from 650°


to 900° C. The UO, powders made by these procedures have low
densities and large particle sizes.
Uranium dioxide itself has been air oxidized to UAOs and again

its
reduced with hydrogen and methane to improve sinterability and
reactivity toward hydrofluorination and chlorination [88–90]. Re
oxidation (400° C) and reduction (600° C) improve

of of
peated cycles
the properties the UO, increasing the specific surface and decreas
ing the bulk density. However, when the air oxidation temperature
powder sintering occurs, and the resulting UO2
to

raised 800°
is

is
C
unreactive, with low density and small specific surface [16, 34, 91].
a

a
by
be

of
Uranium dioxide can also made the reduction UAOs obtained
from uranium metal [32, 34]. Uranium chips, turnings,

be
foil can

or
directly oxidized hydrided UH, which then decom
or
to

to
UAOs

is
posed pyrophoric uranium powder. However, the UO2 powders
by to
a

particular advantages
no

In
made this method have properties.

in
addition, unless large supply scrap uranium available, the use
of

is
a
of

uranium metal not economically feasible.


is

(b) Uranium Dioacide Prepared from UO, Hydrates

has been found that hydrogen reduction of

of
the various forms
It

UO, hydrates results UO, with special properties [32, 34, 92].
in

Crystals UO, H2O may grown hydrothermally

an
of

be

in
autoclave
from slurry uranium peroxide-uranyl nitrate [32, 92].
of

at

250°

C
a

Several crystal forms may result, including bipyramids, rhombohedra,


and needles. The UO, hydrate crystals can reduced with hydrogen
be

UO, which maintains form as its


at

to

1,800° the same external


C

parent, but about two-thirds its size.


is

There are some indications that hydration UO,


of

certain types
of

yields UO, superior sinterability. When UO, water,


of

milled
in
is
a

UO, H.O which, when


be

percent the powder may


98

of

to

converted
UO2, will sinter very high bulk density [88]. The mill
to

to

reduced
a

the UO,
by

ing treatment itself, increasing the specific surface


of

may contribute improved sinterability.


to

the

2.5.4 Preparation from Fluorine Compounds

(a) Uranium Diowide Prepared from UF, and UF,


continuous, one-step fluid-bed process has been developed for the
A

UF, and UF, UO, [93]. Uranium tetrafluoride


of

to

conversion
is

pyrohydrolyzed with steam UO, both steam and hydrogen are used
to

for simultaneous pyrohydrolysis


UF. Reactant ex
of

and reduction
cesses of 300
to

600 percent
for steam and hydrogen and temperature
a

UFs UOs.
of

to

give greater
of

600° than percent conversion


C

95
PREPARATION OF URANIUM DIOXIDE 75

A dry process for converting enriched UF, to UO2 may offer eco
nomic advantages over the wet process now being used (see Sect. 2.4).
Besides being a direct method, the fluidized-bed technique offers good
temperature control, large surface area for reaction, and controlled
mobility of the solids.

(b) Uranium Diocide Prepared from UO, F,

Uranyl fluoride undergoes reduction with hydrogen to a very reac


tive pyrophoric type of UO, [23, 31, 94]. The reaction proceeds
rapidly at temperatures of 600° C and higher. Uranyl fluoride can
be prepared from a variety of starting materials, such as UF, , UF, ,
-
UO,(C.H.O.), and UO, [31, 94, 95].

2.6 HIGH TEMPERATURE PREPARATION OF UO,


W. H. Hedley and R. J. Roehrs
2.6.1 Flame Denitration and Reduction of Uranyl Nitrate
Liquors

(a) Introduction

Uranium trioxide powders produced by pot denitration have always


shown more variation in physical and chemical properties than is de
sirable. Moreover, pot operation presents problems in dust control,
undesirable from a health standpoint, and difficulties in maintenance.
To avoid these problems in operation and to achieve a more constant
product, a new and cheaper method of denitration has been sought.
What follows is the description of a technique which has reached
the pilot plant stage of development in which uranyl nitrate liquors
can be converted directly to uranium dioxide in a single treatment.
This process, called flame denitration, involves atomizing the liquor
in a high-temperature reducing atmosphere to create a UO2 powder
suitable for nuclear use.

(b) The Process

The equipment used to carry out flame denitration is shown dia


grammatically in Fig. 2.19. The feed system handles uranyl nitrate
liquor which has been boiled down to the desired concentration. This
liquor is pumped continuously through a pressurized, steam-traced,
recycle line to ensure against freezing or stoppage. The direct feed
line to the reactor is so constructed that either deionized water or
uranyl nitrate liquor may be forced through
it.

Deionized water
is

used during the preheat cycle


to

prevent localized overheating the


in
º
76 URANIUM Dioxide: PROPERTIES AND NUCLEAR APPLICATIONS
FILTER CYCLONE PRESSURE SAFETY
- Disc.

LEGEND

C. PROPANE AND AIR

OFF-GAS LINE- , NITROGEN


º PREHEATER
OFF-GAS– -
BURNER

|NSULATION
PRESSURE CONTROL VALVE

URANYL NITRATE
FEED TANK

E
NITROGEN

PACKAGING
STATION

FIGURE 2:19. Flame Denitration Flow Diagram.

reactor and at the conclusion of operation for flushing liquor from


the lines.
The cylindrical reactor shell is lined with silicon carbide to provide
good resistance to high temperatures, to corrosion, and to thermal
shock. This liner is backed by thermal insulation to minimize heat
losses. Heat is supplied by a special gas burner mounted at the top
of the vertically positioned chamber. This burner is designed to
operate with premixed propane and air and contains a spark plug
for ignition as well as a flame detector. The temperature within the
reactor is maintained at about 2,000°F.
PREPARATION OF URANIUM DIOXIDE 77

Uranyl nitrate liquor enters the reactor through a pneumatic atomiz


ing spray nozzle mounted at the center of the burner. Both air and
an air-propane mixture have been used as atomizing gases to yield a
fine droplet size that is denitrated in the first few inches of travel
after leaving the nozzle. The reduction to UO, takes place as the
particles travel further down the length of the reactor.
The product powder is conveyed by the entraining reaction gases
through a water-cooled conveyor line to a cyclone separator. A stain
less steel filter equipped for periodic blow-back of the porous tube
sections has by then effectively collected the gas-borne fines. The
powder is protected against contact with air until completely cooled,
first collecting it in a seal hopper and then discharging it by a rotary
valve through a water-cooled screw conveyor to a packaging station.
The off-gases after leaving the filter are mixed with air for two
reasons: (1) to burn the H2 and CO to eliminate explosion and health
hazards; (2) to oxidize the NO in the gases to NO2, making it possible
to recover the nitrogen oxides as nitric acid.

(c) Product Properties

Flame denitrated UO, has a particle size range from one-half to 10


microns and despite a moderately high surface area (2.6 m”/g) is
resistant to atmospheric oxidation at room temperature.
Chemically, the product contains approximately 0.03 percent water
and <0.1 percent nitrate with a carbon content ranging up to 300
ppm. Spectrographic analyses of other impurities in the product
have shown that they are present in essentially the same concentra
tion as in the feed liquor. Results of sinterability tests show that the
flame denitrated UO, compares favorably with comminuted MCW
oxide.
Oxide mixtures may readily be prepared by feeding blended nitrate
solutions to the atomizing spray head. In this fashion, UO2–ThC),
or UO2–Pu(), combinations could be processed without health or
criticality hazards by suitable design of reactor and conveyor line
dimensions.
Flame denitrated UO, is also under study as a source of uranium
metal. The powder exhibits a high reaction rate with HF which is
favorable for complete hydrofluorination.

2.6.2 Preparation of UO, Crystals by Flame Fusion

(a) Introduction

A flame fusion technique has been under study as a possible means


of preparing single crystals of UO, in a manner similar to that em
ployed for synthetic gems [97]. This procedure, known as the Ver
78 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

neuil process, has been successful in the preparation of large single


crystals of refractory materials, such as corundum (Al2O3) and spinel
(Mg,Al,O.), for use in instrument bearings, precision gauges, and as
gems for the jewelry trade [98–102]. The method employs a high
temperature flame (2,200° to 2,400° C) of oxyhydrogen through
which, for instance, pure aluminum oxide powder is dropped to melt
and crystallize on a seed crystal of the same material. The resultant
boule possesses the same crystal structure as the seed and contains no
flaws. Such a procedure applied to UO, offers the possibility of pro
ducing a crystal having theoretical density. The technique contem
plated could apply to crystal growth from a UO, powder feed or to
simultaneous reduction and crystal growth employing UO, powder as
the raw material.

(b) Gas Burner Equipment

Special gas burners utilizing fuel mixtures of hydrogen, oxygen,


acetylene, and methane, and capable of generating high flame tempera
tures were used during early experimental attempts to flame-fuse
UO2. In general, these were found to be inadequate for the very high
(approximately 2,760° C) melting point of UO, although one poly
crystalline boule (Fig. 2.20) was successfully grown. The boule was

FIGURE 220. Polycrystalline UO Boule; Approximately X4.5 [97].


PREPARATION OF URANIUM DIOXIDE 79
jet
color, measured about triangu

1%

its
inch along each side

of
black

in
point

at
tapered

on to
lar base, about inch above the base, and was

1
smoothly glazed Chemical and X-ray analyses yielded
one side.
the following composition: carbon 0.0036 weight percent, uranium
(IV) 87.90 weight percent, uranium (VI) 0.030 weight percent.

A
by
1%
inch square inch thick cut from the glazed surface

of
sample
14

10.96 g/cc

or
the boule yielded density 100 percent

of

of
theoretical
a

density for UO.


raise the flame temperature, lower the melting point
an
In

attempt
to

the resultant oxide mixture, and improve the oxidation resistance


of of

UO, ceramic element


an

the boule,

of
aluminum oxide modification

a
by
was attempted. This was prepared feeding powdered aluminum
through the flame with UO, powder. The outcome was moderately
that weight percent Al–96 weight percent UO, mix
in

successful
4
a

crystal growth

an
ture exhibited favorable properties and improved
UO2. All
as

compared pure

of
oxidation air with
in
to

resistance the
boules, however, exhibited some porosity their principal defect.
as

(c) Atomic Hydrogen Arc


an

More recent work has employed arc shielded with hydrogen

as
a
achieving the necessary temperature for fusion
A of

of
means the UO2.
Fig. 2.21 shows seed rod supported rotat
at

the top
of

Sketch of
a
a

ing water-cooled probe which positions the seed during the growth
operation. may pressed and sintered compact
be

The seed or
a

single crystal from previous flame fusion. Approximately one-half a


a

is B,

the top surface kept molten (Sketch Fig. 2.21) by


of

of

this seed
is

the hydrogen arc while fresh powder awaiting fusion being depos
UO, UO,
on

of

the other half. The feed powder car


or

either
is

ited
by

ried through tube entraining gases which exit above the seed rod
a

directly opposite the hottest zone the arc. As the probe rotated,
of

is

powder
on

deposit top
of
to

the feed continues the the seed and then


moves into position directly under the arc where melting takes place
C,

Fig. 2.21). The rotary action the probe also spreads the
of

(Sketch
powder evenly over the top
of

of

the seed. The surface tension the


molten liquid, with the sharp thermal gradient between the molten top
it,

and the solidified edge below prevent the liquid from flow
is to

tends
gradu
as

ing down the sides formed. The probe


to of

the seed rod


it

is

the growing UO,


as

ally lowered maintain the top


of
so

rod
in
a

equilibrium
an

fixed horizontal plane that


of

favorable for heat


is

As Fig. 2.21,
of

transfer. indicated Sketches and rod


in

is
D

a
by

melting and solidification


the progressive UO, within
of

formed
the gas envelope, which also accomplishes the reduction from UOs
when this type employed.
of

feed
is
80 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

PROBE

FIGURE 2,21. Schematic Representation of the Flame Fusion of a UO, Boule.

Figure 2.22 shows a general view of the present atomic hydrogen


arc equipment. The powder feed unit consists of a flanged gas-tight
chamber on which a Syntron vibratory feeder is mounted. The
vibrator permits a controlled rate of feed downward through a supply
tube which discharges directly above the seed rod of UO2.
its

The positioning of the seed on supporting water-cooled probe


of by in

plane
of

fixed horizontal the arc


at

the hottest zone achieved


is
a

continuously lowering the probe rate coinciding with the rate


at
a

by

crystal growth controlled twin screws


in

the flame. Probe travel


is
by

which are driven variable-speed motor and which support the


a

platform
on

which the probe mounted. This motion may


be

set
at
is

any desired rate within range 1.25 inches per minute.


of

to

0.001
a

Also available are lateral adjustments the probe and


of

of

the fusion
obtaining the desired symmetry crystal growth.
of
to

chamber assist
in

The quality
on
of

of

the boule likewise depends the speed rotation


the probe. The present equipment permits speeds from
of

to

120
5

rpm, but even faster rotation may be desirable.


PREPARATION OF URANIUM DIOXIDE 81

FIGURE 222. General View of Atomic Hydrogen Arc.


82 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

As the fused rod is lowered, it enters a cooling chamber which pro


tects the UO2 from oxidation as the temperature is lowered. Inert or
reducing gases flowing upward through the chamber serve as both
protective and cooling media. The actual reduction and fusion take
place within a refractory-lined chamber which is equipped with an
off-gas removal vent that minimizes such operational hazards as ex
posure to uranium oxide dust, ultraviolet radiation, and fumes.
With this equipment, a considerable number of boules of pure UO.
have been grown. The largest prepared to date with UO, pow
der feed is 746 inch in diameter by 6 inches long, weighing 138
grams (Fig. 2.23, A). The highest density recently achieved for a
boule fused in the hydrogen arc is 10.76 g/cc or 98 percent of the
theoretical density for UO2. The fastest growth rate yet attained
has been approximately 1 pound per hour.
X-ray analysis of boules grown from both UO, and UOs powders
shows that the composition of the product lies between UO2.00 and
UO, phases other than UO, are present. Figure 2.23,
no
os,

and that

B
INCHES
2

|
|

INCHES
2
|

FIGURE 2,23B. UO, Boule Made from UOs, Top View.

FIGURE 2,23C. UO2 Boule Made from UOs, Side View.


FIGURE 2:23. Boules of UO2.
PREPARATION OF URANIUM DIOXIDE 83

and C, shows top and side views of a UO, boule grown from UO,
powder. One of the columnar grains has been checked at several
points along its length for monocrystallinity by means of a Laue
X-ray diffraction pattern. Each region examined was approximately
1 millimeter in diameter, and each pattern obtained was that of a

is,
single crystal. It thus, conjectured that each columnar grain

of
the
boule needle-shaped single crystal.
is
a

2.7 SUMMARY

UO, suitable for fuel


In

nearly all preparation procedures, types

of
by

element use are prepared high temperature reduction hydrogen

in
higher oxide. The properties
of

the final oxide are sensitively


of
a

dependent upon the method preparation the starting material


of

of
and the particle size, crystal structure, surface area, etc., the higher

of
oxide. These characteristics are discussed Chap. in

3.
The bulk UO, powder produced for reactor use has been prepared
of

by

by
from uranyl nitrate hexahydrate pot denitration UO, and

to
hydrogen reduction UO, similar
in
to

continuous screw reactor.

A
a

procedure vibrating tray reactor has been largely employed for


in
a

UF, U.F. production, but could utilized for UO, preparation

as
be
to

UO, UO, fluid-bed reactors yield


of

well. More recent reductions


to

in

excellent material for processing U.F., but the suitability for prepa
is to

UO, for fuel element use still unexplored.


of

ration
Commercial production UO, from UF, employs hydrolysis and
of

precipitation ammonium diuranate, which then decomposed and


of

is

be
hydrogen. any
of

Uranium dioxide
in

reduced enrichment can


prepared, and the physical properties the final powder depend
of

directly
on

the conditions under which the ADU precipitated.


is

Rigorous control precipitation conditions may


be
of

eliminated
through the recently developed hydrometallurgical reduction
of

crude
The ADU intermediate may eliminated en
be

ammonium uranate.
tirely UF, UF, UO, proves
of

of

direct conversion
or

to
if

the method
successful.
uranium metal with high temperature steam,
or

water,
of

Contact
both, offers the possibility UO, but the selec
or

of

of

direct formation
process conditions avoid forming U.O.,
of

tion not well clarified.


to

is

Metal treated with hydrogen create UH, may


be
at

to

about 700°
of C
.

steam oxidized UO, after decomposition the hydride


to

vacuum.
in

Other precipitation techniques have been studied experimentally.


acid uranyl solutions with H.O.
of

These include treatment give


to

uranium peroxide; the use urea, ammonium carbonate, ammonia,


of

or hydrazine place NH, OH


of

to

obtain ammonium diuranate from


in

the hydrolysis UF, and the preparation oxalate precipitates


of

of
;

57.4789 O-61–7
84 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

from a uranyl salt treated with oxalic acid. These precipitates may
be decomposed to a higher oxide and then reduced or may be heated
in H, or CO to get UO, directly. Special crystal forms are possible.
The reduction of higher oxides, such as U3Os, offers the possibility
of obtaining low-bulk densities. Where a hydrate has been formed
(UOs: H2O), especially if combined with milling, high sinterability
is characteristic of the UO, subsequently produced by reduction in
hydrogen.
A new method of direct conversion of uranyl nitrate solutions to
UO, by an atomizing spray into a reducing flame yields a powder
with sinterability and atmospheric stability favorable for fuel element
use and, at the same time, with high reactivity to HF that is auspicious
for uranium metal production.
Direct formation of UO, single crystals from UO, powder may be
achieved by a new technique involving flame fusion in an excess of
hydrogen.
The final choice of the method of preparation of UO, suitable for
fuel element use depends on the enrichment desired (UNH versus
UFs), the sintered density specified, the stability needed against
further oxidation, the impurity content, and the relative cost in
volved.

REFERENCES

1. J.
E. WANCE, “Preparation of Pure UAOs from Crude U3O, and from Sodium
Diuranate” in National Nuclear Energy Series, Div. VII, Uranium Tech
mology, Vol. 2A, Chap. 3, pp. 47–61, U.S. Atomic Energy Commission, 1951.
2. J. E. WANCE, “Preparation of Pure Uranium Compounds from Pure Uranyl
Nitrate Hexahydrate” in National Nuclear Energy Series, Div. VII, Uran
ium Technology, Vol. 2A, Chap. 5, pp. 106–113, U.S. Atomic Energy Com
mission, 1951.
3. C. D. HARRINGToN, “Preparation and Properties of Uranium Dioxide Pow
der” in “Fuel Elements Conference, Paris," TID–7546, Book 2, pp. 369–
383, Mar. 1958.
4. C. D. HARRINGTON, et al., “Uranium Oxide Refinery Ether Process, Current
Commission Methods for Producing UOs, UF, and UFe,” TID-5295, pp.
13–72, Jan. 1956.
5. B. G. RYLE, et al., “TBP Extraction Process—Fernald Refinery, Current
Commission Methods for Producing UOs, UF, and UF,” TID-5295, pp.
73–142, Jan. 1956.
6. M. J. Szuli Nski, “I)evelopment
of an Agitated-Trough Continuous Cal
ciner,” Chem. Eng. Progr. 53, 586–589 (1957).
7 . R. G. GEIER, “Continuous Calcination Equipment for Converting UNH to
UOs,” HW–49652A, Apr. 12, 1957.
8. C. D. HARRINGtoN and A. E. RUEHLE, eds., “Uranium Production Tech
nology,” Van Nostrand Company, New Jersey, 1959.
9. A. A. Jon KE, et al., “Active Process Development Activities for Processing
Feed Materials,” TID-7501, Part 1, pp. 53–72, Feb. 1956.
PREPARATION OF URANIUM DIOXIDE 85

10. E. J. PETKUs, H. M. KATz, A. A. JonkE, and N. M. LEvitz, “Fluidized-Bed


Process for the Production of Uranium Tetrafluoride,” Chem. Eng. Progr.
53, 199–202 (1957).
11. E. F. SANDERs and S. N. Robinson, “Fluid Bed Denitration” in “Process
Development Quarterly Report, Part II,” MCW—1409, Nov. 1957, pp. 33–39.
12. C. W. KUHLMAN, “Reduction of Uranium Trioxide to Uranium Dioxide with
Hydrogen-Reaction Rates at Various Temperatures,” MCW-142, Oct. 1948.
13. R. H. MooRE and R. F. MANEss, “Reduction of UO, to UO, with Hydrogen,”
HW–38321, July 1955.
14. N. C. ORRICK, “Hydrogen Reduction Rates of Uranium Trioxides as Obtained
with a Thermobalance,” K–1081, Dec. 1953.
15. B. A. J. LISTER and G. M. GILLIEs, “The Conversion of Uranyl Nitrate to
Uranium Dioxide and to Uranium Tetrafluoride” in “Process Chemistry,"
F. R. Bruce, J. M. Fletcher, H. H. Hyman, and J. J. Katz, eds., pp. 19–35,
McGraw-Hill Book Co., New York, 1956.
16. J. S. ANDERSON, E. A. HARPER, S. MooRBATH, and L. E. J. Roberts, “The
Properties and Microstructure of Uranium Dioxide; Their Dependence
upon the Mode of Preparation,” AERE-C/R-886, Aug. 19, 1952.
17. C. W. KUHLMAN, “Reduction of Uranium Trioxide to Uranium Dioxide with
Anhydrous Ammonia,” MCW-217, Sept. 1949.
18. C. W. KUHLMAN, “Reduction of Uranium Trioxide with Hydrogen-Nitrogen
Mixtures,” MCW-215, Sept. 1949.
19.

K. PETRETIc and H. W. BERTIAM, “Summary Report on the Long Range


G.

in

Program for the Period Ending February 15, 1953,” C. Rodden, ed.,
NYO—2039, Apr. 15, 1953, J.
p.
6.

W. KRAUSE, “Quarterly Progress Report for the


G.

Goldbeck and
C.

J.

J. in

20.
Period Ending March 31, 1954,” C. Rodden, ed., NYO–2049, June 1953,
p.
7.

W. KUHLMAN, “Note on the Inhibitory Effect Hydrogen Fluoride on


of
C.

21.
the Hydrogen Reduction of Uranium Trioxide,” MCW—147, Oct. 1948.
Sw1NEHART, “Reactivity Test for UO,” “Process Development Quar
B.
A.

in

22.
terly Report, Part I,” MCW—1382, pp. 73–75, Jan. 1956; MCW–1385,
1, 3,

pp. 38–44, Apr. 1956; MCW—1392, pp. 24–33, July 1956; MCW—1399,
1, 1,

pp. 31–33, Jan. 1957.


23.

W. KUHLMAN, “Reduction Uranyl Fluoride with Hydrogen. Re


A.
of
C.

action Rate. B. Reaction Mechanism,” MCW-175, Mar. 1949.


W. HUNTINGToN, K. Notz, and M. MENDEL, “Effect Trace Im
of
in G.
J.
C.

24.
purities on Orange Oxide Reduction” “Summary Technical Report,”
NLCO-670, Apr. 18, 1957,
p.

43.
25.

LUdwig and KENNELLEY, “Hydrofluorination UO,”


A.

of
J.

J.

“Process
in
F.

Development Quarterly Report, Part I,” MCW-1412, Apr. 1958, pp.


1,

55–78.
M. Edwards, al., “Uranium Tetrafluoride Plant, Current Commission
et
R.

26.
Methods for Producing UOs, UF, and UFe,” TID-5295, Jan. 1956, pp.
177–198.
KREss, “UF, Manufacture
A.

Development, Pilot Plant


H.

B.

SMILEY and
S.

27.
Preparation of UF,” K–479, Aug. 10, 1949.
Uranium Trioxide
C.

to

BRATER,
of

H. SMILEY and D. “Conversion


S.

28.
Uranium Tetrafluoride Vibrating Tray Reactors, Current Commission
in

Methods for Producing UOs, UF, and UF,” TII)-5295, Jan. 1956, pp.
161–176.
KERR, W. KRECHT, and MULDRow, “Fluid-Bed Reduction Re
K.
H.
E.
G.

S.

29.
the Destrehan Street Green Salt Plant” “Process Development
at

actor
in

Quarterly Report, Part II,” MCW–1411, Feb.


1,

1958, pp. 63–66.


86 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

30. O. BUCKHEIM, “Aqueous Oxidation of Metallic Uranium ‘Sawdust’,” MCW


185, May 12, 1949.
31. J. J. KAtz and E. RABINowitch, “The Chemistry of Uranium,” National
NuclearEnergy Series, Div. VIII, Vol. 5, McGraw-Hill Book Co., New
York, 1951.
32. J. R. Johnson, S. D. FULKERson, and A. J. TAYLOR, “Technology of Uranium
Dioxide, A Reactor Material,” Am. Ceram. Soc. Bull. 36, 112–117 (1957).
33. J. R. Johnson, “Uranium Dioxide” in “Progress in Nuclear Energy, Series
W, Metallurgy and Fuels, Vol. 2,” p. 209–222, Pergamon Press of London,
1959.
34. J.C. CLAYTON and S. ARonson, “Some Preparative Methods and Physical
Characteristics of UO, Powders,” J. Chem. and Eng. Data 6, 43–51 (1961).
35. T. WATHEN, “Corrosion of Uranium Metal in Air and Steam at Various
Temperatures,” BR-223A ; R/GC/1304, May 13, 1943.
36. C. EICHNER, B. GoLDsch MIDT, and P. VERTEs, “Preparation of Metallic
Uranium at the Bouchet Plant of the Commissariat of Atomic Energy,"
Bull, soc. chim. France 1951, 140–142.
37. B. GoLDsch MIDT and P. VERTEs, “The Preparation of Pure Uranium Metal"
in “Proceedings of the International Conference on the Peaceful Uses of
Atomic Energy, Geneva 1955,” Vol. 8, pp. 152–155, United Nations, New
York, 1956.
38. “Uranium Metallurgy in Belgium” in “Proceedings of the International
Conference on the Peaceful Uses of Atomic Energy, Geneva 1955,” Vol. 8,
pp. 156–161, United Nations, New York, 1956.
39. E. HAwthor N, “A Résumé of the Past Work on Uranium Peroxide Precipi
tation,” PM (S)-7, June 25, 1953.
40. G. W. LLEwely N, “A Report on Work Carried Out to Improve Precipitation
Efficiencies in the Crude Oxide Extraction Plant,” SRS–105, Dec. 1, 1954.
41. J. TERRAzA, J. CERRoi,AzA, and E. APARICIo, “Sintering UO, at Medium
Temperatures” in “Proceedings of the Second United Nations Interna
tional Conference on the Peaceful Uses of Atomic Energy, Geneva 1958."
Vol. 6, pp. 620–623, United Nations, Geneva, 1958.
42. A. BEI, and Y. CARTERET, “Contribution to the Study of Sintering of Uranium
Dioxide” in “Proceedings of the Second United Nations International
Conference on the Peaceful Uses of Atomic Energy, Geneva 1958.” Vol. 6,
pp. 612–619, United Nations, Geneva, 1958.
43. G. WIRTHs and L. ZIEIIL, “Special Problems Connected with the Production
of Uranium Metal and Uranium Compounds" in “Proceedings of the
Second United Nations International Conference on the Peaceful Uses
of Atomic Energy, Geneva 1958.” Vol. 4, pp. 16–21, United Nations, Geneva,
1958.
44. P. MoHR, “Production of Uranium Peroxide,” U.S. Patent 2,551,543, May 1,
1951.
45. C. E. LARson, “Processes of Producing Uranium Trioxide,” U.S. Patent
2,723,181,
Nov. 8, 1955.
46. C. A. KRAUs, “Factors in the Precipitation of UO, Which Influence the
State of the Resulting Oxide Particularly with Respect to its Reactivity
in the Vapor Phase Reaction,” A-2314, July 19, 1945.
47. B. L. KELCHNER, “A Qualitative Survey of the Physical Factors Affecting
the Precipitation of Uranium from Uranyl Nitrate Solution by Hydrogen
Peroxide,” LA-1089, July 7, 1950.

48. J. R. SANDERson, H. E. DIBBEN, and H. Masos, “The Reaction between


Hydrogen
Peroxide and Solutions of Pure
Uranyl Salts. Part I: The
PREPARATION OF URANIUM DIOXIDE 87

Interaction of Hydrogen Peroxide and Uranyl Nitrate,” CI—R–15, May


1951.
49. J. H. PATTERson, “The Gravimetric Peroxide Method for the Determination
of Uranium” in “Manual of Special Materials Analytical Laboratory
Procedures,” ANL–5410, 1955.
50. E. L. ZIMMER, “Preparation and Separation of Uranium Peroxide as a
Stage in the Chemical Purification of Crude Uraniferous Products” in
“Proceedings of the International Conference on the Peaceful Uses of
Atomic Energy, Geneva 1955,” Vol. 8, pp. 120–122, United Nations, New
York, 1956.
51. G. TRIDot, “Uranyl Peruranate,” Compt. rend. 232, 1215–1216 (1951).
52. W. H.A.MAKER and C. W. Koch, “Note on the Composition of Uranium
J.
Peroxides” in “Chemistry of Uranium,” J. J. Katz and E. Rabinowitch,
eds., TID–5290, 1958, pp. 152–153.
53. R. F. LEININGER, J. P. Hu NT, and D. E. Kos H LAND, Jr., “Composition and
Thermal Decomposition of Uranyl Peroxide,” CC–3424, June 30, 1945:
TID–5290, 1958, pp. 704–721.
54. R. E. DEMARCO, D. E. RICHARDs, T. J. Collopy, and R. C. Abbott, “Evidence
for the Existence of Peroxyuranic Acids,” J. Am. Chem. Soc. 81, 4167–4169
(1959).
W. Mogg and LARson, “Factors Affecting the Precipitation
IX.

of
E.
C.

55.
Uranium Peroxide,” AECD–4103, July
8,

1946.
W. WATT, ACHORN, and MARLEY, “Some Chemical and Physical
S.
G.

L.

J.
L.

56.
Properties of Uranium Peroxide,” Am. Chem. Soc. 72, 3341–3343 (1950).
J.

Boggs and M. El-CHEHABI, “The Thermal Decomposition


of
E.

57. Uranium
J.

Peroxide, UO,-2H2O,” Am. Chem. Soc. 79, 4258–4260 (1957).


J.

UKAz1, “On the Thermal Decomposition Uranyl Nitrate, Ammonium


of
R.

58.
Diuranate and Uranium Peroxide,” Atomic Energy Soc. Japan

1,
J.

405–411 (1959).
KITE and D. W. SMITH, “Development High Density Uranium Di
H.
T.

of

59.
oxide Powders,” Y-876 (Del.), May 28, 1952.
50.

KRAUs, “Preparation by Vapor


of
A.

Uranium Tetrachloride Phase


C.

by

Reducing Uranium Trioxide


of

Chlorination Uranium Dioxide Formed


with Ethanol,” A-2313, July 12, 1945.
KRAUs, “Preparation Uranium Tetrachloride by Vapor Phase
C.
A.

of

61.
Chlorination of Uranium Dioxide Formed by Reducing Uranium Trioxide
with Ethanol. No. II,” A–2321, Sept. 17, 1945.
WAtson, “Production Uranium Dioxide for Ceramic Fuels,” CRL–
C.

of
L.

62.
45, Nov. 20, 1957.
Ew INg, KIEHL, and “Investigation
A.

A.

BEARSE,
of
J.
S.

E.

Ammonium
R.

63.
Uranates,” BMI-1115, July 19, 1956.
GELIN, H. MogARD, and NELsoN, “Refining
R.

B.

of

64. Uranium Concentrate and


Production of Uranium Oxide and Metal” “I’roceedings of the Second
in

United Nations International Conference on the Peaceful Uses of Atomic


Energy, Geneva 1958,” Vol. pp. 36–39, United Nations, Geneva, 1958.
4,

SchoNBERG, RUN FoRs, and R. KIEssli Ng, “The Sintering of Uranium


U.
N.

65.
Dioxide” “Proceedings of the Second United Nations International
in

Conference on the Peaceful Uses of Atomic Energy, Geneva 1958,” Vol.


6,

pp. 605–611, United Nations, Geneva, 1958.


A. HERMANs. “The Preparation of Uranium Dioxide Fuel for Sus
E.

66. M.
a

pension Reactor” “Proceedings of the Second United Nations Inter


in

national Conference on the Peaceful Uses of Atomic Energy, Geneva


1958,” Vol. pp. 39–44, United Nations, Geneva, 1958.
7,
88 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

67. L. C. FARRELL and L. F. GRILL, “Ammonium Diuranate Precipitation with


Anhydrous RFP-133, Mar. 2, 1959.
Ammonia,”
68. W. T. Bourns and L. C. WATsoN, “Preparation of Uranium Dioxide for Use
in Ceramic Fuels. Part I. Batch Precipitation of Ammonium Diur
anate,” CRCE–716 (Pt. I), Sept. 1958.
69. F. J. ARM son, H. E. DIBBEN, and H. MAsoN, “An Investigation into the
Formation and Composition of Ammonium Diuranate by Potentiometric
Titration of Uranyl Salts with Ammonium Hydroxide,” SCS-R-30, July
20, 1949.

70. P. S. GENTILE, L. H. TALLEY, and T. J. Colory, “The Chemistry of Uranyl


Nitrate-Hydroxide-Urea Systems,” J. Inorg. & Nuclear Chem. 10, 110–113

(1959).
71. G. TRIDot, “Conditions for Precipitation and Stability in Aqueous Solutions
of Ammonium, Sodium, and Potassium Uranates. Application to a Study
of the Red and Orange Sulfides of Uranium,” Ann. Chim. 5, 358–397

(1950).
G. W. WATT, W. A. JENKINs, and J. M. McCUIston, “Reactions of Some
Uranium Compounds in Liquid Ammonia,” J. Am. Chem. Soc. 72, 2260–
2262 (1950).
73. E. YATABE and L. C. WATson, “Preparation of UO, for Use in Ceramic
Fuels. Part II. Continuous Precipitation of Ammonium Diuranate,"
CRCE–716 (Pt. II), June 1958.
74. G. H. CHALDER, N. F. H. BRIGHT, D. L. PATERson, and L. C. WATson, “The
Fabrication and Properties of Uranium Dioxide Fuel" in “Proceedings
of the Second United Nations International Conference on the Peaceful
Uses of Atomic Energy, Geneva 1958,” Vol. 6, pp. 590–604, United Nations,
Geneva, 1958.
75. C. DUVAL, “The Thermogravimetry of Analytical Precipitates. XXIV. De
termination of Uranium,” A mal. Chim. Acta. 3, 335–344 (1949).
76. H. W. SAFFoRD and A. KUEBEL, “Preparations and Properties of Ammonium
Diuranate,” J. Cnem. Education 20, 88–91 (1943).
77. A. R. BANCROFT and L. C. WATson, “Preparation of Uranium Dioxide for
Use in Ceramic Fuels. Part III. Ammonium Diuranate Reduction
Studies,” CRCE–716, May 1958.
78. J. D. PEDREGAL and R. R. Sol,ANo, “Self-Reduction of Ammonium Diuranate
to Uranium Dioxide” in “Proceedings of the Second United Nations In
ternational Conference on the Peaceful Uses of Atomic Energy, Geneva
1958,” Vol. 4, pp. 85–87, United Nations, Geneva, 1958.

79. “New Route to Reactor Grade UO,” Chem. Eng. Neu's 38, No. 19, 54 (1960).
80 . I. H. WARREN and F. A. Forw ARD, “Hydrometallurgical Production of
Uranium Dioxide for Reactor Fuel Elements,” presented at the American
Ceramic Society meeting, Philadelphia, May 1960.
PREPARATION OF URANIUM DIOXIDE 89

81. G. T. SEAborg and J. J. KATz, eds., “The Actinide Elements," National Nu


clear Energy Series, Div. IV, Vol. 14–A, p. 160, McGraw-Hill Book Co.,
New York, 1954.
W. Ö. DE CoN INck and A. RAYNAUd, “Uranyl Oxalate,” Bull, soc. chim.
France 11, 531 (1912).

G. C. PIMENTEL and F. R. STEve Nso N, “Solubility of Uranyl Oxalate and


the Existence of Undissociated Uranyl Oxalate in Solution” in “Chemistry
of Uranium,” J. J. Katz and E. Rabinowitch, eds., TID-5290, 1958, pp.
168–172.

G. GTBsoN and J. J. KAtz, “The Reaction of Uranium Oxides with Liquid


Dinitrogen Tetroxide Anhydrous Uranyl Nitrate,” J. Am. Chem. Soc. 73,
5436–5438 (1951).
5. A. Bou LLF. R. JARY, and M. Do MIN É-BERGES, “The Oxides of Uranium Re
sulting from the Decomposition of Uranyl Oxalate,” Compt. rend. 230,
300–302 (1950).
I. SHEFT, S, FRIED, and N. DAvidson, “Preparation of Uranium Trioxide,"
J. Am. Chem. Soc. 72,2172–2173 (1950).
. R. B. Holden, “Factors Influencing the Reactivity of Uranium Trioxide,”
TID–5063, Oct. 9, 1951.
D. A. WAUGHAN, J. R. BRIDGE, A. G. ALLIsoN, and C. M. SCH wartz, “Process
ing Variables, Reactivity, and Sinterability of Uranium Oxides,” Ind.
Eng. Chem. 49, 1699–1700 (1957).
D. R. Stex QUist, B. MASTEL, and R. J. ANICETTI, “Note on Correlation of
Surface Characteristics of Uranium Dioxide Powders with Their Sinter
ing Behavior,” J. Am. Ceram. Soc. 41,273–274 (1958).
M. J. Poliss AR, “Process for the Production of an Activated Form of UOz,"
Sept. 24, 1957.
U.S. Patent 2,807,519,
91. D. L. BUNKER, R. C. GREENough, E. H. KALMUs, and R. J. BARD, “The Acti
vation of Low-Reactivity Uranium Dioxide Particles,” LA—1952, Oct.
1955.

J. R. Joh Nso N and A. J. TAYLoR, “Method for Preparation of UO, Particles,"


U.S. Patent 2,905,528, Sept. 22, 1959.
I. E. KN tº psh:N, N. M. LEvitz, and S. LAw Roski, “Preliminary Report on Con
version of Uranium Hexafluoride to Uranium Dioxide in a One-Step Fluid
Bed Process,” ANL-6023, Aug. 1959.
. L. M. FERRIs and R. P. GARDNER, “Recycle of UO.F, in the Fluorox Process:
Reaction of UO.F, with Hydrogen,” ORNL-2690, July 9, 1959.
. L. H. Brooks, E. W. GARNER, and E. WHITE HEAD, “Chemical and X-ray
Crystallographic Studies on Uranyl Fluoride,” IGR–TN/CA–277, Feb. 1956.
W. H. HEADLEY, R. J. Roe HRs, and W. T. TRAsk, “Flame Denitration” in
“Process Development Quarterly Report,” MCW-1451, Aug. 1, 1960.
90 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

97. C. M. HENDERson, “High Temperature Flame Techniques Related to Nu


clear Fuel Element Materials” in “Ceramic Information Meeting Held at
Oak Ridge National Laboratory on October 1, 2, and 3, 1956,” TID–7530,
Part 1, pp. 101–114.

98. A. VERNEUIL, “Growing Synthetic Rubies by Fusion,” Compt. rend. 135,


791–794 (1902).
99. A. VERNEUIL, “Synthetic Reproduction of the Sapphire by the Fusion Meth
od,” Ann. chim. et phys. 3, 20–48 (1904); Compt. remd. 150, 185–187 (1910).
100. A. VERNEUIL, “New Process for the Fusion and Refining of Chrome-Alumina
and the Production of a Material Possessing the Composition and Hard
ness of Ruby,” Compt. rend. 15, 131 (1911 ).
101. A. VERNEUIL, “Making Synthetic Sapphires,” U.S. Patent 988,230, Mar. 28,

1911.

102. A. VERNEUIL, “Synthetic Sapphire,” U.S. Patent 1,004,505, Sept. 26, 1912.
Chapter 3

CHARACTERIZATION OF URANIUM DIOXIDE


J. C. CLAYTON

3.1 INTRODUCTION

Certain physical and chemical properties of a solid, such as uranium


dioxide, are determined largely by preparation. By
its

of
method
varying the properties

of
parent compound, the temperature
of

the
reaction, and the conditions reduction,
of

oxidation and considerable


exercised over the chemical composition, particle shape,
be

control can
size and size distribution, density, and related characteristics

of
the
UO, powder [1–4]. These powder characteristics can, turn, affect
in
the properties the finished product the perform
of of

of

as

as
some well
during processing into solid fuel. knowledge

in of
ance the oxide
A
is,

the physical and chemical characteristics therefore, important


selecting the most suitable UO, powder for particular application.
a

these properties the principal factor obtaining uni


of

Control
in
is

formity among different UO, powder batches.

3.2 PHYSICAL CHARACTERISTICS

The physical characteristics


in
of

uranium dioxide are described


Sects. 3.2.1 through 3.2.6. These characteristics include density, micro
structure, surface properties, particle size and distribution, crys
tallite size, and optical properties (color, spectra, index
of

refraction)
the UO, powder.
of

3.2.1 Density

The theoretical density values UO, vary from 10.95 g/cc


of

is to

10.97
[1, 10]. This ideal, crystallite, X-ray density
5,

or
6,
7,

calculated
its

the unit cell and the atomic weights constituents,


of

of

from the size


all

assuming that lattice points are occupied." Density data obtained


10.96 g/cc for the theoretical density
be of

of

value stoichiometric uranium dioxide


A
*

(UO2 oy will assumed throughout this book.

91
92 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

by the displacement of a fluid that can penetrate all pores and capil
laries of a solid down to molecular dimensions are termed real densi
ties. Closed pores, voids completely surrounded by the solid matrix
and, thus, inaccessible to any fluid except by diffusion through the
solid, are arbitrarily considered as part of the real volume of the ma
terial. The helium atom has a small diameter which enables it to
penetrate into very fine pores; in addition, there is negligible adsorp
tion of helium on solids at room temperature [8]. Therefore, densi
ties obtained by helium displacement most closely conform to the defi
nition of real densities. Measurements obtained by liquid displace
ment are usually lower because of incomplete penetration of the pores
of the solid by the liquid [8].
Published data on UO, densities vary over a wide range (10.0 to
11.0 g/cc), depending on the experimental technique and the method
of preparation of the uranium oxide [1, 2, 5, 9, 10]. Some typical re
sults are listed in Table 3.1. The measured densities of UO, powders
prepared by steam oxidation procedures approach the theoretical
10.96 g/cc [2]. The liquid densities are slightly lower than the helium
values, indicating that these powders consist of dense granules (no
closed porosity) with varying amounts of small open pores. Low
density UO, (10.0 to 10.4 g/cc) is obtained from UOs and UAOs made
by direct pyrolysis of uranyl nitrate hexahydrate crystals [1, 2, 9].

|=–
TABLE 3.1—DENSITY OF UO, [2]

Method of preparation 9"


Density (g,cc)
-
Helium CCl."

High-pressure steam oxidation of uranium-- - - - - - 1.97 10. 92+0. 1 10. 75

Controlled low-pressure steam oxidation of


uranium---------------------------------- 2. 03 10. 85 + 0.1 10. 91
Steam oxidation of uranium from UH3- - - - - - - - - - 2. 03 10. 98+ 0. 1 10. 67
Hydrogen reduction of UO3·2H2O - - - - - - - - - - - - - - 2. 01 10. 25 + 0.05 || 10. 25
Hydrogen reduction of UO3.H2O - - - - - - - - - - - - - - - 2. 01 10, 34+ 0.05 || 10. 37
Air pyrolysis of UO2(NO3)2·6H2O to UOs;
hydrogen reduction to UO.” - - - - - - - - - - - - - - - 2. 02 10. 16+0.05 || 10. 13
Air pyrolysis of UO2(NO3)3·6H2O to UAOs;
hydrogen reduction to U():- - - - - - - - - - - - - - - - - - 2. 02 10. 03 + 0.05 9, 96
Hydrogen reduction of uranyl oxalate-- - - - - - - - - - 2. 01 10. 61 + 0.2 10. 61
Hydrogen reduction of uranium peroxide--- - - - - - 2. 08 11. 00+ 0.1 10. 78
Hydrogen reduction of ammonium hydroxide
precipitated diuranate - - - - - - - - - - - - - - - - - - - - - - 2. 07 11. 1 1 + 0.2 10. Sl
Hydrogen reduction of urea precipitated -
diuranate--------------------------------- 2. 08 10. 74 + 0.2 10. 19

"Precision: +0.2 percent.


** MCW oxide used in PWR Core 1.
CHARACTERIZATION OF URANIUM DIOXIDE 93

For this type of UO, the helium and liquid densities, besides being
considerably below the theoretical values, are in close agreement with
each other. This suggests that these oxides contain little open porosity
-
but an appreciable amount of closed pores.
In general, the real density of a UO, powder preparation depends
upon the density of its parent higher oxide (UO,-2H.O, UO, H.O,
UOs, or UAOs) [1, 2,9]. The density is not dependent on the reduction
temperature up to 1,200° C [1,2]. However, heating low density UO,
its

preparation can cause

an
temperature of its den

in
above increase

by
sity [1,2]. The densities UO, made CO reduction
of

of
UAOs are

by
measurably higher than the densities UO, prepared H, reduction

of
the same material under similar conditions [1]. For many UO,
of

preparations the liquid densities are considerably below the helium


values [2]. This the inability the large liquid molecules

of
to

due
is

penetrate small microcracks and fissures and indicates that consider


to

open porosity are present these powders.


of

able amounts
Grinding low density UO, powders does not increase their fluid den in

sities significantly [1,


2,

11]. Comminution processes, general, are

in
the closed pores. Milling the less abrasive UO,
of

open any
to

unable
its

produces some increase density which retained upon reduction


in

is

UO. Uranium dioxide powders with measured densities above the


to

theoretical value have been found have O/U ratios well above 2.00
to
5,
9,

[2, 10, 12, 13].


Density significant characteristic both green (as-pressed) and
of
is
a

sintered UO, bodies. UO, compact contains both open (connected)


A

pores. The total volume can


be

and closed (unconnected) considered


the open pore volume, the closed pore volume, and the
of
as

the sum
theoretical UO. volume. Techniques have been developed
to

measure
Examples
of

of

each these volumes [14–16]. the relation between


density and porosity sintered UO. pellets prepared from MCW and
in

ADU Figs. Chap. for


in

oxides are shown 3.1 and 3.2 (see more


7

detailed discussion). At the lower densities, the closed porosity


relatively constant, and the main mode de
of

densification the
is

is

open pores. There the closed poros


of
in

in

crease the number rise


is
a

ity when the open porosity approaches zero the higher density
in

range. These curves differ somewhat for bodies prepared from


by

UO,
or

different fabrication techniques [14, 16].


of

different sources

3.2.2 Microstructure

The particle shape UO, powders has important influence


on
an
of

their performance during processing and the final properties


on

of

the
sintered product. Particular particle contours may control powder
pouring, packing, flow characteristics, and both the as-pressed and
addition, particle shape affects the
In

sintered density compacts.


of
94 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I I T i
A L.
Liu
> 15 H.
B
co
>
^A - s–
g
*
H. §
3 #
Gl. c
Q->
5 A
|
to - - to z
º 3
§ e CLOSED POROSITY, 1500 °C ar
O CLOSED POROSITY, 17OO "C #.
É
A OPEN POROSITY, 1500 °C >
:
a.
> A OPEN POROSITY, 17oo *C
º c
2 5 H – 5 5
El
o
0. c
#: O-es N
c

o O
N.
N.
O | | | I N o
75 8O 85 90 95 IOO
SINTERED DENSITY, PER CENT THEORETICAL

FIGURE 3.1. Porosity as a Function of Density: MCW Oxide Pressed to 65


Percent of Theoretical Density and Sintered in Hydrogen [14].

IO - - lo

CLOSED POROSITY, 1500 °C


CLOSED POROSITY, 17OO "C

: OPEN POROSITY, 1500 °C


OPEN POROSITY, 17oo "c

75 8o 85 90 95 too
SINTERED DENSITY, PER CENT THEORETICAL

FIGURE 3.2. Porosity as a Function of Density: ADU Oxide Pressed to 65


Percent of Theoretical Density and Sintered in Hydrogen [14].
CHARACTERIZATION OF URANIUM DIOXIDE 95

results of bulk density, particle size, and surface area measurements.


Microscopic examination (see Appendix B, Microscopy of UO.) of
UO2 powders indicates that their particle size, shape, and structure de
pend upon the method used to prepare them [2–4, 6, 7, 17–26]. These
diversecharacteristics are illustrated by the photomicrographs in
Figs. 3.3, 3.4, and 3.5. The particle sizes range from tenths of
microns to several millimeters. The shapes include spheres, needles,
plates, and irregularly curved and angular particles. The internal
particle structures vary from nearly solid chunks to very loose spongy
masses and include clusterlike and cored structures. In many cases,
the shape and even the size of the powder particles of the original
precipitate or higher oxide have been retained upon conversion to UO2
[2–4, 6, 7, 23, 26, 27]. In fact, the general shape and size of agglom
erates have been preserved through the (NHA).U.O.-UOs–U3Os–
UO2—UF, chain even though several chemical reactions are involved
in these steps [23]. Wide variations in the surface characteristics of
UO, powders derived from various sources are also seen in the electron

A. UO, from Spray Denitrated UOs: B. UO, from Air Oxidized Uranium
×250, Reduction Factor 14. Metal Foil; X250, Reduction Fac
tor. 14.

(". MCW PWR Core 1 UO, ; X250, D. Ball-Milled MCW PWR Core 1
Reduction Factor, 14. UO. : X 500, Reduction Factor, A.

FIGURE 3.3. Micrographs of Typical UO, Powders [2].


96 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

A. Controlled Low-Pressure Steam- B. UO, from UHs; X250, Reduction


Oxidized UO2; X250, Reduction Factor,A.
Factor, 14.

C. UO, from UO, H2O Hydrate Crys- D. High-Pressure Steam-Oxidized


tals; X250, Reduction Factor, 14. UO, ; X250, Reduction Factor, 14.

FIGURE 3.4. Micrographs of Typical UO2 Powders [2].

micrographs of Fig. 3.6. Powder agglomerates, surface texture,


and particle size and shape can be readily observed. Through
the use of electron micrographs, definite relationships have been
noted between the sintering behavior of UO, powders and their sur
face texture and particle size (see Chap. 7) [20, 21, 24, 25].

3.2.3 Surface Properties

(a) Surface Area

Surface area measurements obtained by gas adsorption techniques


determine the total area of a solid. In addition to an external (mac
roscopic) surface, porous bodies possess an internal (microscopic) sur
face because of minute pores and cracks. Both external and internal
areas are important characteristics in the performance of a UO.
powder. Sedimentation rates, resistance to flow of gases, bulk density,
flow, and packing properties are functions of the external surface.
CHARACTERIZATION OF URANIUM DIOXIDE 97

A. UO, from Ammonia Precipitated B. UO2 from Ammonium Carbonate


Diuranate; X250, Reduction Fac- Precipitated Diuranate; X250, Re
tor, 14. duction Factor, A.

C. UO, from Uranyl Oxalate; X250, I). UO, from Uranyl Oxalate; 6300X,
Reduction Factor, 14. Reduction Factor, 12.

FIGURE 3.5. Micrographs of Typical UO, Powders [2].

The total surface is involved in considering chemical reactivity, such


as rates of solution, oxidation, hydrofluorination, and chlorination, as
well as in problems concerned with the effect of surface impurities and
adsorbed gases. Both surface roughness and particle size can account
for variations in the sintering behavior of UO, powders [4, 20–22,
24, 25].
Permeability and microscopic measurements also give an estimate
of the surface area of a powder. In the permeability method, the ve
locity of a fluid (under a constant pressure gradient) through a packed
bed of powder is measured. The porosity of the packed bed is a
factor determining the rate of flow and must, therefore, be known. In
the microscopic procedure, surface areas are determined by viewing
the powder particles under a microscope, ultramicroscope, or electron
microscope and measuring their dimensions. The permeability and
microscopic methods give a good estimate of the total surface area of
a powder composed of smooth nonporous particles of uniform size and
98 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

A. UO, from Ammonium Carbonate B. High - Pressure Steam - Oxidized


Precipitated Diuranate; 6300X, - UO, ; 6300X, Reduction Factor,14.
Reduction Factor, 14.

C. MCW PWR Core 1 UO, ; 9000X, D. MCW Ceramic Grade UO, ; 8500X,
Reduction Factor, 14. Reduction Factor, 13.

FIGURE 3.6. Electron Micrographs of Typical UO, Powders [21, 25].

shape. The surface area measured by these methods will be lower than
that determined by gas adsorption to the extent that the powder does
not conform to the ideal specifications.
The gas adsorption method measures total accessible surface. Ex
ternal surface due to roughness and the internal surface due to porosity
are not measured by the permeability and microscopic methods. How
ever, the ratio of the surface areas determined by the gas adsorption
method to those measured by the permeability and microscopic meth
ods, Sg/Spratio(termed the roughness or porosity factor), may give
some indication of the degree of roughness, porosity, irregularity, and
CHARACTERIZATION OF URANIUM DIOXIDE 99

nonuniformity of the powder particles. When this ratio is unity, a


powder can be considered as essentially smooth and nonporous. The
relationship between gas adsorption and permeability or microscopic

is,
measurements of surface area thus, similar the relationship be

to
by
density liquid displacement.

of
tween measurements helium and
UO, has been measured by gas adsorption, per

of
The surface area
meability, and microscopic techniques [1–4, 11, 14, 16, 17, 19, 20, 24, 26,
m”/g, depending upon

25
The areas varied from

to
28–30]. 0.05
type powder and method
of

of
the measurement. The surface areas
representative UO, powders are presented
of

Table 3.2. Areas

in
some
are reported per unit volume solid material rather than per unit
compare areas of powders having different densities.
order of
in

to

mass
Total surfaces were determined by the gas adsorption method (Sg),
whereas the external surface
was estimated from permeability
measurements (Sp). The total
and external specific surfaces for the
by

UO, prepared the controlled low-pressure steam oxidation tech


nique are low and approximately equal. This agreement with the

in
is
photomicrograph (Fig. 3.4) and density data (Table 3.1) which show
that this oxide powder coarse, dense particles with

no
composed
of
is

by
open porosity. For the other powders prepared steam oxidation
procedures, the external surfaces are smaller than the total surfaces,
again consistent with the electron micrographs (Fig. 3.6) and
density results (Table 3.1) showing relatively coarse, dense particles
in

varying open porosity. The specific surfaces UO. of


of

with amounts
by

the higher uranium oxides


of

powders prepared hydrogen reduction


porosity factors (0.8
or

are also quite low. The low roughness


to

2.1)
are consistent with their helium and liquid densities (Table 3.1) and
photomicrographs (Figs. 3.4 and 3.6) showing the presence
of

little
in

open porosity surface roughness. The Sg/Sp ratio for the


no

or
or

its

UO. obtained from uranyl oxalate 1.5. Although fluid density


is

open porosity,
no

an

of

values (Table 3.1) indicate examination


a

micrograph (Fig. 3.5) this powder shows the presence


in of

of

fine
surface hairs which result slightly greater total surface area. The
a

UO2 powders prepared from uranium peroxide and ammonium diur


anate have the highest external and total surface areas and, thus, the
smallest particle sizes. Porosity factors approximately are con
of

sistent with the density measurements (Table 3.1) suggesting the


in

presence open pores. high roughness


of

in of

some The factor (9.9) the


O. made from ammonium carbonate precipitated diuranate ex
is
U

agreement with electron micrographs this powder (Fig. 3.6)


of

cellent
showing very rough surface texture.
in

UO,
on
In

general, the surface areas preparations depend both


of

their thermal history and


on

of

the surface area the material from


which they are prepared [1–4, 26]. Under approximately the same
UO, powders
or

reduction conditions, the surface areas


of

oxidation
57.4789 O—61—8
100 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 3.2—SURFACE AREAS OF UO, POWDERS [2]

Surface area (m2/cc)


Method of preparation
Total (Sg) | External (Sp) Sg/Sp

High-pressure steam oxidation of uranium--------- 12. 9 3. 1 4. 2


Controlled low-pressure steam oxidation of uranium- 2. 4 3. 2 0.8
Steam oxidation of uranium from UH3- - - - - - - - - - - - 22. 6 5. 6 4. 0
Hydrogen reduction of UO3·2H2O - - - - - - - - - - - - - - - - 8. 8 4. 1 2. 1
Hydrogen reduction of UO3.H2O - - - - - - - - - - - - - - - - - 0. 5 0. 6 0.8
Air pyrolysis of UO2(NO3)2.6H2O to UOs; hydrogen
reduction to UO,”---------------------------- 6. 4 3. 1 2. 1
Air pyrolysis of UO2(NO3)2·6H2O to U3Os; hydro
gen reduction to UO,-------------------------

9. 5. 3.

2. 1, 1.
7. ...

8 0 3
3 7

9 5 7
Hydrogen reduction uranyl oxalate-------------
of of of

Hydrogen reduction uranium peroxide- 28.

4
-
-
-
-
-
-
-

-
Hydrogen reduction ammonium hydroxide pre
cipitated diuranate--------------------------- 37.4

3. 2.
13.

6 9

3 7
Hydrogen reduction urea precipitated diuranate- 42. 12.
of of

1
Hydrogen reduction ammonium carbonate pre
cipitated diuranate---------------------------

4.
40.

9.
7

9
*MCW oxide used in PWR Core
1.

the parent uranium metal higher oxide


of

or
to

are related the areas


(UOs-2H2O, UO, H2O, UOs, UAOs) [1–3, 26]. Large particle
or

higher oxide yield large particle size


or

size samples
of

uranium metal
UO2, whereas UO, particle size can only prepared from
be
of

small
particle Ignition UO, (Fig. 3.7)

or
parent
of

small size material.


of

ammonium diuranate UAOs (Table 3.3) temperatures


to

15 at

above
particle size (up
an

gives
in

to

600° increase times some cases)


in
C

largely maintained UO, [1–3, 26]. Oxida


on

to

which reduction
is

UO, UAO, lower temperatures (300° C) can cause con


of

tion
to

at

siderable increase surface area [2, 24, 28]. The UAOs ignition
in

temperatures, therefore, are important determining its surface area


in

(and particle size) and that the UO, product.


of

The temperature preparation also important determining the


of

in
is

UO, particle size, that


is,

the higher the temperature, the greater the


particle size. This illustrated Figs. 3.8 and 3.9 which show the
in
is

UO, particle U.O, and UO, reduction tempera


of

variation size with


particle UAOs, espe
of

ture. Some breakdown occurs the reduction


in

cially the lower temperatures [1–3]. Both particle breakdown and


at

particle growth have been observed UO, [1–3].


In
of

the reduction
in
all

cases, however, particle growth very high reduction tem


at

occurs
peratures. fact, merely heating UO, powders above their tempera
In

preparation can cause particle growth [1, 2]. Where particle


of

ture
occurs, greater temperatures
at

breakdown the lower because


is
it
CHARACTERIZATION OF URANIUM DIOXIDE 101

2.O I I

TIME OF HEATING-2 HR

O.5

O5 l l I I l
3OO 4OO 5OO 6OO 7OO 8OO 900
TEMPERATURE, oc

FIGURE 3.7. Variation of Particle Size with Tem


perature of Ignition of UO, [3]. (Reprinted with
permission from B. A. J. Lister and G. M. Gillies,
“The Conversion of Uranyl Nitrate to Uranium
Dioxide and to Uranium Tetrafluoride,” Pergamon
Press, Inc., 1956.)

Table 3.3—SURFACE AREAS OF UO, POWDERS PREPARED BY HY


DROGEN REDUCTION OF U20, OBTAINED FROM AMMONIUM
DIU RANATE [2]

Surface area (m3/cc)


U3O8 ignition
Source temperature
(°C Total External (Sg/Spy
(Sg) (Sp)

Ammonium hydroxide precipitated diur


anate------------------------------- 500 59.9 12.9 4.6
Ammonium hydroxide precipitated diur
anate------------------------------- 800 9.7 5.5 1 ..8

Trea precipitated diuranate- - - - - - - - - - - - - 500 60.4 10.3 5.9


Urea precipitated diuranate- - - - - - - - - - - - - 800 6.0 3.7 1.6
Ammonium carbonate precipitated diur
anate - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - 500 36.2 3.9 9.3
Ammonium carbonate precipitated diur
anate------------------------------- 1,000 14. 1 3, 1 4.5

mobility is lower in solids, and strains resulting from the phase trans
formation are less easily annealed out at the lower temperatures. The
breakdown itself can be explained on the basis of differences in crystal
structure and density between the parent higher oxide and the UO.
product. Simply grinding the uranium oxides (U.O., U.O., U.O.) or
partially dissolving the powder particles in acids increases their sur
face area and decreases their particle size [1, 2, 11, 16, 24, 28, 31]. The
102 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I I I I I T

O. H.

O | | | | | |
O 500 6OO 7OO 8OO 900 IOOO
TEMPERATURE OF PREPARATION,”c

FIGURE 3.8. Particle Size of UO, Prepared from UAO, at Various


Temperatures [1].

increase in surface area for the higher oxides is retained upon reduc
tion to UO2.
The surface area variations of UO, powder as a function of tem
perature and atmosphere have been investigated [11, 26]. Reduction
of the powder surface area begins at a lower temperature in vacuum
or in argon than in hydrogen. In hydrogen, the UO, surface area is
constant up to 800° C and begins to decrease only at higher tempera
tures (Fig. 3.10). In vacuum, however, the specific surface remains
stable only up to 600°C, where it begins to decrease. The change is
fastest between 700° and 800° C (Fig. 3.10). This inhibition of sin
tering may be due to the absorption of hydrogen at the surface of the
UO. (see Sect. 3.2.3c) [26]. It is only after hydrogen desorption that
the oxide will sinter.
Surface area measurements, using gas adsorption techniques, have
also been made on a number of sintered UO, compacts of different den

sities (see Fig. 16 of Chap. 9) [32]. Surface area decreases with in


creasing density and is affected by surface treatments, such as grind
ing, as well as by the nature and distribution of pores in the sintered
compact. High-density (95 to 99 percent) as-sintered UO, pellets,
with no measurable open porosity (as measured by water absorption),
had roughness factors of 2 to 3 in most cases (Table 3.4). The pres
ence of small, open pores as well as surface roughness is indicated,
however, since a slow adsorption process was found in some instances
[32].
CHARACTERIZATION OF URANIUM DIOXIDE 103

ol—l l l I I I |
4OO 600 800 i2OO
IOOO 14OO I600 |800
TEMPERATURE OF REDUCTION,”C

FIGURE 3.9. Surface Area of UO, Prepared from UO, as a Function of Reduction
Temperature [2]. -

TABLE 3.4–SURFACE AREAS OF AS-SINTERED UO, COMPACTS [32]

Percenttheo- Surface area Roughness


retical Adsorbate gas (cm2/g)** factor
density."

94.6 | Butane -------------------------------- 3.0+0. 7 3. 3


96.7 Ethane -------------------------------- 1. 0+ 0.5 1. 3
98.0 | Butane - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - 1. 9-1-0. 9 2. 9
98.1 Ethane - - - - ---------------------------- 1. 5+0. 5 2. 2
98.9 | Butane - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - 1. 3 + 0. 6 1. 9

"No measurable open porosity by water absorption.


*"Limits of error based on scatter of experimental data.

The application of surface area measurements as a function of


sintered density ofUO, to the problem of fission gas release is dis
cussed in Chap. 9.

(b) Adsorption from Solution

Adsorption from solution gives a measure of the total surface


reached by the molecules of the adsorbate. the adsorbent contains If
fine cracks and pores, an adsorbate composed of large dye molecules
would be unable to penetrate the smaller openings, but could give
its

the porosity of the adsorbent or particle


of

information on state
104 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

3O I

I

2O -
-
E
on
^. -
ou
E -
<1. -
LL
or -
<!
-
LL
O
<1 -
u
or: -
->
on -
|O -

l | | | | | | | | | |
O
|OO 500 |OOO
TEMPERATURE, *c

FIGURE 3.10. Surface Area of UO, as a Function of Heating Temperature in


Different Atmospheres [26].

aggregation. In addition, it is thought that dye adsorption is a


specific chemisorption process, i.e., the adsorption occurs on certain
active groups of the surface rather than on the entire surface [8].
Langer has studied the adsorption of different dyes on UO, powders
in aqueous solutions [33]. Thionine blue was found to be most ad
sorbed by UO, and was, therefore, studied in some detail. The amount
of thionine blue adsorbed at first increases with the concentration of the
dye in solution but later levels off to a plateau. An increase in tem
perature from 25° to 50° C reduces the amount of dye adsorbed by
almost one half. Breaking up the larger UO. agglomerates by shak
ing, sieving, or milling increases the amount of dye adsorbed. The
amount of dye adsorbed, compared with gas adsorption measurements
on UO2 samples over a wide range of specific surfaces, shows a linear
relationship between dye adsorption and gas adsorption for the UO.
Samples studied.
CHARACTERIZATION OF URANIUM DIOXIDE 105
(c)
Adsorption

of
Gases

Water, oxygen, and reducing gases, such

as
hydrogen and carbon

by
monoxide, are adsorbed UO, powders. Oxygen rapidly chemi

is
by

freshly reduced surfaces UO, -183° and also by

of

at

C
sorbed
mixed crystals UO, and Tho, [34–36].
of

In
of
surfaces the latter
adsorbed only the uranium sites. Only about

on
case, oxygen

30
is

on
the occupied sites become vacant high-temperature
of

percent

on

at
evacuation. Additional oxygen adsorption UO2 commences
138°C. The amount adsorbed increases with time and temperature,
-

the gas pressure (see Chap.


on
and the rate dependent for more
is

8
of

detailed discussion the kinetics). The amounts adsorbed above

of
so

0°C are large that considerable penetration oxygen into the UO2
crystal lattice must occur.
chemisorbed by uranium dioxide temper

at
Carbon
monoxide
is

atures below 20°C [34, 36, 37]. There mutual interference between
is

on
the adsorption oxygen and carbon monoxide
of

uranium dioxide

by
of

low temperatures. The adsorption carbon monoxide uranium


at

higher temperatures All the CO recoverable

as
at

dioxide small.
is

700°C, although some carbon deposited high temperatures is


at

at

CO
is

(>500° C) and high pressures atmosphere).


(1

by

Hydrogen not appreciably chemisorbed uranium dioxide below


at is

400°C, but higher temperatures activated adsorption occurs, and


high surface coverages can attained [34, 36, 37]. may diffi
be

be
It

free the UO, from the last traces hydrogen, even C.


of
to

at

cult 750°
the sorption UO, are summarized
on
of

gases
of

The results Table


in

3.5.

TABLE 3.5–ADSORPTION AND ABSORPTION OF GASES ON UO, [36

Gas Temperature (°C) Reaction Amount gas re- Heat evolved


acting (V/Vm)* (kcal/mole)
(). ();

Chemisorption 0.3 to 0.6___| 55 to 10**


to
––

195 183
-

-
_ -

-
-
-
-
-
-
-
of

Oxidation ~40Å 2.2-------


=

138 to 50_
_
_
_
_

depth
(),

UO2.4-----|-
to

18) Oxidation
to

100 88+
-
-

_ -
-


-
-
-

-
-
-
-
-
-
-
-

C() 183 to 20 Chemisorption 0.7 to 0.8_ 20 to ft


4
_
-
-
-
-
-

-
-
-
-
-
-
-
-

C() 275 Chemisorption (?) 0.08_


--

_
_
_
_
_
_
-
-
-
-
-
-
-
-
-
-

*
- -
-

C() >500. -----| Carbon deposition 2.0


--

-
-
-
-
-
-
-
.
.
.
.

Chemisorption (?) ~ 0.3_


H; H.

183
_
_
_
_
_
-
-
-
-
-
-
-
-
-
-
-

-
-
-
-
-

20--------------|-------- - - - - - - - - - - - - - -
0
-
-
-
-
-
-
-
-
-
II,

1.0 to 1.6 35ft


>

Chemisorption
-

400
f
-
-
-
-
-
-
-
-
-
-
-

-
-
-
-
-
-
-
-
|

multiple Vm, the amount oxygen necessary physically adsorbed monolayer


or

to
of

of

"Fraction form
a
-

at 183°C.

"Directly measuredcalorimetrically.
of

of

Estimated from the heats formation UO2 and UO 2.25.


tº *

Estimated from rates desorption.


of

***Calculatedfrom adsorption isotherms.


106 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

(d) Heats of Adsorption

The measurement of heats of adsorption of gases on solids can


yield useful information about surface structure [8]. Examples of
constant heats of adsorption over the entire surface of a solid, or even
over a large part of the surface, are rare. The heat of adsorption of
an isolated molecule on a plane surface is different from that of a
molecule located in the crevices, pockets, or pores of an adsorbent.
Even on a plane surface the heat of adsorption may vary considerably
because of the packing of the surface atoms. Heats of adsorption of
oxygen on several types of UO, and solid solutions of UO, in Tho.
have been measured by British workers [38–40.]. At -183° C all of
the UO, specimens showed a decrease in the heat of adsorption with
surface coverage. The decrease was most marked for samples of
greatest surface area. The initial heat of adsorption was 55 kcal/
mole. The solutions of UO, in Tho, showed a steady decrease in the
heat of adsorption with uptake of oxygen. The initial heat of adsorp
tion decreased as the uranium content of the solid solution was re
duced. In all cases, both physical adsorption and chemisorption
processes were detected.

(e) Hahn Emanating Power *

The fraction of radioactive inert-gas atoms formed in a solid that


escapes from the solid is called its emanating power. The emanating
power of a solid depends on the composition, crystal structure, specific
surface, and temperature of the solid and also on the half-life and
recoil energy of the radioactive inert gas [41]. Anderson, et al., in
dexed a set of UO, powders with radiothorium and compared their
emanating characteristics by determining the rate of escape of thoron
emanation (Rn “”) as a function of temperature [1]. The results
illustrated the dependence of the microstructure of a UO, preparation
on the nature of the parent substance. The samples with the smallest
particle size had the highest overall emanating power.

3.2.4 Particle Size and Size Distribution

Mean or average particle diameters are usually calculated from gas


adsorption measurements on the assumption that the powder is com
posed of ideal uniform spheres. Particles in a real powder generally
have a variety of sizes and shapes and differ in surface roughness and
porosity. The calculated average particle diameters may, in many
cases, have little physical meaning. The assumption of uniform
spheres is also made in calculating average particle diameters from
permeability measurements. The microscopic determination of pow

* See Chap. 9 for a discussion of Hahn emanation and rare gas diffusion.
CHARACTERIZATION OF URANIUM DIOXIDE 107

der particle size is the only direct method (see Appendix B, Micros
copy of UO2), but is limited by irregularities in particle shapes and
the difficulty of obtaining a truly representative small sample. Both
the permeability and microscopic methods should give a better indica
tion of the physical size of powder particles than gas adsorption meas
urements, since the internal surface area and the surface roughness of
the particles are not measured by these techniques. In practice, how
ever, powders are rarely composed of particles of uniform size, but
exist in a range of different sizes. It is useful, therefore, to determine

is,
the particle size distribution, that the quantitative distribution

of
the particles among the various sizes.
Surface area and average particle diameters can

be
calculated from
sedimentation and elutriation techniques used measuring particle

in
size distributions. One might expect reasonable agreement between
the surface areas and the average particle diameters calculated from
microscopic permeability measurements and from particle size
or

distribution measurements, since all three methods the geometric


in

the particles are primary importance.


of

However, average
of

sizes
particle diameters calculated from particle size distribution measure
ments are generally severalfold greater than those calculated from
permeability and microscopic measurements. The difference can

is, be
attributed the fact that most powders are agglomerated, that
to

they are composed aggregates less loosely bound discrete


of

of

or

more
particles. The different experimental techniques used together, how
ever, can provide complementary information
on

the size, shape, de


porosity, and degree
of

gree aggregation powder particles.


of

of

The particle size UO, powders has been calculated from gas ad
of

sorption, permeability, microscopic, and particle size distribution


measurements [1–4, 17–20, 28, 29, 42]. The particle sizes varied from
to

microns, depending upon the type powder


10

of

0.01 and method


The particle sizes some typical UO, samples are
of

of

measurement.
listed Table 3.6 where particle sizes calculated from gas adsorption,
in

permeability, and air sedimentation measurements are compared.


parti'e with one containing cracks and
or

Since rough surface


a
a

by

pores will particles the gas adsorp


as
be

measured several smaller


tion method, the smallest particle diameters are obtained with this
technique. seen below, the particle size ranges
be

of

As will milled
UO2 powders are relatively narrow, and fair agreement might
be

expected between average particle diameters calculated from perme


ability and particle size distribution data. However, the average
permeability particle diameter for the milled oxides half that cal
is

culated from the particle size distribution. This indicates that UO,
form hard aggregates, and even milled powders exist ag
an
to

tends
in

glomerated form. Electron micrographs support this conclusion [2,


11, 17].
108 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 3.6–PARTICLE SIZES OF UO, POWDERS [2]

Particle diameter (microns)

Method of preparation
Gas adsorp- Permeability Air sedimen
tion tation

Air pyrolysis of UO2(NO3)2·6H2O to UO3;


hydrogen reduction to UO2* - - - - - - - - - - - 0.94 1. 9 6. 6
MCW PWR Core 1 UO, wet milled with
860-g U balls-- - - - - - - - - - - - - - - - - - - - - - - - - 0.33 1. 1 2. 1
MCW PWR Core 1 U(), wet milled with
1500-g U balls - - - - - - - - - - - - - - - - - - - - - - - - 0. 19 1. 0 1. 7
Hydrogen reduction of milled UOa---- - - - - - - 0.43 1. 3 4. 6
Pyrolysis of aqueous ammonia-precipitated
diuranate to U3O8; hydrogen reduction
to UO2------------------------------- 0. 39 1. 8 7. 5
Pyrolysis of urea-precipitated diuranate to
U30s; hydrogen reduction to UO2- - - - - - - 0.35 1. 4 5. 0

*MCW oxide used in PWR Core 1.

Particle size distribution analyses of UO, powders have been deter


mined by air and liquid sedimentation and microscopic counting
techniques [2, 3, 11, 17–19, 27, 42]. The particle size distribution
curves as determined by air sedimentation for several representative
UO, powders are plotted in Fig. 3.11. All the powders have a
wide range of particle sizes, varying generally from 5 to 50 microns.
Although an average particle diameter of around 2 microns is cal
culated from permeability measurements for these UO, powders, the
particle size distributions differ appreciably.
The effect of the precipitant on the particle size distribution of
UO, powders prepared from ammonium diuranate is shown in Fig.
3.12. The particle size analyses are in agreement with microscopie
evidence (Fig. 3.5). The UO, particles prepared from ammonium
hydroxide and ammonia-precipitated diuranate varied widely in size.
Urea-precipitated diuranate yielded oxide with less particle size varia
tion, whereas UO, made from ammonium diuranate precipitated with
ammonium carbonate had a relatively narrow size range. In general,
the particle size distribution of a UO, preparation depends upon the
particle size distribution of the higher oxide from which it has been
prepared by reduction and on the reduction temperature itself [2].
In one study on the reduction of UOs, the uranium dioxide reduction
product was most similar in surface area and particle size distribu
tion to the parent oxide at low temperatures (480° C) [2]. Particle
growth occurred at higher temperatures. No change in particle size
distribution occurred, but an increase in surface area was observed
on low temperature reduction (500° C) of U.O. to UO. There
CHARACTERIZATION OF URANIUM DIOXIDE 109
loo
I | I

so A-HIGH-PRESSURE STEAM oxidATION


B - Mcw PwR core- Uo.2
so c - Mcw PRECIPITATED Uo.2
D - FLUIDIZED-BED DENITRATION
7OH- Uos REDUCTION

60 H
50 H

4OH

3OH

|
| 2 5 IO 20 50 |OO
DIAMETER IN MICRONS

FIGURE 3.11. UO, Particle Size Distribution : Effect of Powder Preparation [2].
100
I I I

90 – A-AMMONIUM CARBONATE
B - UREA
80 | C - AMMONIUM HYDROxIDE
-
D GASEOUS AMMONIA
70

5O

50

40

30

O l | | | |
| 2 5 IO 2O 5O |OO
DIAMETER IN MICRONS
FIG tºRE 3.12. I’ (), Particle Size IDistribution : Effect of Precipitant [2].

was less particle breakdown at higher reduction temperatures. This


suggests that the UO, particles are fractured, reduced UAO, particles
whose strains anneal out at higher temperatures [1,2].
The effect of comminution procedures on the particle size distribu
tion of UO2 prepared from ammonium diuranate is shown in Fig.
3.13. Although the as-prepared powders differed appreciably in
particle size distribution, the milled powders are rather similar. The
influence of wet and dry milling, the amount and type of grinding
media, the milling time, and the grinding processes have been studied,
and little difference in the particle size distribution of the milled
powders was observed [2, 11].
110

IOO

90 H

8O H.
URANIUM DIOXIDE:

-
PROPERTIES AND NUCLEAR APPLICATIONS

70 H

60 H.

50 H.

4OH

A - Mcw PRECIPITATED UO2 -


B - UREA-PRECIPITATED DIURANATE UO2
C - "B" MILLED -
D - "D"MILLED
l l l l
I 2 5 lo 2O 50 iOO
DIAMETER IN MICRONS

FIGURE 3.13. UO, Particle Size Distribution : Effect of Milling Procedure [2].

3.2.5 Crystallite Size

A crystallite
is defined as the smallest unit of a solid which coher
ently scatters X-rays [1]. X-ray line broadening is probably the best
method for the determination of crystallite size, but it may also be
possible to determine crystallite sizes with the electron microscope
[17]. To a first approximation, the crystallite size of UO, powders
depends more on the temperature than on the method of prepara
tion [1, 2]. As illustrated in Fig. 3.14, the crystallite size may
vary from preparation to preparation at any one temperature, but
markedly increases at the higher temperature. Heating UO, samples
above their temperature of preparation can cause crystallite
growth [1]. The crystallite size of UO, was found to remain
essentially constant throughout a number of successive oxidation and
reduction cycles in which the original temperature of preparation was
not exceeded [28]. Part of the X-ray line broadening of some UO.
samples is thought to be due to strain (see Chap. 9) [43].

3.2.6 Optical Properties

(a) Color

The color of uranium dioxide varies from brown to black [5].


In general, UO,
powders prepared by hydrogen reduction procedures
are brown, while those made by steam oxidation of uranium are black.
However, heating the latter preparations in hydrogen causes their
color to change from black to brown [2]. Also some black UO.
CHARACTERIZATION OF URANIUM DIOXIDE 111

10,000 i T- I I -T

IOOOH- -

too § I | | I
4OO 5OO 6OO 7oo 800 900
TEMPERATURE OF PREPARATION, “c

FIGURE 3.14. Crystallite Size of UO, Prepared at Various Temperatures [1].

powders prepared by hydrogen reduction have been reported [2].


Oxidation will
convert brown UO, into a black product.
The color of UO, also varies with its particle size [3]. The
gradation of color with increase in particle size is given in Table 3.7.
The black color of the small particle size oxide is probably due to
reoxidation.

TABLE 3.7—COLOR OF UO, [3]


Particle
diameter*
Color (microns)
Black -------------------------------------- 0.05
Light brown-------------------------------- ca. 0.1
Brown------------------------------------- ca. 0.3
Dark brown-------------------------------- ca. 0.8
Red brown--------------------------------- ca. 1.0
Blood red----------------------------------- ca. 10.0

"Determined by gasadsorption measurements.

(b) Spectra •

The absorption spectra of solid uranium dioxide and UO,-Tho,


solid solutions have been measured in the 2,100 to 8,000 angstrom unit
region [44–46]. Several discrete bands with absorption maxima were
observed. Five different UO, samples were examined for infrared
*See Sect. 5.5.4 of Chap. 5.
112 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

spectra in the 2- to 16-micron region, and all gave nearly identical


spectra, showing no absorption other than that of water [47].

(c) Indea of Refraction

Since unstrained UO, crystallizes in the isometric (cubic) system,


it is isotropic and has only one index of refraction. Sulfur-selenium
reference melts and a lithium light source are required for measuring
the refractive index of UO2. The index of refraction of large-grained
UO, has been reported to be 2.35, but uranium oxide films with refrac
tive indices as low as 1.93 have been observed [6, 48].
The index of refraction of UO, films evaporated onto fused silica
varied from 2.29-4-0.03 at 8,000 Å to a maximum of 2.58+ 0.22 at
4,500 Å [46]. Some correlation was noted between the UO, hydrogen

reduction temperatures and the refractive indices of the resulting


UO." At reduction temperatures of 500°, 700°, and 900°C, the UO.
powders had refractive indices of 2.24, 2.26, and 2.28, respectively.”
At Oak Ridge National Laboratory the same refractive index (2.35)
was obtained for UO, prepared by reduction at both 1,000° and 1,750°
C. Because of the inherent lack of precision in the range above 2.0,
refractive index measurements are not too good a method of charac
terizing UO, powders.

3.3 CHEMICAL CHARACTERISTICS

The chemical characteristics of uranium dioxide are described in


Sects. 3.3.1 through 3.3.3. These characteristics include nonstoichi
ometry, oxidation resistance, and chemical reactivity of UO2.

3.3.1. Nonstoichiometry"

The ability of uranium dioxide to deviate readily from the stoichio


metric UO.co composition accounts for variations in some of its phys
ical properties. The U/O ratio of UO can be accurately determined
by several methods. The oxygen content can be determined directly
by reaction with bromine trifluoride, or the oxygen in excess of compo
sition UO, on can be measured by reduction with hydrogen or carbon
monoxide, or by a thermogravimetric oxidation-reduction technique
|49–53]. Oxygen to uranium ratios can also be calculated from total
uranium, U (VI), and U (IV) analyses [53–55].

* See Sect. 5.5.4 of Chap. 5.


* A. G. Allison and E. A. Durbin, personal communication.
* See Chap. 6.
CHARACTERIZATION OF URANIUM DIOXIDE 113

3.3.2. Oxidation Resistance

(a) Pyrophoricity and Storage Stability

Uranium dioxide powders prepared by steam oxidation methods


are usuallynot pyrophoric unless the starting uranium metal is very
finely divided [2, 6]. In the latter case, heating the UO, in hydrogen
at elevated temperatures Even low temperature re
will stabilize

it.
the higher uranium oxides (UO,-2H2O, UO, H2O, UOs,
of

duction
by

U.O.) prepared the ignition uranyl nitrate, uranium metal,

of
or or

some other nonprecipitation procedure generally results non

in
a
pyrophoric UO, [1, 2]. Uranium dioxide preparations made from
uranium precipitates (ammonium diuranate, uranium peroxide,
by

by
uranyl oxalate, etc.) either first pyrolyzing

of
direct reduction
UO, UAO, with hydro
or

precipitates air reducing


in

to

the and then


ethyl alcohol are frequently pyrophoric [1–4, 56]. These oxide
or

gen
by

higher

or
powders can using longer reduction periods
be

stabilized
[1,2].
or

reduction temperatures, both


Since the reactivity UO, preparation toward oxygen func
of

is
a

a
particle size rather than crystallite size
or
density, processes
of

tion
small UO, particles, such low temperatures prep

of
as

which result
in

even grinding, favor the production pyrophoric oxide


of
or

aration
[1–3]. Oxygen adsorption, extent proportional
an
to

to

the surface
reduced, degassed sample UO, exposed
of
as

as

area, occurs soon


is
a

air [1]. The process


exothermic and, the particles are small
to

if
is
so

that the heat evolved per unit volume the sample due
of

enough to
oxygen adsorption higher than can the surroundings, the
be

to

lost
is

temperature the particles rises and the reaction proceeds faster.


of

Larger UO,
particles are not pyrophoric, but the O/U ratio increases
steadily with time exposure oxygen.
of

to

UO, under storage


of

of

Studies have been made the air oxidation


[1, 58]. Uranium dioxide powders with large sur
3,

conditions 57,
rapidly
do

face areas oxidize more and extensively than those with


the susceptibility UO,
of

small ones [57, 58]. This dependence


of

to

atmospheric oxidation upon the particle size illustrated Fig.


in
to is

particle fairly
of

Above diameters about 0.2 0.3p, U.O.,


is

3.15.
oxidation, but
of
up to

at

to

stable diameters about 0.05 0.08p, U.O. can


appreciable oxygen [3, 26]. has also been ob
of

amounts
It

take
served that UO,
oxidizes more rapidly hot, humid weather [58]."
in

However, other factors influence the storage stability UO. The


of

UO, increase rapidly with rising tem


of

of

rate and extent oxidation

One study
of

of

the effect moisture indicated that water has no effect on the oxidation
*

Tº), at 60° and 80° However, other workers have reported that moisture
of

[59].
C

accelerates the oxidation of UO2 at 113° air and at 180° oxygen (see
in

in
C

Chap. [60, 61).


8)
114 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I I I I I I i. I T

2.7 H.
-
H---U30s

©—l l I --O- 1
I.O
``O O. o.2 o3 O.4 O.5 O.6 O.7 O.8 O.9
PARTICLE DIAMETER, d (, )
FIGURE 3.15. Relation between Particle Size and Oxygen Content of Uranium
Oxides [3, 26].

perature; the extent of oxidation increases while the rate of oxidation


decreases with time; and oxidation depends on access of air to the ox
ide [57, 58]. The amount of oxygen in contact with the UO, deter
mines the extent and probably the rate of oxidation. Because of these
factors, optimum handling conditions require storing UO, in airtight
containers at a temperature of 25°C or less.

(b) Thermogravimetric and Differential Thermal Analyses


Thermogravimetric and differential thermal analysis (DTA) data
have been found useful in characterizing UO, as to oxidation resist
ance, in detecting higher oxides, in calculating O/U ratios, and in esti
mating the amount of adsorbed water [2, 6, 7, 13, 52,62, 63]. Figures
3.16 and 3.17 show some DTA weight-loss data for two UO, powders.
A slightweight loss occurs around 100° C from the evolution of water.
The amount of adsorbed water and the initial O/U ratios calculated by
thermal techniques are in fair agreement with the values obtained by
chemical analyses. On further heating, both the DTA and weight
change curves indicate two exothermic reactions corresponding to
the oxidation UO2–UAO,->UAOs. For UO, powders of small particle
size, the IOTA peaks occur at lower temperatures and with a higher
degree of separation than for oxides having small surface areas which
react more slowly at higher temperatures. An additional intermediate
exothermic DTA peak, observed in some cases, has been attributed to
disproportionation of the U.O., into UO,--U.O., followed by rapid
oxidation of the UO, to U.O, [13, 63].
CHARACTERIZATION OF URANIUM DIOXIDE 115

+0040 - -

10020
-

+OOH
o - -

SL
2OO 6OO 800 IOOO
O Hº-> I l
TEMPERATURE,”C

-OOOH-
-

l I I | | | l
-
-002O
O 15 3O 45 6O 75 90 IO5
TIME, MINUTES

-0030

FIGURE3.16. Ammonium Diuranate Precipitated UO. : DTA and Weight Change


from UO, to U20, [6, 7, 62].

+ OO60

+ 0.050 H. -
+ 0.040 H. -
+ 0.030 - -
-
.
* 0.02OH. -
;
º

-
- + 0Oio H. -
#
2OO 4OO 6OO 80O IOOO
#
O L L I l
TEMPERATURE,°C

~ 000 H. -

- Oozo I l l ! l !
O 15 30 45 6O 75 90 iO5
TIME, MINUTES
- OC50

FIGURE 3.17. Fused UO, ; DTA and Weight Change from UO, to U.O.
[6, 7, 62].

57.4789 0–61–9
116 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

3.3.3 Chemical Reactivity

(a) Hydrofluorination

The hydrofluorination of UO, is its most important chemical re


action, since the product UF, is a key intermediate for the production
of both uranium metal and uranium hexafluoride. The chemistry of
the hydrofluorination reaction has been and continues to be the subject
of extensive investigations. Large-scale hydrofluorination operations
have been harassed by a variety of difficulties resulting from occa
sional poor yields of UF. The factors that influence the rate and
extent of the reaction between UO, and HF have been found to be
complex and are dependent on the previous history, surface area, and
state of agglomeration of the UO, as well as the time, temperature, and
gas concentration [3, 64–68]. Initially, the rate of conversion of UO,
to UF, is relatively rapid, then becomes slower as the extent of con
version increases [69]. Both the rate and extent of hydrofluorination
are temperature dependent, increasing up to a certain optimum tem
perature and then frequently decreasing with further increase in
temperature [68, 69]. Decreasing the size of the UO, charge and
using thinner layers and a rippled surface of UO, as well as a smaller
UO, particle size result in a marked increase in the rate of conversion
[65, 69]. The rate of reaction has been found to depend in an im
portant way upon the UO, preparation temperature. Figure 3.18
IOO
| I

80 H. -


g so
U03 REDUCED AT
A—450 °C -
uu B-540° C.
5 C-780° C.
(HYDROFLUOR INATION TEMP 575° C)
or

$40
-
->

20

O
|
40
|
8O
l
120
|
160
|
2OO
|
240 280
TIME IN MINUTES

FIGURE 3.18. Effect of Reduction Temperature on Hydrofluorination Rate [64].


(Courtesy, American Chemical Society.)
CHARACTERIZATION OF URANIUM DIOXIDE 117

shows the rate of hydrofluorination of UO, made by hydrogen re


duction of UO,
at several temperatures. The higher the reduction
temperature, the slower was the hydrofluorination process. This
dependence of UO, reactivity on reduction temperature can be cor
related with its surface area since, as was illustrated in Figs. 3.8
and 3.9, the UO, surface area has been found to decrease as the reac
tion temperature increases [1,2].
The particle size and the degree of agglomeration of the parent
higher oxide influence the specific surface, state of aggregation and,
hence, the hydrofluorination reactivity of the UO, product. Several
contrasting types of behavior observed during the UO, hydrofluorina
tion reaction have been explained on the basis of distinct powder types
which differ in the degree of aggregation of the intrinsic particles
{3,64].

One type of UO, consists of large, closely packed aggregates of very


small particles. This material is reactive and hydrofluorinates readily
at lower temperatures, but the extent of the reaction starts to decrease
at temperatures above 500° C. A second type of uranium oxide is
composed of loose aggregates of a smaller number of larger individual
particles and, although the large particle size considerably reduces
UO, hydrofluorination,
the

in of

the degree
of

rate conversion increases


with rise temperature until complete conversion occurs around 650°
hydrofluorination decreases.
of

Above the degree This


C.

650°
higher temperatures complete
in

to

effect, which are unfavorable


hydrofluorination, called thermal damage, since the properties of
is

reacting solids are altered high temperatures prejudice


the

at

to

completion
of

the reaction [64].


Thermal damage has been interpreted competition
of

terms
in

between two rate processes: the chemical reaction between UO, and
HF and the sintering the product UF, [3, 64]. As UO, particles
of

HF, outer layer UF,


an

formed, and the subsequent


of

react with
is
by

the HF gas
of

of

influenced the relative rate diffusion


is

reaction
through this UF, layer. high initial reactivity will result high
in
A

particle temperatures which promote sintering the UF, layer. If,


is of

addition, the state particle aggregation considerable, large,


of
in

relatively highly sintered UF, particles are produced, often before


impervious coating
an

complete.
of

of

reduction The formation


is

UF, trap
an

appreciable fraction unreacted UO.


of
to

sintered tends
Microscopic such products
of

within the sintered shell. examination


shows very clearly particles with green, smooth, glassy exteriors and
brown unhydrofluorinated interiors [3]. Larger individual particles
clustered into smaller aggregates have less tendency sinter together,
to

sufficiently
of
at

since the rate which the heat reaction evolved


is

is

into the surroundings. Thus, al


so

slow, may dissipated


be

that
it
118 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

though the rate of hydrofluorination for such UO, powders is slower


at lower temperatures, the predominating reaction at higher tempera
tures is now hydrofluorination rather than sintering, and it is not until
about 650° C that the sintering process prevails.
Thermal damage may be minimized by carrying out the hydro
fluorination reaction in such a way that the bulk of reaction occurs
before particle temperatures sufficient to cause sintering are attained.
This can be done by programming either the reaction temperature or
the composition of the hydrofluoric acid-water reaction gas [64]. Pre
early stages, thus,

its
sumably, programming slows the reaction in
allowing the dissipation

of

of
more the heat evolved.
Some evidence has been obtained that different crystalline forms

of
the parent UO, resulting from variations its preparation may part

in
UO, [64, 66]. At least
ly

govern the hydrofluorination reactivity

of
the five distinct X-ray diffraction patterns have been reported for UO,
(see Chap. [5, 64]. The UO, produced commercial scale denitra
6)

in
very wide variation reactivity, de

of
tions has been found
to

have
a

pending upon the manner which the denitration was carried out
in

denitration, the presence

of
[64, 66, 70]. Rate and temperature of
impurities, and the type important in

an an
of

equipment used have


the UO, properties. The possibility inhibiting film
on

of
fluence
the UO, surface has been suggested account for
on

being present

to
poor hydrofluorination reactivity [66]. Experiments

on
the activation
UO, particles which displayed low reactivity towards hydrofluori
of

nation showed that partial dissolution the particles nitric acid


of

resulted some activation, and that successive low temperature oxi in


in

dation and reduction produced highly activated UO, [28, 66].


UO, with UF, has been investigated the tempera
to of

The reaction
in

Although the reaction products

at
ture range 25° 500° [71].
C

500° are UO.F. and UF, the reaction apparently proceeds through
C

pentavalent oxyfluorides.
of

the formation unstable

(b) Other Halogenation Reactions

Katz and Rabinowitch have extensively reviewed the literature


UO, halogenation reactions [5]. The chlorination
of on

of

number
a

types UO, has been studied some detail [5, 72, 75,
of

different
in

by

76]. reactive UO.,


suitable for preparing UCl, vapor phase
A

chlorination with CC1, SOCl, obtained from uranium peroxide


or

is

[72, 74]. The chlorination reactivity the UO, largely dependent


of of

is

the preparation
on

the method used the uranium peroxide source


in

material. As was found with the hydrofluorination reaction, U.O.


prepared lower temperatures chlorinated better than oxide made
at

higher reduction temperatures [73]. Several processes for pro


at

ducing UCl, from UO, have been patented [77–79].


CHARACTERIZATION OF URANIUM DIOXIDE 119

(c) Dissolution Reactions

Reagents which might be effective in dissolving UO, are of con


siderable interest, since the dissolution of UO, is the usual first step
its

in chemical reprocessing. The dissolution chemistry concerned

is
with reaction rates for UO, various solvent systems. The particle

in
as

as
the temperature
of

of
size the UO2 well reaction and concen
the dissolving agent affect the rate The develop
of

tration data.
adequate dissolving techniques for
of

ment nuclear reactor fuel


the acquisition substantial body

of

of
in
elements has resulted

a
chemical and engineering data [80]. Both batch pot and continuous
dissolution techniques and equipment have been successfully developed.
Early workers studied the solubility UO,
of
various acids. Near

in
by

solubility weight acid, HBr, HCl, H.S.O., aqua


in

18°C the acetic


HNO,
were reported UO, acid, 1:12000, 1:4650,
be
to

regia, and

;
13100, 1:2200 (66° Bé), 1:29.6, and 1:8 (36° Bé) [81]. Better dis
HCl was obtained by using longer reaction times [82].
in

solution
Cold, dilute HNO, hydrolytic effect upon
to

was claimed have


a

UO, [83–85].
Uranium dioxide dissolves aqueous HF forming UF,
is in

insoluble
UO, particle
on

dependent size,
of

[5]. The rate dissolution the


acid concentration, and temperature. Hot concentrated HaPO, dis
UO, slowly but completely [86]. Fusion with KHSO,
or
solves the
HNO3–HF mixture will also dissolve UO, [86].
In

general,
of

use
a

almost any strong oxidant (H2O, KMnO,) combination with an


in

dissolving UO, [87]. Because


be

acid will
to

effective some extent


in

low solubility
aqueous solutions, the addition var
of
its

in

to

ozone
of

ious reagents was not effectiveenhancing the solubility UO, [87].


of
in

Uranium dioxide dissolves readily aqueous alkaline peroxide so


in

Na2O, H2O, with NaOH, K.COs,


of
as

lutions such and mixtures


(NH,)2CO3, and (NH,)2C2O, [5, 87]. The kinetics
of

the dissolu
UO, Na2CO3–NaHCO, and H2SO, solutions have been
of

in

tion
investigated area, oxygen
of

88–91]. The effects the surface


|

partial pressure, temperature, reagent concentration, and pH were


determined.

REFERENCES

ANDERson, MooRBATH, and Roberts, “The


A.

HARPER,
S.
E.

E.
L.

J.
1.

S.
J.

Properties and Microstructure Uranium Dioxide; Their Dependence


of

Upon the Mode Preparation,” AERE C/R 886, Aug. 19, 1952.
of

ARonson, “Some Preparative Methods and Physical


C.

CLAYTON and
2.

S.
J.

Uranium Dioxide Powders,” Chem. and Eng. Data


6,
of

Characteristics
J.

43–51 (1961).
M. GILLIEs, “The Conversion Uranyl Nitrate
G.

ListER and
B.

to
A.

of
J.
3.

Uranium Dioxide and Uranium Tetrafluoride” “Progress Nuclear


in

in
to

Energy, Series III, Process Chemistry,” Bruce, M. Fletcher,


H.
R.

H.
F.

J.
120 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Hyman, and J. J. Katz, eds., pp. 19–35, McGraw-Hill Book Co., New York,
1956.
4. L. C. WATSON, “The Production of Uranium Dioxide for Ceramic Fuels,”
CRL–45, Nov. 20, 1957.
5. J. J. KATz and E. RABINow Itch, “The Chemistry of Uranium,” National Nu
clear Energy Series, Div. VIII, Vol. 5, McGraw-Hill Book Co., New York,
1951.
J.
R. Johnson, S. D. FULKERson, and A. J. TAYLOR, “Technology of Uranium
Dioxide, A Reactor Material,” Am. Ceram. Soc. Bull. 36, 112–117 (1957).
. (a) J. R. Joh Nso N, and C. E. CURTIs, “The Technology of UO, and Tho..”.
in “Proceedings of the International Conference on the Peaceful Uses of
Atomic Energy, Geneva, 1955,” Vol. 9, pp. 169–173, United Nations, New
York, 1956.
(b) J. R. Johnson, “Uranium Dioxide” in “Progress in Nuclear Energy,
Series V, Metallurgy and Fuels,” Vol. 2, J. P. Howe and H. M. Finniston,
eds., pp. 209–222, Pergamon Press, New York, 1959.
S. BRUNAUER, “The Adsorption of Gases and Vapors,” Oxford University
Press, London, 1943.
J. WILLIAMs, E. BARNEs, R. W. THACKRAY, and P. MURRAY, “The Nonstoi
chiometric Oxides of Uranium.” Trans, Brit. Ceram. Soc. 57, 608–623
(1958).
10. F. GRoN vold, “High-Temperature X-ray Study of Uranium Oxides in the
UO2–UAO, Region,” J. Inorg. d: Nuclear Chem. 1, 357–370 (1955).
11. J. C. CLAYTON and S. ARONSON, “The Effects of Mechanical and Thermal
Processes on Some Physical Characteristics of UO, Powders,” Bettis Tech
nical Review, WAPD—BT-15, Sept. 1959, pp. 50–60.
12. J. C. CLAYTON and S. ARoNsoN, “Densities of Uranium Oxide in the Region
UO,-U.O,” Bettis Technical Review, WAPD—BT—10, Oct. 1958, pp. 96–100.
13. P. MURRAY and D. T. LIvey, “The Technology of Urania and Thoria” in “Prog
ress in Nuclear Energy, Series V, Metallurgy and Fuels,” H. H. Finniston
and J. P. Howe, eds., pp. 448–478, McGraw-Hill Book Co., New York, 1956.
14. J. BELLE and B. LUsTMAN, “Properties of Uranium Dioxide” in “Fuel Ele
ments Conference, Paris,” TID–7546, Book 2, Mar. 1958, pp. 442–515.
15. J. D. EICHENBERG, P. W. FRANK, T. J. KIsIEL, B. LUsTMAN, and K. H. Voge:L.
“Effects of Irradiation on Bulk Uranium Dioxide” in “Fuel Elements Con
ference, Paris,” TID–7546, Book 2, Mar. 1958, pp. 616–716.
16. G. H. CHALDER, N. F. H. BRIGHT, D. L. PATERSoN, and L. C. WATsoN, “The
Fabrication and Properties Dioxide Fuel” in “Proceedings of
of Uranium
the Second United Nations International Conference on the Peaceful Uses
of Atomic Energy, Geneva, 1958,” Vol. 6, pp. 590–604, United Nations,
Geneva, 1958.
F. D. LEIPzIGER, “The Particle Size of Some Uranium Dioxide Samples.”
KAPL-1743, Mar. 8, 1957.
W. A. OPPOLD, “A Report on the Examination of Uranium Dioxide with the
Electron Microscope,” NYO-5195, Aug. 27, 1946.
F. A. SMITH and C. S. LEEs, “Electron Microscope Studies of Particle Sizes
and Shape in Uranium Oxide,” AERE G/M 136, June 24, 1952.
D. R. STEN QUIST,B. MASTEL, and R. J. ANICETTI, “Correlation of Surface
Characteristics with the Sintering Behavior of Uranium Dioxide Powders.”
HW-51712, July 25, 1957; J. Am. Ceram. Soc. 41, 273–274 (1958).
21. E. A. Eva Ns, “Fabrication and Enclosure of Uranium Dioxide,” HW-52729.
Sept. 18, 1957.
CHARACTERIZATION OF URANIUM DIOXIDE 121
22.

“Preparation and Properties


in D. HARRINGTON,

of
Uranium Dioxide Powder"
C.

“Fuel Elements Conference, Paris,” TID–7546,

2,
Book Mar. 1958, pp.
302–365.
KIEssli Ng and RUN FoRs, “Sintering

of
Uranium Dioxide” “Fuel

U.
R.

in
23.
Elements Conference, Paris,” TID–7546, Book

2,
Mar. 1958, pp. 402–413.
ANICETTI, “Fabrication Behavior Some Uranium
R.

R.

of
J.
D.

24. STEN QUIST and


Dioxide Powders,” HW–51748, Dec.

1,
1957.
BERRIN, “A Study the Sintering Behavior

of

of
CLAYToN and L. Some
C.

25.
J.

Uranium Dioxide Powders,” Bettis Technical Review, WAPD—BT-20,


Sept. 1960, pp. 23–38.
the Study Sintering
in Y.

to
CARTERET, “Contributions

of

of
BEL and Uranium
A.

26.
Dioxide” “Proceedings the Second United Nations International
of
Conference on the Peaceful Uses of Atomic Energy, Geneva, 1958,” Vol.

6,
pp. 612–619, United Nations, Geneva, 1958.
SCHONBERG, U. RUNFORs, and R. KIEssli NG, “The Sintering of Uranium
N.

27.
Dioxide” “Proceedings the Second United Nations International
of
in

Atomic Energy, Geneva, 1958,” Vol.

6,
of

Conference on the Peaceful Uses


pp. 605–611, United Nations, Geneva, 1958.
28.

BUNKER, R. GREENough, H. KALMUs, and BARD, “The Activa


C.

R.
L.

J.
E.
D.

tion of Low-Reactivity Uranium Dioxide Particles,” LA—1952, Oct. 1955.


W. LAUTERBACH, LAsKIN, and LEACH, “Specific Surface Determina
L.
K.

S.

29.
by Low Temperature Adsorption Ethane,”
of

of
tions Uranium Dusts
Franklin Inst. 250,
(1950).
J.

13-24
RULLI, “Measurement of Low Surface Area of the
C.

E.

CLAYTON and
J.

30.
J.

Uranium Oxides by the Innes Method,” Chemist-Analyst 47, 62–64 (1958).


M. McCoNNELL, “Small-Scale Cyclone Micronizer,” AERE C/M 319,
D.

31.
J.

Sept. 1957.
32.

ARonson, CLAYTON, RULLI, “Surface Areas


T.
C.

R.

PADDEN, and
E.
J.

J.
S.

Sintered UO, Compacts,” Bettis Technical Review, WAPD—BT-19,


of

June 1960, pp. 83–92.


LANGER, “Specific Surface Determination Uranium Oxide Powders by
A.

of

33.
Dye Adsorption,” Westinghouse Research Report 100FF1009–R1, Sept.
21, 1956.
Roberts, Adsorption and Absorption Gases by Uranium
of
J.
E.

“The
L.

34.
Dioxide,” AERE C/R Mar.
5,

887, 1953.
Roberts, “The Oxides The Chemisorption
of

Part
V.
J.
E.

35. Uranium.
L.

of Oxygen on UO, and UOz–Tho, Solid Solutions,” Chem. Soc., 3332–3339


J.

(1954).
36.

Roberts, “The Surface Chemistry


D.

M. McCoNNELL and
in L.

of
E.

Uranium
J.
J.

and Thorium Oxides” “Chemisorption: Proceedings of Symposium


a

held at the University College North Staffordshire, Keele, Staffordshire,


E. of

by the Chemical Society,” W. Garner, ed., pp. 218–226, Academic Press,


New York, 1957.
37.

Roberts, “The Oxides Uranium. VI. The Chemisorption Reduc


of

of
E.
L.

J.

ing Gases on Uranium and Thorium Dioxides,” Chem. Soc. 3939–3946


J.

(1955).
38.

M. McCoNNELL, “Heat Adsorption Oxygen


of
at D.

on
of

Uranium Dioxide
J.

–183° Part Measurements with Liquid Oxygen Calorimeter,”


C.

I.

AERE C/R 1484, Sept. 16, 1954.


M. McCoNNELL, “Heat Adsorption Oxygen on
of
D.
F.

of

FERGUson and
J.

39.
I.

Uranium Dioxide –183°,” Proc. Roy. A241, 67–79 (1957).


at

Soc.
M. McCoNNELL, Adsorption Oxygen
of
D.

on
of

of

“Heats Solid Solutions


J.

40.
Uranium Dioxide Thorium Dioxide –183°,”
at

J.

Chem. Soc. 947–950


in

(1958).
122 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

41. A. C. WAHL and N. A. Bon NER, eds., “Radioactivity Applied to Chemistry,”


John Wiley and Sons, New York, 1951.
42. F. H. GUNZEL, JR. and W. A. LAMBERTson, “Urania Bodies: Fired Density
vs. Particle Size,” ANL-5094, Aug. 1954.
43 . E. R. Boyko, J. D. EICHENBERG, R. R. Roof, JR., and E. K. HALTEMAN, “X-ray
Examination of Irradiated Uranium Dioxide,” Bettis Technical Review,
WAPD—BT—6, Jan. 1958, pp. 64–74.
D. M. GRUEN, “Absorption Spectra and Electrical Conductivities of UOr
Th9, Solid Solutions,” J. Am. Chem. Soc. 76,2117–2120 (1954).
. R. J. AcKERMAN, “The High Temperature, High Vacuum Vaporization and
Thermodynamic Properties of Uranium Dioxide,” ANL–5482, 1955.
46. R. J. AckerMAN, R. J. THoRN, and G. H. WINSLow, “Visible and Ultraviolet
Absorption Properties of Uranium Dioxide Films,” J. Opt. Soc. A m. 49,
1107–1112 (1959).
47. J. H. LADY and R. E. AdAMs, “Infrared Spectra of Uranium Compounds and
the Characteristic Frequencies of the Uranyl Ion,” Westinghouse Research
Report 6–94602–9–R12, July 31, 1957.
48. A. E. STEBBENs and L. L. SHREIR, “Refractive Index of Uranium Oxide Pro
duced by Anodic Oxidation,” Nature 183, 1113–1114 (1959).
49. H. R. Hoekstra and J. J. KATZ, “Direct Determination of Oxygen in Less
Familiar Metal Oxides,” A mal. Chem. 25, 1608–1612 (1953).
. L. E. J. Roberts and E. A. HARPER, “The Determination of Oxygen in Uranium
Oxides,” AERE C/R 885, May 5, 1952.
. N. F. H. BRIGHT, L. G. RIPLEY, J. F. Row LAND, and R. H. LAKE, “The Deter
mination of the Oxygen-Uranium Atomic Ratio in Nonstoichiometric
Uranium I) ioxide and Other Oxides of Uranium,” MID–207, July 5, 1956.
. L. G. Ston HILL, “The Determination of Atomic Ratios in the Uranium-Oxygen
System by a Thermogravimetric Method,” Can. J. Chem. 37, 454–459
(1950).
. C. J. Roddex, ed., “Analytical Chemistry of the Manhattan Project,” National
Nuclear Energy Series, Div. VIII, Vol. 1, McGraw-Hill Book Co., New
York, 1950.
. R. M. BURd and G. W. Gow ARD, “The Polarographic Determination of Hex
avalent Uranium in Uranium Oxides : The Determination of Oxygen Ura
nium Ratios,” WAPI)—205, Apr. 1959.
. J. MIN czEws KI, R. PRzYTYCKA, and J. Koh MAN, “Potentiometric Deter
mination of Small Amounts of Hexavalent Uranium in Uranium Dioxide."
Chem. Anal. (Warsaw) 3, 27–32 (1958).
. B. M. HAINEs and V. P. CALKINs, “The Reduction of Uranium Trioxide by
Ethyl Alcohol,” AECI)—4006, Nov. 18, 1946.
. J. STEvex son and J. Boyd, “The Oxidation of UO, in Air and under Conditions
of Storage,” NYO-5224, Apr. 1, 1948.
58. G. L. MARTIN, “The Deterioration of U (), in Storage,” NYO-5228, May 1, 1948.
59. S. M. Hurston, “A Study of Water in Brown Oxide [UOs] and Its Influence
on the Oxidation of that Material." N YO-5222, Mar. 11, 1948.
. S. M. LANG, F. P. KN Upsr. N. C. L. FILLMoRE, and R. S. Roth, “High Tempera
ture Reactions of Uranium Dioxide with Various Metal Oxides,” Natl.
Bur. Standards Circ. 568, Feb. 20, 1956.
. D. A. WAUGH AN, J. R. BRIDGE, and C. M. SchwARtz, “Comparison of Active
and Inactive Uranium Dioxide-Oxygen System.” BMI–1241, Dec. 10, 1957.
. J. M. WARDE and J. R. Jon Nsox, “Recent Developments in the Technology of
Ceramic Materials for Nuclear Energy Service,” J. Franklin Inst. 260,
455–466 (1955).
CHARACTERIZATION OF URANIUM DIOXIDE 123

63. P. MURRAY and R. W. THACKRAY, “Differential Thermal Analysis of the


Uranium Oxides,” AERE–M/R—632, Jan. 1951.
64. C. W. KUHLMAN, JR., and B. A. Sw1NEHART, “A Thermal Damage Effect in
the Production of Uranium Tetrafluoride,” Ind. Eng. Chem. 50, 1774–1776
(1958).
65.

MILLER, “The Reaction Uranium Dioxide with Hydrogen Fluoride

of
E.
to I.

Produce Uranium Tetrafluoride and Related Topics,” MCW-24, Apr.

1,
1946.
WILLsoN, “The Reactivity Uranium Oxides,” NYO-1466, Feb. 15, 1952.
R. K.

of
E. S.

66.
BoNFER, and Abbott, “A Kinetic Study the Re
C.

C.
R.
DEMARCO, D.

of
67.
action, U.O.--4HF-> UF,4-2H.O. Part Development the Rate Equa

of
I.
tion,” NLCO-675, Apr. 18, 1957.
W. KUHLMAN, “Reaction Rate

of
the Fluorination

at
of
Uranium Dioxide
C.

68.
Continuously Increasing Temperature,” MCW-121, July 19, 1948.
a

W. KEw FRY, “Effects Certain Variables upon the Reac


R.

of
E.

and O.
Is

o!}.
H

tions for Preparation, by the Dry Method, UF, from U.Os,” LA—1860,

of
Nov. 1954.
Holden, “Factors Influencing the Reactivity Uranium Trioxide,”
R.

of
B.

70.
TID–5063, Oct.
9,

1951.
A. RAMPY, “The Reaction of Uranium Dioxide with Uranium Hexafluor
G.

71.
ide,” GAT-265, June
5,

1959.
KRAUs, “Preparation of Uranium Tetrachloride by Vapor Phase Chlor
A.
C.

72.
ination of Uranium Dioxide Formed by Reducing Uranium Trioxide with
Ethanol,” A-2313, July 12, 1945.
KRAUs, “Preparation Uranium Tetrachloride by Vapor Phase Chlor
of
A.
C.

73.
ination of Uranium Dioxide Formed by Reducing Uranium Trioxide with
Ethanol, Report No. II,” A–2321, Sept. 17, 1945.
KRAUs, “Factors the Precipitation UO, which Influence the State
A.

of
in
C.

74.
of the Resulting Oxide Particularly with Respect its Reactivity
to

the in
Vapor Phase Reaction,” A-2314, July 19, 1945.
HARRIson, “Preparation Uranium Tetrachloride,” AERE–GP/R–2409,
of
R.
E.

75.
-
1951.
76.

Volsky, “The Chlorination


N.
V.

A.

of

BUDAYEv and Uranium Dioxide and


I.

Plutonium Dioxide by Carbon Tetrachloride” “Proceedings of the Second


in

United Nations International Conference on the Peaceful Uses of Atomic


Energy, Geneva, 1958,” Vol. 28, pp. 316–330, United Nations, Geneva, 1958.
77.

Producing Uranium Tetra


of
H.
R.

WAGNER, “Method
E.

McCoM BIE and


L.

chloride,” U.S. Patent 2,735,746, Feb. 21, 1956.


78.

“Process for Producing Uranium Tetrachloride,” British Patent 812,121, Apr.


22, 1959.
79.

“Process for Troducing Uranium Tetrachloride,” British Patent 812,792, Apr.


29, 1959.
80.

Foster, “Nuclear Reactor Fuel Dissolution”


G.

D.

in
R.

WYMER and
L.

“Progress Energy, Series III, Process Chemistry,” Bruce,


R.

Nuclear
in

F.

ed., pp. 85–96, McGraw-Hill Book Co., New York, 1956.


$1.

RAYNAUD, “Solubility
of

Several Acids,” Compt. rend.


A.

Uranous Oxide
in

153, 1480 (1911).


Col.ANI, “Action
of

Acids on Uranous Oxide,” Compt. rend. 155, 1249


A.

82.
(1912).
83.

CoNINck, “Uranium Compounds,” Bull. classe sci., Acad. roy.


DE
Ö.

W.
Belg., 992 (1908).
Jolibois and Bossuet, “Relations between the Different Oxides
of
R.

84.
P.

Uranium,” Compt. rend. 174, 386–388 (1922).


124 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

85. P. LEBEAU, “Oxides of Uranium,” Compt. rend. 174, 388–391 (1922).


86. R. W. BANE, “Experiments in the Dissolution of Uranium Dioxide,” CC–2670,
Jan. 1, 1945.
87. F. J. Joh Nston,
P. W. WILLs, and J. J. KATz, “Dissolution of Uranium
Oxide Arising from Slug Failure,” ANL–5084, July 1953.
. W. W. SCHORTMANN and M. A. DESESA, “Kinetics of the Dissolution of
Uranium Dioxide in Carbonate—Bicarbonate Solutions” in “Proceedings
of the Second United Nations International Conference on the Peaceful
Uses of Atomic Energy, Geneva, 1958,” Vol. 3, pp. 333–341, United Nations,
Geneva, 1958.
89. R. L. PEARson and M. E. WADsworth, “A Kinetic Study of the Dissolution
of UO, in Carbonate Solution,” Trans. AIME 212, 294–300 (1958).
90. T. L. MACKAY and M. E. WADsworth, “A Kinetic Study of the Dissolution
of UO, in Sulfuric Acid,” Trans. AIME 212, 597–603 (1958).
91. J. L. SWANSON, “Rates of Reaction of Irradiated UO, and Uranium Metal
with Sulfuric Acid,” HW–61482, Aug. 13, 1959.
Chapter 4

FABRICATION OF URANIUM OXIDE


T. J. BURRE AND J. GLATTER, Editors

4.1 INTRODUCTION

Available techniques, both conventional and unconventional, for the


fabrication of fuel and fuel elements containing uranium oxide are
described in this chapter." Fabrication problems for specific applica
tions are only highlighted; process details and illustrative material
are kept to a minimum. References at the end of this chapter provide
detailed information in specific areas of interest.
Three forms of uranium oxide are described: (1) bulk, (2) dis
persion or mixture, and (3) solid solution. A bulk form of uranium
oxide fuel element is one in which the fissile and/or fissionable ma
terial consists of a single component, for example, UO2. A dispersion
type of uranium oxide fuel element is one in which the fissile phase is
dispersed in a material which is nonfissile.” The third type of ura
nium oxide fuel element is one in which the fuel and nonfissile ma

terials form solid solutions of one or more phases.


Conventional ceramic techniques, such as cold pressing, hot press
ing, hydrostatic pressing, slip casting, and extrusion, have been suc
cessfully applied to fabricate both bulk and solid solution forms of
UO. Bulk oxide has also been made by using such nonconventional
all

techniques as swaging and vibratory packing.


Although these
methods have been employed some extent, the only method that has
to

date for large quantity, close tolerance produc


to

proved succesſful
high-density UO, fuel cold pressing and sintering.
of

tion
is

Powder metallurgical techniques have been used combination with


in

hot working metallurgical techniques, such rolling and extruding,


as

form dispersions
of

uranium oxide ductile metallic matrices.


in
to

The conventional ceramic methods mentioned previously are equally


applicable uranium oxide dispersions
of

the fabrication
to

ceramic
in

matrices.

this chapter the general term “uranium include both UO2 (stoichi
to
In

oxide" used
is
*

ometric and nonstoichiometric) and UsOs.


be

This material may, course,


of

fissionable.
*

125
126 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

4.2 BULK FORM URANIUM DIOXIDE


R. J. Anicetti, T. J. Burke,
H. R. Hoge, and B. E. Schaner

4.2.1 Cold Pressing


B. E. Schaner

Various uranium dioxide powder preparations are readily formed


by the conventional cold-pressing methods. The right cylinders, tubes,
wafers, and spheres shown in Fig. 4.1 represent a few of the shapes
which may be produced by cold pressing. Because high-sintered
density and close dimensional tolerances are requisites for some fuel
element components, certain special considerations are necessary in
the fabrication of UO, components by cold-pressing techniques. The
process flow diagram in Fig. 4.2 is a summary of the basic unit
operations required in the production of UO, compacts by cold
pressing.
Uranium dioxide powders, regardless of origin, are normally com
posed of finely divided particles in the form of agglomerates or dis
crete crystallites (see Chap. 3). As prepared, these powders are not
free-flowing; therefore, the powder is usually converted to free-flowing
granular feed material for compacting presses because reproducible
volumetric die fill is essential for the control of the sintered size of cold
pressed shapes. Several binders may be combined with UO2 powders
for conversion to granules. Among these are polyvinyl alcohol, poly
ethylene glycol, camphor, and paraffin [1–5]. These binders may be
incorporated dry, in aqueous solution, in alcohol solution, or in carbon
tetrachloride solution [1–3, 5]. As well as serving as binders, these
materials lend strength to the green compacts for handling and pro
vide a degree of internal lubrication during pressing, thereby min
imizing pressing and ejection defects. The presence of the binder
also decreases the radioactive dust hazard which exists in the process
ing of finely divided uranium compounds.
Polyvinyl alcohol, although an excellent binder, has the disadvan
tage of producing hard granules (when agglomerated) which do not
break down under compacting pressures of less than 50 tsi. As a
result, the granular shape is retained even after sintering and can
produce open porosity as high as 2 percent of the compact volume.
Uranium dioxide granules made with polyvinyl alcohol usually re
quire the addition of a material such as powdered, hydrogenated
poly
oil

vegetable provide lubrication for cold pressing. The


to

up

ethylene glycol binders, when used weight percent -


to

amounts
in

3
be

as

*Open porosity may


to

defined voids connected the surface; the amount open


of
by

porosity usually determined water absorption (see Chap. 3).


is
FABRICATION OF URANIUM OXIDE 127

3 INCHES

FIGURE 4.1. High-Density UO, Shapes Formed by Cold Pressing.

-e
UO2 POWDER

I. GRIND
ORGANIC
w BINDER
2. DRY MIx

WATER
3. WET MIx

4. GRANULATE

5. DRY

6. SCREEN

7. COLD COMPACT

8 co, PRETREAT

9. HYDROGEN
SINTER

y
IO. GRIND SINTERED BODY
TO SIZE

FIGURE 4.2. Process Flow Diagram of Principal Operations for Fabrication of


Cold-Pressed UO, Fuel Components.
128 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

produce a free-flowing granular powder which breaks up readily

tsi
when compacted at pressures above 15 and produces compacts
having

an
excellent surface finish [2].
Two-piece dies composed soft steel shrink ring

of
liner and

a
a
have been found most satisfactory for pressing UO, powders. High
chrome, high-carbon steel hardened Rockwell C60–62 satis

to

is
a
factory die liner material when compacting pressures below

50
tsi
are used. However, die wear becomes excessive after pressing 5,000

to
10,000 compacts. Tungsten carbide die liners and punches, on the
other hand, have shown only slight wear after pressing more than

up
300,000 compacts pressures 125 tsi. Typical dies and punches
at

to
for laboratory manual pressing and punches used for long-run auto
matic pressing are shown Fig. 4.3. Diametral clearance between
die and punches in
maintained between 0.0005 and 0.0015 inch. In
is

corporation taper allow gradual compact expansion


of

to
the die
in
a

during ejection helpful preventing occurrence

of
transverse
in
is

laminations green compacts.


in

in
6
|

FIGURE 4.3. Compacting Die and Punches Used for Cold Pressing: Punches for
Background, Punches for Manual Pressing
to

Automatic Pressing Shown


in

the Right and Left Die, Unsintered and Sintered UO, Wafers Foreground
of

in

[2]. (Courtesy, The American Ceramic Society, Inc.)


FABRICATION OF URANIUM OXIDE 129

A direct correlation exists between compacting pressure and sintered


density for unmilled powders such as MCW UO. As indicated in
Fig. 4.4, compacts formed at successively higher compacting pres
sures will sinter to successively higher densities under identical sinter
ing conditions. This relationship, however, is not evident for the same
powder which is mechanically comminuted. Figure 4.5 illustrates
the comparative behavior of the two powders and shows that the effect
of compacting pressure on sintered density is reduced by mechanical
comminution treatment. Consequently, the milled powder can be com
pacted in the range of 20 to 40 tsi, whereas the unmilled powder re

75
T | |

U02 Powder (Mcw) + 1 wo Pva


70 H 0.400 DIA. RAM, CARBIDE DIE
POWDER FROM BATCH # 5 - 10

65

60

55

50 | | | | L
25 50 75 IOO 125 150 175
LOG PRESSURE, TSI

Figure 4.4. Compact Density as a Function of Compacting Pressure [1].


(Courtesy, American Institute of Mining, Metallurgical, and Petroleum
Engineers, Inc.)

IOO I I i I-T
-01750°c-14 HR
(
*
1- 95
NHAMMER-MILLED McW u02

º
2.
I725°C-IO HR

º
ial
º
Lau
or 90
AS-RECEIVED McW UO2
_c^
-
Lu
2

!---
7,

85 l l I l 1–
IO 20 30 40 50 60 7O 80 90 IOO IIO I2O 130
COMPACTING PRESSURE, TSI

FIGURE 4.5. Relationship between Compacting Pressure and Sintered Density


for Mineral Source (As-received MCW UO.) and Mechanically Comminuted
(Hammer-milled MCW UO2) Powders [2]. (Courtesy, The American Ceramic
Society, Inc.)
130 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

quires compacting pressures in excess of 100 tsi. In either case,


excessive compacting pressures cause fracture planes in the sintered
shapes. These fracture planes occur at the pressed faces of the com
pact and are inclined at 45 degrees to the axis of compaction. Extreme
cases have been observed in right cylinders where a thin cone separated
from both ends of the compact. In some cases, however, this defect
is subsurface and does not appear unless the part is ground.
Cold pressing can be performed with either single- or double-acting
punches. For most UO, powders, successful single-action pressing
appears to be limited to sections less than 0.050-inch thick. Right cy
linders of less than 0.5-inch diameter have been pressed with a length
to-diameter (L/D) ratio of 1.0 from unmilled mineral source UO2.
For this type of powder, high compacting pressures are required to
produce high-green densities. Pressing right cylinders at L/D
ratios greater than 1.0 results in green compacts having low densities
in the center; this nonuniform density produces nonuniform diametral
shrinkage during sintering. Improvement in sintered size control can
be obtained at L/D ratios greater than 1.0 by use of mechanically com
minuted UO, powder.
The control of shrinkage is important in achieving close dimensional
tolerances because variation in size is related to dimensional changes
between the green and the sintered compact. To attain the maximum
control of shrinkage, the powder is compacted to the highest practical
green density. Since pressures between 20 and 40 have little tsi
effect upon sintered density when milled powders are used (see Fig.
by

4.5), size green density with different


of

controlled the variation


is

pressure, the green


an

pressures within this range. With increase


in

compacts become more dense and, consequently, shrink less during


sintering." Large-scale sintering UO, has been done molybde
of

in

hydrogen atmosphere furnaces full description of


(a

num-wound
the equipment and methods given Ref. 1).
in
is

4.2.2 Hot Pressing


B. E. Schaner

The hot-pressing technique for fabrication bulk UO, involves the


of

application pressure graphite die


of

simultaneous heat and within


a

UO, for
of

form powder shapes. this technique


to

into dense Use


UO, fabrication has been investigated limited laboratory scale.
on
a

In some unpublished work Armour Research Foundation, U.O.


at

A by

bodies were hot pressed density percent theoretical using


95
of
to
a

hot
C.

pressure 17,000 psi and temperature


of

of

1,650°
a

pressing unit was used which consisted basically graphite die


to
of
a

See Chap. for the sintering


of

of

discussion UO2.
7

a
*
FABRICATION OF URANIUM OXIDE 131

contain the UO2 powder, a high-frequency induction furnace to heat


the die, and a hydraulic press to apply the compacting pressure. It
was found best to apply pressure quickly and continuously once
the temperature range was reached; any interruption in pressure ap
plication tended to produce a set which often could not be overcome
without breaking the graphite punch or UO, compact.
Murray, et al., found that lower temperatures may be used to com
pact nonstoichiometric UO, to densities equivalent to those obtained
with stoichiometric UO, [6, 7]. The oxide UO, is sinters appreciably
at a temperature as low as 1,000°C to give a bulk density of 7.89 g/cc.
The maximum density is nearly reached at 1,400°C, when a density of
10.0 g/cc is obtained. At temperatures above 1,800° C the density
again increases (but only slightly) to 10.55 g/cc. Samples pressed
above 1,800° C are weak and brittle and usually break on extraction
from the die. Reaction of the graphite die with the oxide reduces the
outer layers of oxide to stoichiometric composition, whereas the center
remains nonstoichiometric in composition; the nonuniform composi
tion results in subsequent cracking.
The warm-pressing process at 800° C under a pressure of 10

tsi
in
produce closer
to
titanium carbide nickel-bonded die was found
a

UO, than the hot-pressing


of

bulk

at
sintered dimensional tolerances
graphite die [8].
tsi

pressure
to

1,800° under
in

1,400°
C

a
1

4.2.3 Hydrostatic Pressing


B. E. Schaner

Hydrostatic pressing forming which pressure


of

method
in
is

a is
a

isostatically applied [9, 10]. The oxide, powder


of

or

the form
in

loosely pressed shape, placed rubber mold, immersed oil, and


in

in
is

high pressures. Since the compaction forces


on
to

subjected the oxide


applied equally all directions, the green body has uniform den
in

are
a

sity and, consequently, shrinks uniformly during sintering.


Several types UO, powders have been hydrostatically pressed:
of

MCW oxide, UO, precipitated from ammonium diuranate, and fused


UO. [4, 11, 12]. To improve the sintering properties, the powder
is

usually ball milled prior pressing. The addition gen


of
to

binders
is

erally not necessary for this pressing method unless granular feed
a

ensure uniform mold fill give strength for


or
to

to

material
is

needed
handling between the preforming and hydrostatic pressing operations.
by

The powder usually preformed steel dies cold pressing low


in
is

at

pressures, between 10,000 and 24,000 psi [12]. The preformed part
then placed rubber mold and pressed pressures between
in

at
is

30,000and 50,000 psi [11,12].


Sintering may
or

argon, helium, hydrogen [11, 12].


be

done
in

Prismatic and cylindrical shapes have been sintered static argon


in

57.4789 0–81–10
132 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

after pressing fused UO, powder at 10,000 psi in a hardened steel


die [12]. Hydrogen has been used to sinter hydrostatically pressed
shapes at temperatures as high as 1,800° C in molybdenum-wound
surfaces. Fine-grained ammonium-diuranate-precipitated UO2,
pressed into cylindrical shape at 30,000 psi, was sintered in hydrogen
at 1,800° C for 4 hours to 95 percent theoretical density [11]. New
kirk and Anicetti obtained similar densities with lower sintering tem
peratures for MCW UO, which was milled for 48 hours and hydro
statically pressed into tube shapes at 40,000 psi [13]. Schaner ob
tained densities of 97 percent theoretical with ball-milled MCW UO.
which was pressed at 24,000 psi to form short cylinders, then hydro
statically pressed at 50,000 psi and sintered at 1,725°C for 16 hours.”

4.2.4 Slip Casting


T. J. Burke

The slip casting process consists of making a stable water suspension


of insoluble particles which, when poured into a plaster mold, be
comes solidified by means of the liquid extraction mechanism. After
a given setting-up time, the solid has sufficient rigidity and strength
to be removed from the mold in the green state.
Several workers have reported the results of studies on the fabrica
tion of UO, bodies by slip casting [14–16]. Figure 4.6 summarizes
* Unpublished data, Bettis Atomic Power Laboratory.

UO2 POWDER

I. BALL - MILL

2. LEACH
(5N HCl)

3. WASH

4. ADJUST pH AND Sp. G.

5 CAST

6. DRY

7. HYDROGEN-SINTER

FIGURE 4.6. Flow Diagram of Principal Operations for Slip Casting Urania
Refractories.
FABRICATION OF URANIUM OXIDE 133

by
applied successfully Eyerly, the preparation
of the
process al., for

et

by
UO, refractories The slip was formed
[17]. suspending ball
milled UO, dilute hydrochloric acid solution. The leaching op
in
a
eration was performed remove the iron contaminant and aid

in
to

to
the UO, particles the slip. was found helpful
of to the

suspension of

in

It
binder the slip and adjust the particle size distribution

to
by in
omit
a

slip unmilled powder. Addition

of

of
the final addition coarse
clogging introducing open

of
to

material avoid the mold had the effect


porosity.

by
different process was developed form high-density

in to
Gamble
A

rod shapes [18]. This process summarized Fig. 4.7. Gamble


slip is
provides green
clay
of

reported that addition Bentonite

to
the
polyvinyl alcohol provides optimum
of

strength and that addition


mold release properties. Sodium alginate and sodium hexameta
effective deflocculants and yielded good
be

phosphate were found


to

flow and casting characteristics for the low-water slip. Deairing

of
minutes prior
of the

slip for
to

casting was essential for the attainment


3

green body high density. Optimum mold design allowed

18
of
a

percent linear shrinkage. Air drying, usually about hours,


20
to

at 6
by

drying
an

byC.
air convection oven for hours
in

was followed 80°


4
be
percent
90
of

to of

theoretical can obtained


at in

Densities excess
sintering temperatures 1,725°C for
of

in
to

1,700°
10

hours an
8

flowing dry hydrogen.


of

atmosphere

DRY Mix

(uo, Powder, clay, Binder)

DRY BALL-M.ILL

we Mix -s— waſ ER, weTTING AGENT


T

DEFLOCCULENTSOLUTION

of AIR SLIP

cAsT

DRY

st NTER

GRINDSINTEREDBODY
To size

Flow Diagram Principal Operations for Slip Casting UO, Fuel


of

FIGURE 4.7.
Components.
134 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

4.2.5 Extrusion
T. J. Burke

The fabrication of UO, by extrusion is attractive because the pro


cess does not have the length-to-cross-section or the cross-sectional
shape limitations inherent in other conventional ceramic forming pro
cesses, such as dry pressing or hydrostatic pressing. The successful
extrusion of long, high-density rod, tubular, and ribbon-shaped fuel
cores from various types of UO, powder has been reported by investi
gators at several AEC laboratories [3, 19–22].
In general, the extrusion of UO, involves the preparation of the
raw powder, the incorporation of a plasticizer to form a mass capable
of being extruded,the extrusion operation, controlled drying, and
densification by high-temperature sintering in an inert gas or hydrogen
atmosphere.
Several types of UO, have been used as the raw material to produce
high-density extruded shapes. The raw powders normally require
activation to improve sinterability; this can be accomplished either by
mechanical comminution or by controlled oxidation-reduction treat
ment to produce the required particle size distribution, surface area,
and particle structure (see Chap. 3). The incorporation of a densi
fier, such as TiO2, has been found beneficial in increasing the sintered
density of UO, shapes (see Chap. 7). Table 4.1 shows a density com
parison of various extruded UO, shapes [23].

TABLE 4.1–DENSITY COMPARISON OF EXTRUDED UO, POWDERS [23]

Average diam- Percent of theoretical Density


Surface eter; Fisher density increase,
UO2 powder area subsieve sizer green-to
(m2/g) (microns) sintered
Green | Sintered"

1. Ceramic grade--------- - - - - - - - 2.0 1.25 56 96 40

2. Ceramic grade------ - - - - - - - - - - 1.4 1.65 56 95 39


3. Ceramic grade---------------- 0.4 4. 10 57 80 23
4. MCW—PWR (milled 9 hr)------ 0.8 1 : 50 56 83 27
5. MCW—PWR (milled 64 hr)**__| 2.3 0.90 59 93 34
6. MCW—PWR (milled 8 hr)**____| 2.1 1.28 59 93 34

7. Hanford depleted”-- - - - - - - - - - 5.2 4.80 52 90 38

(From Ceramic Industry, October 1958,Reproduced by Permission.)


*Sintered in hydrogen atmosphere for 8 hours at 1,750°C.
**Treated by a controlled oxidation-reduction process.

To extrude bulk UO, it is necessary to incorporate an agent which


will impart plasticity to the powder mass and render it capable of be
ing forced through a die to form a uniform, coherent shape which
possesses sufficient green strength for handling during postextrusion
FABRICATION OF URANIUM OXIDE 135

drying and densification Allison and Duckworth sur


treatments.
veyed several plasticizers for of UO, powders
use in the extrusion
[3]. In addition, other compounds, such as polyvinyl alcohol, corn
starch, polyethylene glycol, and methylcellulose, have been evaluated.
Polyvinyl alcohol was found difficult to work with in amounts exceed
ing 1 weight percent because of the tacky character it imparts to the
wet mix and to the wet extruded shapes. Cornstarch was found inade
quate because mixes plasticized with this material produced fragile
extrusions. The polyethylene glycols, although attractive because as
waxy materials they serve both as plasticizer and lubricant, proved
unsatisfactory because excessive amounts were required to produce
bodies of practical green strength. Optimum results have been ob
tained with 1 to 2 weight percent methylcellulose and with 1 to 2.5
weight percent Carbopol 934 [20, 22].

4.2.6 Swaging
D. R. Stenquist, P. L. Farnsworth, and R. J. Anicetti

Swaging is a method of fabrication whereby ceramic materials can


be simultaneously compacted and clad into relatively inexpensive,
high-density, rod-type fuel elements. Exploratory experiments on the
compaction of UO, in tubes by swaging were first reported by Quinlan
and Roake [24]. Their experiments demonstrated that swaging is an
attractive method for rapidly fabricating tubular or rod-type UO,
fuel elements of any length for nuclear reactors. Compacted UO, den
sities greater than 87 percent of theoretical were obtained by cold
swaging and up to 94 percent by hot swaging.
At Chalk River [25] cold rotary swaging has been used to obtain a
high degree of densification of UO, powder within a metal sheath.
Densities as high as 95 percent of theoretical were obtained in tests on
various UO2 powder preparations sheathed in aluminum, austenitic
stainless steel, and Zircaloy—2.
The action of the swage machine consists of a series of very rapid
blows which reduce the diameter of a metal tube or rod without chang
ing the shape of the cross section. By this process, an element contain
ing a ceramic fuel core is produced to very close tolerances with an ex
cellent surface finish (see Fig. 4.8) [26]. The main working parts of
the swage machine are shown in Fig. 4.9. In general practice, the
dies are caused to reciprocate as they pass between opposing rollers of
the roll cage. As the dies pass from under the rollers, they are thrown
outward, allowing the work to be inserted further into the die, ready
for the next forging stroke.
The feasibilityof manufacturing ceramic fuel elements by the
swaging process depends primarily on the densities that can be ob
136 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

& º,
- -
& Sº
G S
3 INCHES

FIGURE 4.8. Swaged UO, Fuel Rods; 1X [26]. (Courtesy, American Institute
of Mining, Metallurgical, and Petroleum Engineers, Inc.)

tained and on the reproducibility of the product. Parameters which


affect the density include percentage reduction in area, particle density,
cladding material, and swaging technique.
The physical properties of the cladding material markedly affect
UO, density. This is apparent when one considers the action which
occurs during swaging. The diameter of a metal tube when swaged
is reduced with little elongation. The displacement of metal is accom

If,

the
modated by an increase in the tube-wall thickness. however,
tube swaged over mandrel, the tube walls are thinned, and the flow
is

by

lengthening
of

of

metal revealed the tube. Powdered UO.


is

offers little resistance during the early stages swaging. As the clad.
of

ding-tube diameter reduced, the wall thickens with little elongation.


is

Thus, the contained volume

the
reduced, and the bulk density

of
is

UO, increased. The UO, acts


as

mandrel when sufficiently


is
it
is

densified and offers increasing resistance further deformation.


to

Further swaging only reduces the cladding thickness because the

forces required promote further UO, densification exceed the yield


to

strength cladding. Swage compaction studies have included


of

the
In

UO. Zircaloy-2, and stainless steel.


aluminum,
in

clad general,
cold-swaged UO. densities aluminum and Zircaloy-2 are lower than
in

are

those obtained stainless steel when identical swage reductions


in

made.
Uranium dioxide densities after swaging depend large degree
to
a

upon the physical characteristics the UO, powder used.


In
of

general,
high-swaged densities are obtained with powders having low-surface
areas, high-particle densities, and large average particle sizes. These
powder are opposite for
to

characteristics those suitable forming


by

bulk UO, cold pressing and sintering.


The particle density the UO, powder probably the best single
of

is
its

property from which predict swage-compacted density.


to

Elec.
tron microscopy helpful predicting swaging
in

of

the behavior
is

a
··· ---…
…………………
\\ --daerae

na
wae.……/"
… mae

o
awu-/ aalwo aamma

aswawo onulua

…o.s…--
- ºmaes
ºn-le
FABRICATION

×ove mae

1
as 3 0 –
OF URANIUM

v
l.
ºras 130 ºm...……d.

n.
~--–
— 310
OXIDE

– ~ www.º.
83 xoongºniniw138ºnlae
»

~ aer -ſ+ donae


→~ – www.º. xoongwannos
~ – les mqae
– aennom
minou~ Bºwo
– Bowºalvind
→ 0Ni*
IV

ſº

I
nÐIJI 38 '6"# pºļJòAuI ĐÔIKJ, Itºno? baaS ºu ºuȚqob
137
138 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

UO, powder. Carbon replicas readily reveal the morphology of UO.


particles (see Appendix B, Microscopy of UO2) [27, 28]. Powers hav
ing smooth surfaces and few pores are most likely to be swaged to high
densities.
Of the powders evaluated, fused UO, with a particle density essen
tially theoretical compacts tothe highest density. Cold-swaged densi
ties of approximately 90 percent of theoretical are obtained with swage
reductions of 45 percent of cross-sectional area. This density can be
compared with a density of about 74 percent of theoretical obtained
with untreated UO, powder. Uranium dioxide particles prepared
from sintered compacts of untreated UO, swage to about 86 percent
of theoretical density.
Swage compaction studies with nonactivated UO, illustrate the
effect of temperature on density. The UO, packed in 0.625-inch OD
by 0.060-inch wall stainless-steel tubes, swaged to 80 percent of theo
retical at a reduction in cross-sectional area of 60 percent at room
temperature. At 600°C,
under similar swaging conditions, a density
of 89 percent of theoretical was obtained, while at 800°C, 94 percent
of theoretical Tubes containing UO, can be heated with
was achieved.
either resistance or induction heating techniques. In resistance heat
ing, the rod is passed through two brushes which cause electricity to
flow in the fuel rod cladding. In induction heating, the fuel rod is
passed through an encircling induction coil. Cladding temperatures
are controllable within +50° C at 800° C.
Several steps in fabricating a swaged fuel element are illustrated
in Fig. 4.10 [26]. The UO, is vibrated into a thin-walled metal tube,
one end of which has been sealed with absorbent cotton. The remain
ing end is similarly sealed after filling the tube. For hot swaging,
asbestos rope is a convenient substitute for the cotton. The swaged
rods are straightened, when necessary, in a rod straightener. The
fuel elements are then cut to the specified length, and end caps are
welded into place. The welded fuel rods are nondestructively tested,
etched, autoclaved, and assembled into a rod cluster fuel element.
Swaging offers promising method for economical fabrication of
a
UO, fuel elements. Most cold compaction techniques, if not followed
by subsequent sintering, result in UO, cores of low density. With
swaging, however, certain UO, powders can be compacted to densities
of 90 percent of theoretical. Higher densities are expected if very
dense oxide particles having optimum particle size and shape are used.

4.2.7. Vibratory Packing


H. R. Hoge

Vibratory packing is another fabrication method which has been in


vestigated for application to the densification of bulk UO, powders.
~
FABRICATION OF URANIUM OXIDE 139

A. Before Swaging,
UO, Powder Is B. End Caps Such As This Typical
Poured into a Metal Capsule Which Design Are Welded in Place for Hot
Is Then Sealed at Each End. Swaging.

| 18 NCHES |

C. After Swaging, Fuel Elements Are D. A Complete Three-Rod Cluster


Cut to Length and a SmallAmount Fuel Assembly Consisting of Swaged
of UO, Is Removed to Allow Weld- UO, Rods 18 Inches Long by 0.570
ing of End. Inch OID.

Figure 4.10. Stages in Fabricating a Swaged UO, Fuel Element [26]. (Cour
tesy, American Institute of Mining, Metallurgical, and Petroleum Engineers,
Inc.)

The process presents another means of producing long, integral length,


rod-type fuel elements. This process, originally developed by Bell
and coworkers, was investigated by Hoge as a possible method of pro
during fuel rods containing MCW UO, [29–31]. Although relatively
long, thin rods, both clad and unclad, were produced by this method
with laboratory scale apparatus, the highest density obtained was only
70 percent of theoretical. The process does, however, show promise
140 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

UPPER
VIBRATOR
+

—L–TLTAMPING
ROD

FEED
HOPPER

COMPACTION
DIET--
|

for
LOWER
VIBRATOR

FIGURE 4.11.
~Laboratory
U

Device for Compaction


º
of UO, Powder.

the fabrication of fuel cores from both fused and sintered, high
density UO, powders.
The powder compaction is accomplished by overcoming particle
inertia with low-frequency, high-energy shock waves. By impacting a
powder mass with high energy, the individual particles tend to pack
to the highest possible density consistent with the geometry of the
system. Unlike cold pressing, this process does not fracture the ma
terial to effect densification, but permits the particles to assume posi
tions favorable for closest packing. Vibration amplitude and fre
quency, along with the kinetic energy of the tamping mechanism, are
the chief variables which must be adjusted to produce a body of opti
mum density. The particle size distribution of the powder is an im
portant factor in the attainment of maximum density by this method.
Powders of mixed particle size are, in general, capable of being com
pacted to higher density than powders having a narrow size distribu
tion because the finer particles can be vibrated to fill the voids which
exist between the coarse particles.
A typical laboratory vibratory packing apparatus, shown in Fig.
4.11, is composed of three principal parts:
for holding the powder
a die
to be packed and two opposing pneumatic vibrators which independ
FABRICATION OF URANIUM OXIDE 141

ently activate upper and lower tamping rods. The lower rod is fixed
and the upper rod is retractable. The apparatus is designed to allow
the three components to float freely on common guides and direct op
posing shock waves from the vibrating tamping rods to the powder
in the die. All of the system components are isolated from each other
in order to minimize energy losses. Compaction dies should be made
of a ductile material having properties similar to those of low carbon
steel. Designs should incorporate extra heavy cross sections wherever
possible to prevent rapid failure of the component by fatigue. Dies
for packing long, thin rods should be made of three, rather than two,
sections to facilitate compact removal. Dies and die holders should
be mounted on shockproof rubber or springs to prevent damping and
consequent loss of compaction energy to the structural supports.
Long, thin UO2 rods are compacted by continuously feeding con
trolled quantities of damp (less than six weight percent water), free
flowing powder into the packing die, which contains an upper and
lower tamping rod. The powder is metered into the die past the
upper tamping rod while the rod is vibrating. The vibrations pack
the powder as it is fed into the die. Adjustment of the continuously
vibrating upper tamping rod to compensate for change in impact
level is accomplished by periodically raising and lowering the rod at
intervals throughout the operation. This procedure is continued un
this manner, UO, rods
In

desired length produced.


of
til

rod
is
a

or

may be formed within the steel die, the die may designed
be

hold to
tubes and the powder fed into and compacted with the tube produce
to

clad, green compact directly.


a

In

recent work, vibrational compaction has been shown rapid,


be
to

simple, and extremely flexible method simultaneously compacting


of

and cladding ceramic fuel materials [32]. This method provides


a

fabricating having core materials and cladding


of

means fuel elements


geometries more conventional techniques.
to

not amenable
the Hanford work the equipment used for vibrational compac
In

by

electrodynamic, moving-coil vibrator driven


an
of

tion consisted
two, 1-kw amplifiers [32]. Uranium dioxide eight-foot long Zir
in

G,

no
as

caloy tubes was vibrated great with


at

as

accelerations 100
or

dimensional changes
to

adverse effects the tubes.


Important variables vibrational compaction, addition vibra
to
in

in

tion condition, particle density, and particle size distribution


of

the
fuel, were found cladding geometry and method loading the
of to
be

of

Each type fuel element requires particular combination


of

fuel.
a

by
of

these interrelated variables. Some the conclusions reached the


Hanford work are summarized below [32]:
UO, compacts are function the particle density.
of

of

Densities
1.

UO, having particle density 100 percent


98
of

of

Fused theoreti
to
a
FABRICATION OF URANIUM OXIDE 141

ently activate upper and lower tamping rods. The lower rod is fixed
and the upper rod is retractable. The apparatus is designed to allow
the three components to float freely on common guides and direct op
posing shock waves from the vibrating tamping rods to the powder
in the die. All of the system components are isolated from each other
in order to minimize energy losses. Compaction dies should be made
of a ductile material having properties similar to those of low carbon
steel. Designs should incorporate extra heavy cross sections wherever
possible to prevent rapid failure of the component by fatigue. Dies
for packing long, thin rods should be made of three, rather than two,
sections to facilitate compact removal. Dies and die holders should
be mounted on shockproof rubber or springs to prevent damping and
consequent loss of compaction energy to the structural supports.
Long, thin UO2 rods are compacted by continuously feeding con
trolled quantities of damp (less than six weight percent water), free
flowing powder into the packing die, which contains an upper and
lower tamping rod. The powder is metered into the die past the
upper tamping rod while the rod is vibrating. The vibrations pack
the powder as it is fed into the die. Adjustment of the continuously
vibrating upper tamping rod to compensate for change in impact
level is accomplished by periodically raising and lowering the rod at
intervals throughout the operation. This procedure is continued un
this manner, UO, rods
In

desired length produced.


of
til

rod
is
a

or

may be formed within the steel die, the die may designed
be

hold to
tubes and the powder fed into and compacted with the tube produce
to

clad, green compact directly.


a

In

recent work, vibrational compaction has been shown rapid,


be
to

simple, and extremely flexible method simultaneously compacting


of

and cladding ceramic fuel materials [32]. This method provides


a

fabricating having core materials and cladding


of

means fuel elements


geometries more conventional techniques.
to

not amenable
the Hanford work the equipment used for vibrational compac
In

by

electrodynamic, moving-coil vibrator driven


an
of

tion consisted
two, 1-kw amplifiers [32]. Uranium dioxide eight-foot long Zir
in

G,

no
as

caloy tubes was vibrated great with


at

as

accelerations 100
or

dimensional changes
to

adverse effects the tubes.


Important variables vibrational compaction, addition vibra
to
in

in

tion condition, particle density, and particle size distribution


of

the
fuel, were found cladding geometry and method loading the
of to
be

of

Each type fuel element requires particular combination


of

fuel.
a

by
of

these interrelated variables. Some the conclusions reached the


Hanford work are summarized below [32]:
UO, compacts are function the particle density.
of

of

Densities
1.

UO, having particle density 100 percent


98
of

of

Fused theoreti
to
a
FABRICATION OF URANIUM OXIDE 143

Of the various uranium oxide-metal systems, two have been studied


in some detail. Fabrication techniques for uranium oxide dispersions
in aluminum and stainless steel are described below.

(a) Fabrication of Uranium Oaside-Aluminum. Fuel Elements'

Aluminum-base fuel elements have been generally used for high


flux, low-temperature research reactors which operate at relatively
high power densities. Compactness and high operating performance
characteristics are achieved by fully incorporating enriched uranium
into extended surface fuel elements to facilitate heat removal. For
most applications the conventionally accepted method for incorporat
ing the uranium in the aluminum has been by alloying. In special
cases where a very accurate knowledge of the content is manda U*
tory or where a high uranium content (45 to 50 weight percent) is
required, uranium oxide as a fuel dispersant in aluminum is consid
ered. Initial preparation of the UO, or UAOs cermet is easily ac
complished by conventional powder metallurgy techniques. Sub
sequent forming of the fuel-bearing core into a finished element may
be accomplishedby a variety of fabrication techniques. The chemical
compatibility problem which exists with combinations of both UO,
and UAOs with aluminum is discussed in Chap. 7. In general, the
combination of UO, with aluminum may be a serious problem if the
fabrication process involves much time at elevated temperature. The
combination of UAOs with aluminum, on the other hand, is more stable
at the elevated temperatures required for processing composite
products.
Aluminum-clad dispersions of aluminum and uranium oxide
(UO2 and U.Os) are amenable to fabrication into fuel elements by a
wide range of methods. Roll cladding and brazing, extrusion, swag
ing, and cold drawing have been developed into satisfactory techniques.
The first major use of a UO, dispersion fuel in an aluminum matrix
was the manufacture of the core loading for the Geneva Conference
Display Reactor by Oak Ridge National Laboratory in 1955. The
procedure used was a modification of that developed at ORNL for
producing MTR-type fuel elements [33]. The billet is prepared for
hot rolling by use of the picture-frame technique (seeFig. 4.12).
Metallurgical bonding is achieved by hot rolling at 590° C to
a total

reduction in thickness of 91 percent. Plates are flux annealed in a


furnace for 1 hour at 600° C with an aluminum brazing flux to
minimize blister formation. They are cold rolled to a final thickness
of inch and stress annealed for 11% hours at 500° C. The plates
0.060
are then machined, annealed for 1 hour at 600°C, and formed cold to a

* R. J. Beaver, J. E. Cunningham, and R. C. Waugh.


144 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FUEL-BEARING
ALLOY OR
PRESSED-POWDER
CORE

FIGURE 4.12. Exploded View Showing Makeup of Composite Fuel Plate Prior
to Rolling [33].

5.5-inch radius of curvature. Final joining of these plates into an in


tegral assembly is accomplished by brazing with a 12 percent silicon
aluminum alloy. The assembly is brazed at 600° C for 35 minutes in
a forced air circulation furnace. Figure 4.13 shows a cross section
of a brazed fuel element assembly. Note the outline of the fuel
bearing core of each plate.
The aluminum-UO, compatibility problem at elevated temperature
caused considerable difficulty in the fabrication of these elements [34].
The fuel plate growth which occurred during the brazing operation
caused many elements to be rejected because of unacceptable fuel
plate spacing. On the other hand, fuel elements containing 56 weight
percent UAOs (minus 100, plus 325 mesh) and aluminum (minus 100
mesh) have been successfully fabricated at Oak Ridge National
Laboratory by the same technique.
A modification of the picture frame technique has been developed at
Argonne National Laboratory [35]. It involves encasing the fuel
FABRICATION OF URANIUM OXIDE 145

| 2 INCHES |

FIGURE 4.13. Cross Section of a Brazed Fuel Element Assembly [33].

core A highly compacted core


with aluminum by a casting operation.
of 55 weight percent U.O,
and aluminum is placed in a specially de
signed mold and held in position by three steel locating pins. Molten
aluminum alloy is poured in air about the core, completely encasing
The locating pins are removed after solidification, and the pin
it.

holes are sealed during subsequent hot rolling. uniform dispersion


A

U.O. uniformly and completely clad


obtained, and the core
is
of

is
the

with aluminum alloy.


by

Uranium dioxide-aluminum fuel elements have been fabricated


hot

8.5

pressing and extrusion [36]. The weight percent UO, (minus


$25 mesh) and atomized 2S aluminum powder (minus 100 mesh) are
200 hot

double acting graphite die pressure


of
in
at

95 at

pressed 645°
C

1,000 psi.Billets produced percent


99

this manner are


to
in
to

dense. The most satisfactory extrusions are made by using shear


a
146 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

die with a 146-inch land. A temperature of 510° C and a slow extru


sion speed of the order of 5 inches of rod travel in 10 seconds give best
results. By placing a soft steel dummy block in back of the billet, ap
proximately 95 percent of the billet is converted into satisfactory rod.
Rods exhibiting a variation of +0.1 percent in the average UO2 con
centration of 8.5 weight percent were produced. The extruded rod has
considerable ductility and can sustain extensive swaging. A reduc
tion in area of 13 percent by swaging produces a fuel rod accurate to
within +0.2 mil in diameter and with a warp of 3 to 5 mils over an
8-inch length.
Aluminum-U,Os fuel elements have been successfully fabricated by
hot extrusion. The first core loading for the Argonne Zero Power re
search reactor (Argonaut) was fabricated at Argonne National Lab
oratory. The plates were made by coextruding a mixture of 39 to 40
weight percent partially enriched UAOs and aluminum powder con
tained in a thin-walled aluminum can with end plugs at approximately
440°C. The extruded strip was cut into the desired lengths in a press
mounted die, exposing the core at both ends of each plate. The good
corrosion resistance of the fuel core under the operating conditions
anticipated in the Argonaut made end sealing unnecessary.
More recent work indicates that UO,-aluminum fuel elements of
tubular cross section can be fabricated by a technique involving a final
diffusion annealing operation to secure metallurgical bonding of the
cladding to the core [37]. The process entails the bonding of a fuel
bearing section of tubular cross section with two concentric aluminum
tubes to form the composite product. The inner and outer cladding is
1100-grade aluminum, and the core is a dispersion of 65 weight percent
UO, in an aluminum powder matrix. In cores where a high uranium
investment is required, a mixture of 95 percent aluminum and 5 per
cent aluminum-silicon alloy of eutectic composition is used to gain
additional densification.
In the initial preparation of the fuel-bearing core, two different
techniques have been developed. In one, the core mixture is cold com
pacted at 3.8 tsi into a thin billet, hot-rolled at 620° C into a dense
sheet, trimmed to size, and formed into tubing by cold swaging on a
mandrel. Tubular-shaped cores containing 30 to 50 weight percent
UO, and having a wall thickness of 0.010 to 0.015 inch have been suc
cessfully prepared in this manner. The other technique involves cold
pressing cylindrical bushings 1% inch high at 42 and sintering
tsi

in

hydrogen for 620°C for added densification. Bushings


at
15

minutes
containing weight percent UO, and having 0.015
65

75

0.025-inch
to

to
by

wall thickness have been successfully prepared this method.


by

Fuel-bearing cores prepared either method follow essentially the


same procedure the makeup the composite fuel tubes.
of

The
in
FABRICATION OF URANIUM OXIDE 147

wrought aluminum cladding is chemically cleaned and activated for


bonding. The core is assembled between the inner and outer cladding
and the assembly slipped over a steel mandrel. The fuel assembly is
given an elevated-temperature vacuum degassing treatment of 45 min
utes at 630°C to attain a vacuum of less than 50 microns. This step is
followed by a reduction of 35 percent in cross-sectional area by cold
drawing under dynamic vacuum to effect contact between the compo
nent parts and achieve the desired dimensions. The mandrel is re
moved and a metallurgical bond is attained by a final diffusion-anneal
ing operation of 30 minutes at 630°C. The fuel tube is radiographed,
cut to proper length, and inspected.
A density of 4.74 g/cc has been measured on bushings of 65 weight
percent UO2. After sintering, the density is slightly lower as a result
of burn-off of the stearic acid lubricant. Fuel tubes containing these
bushings can be cold worked to give reductions in area greater than 50
percent and elongations of approximately 100 percent without any
deleterious effects.

(b) Fabrication of Uranium Oaside-Stainless Steel Components *

Stainless-steelbase components which contain a dispersion of en


riched UO,
have been successfully developed for use in power reactors.
The uranium dioxide fuel material is uniformly dispersed in stainless
steel powders by blending, then formed to a desired intermediate
shape by the pressing, sintering, and coining procedures employed in
powder metallurgy. The final fuel element form and metallurgical
bonding of cladding to the dispersion component are attained by roll
cladding, extrusion, or swaging.
The stainless-steel powders selected as the matrix material and the
cladding should be chosen on the basis of good corrosion resistance to
the specific reactor coolant expected to be encountered in service.
Normally, austenitic stainless steels are selected because of their good
corrosion resistance properties. On the other hand, ferritic steel,
which has a lower coefficient of thermal expansion than austenitic
steel, more readily matches the expansivity of UO. Also, the ferritic
stainless steel, which has a higher thermal conductivity than austenitic
is more desirable since it prevents overheating of the UO2 phase
steel,
and increases the rate of heat transfer to the reactor coolant. All
these factors must be carefully considered in the selection of the
matrix alloy. Either angular or spherical prealloyed stainless-steel
powders are commercially available, and both have been used suc
‘essfully. The particle size of the stainless-steel powders or elemental
metal powders should be as small as possible (minus 325 or minus 400
mesh) to facilitate blending and sintering operations. When ele

* G. L. Ploetz.

57.4789 O–61–11
148 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FIGURE 4.14. Photomicrograph of Rounded UO, Particles of 50- to 100-Micron


Size Sintered at 1,725° C for 1 Hour in Hydrogen Atmosphere; Unetched :
X250.

mental powders are selected, electrolytic- or carbonyl-iron, carbonyl


nickel, and electrolytic-chromium powders are used.
The shape, size, bulk density, and the internal grain growth of the
UO, particles used as the dispersed fuel phase are the principal vari
ables to be controlled in order to secure the optimum physical proper
ties of this material. It is possible to increase the sinterability of UO.
by small additions of other oxides (see Chap. 7) [38]. Such an ap
proach, however, has been found to be unnecessary if a large surface
area and a small ultimate crystallite size of the UO, are utilized to
provide a readily sinterable material. To prevent particle breakup
or stringering, it has been found that rounded or approximately
spherical particles (see Fig. 4.14) are preferable to angular, as
screened UO, particles.
The optimum particle size of the dispersed UO, particles is a com
promise among several desiderata: (1) Particles as large as possible
to retain the maximum amount of the fission products (both solid and
gaseous) with the UO, thereby reducing the possibility of dimensional
change of the fuel element; (2) particles as large as possible to pro
vide minimum surface area in contact with the matrix material,
thereby decreasing the radiation effects on the matrix; (3) particles
as small as possible to reduce self-shielding effects and to prevent
overheating of the UO, particles as a result of the low thermal con
FABRICATION OF URANIUM OXIDE 149

ductivity of UO, ; and (4) particles of an optimum size to provide a


continuous metallic matrix, thereby improving the heat transfer rates
and the structural stability of the fuel elements (see Chap. 9).

The selection of an optimum size of UO, particles is governed by the


shape, the physical dimensions, and the fuel loading of the elements.

For example, an upper limit for UO, particle size has been chosen as
10 percent of the fuel element core diameter in the case of cylindrical
elements [39].

Another guide to proper particle size selection is the fact that for
a fuel element containing 20 volume percent UO, in stainless steel,

radiation damaged zones surrounding each UO, particle will touch


the

UO, particle diameter approximately microns.” Therefore,

25
to of
at
a

particle size range 100 microns (minus 140, plus 270 mesh)
50
to of
a

sometimes selected prevent radiation damaged zones from touch


is

providing continuous areas undamaged matrix.


of
ing, thereby
UO, particles high

be

is to as
density
of

The bulk

as
the fissile should
UO,
of

reduce the volume fraction


in
to

possible the stainless steel


minimum. For this reason, the bulk density the UO, particles
of of
a

be

specified percent theoretical density.


90
of

minimum
of to

usually
a

grain UO, particles be


as

as
The size the sintered should small
large grain growth increases the brittleness

of
possible because the
particles. Large grain growth can result stringering
or

breakup

of
in

these particles during metallurgical fabrication processes.


It

has been
average grain size approximately
an

microns, with
of

found that
a
6

maximum grain size approximately microns within the particles,


by 10
of

safe range that readily obtained conventional UO, sintering


is
is
a

techniques.

Uranium dioxide particles with these desirable features can pre


be

pro
by

pared conventional ceramic processes. The as-received UO2,


by

or

duced either the ammonium diuranate the hydrogen peroxide


precipitation process, ball-milled with acetone for hours stain
in
is

mill with mixed sizes stainless-steel balls. After drying,


of

less-steel

this fine-grained UO, pelletized under 2,000 psi pressure without


is

binder. The pressed pellets are broken into angular, oversized parti
The oversizing necessary percent firing
to

20

allow about
is

cles.
correctly particles
or

shrinkage. The sized compacted are rolled


all
on

to

swirled fine mesh screen eliminate sharp corners, then given


a

The particles are sintered


C, in

final screening. molybdenum boat


in
a
a

dry hydrogen atmosphere temperature


of
to

maximum 1,725°
a
a

this temperature for hour, and furnace cooled


at

to

held room
1

The particles are screened final size and inspected


to

temperature.
shape, bulk and pack density, and grain size.
for

Barney Seymour, Laboratory, personal


E.

W. Knolls
K.

"W. and Atomic Power


*ommunication.
150 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Several types of UO2-stainless steel dispersion fuel elements have


been successfully fabricated:
1. Cold-pressed and sintered powder cores with picture-frame clad
ding, hot-rolled to final size [40, 41].
2. Cold-pressed powder cores in a stainless-steel jacket, hot extruded
and hot drawn to final size.”
3. Cold-pressed and sintered powder cores in a stainless-steel tube,
hot coextruded to nearly final size.”
4. Cold-extruded and sintered powder cores inserted in stainless-steel
tubing, hot swaged or hot Turk's-headed to final size.”
5. Powder-rolled or direct-rolled and sintered cores formed into flat
plates and clad by picture framing. In an adaptation of this proc
ess, the plates are formed into tubes that are clad inside and out
with stainless-steel tubing” [42].
6. Slip-cast and sintered powder cores, with or without powder clad
ding, hot-rolled or hot Turk's-headed to final size. If only the core
is slip cast and sintered, the cladding can also be performed by con
ventional picture-frame techniques.
7. Cold-pressed and sintered cores hot extruded by assembly of up to
12 into an extrusion billet and extruding with the cladding to nearly
final dimensions.”
In any of the above methods, the objective is to achieve full metal
lurgical bonding between the stainless-steel cladding and the two
phase UO2-stainless steel core. At the same time, complete random
distribution of the fissile UO, particles within the metal matrix is
necessary. The various metal working processes sometimes cause par
ticle breakup or stringering. Such defects must be kept to a minimum.
To achieve random distribution, careful attention to powder blend
ing or mixing operations is essential. With reasonable precautions,
the small density difference between the two phases will not cause
phase segregation.
In all ofthe methods except (2) above, the powder is sintered to
achieve partial densification of the UO2-stainless steel cores before
the final cladding operations. Thus, the sintering operation is an im
portant step that requires special attention. Both phases require non
oxidizing atmospheres for sintering; therefore, a hydrogen atmos
phere is usually used. It is important that the hydrogen be as free
from moisture and oxygen as possible.

10H. H. Hirsch, Knolls Atomic Power Laboratory, personal communication,


11W. E. Ray, C. J. Beck, and H. G. Sowman, Knolls Atomic Power Laboratory, personal
communication.
12H. G. Sowman, C. E. Lacy, G. L. Ploetz, and C. J. Bier, Knolls Atomic Power Labora
tory, personal communication.
* L. Frank, The Martin Co., unpublished data.
* R. N. Honeyman and J. R. Peloke, Knolls Atomic Power Laboratory, personal
communication.
FABRICATION OF URANIUM OXIDE 151

For most stainless-steel powders,


either prealloyed or elemental,
Depend
ing the

sintering temperatures
range from 1,200°

at C.
to
1,350°
the fuel elements, soaking times

of
shape and size
on

maximum
temperature usually vary from For each particular

to
hours.

6
conditions the optimum time-temperature relationships should
be set

of

the process development stage. The sintered den


in

determined
usually within the range percent

70

85
of

of
to
of

sity the cores the


is

UO,-stainless
density More im
of
theoretical the steel mixtures.

be
portant than high-sintered densities density value that can

is
a

In
achieve dimensional uniformity.
to

consistently obtained some

by
cases, the sintered cores are further densified coining operation

a
before the cladding applied. This operation also corrects small
is

dimensional nonuniformities caused by differences shrinkage rate

in
during sintering.
cladding processes depend upon the desired final shape, size,
The
the fuel elements. Examples typical shapes are
of

of
and dimensions
flat

plates, tubes, ribbons, pins, and rods.


of
Most the conventional
metal working processes have been adapted

of
solve the problems
to

cladding fuel cores with complete metallurgical bonding. For flat


plates, the picture-frame technique usually used. The core com
is

is
pletely covered with the cladding layers, seam welded, evacuated, and
to

final size. The percentage per pass kept


of

hot-rolled reduction is
prevent UO, particle breakup.
to

minimum
to
a

the various hot-extrusion processes, the cladding


In

extruded
is

with the core, that further cladding operations are not


so

integrally
to In

further reduction final size required


to

required. some cases,


is

subsequent the hot-extrusion operation.


For the cold-binder process, after welding end plugs
to

the sintered
by

core, cladding inserting the sintered cores stainless


in

achieved
is

steel tubing. To facilitate bonding, thin layer


be
of

nickel can elec


a

Bonding
or

the tubing.
of

the inside
to

troplated onto the sintered core


by

hot working swaging machine


to or

Turk's-head
in

achieved
is

Size reductions per pass are again held


to

machine. minimum
a

prevent breakup the dispersed UO, particles.


of

cladding operations, the fuel elements are annealed


all

develop
In

to

homogeneity within the matrix alloy. The fuel elements are usually
by

cleaned after the metal working operations firing dry hydrogen


in

electropolishing produce
of to

and smooth surface.


a

Corrosion testing UO2-stainless fuel elements the en


in

steel
important criterion
be

an

of
to

encountered service
in

is

vironment
successful future performance. Therefore, corrosion tests conducted
the applicable coolant medium are essential under both static and
in

dynamic conditions.
152 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

4.3.2 Dispersions in Ceramics

fuel material based on a combination of UO, with a metal oxide,


A
both mutually insoluble in the solid state, would be a true ceramic
dispersion type fuel. Two such metal oxides of low cross sec
tion (alumina and beryllia) which do not form solid solutions with
UO, (see Chap. 6) have been investigated as possible fuel materials
for power reactors. The fabrication of ceramic bodies of UO.-Al2O,
and UO,-BeO is described below; irradiation behavior of these mate
rials is discussed in Chap. 9.

(a) Fabrication of Aluminum Oaſide-Uranium Dioxide Dispersions *


Dispersions of UO, in Al2O, in the form of plates (or wafers) up to
2 inches long, 1 inch wide, and 0.250 inch thick containing 21 weight
percent UO, have been fabricated by cold pressing and sintering
techniques similar to those utilized for the fabrication of bulk UO.
[19]. The Al2O3–UO, mixtures are ball milled in a rubber-lined mill
using uranium slugs as the grinding medium, then agglomerated with
a 1 weight percent polyethylene glycol binder. The wafers are formed
by cold pressing, and densification is attained by hydrogen sintering.
The source of the alumina was found to have a distinct effect on dis
tortion of Al2O3–UO, compacts during sintering. Densities obtained
on sintering were found to be substantially affected by the particle
size of the UO, powder. Highest densities are obtained with a feed
material composed entirely of minus 325 mesh UO, powder made by
crushing and grinding presintered UO, pellets.
Wafers of 21 weight percent UO,-Al2O, composition, approximate
ly inch square by 0.040 inch thick, and with a sintered density over 4.5
1
g/cc (98.3 percent theoretical), have been fabricated on a semipro
duction basis [43]. The raw materials used were natural ADU-type
UO, and grade A–14 aluminum oxide. Mixing of the raw powders
was accomplished by ball milling. A free-flowing feed material was
formed by mixing with a 4 percent polyethylene glycol water-acetone
solution, drying, and forcing through a 35-mesh screen. The granu
lated powder was dry pressed at 20 pieces per
of
tsi

16

25
at

to

rate
a

tabletting press. Hardened steel dies and


on

minute Stokes 280–G


a

punches were replaced with carbide dies and punch tips prevent
to

the powders. Because the as-pressed


of
to

wear due the abrasive nature


wafers were only 0.055 inch thick, the fill control was extremely crit
ical. The wiping action nonuniform filling
of

the feed shoe resulted


in
by

the cavity and was compensated for tapering the lower


of

by
fill
on

punch increase the depth


of

of

the die affected


to

the side
15

B. E. Schaner.
FABRICATION OF URANIUM OXIDE 153

the shoe motion. The wafers were sintered to densities between 4.50
and 4.56 g/cc in flowing hydrogen at 1,750° C for 8% hours.

(b) Fabrication of Beryllium Oaside-Uranium Dioacide Dispersions *


High density bodies of BeO-UO, have been fabricated by cold press
ing and sintering mixtures of BeO and UO, powders. Similar to
Al-O,-UO2, BeO and UO, form a simple eutectic system (see Chap. 6)
[44]. The resultant product is a dispersion of UO, particles in a BeO
matrix.
Best results were obtained by using finely divided BeO powder.”
For small particle size UO2 powder (<10a) dispersions, the two
powders are mixed, agglomerated with 10 weight percent polyethylene
glycol (Carbowax 20M), pressed at 16

tsi
steel dies, and sintered

in
H, for

in
of at

14

1,750° The sintered bodies have densities


hours.
in

percent open porosity


97

of

of
excess theoretical
and show absence
(see Fig. 4.15A). Difficulties were experienced fabricating disper
coarse UO, particles (approximately 150p) BeO according in
of

in
sions
UO, particles be
as

procedure segregation
of

of

result
to

this the
a

higher particle size. Segrega


of

cause their density and increased


by be

prevented large extent and approximate homogeneity


to

tion can
a

forming paste the powders with percent dextrin,


of

achieved
2

2
a

percent Carbowax 20M, 0.5 percent sodium alginate, and water. The
allowed partially dry before forcing through
to

paste 14-mesh
is

it

form granules. The granules are then thoroughly dried,


to

screen
pressed, and sintered under the same conditions used for the fine UO,
dispersions. mixture containing about weight percent UO,
48
A
an

average
density percent theoretical with
96
of
2 to

was sintered
open porosity (see Fig. 4.15B). How
of

approximately percent
greater amounts coarse UO, particles tend
of

of

additions
to

ever,
reduce the density and increase the amount open porosity.
of

4.4 URANIUM DIOXIDE SOLID SOLUTION CERAMICS


H. Handwerk, Hoenig, Ploetz, and
G.
J.

B.
L.

L.

E.

Schaner
C.

4.4.1 Systems under Investigation

Among the UO, solid solutions that have been and are being con
sideredfor nuclear fuel application are the systems Tho, U.O., ZrO,
CO, ZrO,-UO,-CaO, and CeO.—UO. Phase equilibria information
presented Chap. and irradiation behavior
in
is on

these oxides
is

Chap.
of

discussed Fabrication these materials discussed


in

9.

is

below.

Berrin and B. E. Schaner.


L.
* *

COX grade, Brush Beryllium Corp.


154 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

A BeO, 35.8 w/o UO, (Fine Particles); Sintered 10 Hours at 1,700° C in H.;
4.04 g/cc (98.9 Percent T.D.); No Open Porosity.

- - --
"
º
- --
-- -
…,

...
*

:
-
--- -
-

4
.

|
|
S.

BeO, 48.6 w/o UO, (Coarse Particles); Sintered


14
B.

at

Hours 1,750°
in
C

He; 4.48 g/cc (96 Percent T.D.); 2.1 Percent Open Porosity.

FIGURE 4.15. Microstructure of BeO-UO, X.400.


;
FABRICATION OF URANIUM OXIDE 155

4.4.2 Fabrication of Thorium Dioxide-Uranium Oxide


J. H. Handwerk and C. L. Hoenig

Usable fuel compositions have been fabricated from Thoa-UO,


solid solutions; these fuels were primarily Thoa spiked with suf
ficiently enriched UO, to cause criticality [45].

The urania-thoria solid solutions as prepared by Hund and Niessen


were formed by dissolving the nitrates of uranium and thorium in
water, evaporating to dryness, and calcining at 1,200 C [46]. Ander
son, et al., followed the precipitation techniques of Trzebiatowski and

Selwood, which consisted of running aliquot uranium and thorium


solutions into excess ammonia solution [47, 48]. The urania-thoria
solid solutions et al., were formed by sinter
prepared by Lambertson,
and Tho,
UO,
ing

of

or
hydrogen vacuo, whereas
in

in
mixtures

by

of
Handwerk, al., formed the solid solutions heating mixtures
et

U.0, and Tho, air temperatures between 1,300° and 1,800°


at
in

Thus, solid solutions UO, and Tho, can pre

be
of

[49, 50].
C

by

the following methods: (1) coprecipitation

of
any
of

pared

uranium and thorium salts, (2) sintering mixtures UO, and

of
soluble

(3) sintering mixtures


or

or

Th0, protective atmosphere vacuum,


in
a

U.0, and Tho, without the use protective atmosphere.


of
of

Initial work involving solid solutions UO, Tho,


of

and was done


by

by

with UO, prepared U.O.


of

the reduction dissociated ammonia.


to

This material was screened pass 325 mesh. The UAOs used form
in

UO,-Tho, solid solutions


by
ing

air-fired was prepared heating UO.


air. This calcination was continued until the brown
in

1,000°
C
to

UO,
of

changed the U.O. black


to
of

color the the characteristic


oxide. X-ray diffraction patterns and after heating made before
indicated complete conversion U.O. Equally good U.O. was also
to
by

oxidizing uranium metal chips the same manner. How


in

prepared

the conversion of uranium dioxide or uranium metal


to

U3Os
as

ever,

always convenient, UO, air


to

was dissociated obtain the


in

was not

U0. This material was usually fine, black powder


of

the form
in

325-mesh screen. However, some lots had


an

passing
of

capable
a

olive green cast. Although X-ray diffraction patterns indicated that


this material was U3Os, noticeable weight change occurred when
a

these samples
all of

air. The color


in
to

1,000°
were heated
C

Samples

from olive green black; therefore, U.O. was cal


to

also changed
prior use. The Tho, also exhibited some shrink
to

1,000°
at

timed
as
to

air and,
in

age when heated 1,000° the calcined material was


C

Tho,
all

have improved pressing characteristics, was cal


to

found
prior pass 325 mesh.
of to

to

"ined use and screened


The first the three methods that have been used preparing
in
of

urania-thoria solid solutions involves the soluble salts uranium


156 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

and thorium, such as UO, (NOA), .6FI.O and Th (NO3), which are
precipitated by such materials as ammonia and oxalic acid. However,
concentrated solutions of these materials can be evaporated to dryness
and the residue dissociated by heating to 600° C. This residue, as
indicated by X-ray diffraction patterns,
appeared to be in the form of
a fine-grained solid solution UO,-ThC).
The low calcination temper
ature produces a material which can be easily crushed to any particle
size for fabrication.
In the second method of preparing urania-thoria solid solutions,
UO, and Tho, powders are mixed by grinding in porcelain-lined
pebble mills. There are some indications that dry grinding the UO2
and Tho, for periods of 3 to 4 hours is not sufficient to produce mix
tures which will form good solid solutions on sintering. Therefore,
these materials are usually mixed with sufficient water to form a
thin slurry, then milled in a pebble mill for a period of 3 to 4 hours.
This treatment is usually sufficient to produce an intimate mixture
which will sinter to a solid solution. The slurry is dewatered and
dried, and the cake is granulated to form a powder suitable for
fabrication.
In the third method, UAOs and Tho, are mixed by grinding in
porcelain-lined pebble mills. Dry grinding for periods of 3 hours
results in an intimate mixture which produces a homogeneous solid
solution when fired in air. Wet milling may also be used; however,
as the UAOs appears to be a very soft material, good mixes with Thos
are obtained without the need of forming a slurry.
The powdered, coprecipitated urania-thoria solid solutions or the
mixtures of either UO, and Tho, or U.O, and Tho, are formed by
common ceramic means such as dry pressing, hot pressing, and slip
casting.
Two general dry pressing procedures have been used: one for pro
ducing a small number of pieces, the other for a large number of
pieces. A small number of pieces can be formed by merely pressing
the dry powders in a steel die under pressures from 10,000 to 20,000
psi. Dry strength can be improved by adding a small amount of
organic binder, usually from 1 to 3 weight percent. A simple pro
cedure is to dissolve a small amount of polyethylene glycol in hot
water and then to temper the dry powder with about 3 percent by
weight of this solution. Compacts formed have been isostatically
pressed at pressures as high as 54,000 psi to improve green density
prior to firing.
Automatic pressing of a large number of pieces requires the feed
material to be free flowing. Mixtures of UAOs and Tho, or the
powdered coprecipitated solid solutions are not free flowing and must
be granulated. To accomplish granulation, the oxides are mixed with
from 1 to 3 percent polyvinyl alcohol. The mixture is dampened with
FABRICATION OF URANIUM OXIDE 157

approximately 12 percent water containing a small amount of wetting


agent, and the damp mixture is forced through a 16-mesh screen to
form granules which are dried at 80° C. Before pressing, 1 percent of
a mixture of 50 percent kerosene and 50 percent oleic acid is added as
a lubricant. This procedure produces a free-flowing material which
can be pressed in steel dies at pressures from 13,000 to 20,000 psi.
Mixtures of UO, and Tho, and UAOs and Tho, have been slip cast
by a procedure similar to that used for slip casting UO, [14]. The
mixtures of UO, and Tho, were ground with water in steel pebble
mills with steel balls for 72 hours. The slurry was acid leached with
hydrochloric acid to remove the iron contaminate, then washed with
water. The resultant slip has a pH of 2 to 4 and is suitable for casting
into plaster molds. The mixtures of UAOs and Tho, were cast by
grinding the ThC), in steel pebble mills, followed by an acid leach to
remove the iron contaminate. The pH of the acid slip was adjusted to
approximately 6 by adding a small amount of ammonium hydroxide;
U.Os was added and the mixture was wet milled in a porcelain-lined
pebble mill for 3 hours, then cast into plaster molds.
Mixtures of UO, and ThC), and UAOs and Tho, as well as dried,
coprecipitated solid solutions have been hot pressed. Maximum pres
sures with the use of graphite plungers are 1,200 to 1,500 psi. Higher
pressure (up to 5,000 psi) can be reached by using tungsten car
bide plungers. Mold temperatures of 1,400° to 1,800° C have been
used with no evidence of carbide formation in the urania-thoria bodies.

If tungsten carbide plungers are used, their temperatures should be


held below 1,400° C outside the mold. This precaution is necessary
because at sufficiently high temperatures the cobalt matrix used in
the tungsten carbide plungers begins to soften, and that portion of the
plunger outside the mold no longer transmits pressure but begins to
buckle. Hot zones at the sample level can be maintained without ex
ceeding the temperature limitation of the tungsten carbide by using
a long graphite mold. The extremities of a mold 18 inches long are
cool enough adequately to contain the tungsten carbide plungers.
Within the mold the pressure is transmitted by means of graphite
down to the sample and, as the mold case is supported on a separate
plate, the sample is pressed from both ends.
The urania-thoria ceramics may be sintered in hydrogen, in vacuo,
in an inert atmosphere, or in air, depending upon whether or not
U.O. or UO, is used in their preparation. Mixtures of UO, and
Tho, must be fired in a protective
atmosphere or in vacuo to prevent
the oxidation of the UO, to
UAOs. When firing in hydrogen is used,
homogeneous solid solutions are achieved with some difficulty. Inti
mate mixing, maximum temperatures of 1,700° to 1,800°C, and long
soaking periods are frequently necessary to attain a homogeneous
158 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

solid solution. Vacuum firing has the limitation in that the vapor
pressure of uranium dioxide is high enough to cause some loss in
material at temperatures sufficiently high to form solid solutions
(see Chap. 5). When used in forming these mixtures, organic binders,
such as polyethylene glycol, are removed during the initial sintering
stages by slowly heating the furnace to 600° C.
The coprecipitated solid solutions and the mixtures of UAOs and
ThC), up to 78 mole percent UO, may be fired in air or other con
venient atmosphere. Ceramic bodies formed from these materials are
first slowly heated to 500° to 600° C to dissociate organic binders
which may have been used in forming, then fired at 1,400° to 1,800° C.
Probably the most important property of a fired urania-thoria body
is bulk density. Such factors as particle size, forming pressure, binder
content, and sintering temperature influence bulk density in urania
thoria fabrication as much as they do for other ceramic bodies. There
are other considerations, however, which aid in further densifying a
sintered urania-thoria compact. The fired bulk density is also de
pendent on whether or not the starting materials are mixtures of the
oxides or are in the form of solid solutions. It appears that the use of
U.O., in forming the solid solutions results in higher densities at lower
sintering temperatures. However, as the UAOs content decreases, its
effect on density is less pronounced, and smaller particle size, as ex
emplified by the coprecipitated solid solutions, becomes more impor
tant in attaining high densities.
Both types of urania-thoria solid solutions were made: UO, with
ThC), and U.O, with Tho. The U.O, and Tho, mixtures were pre
pared by wet milling in a porcelain mill for 3 days, then calcining at
600° C. Preforming without binder was performed at 6,000 psi to
produce 34-inch-diameter pellets approximately 1 inch long. These
pellets were then pressed isostatically at 40,000 psi prior to firing.
The codigested solid solution, formed by dissolving uranyl and tho
rium nitrates in water, was evaporated to dryness, then calcined slowly
to 600° C in air overnight to decompose the nitrates. The residue,
in the form of a coarse aggregate, was ground to pass a 60-mesh screen
X-ray diffraction patterns showed line broadening, characteristic of a
very fine particle size distribution. It would appear that the agglom
erates had sintered somewhat at 600° C. This material was ground
to pass a 325-mesh screen, preformed without binder at 6,000 psi, and
isostatically pressed at 54,000 psi prior to firing.
Mixtures of U.O.s and Tho, were fired in air, and draw trials were
made at various temperatures (1) to determine the temperature at
which solid solution is complete and densification most rapid and (2)
to ascertain whether or not the solid solutions lose oxygen at high tem
peratures at a rate sufficient to cause cracking of the sintered body.
Six pellet specimens, 34 inch in diameter and approximately 34 inch
FABRICATION OF URANIUM OXIDE 159

high, were pressed at 20,000 psi from mixtures of 70 weight percent


UO, (added as U.Os) and 30 weight percent Tho. The mixture
contained 3 percent polyethylene glycol, which was carefully purged
from the pellets by slowly heating them to 600° C. No cracking was
visible at this point. All specimens were slowly heated to 1,000° C.
The first specimen was removed from the furnace after 3 hours at this
temperature and allowed to cool in air to room temperature. Similar
draw trials were made at 100° intervals up to 1,500° C. X-ray anal
yses were made of all specimens; the information from these tests is
tabulated in Table 4.2. The data in Table 4.2 indicate that the follow
ing changes take place in firing a UAOs-Tho, mixture to 1,500° C or
higher:

TABLE 4.2–PROPERTIES OF A 70-MOLE PERCENT UO,4–30-MOLE PER


CENT Tho, MIXTURE UNDER WARIOUS FIRING CONDITIONS

Diametrical Percent T.D. on


Drawtrial tempera- firing shrink- | Bulk den- basis of 10.67 Crystal phasespresent
turein air (° C) age (percent) sity (g/cc) g/cc for
U0.1Tho.3O2.0

1,000----------- 0 4.2 39 || Orthorhombic and cubic


fluorite
1,100--- - - - - ---- 0. 5 4.7 44 | Orthorhombic and cubic
fluorite
1,200--- - - - - ---- 4.5 5.9 56 | Orthorhombic and cubic
fluorite
1,300-- - - - - 10. 7 8.5 80 | Cubic fluorite
1,400**_ _ _ _ _ _ _ _ _ 12.2 9.7 92 | Cubic fluorite
1,500**_ _ _ _ _ _ _ _ _ 12.9 9.8 92 | Cubic fluorite

"U0: added as U3Os.


"The specimens had 2 longitudinal cracks, but could still be evaluated.

1. Solid solution is completed somewhere between 1,200° and 1,300°


C, and densification is most rapid in this temperature range.
2. The UAOs apparently forms a solid solution with Tho, and in doing
so releases oxygen. This is in accord with the results of Anderson,
et al., who found the oxidized composition limit of Uo. Tho...O.0 to
be Uo Tho...O.s, [47]. Since the O/U ratio for UAOs is 2.67, only
partial decomposition of UAOs occurred at this point. The release
of oxygen apparently does not cause cracking because a consider
able number of open pores still remain in the partially sintered
body.
3. From 1,300° to 1,400°C, sintering nears completion, and the body
is essentially nonporous.
4. Above 1,400°C, cracks occur in the specimens. Biltz and Mueller
have shown rapid dissociation of UAO, under a constant O. pressure
of 10 mm Hg between 1,400° to 1,500°C, the reaction being UO, an
160 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONs

to UO2.2 (see Chap. 6). Assuming that the nonstoichiometric


UO2–ThC), solid solutions are likely to dissociate in a manner simi
lar to that of UAOs, the rapid oxygen release within an already
dense body could then be responsible for the cracking of the speci
mens fired to 1,400° and 1,500° C.
The fired shrinkage and bulk density of two UAOs—Tho, composi
tions were studied. One composition was 70 weight percent UO,
(added as UAOs) and 30 weight percent ThC), : the fabrication and fir
ing procedures were those previously cited. The second composition
was a mixture of 10 weight percent UO, (added as U3Os) and 90
weight percent Tho. The latter composition was dry blended for 3
hours. The dry powder was moistened with a water solution contain
ing 10 percent polyethylene glycol. The damp powder was then
pressed into cylinders 14 inch in diameter by 1% inch in length under
a pressure of 15,000 psi. Some of the pellets were fired in electric
furnaces at various temperatures up to 1,500° C. After a 1-hour soak
at peak temperature, the pellets were allowed to cool in the furnace to
room temperature. For from 1,600° to 1,800°C, a gas
temperatures
fired kiln was used. The linear firing shrinkage and bulk density
were evaluated for each temperature. These data are plotted in Figs.
4.16 and 4.17.

The shrinkage for the 10-percent UO, body appears to be linear up


to 1,300° C (Fig. 4.16), where the rate changes. Above this tem
perature the shrinkage rate is essentially constant. An inflection is
also apparent in the shrinkage curve for the 70-percent UO2 body.
After 1,300° C the shrinkage is again linear and appears to be at
approximately the same rate as the high-temperature shrinkage of the
10-percent body. This rapid initial shrinkage in the 70-percent UO,
body is not fully understood. It could, however, be attributed in part
to the more intimate mixing and fine particle size of this body (due to
wet milling) and also to the high forming pressure used in fabricating
these specimens.
The bulk density of the 10-percent body was found to increase with
firing temperatures up to 1,800° C. A slight change in the rate of
increase in bulk density of this body is apparent at 1,300° C (Fig.
4.17), whereas above this temperature the rate appears to be essentially
constant. The rate of change of the 70-percent body is greater than
that of the 10-percent body, and this rate again appears to change at
1,300° C. This difference could also be due to the different prepara
tion procedures for the two bodies.
The final densities of the urania-thoria solid solutions are appar
ently dependent upon the processing variables and, based on the lim
ited knowledge presently available, the following general conclusions
appear to be justified:
FABRICATION OF URANIUM OXIDE 161

2O I I
O - loº.'uoz* -
18H-
A— 70% uoz*
16

14H. -
i2

io | -
i 8

6H- -
*ADDED As AN EQUIvaLENT
4
AMOUNT OF U3O8
2H -
O
iOOO I loo |2OO 1300 |4OO 1500 16OO 17OO 18OO
FIRING TEMPERATURE, "C

FIGURE 4.16. Shrinkage vs Firing Temperature for Two Urania-Thoria


Compositions.

IO

–—T’
9 —r

§
8
Pr
7 2%
O – Io ". Uoz *
A – 70 % UO2 *
| e

* ADDEDAS AN EQUIVALENT
AMOUNT OF U30s

4
rooo lioo 12OO 13OO |400 1500 I6OO 17OO 18OO
FIRING TEMPERATURE, *c

FIGURE 4.17. Bulk Density vs Firing Temperature for Two Urania-Thoria


Compositions.

1. The coprecipitated or codigested solid solutions are fine grained


and will result in ceramic bodies having higher densities than those
formed from UO2–ThC), or UAOs—Tho, powder mixtures.
2. Fine-grained raw materials, absence of organic binders, and high
forming pressures usually result in higher final densities.
162 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

3. UO,-Tho, mixtures sintered in a protective atmosphere are not


resistant to oxidation and will when reheated in air.
disintegrate
This effect is thought to be due to the incomplete solid solution
formed during the single sintering operation.
4. Codigested solid solutions or solid solutions formed by sintering
mixtures of UAOs and Tho, are resistant to oxidation and may be
reheated for long periods at temperatures of 1,200° to 1,300° C
without detectable dimensional or weight changes.
5. In practice, densities of 90 percent of theoretical are readily ob
tainable, while densities of 95 percent of theoretical can be obtained
with careful process control.

4.4.3 Fabrication of Zirconium Dioxide-Uranium Dioxide


B. E. Schaner

In early investigations of the use of ZrO2—UO, solid solutions as


possible fuel materials, samples were prepared with UO, content rang
ing up to 15 weight percent on the assumption that the UO, would sta
bilize the zirconia in the tetragonal phase and would, thus, prevent
phase transformations during further exposure to heating and cooling.
However, during exposure of these samples to high-temperature water
and steam, complete disintegration resulted. X-ray examination of
the samples showed the presence of a monoclinic phase along with the
tetragonal. Later phase diagram work indicated the presence of a
two-phase field up to about 25 weight percent UO, and another two
phase region consisting of a tetragonal and cubic phase from 25 weight
percent to 93 weight percent UO, [52, 53]. Mixtures with UO, con
tent over 23.5 weight percent showed excellent resistance to high
temperature water and steam.
The ZrO2–UO, ceramics are prepared by first mixing hafnium-free
ZrO2 and UO, powders in the amounts desired. The mixture is then
loosely compacted into blocks which are treated at 1,700° C in hy
drogen for a minimum of 14 hours to react the components thoroughly.
The treated blocks are crushed, ball-milled, and granulated with 2
weight percent polyethylene glycol binder. Alumina has been used
to ball-mill ZrO.—UO, powder, but abrasion by the hard powder re
sulted in alumina contamination of up to 2,000 ppm. Uranium metal
cylinders are effective in grinding, but care must be taken to minimize
uranium pickup which can alter the composition by a few percent.
the de
tsi

The powder is then cold pressed at approximately 20


to

hydrogen atmos
at

sired shape and sintered


in
to

1,650° 1,750°
C

phere high density (see photomicrographs Fig. 4.18).


to

in

ZrO2–UO, can radically change its microstruc


of

Heat treatment
ZrO2–34 weight percent UO, annealed
of

ture [54]. Sintered bodies


FABRICATION OF URANIUM OXIDE 163

A. ZrOr-18 w/o UO, ; Sintered 15 Hours at 1,655° C in He, Annealed 89 Hours


at 1,740° C in Hs; 97.2 Percent T.D.

fºr
B. ZrO2-26 w/o UO, : Sintered 14 Hours at 1,690° C in Hz, Annealed 89 Hours
at 1,740° C in H, ; 98.3 Percent T.D.

C. ZrOr—46 w/o UO, : Sintered 15 Hours at 1,605° C in Hz, Annealed 89 Hours


at 1,740° C in Ha; 98.8 Percent T.D.

FIGURE 4.18. Microstructure of ZrO2—UO, ; X400, Reduction Factor 2/5.


-----a - a-- -a
164 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

at 1,750° C for 89 hours increase in grain size from


10 microns to 75
microns. Annealing ZrO2–34 weight percent UO, at 1,400° C results
in the fine precipitation of another phase.
Another method of stabilizing ZrO2–UO, (even with UO, contents
lower than 15 weight percent) is by the addition of a stabilizing
agent, for example CaO [55]. As little as 4 weight percent CaO is
sufficient to stabilize the fuel against attack by high-temperature
water. However, the addition of CaO does not, in all cases, preserve
the single-phase structure obtained with ZrO2—UO, compositions, but
introduces another phase.
The preparation of this ternary composition is generally the same
as that for the binary except that the ZrO, is first reacted with CaCO,
at 1,700° C. This material is then crushed prior to mixing with
UO2. The briquetting, calcining, granulating, pressing, and sintering
in H, are identical to those used for the ZrO2—UO, system. A liquid
phase, probably due to the presence of alumina impurities picked up
from the furnace atmosphere, is present during sintering and accounts
for the high-density, pore-free structure (Fig. 4.19A). Single-phase
face-centered cubic structures have been observed in samples of
ZrO2–CaO—UO, prepared with stabilized zirconia (stabilized with 10
weight percent CaO) and sintered in vacuum at 1,800° C (Fig. 4.19,
B) [56]. Samples containing zirconia stabilized with 15 and 20
weight percent CaO exhibited, in addition to the cubic phase, a
twinned calcium zirconate phase [56].
It is interesting to note that both ZrO,-UO, and ZrO2–CaO—UO.
have higher strength and hardness than UO. Even thin (0.040 inch)
sintered platelets may be handled without breakage; however, grind
ing is more difficult because of the increased hardness.

4.4.4 Fabrication of Cerium Dioxide-Uranium Oxide


B. E. Schaner and G. L. Ploetz

Fabrication studies [19, 57] have been made to determine the feasi
bility of using CeO2 as the diluent in a binary solid solution fuel system
with uranium oxide. Ploetz, et al., [57] undertook the development of
stable CeO2–UO, of 60 to 70 percent of theoretical density. Two
methods of fabrication were investigated: cold pressing and extru
sion. For cold pressing, the oxide powders were mixed with 2 weight
percent Ceremul–C emulsified wax binder, and pressing was done in
FABRICATION OF URANIUM OXIDE 165

/* *::/

&
‘ ‘e *;
y

-- > 3* A

(
/
A. ZrOr–13 w/o CaO-15 w/o UO2: Sintered 14 Hours at 1,750° C in H, ; 5.49
g/cc (98 Percent T.D.); No Open Porosity; X.400.

- •, - -
.e.”
.

'"
...
• "

º

th

*-
..
.

*
,

-
-
.

º,
e

"

-
.

-

©
-
~

- * -
-
e

O f,
•.

-
.*
O
º
e.
º

O O
e

2-

-
C. **
{{

*e
(;
.

*
;

O
*

iſ

wº-
- O
"e

o

O v-->

.* -
.*
.
.

©
*
-
-

- •".
">

-
*
-
.e.
-

's
--

*-

e” .* .* -
-
º

e.
-
2

~
*

-
-

*. Q
\"

- ->
2,

*.* -

- -
*
.
.
.

e
*

*
*

-
-
-

/•

-
-
-

O ©
*.* Tº
...
2
Q
&

‘.
e

-
e.
-

.
.

-

2.

- -Tº
- ->
-

.

o
*
*

".
"

-
O.
.."
º
ſº
.

*

"a
•.

©
g
2

- O •*
:

/ry ©
-
-
e
-
.. -


O
º

-

* ,

-
-
0.

o ...
.

Ce
~

-
~
o

*
-

-

º
-

2_* -
&

-

-
-
O

- GP
- 10
-
-
-
*
*-


-

f
@

-
*
-
-
º

© -
º
o

ZrO2–9 w/o CaO-10 w/o UO.


B.

at

Sintered 1,800° Vacuum: Etchant


in
C
:

45 v/o HNO3, v/o HF, 50 v/o H2O: X500 [56]. (Courtesy, G. Zuromsky and
5

W. Chernock, Nuclear Division, Combustion Engineering, Inc.)


P.

FIGURE 4.19. Microstructures of ZrO2–CaO–UO,.


166 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

double acting steel dies under a pressure of 16.5 tsi. The plasticizer
used for the extrusion process was methylcellulose (1.1 weight per
cent). Extrusion, however, was not as successful as the cold-pressing
technique.
Sintering was carried out in a hydrogen atmosphere for CeO2–UO.
mixtures or in air when CeO2–UAOs mixtures were used. The CeO2–
UO, bodies that were fired in hydrogen resulted in a face-centered
cubic single-phase solid solution, as determined by X-ray diffraction.
These solid solutions were tested for stability in an oxidizing atmos
phere by firing compositions of 15 and 30 volume percent UO2 to

determine weight gain at temperatures up to 1,500° C. The maximum


gain in weight was only 0.062 percent, an indication of the excellent
stability of these solid solutions under oxidizing conditions. Appar
ently, the CeO2 stabilized the UO, structure, preventing oxidation
to U.Os.
Single-phase solid solutions having the same X-ray diffraction pat
tern as that of hydrogen-fired CeO2–UO, may be readily formed by
air firing U.O., mixtures in an oxidizing atmosphere. Handwerk, et
al., [58] reported that it was possible to form single-phase solid solu
tions with the Tho, UO, system by starting with Thos-U,Os mix
tures. Since CeO2 has the same crystal structure as Tho, and UO,
it is reasonable to expect that the CeO2–U,Os system would be compar
able to the Thoe-U,Os system under oxidizing conditions. This sim
ilarity was borne out by air firings on materials containing 75, 50, 30,
and 15 volume percent of UO2. The UAOs reduces to UO, above
1,150° to 1,400° C by loss of oxygen, and UO, then becomes stabilized
by the solid solution formation with the CeO2. Thereafter, upon
cooling, the UO, does not convert back to U.O.. Hydrogen sintering
reduces the CeO2 to Ce2O, and, although a dense body can be produced,

the Ce2O3 reverts to CeO2 upon exposure to air at room temperature,


resulting in weakening of the body.” This effect is evident from a
comparison of the mechanical strength of bodies made by the two
methods.

* According to Bevan, a number of ordered phases exist in the system CeO2–Ce2Os [59].
FABRICATION OF URANIUM OXIDE 167

Specimens with a bulk density of 67.7 percent of theoretical had an


average modulus of rupture of 11,230 psi, whereas specimens of
comparable density fired in hydrogen had an average modulus of
rupture of 1,500 psi. This strength appears to be very sensitive to
high strength
the

Specimens
of

of
oxygen content the solid solution.
hydrogen annealed 1,000°C, and the strength was redeter

at
were

The modulus of rupture decreased from 550 psi,

to
mined. 9,670

bulk density decreased from 59.5 51.8 percent

of
to
and the theoretical.

on
this strength dependency oxygen content, the air-fired
To

check

original modulus rupture psi and


an

of
of
Specimens with 9,670

at a
density percent
of

bulk 59.5 theoretical were vacuum annealed


The strength psi and the bulk

be
to
1,000°C. was determined 9,770

percent
67
be

of

theoretical.
to

density

Wagner, and Peetz studied the crystal structure

of
Hund, CeO2–

up
by

heating the mixed oxides air [60].


in

to
U0, mixtures 1,200°

C
reported that the products air showed
heated single-phase
in

They
a
mole percent urania (see Chap. 6). Beyond
63

63
of
to

region from
0

percent, two-phase region which consisted the CeO2–UO,


of

mole
a

orthorhombic U,Os was en


of

mixture with fluorite structure and


a

Ploetz, al., extended the single-phase


of

The work solid


et

countered.
up

75

of
at

range mole percent urania when fired


in
to

solution least
air,

undoubtedly because the higher temperatures employed


of

[57].
CeO,-UO, having
of

Schaner produced samples


95
of

densities
by

This accomplished
of

percent theoretical [19]. was first


in by

briquetting mixed CeO,-UO, powders, followed calcining air


in

1,500°C for hours with heating and cooling argon. This


14

is
at

by

followed crushing, ball milling, granulation with weight per


1

20

Carbowax binder, and pressing The


in

at

cent steel die tsi.


a

1,500°C for
at

14

then sintered air hours.


in

Samples are
by

Cerium oxide-uranium dioxide bodies have been ground lapping


with abrasive grit. While such bodies are softer than UO, and, thus,
readily, care must handling because
of
be

in

grind more exercised the


strength
of

lower such material.


168 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

The CeO2—UO, bodies were found to be very reactive with the


usual ceramic refractory setting materials. Reactions were noted with
Al2O3, ZrO2, molybdenum, and tungsten in the hydrogen firings and
with MgO in the air firings. Platinum foil was fairly satisfactory
as a setter in the air firings, but some adherence of the samples was
noted. Best results were obtained with BeO setters below 1,600° C.;
no visible reaction or sticking occurred. Above 1,600° C, however,
reaction occurred with BeO. Rhodium was used satisfactorily above
this temperature in both oxidizing and neutral atmospheres.

REFERENCES

1. T. J. BURKE, J. GLATTER, H. R. HoGE, and B. E. SchANER, “Fabrication of


High Density Uranium Dioxide Fuel Components for the First Pressurized
Water Reactor Core” in “Nuclear Metallurgy (Vol. IV), A Symposium
on Uranium and Uranium Dioxide,” pp. 135–143, New York, AIME, 1957.

2. B. E. SchANER, “Fabrication of High Density UO, Fuel Platelets,” Am. Ceram.

Soc. Bull. 38, 494–498 (1959).


3. A. G. ALLISON and W. H. DUCKworTH, “Ceramic Investigations of UOs,”
BMI–1009,
June 1955.
4. E. A. Evans, “Fabrication and Enclosure of Uranium Dioxide” in “Fuel
Elements Conference, Paris,” TID–7546, Mar. 1958, pp. 414–431.
5. L. C. WATSON, “The Production of Uranium Dioxide for Ceramic Fuels,”
CRL–45, Nov. 1957.

6. S. F. PUGH, P. MURRAY, and J. WILLIAMs, “Uranium Dioxide as a Reactor


Fuel” in “Fuel Elements Conference, Paris,” TID–7546, Mar. 1958, pp.
432–441.

7. P. MURRAY, D. T. LIVEY, and J. WILLIAMs, “The Hot Pressing of Ceramics"


in “Ceramic Fabrication Process,” W. D. Kingery, ed., pp. 147–171, John
Wiley and Sons, New York, 1958.
8. P. MURRAY and J. WILLIAMs, “Ceramic and Cement Fuels” in “Proceedings
of the Second United Nations International Conference on the Peaceful
Uses of Atomic Energy, Geneva, 1958,” Vol. 6, pp. 538–550, United Nations,
Geneva, 1958.

9. C. G. GoetzEL, “Treatise on Powder Metallurgy,” Vol. I, Interscience Pub


lishers, Inc., New York, 1950.

10. W. D. KINGERY, ed., “Ceramic Fabrication Processes,” John Wiley and Sons,
New York, 1958.
FABRICATION OF URANIUM OXIDE 169

11. J. R. Johnson, “Résumé of UO, Studies at ORNL Ceramic Laboratory” in


“Résumé of Uranium Oxide Data-II,” WAPD—PMM-167, June 1955, p. 79.

12. M.D.. BURDick and H. S. PARKER, “Effect of Particle Size on Bulk Density
and Strength Properties of Uranium Dioxide Specimens,” J. Am. Ceram.
Soc. 39, 181–187 (1956).
13. H. W. NEwkIRK, Jr., and R. J. ANICETTI, “Fabrication of Uranium Dioxide
Fuel Element Shapes by Hydrostatic Pressing,” HW–51770, Aug. 1957.
14. R. E. Corwin and G. B. EYERLY, “Preparation of Refractories from
Uranium Dioxide,” AECD 3349; J. Am. Ceram. Soc. 36, 137–139 (1953).
15. P. MURRAY, “A General Interpretation of the pH-Wiscosity Relationships for
Amphoteric Ceramic Oxide Slips,” AERE-M/R—507, Apr. 1950.
16. P. J. ANDERSON and P. MURRAY, “Zeta Potentials in Relation to Rheological
Properties of Oxide Slips,” AERE–M/R–2482, Jan. 1958.
17.

EYERLY, W. A. LAMBERTson, KRAFT, and Corwin, “Prepara


G.

E.
R.

R.
B.
G.

tion of Urania Refractories,” ANL–5038, Mar. 24, 1953.

18. GLATTER, “Fabrication of Bulk Form Uranium Dioxide for Use as Nuclear
J.

Reactor Fuel” in “Nuclear Metallurgy (Vol. IV), Symposium


A on Uranium
and Uranium Dioxide,” pp. 131–134, New York, AIME, 1957.
19. “Pressurized Water Reactor (PWR) Project Technical Progress Report for
the Period April 24, 1957, June WAPD–MRP–68.
to

23, 1957,”

M. SHAPIRo, “Résumé Uranium Dioxide Data–IV,” WAPD—PWR—PMM


of
Z.

20.
5,

417. Dec. 1955.

BELLE and Jones, “Résumé of Uranium Dioxide Data—IX,” WAPD


J.

L.
J.

21.
TM-44, Mar. 15, 1957.

ANICETTI, “Extrusion Fuel


R.

R.

of
D.

STENQUIST and Uranium Dioxide


J.

22.
Cores,” HW–51747, Dec.
1,

1957.
ANICETTI, “How UO, Cores Are Extruded,”
R.

J.
D.

23. STENQUIST and R. Ceram.


Indust. pp. 102–103
4,

71, No. (1958).


24.

RoAKE, “Compaction UO, by Swaging,” HW–


B.

of
E.

QUINLAN and W.
F.

45631, Sept. 20, 1956.

H. BRIGHT, PATERson, and WATSON, “The


D.
N.

C.
L.

H. CHALDER,
G.

F.

L.

25.

Fabrication and Properties of Uranium Dioxide Fuel” “Proceedings of


in

the Second United Nations International Conference on the Peaceful Uses


Energy, Geneva, 1958,” Nations,
6,

pp. 590–604,
of

Atomic Vol. United


Geneva, 1958.
28.

R.

ANICETTI, “Fabrication
of

Ceramic Fuel Elements


R.
D.

STENQUIST and
J.

by Swaging.” Metallurgy AIME


5,

Nuclear pp. 1–12, (1958).


27.

MARRIott, “A Replica Technique for Particle Surfaces,” Brit. Appl.


J.
J.

Phys. 7,373–374 (1956).


170 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONs

28. D. R. SteNQUIST, B. MASTEL, and R. J. ANICETTI, “Note on Correlation of


Surface Characteristics of Uranium Dioxide Powders with Their Sintering
Behavior,” J. Am. Ceram. Soc. 41, 273–274 (1958).
29. W. C. BELL, R. D. DILLENDER, H. R. LoMINAc, L. W. LoNG, and E. G. MANNING,
“Vibratory Compacting of Metal and Ceramic Powders,” WADC-TR-53–
193, Apr. 1953.

30. W. C. BELL, R. D. DILLENDER, H. R. LoMINAc, L. W. LoNG, and E. G. MANNING,


“Vibratory Compacting of Metal and Ceramic Powders,” WADC-TR-53–
193 (Pt. II), Apr. 1954.
31. H. R. Hoge, “An Interim Report on Vibratory Packing for the Period of
October 1954 to April 1955,” WAPD–FE–958.
32. The Staff of Ceramic Fuels Development, Hanford Laboratory Operation,
“Novel Ceramic Fuel Fabrication Processes,” HW–64629, Apr. 15, 1960.
33. J. E. CUNNINGHAM and E. J. BoyLE, “MTR-Type Fuel Elements” in “Pro
ceedings of the International Conference on the Peaceful Uses of Atomic
Energy, Geneva, 1955,” Vol. 9, pp. 203–207, United Nations, New York,
1956.

34. R. C. WAUGH and J. E. CUNNINGHAM, “The Application of Low-Enrichment


Uranium Dioxide to Aluminum Plate-Type Fuel Elements,” CF–56–8–128,
Aug. 20, 1956.

35. H. BERGUA, R. FRIDDLE, and J. BAIRD, “Fabrication of an Aluminum Clad


Plate Type Fuel Element Containing a Core of 55 w/o U20s–45 w/o Alumi
num,” paper presented at Am. Nuclear Soc., June 1958.
36. H. H. HIRscH, “A Powder Metallurgy Technique for Combining Insoluble
Materials as Applied to Al-UO, Fuel Elements” in “Proceedings of the
Metallurgy and Materials Information Meeting, Oak Ridge, Tenn.,” TID
5061 (Del.), Vol. I, Apr. 1951, pp. 527–537.

37. J. E. CUNNINGHAM, R. J. BEAver, W. C. THURBER, and R. C. WAUGH, “Fuel


Dispersions in Aluminum-Base Elements for Research Reactors” in “Fuel
Elements Conference, Paris,” TID–7546, Mar. 1958, pp. 269–296.

38. H. G. Sow MAN and G. L. PLOEtz, “An Investigation of the Sintering of Ura
nium Dioxide,” KAPL-1556, Aug. 17, 1956.

39. C. E. WEBER, “Progress on Dispersion Elements” in “Progress in Nuclear


Energy, Series V, Metallurgy and Fuels, Vol. 2,” J. P. Howe, H. M. Fin
niston, eds., pp. 295–362, Pergamon Press, New York, 1959.
40. C. T. ARM ENoFF and M. H. BIN stock, “Fuel Elements for the Organic Mod
erated Reactor Experiment,” NAA—SR-1934, Dec. 15, 1957.
41. J. E. CUNNINGHAM, R. J. BEAVER, and R. C. WAUGH, “Fuel Dispersion in
Stainless-Steel Components for Power Reactors” in “Fuel Elements Con
ference, Paris,” TID–7546, Mar. 1958, pp. 243–268.
FABRICATION OF URANIUM OXIDE 171

42. S. StoRCHHEIM, “Metal Powder Rolling—A New Fabrication Technique,”


Metal Progr. 70, 120–126 (1956).
43. F. A. BIckford, F. E. BARR, F. I. PETERs, and L. W. PokALLUs. “The Fabrica
tion of UO,-Al-O, Fuel Plates,” Final Report, Corning Glass Works, Sub
contract No. 74 (14–423), Oct. 8, 1957.
44. L. F. EpstEIN and W. H. HowLAND, “Binary Systems of UO, and Other
Oxides,” J. Am. Ceram. Soc. 36,334–335 (1953).
45. J. H. HANDwekk and R. A. NoLAND, “Oxide Fuel Elements for High Tempera
tures,” Chem. Eng. Progr. 53, 60F-62F (1957).
46. F. HUND and G. NIEssen, “Anomalous Solid Solution in the System Thorium
Oxide-Uranium Oxide,” Z. Elektrochem. 56,972–979 (1952).
47. J. S. ANDERson, D. N. EDGINGTON, L. E. J. Roberts, and E. WAIT, “The
Oxides of Uranium. Part IV. System UOz–Thoz-O,” J. Chem. Soc., 3324–
3331 (1954).
48. W. TRzeBIAtowski and P. W. SELwood, “The Magnetic Susceptibilities of
Urania-Thoria Solid Solutions,” J. Am. Chem. Soc. 72, 4504–4506 (1950).
49. W. A. LAMBERTson, M. H. MUELLER, and F. H. GUNzel, Jr., “Uranium Oxide
Phase Equilibrium Systems. IV,” J. Am. Ceram. Soc. 36, 397–399 (1953).
50. J. H. HANDweRK, L. L. ABERNETHY, and R. A. BACH, “Thoria and Urania
Bodies,” Am. Ceram. Soc. Bull. 36, 99–100 (1957).
51. W. BILtz and H. MüLLER, “Systematic Affinity Principle. XLI. Uranium
Oxides,” Z. anorg.u. allgem. Chem. 163, 257–296 (1927).
52. G. M. Woltex, “Solid Phase Transitions in the UO,-ZrO, System,” J. Am.
Chem. Soc. 80, 4772—4775 (1958).
53. N. M. WoRoNow, E. A. VoITEKHowA, and A. S. DANILIN, “Phase Equilibrium
Diagrams of the UOz-ZrO2 and Thor—ZrO, Systems” in “Proceedings of
the Second United Nations International Conference on the Peaceful USes
of Atomic Energy, Geneva, 1958,” Vol. 6, pp. 221–225, United Nations,
Geneva, 1958.

54. B. E. SchANER, “ZrO,-UO, Ceramic Fuels: Some Microstructural Changes


Resulting from Phase Transformation,” paper delivered at the American

Ceramic Society Meeting, Philadelphia, Apr. 27, 1960.


55. P. DUwez, F. ODELL, and F. H. Brown, Jr., “Stabilization of Zirconia with
Calcia and Magnesia,” J. Am. Ceram. Soc. 35, 107–113 (1952).
56. G. ZURomsky, “The Development and Testing of Homogeneous Ceramic
Fuels. Progress Report for Period December 1, 1959–February 29, 1960,"

CEND-84.
57. G. L. PLoetz, A. T. MUCCIGRosso, and C. W. KRYSTYNIAK, “Properties of
Urania-Ceria Bodies for Nuclear Fuel Application,” KAPL–1918, Jan. 1958.
172 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

58. J. H. HANdweRK, C. L. HoeNIG, and R. C. LIED, “Manufacture of the Thor


UO, Ceramic Fuel Pellets for Borax–IV,” ANL–5678, Aug. 1957.
59. D. J. M. BEVAN, “Ordered Intermediate Phases in the System CeO2–CeOs,”
J. Inorg. Nuclear Chem. 1, 49–59 (1955).
60. F. HUND, R. WAGNER, and U. PEETz, “Anomalous Mixed Crystals in the Sys
tem Cerium Dioxide-Uranium Dioxide,” Z. Elektrochem. 56, 61–65 (1952).
Chapter 5

PHYSICAL PROPERTIES OF URANIUM DIOXIDE


W. H. DUCKworth, Editor

5.1 INTRODUCTION

The unirradiated physical properties of uranium dioxide are dis


cussed in this chapter. This information is organized under four
major groupings: crystal structure; thermal properties; mechanical
properties; and electrical, magnetic, and optical properties. Factors
that affect each property are discussed; these include temperature,
Stoichiometry, impurities, pore and grain structure, method of meas
urement, etc. Specific effects of irradiation are examined in detail
in Chap. 9.

5.2 CRYSTAL STRUCTURE


D. A. Vaughan

5.2.1 Lattice Parameter Variations

In the uranium-oxygen system, oxides over the composition range


UO, is to UO, have been reported to exhibit a face-centered-cubic

He–5.4704A–
FIGURE 5.1. Structure of UO2.

structure with the CaF2 arrangement of atoms (Fig. 5.1) [1].


Investigations by Perio, Gronvold, and Vaughan, et al., indicate that

173
174 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

solid solution does not occur below the stoichiometric composition of


UO2.00 [2–4]. Phase studies in the composition range above UO, on
show two distinct compounds, UO, and U.O., both having the fluorite
type structure' [2–5]. Room-temperature lattice parameters for the
two compounds are given in Table 5.1.
The theoretical density for essentially stoichiometric UO, calculated
from the lattice parameter measurements varies from 10.95 to 10.97
g/cc [3, 8,9]. A value of 10.96 g/cc for the theoretical density of
stoichiometric UO, oo is assumed throughout this book (see Chap. 3).

TABLE 5.1—ROOM-TEMPERATURE LATTICE PARAMETERS OF THE


FLUORITE-TYPE PHASES IN THE URANIUM-OXYGEN SYSTEM
Composition Ao
(anº units) Reference

UO,-, (UO2+ U) 5.4720 + 0.0007 6


UO1.9% 5.4699 + 0.0004 and 12
5.4701 + 0.0003
UO2 5.4691 + 0.0005 6
UO2 5.4678 + 0.0005 7
UO2.00 5.4690 + 0.0001 8
UO2.00 5.4704 + 0.0008 3
UO2.01 5.4708 + 0.0005 4
UO2.05 5,4682 + 0.0002 9
UO2.03 5.46852 + 0.00011 10
UO2.04 5.468 + 0.002 11
UO2.07 5.4662 + 0.0005 4

U40,

UO<2.3 5.4407 -- 0.0008 6


UO2.2 5.4462 + 0.0008 3
UO2.2, 5.4431 + 0.0005 4
UO2.25 5.438 + 0.003 8
UO2.25 5.4411 + 0.0008 3
UO2.25 a = 4Ao = 21.73 (*)
UO2.27 5.4402 + 0.0005 4
UO2.30-UO2.41 (active oxide) 5.437 + 0.001 4

*P. Perio, personal communication (seeChap. 6).

The variation in lattice parameter depends upon three factors: the


oxygen content, the degree of approach to phase equilibrium in the
sample, and the temperature. During low-temperature (180° C)
oxidation of UO, the lattice parameter decreases (see Chap. 8).
Perio [8] reported that during homogenization the cell size of the
single phase UO2.x, a metastable state, continues to contract according
to the equation
ar = 5.4690 – 0.12a: Eq. (5.1)
* A more complete discussion of phase relationships and stoichiometry in the uranium
oxygen system is presented in Chap. 6.
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 175

However, extended heat treatment (12 hours at 800° C, 20 hours at


700°C, or 3,000 hours at 140° C) of samples analyzing between UO2
and UO, as was found to disproportion the apparent solid-solution
oxide into two distinct phases, UO, and U.O., with no solubility
of U.O., in UO, at room temperature. These results agree with those
of Gronvold, whereas data obtained by Vaughan, et al., indicate that
after homogenization for 100 hours at 450° C the room-temperature
lattice parameter of the UO, phase decreases with increasing oxygen
content from 5.4720 + 0.0005 to 5.464 + 0.001 angstrom units at the
solid solubility limit of UO2,0s [3,4].
Schaner reported room temperature lattice parameter measurements
of the single-phase UO2.x structure as a function of O/U ratio on
samples that had been annealed at 900° C and quenched to room tem

H-H
perature (see Chap. 6) [5]. Metallographic examination of the micro
structure substantiated the presence of the single-phase UO2, only.
These results are shown in Fig. 5.2.
The temperature dependence of the cell size was determined by
Gronvold, by Vaughan, et al., and by Kempter and Elliott [3, 4, 10].
Gronvold obtained a thermal expansion coefficient of 10.8×10°/°C
for the temperature range 20° to 946° C, Vaughan reported

5480

...º. -

º 5.4 60 H -

5.450 H -

5.440 H. —

5.430 -

1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1
2.00 O2 O4 O6 O8 2.I.O. 12 .14 .16 .18 2.2O .22 24 2.25

u02 U/O ATOM RATIO U499

FIGURE 5.2. Lattice Constants of the UO. Phase at Different O/U Ratios;
Samples Annealed at 900° C for 67 Hours and Quenched to Room Temperature
[5]. (Courtesy, Journal of Nuclear Materials.)
176 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

9.9 × 10−"/° C over the range 25° to 800° C, and Kempter and
Elliott calculated an average value of 10.52×10°/°C for the range
26°C to at least 1,000° C. These results are in general agreement
with dilatometric measurements, discussed in Sect. 5.3.4.
Oxygen solubility in UO, was determined by Gronvold and by
Vaughan from high-temperature X-ray diffraction measurements and
by Schaner from metallographic studies. An increase in the oxygen
solubility limit with increasing temperature was found in all these
studies (Table 5.2). The results obtained by Vaughan, et al., are
somewhat higher than those obtained by Gronvold and by Schaner.

-
TABLE 5.2—SOLID SOLUBILITY OF OXYGEN IN UO, AS FUNCTION
OF TEMPERATURE

Solid solubility limit (O/U ratio)


Temperature (°C)

Vaughan, et al. [4]| Gronvold [3] Schaner [5]

25-------------------------------- 2. 08 2. 00 2. 00
200------------------------------- 2. 08 2. 00 2. 00
400------------------------------- 2. 11 2. 03 2. 02
600------------------------------- 2, 17 2. 11 2. 14
800------------------------------- 2. 20 2. 14 2. 17

5.2.2 Atomic Positions

The structure of UO, was determined by Cuy in 1927. It has the


space group O3–Fm3m with four (UO2) per unit cell [13]. This
information is given by Swanson and Fuyat along with a diffraction
powder pattern listing the interplanar spacings, intensity measure
ments, and Miller indices of the 16 peaks of largest spacing [9].
The atomic positions in the fluorite structure of UO, consist of four
uranium atoms at 000, #40, 30}, and 0}} plus 8 oxygen atoms at
+ (H+), with cyclic interchange. Interatomic distances based
upon the room-temperature cell size of 5.4704 angstrom units for
stoichiometric UO, are [3] :
U–12U = 3.868 A.

O – 6O = 2.735 Å
U–8() = 2.368 A.

Thus, the oxygen ion radius of 1.37 angstrom units is in good agree
ment with accepted values (1.32 and 1.40 angstrom units) [14]. The
largest interstitial holes in this structure are at the ###, H-400, 3-0%0,
and 00% positions (the holes are equal in size and equidistant from
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 177

eight oxygen atoms). The radius of the hole, 1.00 angstrom unit,
is smaller than the 1.37 angstrom units radius of a fully ionized
oxygen ion. This difference presents an interesting problem as to
be, the

the oxygen solid solutions. may

of

It
structural characteristics
however, that the strain energy which results from the introduc

by
the lowering

of
interstitial oxygen anion
an
of

overcome

is
tion
crystal energy through the formation the U" U" ion.
the

or
of
5.3 THERMAL PROPERTIES

5.3.1 Thermal Conductivity


Kingery,
A.
D.

W. M. Ross, and Belle

Factors Affecting Thermal Conductivity J.


(a)

Above room temperature the thermal conductivity

of
oxide ceramics
by

the scattering

to
of

mainly determined thermoelastic waves due


is

lattice vibrations. The thermal conductivity can


in

anharmonicities
the mean free path for phonon scattering according
to

to
be

related
the relation

Eq. (5.2)
C,

K=1%
vl

the thermal conductivity,


(',

the volume heat capacity,


K
is

is

where

v
the

velocity, and the mean free path [15–17]. The wave


is

wave
is

(E/p)”, where v-
by

approximated
be

the modulus
is
E

Velocity can
elasticity and the density. From experimental data (see below
is
p
of

and Sects. 5.3.2 and 5.4.2) for heat capacity, elastic properties, and
conductivity, the mean free path UO, corre
in

at
on

thermal 100°
approximately angstrom units. With increasing tempera
27
to

sponds

measured thermal conductivity decreases proportionally


to

ture the
at

17, until 1,000° the phonon mean free path length calculated
C

Eq. (5.2) approximately angstrom units. These results are


is

from
7

agreement with theoretical expectations and with experimental data


in
for

other materials.

decrease with temperature, the mean free path for


to
In

addition
a

thermoelastic waves may the lattice periodicity de


be

reduced
is
if

For example, glassy materials with nearly random ar


in

“reased.
a

ions, the mean free path may


is be
of

the order
to

rangement reduced
angstrom units [18].
of

for this reason that


It

magnitude
ºf

thermal conductivity characteristically low [19].


the

of

glass
in of is

periodicity UO, lattice may


be

way which the the reduced


in

One
-

the

Similarly, the
by

oxygen ions
of

addition solid solution.


is

by

conductivity can
be

reduced substitutions the cation


in

thermal
Experimental observations
of

in

lattice. these effects are discussed


Sect. 5.3.1(b).
178 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

The combined effects of oxygen content and cation substitutions on


the thermal conductivity of urania ceramics are important in selecting
a suitable conductivity value for nuclear reactor engineering pur
poses. Fission products form both substitutional and interstitial scat
tering centers. A discussion of the experimentally observed effects of
irradiation on the thermal conductivity of UO, and recommended
values for nuclear reactor applications are presented in Chap. 9.
Most polycrystalline ceramics have a certain amount of porosity.
Experimental and theoretical investigations indicate that in the
porosity ranges of densely sintered ceramics, the true conductivity
(Kr) is related to the measured conductivity (KM) to a good approxi
mation by the simplified Loeb equation,

KT = #. Eq. (5.3)

where P is the volume fraction of pores [20]. The effect of porosity


on the thermal conductivity of sintered UO, is discussed in Sect.
5.3.1(b).

(b) Measured Conductivity Values

(1) Stoich.IoMETRIC UO. Some of the earlier thermal conduc


tivity data obtained on UO, both in powder and ceramic form, are
quoted by Katz and Rabinowitch [1]. Summaries of this earlier work,
together with more recent data, are presented in reviews by Ross and
Tennery [21, 22].
Kingery, et al., obtained thermal conductivity data over the tem
perature range 200° to 1,000°C on a polycrystalline ceramic sample of
stoichiometric UO, fabricated
oo from a suspension and fired at
1,980° C in vacuum to a density of 8.00 g/cc (73.3 percent of theoreti
cal) [23]. The measurements were made in vacuo. Subsequent meas
urements by Kingery were made on a sample of UO2.00 (sintered in
hydrogen) which had a bulk density of 10.08 g/cc, corresponding to
a porosity of 8.0 percent [18].
All these measurements were carried out by the comparative method
described by Francland Kingery [24]. In this method, heat flow
through the UO, was maintained identical to heat flow through an
Al2O, or ZrO2 standard and the temperature gradients compared.
Precision of measurements was estimated to be approximately +2
percent. Comparison with other measurements indicated an estimated
absolute accuracy of approximately +4 percent.
Experimental determinations of thermal conductivity were made
on porous UO, and calculations were made (Eq. 5.3) for nonpor
ous UO, (see Table 5.3). As illustrated in Fig. 5.3, the measure
ments of Kingery, et al., for samples of similar composition made at
different times and prepared from different starting materials are in
good agreement.
PHYSICAL PROPERTIES of URANIUM DIoxIDE 179

Other thermal conductivity measurements for stoichiometric UO2


as a function of temperature have been made by Hedge and Field
house, by Scott, and by Deem * [25, 26]. In the work of Hedge and
Fieldhouse, UO, disks were placed in a helium atmosphere, and a
radial heat flow method was used to determine the conductivity. The
U0, disks, 3 inches in diameter, were fabricated by cold pressing
MCW oxide at 20 tsi and sintering steam-hydrogen cycle (see

in
a
density 8.17 g/cc (74.5 percent
of

of
theoretical).
to

Chap.

It
be 7)

should noted, however, that these conductivity measurements were


cracked samples.
on

made
conductivity measurements

on
one specimen

he of
Scott made thermal
Stoichiometric UO, over the temperature range 800° 1,150°C;

to
oo

used steady-state, radial flow method involving direct electrical


a

the specimen.
of

heating Measurements were conducted with the


flowing hydrogen. The specimen
an

placed atmosphere
of
in

specimen
10 by

wasformed hydrostatically pressing nonstoichiometric powder


a

and sintering
tsi
is)

at at
(UO, nitrogen atmosphere
at

in
it

a
by

1,400°C for hours, followed


reduction hydrogen for hours
in
2

2
1200°C. The O/U ratio specimen was 2.000+0.005,
of

the sintered
was 10.5 g/cc (96 percent
of

and the density theoretical). The


the conductivity measurements was estimated
of

10
be
to
accuracy

+
percent.

the Hedge and Fieldhouse and Scott measurements


of

The results
(corrected for nonporosity) are also listed Table 5.3 and are plotted
in

Kingery, Fig. 5.3. Within the


of

(together with the data al.)


et

in

experimental errors, Kingery and Scott data are good


in

claimed the
and Fieldhouse (ap
by

agreement. The low results obtained Hedge


at
30

Kingery
be

proximately percent lower than the values) may


the fact that the samples had some cracks.
to

tributed
empirical expression for thermal conductivity
an

Scott presented

-###, watt/cm "C Eq. (5.4)



*

theoretical density for UO, and


of

the fraction the


is

T
is

where
f

degrees centigrade [26]. According Scott, this


to

temperature
in

equation satisfactorily represents the thermal conductivity unir


of
of

the experimental measure


of

radiated UO2 within the limits error


ments on ceramic UO2.

*H. W. Deem, personal communication.


Recently. Reiswig [R.
m.

Reiswig, “Thermal Conductivity 2100° C,”


D.

to

A.
of

UO2
J.
*

1,

Ceram.Soc. No. pp. 48–49 (1961)] reported the equation


,
,

K=–watt/cm “C,

17.3 +0.016
T

“K, represent his measurements the thermal conductivity theoretically


to

of

of

*here
in
is
T

*nse UO, from 833 2,112° radial heat flow method was used on disks
C.
of to

of

UO2
A
85

percent theoretical density.


to

*intered
57.4789 O-61—13
180 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
O.14. I i I I I I I

O.13 – -
© Tho.souo.1002+x
O.12H.
- (KINGERY)
U02.0 D Thorse Uo.26402
-
kingERY
O.l I H A Thorse Uo.264024x -
$’ V Tho.69Uo.5102+x
t
->
O.IO – -
*
# 0.09 H.
U
U02.0 (KINGERY, et al.) -
>
* 0.08 H —
:
> O.O.7H. -
5
# 0.06 |992.09
z
-
8
-
D.
H.
...

o.os
Thoz (KINGERY,

et
al.)
-
* §s

- 0.04
H

&\-uo.,
so,

O.os -
O.O2 H.

O.O. H.
--4--4-------------4----9––––––––––– U02.0
----
(HEDGEAND FIELDHouse)

\
O.OO
l

l
2OO 4OO 600 Boo IOOO 200 14Oo |600 isoo
o

|
TEMPERATURE, *c

FIGURE 5.3. Thermal Conductivity


of

Stoichiometric and Nonstoichiometric


Uranium Dioxide of Theoretical Density as Function of Temperature; Data
a

for Thoz and UOg—Th9, are also shown.

Porosity AND MICROSTRUCTURE. Although the


of

(2) EFFECTs
Kt Kingery, al., appear similar for two samples
be
of

of

to

values
et

having different porosities (Fig. 5.3), extrapolation


an

temperature
is in
of

density sample necessary

to
the values obtained for the lower
obtain comparison [18, 23].
a

Deem “observed that fabrication techniques appear affect the


to

thermal conductivity An absolute steady-heat-flow type


of
of

UO2.
thermal conductivity apparatus was used; measurements were made
10−" mm Hg) cylinders
on

diameter and
(5
in

14

vacuo inch
in
×

inches long. Values Kr for three UO, samples which were extruded,
of

isostatically pressed, and sintered porosities and 4.5 per


of

8,
to

22.5,
cent were similar throughout the temperature range 50°
to

800°
C

(Fig. 5.4). fourth sample, however, gave values


of

AT which were
A

percent higher the temperature range 100° This


C.
to
10

some 600°
in

sample was isostatically pressed without extrusion and sintered


to
a

porosity 6.2 percent, and had larger grain size than the other
of

three specimens.

H. W. Deem, personal communication.


*
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 181

TABLE 5.3—THERMAL CONDUCTIVITY DATA FOR STOICHIOMETRIC


UO, CERAMIC SPECIMENS

Thermal con
ductivity
Density Temperature
Investigator Composition (g/cc) (° C)
Calculated for
nonporous body
(watt/cm-* C)

-
Kingery [23]--------------- UO.00-------- 8. 00 200 0.0815
400 0.0590
600 0. 0452
800 0. 0376
1, 000 0.0351
Kingery [18]--------------- UO2.00-------- 10. 08 102 0. 105
149 0. 0908
Kingery [18]---------------| UO.00*------- 10. 55 58 0. 115
70 0. 114
82 0. 114
Hedge and Fieldhouse [25]--| UO2.00-------- 8. 17 202 0.0524
214 0. 0510
391 0. 0428
638 0.0348
867 0.0264
1, 104 0.0234
1, 319 0.0209
1, 404 0.0199
1,563 0.0190
1, 668 0.0193
Scott (26]----------------- UO2.00--------| 10.5 800 0. 0340
900 0. 0310
1, 000 0.0275
1, 100 0.0260
1, 150 0. 0255

"Reduced from nonstoichiometric UO2.00 (seeTable 5.4).

The values of Kr obtained by Kingery, et al., are approximately


24 percent higher at 100° C than those obtained by Deem on the
sample of 6.2 percent porosity; this difference, however, decreases
with increasing temperature and becomes very small near 800° C.
Loeb calculated the theoretical relationship between the thermal
conductivity of a porous material and that of the same material of
theoretical density [20]. The Loeb theory considers the amount of
porosity, the heat radiated across the pores, whether the pores are
cylindrical, laminar, or spherical, and the orientation of the aniso
metric pores to the direction of heat flow. Francland Kingery have
determined experimentally the applicability of Loeb's equations to
alumina, graphite, and nickel in the temperature range 0° to 1,000°
C (27]. These authors also reviewed previously published theories,
182 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

- i I
SPECIMEN DATA
DENSITY,
SPECIMEN DENSITY PER CENT FABRICATION
O.I.O H. NUMBER g cm-3 OF THEORETICAL METHOD
65 9.58 87.4 (I)
68 |O.O7 91.9 (.)
7O |O.45 95.3 (t )
|OOO |O.27 93.7 (2)
O.O8 H -

(1)
EXTRUDED AND ISOSTATICALLY PREssed
AT 40,000 psi
(2) ISOSTATICALLY PRESSED AT 40,000 psi

SPECIMEN IOOO

ALL THERMAL CONDUCTIVITIES SPECIMENS 65, 68, AN


AND 70– -

,
0.02 ARE ADJUSTED To Too PER CENT
H

THEORETICAL DENSITY
O

!
|
I

l
l

l
OO 2OO 3OO 400
O

500 6OO 7OO


|

TEMPERATURE, *c

FIGURE 5.4. Thermal Conductivity of Unirradiated UOs. (H. W. Deem, Battelle


Memorial Institute, personal communication.)

particularly those Russell and Eucken, but concluded that the Loeb
of

equations provided the best


fit

their results [28, 29].


to

Ross investigated the applicability the Loeb theory polycrys


of

to

talline UO, [30]. Thermal conductivity measurements were made


UO, samples having
on

60°C wide range porosities. As


of

near

a
a

the preparation methods used, the grain size and pore geom
of

result
comparative, longitudinal heat flow apparatus
A

etries also varied.


The absolute accuracy the thermal conductivity meas
of

was used.
was not considered high; however, the reproducibility
of

urements
values obtained from single specimens was reported +2 percent.
be
to

Thus, useful relative values thermal conductivity were obtained.


of

Fig. 5.5) showed that


of

Results these measurements (shown


in

A.
r

decreased with decreasing density, disagreement with the simplified


A in

of

equation independent
be

Loeb which states that should the


T

porosity.
These results suggested either: (1) that other
of of

amount
sources thermal resistance were present the samples examined
in

of or
as

and amounts which decreased the density increased; (2) that


in

Eq. (5.3) did not adequately account for the effects porosity
conductivity density samples.
of

the thermal the low


in

significant thermal resistance other than porosity was


of

source
A

impurities, since the chemical analyses and


be
to

to

believed not due


PHYSICAL PROPERTIES OF URANIUM DIOXIDE 183
O OB I I I I I I I-T— I I I

SYMBOL SAMPLE TYPE


D D
C.O7 H
[] LONGITUDINAL D
-
D
O TRANSVERSE []

O O6 H

O O 5 H.

O.O.4 | | | | l | I l | —l l l
8.2 8.4 8.6 8.8 9.0 9.2 9.4 96 9.8 IOO 10.2 IO.4 |0.6 |O.8
BULK DENSITY, g/cm3

FIGURE 5.5. Variation of Thermal Conductivity (Corrected to Zero Porosity)


with Bulk Density; Longitudinal, Direction of Heat Flow Parallel to Original
Pressing Direction; Transverse, Direction of Heat Flow Perpendicular to
Original Pressing Direction [30].

metallographic and X-ray examinations indicated that the samples


consisted only of single phase polycrystalline UO2, very near to stoi
chiometric composition. Ross attributed the low results for the low
density samples to irregularly shaped or anisometric pores at the
grain boundaries, such pores being more effective as obstacles to heat
flow than spherical pores having the same volume and distribution
[30]. The more dense samples, shown to contain mainly spherical
pores, gave results which appeared to be those described by the Loeb
theory. Although the grains in the low-density samples were consid
erably smaller than those in the dense samples, the effect of grain size
was thought to be small compared to the effect of pore shape [30].
(3) EFFECT of NoNSTOICHIOMETRY. In nonstoichiometric UO2, the
extra oxygen may be accommodated by solid solution formation or by
formation of the second U.O., phase (see Chap. 6). Measurements of
the effect on thermal conductivity of the latter were made by Kingery
and by Ross; some measurements were also reported by Nichols [18,
30, 31]. In the work of Kingery, the conductivity of the following
samples was measured by using the comparative method: (1) A sample
of stoichiometric UO, that had been heated in air at a temperature of
approximately 150° to 175° C to an unknown composition, U.O., ;
ºº
184 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

*º --
*.*.*.*.*.*.*.*.*.*
-
* -

;3
- :
; :

A. Sintered in Steam (1,400° C, 2 B. Same Sample Reduced in Hydro


Hours) to Nonstoichiometric UO,...is; gen (1,200°C, 12 Hours) to an O/U
X750, Reduction Factor, 14. Ratio of 2.009; X750, Reduction
Factor $4.

*
SQ

C. Hydrogen Sintered at 1,750° C to D. Same Sample Reduced in Hydro


96 Percent Theoretical Density and gen (1,200°C, 12 Hours) to an O/U
Steam Oxidized at 1,450° C to Non- Ratio of 2.006; X750, Reduction
stoichiometric UOz.oo; X750, Reduc- Factor, $4.
tion Factor, 14.

FIGURE 5.6. Microstructure of Uranium Dioxide Ceramic Specimens Used for


Thermal Conductivity Measurements.

(2) a nonstoichiometric sample, prepared by steam sintering, that had


an oxygen content corresponding to a composition of UO, 1a; (3) an
by

other nonstoichiometric sample (UO, steam oxidizing


on)

prepared
dense stoichiometric UO, ceramic; and (4) this same nonstoichio
a

metric sample (UO2.00) after reduction oxide


to

stoichiometric
in

hydrogen.” presence U.O., samples


of

The the nonstoichiometric


in

compositions UO, Fig.


of

illustrated
in
on

and UO2.1% 5.6 (see


is

Chap. and Appendix


of

for other microstructures nonstoichiomet


B
6

ric UO2.x, both single and two phase). Results these measurements,
of

shown Table 5.4, and plotted Fig. 5.3, show significant lowering
in

in

conductivity with increased oxygen content. The composi


of

thermal
substantially lower conductivity (about percent
30

of

tion UO2.19 has


a

compositions UO, and UO,


on

samples
of

The nonstoichiometric and the reduced sam


is
*

ple were prepared Bettis Atomic Power Laboratory; personal communication, L. Berrin.
at
Nº PHYSICAL PROPERTIES OF URANIUM DIOXIDE 185
*
º TABLE 5.4—THERMAL CONDUCTIVITY DATA FOR NONSTOICHIO
,
METRIC UO, CERAMIC SPECIMENS*

Thermal conductivity
Composition Density Temperature
(g/cc) (° C.) Measured (cal/ Calculated for non
º Sec-cm-*C) porous body (wattſ
cm-° C)

. "*------------------ 10. 08 79 0. 0211 **0. 096

*- "...----------------- 9. 83 33 0. 0071 ***0. 034


45 0.0066 ***0. 031
53 0 0073 ***0. 035
68 0. 0070 ***0. 033
81 0. 0072 ***0. 034
85 0. 0071 ***0. 034
"ha------------------ 10. 55 38 0. 0164 f0. 072
*
50 0.0167 f0.074
57 0.0161 f0. 071
66 0. 0166 #0 073
76 0 0 165 f0. 073
91 0. 0157 f0. 069

"Kingery
[18].
for

10.96 theoretical density unknown composition.


of

"Assumed
for theoretical density UO,
of

"Calculated11.22
i.
for

11.08 theoretical density


of

"Calculated UO1.0.

value), corresponding
the

mean free path


of
to

stoichiometric about
a

units. Compositions
of

"angstrom intermediate oxygen content have


UO, but
of

thermal conductivities lower than that stoichiometric


that for UO,
not

19.
as
so

low
on
at

composition
of

Measurements made uranium oxides


C

60°
T0, and UO,
by

have been reported Nichols [31]. Again, the


is

of

conductivity the nonstoichiometric samples was found


to

thermal
percent that for the stoichiometric compositions.
38

of

belower, about
the The

methods and the resulting microstructures


as

as

preparation well
of

method measurement were not described.


Ross, thermal conductivity measurements were made
50° the

of

work
In

on

two sintered uranium dioxide samples which were


C

lear
stoichiometric UO,
to

composition and which were


in

initially close
then

several nonstoichiometric compositions


to

progressively oxidized
"]
0.1

the original stoichiometric samples contained


of

(One
Samples composition UO,
of

weight percent TiO2.) were measured


as

which there was some slight indication the presence UO,


of

of

as
in

the major constituent U.O. The comparative longitudinal


as

well
heal flow method was used.
The

Kingery, Nichols, and Ross are


of

of

results the measurements


Table graphically Fig. 5.7 [18,
in

5.5 and are shown


in

summarized
186 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

30, 31]. Note (Table 5.5) that the absolute values reported by Nichols
and Ross are considerably lower than those of Kingery. The thermal
conductivity values for the nonstoichiometric samples (corrected to
zero porosity) are shown in Fig. 5.7 as percentages of the corre
sponding values for the stoichiometric samples.

TABLE 5.5–EFFECT OF COMPOSITION ON THE THERMAL CONDUC


TIVITY OF URANIUM OXIIDES

Theoretical Thermal conduc- | Relative


Composition (O/U ratio) Density density 2 tivity corrected | conductivity | Investigator
(g/cc) (g/cc) to zero porosity (percent)
(wattſcnn-*C)

10.55 10.96 30. 115 100 Kingery [18]


10.0 10.96 0.092 100 Nichols [31]
10.30 10.97 0.071 100 Ross (30)
10.25 10.97 0.0535 100 Ross (30)
10.30 11.02 0.064 90 Ross [30]
10.30 11.03 0.061 86 Ross (30)
10.25 11.03 0.043 80 Ross (30)
10.30 11.06 0.054 76 Ross (30)
10.30 11.07 0.0475 67 Ross (30)
10,25 11.08 0.031 58 Ross (30)
10.55 11.08 40.071 62 Kingery [18]
10.30 11.09 0.0375 53 Ross (30)
10.30 11.11 0.033 46.5 Ross (30)
10,25 11.15 0.0225 42 Ross (30)
10.8 11,18 0.035 38 Nichols [31]
9.83 11.22 § 0.035 30 Kingery [18]
10.6 11.25 60.018 25 Ross (30)
2.66(U3O8plus trace amount 8.05 8, 40 70.018 25 Ross (30)
of UO2)

1Measurements made at 60° C except where noted.


* Calculated by linear extrapolation between 10.96for UO2 and 11.30for U4O9.
3Value at 58° C.
* Value at 57°C.
* Value at 53°C.
* Average of 3 specimens.
* These samples contained 0.1 weight percent TiO2.

It is seen that the thermal conductivity decreases with increasing


oxygen content in the composition range UO2.0 to UO,
21.

This de
probably due
to

crease the formation and increase concentration


in
is

the second U.O., precipitate phase which has


of

lower thermal con


a

ductivity than UO. Thus, the resulting two-phase structures ex


thermal conductivity
of
as

hibited decreases the amount the U.O.,


in

phase increased. From the UO,-U,O, phase diagram (Fig. 6.11), the
single-phase intermediate composition solid solution (O., present
is

only higher temperatures.


at

The thermal conductivity


of

nonstoichiometric uranium dioxide


with the oxygen solid solution has not been measured. Since mas
in

sive samples suitable for thermal conductivity measurements may


be
by

difficult the necessary quenching technique (see Ref. 5),


to

fabricate
conductivity measurement high temperatures, where the excess
at
I-
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 187
iod T-I I I I I i I N-N i I I

90 H. O —

8 O H. A LEGEND: -
O KINGERY
D NICHOLS
O ROSS
H. 70 H A ROSS MEAN VALUE -
:
O
L ROSS MEAN VALUE
(MAJOR CONSTITUENT Usos)
:
a.
A ROSS SAMPLES CONTAINED

× 60 H
O O.I wo TiO2
-
->
F

~
2
3 50 H
Q
-

<

-
# A
u
# 40 H
*.
D
>
F
<
º
# 30 H O -
AN
2O H. -

io H -

O
0.0 002
1–1–1–1–1–
0.04 OO6 0.08 O.IO O.I2
|→l—l
O.[4 O 16 O.18 O.20
| | |
64 66 68 7O
CHANGE IN O/U RATIO
FIGURE 5.7. Thermal Conductivities of Nonstoichiometric Uranium Oxides
Relative to Stoichiometric UO2.

oxygen is dissolved, would indicate the effect of solute oxygen atoms


on the periodicity of the UO, lattice.
(4) EFFECT of CATION (EQUIvaLENT) SUBSTITUTION. A lower limit
for the thermal conductivity of UO, due to impurity scattering has
been calculated by Kingery [18]. For example, the specific heat (Cp)
at 300° K of UO, is 0.63 cal/* C-cc (see Sect. 5.3.2); the wave
velocity is approximately 4.0 × 10° cm/sec. If we assume that ran
domness leads to a mean free path the order of lattice dimensions, say
4 angstrom units, the expected thermal conductivity is approximately
0.0034 cal/sec-cm-º C (0.014 watt/cm-2 C).
188 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Kingery has reported measurements on Tho.-UO, solid solu

all

an
tion [18]. These samples,

of
but one which had undetermined
oxygen content, consisted mole percent UO, and had

31
of
10, 26, and
Experimental

of
porosities 9.4, 5.0, and 18.0 percent, respectively.
results, Eq. (5.3), are shown density by

to
corrected theoretical
Fig. 5.3. The value found for the ThocoLos O., composition
in

is
approximately the lower limit the expected thermal conductivity

to
impurity scattering.
to

due

of

In
(5) EFFECT ALTERVALENT CATION SUBSTITUTIONs. solids con
taining free charge carriers the thermal conductivity will include

a
contribution (Ken) due the charge carriers addition the phonon

in
to

to
metals which have high carrier concen
In

of
contribution. the case
trations, the phonon contribution decreases 1/T above the Debye

as
temperature and small comparison Kei which
in proportional

in
is

is
For semiconductors, where the carrier concentrations are lower,
T.
to

(phonon) decreases again 1/T, but Kei does not become signifi
as
K

cant until very high temperatures are attained. Although the carrier
concentration relatively small, UO, semiconductor (see Sect.
is

is
a
5.5.2) and doping with altervalent ions should increase the carrier
concentration and, hence, the electrical conductivity. This doping
by

introducing
an

accomplished
be

can ion (as the oxide) into the


lattice whose size would permit replace U" and whose valence
to
it
3.

ion the lattice would gain


be

In
or

3+
of

could either the case


5’

one-half positive hole per substituted ion, whereas ion would


5’
a
a

These charge
in of
in

to

result the release free electron the lattice.


a

carriers could result sizable electronic component to thermal


a

conduction.
by
improve the thermal conductivity UO,
In

attempt
an

of

small
to

oxide additions, Shapiro and Powers studied the effect


of

Y.O.s and
Nb.O., additives [32]. Thermal conductivity measurements were
made using the comparative apparatus described by
of

version
a

Francland Kingery [24]. The


of

comparative method gave ratio


a

the thermal conductivity the UO, plus additive the UO, stand
of

to
by

given temperature; this ratio was multiplied the Kingery.


at

ard
a

al., value the thermal conductivity UO,


of

of
et

to

obtain the absolute


value.
The results for UO, with the Y.O., additions showed increasing
conductivity for increasing amounts Y.O. The sample
of

thermal
mole percent Y.O. showed percent increase over UO.
25

30

with
to
4

a
C,

between 650° and 850° whereas the Nb, O, and CeO2 additions
lowered the measured thermal conductivity. these re
In

of to

contrast
sults, however, some cursory measurements
of

the effect one and


two mole percent Y.O., additions the thermal conductivity UO.
on

of

progressive decrease with amount Later,


of

showed additive."
a

Hartwig, personal communication.


F.
J.
*
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 189

Powers reported that subsequent thermal conductivity measurements


failed to confirm the original results and that the criteria for balance
in the comparative thermal conductivity apparatus were found to be
in error [33]. Although 3.85 mole percent additions of Y.O., increased
electronic carrier content somewhat, the thermal conductivity was
reduced. Moreover, an in-pile experiment (see Chap. 9), in which the
comparative heat ratings of UO, containing 3.85 mole percent Y2O,
and pure UO, were compared, showed that this addition did not im
prove the mean thermal conductivity of UO, over the temperature
range 350° to 2,800° C [34].
From these results it appears unlikely, at least at this time, that
anysubstantial electronic contribution to the thermal conductivity of
UO, can be expected from altervalent cation additions.

Summary
(c)

made concerning the application


to be

of
Several conclusions can these
thermal conductivity data UO, fuel.
Polycrystalline sintered UO, fuel should high density
or

be

of
1.

porosity) pore shape, shown by Ross,


of

(low where the effects as


are minimized.
as

possible
be

Uranium dioxide fuel should stoichiometric


to
as

close
2.

composition because has been shown


that the experimentally
in

it

second U.O,
phase reduces the thermal conductivity drastically.
Solid solution Tho, UO, compositions appear have thermal
to
3.

to by

conductivities corresponding that predicted theory.


to

Altervalent cation additions (specifically, Y") UO, were shown


4.

experimentally both out-of-pile and in-pile


to

decrease the thermal


conductivity.

5.3.2 Specific Heat


H. D. Sheets

Heat capacity (Cp) measurements


to of

uranium dioxide have been


over the temperature range The specific
15

made 300° [35].


K,

UO, has peak between


of

heat curve with maximum


15

and 50°
a

a
K.

This anomaly has been related the magnetic be


at

6) to

value 28.6°
Chap.
of

havior the material (see Sect. 5.5.3 and [36].


the thermodynamic properties UO, (see Chap.
at of

of

Detailed studies
Mines [37, 38]. The specific
of

have been made the U.S. Bureau


6)

heat equation calculated from these data,


(', (cal/mole-9 C) T-2.272×10° T-4,'
=

18.45+2.431
×

10−8

Popov. al., report that the heat capacity given by the


et

to
of

UO2 from 160° 603°


is
C
7

expression,

Cp

=

15.29 1.716 10-2 1.41 10-5 T2 [39].


+

×
T
190 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

given in Katz and Rabinowitch [1], was used by Hausner and Mills
to calculate the specific heat values given in Table 5.6 [40]. These
calculated values are greater than the approximate maximum for the
molar heat capacity, 6m cal/deg, where n is the total number of atoms
in the molecule. Deviations from the law of Dulong and Petit are
not unreasonable, however, at high temperatures.

TABLE 5.6—SPECIFIC HEAT OF UO, [40]

Temperature(°C) Specific heat C, (callg-° C)


100 0.063
200 0.067
500 0.074
1, 000 0.078
1, 500 0.082

5.3.3 Debye Temperature


J. Belle and V. J. Tennery
As discussed in Sect. 5.3.1, at temperatures higher than the Debye
characteristic temperature, 9, the thermal conductivity of dielectric
crystals is governed only by the temperature dependence of /, the
phonon mean free path. If
A (phonon) = 1%('el and if it is assumed
that heat capacity and phonon velocity remain essentially constant
with temperature when T-6,
A (phonon) is proportional to 1/7.
Tennery used this approximation to estimate the magnitude of the
Debye temperature 6 for UO, [22]. He obtained a value of about
870° K from a plot of the Kingery, et al., conductivity data versus
reciprocal temperature, since the plot becomes reasonably linear above
850° to 870° K [23].

is,
This method of determining the Debye characteristic temperature
however, not the classical one and certainly open question. The
to
is

Debye temperature can determined from heat capacity data.


be

Either approximation
7”

the used near absolute zero, which the


in
is

Debye specific heat equation

C,-3R º/T rºd.r 30/T


Eq.
[2% Jo ez-i Taºſ T-I (5.4)

reduces to

Eq.
cº-º-R(ſ) Teo (5.5)
or
or

the specific heat curve the compound


to

the Debye
of

matched
is
by

specific heat curve choosing the proper value for Eq.


in
6

Jones, al., UO,


be

(5.4). for from their


to

determined
et

160°
by 9

specific heat data below 20° using the approximation [35].


7”
K
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 191

Due to the high specific heat (C.), it is not possible to match the
Debye function to the data on the rising part of the specific heat curve
of Jones, et al., to obtain another estimate of 6 [35]. According to the
Debye theory, however, at T=0, C, has attained 94 percent of

its
max
imum value. Assuming that the maximum value for the molar specific
heat (Cr) 9/P (17 cal/deg),

be
of

of
UO2 can seen from the data

it
is C.), that the Debye temperature

('p
Jones, al., (assuming equal

to
et

is
greater than 300° K.” Note also that the heat capacity has only
is

K.
its
percent room temperature value
65

of

at
attained about 160°
Debye temperature can

be
of

Another estimate the made from the


Lindemann approximation the melting

at
which assumed that
in

is
of it

or
point the amplitudes
of

vibration the solid

in
the atoms ions
are approximately equal their mean distance apart. The maxi
is to

by
va,

mum Debye frequency given

T.
v,

-2.8×10”
MV2/3

weight, the molecular volume, and Tm


J/

where molecular

is
is

is
V

win

the melting point For UO2,


in

[41]. 10” sec and


K

3.22 ×

-
*

hym

6–4. =
154°K,
by

surprisingly close the value found Jones, al., from


et
to

which
is

low temperature specific heat data [35].


may questionable whether the concept single Debye tem
be

of
It

very meaningful for UO, that


vo.,
is,

perature the maximum lattice


is

by

vibrational frequency, cannot such simple approxi


be

determined
a

The frequency mode distribution for UO, must


be

known
in

mation.
specified temperature. This may ob
be
to

at

order determine
9

a
by

tained using the Born-von Karman method for the UO, structure,
although the mathematical complexity the calculation may pro
be
of

hibitive [42].
Debye's theory predicts specific heats which seem
fit
to

the data best


for diatomic solids which the ratio of the atomic masses of the
in

two species present does not vary more than 2:1 (such
as

the alkali
halides, CaF2 and Al2O3).
of

Blackman has made calculation the


a
of

specific heat linear diatomic lattice and has shown that varies
6
a

of

considerably with temperature [43]. As the ratio the atomic


masses increases, the deviation with temperature becomes more
6 of
6

extreme; for high mass ratio, increases rapidly with temperature


a

fairly constant high temperatures. Blackman also cal


at

but becomes

Scott reported [26].


to

be less than 600°


K
6
*
192 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

culated the variation of 6 as a function of temperature for a simple


cubic monatomic lattice; in this case, considerable variation of 6 with
temperature was also predicted.
For UO2, where the mass ratio is approximately 15, a considerable
variation of 6 with temperature could be expected. If this is the
case, the value of 6 determined by Jones, et al., from the Tº
extrapolation below 20° K may be valid only in this temperature
range [35].
The distribution of vibrational modes has not been calculated for
the fluorite lattice, nor have the expressions been derived which
give 6 (T) as a function of the mass ratios. If
the electronic specific
heat is negligible at elevated temperatures, the derivation of these
functions should permit the determination of 6 (T) for UO2.

5.3.4. Thermal Expansion


H. D. Sheets

Several investigators have reported results of thermal expansion


measurements on UO, over the temperature range 20° to 1,200° C.
Murray and Thackray were the first to report the coefficient of ther
mal expansion for both sintered UO, and sintered nonstoichiometric
UO2 is [44]. (It should be noted that since analyses were not re
ported, the latter oxide may have been reduced at higher tempera
tures.) In these dilatometric measurements made in a purified argon
atmosphere, two heating rates (3%" and 10°C per minute) were used.
Results (Fig. 5.8) show that the expansion coefficient increases up to
400° C approximately and remains constant from 400° to 900° C.
In this temperature range, the coefficient of thermal expansion is
10 × 10−"/° C. Schaner * determined the linear thermal expansion of
high density UO, (about 98 percent theoretical density) to be 10.1 ×
10"/* C from room temperature to 1,100° C. The measurement
was made in vacuum by the autographic optical-lever method. The
sample was heated and cooled at 4° C/min and the data from the
cooling curve were used to compute the expansion coefficient. The
percentage linear expansion between room temperature and various
temperatures up to 900° C together with the expansion coefficients
over the range 400° to 900° C are given in Table 5.7. Note that bulk
density has little effect on the measured thermal expansion.
Other results on sintered UO, reported by Burdick and Parker, by
Bell and Makin, and by Lambertson and Handwerk are shown in
Fig. 5.9; these data agree up to 800° C [45, 46, 47]. (The data of

* B. E. Schaner, unpublished work.


PHYSICAL PROPERTIES OF URANIUM DIOXIDE 193

O8 I i i i i i -T I I


0.7 H. -

O6 H -
+
2.
u
C
§ 0.5 H -
a.
2
Q
ºn
z 0.4 H -
§
x
u
-< -
u 0.3 H
2
L

loº
O.2 H. A-UO2 (DENSITY 7.2 g/cm3) C/MIN

72
o-UOz (DENSITY g/cm3).3%" c/MIN

8.5
m-UO2.s(DENSiTY g/cm3, 3}•c/MIN
A-UO2 isſDENSITY 101 g/cm3) o'c/MIN
O.

&
H

A.
:

cºlt
O

l
IOO 2OO 3OO 4OO 500 6OO 7OO 8OO 900
TEMPERATURE.2c

FIGURE 5.8. Thermal Expansion of UO2.

TABLE 5.7–THERMAL EXPANSION DATA FOR SINTERED UO, [44]

Percentage linear expansion Expansion


coefficient
Specimen Heating rate (400°–900°
20°–200° 20°–400° 20°–600° 20°-900° C)
C

C
C

UO, 7.2g/cm3.-------- ------ - 10°/min. 0.07 0.22 0.42 0.72 10x10-8


..
.
.
.
..
.

TO:, 7.2g/cm3........ -------- 3.5°/min--------- 0.09 0.25 0.44 0.73 9.5x10-4


||

TO:, 10.71g/cm3.--...-------. 4.0°/min.--------|----------|----------|---------|---------- 10.1×10-8


is, 11,

TO: 8.54g/cm3.------------ 10°/min--------- 0.08 0.25 0.45 0.75 10x10-8


||

UO, 10.1g/cm3. -----------| 3.5°/min.-------- 0.08 0.25 0.47 0.78 10.5×10-8

"B. Schaner, unpublished work.


E.

Murray and Thackray are shown for comparison.) Burdick and Par
ker fitted their data for the temperature range 27°
to

to

1,260°
C

curve of the form


y=aT-4 b/'4-e
where
on

percent thermal expansion length 27°C


at
a y

based
= ==

2.1481 10-7
X

8.4217 10-4
×
b

c=-3.0289 10-2
×

T=temperature (°C).
194 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

H-T-T—
-
20
D BURDICK AND PARKER
O BELL AND MAKIN
e LAMBERTSON AND HANDWERK
MURRAY AND THACKRAY

O IOO 200 300 400 500 600 700 800 900 OOO IOO i2OO 3oo
TEMPERATURE,”C

FIGURE 5.9. Thermal Expansion of UO,.

It should be noted that the experimental thermal expansion data


agree in general with values calculated from X-ray parameters (see
Sect. 5.2.1).
Recent measurements of Halden, et al., have extended thermal ex
pansion data from the region of 900° to 1,000°C to the melting point
of UO, [48]. The linear expansion of dense (93 percent of theo
retical) UO, was found to follow closely the equation
l=l, (1+6.0×10+ T-1-2.0×10-9 T-4-1.7×10-14 Tº)

in which T is in * C. An anomalous expansion was noted in the tem


perature range of 1,000° to 1,500° C. These results are shown in
Fig. 5.10; the data of Bell and Makin and of Murray and Thackray
to 1,000° and 900°C, respectively, are shown for comparison [44, 46].
The volume expansion of UO, up to the melting point, determined
by Halden, et al., is shown in Fig. 5.11 for two types of dense UO.
specimens [48]. These data were calculated from linear expansion
measurements and illustrate the large increase in volume which oc
curs at temperatures above 2,000° C.
Above 2,450° C excessive vaporization and crystal growth on the
surface of the specimens were observed. An attempt was made to ex
plore the region near the melting point by means of a solar furnace,
but only qualitative information was obtained. Some preliminary
results suggested that the volume change incident on melting was
probably less than 3 percent. Note that this compares with an esti
mate of from 1 to 7 percent obtained from metallographic examina
tion of irradiated UO, that had melted (see Chap. 9).
|
* TTT-H PHYSICAL

A - BELL
PROPERTIES

AND MAKIN
OF URANIUM DIOXIDE 195

B - MURRAY AND THACKRAY

O3

O|

oss

FIGURE
H--ll
O

5.10.
500 IOOO |500
TEMPERATURE , "c
2OOO

Linear Thermal Expansion of Dense (10.19 g/cc)


| |
2500
|

UOa
| |

upon
3OOO

Reheating after Initial Heating to 2,000° C [48].

5.3.5 Melting Point

In view of the difficulties involved in measuring and controlling


very high temperatures, the effect of small quantities of impurities on
melting points of oxides, and the rather high vapor pressure of UO, at
elevated temperatures (see Sect. 5.3.6), it is not surprising that widely
different values for the melting point have been reported. Some of
the earlier values of the melting point are those of Ruff and Goecke,
2,176°C, and Friederick and Sittig, 2,500° to 2,600° C [49, 50].
Lambertson and Handwerk, using specimens heated in a purified
helium atmosphere in a tungsten crucible, reported a melting point
of 2,880°= 20°C for UO, [47]. Spectrographic analysis after melt
57.4789 O-61—14
196

-
URANIUM
1.24

1.2O H
H-Hº
O
()


DIOXIDE:

10.19 g/cc3
PROPERTIES

NORMAL GRAIN SIZE


Io.19 g/cc 3 LARGE GRAIN SIZE
(SOLID SYMBOLS INDICATE MEASUREMENT
AND NUCLEAR APPLICATIONS

ON COOLING)

3|> 1.16 –
<!| -
+
+
| z
à 2
1.12H # T
# +
º: d
Lu >
Lil tw.
>
B 1.08 H 9 –
o

_>
Ö
> O-

©
1.O4 H -

I.OO T----|--|--|--|--|--|--|--|--|--|--
O 500 IOOO 1500 2OOO 2500 3OCO
TEMPERATURE, "C

FIGURE 5.11. Volume Expansion on Heating UO, to the Melting Point; Solid
Circles Indicate Measurements on Cooling [48].

ing showed greater than 99.95 weight percent UO2. The lattice
parameter to be 5.4725+0.0001 angstrom units.
was found
Ackermann found the melting point of UO, in a vacuum to be
2,407°-E 19°C [51]. (Later Ackermann, et al., stated that the UO, ap
peared to have melted at this temperature [52].) Spectrographic
analysis of the sample showed less than 0.5 percent silicon and 0.2 per
cent titanium after melting. -

Wisnyi and Pijanowski determined the melting point of UO, in


hydrogen, argon, or helium atmospheres [53]. The samples were
held in a V-shaped tungsten ribbon; the ribbon itself was used as the
heating element. The UO, samples had lattice parameters of 5.472+
0.005 angstrom units before melting; after melting, the lattice param
eters were identical within experimental error. The melting point was
found to be 2,760°=30° C. Differences in melting point in the dif
ferent atmospheres were within the experimental error.
Ehlert and Margrave determined the melting point of UO, in
vacuum [54]. They reported that the UO, specimens melted at
2,530° = 55° C, brightness temperature, and calculated the melting
point to be 2,860°,+45° C from emissivity data. By comparison with
the surface of a tungsten wire which supported the UO, the spectral
emissivity at 6,500 angstrom units of the polished UO, surface within
100 degrees (C) of the melting point was found to be 0.416--0.026.
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 197

Using (polished UO, containing powdered UO,


a different method
heated in graphite enclosures) the spectral emissivity of solid UO.
and of powdered UO, at 6,500 angstrom units was found to be
0.40+0.02 and 0.51+0.03, respectively, between 1,800° and 2,100° C.
The spectral emissivity of nonpolished specimens of UO, as a function
of temperature was also determined by Claudson [55]. The data of
Claudson and of Ehlert and Margrave are shown in Table 5.8 [54, 55].

TABLE 5.8—SPECTRAL EMISSIVITY OF UO, [55]


Temperature(°C) Emissivity
727 0. 850
1,047 0. 798
1, 320 0. 628
1, 482 0. 510
1, 522 0. 515
1, 580 0. 417
1,647 0. 402
1,682 0. 484
1, 780 0. 446
1, 947 0.370
1, 800°–2, 100° C 0.40%
*Reference54.

5.3.6 Vapor Properties


C. A. Alexander

While no systematic determination of vapor pressure as a function


of composition has been made in the UO,
phase region, there is ample
evidence that more than one vaporization process occurs. Factors
affecting the volatilization are conditions of extreme temperature, non
stoichiometry, and the existence of relatively low oxygen concentra
tions in contact with UO, (c).
Vaporization of stoichiometric UO, occurs predominantly as con
gruent sublimation to UO, (g), at least at temperatures below 2,000°
K. Ackermann, et al., reported that the vapor pressure (mm Hg)
from 1,600° to 2,000° K is given by [52]

p--º'-402 Eq. (5.6)


log

log

T-25.6sº
The same authors observed positive deviations from the above equa
higher temperatures; apparently, these deviations were real
at

tion
additional vaporization process became signifi
an

and indicated that


higher temperatures. The total volatility UO,
of

cant was ex
at
as

pressed

–7– T2
1”
3.7195X 104 3.5162× 10^
log p(mm Hg)=13.298
+

2.617 109

198 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

From considerations of the available thermodynamic data, Acker


mann, et al., postulated that the additional vapor species were the
dimer (UO2), (g) or possibly UO, (g) and UO (g) [52]. There is
evidence that both UO (g) and UO, (g) can become the principal
oxide vapor species under certain conditions. In an initial mixture of
molten uranium metal and UO, UO (g) was observed and a value of

its
was determined for dissociation energy. was further

It
7.5 ev
determined that UO more volatile than UO2 but less volatile than

is
metallic uranium.”
low partial pressures has been ob
In

the presence oxygen,


of

of

it
subliming

of
served that the amount volatile oxide much more than

is
by
the vapor pressure UO, [56]. The probable
be

of
can accounted for
mechanism for this vaporization

is
UO,(c) +1/2 O,-UO,(c)

UO,(c) →U04(g)

From the thermodynamic data for UO, (c) and for UO, (g), the

by
equilibrium pressure

be
UOs (in atmospheres) would given
of

[38,57]
— 13,1
-1/2

log p==}"+4.73 log po, Eq. (5.8)

Equation (5.8) would indicate that


is at

oxygen pressures

of
the order
10" atmosphere, UO, (g) the principal vapor species
of

to

at
10-
temperatures below 2,000°K. At oxygen pressures much higher than
these, Eq.
be

account for the


to

to

(5.8) would have modified


change the composition the condensed phase. Although Eq.
of
in

(5.8) has not been experimentally verified, unpublished Battelle data


equilibrium pressure
an

10−" atmosphere
of

at

1,520°
K

show over
×
3
an

argon atmosphere contaminated with traces oxygen


of
in

UO2.02
''

and nitrogen. This value magnitude greater than


of

several orders
is

vapor pressure UO, temperature.


of

the the same


at

Thus, the vaporization the UO. phase range may


be

summarized
in

principally congruent sublimation


an

oxygen-free system con


as

in

taining stoichiometric (). slightly oxidizing atmosphere, in


In
U

vaporization UO, (g) and also


an
of

formation
to

to

creased due
is

volatility through UO (g) under condi


of

increase the formation


in

tions which uranium metal and U(), are contact.


in

in

Ackermann, private communication.


R.
J.
*

See Chap. for discussion of oxygen partial pressures equilibrium with non
in
6

a
*

stoichiometric uranium oxides.


PHYSICAL PROPERTIES OF URANIUM DIOXIDE 199

5.4 MECHANICAL PROPERTIES


H. D. Sheets and W. H. Duckworth

5.4.1 Strength

Factors Affecting Strength


(a)

dioxide ceramics Chap. exhibit brittle behavior;

4)
Uranium (see
fracturing, with little plastic

no
or
to
failures under stress are due

do
deformation. Such brittle materials not have unique strengths.

on
or
strength depends

of
The mean fracture stress conditions measure
ment, including specimen size, stress system, and rate loading.

of
No
accepted relationships exist between the mean fracture stress given

of
a
these factors, although some progress has
of

brittle material and any


been made toward predicting the effect

of
specimen size and stress
strength through so-called statistical theories
on

strength.

of
state
Qualitatively, mean fracture stress increases with decreasing specimen
increasing loading rate. The effect

be
stress state appears
of
or

to
size
related to the volume of material under tensile stress. The nature
temperature
of

the effect varies with different brittle materials and


of

with the temperature testing. At certain high temperature,


of

level
a

brittle materials lose strength rapidly, probably coincident with the


appreciable plastic
of

onset flow.
Theory and experiment indicate that the mean fracture stress
of
a
quite sensitive processing changes. Strength
to

brittle material
is

by

given testing conditions apparently determined structural


is

under
by

which, turn, are fixed the processing conditions.


It
in

is
factors
known that fracture strength decreases rapidly with increased porosity
and, essentially nonporous polycrystalline bodies, the strength has
in

founddecrease with increases average crystal size.


to

in

been

some brittle materials, glass, for example, fracture normally


In

originates As consequence, surface treatments and


at

the surface.
a

strength
an

on

testing atmosphere used can have appreciable effect


values.

the above considerations, strength values can


be
to of
In

view used
only compare the resistance
of

different brittle mate


to

safely stress
be
In

making such comparisons, essential that the data


is

rials.
it

obtained under standardized conditions. Such values, for example,


group UO, ceramics prepared
be

of

of
to

can used determine which


to a

differently has greater resistance fracture.


200 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

(b) Strength Data

Lambertson and Handwerk reported modulus of rupture data from


three high-density (10.6 g/cc) UO, specimens, 0.25 inch in diameter
and 2 inches long. One specimen was loaded at a fast rate, and failed
at 10,000 psi [47]. The two remaining specimens were loaded at
the rate of 0.05 inch per minute and failed at 24,800 psi and 27,500
psi, respectively.
Burdick and Parker and Knudsen, Parker, and Burdick measured
moduli of rupture at room temperature and at 1,000° C of UO, speci
mens made from electric-furnace fused UO, and from four varieties
of chemically prepared UO, [45, 58]. Fabrication history and
strength data are given in Table 5.9. The specimens were broken by
quarter-point loading. Stress was applied in 1,000-psi increments
at 20-second intervals.
The difference in the grain size of the fused UO, powder used in
the early investigation did not remain after sintering [45]. The

5p.
average grain size of the sintered specimens made from the 0 to
10p, powders was 48p and 46p, respectively. The specimens
to

and
5

differed the total porosity and the size the pores. The stronger of
in

5p

material, had less total porosity and


Op

specimens, from the


to

smaller pores.
The data suggest that the fracture strength UO, ceramics of

be is
slightly higher room temperature. This could
at

1,000° than
at
C

annealing stresses present the specimens


of

room tem
in
to

at
related
perature plastic flow Similar behavior has
or
to

the material.
in

been reported for silicon carbide, graphite, and several other ceramic
materials. Scott, al., have recently found that dense UOzoo has
et

plasticity but becomes plastic higher


no
or

little
at

at

about 1,000°
C

temperatures (1,600° C); dense nonstoichiometric oxides (UO2.06 and


UO, 1a), deformed plastically
on

at

the other hand, can


be

about
800° [60].
C

the later NBS work,


In

was found that the strength (S) could


it

by

porosity (P) and grain size (G)


be

specimen
to

related the the


proposed equation
S=}:G-de-l'I’
are empirical constants which depend, part,
on
a,
k,

of in

where and
b

the test temperature (see Table 5.10) [58, 61]. The effects porosity
and grain size rupture UO, room temper
on

of

to of

at

the modulus
according the equations given
at

ature and 1,000° (calculated


in

Table 5.10) are shown Fig.


in

5.12.
-
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 201

45 I T I I i I
IO
\
\ -
\ uo, SPECIMENs
\ -
---
AT ROOM TEMP.

\\
40 H
AT IOOO°C

NUMBERS INDICATE GRAIN SIZE


\ IN MICRONS
-
35 H. \
\
\
\
\ -
30 H. \
\
: \
g \\
2
-
25

\
g ...

20
\\
\\
: :

N.
\\

\ N.

*
-
N.
20

N.
3. *

N
\\

N
É

30

N.

SS
N
;

N.

15
N.
-

10
-
H

-
H-
5

1—
l
|
|
O

l
l

iO 15 2O 25 3O 35
O

POROSITY, PER CENT


of

Rupture
of

the Moduli
on

Porosity and Grain Size


of

FIGURE 5.12. Effects


C,

According
to

Temperature the
at

U0, 1,000° Calculated


at

Room and
Equations Table 5.10 [58].
in
TABLE 5.9–BEND STRENGTH AND BULK DENSITY OF DIFFERENT UO, CERAMICS [45, 58]

Sintering conditions Strengthdata


Starting UO, powder
Atmosphere tempº Nº:* sºle.
º * Bulºus
rupture (psi) (psi)

At Room Temperature

Fused, 5put----------------------- Argon---------- 2,000 11, 990 370 10.10

to to
05
9,
2, 1,

Fused, 101----------------------- Argon- -------- 2,000 480 180 9.68


Fused, 10 15u---------------------- Argon---------- 2,000 10, 270 1,040 9. 14

to to
15
8,

Fused, 2014---------------------- Argon---------- 2,000 610 380 8.48

to
%% %%%

0
1, 1,

Fused, 51------------------------ Helium--------- 2,000 12, 920 261 9.89


Steam oxidized”---------------------- Helium--------- 2,000 340 560 8. 34
7, 9,

..Hydrogenated steam oxidized”--------- Helium--------- 2,000 580 637 8.82


4464 9876

--
Helium---------

---
-
--
--
Peroxide precipitated” 2,000 12,400 610 10.04

1,
Ammonia precipitated” --------------- Helium--------- 900–2, 000 12 11, 100 230 9.84
7

11 1 1 1

1,

-H
-
--
--
1, 1, 2,

Steam oxidized 0.75 w/o TiO2- Helium--------- 900 12, 700 180 9.66
Hydrogenated steam oxidized --0.50 w/o
3,

TiO,------------------------------- Helium--------- 1,800 14, 500 180 10. 10


1 1
67

1,
Peroxide precipitated +0.25 w/o TiO2----| Helium--------- 600 10, 400 2,800 10.39
Peroxide precipitated +0.2 w/o alumi
1
8
1,

num stearate----------------------- Helium--------- 2,000 12, 700 730 10.37


Ammonia precipitated +0.2 w/o alumi
1

-
---
--
Helium
--
1,

num Stearate----------------------- 2,000 11 10, 870 164 10.26


C.
At 1,000°

1
17, 980 900 02
Fused, 511------------------------
15, 790 1,050 56

to to
Fused, 10u -----------------------

3,
12, 590 170 17

0 5 10
Fused, 15u ----------------------

8,
280 1,080 .48

to to
--
--
---

---
---
--
---

-
Fused, 15 2011

%%% % %
89

to
14, 900 2,020

0
Fused, 5u ------------------------

7,
Steam oxidized------------------------ 700 897
1, 81
Hydrogenated steam oxidized----------- 10, 800 210
1

21, 700 3,770 05


Peroxide precipitated.------------------

9,
380 1,440 85
Ammonia precipitated.-----------------

.
846 .62

--
--
--
Steam oxidized --0.75 who TiO2- 15, 600

-H
Hydrogenated steam oxidized 0.50 w/o
800 19, 500 1,480 10. 13

1, 1,
Peroxide precipitated +0.25 w/o TiO2---- 600 13, 600 3,060

Peroxide precipitated +0.2 w/o alumi 5:


11
1,

num Stearate------------------------ 2,000 11, 800 820 10. 36

-H
Ammonia precipitated 0.2 w/o alumi
1
6
1,

9,

num stearate----------------------- 2,000 380 770 10. 30

at

a
in
Notes.—Forming: Preformed steel mold and then hydrostatically repressed 45,000psi; no binder used.
Sintering: Tungsten setter plates, inductively heated, graphite susceptorſurnace.

to
X
Finishing: Specimensground and lapped 1% inches.

9%
%4
*Norton Co. electric-arcfused UOz.

of
**Prepared by ORNL; for description materials,see Ref. 59.
§
204 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 5.10—CALCULATED AND MEASURED MODULI OF RUPTURE OF


UO,” [58]

Steam Hydrided | Peroxide Ammonia


Type of UO2 powder used in specimen preparation oxidized Steam precip- precip
oxidized itated itated

Temperature at which the specimenswere sintered


o C. 2,000 2,000 2,000 2,000
Specimen porosity---------------------------- percent-- 23.9 19.5 8.4 10.2
Specimen grain size -------------------------- microns. . 23.2 17.9 19.6 43, 9

Strength at room temperature as measured-------- psi- 7, 340 9,580 12,400 11.100


As calculated.-------------------------------------- psi -- 7,648 9,078 12,753 10, 943
Difference as percent of calculated strength------------- –4 +6 –3 +1

Strength at 1000°C as measured.------------------ psi-- 7,700 10,800 21,700 9,380


As calculated.-------------------------------------- psi -- 7,284 11,635 20,875 9, 560
Difference as percent of calculated strength .. . --------- +6 –7 +4 –2

*Calculated using the equation


S=k G-se-bp
where S is the strength (psi)
G is the averagegrain size (microns)
P is the porosity (expressedas a fraction)
k, a, and b are empirical constants
At room temperature, strength S=23,700G-0.11%-3.17 P
At 1,000°C, strength S=409,000G-0.437e-s
sºP

Compressive strengths of specimens made from the fused UO, also


were reported by Burdick and Parker [45]. Average compressive
strengths of specimens having a length-to-diameter ratio near 2:1
were 140,000 psi with the Op to 5u starting material, and 70,000 and
60,000psi with the 10p to 15u and 15u to 20p material, respectively.
Specimens having a length-to-diameter ratio of 1:1 fractured at some
what higher stresses.
The effect of nonstoichiometry on the room-temperature bend
al.

strength has recently been determined by Scott, et [60]. Measure


sintered UO, similar grain size
on

of

ments were made and UO2.0


is

and porosity percent) using the quarter-point loading technique.


(5

Average values 11,400 and 8,800 psi were obtained for UO,
of

and
is

UO2.0, respectively, for six specimens


of

each with standard deviations


1,315 psi, respectively.
of

+950 and
+

Peters reported modulus rupture data the range from room


of

in

temperature The specimens, bars approximately


to

1,000°
[62].
C

by

midpoint loading
on

0.25 0.52×414 inches, were broken 4-inch


×

span. Room-temperature measurements were made air. Higher


in

20
an

of

temperature measurements were made atmosphere percent


in

hydrogen percent argon. specimens


80

and The were made from


MCW UO, powder which
20

had been dry ball milled for hours.


Specimens were pressed psi
in

in
at

6,400 steel die and sintered


a
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 205

hydrogen using a continuous, molybdenum-resistor furnace. The


average modulus of rupture of the UO, ceramics prepared in this
fashion ranged from about 14,000 to 16,000 psi. The data (sum
marized in Table 5.11) agree with those obtained in the NBS work:
that fracture strength of UO, ceramics is somewhat higher at 1,000°C
than at room temperature [45]. Otherwise, the data do not show any
appreciable effect of variables investigated on strength. As with the
Bureau of Standards' bodies, porosity effects probably were a domi
nant factor in their strength. All specimens had significant porosities.

TABLE 5.11—EFFECT OF PROCESSING WARIABLES ON MODULUS OF


RUPTURE OF UO, CERAMICS [62]

Strength data
Sinter
ing tern- || Sinter
perature ing time Surface treatment Test tem- Number | Average | Standard Bulk
(° C) (hr) perature of speci- modulus deviation density
(° C) mens of rupture (psi) (g/cc)
(psi)

1,750 1 ||Ground with 400-gritAl2O3--| Room----- 6 14,850 1,302 10.4


1.750 1 ||Ground with 400-gritAl2O3--| Room----- 6 13,900 1, 729 10.4
1,750 1 ||Ground with 400-gritAl2O3--| Room----- 6 14,800 1, 186 10.3
1,750 1 | Untreated.------------------ Room----- 6 12,500 1,285 10.3
1,750 1 || Ground, then annealed at Room----. 4 14,400 640 10.4
500°C.
1,750 1 || Ground, then annealed at Room----- 5 16,800 742 10.2
500°C.
1,750 1 | Fine ground-----------------| Room----- 6 15,800 1,337 10.4
1.750 1 | Polished.-----------. - - - - - || Room----- 6 14,700 1,255 10.3
1,750 1 ||Ground with 400-gritAl2O3--| 500-------- 6 12,950 1, 138 10.2
1,750 1 ||Ground with 400-gritAl2O3--| 500-------- 6 16,150 806 10.4
1,750 1 Ground with 400-grit Al2O3--| 1,000-- - - - - 6 14,500 749 10.3
1,750 1 ||Ground with 400-gritAl2O3--| 1,000------ 6 12,800 2,662 10.2
1,750 5 || Ground with 400-gritAl2O3--| Room---- 6 13,450 865 10.4
1.750 5 ||Ground with 400-gritAli03--| 500-------- 6 13,200 2,446 10.5
1.750 5 || Ground with 400-gritAl2O3--| 1,000-----. 6 17,000 3, 159 10.5
1.750 5 Ground with 400-gritAl2O3--| 1,000----- 6 13,650 1, 177 10.5
1,750 5 || Ground with 400-gritAl2O3--| Room---- 6 12,000 1,070 10.5
*1,750 5 ||Ground with 400-grit Al2O3--| Room---- 5 15,400 1,502 10.4
1,750 5 || Ground with 400-gritAl-O2- 500-------. 4 13,300 989 10.5
*1,750 5 Ground with 400-gritAl2O3--| 500------ -- 4 14,450 1,462 10.4
1.750 5 || Ground with 400-gritAl2O3--| 1,000-- - - - 5 13.550 745 10.5
*1.750 5 || Ground with 400-gritAl2O"--| 1,000------ 5 16,750 1,030 10.4
1,750 20 ||Ground with 400-gritAl2O3--| Room----- 6 12,250 1,496 10,5
1,750 20 || Ground with 400-gritAl2O3.- 500-------- 5 12,250 1,935 10.6
1,750 20 ||Ground with 400-gritAl2O3--| 1,000------ 5 14,200 1,721 10.6
1,500 20 Ground with 400-gritAl2O3--| Room----- 6 14,050 714 10.3
1.550 20 || Ground with 400-gritAl2O3--| Room---- 6 15,500 1,029 10.4
1,600 20 ||Ground with 400-gritAl2O3--| Room---- 4 13,200 912 10.5
1.550 20 || Ground with 400-gritAl2O3- 500-------. 6 14,200 820 10.5
1,550 20 || Ground with 400-gritAl2O3--| 1,000-- - - - - 5 15,500 1,390 10.6

"CO, calcined 1 hour at 1,400°C prior to pressing.


Note.—Specimens fabricatedunder the same conditions, but reportedseparately, were prepared at differ.
enttimes.
206 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Bowers, et al., measured the strength of 1/2-inch-diameter, 1/2-inch


long solid cylinders by measuring the compressive load required to
break them (load applied perpendicularly to the axis) [63]. Frac
ture resulted from a tensile stress perpendicular to the direction of
load application. The fracture strength was calculated as follows:

Strength (psi) 2P. Eq.


TrLD (5.9)

where
P=breaking load (lb)
L=length (in.)
D= diameter (in.)

Specimens were fabricated by hydrostatic pressing at 100,000 psi


and sintering in hydrogen at 1,750°C; MCW UO, powder was used.
Strength data are tabulated in Table 5.12.

5.4.2 Elastic Properties

The elastic moduli of uranium dioxide, as determined by the sonic


method, were reported by Lang; these data are shown in Table 5.13
[64]. Modulus of elasticity values reported by other workers are
summarized in Table 5.14 [65]. Other than experimental error, these
data are expected to be affected significantly only by porosity. The
approximate effect of porosity is given by

Polis-É, I'meas
Eq. (5.10)

where P is the volume fraction of pores.


The modulus of elasticity of a specimen of UO, ceramic having a
density of about 10.2 g/cc decreased from 26.5× 10° to 24×10° psi
when heated from room temperature to 800° C [66].

TABLE 5.12—STRENGTH PROPERTIES OF UO, SOLID CYLINDERS (63)

Material
Fracture
Bulk density strength"
Average (g/cc) (psi)
Powder treatment particle
size (u)

As received------------------------------ 25 9. 1 5,400

Screened through 325 mesh -- - - - - - - - - - - - - - - 9 9.3 6,090


Dry milled------------------------------ 5 10.2 2, 320

*Average of about 5 specimens.


TABLE 5.13–ELASTIC PROPERTIES OF UO, AT ROOM TEMPERATURE [64]

Percent Young's modulus Shear Poisson's ratio Bulk modulus


bulk of theo- Inodulus

of
Type UO, density retical
g/cc density

G
El Elw Eſl

w
All Auf Kl Krw

||

|
MCW oxide cold pressed and 10.37 94.6 1,930kilobars" 1,940kilobars 1,930kilobars 745kilobars 0.302 0.306 1,620kilobars 1,660kilobars

at

in
sintered H1 1750°C; O/U (28.0×108 (28.1×108 (28.0×108 (10.8×108 (23.5x108 (24.1×108
ratio =2.02** psi) psi) psi) psi) psi) psi)

|
NH3-precipitated oxide, cold 10.19 93.0 1,820kilobars 1,740kilobars 1,840kilobars 705kilobars 0.291 |---------- 1,460kilobars

at

in
pressedand sintered H2 (26.4x10% (25.2×108 (26.7×108 (10.2×108 (21.1×108
1750°C; O/U ratio not givent psi) psi) psi) psi) psi)

El-Young's modulus calculated from the longitudinal resonant frequency.


Ff Young's modulus calculated from the flatwise flexural vibration.
Efl Young's modulus calculated from the edgewiseflexural vibration.

===
w G
Shear modulus calculated from the torsional resonant frequency.

u-j-1

to
K= E/3 211)

(1 =

*1
kilobar 10° dynes/cm2.

5
of
**Average specimens.

fl
specimen.
§
208 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 5.14–ROOM-TEMPERATURE MODULUS OF ELASTICITY OF


UO, CERAMICS

Investigator
Emea,X10-0
(psi)
Bulk density
(g/cc)
*
Approximate"

F-idpSl

Lambertson and Handwerk [47]-- - - - - - - 21 10. 6 23


Lang (64]---------------------------- 28 10. 4 31
Bowers (63]-------------------------- 21. 3 10. 0 26
22. 1 9. 6 30
Johnson [65]------------------------- 25 10. 0 31

*Calculated from Eq. (5.10).

The effect of nonstoichiometry on Young's modulus was recently


[60]. Measurement was made by the sonic
al.

determined by Scott, et
rectangular composition UO,
on

and
A of

of
method bar section

is
a

density 10.5 g/cc. 18×10° psi was obtained.


of

value The value

to
for the same specimen, after reduction hydrogen

in in

at
1,200°

C
stoichiometric oxide, was 10° psi, agreement with the data
27
×

shown in Table 5.14.


The value for Poisson's ratio (see Table 5.13) somewhat higher

is
on

of

expected These
of

than the basis values other brittle materials.


materials usually have Poisson's ratio
of

around 0.20.
a

the compressibility (recip


of

From the data Table 5.13, the value


in

UO, 0.62× 10-14 cm”/dyne. On the


of

of

rocal bulk modulus)


is

assumption that the atomic binding forces UO, are predominately


in

ionic character,” the compressibility absolute zero can also be


at
in

computed from the expression [67]:

Be?A, Eq. (5.11)

ionic
in

where the Born constant for the b/r" repulsive interaction


is
n

crystals, assumed equal 10.5 for UO, constant characteristic


to

is is
a
a
;

V/r.
W,
of

the lattice equal the molar volume,


V
to

where
is
V

Avogadro's number, and the cation-anion distance (2.368 ang


r,
is

strom units for UO.); the compressibility cm"/dyne:Are


in

is
is
B

the Madelung constant equal 5.0388 for the fluorite lattice; and
to

the electronic charge. The calculated value 0.79×10−14 cm.”/dyne,


in is

is

good agreement with the value 0.62×10−" cm°/dyne


of

- obtained
from the bulk modulus measurement.

*The justification for this assumption more fully explored Chap.


6.
is

in
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 209

looo [-T-I-T-T—T-T—T-T—r—r

9 OO

800

7oo

666 tº
LAMBERTSON AND HANDWERK

2.OO 2O5 2.10 2.15


O/U ATOM RATIO

FIGURE 5.13. Effect of Oxygen Content on Hardness of UO,.. Solid Solution;


Samples Annealed at 900° C for 67 Hours and Quenched to Room Tempera
ture; Knoop Diamond Indentor, 500 g Load [5]. (Courtesy, Journal of Nuclear
Materials.)

5.4.3 Hardness

Lambertson and Handwerk reported the Knoop hardness (load


unspecified) of an inclusion-free sample of electrically fused UO, to
be 666+14 [47]. This corresponds to a hardness of 6 to 7 on Moh's
scale and is in agreement with unpublished Battelle measurements.
Schaner the effect of oxygen content on the hardness of
reported
nonstoichiometric single-phase UO... [5]. The hardness of annealed
and quenched samples of single-phase UO., was measured by using
a Knoop diamond indentor with a load of 500 grams. The average
of 15 hardness impressions is plotted as a function of connposition in
Fig. 5.13. A small amount of excess oxygen results in a large in
crease in the hardness of UO2.x.

5.4.4 Thermal Stress Resistance

When heat flows through a body, the concomitant temperature


differences create a stress system. The nature of this stress system
depends on conditions of heat flow, the shape and size of the body, and
210 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

certain properties of the material. With brittle-state materials, when


the level of such a stress becomes excessive, relief occurs through
fracturing. It follows that an understanding of factors affecting
fracture stress is important to an understanding of the thermal
stress resistance of brittle materials. Some comments on this subject
were given previously in Sect. 5.4.1.
In the case of steady-state heat flow, the following material factor,
M, can be used to indicate the resistance of a material to thermal
fracture:

M_{{
Eo.
Eq. (5.12)

where
P=fracture stress
K=coefficient of thermal conductivity
E=Young's modulus
of thermal expansion
a = coefficient
The material factor, M, is temperature dependent.
Approximate values of M were determined for two UO, ceramics in
some cursory experiments at Battelle [68]. The ceramics were pre
pared by hydrostatically pressing at 100,000 psi both MCW PWR
type powder and a Battelle-prepared sinterable powder. The MCW
powder was then sintered in hydrogen for 2 hours at 3,200°F, while
the Battelle powder was sintered in hydrogen for 2 hours at 3,000°F.
The specimens were hollow cylinders, approximately 14 inch ID, 12
inch OD, and 1% inch long. Both ceramics were porous: the density
of the one prepared from MCW powder was 9.48 g/cc and for the
other, it was 10.4 g/cc.
The specimens were fractured by radial steady-state heat flow, and
the heat flow at fracture was measured. Using this measured value
and the dimensions of the specimen, an approximate value of M was
calculated. The results were as follows:

Number of Ave fracture Ave heat flow at fracture Approximate


Material specimens temperature (watts/cm) material factor
• C) M (watts/cm)

MCW powder. . . .. . 4 ~800 28.4 + 4* - - - - - - - - - - 1. 19

BMI powder --- - 6 <760 24.3 + 4.8°-- - - - - - - - 0. S7

"90 percent confidencelimits.

For comparison, the material factor of commercial high-alumina


porcelain was found to be about 2.76 watts per cm. This can be con
sidered to mean that, under similar conditions of steady-state heat
transfer, the alumina porcelain can be expected to withstand about
three times the heat flow of these UO, ceramics without fracturing.
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 211

By substituting the above values of M and values for K, a, and E


from Sects. 5.3.1, 5.3.4, and 5.4.2 in Eq. (5.12), the computed frac
ture stress P is 7,300 psi for the specimens made with the MCW
powder and 6,700 for those made with the BMI powder. These values
of P are reasonably in line with mechanical strengths measured on
UO, ceramics (see Sect. 5.4.1).
Both theoretical and experimental work at Massachusetts Institute
of Technology and at Battelle" have shown that the material factor
decreases markedly with porosity, primarily because of the effect of
porosity on fracture stress [69]. However, the cursory experiments
at Battelle do not reflect this expected effect of porosity on M for the

two UO, ceramics. The reason for this apparent insensitivity to


porosity is not known. Possibly, a temperature sensitivity in the
thermal-fracture resistance of UO, ceramics is the cause; the two
ceramics failed at different temperatures. In view of the porosity of
these two ceramics, however, it can be assumed that the data do not
represent the level of thermal-fracture resistance potentially attain
able in UO, ceramics.
In the case of transient heat flow, the effect of properties of a mate
rial on its resistance to thermal fracture is different from that in the
case of flow. For an extreme transient (shock) condi
steady-state
tion, the conductivity and other heat-transfer properties of the mate
rial do not enter, and for a crude indication of a material's resistance,
the material factor (M) can be divided by the coefficient of thermal
conductivity. The unknown effect of rate of stressing on fracture
stress P, for example, causes uncertainty in the use of such a thermal
shock material factor. If fracture occurs in a heat transfer situation
between steady-state and extreme transient where the stress system is
changing with time, the diffusivity and the manner in which Poisson's
ratio enter the expression for thermal stress become important. No
general method of satisfactorily assessing a material's resistance to
thermal fracture under these conditions is known.

5.5 ELECTRICAL, MAGNETIC, AND OPTICAL PROPERTIES


R. K. Willardson and J. W. Moody

5.5.1 Introduction

Several investigators have reported measurements of the electrical


conductivity of the various stable uranium oxides [50, 70–80]. Both
TO, and U2O, are reported to be metal-excess semiconductors with
electronic conduction. Uranium dioxide, on the other hand, has been
considered to be a metal-deficit (oxygen-excess) semiconductor with

* Battelle Memorial Institute, unpublished data.


57.4789 0–61 15
212 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

the electric current being carried by positive holes; however, it is not


known whether the positive carriers arise from excess oxygen atoms in
interstitial positions or from uranium atom vacancies.”
The magnitude of the room-temperature conductivity of UOs is
quite low, ranging from 10-7 to less than 10-5 ohm' cm-". Part of
the conductivity of U.O., is believed to be ionic in nature. Reported
values of the total room-temperature conductivity of UAOs range
from 2×10−" to 1 × 10−7 ohm-1 cm−". Values for the p-type con
ductivity of UO, range from 3× 10- to 4×10-5 ohm-' cm-". These
wide ranges of conductivity are caused by variations in the composi
tion, purity, density, and crystalline perfection of the specimens
studied.
Previous summaries of the electrical properties of UO, have been
published by Meyer and, more recently, by Katz and Rabinowitch
[1, 79]. Meyer studied the electrical conductivity of UO, as a func
tion of temperature and determined “activation energies” for free
carriers in samples of widely different conductivities [72]. He ob
tained values ranging from 0.3 to 0.9 ev. Chiotti also studied the con
ductivity of UO, as a function of temperature, but extended the
measurements to higher temperatures where intrinsic conductivity
becomes predominant [76].
Hartmann is the only investigator to report Hall-coefficient measure
ments on UO, [73]. A room-temperature conductivity of 1.3 × 10-4
ohm-' cm-' and a Hall coefficient of +770 cc per coulomb were re
ported for a sintered specimen having a bulk density of 6.5 g/cc.
Although this density value is far too low for sintered UO2, the mag
nitude and temperature dependence of the electrical conductivity are
consistent with those reported for dense UO, specimens.
Willardson, et al., studied the dependence of the electrical conduc
tivity and thermoelectric power of UO, on composition, temperature,
and thermal history. They related the magnitude and type of con
ductivity found to the phase relationships in the UO2–UAOs system
(see Chap. 6) [78].
The magnetic and optical properties of UO, have not been reported
on as extensively as the electrical
properties. Dawson and Lister and
Arrott and Goldman studied the magnetic susceptibility of uranium
oxides in the region of UO, to U.O., as a function of temperature from
2° to 570° K [81, 82]. Arrott and Goldman concluded that, when oxy
gen is added to UO,
the anion lattice is undisturbed and the oxygen
ions enter into interstitial positions in the fluorite structure. With
increasing oxygen content, the U" ions are probably replaced by
U” ions. Low-temperature magnetic susceptibility behavior of UO,
indicates a transition to an antiferromagnetic state at 28° K.

* See Chap. 6 for a further discussion of this point.


PHYSICAL PROPERTIES OF URANIUM DIOXIDE 213

Ackermann, et al., the optical absorption of UO2.


investigated
films as a function of heat treatment and oxygen partial pressure [83].
Their data indicate a region of strong absorption at about 2.7 ev and
an absorption edge at 5 ev.

5.5.2 Electrical Properties

To a large extent, the extrinsic electrical properties of uranium


oxides are determined by the relationships among the various phases
in the UO2–UO, system. Indeed, a discussion of any structure
sensitive property of UO, should include information about the com
position and heat treatment of any particular specimen. It is to be
expected that a wide range of values of electrical conductivity, mag
netic susceptibility, etc., can be obtained among several specimens and
that the values for any one specimen may not be reproducible after
heating to elevated temperatures.
Some typical values of the electrical conductivity, and its de
pendenceon temperature, of various specimens of UO, are shown by
the curves of Fig. 5.14 [78]. The specimens used for this study were
lo

-
iO 4.

10-5

10-6
I 2 3. 4
3
loº/ T(*k)
FIGURE 5.14. Conductivity of Sintered Specimens of n- and p-Type UO, [78].
(Reprinted with permission from R. K. Willardson, et al., “The Electrical
Properties of Uranium Oxides,” Pergamon Press, Inc., 1958.)
214 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

cut from rods of UO, which had been sintered in a hydrogen atmos
phere. The magnitude of the room-temperature conductivity of the
p-type specimens (solid symbols, Fig. 5.14) ranges from about 3 × 10−"
to nearly 10-" ohm-' cm-". Over the extrinsic range, the activation
energy to free a positive carrier is about 0.4 ev. These specimens
were of near-stoichiometric composition. The extrinsic portions of
the n-type specimens (open symbols) show a greater dependence on
temperature than do the p-type specimens. These specimens con
tained somewhat less excess oxygen than did the p-type specimens.
Values of activation energies in n-type specimens ranging from 0.4 ev
for high-conductivity specimens to 0.9 ev for low-conductivity ma
terial have been measured.
In general, these activation energies agree with those reported by
Brabers who measured the temperature dependence of the conductivity
of sintered specimens of UO, that contained various amounts of excess
oxygen [80]. Brabers noted a slight decrease of activation energy with
increasing oxygen content. The activation energy of all specimens
was observed to change between 400° and 600° C. The change in
activation energy at elevated temperatures was possibly due to a phase
change as discussed below.
The anomalous behavior of the conductivity of specimen A (open
triangles of Fig. 5.14) between 250° and 400° C is sometimes ob
served in sintered specimens of UO. Although the reason for this
behavior is not understood, it is believed to be associated with the
presence of a metallic impurity, such as molybdenum, or a nonmetallic
impurity, such as nitrogen. Nitrogen is a possibility, since the sample
was prepared from uranyl nitrate, and it is possible that nitrogen does
persist through such a preparation. Subsequent analysis of the sample
revealed the presence of molybdenum resulting from contamination
of the sample by the molybdenum setter used during the sintering
process. This anomaly was not observed in samples of lower impurity
content.
Since the high-temperature portion of Fig. 5.14 appears to be
independent of the starting material, it is reasonable to associate this
behavior with the intrinsic conductivity of UO. A value of 3.0 ev
for the energy-band separation of UO, can be calculated from the
slope of this portion of the curve.
The thermoelectric power of UO, displays an anomalous tempera
ture dependence. This behavior is illustrated by the curves of Fig.
5.15 [78]. Here the thermoelectric power of some of the same samples
used for the previous conductivity measurements is plotted as a func
tion of reciprocal temperature. For carrier concentrations which
increase with temperature, simple theory predicts a decrease of ther
moelectric power. This is obviously not the case for the p-type curves
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 215
14OO

2OO H n
* \

\
|OOO |

800 I
ſ m
$2 º,
3. | |-s
~ 600
p-TYPE
§o
*
#
H.
400
6
fº*ee-e
3 d
ºf 200
o
>
ar
#
*- O

B
Ooooooo
-2OO
n-TYPE

-400
C

Dººse
TEMP (*C)

-6OO
; ; ; ; ; ;
§ 31–1–i-1–11–4–1–1–1 *.*
l 2 3

loº/ T(ok)

FIGURE 5.15. Thermoelectric Power of Sintered Specimens on n- and p-Type


U0, [78]. (Reprinted with permission from R. K. Willardson, et al., “The
Electrical Properties of Uranium Oxides,” Pergamon Press, Inc., 1958.)

where an initial sharp rise in thermoelectric


power (although the
conductivity increases) is noted. Even the n-type curves are rela
tively flat and do not decrease as would be expected. The p-type
rurves are typical of specimens of near-stoichiometric composition.
The thermoelectric power of p-type specimens with greater oxygen
content behaves similarly to the n-type specimens of Fig. 5.14, that
the

thermoelectric power relatively independent temperature


of
is,

is

over wide range.


a

Meyer considered UO,


an

Although amphoteric
be

semicon
to

subsequent investigations failed


of
to

ductor, confirm the existence


a

phase UO, (see Chap.


of

(or substoichiometric)
6)

metal-excess [79].
by

The n-type conductivity found some specimens has been shown


in
by

Willardson, al., and Vaughan and Willardson


be
to

associated
et

distinct phases within the UO,-U.O., system


of

with the existence


The dependence the room-temperature electrical conduc
of

[78,84].
(both type and magnitude) UO,
on

excess oxygen content


of

tivity
216 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
200 r- I
T

A. AS PREPARED, ALL p-TYPE


B. STABILIZED AT 200 °C
C. STABILIZED AT 400°C

150

IOO

50}.

O
2.O. 2. 2.2 2.3
OXYGEN-URANIUMRATIO

FIGURE 5.16. Room-Temperature Electrical Conductivity as a Function of


Oxygen-Uranium Ratio of Pressed Specimens of UO, [78]. (Reprinted with
permission from R. K. Willardson, et al., “The Electrical Properties of
Uranium Oxides," Pergamon Press, Inc., 1958.)

and the phase relationships in the UO2–UAOs system are illustrated


in Fig. The initial samples, prepared by oxidizing MCW
5.16 [78].
325 mesh UO, powder in dry oxygen at 180°C, were
all

p-type. The
very slight rise conductivity with increasing oxygen content sug
of

gests that only part the absorbed oxygen actually entered the UO.
of

was necessary
to
of

oxidation.
It

lattice under the conditions anneal


samples temperatures homogeneous
of at

to

the elevated order ensure


in

oxygen atoms. Along with the process


of

distribution diffusion
ordering process evidenced
by
an

and homogenization, there occurs


conductivity changes and the appearance U.O., and U.O. Curves
of

Fig.
on

specimens homogenized and


of

and 5.16 were obtained


B

quenched from 200° and 400°C, respectively. The p-type conductiv


up
ity

UO, increases with excess oxygen content the solubility


of

to
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 217

limit of oxygen in the UO, lattice (O/U ratio of 2.08 at 200° C).”
Above this limit, the conductivity decreases as the amount of the U.O.,
phase increases. When U.O, predominates, the conduction mechanism
changes and the material becomes n-type. At still higher oxygen
content, the n-type, metal-rich phase of UAO, appears and the conduc
tivity decreases to a low value as the stoichiometric composition is
reached.

Because of their electrical behavior, the phases within the UO2–


U.O., system can be identified, according to Willardson, et al., as:
1. UO2.x, a metal-deficit semiconductor (p-type).
2. U.O.,
a metal-excess semiconductor (n-type).
3. UsO,..., a metal-excess semiconductor (n-type).
Although UO, is classified as a metal-deficit semiconductor, it should
be emphasized that is not known whether the p-type character of
it
the material arises from interstitial oxygen or uranium vacancies.
Either possibility would give rise to p-type conduction and would be
consistent with the results presented here.
The effects of impurity additions of CaO and MoC), on the electrical
resistivity of sintered specimens of UO, are illustrated in Fig. 5.17.”
Because of the loss of material during sintering, the uncertainty of the
oxygen content of the UO, and the difficulty of achieving homogeneous
doping by diffusion during a short sintering period, the slightly higher
resistivity of the sample containing molybdenum is probably not sig
nificant. Additions of calcium, however, reduce the resistivity by an
order of magnitude.
The effect of calcium would be expected, since a metallic impurity
of less than +4 valence, if substituted for uranium, should increase
p-type conductivity UO,
the

of

of

manner similar indium


to

that
in
a
or

p-type germanium magnesium p-type AlSb. The substitu


in
in

U"
expected reduce the p-type con
be
of

+6 ion for
to

tion would
a

ductivity UO, and even make n-type present


of

if

sufficient
in
it

quantity. This was not observed when Mo(), was added UO2,
to

an

indicating the probable reduction the +6 ion. Nevertheless,


at of

resistivity
of

anomalous increase about 500° was noted for


C

molybdenum-containing samples.
Attention the similar
to

called
is

this anomaly with that previously discussed.


ity
of

unfortunate that detailed studies of the effect of nonmetallic


It
is

of on

impurities the electrical properties UO, As


of

have not been made.


metallic impurities, elements such
as

the case nitrogen (which


in

may persist through preparation techniques involving uranyl nitrate)


profoundly physical and chemical properties UO.
of

may affect the


by

Attempts directly dope UO, with nitrogen exposure gaseous


to

to

... however, Sect. 5.2 and Chap.


6.

Battelle Memorial Institute, unpublished data.


218 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

IO-I
O | 2 s
loº /T ("k)

FIGURE 5.17. Effect of Metallic Impurities on the Electrical Resistivity of


Sintered Specimens of UO,.

nitrogen at elevated temperatures or by the addition of uranium


nitride or uranyl nitrate to UO, powder have been inconclusive * due,
in part, to the difficulty of proving the presence of nitrogen after
ductivity of UO2, and the addition of 5 mole percent uranyl nitrate
doping. All three procedures result in a reduction of the p-type con
actually produces n-type material even when prepared at room tem
perature. Since, however, X-ray examination of the specimen re
vealed that it consisted of UO, and a second phase identified as UOs,

* Pattelle Memorial Institute, unpublished data.


PHYSICAL PROPERTIES OF URANIUM DIOXIDE 219

it is not clear whether the changes in electrical properties were due


to the presence of nitrogen in the UO, lattice or to the partial oxida
tion of UO, by the nitrate.
An approach to a better understanding of the electrical properties
of uranium dioxide which is different from that discussed above is
suggested by the recent results of Aronson.” In this work measure
ments were made of the electrical conductivity and thermoelectric
power of nonstoichiometric uranium dioxide at temperatures of 500–
1,150° C by using a d-c current-potential method. The measurements
were made on plates of high density (99 percent of theoretical) and
large grain to minimize the effects of grain boundaries
size (601)
and pores on the electrical properties. The compositions of the plates
were in the range UO, UO, The conductivity data
23.
ool

in
to

the
single phase UO2.x region were represented by the equation

=ºne)
3.8X10°
(ohm—cm)−"— (2x) (1–2x) exp Eq.(5.13)
a

Q,
An approximate equation for the thermoelectric power, was given
as

**)
k

=; Eq. (5.14)
in

2x
Q

where Boltzmann's constant and the electronic charge.


is

is
k

an
Aronson suggests that the electrical data are agreement with
in

UO, which was postulated


of

ionic mechanism nonstoichiometry


in

thermodynamic evidence [85, 86]. The mechanism


on

of

the basis
UO, lattice consists U" and U"
of

assumes that the nonstoichiometric


lattice cations and O-4 lattice and interstitial anions. The electrical
properties result from the activated transfer electrons between U"
of

and U" ions. The thermoelectric power arises from the temperature
dependence the chemical potential
of

of

the electronic carriers. The


U" ions rather than U" ions must bring
of

presence
be

to

assumed
the theory and the experimental data into good agreement.

5.5.3 Magnetic Properties

the magnetic susceptibility ura


of

of

of

Measurements the oxides


by

by

nium have been made Dawson and Lister and Arrott and Gold
man [81, 82]. both investigations indicate that U"
of

The results
ions, rather than U"
ions, replace U" ions oxygen dissolved by
as

is

UO. These results are not conclusive; Haraldsen and Bakken have
interpreted their magnetic measurements favoring the U"
on

as

UAOs

*S. Aronson, personal communication, Bettis Atomic Power Laboratory.


220 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

formulation [36]. After correcting for the diamagnetism of the ura


nium and oxygen atoms, Dawson and Lister found a magnetic moment
of 3.11 Bohr magnetons for UO,. This value corresponds to a 5f.
(or possibly 5f 6d') electronic configuration for the U" ion.
The measurements of Arrott and Goldman, which were extended
down to 2° K, suggest a transition to an antiferromagnetic state at
about 29° K for oxides close to stoichiometric UO2. The temperature
of the transition corresponds to the temperature at which an anomaly
in the specific heat of UO, was observed by Jones, et

al.
(see Sect.
5.3.2) [35].
From the general increase the interaction parameter in

of
with
creasing oxygen content, Arrott
and Goldman conclude that excess
oxygen enters interstitial positions the UO, lattice. The presence

in
positions UO,
of

oxygen interstitial consistent


in

in
ions

is
excess
with the p-type character the electrical conductivity com
of

of
the
pound, although p-type conduction usually

in
metal-deficit oxides

is
associated with cation vacancies.
magnetic susceptibility with composition well
of

The variation

is
a
When such data are employed

in
recognized tool for phase analysis.
conjunction with data obtained from independent measurements, such
electrical conductivity, X-ray analysis, thermal analysis, con
as

or

light may phase relationships par


be

on

siderable shed the within

a
susceptibility compli
In

ticular system. general, the magnetic

is
a
cated and unpredictable function composition within single-phase
of

a
susceptibility will vary
In

region. two-phase region, however, the


a

linearly with composition the composition the two phases are


of
if

fixed and only the relative abundance the two phases changes.
of

the magnetic susceptibility with composition


of

The variation
within the UO,-U3Os region Fig. 5.18 [81, 82]. The
in

shown
is

strange behavior the susceptibility within the single-phase region


of

UO,
In
extending presumably magnetic transitions.
to

to

due
is
or

the two-phase region from UO, UO, the magnetic susceptibility


to

s,
a

decreases, and second two-phase region appears above UO2.s.


a
At

high oxygen contents, the data Dawson and Lister agree well
of

Discrepancies
of

with those Arrott and Goldman. lower


at

are noted
oxygen concentrations, particularly the two-phase region cor
in

responding oxygen-to-uranium about 2.5. Part


to of of

to of
to

to

ratios 2.3
the discrepancies between the two sets
be

data can ascribed


of

uncertainties the oxygen content and the past histories the


in

various samples (see discussions Chaps. and 8).


in

5.5.4 Optical Properties

the optical properties UO, prob


on

of

There little information


is

ably owing the difficulty preparing crystalline samples suitable


of
to
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 221

3o

O ARROTT AND GOLDMAN


O DAWSON AND LISTER
/N
/ N.
N.

/& N.
/ PS.
>~ /
/ \\
º,
a
\
\
20
g P- * >5 N. /
/
7. 2°K
s Z / *b N
É &- AC N, / 8
>
*-
=!
^
N.
N.
\ſ e &r
**
N.
N.
\
\
\
sºl. Q N. 2^
N.* \
* Y->
oK

#
\
g 10
*i-s-s-ºs-
e-J ><- _--Q-TN as—b
t N.

#
O-
•- L *O-
T~
O
-*S * NN IOoºk
NO
233+-e-Tº- 300-k
T*-Sse
*-es - Q-T-e-200°K's
Ye
--~ JN * YOS.
*

°ºo 2.1 2.2 2.3 2.4 2.5 2.6 2.7

oxygen-URANIUM RATIO

FIGURE 5.18. Magnetic Susceptibility as a Function of O/U Ratio [81, 821.

for measurement. Ackermann, et al., have reported on the optical


absorption of UO., films deposited by sublimation and condensa
tion on fused (silica disk) substrates [83]. Measurements were made
in the wavelength region from 220 to 800 mp. The amount of excess
oxygen in the films was varied by exposing the films to an oxygen
partial pressure at high temperature. The actual values of a were
not determined, but were estimated by reference to the aging condi
tions and to phase diagram studies (see Chap. 6).
As shown in Fig. 5.19, the absorption coefficient of films of low
excess oxygen content is small down to 650 mp, then rises to a maxi
mum at about 320 mp. Beyond the maximum, the absorption co
efficient drops to a minimum at 245 mp, then rises rapidly. Approxi
mate values for the refractive index show a maximum of 2.7 at 650 mp.
and a minimum of 1.2 at 375 mp.
Excess oxygen reduces the absorption in the 340 to 480 mp region
other regions (Fig. 5.19). This
all

and increases the absorption in


by

suggests that the one unresolved band oxygen


at

430 mp removed
is
at

and the one 300 mp becomes more intense.


222 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TREATMENT AT ~950°C
CURVE (THICKNESS=73 m.)

A - VACUUMOF O-6 mm Hg FOR 2 weeks


B – 25 x 0 ° mm Hg of oxygen FOR 8 Hours

c - 0 °mm Hg of oxygen For 74 Hours


>
D - 4 x 107* mm Hg of oxygen FOR 24 Hours

;
5
co
| |
OxYGEN - URANIUM RATIOs
|


I.O. H. A - O/U =2.OO

§
§
O.8 H B - O/U= 2. 14). UO2+x FLUORITE PHASE
-
C O/U =2, 18
D-O/U= 2.23 —U4 Os PHASE
O.6

-
F-E-1–
O.4

O.2 -

O
2OO 3OO 4OO 500 6OO 7oo 8OO

WAVELENGTH(mu)

FIGURE 5.19. Effect of Oxygen Partial Pressure on the Optical Density of Thin
Films of UO2,... [83]. (Reprinted with permission of R. J. Ackermann, et al.,
and the Journal of the Optical Society of America.)

Results from the dependence of the optical density on wavelength


with increasing oxidation at low temperature (160° C) show qualita
tive agreement with the high temperature data shown in Fig. 5.19.
At low temperatures, however, oxidation causes the growth of a
broad band near 600 mp.
To determine if the optical constants observed for thin films of UO.
apply as well to bulk material, Companion and Winslow made diffuse
reflectance measurements on UO, and nonstoichiometric uranium diox
ide powders (1771, particle diameter) [87]. The optical properties
were found to be in general agreement with those obtained on thin
films. Moreover, the same effects of excess oxygen were observed. A
small peak found near 665 mp and also detected by Gruen was not
observed in the film studies of Ackermann, et [83,88].
al.

Measured values of the index of refraction and the extinction


"

coefficient have been compared with values calculated using the


Kronig-Kramers expression relating the real and imaginary parts
of

the complex dielectric constant [89]. The experimentally derived


by

values are represented the points Fig. 5.20, while the solid
in

lines give the calculated values.


Compare with data presented Chap.
3.
in
*
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 223
4 I I T
I

3 -
INDEx OF
REFRACTION

| H ExTINCTION -
COEFFICIENT

ENERGY, ev

FIGURE 5.20. Index of Refraction and Extinction Coefficient of UO, as a Func


tion of Energy [83]. (Reprinted with permission of R. J. Ackermann, et al.,
and the Journal of the Optical Society of America.)

It
is interesting to note that the region of strong absorption at
300 to 400 mg corresponds roughly to the value of 3 ev for the energy
gap of UO, calculated from conductivity data. The absorption be
yond 245 mp, however, suggests a higher value of energy-band sep
aration (about 5 ev). If the higher value corresponds to the true
intrinsic energy gap in UO2, it is difficult to understand the magnitude
and sign of the high-temperature electrical conductivity of UO2.
Actually, calculations using measured values of the conductivity
and reasonable values of carrier mobility and effective mass
(u-1 cm”/volt-sec,
m*=1) indicate energy-band separation much less
than 3 ev. Measurements of conductivities, carrier mobilities, and
optical absorption of single crystals of UO, probably would resolve
these discrepancies.

REFERENCES

1. J. J. KAtz and E. RABINow ITCH, “The Chemistry of Uranium,” National


Nuclear Energy Series, I)iv. VIII. Vol. 5, McGraw-Hill Book Co., New York,
1951.
2. P. PERIo, “The Oxidation
of Uranic Oxide at Low Temperatures,” Bull. soc.
chim. France 20, 256–263 (1953).
3. F. GeoN vold, “High Temperature X-Ray Study of Uranium Oxides in the
UOI-U-O, Region,” J. Inorg. Nuclear Chem. 1, 357–370 (1955).
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 223
4 I I T
I

3 -
INDEx OF
REFRACTION

| H ExTINCTION -
COEFFICIENT

ENERGY, ev

FIGURE 5.20. Index of Refraction and Extinction Coefficient of UO, as a Func


tion of Energy [83]. (Reprinted with permission of R. J. Ackermann, et al.,
and the Journal of the Optical Society of America.)

It
is interesting to note that the region of strong absorption at
300 to 400 mg corresponds roughly to the value of 3 ev for the energy
gap of UO, calculated from conductivity data. The absorption be
yond 245 mp, however, suggests a higher value of energy-band sep
aration (about 5 ev). If the higher value corresponds to the true
intrinsic energy gap in UO2, it is difficult to understand the magnitude
and sign of the high-temperature electrical conductivity of UO2.
Actually, calculations using measured values of the conductivity
and reasonable values of carrier mobility and effective mass
(u-1 cm”/volt-sec,
m*=1) indicate energy-band separation much less
than 3 ev. Measurements of conductivities, carrier mobilities, and
optical absorption of single crystals of UO, probably would resolve
these discrepancies.

REFERENCES

1. J. J. KAtz and E. RABINow ITCH, “The Chemistry of Uranium,” National


Nuclear Energy Series, I)iv. VIII. Vol. 5, McGraw-Hill Book Co., New York,
1951.
2. P. PERIo, “The Oxidation
of Uranic Oxide at Low Temperatures,” Bull. soc.
chim. France 20, 256–263 (1953).
3. F. GeoN vold, “High Temperature X-Ray Study of Uranium Oxides in the
UOI-U-O, Region,” J. Inorg. Nuclear Chem. 1, 357–370 (1955).
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 225

26. R. Scott, “Thermal Conductivity of UO,” AERE–M/R–2526, Mar. 1958.


27. J. FRANCL and W. D. KINGERY, “Thermal Conductivity: IX, Experimental
Investigation of Effect of Porosity on Thermal Conductivity,” J. Am.
Ceram. Soc. 37 (No. 2, Part II), 99–107 (1954).
28. H. W. RUssel.L., “Principles of Heat Flow in Porous Insulators,” J. Am.
Ceram. Soc. 18, 1–5 (1935).
29. A. EUCKEN, “Thermal Conductivity of Ceramic Refractory Materials; Cal
culations of Thermal Conductivity from the Constituents,” Forsch. Gebiete
Ingenieuruc., B3 Forschungsheft No. 353, 16 pp. (1932).
30. A. M. Ross, “The Dependence of the Thermal Conductivity of Uranium
Dioxide on Density, Microstructure, Stoichiometry, and Thermal Neutron
Irradiation,” CRFD–817, Sept. 1960.
31. R. W. NICHOLs, “Ceramic Fuels—Properties and Technology,” Nuclear Eng. 3,
327–333 (1958).
32. H. SHAPIRo and R. M. Powers, “High Conductivity UO. Terminal Report,”
SCNC–294,
Oct. 1959.
33. R. M. PoweRs, Letter to Editor, Nucleonics 18, No. 10, 6 (Oct. 1960).
34. J. A. L. Robertson, A. S. BAIN, and A. RIDAL, “The Effect of 4 Mole Per
Cent Y.O., on the Thermal Conductivity of UO,” CRFD-933, June 1960.
35. W. M. Jones, J. GoRDON, and E. A. LoNg, “The Heat Capacities of Uranium,
Uranium Trioxide and Uranium Dioxide from 15 to 300° K,” J. Chem.
Phys. 20, 695–699 (1952).
H. HARALDse N and R. BAKKEN, “Magnetic Properties of Uranium Oxides,”
Naturiciss. 28, 127 (1940).
E. A. LoNG, W. M. Jon Es and J. GoRDON, U.S. Bureau of Mines, A-329, 1942;
G. E. Moore and E. A. LoNG, U.S. Bureau of Mines, A-329, 1942.
J. P. Cough LIN, “Heats and Free Energies of Formation of Inorganic
Oxides,” U.S. Bureau of Mines, Bull. 542 (1954).
39. M. M. Popov, G. L. GAL'CHENKo, and M. D. SENIN, Zhur. Neorg. Khim. 3,
1734–1737 (1958); Chem. Abst. 53, 21120c (1959); Ceram, Abst., 103.j
(Apr. 1961).
40. H. H. HAUSNER and R. G. MILLs, “Uranium Dioxide for Fuel Elements,”
Nucleonics 15 (7), 94–103 (1957).
41. S. GLAssTone, “Textbook of Physical Chemistry,” 2nd ed., p. 421, D. Wan
Nostrand Company, Inc., New York, 1946.
42. J. DE LAUNAY, “The Theory of Specific Heats and Lattice Vibrations” in
“Solid State Physics,” F. Seitz and D. Turnbull, eds., Vol. 2, pp. 220–303,
Academic Press, Inc., New York, 1956.
43. F. SEItz, “The Modern Theory of Solids," p. 133, McGraw-Hill Book Com
pany, Inc. New York, 1940.
44. P. MURRAY and R. W. THAcKRAY, “The Thermal Expansion of Sintered UOs,”
AERE MAM 22.
45. M. D. BURDIck and H. S. PARKER, “Effect of Particle Size on Bulk Density and
Strength Properties of Uranium Dioxide Specimens,” J. Am. Ceram. Soc.
39, 181–187 (1956).
46. BELL and S. M. MAKIN, “Fast Reactor—Physical
I. P. Properties of Materials
of Construction. Review of Progress from September 1, 1953 to April 1,
1954,” RDB (C)/TN–70, July 5, 1954.
47. W. A. LAMBERTson and J. H. HANDw ERK, “The Fabrication and Physical
Properties of Urania Bodies,” ANL–5053, Feb. 1956.
48. F. A. HALDEN, H. C. Woh LEks, and R. H. REINHART, “Thermal Expansion
of Uranium Dioxide,” Stanford Research Institute Report SRI A–6, Apr.
15, 1959.

49. O. RUFF and D. GoecKE, “Melting and Vaporization of High Temperature


Refractories,” Z. angew. Chem. 24, 1459 (1911).
226 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

50. E. FRIEDERICK and L. SITTIG, “Preparation and Properties of High Melting


Lower Oxides,” Z. anorg. u. allgem. Chem. 145, 127–140 (1925).
51. R. J. Acker MANN, “The High Temperature, High Vacuum Vaporization and
Thermodynamic Properties of Uranium Dioxide,” ANL–5482, Sept. 14, 1955.
52. R. J. Acker MANN, P. W. GILLEs, and R. J. THoRN, “High Temperature Ther
modynamic Properties of Uranium Dioxide,” J. Chem. Phys. 25, 1089–1097
(1956).
53. L. G. WISNYI and S. W. PIJANowski, “The Thermal Stability of Uranium
Dioxide,” KAPL–1702, Nov. 1, 1957.

54. T. C. EHLERT and J. L. MARGRAVE, “Melting Point and Spectral Emissivity


of Uranium Dioxide,” J. Am. Ceram. Soc. 41, 330 (1958).
55. T. T. CLAUDsoN, “Emissivity Data for Uranium Dioxide,” HW–55414.
Nov. 5, 1958.
56. R. J. ACKERMANN and R. J. Yielding Volatile Oxides at
THoRN, “Reactions
High Temperature,” XVI International Congress of Pure and Applied
Chemistry, Paris, 1957; ANL–5824.
57. R. J. ACKERMANN, R. J. THoRN, C. A. ALExANDER, and M. TeTENBAUM, “Free
Energies of Formation of Gaseous Molybdenum,
Tungsten, and Uranium
Trioxide,” presented at the American Chemical Society meeting, Apr. 1959.
58. F. P. KNUDseN, H. S. PARKER, and M. D. BURDIck, “Flexural Strength
of Specimens Prepared from Several Uranium Dioxide Powders; Its De
pendence on Porosity and Grain Size and the Influence of Additions of
Titania,” J. Am. Ceram. Soc. 43, 641–647 (1960).
59. J. R. Joh Nson, S. D. FULKERson, and A. J. TAYLOR, “Technology of Uranium
Dioxide, a Reactor Material,” Bull. Am. Ceram. Soc. 36, 12 (1957).
60. R. Scott, A. R. HALL, and J. WILLIAMs, “The Plastic Deformation of Uranium
Oxides above 800° C,” J. Nuclear Materials 1, 39–48 (1959).

61. F. P. KNUDseN, “The Dependence of the Mechanical Strength of Brittle Poly


crystalline Specimens on Porosity and Grain Size,” J. Am. Ceram. Soc. 42,
-376–387 (1959).
62. F. I. PETERs, Corning Glass Works, R-1068, 1956.
63. D. J. Bow ERs, W. A. HEDDEN, M. J. SNYDER, and W. H. DUckworth, “Effect
of Ceramic or Metal Additives in High UO, Bodies," BMI-1117, July 1956.
64. S. M. LANG, “Properties of High Temperature Ceramics and Cermets, Elas
ticity and Density at Room Temperatures,” U.S. Nat. Bur. Standards
Monograph 6, Mar. 1, 1960.

65. J. R. Joh Nson, “Ceramic Fuel Materials for Nuclear Reactors,” Nuclear Engi
neering and Science Congress, Preprint 110, Dec. 1955.

66. S. M. LANG, U.S. National Bureau of Standards, reported in “Properties of


UO2,” J. Belle and B. Lustman, “Fuel Elements Conference, Paris,” TII)—
7546, Mar. 1958, pp. 442–515.

67. A. J. DER RER, “Solid State Physics,” p. 120, Prentice-Hall, Inc., Englewood
Cliffs, N.J., 1957.
PHYSICAL PROPERTIES OF URANIUM DIOXIDE 227

A. K. SMALLEY, E. M. BARoody, E. M. SIMons, and W. H. DUckworth, “The


Thermal Fracture of Ceramic Cylinders,” BMI-1102, June 28, 1956.
. R. L. Coble and W. D. KINGERY, “Effect of Porosity on Thermal Stress
Fracture,” J. A. m. Ceram. Soc. 38, 33–37 (1955).
. M. LEBLANC and H. SACHSE, “The Electron Conductivity of Solid Oxides with
Different Valencies,” Z. Physik 32, 887–889 (1931).
. M. LEBLANC and H. SACHSE, “Electron Conductivity of Solid Oxides of Vary
ing Valences,” Ber. Verhandl. sachs. Akad. Wiss. Leipzig Math.-phys. Klasse
82, 153–158 (1930).
. W. MEYER, “The Electric Conductivity of Inorganic Materials with Electronic
Conduction,” Z. Physik 85,278–293 (1933).
W. HARTMANN, “Electrical Investigation of Oxide Semiconductors," Z. Physik
102,709–733 (1936).
. W. MEYER and H. NELDEL, “Relationships between the Energy Constants and
the Activation Energy in the Conductivity-Temperature Relationship in
Oxide Semiconductors,” Z. tech. Physik 18, 588–593 (1937).
. W. AMREIN, “Negative Resistances," Schweiz. Arch. angew. Wiss. Tech. 8,
85–89, 109–122 (1942).
. P. CHIoTTI, “Technological Research-Metallurgy,” CT-1985, 1945.
. J. PRIGENT, “Study of the Thermistors Made of Oxides of Uranium,” J. phys.
radium 10, 58–64 (1949).

. R. K. WILLARdson, J. W. Moody, and H. L. Goeri Ng, “The Electrical Proper


ties of Uranium Oxides,” J. Inorg. Nuclear Chem. 6, 19–33 (1958).
. W. MEYER, “Electron Conduction in Solid Chemical Compounds,” Z. Elek
trochem. 50, 274–290 (1944).
. M. J. BRABERs, “Electrical Conductivity of Uranium Oxides” in “Proceed
ings of the Second United Nations International Conference on the Peace
ful Uses of Atomic Energy, Geneva, 1958,” Vol. 6, pp. 122–123, United
Nations, Geneva, 1958.
81. J. K. D.Awson and M. W. LISTER, “Magnetochemistry of the Heaviest Ele
ments: I. Sensitive Magnetic Susceptibility Balance for Small Samples,”
J. Chem. Soc., 2177–2181 (1950).
2. A. ARRoTT and J. E. GoLDMAN, “Magnetic Analysis of the Uranium-Oxygen
System,” Phys. Rev. 108,948–953 (1957).
R. J. Acker MANN, R. J. THORN, and G. H. WIN SLow, “Visible and Ultraviolet
Absorption Properties of Uranium Dioxide Films,” J. Opt. Soc. A m. 49,
1107–1112 (1959).
D. A. WAUGHAN and R. K. WILLARDsoN, “Comparison of the Active with the
Inactive Uranium Dioxide-Oxygen System,” Nuclear Engineering and
Science Conference, Preprint 13, Mar. 1958.
. S. ARoxson and J. BELLE, “Nonstoichiometry in Uranium Dioxide,” J. Chem.
Phys. 29, 151–158 (1958).
86. S. ARonsox and J. C. CLAY!toN, “Thermodynamic Properties of Nonstoichio
metric Urania-Thoria Solid Solutions,” J. Chem. Phys. 32, 749–754 (1960).

57.4789 O-61-4-16
228 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

87. A. CoMPANION and G. H. WINSLow, “Diffuse Reflective Measurements on

Bulk Uranium Dioxide,” J. Opt. Soc. Am. 50, 1043–1045 (1960).


88. D. M. GRUEN, “Absorption Spectra and Electrical Conductivities of UO
Tho, Solid Solutions,” J. Am. Chem. Soc. 76, 2117–2120 (1954).
89, A. J. DEKKER, “Solid State Physics,” pp. 499–522, Prentice-Hall, Inc., Engle
wood Cliffs, N.J., 1957.
Chapter 6

PHASE RELATIONSHIPS IN THE URANIUM


OXYGEN AND BINARY OXIDE SYSTEMS
H. R. Hoekstra, Editor

6.1 INTRODUCTION

Uranium is the fourth member of the actinide series of elements.


In contradistinction to the lanthanide series, in which the trivalent
oxidation state is encountered almost exclusively, the actinide elements
show a much greater tendency towards further oxidation. Uranium
exhibits this characteristic most markedly, since the stable oxides of
the element are found in the UO2–UO, composition range. Stabili
zation of the higher oxidation states in the actinides can be attributed
to the lower binding energies and less effective shielding of the 5f
electron shell as compared with the 4f shell in the lanthanide series.
The uranium-oxygen system is one of the most complex of the metal
oxide systems. The existence of at least four thermodynamically
stable uranium oxides (UO, U.O., U.O.s, and UO,) has been defi
nitely proved, and several additional oxides have been reported.
The system is further complicated by polymorphism, solid solution,
and metastability phenomena.
Although the main emphasis throughout this chapter is placed on
the composition range UO, to UAOs, the remainder of the uranium
oxygen system is briefly summarized. A discussion is also presented
of phase relationships in binary oxide systems containing uranium
oxides.

6.2 BONDING IN URANIUM OXIDES

Perhaps one of the most fundamental problems to be encountered


in a consideration of the uranium oxides is the nature of the bond
between the uranium and oxygen atoms. Are the forces responsible
for the crystal energy principally ionic or covalent? On the basis
of general chemical evidence, the forces would appear to be largely
ionic in UO2. Thus, the estimated Pauling electronegativity for
229
230 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONs

uranium is 1.5, comparable with that of beryllium or aluminum [1].


In combination with oxygen (3.5) an electronegativity difference of
2.0 is obtained for the bond, sufficient to give a 90 to 95 percent ionic
bond.
Zachariasen, however, found a disparity of approximately 0.1
angstrom unit between the observed and calculated bond distances
in uranium and thorium dioxides when the ionic radii of the respec
tive ions were used [2]. The correct oxygen-metal distances are ob
tained when covalent radii are substituted. On this basis Zacharia
Sen concluded that the bonding in uranium and thorium dioxides is
mainly covalent.
The first attempt to arrive at an answer from theoretical calcu
lations was made by Childs [3]. By means of a conventional Born
Haber cycle, theoretical values were calculated for the heats of for
mation of uranium and thorium dioxides. Good agreement was
found between the theoretical and observed values on the assump
tion that the atomic binding forces in these solids are predominantly
ionic. Childs also pointed out that a comparison of observed bond
lengths with calculated ionic and covalent radii leads to anomalous
results in several other systems known to possess ionic bonding in
addition to the uranium and thorium dioxides cited by Zachariasen.
On the basis of the present information, it must be concluded that
the lattice forces in uranium dioxide are largely ionic in character.
The higher oxides of uranium, UAOs and UOs, show a much greater
tendency towards covalent bonding. The characteristic hexavalent
uranium species in aqueous solution is the divalent utranyl ion
(UO, *). Exchange experiments have proved the covalent nature of
the uranium-oxygen bonds [4]. Crystallographic studies confirm the
retention of uranyl bonding in the solid U" oxides as well [5].

6.3 URANIUM-UO, SYSTEM

The solubility of oxygen in uranium metal is low. At the melting


point of the metal (1,133° C), approximately 0.05 atom percent O
can be dissolved; at 2,000° C the solubility has increased to only 0.4
atom percent [6]. No compound formation or solid-solution region
between uranium metal and the dioxide has been conclusively proved.
The preparation of UO has been reported, but the compound has
never been prepared in quantities sufficient for isolation and analysis
[7, 8]. The monoxide has appeared only as a minor impurity in
preparations containing oxygen and uranium in ratios less than 2:1.
It has been suggested that pure UO is not a thermodynamically stable
compound and that it exists only as the compound U(O, C, N), with an
unknown fraction of carbon or nitrogen, or both, being required to
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 231

stabilize the structure [7]. Battelle investigators have prepared a


single phase with an analysis U(Cox.O.,s) [9]. Uranium monoxide
is reported to crystallize in the cubic NaCl-type structure with
a=4.92 Å, [7,8].
Extension of the UO, fluorite structure to compositions below the
stoichiometric 1:2 ratio has been reported on the basis of several
X-ray diffraction studies [10–12]. Anderson, et al., reported the pres
ence of substoichiometric uranium dioxide at high temperatures [12].

Uranium dioxide was fused in an argon arc furnace at 2,800°-E 100° C.


X-ray analysis of a fine chip of the fused product (stable in air)
revealed the presence of two phases: (1) An oxygen-deficient phase
(94 to 98 mole percent) face-centered cubic with an a, of 5.4718-1-0.0004
angstrom units; (2) a minor phase (2 to 6 mole percent) determined to
be 3 uranium metal. The oxygen-deficient phase was estimated to
have the composition UO, as to UO, or from an extrapolation of the

data of Ackermann for the cell dimensions of oxygen-excess uranium


dioxide [13]. (Note, however, that use of the Gronvold data on lattice
parameters indicates a composition more nearly 1.98 or 1.99.) Ander
son, et al., conclude that the small but significant expansion of the
uranium dioxide lattice (from 5.4700+0.0004 angstrom units to
5,4718-1-0.0004 angstrom units) indicates that uranium metal is slightly
is,

soluble in uranium dioxide, that region UO2-,


of
that there

at
is
a

high temperatures[12]. These temperatures may the range 1,600°


be
of in

2,500°C, the estimated quenching temperature the arc furnace.


to

up
by

Later work, however, Anderson and Sawyer showed that


to
at at

least 1,700° stoichiometric uranium dioxide hydrogen


in

stable
is
C

atmosphere [14].”
1

these observations, Wisnyi and Pijanowski concluded


In

contrast
to

from lattice constant measurements that uranium dioxide does not


during melting
or

lower oxide before reducing atmos


of to

in

reduce
a

a
an
or

or

hydrogen inert atmosphere argon [12,


of

phere helium
in

15]. In this work, the uranium dioxide, after being melted, was
an
in

ground oxygen-free atmosphere fine powder and mounted


to
a

small capillary tubes for X-ray analysis. The measured lattice


in

constants were 5.472+0.005 angstrom units (in hydrogen), 5.473

All crystal parameters given this chapter are angstrom units; literature values
in

in
*

reported kx units have been converted angstrom units (1k 1.00202 Å).
in

to

=
X

•work by (“Grain Growth Sintered Uranium Dioxide,” CRFI)—999,


R.

MacEwan
J.

in

Jan. 1961) suggests that UO, oxygen-deficient atmospheres high tempera


at

not stable
in
is

tures. He reported that uranium metal inclusions were found UO, that had been
in

argon at temperatures over 2,000° The shape and intergranular location


C.

annealed
in

the uranium metal inclusions suggest that they were precipitated when the temperature
of

exceeded the melting point One proposed explanation that UO, loses
of

uranium.
is

oxygen at high temperatures form UO2-x which, upon cooling, disproportionates


to

to

UO, and free uranium.


232 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

+0.005 angstrom units (in helium), and 5.473+0.005 angstrom units


(in argon), as compared with the lattice constant of the unmelted
oxide, 5.472+0.005 angstrom units.

6.4 UAOs—UO, SYSTEM

The phase relationships in the uranium-oxygen system become more


complex as uranium oxidation approaches its maximum value at
U(VI). Thus, four and possibly five phases of UAOs have been re
ported (these will be discussed in detail later), and five crystalline
modifications of UOs are known [16]. The trioxide may also be pre
pared in a form which is amorphous to X-rays (crystallites <100 Å.
in diameter). As a result of the multiplicity of UAOs and UO, crystal
modifications,the phase relationships in this composition range are
exceedingly complex, and only incomplete data are available at the
present time. Some of the properties of the UO, phases are sum
marized in Table 6.1. Uranium also forms a peroxide, which is stable
only as a hydrated compound, UO, .2H2O [17].

TABLE 6.1—SOME PROPERTIES OF CRYSTALLINE URANIUM


TRIOXIDES [16]

Decomposition
Phase Color Structure and dimension temperature
in air (°C)

o: Brown----------- Hexagonal; a =3.971 Å, c=4.168 A.--- 450


8 Red, orange- - - - - - Orthorhombic 7 c = 14.3 Å------------ 530
"Y Yellow--- - - - - - - - - (a = 13.01 Å, b=10.72 Å, c=7.51 Å) 7. 650
6+ a =4.15 Å--------------------
Reddish-brown----| Cubic; 400
e Red------------- *---------------------------------- 400

*Only known U(VI) oxide in which all O-U bonds are of equivalent length and strength.

6.5 UO,-U,Os SYSTEM

in the UO2—UAOs system reported by vari


The phase relationships
ous investigators are discussed in essentially chronological order in
this section. However, work dealing primarily with the UO2.s—UAOs
portion of the composition range is taken up separately in Sect.
6.5.2, and thermodynamic properties are discussed in Sect. 6.5.3.
In conclusion, an attempt is made to evaluate results and to resolve
discrepancies.
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 233

6.5.1 Phase Equilibria

The only attempt to study phase relationships in the UO,-U3Os


portion of the uranium-oxygen system prior to 1940 was the work
of Biltz and Mueller [18]. Their equilibrium oxygen dissociation
pressure measurements over uranium oxides indicated that the solid
solution region below UAOs extended to the composition UO2.62 at
1,160° C, followed by a two-phase region between UO.ca and
UO2.s. Pressure measurements below this composition were too low
to permit further conclusions. The equilibrium composition of UO.
as a function of temperature at 10-mm Hg oxygen pressure, as cal
culated by Biltz and Mueller, is shown in Fig. 6.1. Crystal structure
data revealed the presence of the fluorite UO, structure (see Chap. 5)
between the composition limits UO,-UO.s, and density measurements
showed no marked change in this region. A rapidly decreasing
density was observed at higher O/U ratios. Biltz and Mueller ex
plained the structure and density data by assuming that oxygen goes
into interstitial solution in uranium dioxide to the composition UO2.s.
Rundle, et al., reported a considerable solid-solution range for the
UO, fluorite structure on the basis of X-ray investigations [7]:
UO2-, (U+UO, at 2,000° C) 5.4610+0.0007 Å
UO, (Pure UO,) 5.4581--0.0005 Å
UO2, (UO,--U,0s—-UOz, 5.42.97+0.0008
A
a.)
§

400 6OO 8oo IOOO 1200 |400 16OO


TEMPERATURE,
c
*

FIGURE 6.1. Equilibrium UO. Composition


at
as

Temperature
of

Function
a

10-mm Hg Oxygen Pressure [18]. (Courtesy, Zeitschrift für anorganische


und allgemeine Chemie.)
234 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Although the results are reported in angstrom units, the numerical


values given suggest that the data were actually obtained in kx units.
Extrapolation of a straight-line plot of cell parameter as a function
of composition from UO., (at approximately UO, as through UO2
to UO2 x sets the lower limit at UO,

os.
be The anion lattice was assumed
perfect, with uranium being introduced into the vacant cubic
to

the uranium vacancies existing


of
interstices the fluorite structure

in
form UO., (U, O.). The disagreement between this interpreta
to

Biltz and Mueller's density measurements was

of
tion and the results
noted, but Rundle maintained that the decrease cell parameter with

in
by
increase O/U ratio was more readily explained vacant uranium
in
by

interstitial oxygen atoms. Rundle

of
sites than the introduction
U.O, lower

as
also reported extension the UAOs-type phase of

to

a
limit, disagreement with Biltz and Mueller's reported UO2.63.
in

Jolibois UO, heated slowly

in
as

reported that air oxidizes

in
is

it
two distinct steps” [19]. The first oxidation leads the formation

to
U,O., stable below 300° Above 300°C, U.O., The
of C.

produced.
of

is
X-ray diffraction pattern UAO, resembles UO, but contains addi
tional lines.

to
Gronvold and Haraldsen found that the low-temperature (120°
C) oxidation UO, proceeds the composition UO2.s. (3
of

to

150°
lower
In

phase) [20]. spite


of

of

some indications

to
conversion
a

symmetry, the lattice

(a
be
of

=
to
the structure was considered cubic
5.41 Å). The authors were unable not the

or
to

determine whether
parameter change during oxidation was continuous.

in
An increase
density during oxidation from 10.80 11.15 g/cc indicated that at
to

part

at
the oxygen enters interstitial positions.
of

least Oxidation
a

higher temperatures (200° 250° C) produced tetrag


= to

somewhat
a

onal phase UO, phase) with Den


Å.
at

and c=5.55
(8
so

5.38
Å
a

sity measurements this compound (10.00 g/cc) led them sug


on

to

gest partial substitution oxygen for uranium lattice, that


of

the
in
a

Observed and calculated X-ray intensities were


to is,

UossO2.12. found
be

good accord with the postulated structure.


in

of

Alberman and Anderson reported that the fluorite structure UOs


takes up oxygen temperatures below 230° the composition
at

to
C

U(), UO.) without change this


(8

in

cell dimensions [21]. Oxides


in
2

composition range disproportionate upon annealing UO, and UOss.


to

Oxidation above UO, results the gradual development tetrag


of
in

a
a

with c/a increasing with oxygen content UO, (see


to

onal lattice
a

Fig. 6.2). I)isproportionation annealing


on

this composition range


in

high temperatures gave UO, and UAOs the final products.


as
at

Cell dimensions given for the oxides were: UO, 5.468+0.002 angstrom
Kinetic measurements and suggested mechanisms UO2 are described
of

of

the oxidation
*

Chap.
8.
in
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 235

560 H.
H-T-T—
CUBIC
I-I-I TT I-I-I-I-I
TETRAGONAL -

|
H–1–1–1–1–1–1–1 || | | | | |
2.00 2.04 2.08 2.12 2.16 2.2O 2.24 2.28
VALUE OF n IN Uon

FIGURE 6.2. Cell Dimension as a Function of Composition According to Alber


man and Anderson [21]. (British Crown Copyright, reprinted with permis
sion of the Controller of Her Britannic Majesty's Stationery Office.)

units;UO2.2 (annealed), 5.441-E0.003 angstrom units; UO, s, a = 5.397


angstrom units, c=5.565 angstrom units.
The possibility of a substitutional solid solution for the nonstoi
chiometric oxides (as suggested by Gronvold and Haraldsen for
UO2.40) was ruled out by Alberman and Anderson on the basis of the
polar nature of the compound [20, 21]. No X-ray evidence was found
for the presence of a diphasic region in the UO,-UO, a composition
range, and the kinetic data suggested a reaction rate determined by
diffusion through a solid phase rather than one determined by an inter
face between two phases. Alberman and Anderson indicated that
the tetragonal cell given above may not be the true cell, but observed
no diffraction lines corresponding to a larger cell. The excess oxygen
atoms taken up during oxidation were assumed to occupy (400) sites
at random until the composition UO2.2 was attained and then to segre
gate preferentially into one set of positions to form the tetragonal
structure. The subtractive (cation) solid solution suggested by
Rundle, et al., was ruled out on the basis of density measurements and
previous data on the fluorite-type lattice [7]. I)efect lattices based on
the fluorite structure have been shown to have substantially perfect
cation lattices; the variation in cation-to-anion ratio is produced
either by incomplete occupation of anion lattice sites or by incorpora
tion of excess anions into interstitial positions [22]. A possible ex
planation for the changes observed during annealing of UO2.2 was
given by the equation
U0,1, A. U1-,0.

Alberman and Anderson conceded, however, that density measure


ments do not agree with this interpretation [21].
Boullé, et al., studied the low-temperature oxidation of UO, by
thermogravimetry and X-ray diffraction [23]. They noted the ap
pearance of a tetragonal phase early in the oxidation process, followed
236 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

by reversion to cubic symmetry at UO2.s. The formation of a second


tetragonal phase during oxidation to UO.s, occurred as the sample
was heated from 185° to 290° C. The appearance of the orthorhombic
UAO, phase was not observed until the empirical composition reached
UO, comparing the results Boullé,

In
so. al., and Jolibois with

of

et
remaining

be
the low-temperature studies, should remembered that

it
the thermogravimetric studies involve rapidly changing temperature
achieve equilibrium [19, 23].

no
as

as

composition, with attempt

to
well

by
The results reported Hering and Perio disagreed some re

in
spects with those Alberman and Anderson [11, 21]. was pointed
of

It
out that the starting material used the latter work had been UO2.0,

in
rather than stoichiometric UO2. Thus, the composition limits should
have been UO2.2, (cubic) and UO.s, (tetragonal). Hering and Perio
listed the following phases identified the UO,-U3Os region [11]:

in
Cubic extending from UO, 5.468 A) UO, (a5.427A).

to

(a
as
=

=
1.

Thus, the upper limit


of

the cubic structure (but not the cell dimen


agreement with the findings

of
sion) Gronvold and Haraldsen
in
is

[20]. Near the upper composition limit some indication change

of
a
lower symmetry was noted.

of
Evidence for extension the
to
a

fluorite structure UO, lower limit was claimed from exper


to

as
is

iments employing mixtures


of

uranium metal and the dioxide


600° C. The conclusion was based on identifica
at

to

heated 400°
fluorite structure with cell edge larger than stoichiometric
of

tion
a

UO, this observation was later stated


be
by to

error.”
in
;

UO,
A)

annealing phase

of
Cubic the
(a

obtained
as

=
2.

5.438
B

1
appropriate composition
C.

in
above 280° The additional oxygen
each unit cell was postulated occupy the position.
(;
to

})
;

c=5.55 Å), having composition


Å,

tetragonal phase,
(a
A

=
3.

5.38
an y

limits between UO, and UO, Temperatures


of
to.

500°
in

excess
were required induce disproportionation U.O., and U3Os.
to

to
C

Hering superstructure
of

and Perio found some indication


a

Å),
Å,

its
(a

to

c=11.1 but were unable confirm existence.


=

5.39
The c/a ratio decreased with increase temperature until,
in

at

500°
C,

the structure became pseudocubic with


Å.
=

5.41
a

Orthorhombic UAO-type structure whose lower composition limit


4.

UO,
50.

about
is

Somewhat later Perio reported further results the uranium


on

oxygen phase relationships (Fig. 6.3) [24]. X-ray data were obtained
with recording spectrometer. The cubic UO., phase was found to
a

UO, and U.O.. Compositions


be

to

metastable toward conversion


below UO, showed the following changes after annealing: (1) The
as

crystal parameter (in angstrom units) diminished according the


to

Perio, personal communication.


P.
*
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 237
looo I I I I I

— -H CUBIC & -
8 OO H. U02 + CuBIC 8 -

O U4 Os 4 ORTHO. U3 Os
*- 6 OO H. -
#
P
« |-
ORTHO.
U3 Oe -
ºr
u
al
# 4 oo H -
*- TETR, y +
U4 OS +
|- UO2 + U4 O9 TETR. y ORTHO. U3 Os

2 Oo H |
-

%95%
%gusſºft
|
|
|
- |
-
o | I L I | |
2.O. 2.1 2.2 2.3 2.4 2.5 2.6 2.7
MOLES O/ MOLE U

FIGURE 6.3. Revised UOz-U2O, Phase Diagram According to Perio [24]. (Re
printed with permission of Société Chimique de France.)

equation a = 5.468–0.12a, where a is given by UO,...; (2) the rate of


disproportionation to U.O., and UO, varied with annealing tempera
ture (20 hours at 700° C, 4 months at 140° C). The phase limits for
U.O., at 1,000° C were given as UO, 21–UO2 Its density, 11.3 g/cc,
as.

was stated structure containing interstitial oxygen.


to

indicate
at a

UO, temperatures
of

Oxidation below 175° resulted the


by in
C

appearance tetragonal phase (a) about UO, oxi


of

(UO2
at

or
to
is a
by

dimetry, UO2 weight). The c/a ratio (0.991) remained constant


during completion UO,
Å,

c=5.400 Å).
of

the oxidation
to

(a
as

5.447
lattice containing uranium
no its

Perio reported that density suggested


a

vacancies, but gave supporting data until later publication [25].


a

Annealing this phase for 500 hours produced


at

to

120° 170°
140°C, the structure had reverted
at

change, but
no

y to

2,000 hours
= in

cubic UO2., 5.425 Å). Above 180° transformation


(a

to

the
C

tetragonal phase occurred.


Two additional lines observed with the tetragonal phase suggested
y
Å,

that the axis doubled, giving The measured


Å.
=

10.74 c=5.54
is
a

density, 11.5 g/cc, does not agree with structure containing uranium
a

The previously reported approach the crystal parameter


of

vacancies.
ratio (c/a) unity increasing temperature
be

with
in
to

to

was stated
error: c/a remained virtually constant
C.
to

400°
cubic UO., phase towards dispropor
of

The metastability the


a

UO, U.O.,
to

tionation and together with the invariant lattice dimen


238 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

sions of the tetragonal phases (a, y), were taken as evidence that the
UO,-UAOs region below 400° C is made up of two-phase systems
rather than of solid-solution regions. The composition of the tetrago
nal phase with c/a = 1.031, prepared above 180°C, was stated to be
UO,...o instead of UAO, as reported by Jolibois and Boullé, et al. [19,
23].
Shortly after Perio's publication on this system Anderson reported
some additional work on phase relationships in the UO2–U,Os system
at low temperatures [26]. Analysis of X-ray line profiles showed
five phases (UO,
all

on
spectra could interpreted

be

of
that the basis
U.O., U.O.s, and two tetragonal structures). Anderson concluded
that low-temperature the composition

of
oxidation the dioxide

in
range UO,-U.O., results initially oxygen

of
statistical distribution

in
a
throughout the lattice, conserving the fluorite structure and perhaps
the parameter UO.
Annealing causes ordering
of

the lattice

to
in
give two phases, UO,
and U.O. Some indication was observed for
phase lower symmetry during the annealing proc
of

of

the formation
a

ess. Above the U.O., composition, the solid appeared

to
retain cubic
symmetry (with broad diffraction lines) when annealed below 150°C,
C,

two tetragonal phases could

y’
be
but above 180° identified:
(UO, sa), c/a+ 1.016; (~UO, ar), c/a+ 1.030. The densities were
y”

accord with the hypothesis interstitial oxygen (see


be

of
in
to

stated

T-
Fig. 6.4 and discussion below). Anderson suggested that these phases
12
T
I

I
I


Tuoz.x
~

.”
...

(OxYGEN ADDITION)
.

*

U-x 92+x
(OXYGEN SUBSTITUTION)

N.
U-x02
(URANIUM suBTRACTION)

O-O GRONVOLD
Dºulliſ] ANDERSON
Ay-----A SCHANER
1
|

|
l

|
8

2.O 2. 2.2 2.3 2.4 2.5 2.6 2.7 2.8


ATOM RATIO
O
U
/

Uranium Oxide Density as Composition.


of

FIGURE 6.4. Function


a
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 239

could be interpreted on the basis of an enlarged cell, with a’=2a [26]:


y’ U16Osa-H O=UO2.sia
y” UſeOas +20=UO, ars
Or y” U.Oas-H3O=UO,..,as.
No evidence was found for reversibility of the reaction y”—'y'-HUAOs.
Anderson, Roberts, and Harper investigated the oxidation of uranium
dioxide as a function of temperature and oxygen pressure [27]. The
final composition and the X-ray characteristics observed in this study
are important to this discussion. At 155° and 185° C in 40 to 50 cm
Hg pressure the final products were UO, ag-UO, as with the y” tetrag
onal structure. At lower pressures (0.01 to 1 cm Hg) the final prod
ucts (UO2.21–UO2.2%) were cubic. Well annealed U.O., was found to
oxidize extremely slowly below 200° C.
Perio subsequently reported the identification of three distinct tet
ragonal phases in the U.O,-U.O. region in addition to the low-tem
perature a modification [25]:
yi,

c/a-1.031 (UO,...,0), stability range 180°


to
400°C
(Anderson's y”)
ye, c/a+1.016 (UO, sa), stability range 420° 460°C
to

(Anderson's y’)
ya, c/a+1.010 (UO2 ass), stability range 460° C.
to

520°

No intermediate parameters were observed. The reactions

yº. ^2+U30,
20°

460°
^2.", Ya-HU30s

are obtained readily, but not the reverse. However, heating mixtures
UO2 and U.O, for several weeks U.O.,
of
at

gave
of

300° mixture
C

U.O.,
ya.
or
y,

and
Perio suggested that Anderson's interpretation the tetragonal
be of
on

of an

phases the assumption enlarged cell could


of

two
in

modified
U,
O,

ways, one based multiple [24, 26],


on
a

Y1 (U.O.) sO, =UOzars


Y2 (U.O.) sO3= UO2.sa,
als

Ya (U.O.) sO2=UO2
structure containing uranium vacancies:
on
or

second based
a

U.O. =UO.,
U1:Oas =UO2
as

U15Os, =UO2.21.
by

Admittedly, doubling the lattice parameters was not justified


of

X-ray results. Although the high densities annealed samples sug


of
240 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

gest interstitial oxygen, Perio reported that unannealed samples pre


pared by oxidation of the dioxide at 120° C gave relatively low
densities (10.5 to 10.7 g/cc), indicative of uranium vacancies [25].
On the basis of surface area and crystallographic data, Perio rein
terpreted the low-temperature oxidation kinetics of Anderson, et al.,
in terms of the formation of a new phase (tetragonal a) whose surface
advances through the crystal.
Gronvold reported a thorough study of the high-temperature phases
in the UO,-U,O, region [28]. Samples in this composition range were
prepared from mixtures of UO, and U.O.s, annealed for 7 days at
1,000° C in evacuated tubes, then cooled slowly to room temperature
over a 2-month period.
The following phases were identified:

1. UO. : no measurable composition range at 25°C, but extending


to UO, I, at 950° C (a–5.4704 Å at 20°C, d-10.793 g/cc). The
average linear thermal coefficient of expansion was reported to be
10.8×10−"/°C between 20° and 946° C (see Chap. 5).
2. U.O. narrow homogeneity range at elevated temperatures (a =
: a

5.4411 Å at20°C, d-11.159 g/cc). The lattice constant decreases


between 20° and 86° C, then increases linearly to 960° C with
&=11.6×10−"/°C. The presence of many weak additional diffrac
tion maxima led Gronvold to conclude that the uranium atoms are
shifted from their exact face-centered positions.
3. UAOs-s: a homogeneity range between UO, ea and UAOs at 25° C.
At 500° and 750° C, the lower limit is extended to UO2.ss.
C
Density measurements in the UO,-U3Os system have been cited by
Gronvold, Anderson, and others as proof that oxidation of UO.
occurs by the addition of interstitial oxygen rather than by
uranium subtraction from the fluorite lattice as suggested by
Rundle [7,26, 28]. Gronvold and Haraldsen, however, did suggest
partial substitution of oxygen for uranium in their UO2.40 phase,
as did Perio for his a tetragonal phase [20, 24].
The measured densities given by Gronvold (Fig. 6.4) were obtained
on samples prepared at high temperatures and point to U.O., as the
end point of the fluorite-type phase field. The data of Anderson, et
al., indicate extension of the oxygen addition process to UO2.ss by the
low-temperature oxidation process [27]. Density measurements have
not been reported between UAO, and UAOs on low-temperature oxides:
presumably they fall from a maximum at U.O. to join Gronvold's
U.O.-U.O., curve at or near U.Os. The anomalous low density (10.0
g/cc) reported by Gronvold and Haraldsen for their tetragonal
phase may be due to the fact that oxidation had progressed beyond
UAO, to UO, a [20].
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 241

Lattice constants of the single-phase UO., structure at different


O/U ratios were reported by Schaner [29]. Densities calculated from
these lattice parameter measurements (Fig. 6.4) are in fairly good
agreement with Gronvold's direct measurements but do not show a
straight-line relationship. This is probably due to the fact that
Schaner studied the variation of density with composition in UO,...,
while Gronvold's annealed samples were UO, +U,O, mixtures.
The phase diagram derived by Gronvold from a combination of
room-temperature and high-temperature X-ray studies with density
measurements over the entire composition range is given in Fig. 6.5
[28].
Arrott and Goldman carried out a study of magnetic susceptibility
in this region (UO,-U3Os) of the uranium-oxygen system (see
Chaps. 5 and 8) [30]. Susceptibility measurements on partially
oxidized UO, suggest incorporation of interstitial oxygen in the
fluorite-type lattice, but phase boundaries cannot be located with any
degree of precision.
Vaughan, Bridge, and Schwartz in their study of active and in
active UO, preparations arrived at the following conclusions: (1) The

I I I I I T I
iOOO.H.

|
-
900 H

800 H
992:
-
|
700 H. -
|
|
|
$’soo H —
u |
ar
->
3
| -
ar 500 H |
-x
U924xt U.9,
#
* 4oo H
|
949et U3Oe-x | -
300 H. | -

- --- -
200 H. || -
- ––––ll |

IOO }- || ||
|
-
UO2+ U307
| II* ot U39s-x
|
l l Lll ll l 1| | |
20 2.1 2.2 2.3 2.4 2.5 2.6 2.7
U/O ATOMRATIO

FIGURE 6.5. Phase Diagram for Uranium Oxides in the UO, Region According
to Gronvold [28]. (Reprinted with permission from F. Gronvold, “High
Temperature X-ray Study of Uranium Oxides in the UO,-U-O, Region,”
Pergamon Press, Inc., 1955.)
242 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

presence of moisture accelerates oxidation of UO, (see Chap. 8), in


agreement with the observation of Lang, et

al.
[31, 32]. (2) Active
UO, oxidizes room temperature with retention the cubic struc

of
at
ture, and the extent

of
oxidation variable. The maximum observed

is

an
was UO2.41. (Hoekstra and Siegel reported that active UO.

by
UO.C.O.,

at
prepared thermal decomposition

at of
300° closed

in
C

a
system will oxidize room temperature [33].) (3)

to
cubic UO2.s
tetragonal phase observed upon oxidation

of
A

inactive oxide above

is
This phase disappears UO2,

of
UO.or. (with the formation and
U.O.) upon annealing for the empirical com

at

if
450° 100 hours

C
position less than UO2.2s. The reaction irreversible. (4) The
is

is
upper limit the UO2, phase UO, room temperature.
of

Data

at
is

os
as

on cell parameter temperature were correlated with

of
function
a

conductivity measurements (see below and Chap.

5)
to
electrical
obtain the diagram shown Fig. 6.6.
in

Aronson, Roof, and Belle studied the kinetics UO, oxidation

at
of
[34]. Their X-ray diffraction results indicated that
to

350°
C

160°

5.52
|

S.
|

S
|

5.5|
I

N
25°C
eA
|

UO
+

200°
2
x

C c

SST
|

550 |A
*s
Ooo
4

\s,-4–1
600°C
v A
N


|

soo-c
|

-
..a

549
/z
|

*Hº-
|
ă

>

548
%

|
I

Lu
|
>

Tºri.
|

—l O
A.

S.
5.47
5

O
Hºrār

H
i

H.
|

Hº-Ax–
|
ZN
2

|U40s U307
t

-)
5.46

"les
|

5.45

"----
}

O
A-A—AHAHAH
5.44 -
|
|

5.43
l

2.O. 2. 2.2 2.3 2.4


I

OXYGEN-TO-URANIUM RATIO

Diagram the UO-U.O. Region According Vaughan.


in

to

FIGURE 66. Phase


et al. [31].
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 243
looo I i I
I i i

900 H. U499
-
GRONVOLD's
BOUNDARY

800 T (1000 •c
-
ANNEAL)

700 H UO2+x
| x-RAY DATA
V-(400°C ANNEAL)
600
$’
U409 + U3Os

; 4OO

300 II

200

IOO H.

l
2.3 2.4 2.5 2.6 2.7
OxYGEN/URANIUM RATIO

FIGURE 6.7. UOz-U.O. Phase Diagram According to Willardson, et al. [35].


(Reprinted with permission from R. K. Willardson, et al., “Electrical Prop
erties of Uranium Oxides,” Pergamon Press, Inc., 1958.)

at 230° and 275° C the cubic UO., structure persists, with very little
change in parameter, until the composition reaches UO2.1a–UO2.16.
Further oxidation produces a second (tetragonal) phase whose c/a
ratio gradually increases to 1.03. Their results were very similar to
those of Alberman and Anderson except that deviation from cubic
symmetry was found at UO, is instead of UO, as [21].
Willardson, Moody, and Goering studied the electrical properties of
the uranium-oxide system (see Chap. 5) [35]. Electrical conductivity
and thermoelectric power of oxide samples were measured at tem
peratures between 27° and 400° C. Samples were prepared by con
trolled oxidation of UO, at 180° C, followed by an anneal at 200°
or 400° C. Interpretation of conductivity data was aided by X-ray
diffraction techniques. Compounds identified in the study were
UO,.. as a metal-deficit semiconductor, U.O., and UAO,-, as metal
excess semiconductors. At 200° the upper limit to the UO., phase
C
was found to be UO,
on on.

oxygen over the


of

The further addition


composition range UO, UO, caused the appearance
to

of

second
2,

phase, U.O. Between UO, and UO2.s., tetragonal phase was pre
as

57.4789 O–61–17
244 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

dominant. Figure 6.7 depicts the phase diagram derived from the
data.
As mentioned in Sect. 6.3, Wisnyi and Pijanowski concluded from
lattice constant measurements made before and after melting UO,
that the melted material was UO, and not a lower oxide [15]. Ther
mal analysis experiments in vacuum showed that the fluorite structure
is the only stable phase of UO, up to 2,500° C.
Blackburn studied the high-temperature phase equilibria in the
UO,-U3Os region, using the Knudsen effusion method [36]. Equilib
rium oxygen pressures were determined for oxides of compositions
UO2.15 to UAOs at 950° to 1,150° C. Three phases were identified:

1. UO2, in which the oxygen solubility increases with temperature


from UO, is at 990° C to UO, as at 1,080° C.
2. U.O., with a small homogeneity range. Its upper stability limit
was estimated to be 1,167°C.
3. U.O.s (UO2.6), in which oxygen dissolves. The lower limit of this
phase remains effectively constant over the temperature range in
vestigated, in substantial agreement with the early results of Biltz
and Mueller [18].

The phase diagram derived by Blackburn from a combination of


his data with that of Gronvold is illustrated in Fig. 6.8 [28, 36]. The
lower limit of UAO, phase (UO2.ss) obtained by Gronvold at 500° and
750° C was omitted in favor of an invariant boundary between room
temperature and 1,150° C.
Blackburn, Weissbart, and Gulbransen reported that UO, oxidizes
at 200° to 300° C by the formation of a tetragonal UAO, phase (c/a
1.031) which moves through the oxide particles at a rate determined
by the diffusion of oxygen through the UAO, (see Chap. 8) [37]. Both
their X-ray and kinetic studies agree on this mechanism. The solu
bility of oxygen in the dioxide below 300° C is believed to be less than
1 atom percent per mole of UO. Lack of oxygen solubility is evi
denced by the constancy of the dioxide cell dimension throughout the
oxidation process. The authors conclude that failure to observe the
tetragonal pattern during the early stages of the oxidation (before the
composition reaches UO, on in pellet UO, and UO2.11 in powder UO2)
is due to a limitation of the X-ray method which necessitates the
formation of a layer approximately 1,000 angstrom units thick before
the tetragonal lines become visible. Oxidation studies at 150° and
200° C showed a variation in c/a ratio with composition, although the
cell volume remained constant. This variation in c/a ratio is attrib
uted to the incomplete ordering of the structure at the lower tempera
tures rather than to the existence of more than one tetragonal phase.
Oxides in the composition range UO2–U.O., anneal readily to U.O.
and UO, at 300° C, presumably because residual unreacted dioxide
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 245

i2OO I i I I I I I

*
§
IOOO
O
-
D
U409+x |
* |
800 | –
|
U409+y
+
| –
U-O |
5*13
soo
#. 3. g ...

-
|
-

|
ar

|
;

|
400
H

-
|
|
A.

|
BLACKBURN
O

A -
|
2OO
GRONVOLD
O

|
|


|
|b
1
O

|
|

2.4 2.6
U/O ATOMRATIO

Phase Diagram the UO,-U.O, Region According


to

FIGURE 6.8. Blackburn


in

[36]. (Reprinted with permission from the American Chemical Society.)

the U.O, phase. Disproportionation


of
of
in

assists nucleation
by

U.O. U.O., and UAO, annealing


to

elevated temperatures slow


at

is

the U.O., phase.


of

because no dioxide remains


to

assist nucleation
All attempts prepare UAO, from mixtures U,O, and UAOs
of
to

at

temperatures between 165° and 475° were unsuccessful.


C

Aronson and Belle used another method study the uranium-oxy


to

gen system the UO2–UO2.2 composition range [38]. Some work


in

higher oxygen content. Elec


of

was also done with uranium oxides


tromotive force measurements were made on uranium-oxide half cells
the type Fe, Fe0 (ZrO2, CaO)|UO,...,
Pt

the temperature range


of

in

the emf curves occurred only


C.
to

at

*75° 1,075° break the


in
A

C,

composition UO2.20 temperature indicating phase


of
at

940°
a

transformation. Similar emf values were obtained UO,


an

and
at

TO, indicating the presence


so,

two-phase region.
of

Partial molar
a

free energies, entropies, and enthalpies oxygen UO2.


of

of

solution
in

were calculated from the emf data. Oxygen pressures calculated


from the free-energy data were shown
to

good agreement with


be
in

Blackburn's measurements [36].


246 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Aronson and Belle discuss the physical picture of the UO2–U.O.,


region from the standpoint of a simple addition of oxygen in inter
stitial sites of the fluorite lattice [38]. They conclude that available
data do not agree with the assumption that U.O., is formed by the
addition of an oxygen ion into the center of each UO, cell, but rather
that the oxygen lattice in U, O, is complex.
Siegel has reported that a study of U.O., structure with single
crystal X-ray patterns and powder diffraction tracings with neutrons
indicates a tetragonal cell with axes which are multiples of the pseudo
cubic axis (a = 5.440 Å) and the face diagonal [39]. Rotation pat
terns required that the multiplicity be four or more. (See Sect. 6.5.4
(a) for additional information on the structure of U.O.)
Ackermann and Thorn measured equilibrium dissociation pressures
of oxygen over the compositions U.O., and UO, to by a static method
[40]. Their U, O...y-UO, a curve is in good agreement with Black
burn's results (see Fig. 6.9) [36]. They pointed out, however, a rather
large discrepancy between their equilibrium pressures over U.O., and
the values reported by Blackburn. The effusion data for U.O., pro
duce a curve that intersects the UO., U.O., , system at about 1,080°C,
it,

an
and even extends beyond which impossible situation. The
is
by

pressures non

to
anomalous obtained Blackburn were attributed
equilibrium conditions the effusion experiments.
in

TEMPERATURE, *C
OO |2OO |OO OOO 900
-
3

|
|
|

I
I

I
I

–3 - -
*
–4 H -

SS
E~RS
D

–5 H-
BLACKBURN
C A
B

U409+y
-
C. B. A.

U02.6
- -

–6 H UO2+x U409-y
U4O9

-
F

ACKERMANN AND THORN


~7

U409+y UO2.6
E. D.
F.

U409
H

ARONSON AND BELLE


-8
UO2.198
1. H. G. F.

U02.18
–9 H. UO2.142
U02.112
UO2.o.
J.

-1
l

-IO
|
|
l

6.5 7.O 7.5 8.0 8.5

10°/T ("K)

FIGURE 6.9. Equilibrium Oxygen Dissociation Pressures over Uranium Oxides.


URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 247

Roberts and his associates reported additional high-temperature


equilibrium dissociation pressure data in the composition range
UO,-UO, at 1,100° to 1,500° C by a static method [41].
, obtained
Isotherms at temperatures between 1,077° and 1,450° C show a slight
discontinuity at UO, 23–UO, 2. In agreement with Blackburn's re
sults, however, they found no certain evidence for the existence of the
U.O., phase at these high temperatures. The UO., phase extends to
UO, as and to UO, as at 1,450° C.
at 1,077° C
In later work Roberts and Walter “ determined equilibrium oxygen
pressures directly by tensinetric measurements at 1,000°C to 1,450°
C over the composition range UO, ºn to UO,

an.

In
common with
investigators, they found the following: (1) two regions

in
other
which the equilibrium pressure any temperature does not vary with
at

composition and which two solid phases must coexist; these are the
in

UO...-U.O., region below 1,123° UO, region

at
and the ex-UO2.6
in C

temperatures; (2) two regions


all

which the pressure varies con


tinuously with compositions and which only one solid phase can
in

present; these are the UO., region (03:30.26 above 1,123° C)


be

and the U.O., region.


The phase diagram constructed from the equilibrium oxygen meas
shown, together with the phase
of

urements Roberts and Walter


is

Blackburn, Fig. 6.10 [36]. Note that agree


of

boundaries
in

ment with Blackburn's results good except for two points the
in
is

UO...—U,O,-, region. Roberts and Walter disagree with Blackburn,


the U.O., phase.
on

In

however, the temperature range


of

the limits
1,450°C, Roberts and Walter find that the U.O., phase does
to

1,000°
not extend richer oxygen than UO2.25. Of special
compositions
in
to

interest 1,123°--5°C,
the temperature which Roberts and Walter
is

at

invariant point with the three phases, UO, 2., U.O. 250, and
in an

show
UO, equilibrium.
e,

by

The equilibrium dissociation pressures obtained Blackburn and


by

Thorn, the compositions investigated


of

Ackermann and and some


Aronson and Belle have been combined Fig. 6.9 [36, 38, 40]. From
in

the representative results given here apparent that the data (with
is
it

the one exception noted earlier) are substantially agreement; only


in

minor variations equilibrium pressures and slopes were obtained


in
in

from the widely different approaches employed.


the UO,-U3Os system part dis
on

of

Helle summarized data


as
a

the properties
of

of

cussion uranium dioxide given the 1958 Geneva


at

Atomic Energy [42].


on

of

Conference the Peaceful Uses


by

Still another technique delineate phase boundaries was used


to

study UO,-U,O, portion the uranium-oxygen


of

of

Schaner
in

the
a

Walter, “Equilibrium Pressures and Phase Relations


L.

E. Roberts and A.
J.

in
J.
*

the Uranium Oxide System,” report course of publication, AERE, Harwell.


in
248 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
14OO I I-1 I I TI () I I

LEGEND

A AND O – LIMIT OF UO2 ºx PHASE


oAND C – LIMITS OF U20s-y PHASE
13OO H OPEN SYMBOLS – ROBERTS -
AND WALTER
Ö
FULL SYMBOLS – BLACKBURN
O
o

#
> 12OO H.
H.
<I
or
LL
0.
>
Lu
H.

|IOO H.

|OOO -A
A

2,2O 2.22 2.24 2.26 2.28


COMPOSITION, O/U

FIGURE 6.10. Portion of the UO,-Oxygen Phase Diagram. (L. E. J. Roberts


and A. J. Walter, “Equilibrium Pressures and Phase Relations in the Uranium
Oxide System,” report in course of publication, AERE, Harwell.)

system [29]. In this work, limits of homogeneity and phases present


were determined by metallographic examinations
of the microstruc
ture of dense, sintered uranium oxides having various oxygen contents
(see Appendix B). Thin, solid platelets of large-grain-size (average
of about 60 microns), high-density (98.7 percent of theoretical) UO.
were used as the starting material to facilitate microstructure interpre
tation. To obtain the range of composition(O/U ratio), the UO.
platelets were oxidized in steam at various temperatures and then an
nealed in helium-filled Wycor or quartz bulbs at sufficiently high tem
perature to ensure the presence of only the homogeneous single-phase
UO... The temperature at which precipitation of the second phase
(U.O.) occurred for each composition was determined from metal.
lographic examination of quenched samples. Photomicrographs o
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 249

-
-
- . .. . . *. . .

...
. .. --

- . . - ~~ T-;

- -
.

.
.

.
.
.
...

- •"

-
"...’ UO, oo, 98.7 Percent of Theoret

A.
...

.
.
-
>
.
.

ical Density; Etched; X400, Re

."
.
.
.
.

duction Factor, $4.

...
..
-.
.

- ."

-
-
.

- ...

-
...


..
.

-
-
º

C,
at
UO,..., Annealed
B.

900°
.

.
.

C,
`... Slowly Cooled and

to
360°
.

º,
:

. .
.
.

-
.
.

Quenched; Etched; X400, Re


...


*

"
.
.
* .

- duction Factor, 14.


.
.
•. .
".

.
.. .

•.
.
.

**
.
.
.

-
-

--
.

-- -
-
-
--
-
-

..".
-

:
.
*
.

/
.
.
.

-- *- - - -
.
t

-
-

1: 'i'..…
". •.
v.

.. ".
...
‘.

.
.
‘.

-
.

&:
'
"
:
.
.

--
.*
.
.

…,
“.
-

•.

C. UOz.o.o. Annealed C.
at
.”

900°
2.
.
.
.

.
.
.
.
.
.

Slowly Cooled
C,

and
to
".

350°
'..

-
*
a

".
- ..

....
-.
..

Quenched; Etched; X400, Re


-
...

.
.

duction Factor, 1%.


*
.
..

.

.
.
.

.
.

Photomicrographs UO., and UO,--U.O, Microstructures


of

Fictºr. 6.11. [29].


250 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

typical UO., single-phase and UO, +U,O, two-phase microstructures


are shown in Fig. 6.11, A, B, and C. Figure 6.11, B, illustrates a sam
ple of UO,

oo,
quenched from 360° after annealing for

at
900°

C
C
(Fig.
48
hours and displays structure identical with sintered UO.000

C,
A). Figure 6.11, sample the same composition

of
6.11, but

a
C,
quenched from 350° shows precipitation

of
acicular and massive
U.O. The phase diagram determined from these observations
com

is
pared with the Gronvold and other phase diagrams Fig. 6.12. The

in
following features can noted: (1) Two phases, UO2 and U.Oe-y, are

be
present room temperature between O/U 2.00 and 2.22; (2)

of
at

ratios
only
at

temperatures over 940° single phase exists; (3) broad

a
range UO., present between O/U ratios
of

of
tem

at
2.00 and 2.194
is

at of
peratures between 200° and 950° C.; and (4) range substoichio

a
U,O, (U.O.,) O/U ratio
an

of
metric extends to 2.20 940°C and
room temperature.
to

at

2.22
These metallographic results are good agreement with earlier in

emf measurements which indicated phase boundary between UO.


a
gos

and U.O., UO, [38]. The results, however, disagree


at

at

950°
C
,

part with Blackburn's effusion cell measurements, the tensinetric


in

Roberts and Walter, and with Gronvold's high-tem


of

measurements
perature X-ray study [28, 36].
the single-phase UO2.
on

Lattice parameter measurements made


quasilinear decrease with increase oxygen con
in

structure indicate
a

tent (see Chap. 5). from the lattice parameter


Densities calculated
(Fig. density with O/U ratio
an

measurements 6.4) show increase


in

general agreement with the data


of

Gronvold [28].
in

Hoekstra, Santora, and Siegel have reported investigation


an

of

the
during UO, tempera
as

phases formed
of

of

the oxidation function


a

oxide particle size [43]. X-ray


of

ture and measurements were made


with GE recording spectrometer. The authors recommend elimina
a

is,

commonly used for this system, that


of

tion the nomenclature


phase, U.O. phase, U.O. =y phase, since each
of

UO2=a these
=
3

as

phases represents composition different


as

different well
a
a

Structure.
Two tetragonal modifications UAO, were identified: «UAO, with
of

A,

The exist
Å,

U.O. with
Å.
=

a=5.46 c=5.40 and 5.38 c=5.55


Å

a
uRANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 251

H-H++++++++++++++++++++
I2OO O O O

iOO

IOOO

900

Boo

7oo

SCHANER
O UO2+x
© UO2 +x + U4O9-y

GRONVOLD

BLACKBURN
ARONSONAND
BELLE
PHASE BOUNDARY
BETWEEN UO2+x
-
AND U4O9–Y
|UO2+x+U40s-y
UO2+x
WILLARDSON,
MOODYAND
GOERING

U02 +U40s-y
O
I O O O
1 | | | Ll1
OO
| | L. Lºº ºl,
2.05 2.IO 2.15 2.2O 2.25
O/U ATOM RATIO

FIGURE 6.12. UOz—U.O., Phase Diagram.

ence of a UAO, has previously been reported only by Perio, while 8


U.O. is identical with the structure commonly referred to as the y
phase [25]. Little solid solution exists in the UO,-UAO, composition
range, since a U.O., the first new phase to appear during oxidation,
was identified in samples of low surface area at the empirical compo
sition UO2.04. No significant change in lattice parameter was observed
-O
O

v
--
Bddw BONvb 3.SvHd
252 URANIUM DIOXIDE:
PROPERTIES

60

zonworzon zon90
AND NUCLEAR

~
~
2~
•99•99•99
2
8

--
APPLICATIONs

»
Jo

Jo

ºranoiſ "EI"9 īņwouÐ ſoºn ºuļump pļxo uoņu '[&#]*On


URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 253

I l —l l l
2.O. 2. I 2.2 2.3 2.4 2.5 2.6
O/U RATIO

FIGURE 6.14. Observed Parameter Limits during Oxidation of UO, to U20, [43].

over the UO2.0,-UO2 as composition range. Figure 6.13 illustrates the


growth of
the a U.O. absorption peak at 20=56 degrees during the
oxidation process.
Pure a U.O. may be obtained by oxidation of small-particle-size
UO, at low temperatures (100° to 175° C). At higher temperatures
conversion of a to 8 begins before oxidation is complete. The upper
limit of the tetragonal phases is near U.O., since the orthorhombic
U.O.-type structure appears in samples oxidized above UO, Since
ss.

«U.O, and UAOs have never been observed together, assumed that
is
it

oxidation occurs via the sequence


UO2–-a UAO,
!

UAO,->U3Os.
B
of

Measured densities and BUAO, were identical and were higher than
a

UO, these increased densities indicating oxygen ad


of

that the parent


to

dition the fluorite structure. The observed lattice parameter limits


for UO., U.O., and U.O. are shown Fig. 6.14. The constant
in
8
a

the three phases over the composition range given


of

cell dimensions
254 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

are in disagreement with Alberman and Anderson and Aronson, et

al.
[21, 34].
Annealing experiments with U.O. between 300° and 600°

C
8

C
favor the stepwise disproportionation reported by Perio, but place the
limits between UO,
composition and UO, rather than between

as

as
UO2.0 and UO, [25]. Two tetragonal phases were identified:

as
At

C.
c/a+1.016 prepared and c/a+ 1.012 prepared

at

at
400° 500°

C
and above, disproportionation U.O., and UAOs complete.

to
600°

is
C

Attempts find evidence for reversibility the disproportionation

of
to

an
lower temperatures were unsuccessful, even after
at

reaction anneal
of weeks at 350° C.
6

6.5.2 Phase Relationships and Structures: UO,

to
U20s

a
As was indicated this chapter, Biltz and Mueller found
earlier
in

that solid-solution region extends below UAO,


1,160°C the com
at

to
a

position UO2.62 [18]. Rundle, al., reported extension

of
the U3Os
et

type phase U.O, lower limit [7]. The U.O., structure was listed
to

= as
a

Å,

Å,
as

32
orthorhombic c=8.27A) with
of an (a

b=31.65 atoms

U
6.72
-
per cell give X-ray density g/cc.
of
to

8.25

by
The structure U,O, (Fig. 6.15) was first determined Zacharia
6.71+0.01 Å, b=
an

sen [44]. He found orthorhombic cell with


=
a
Å,

Å.

11.96+0.01 c=4.15+0.01 The unit cell contains two molecules


UI

UAO, with
of

(4:0).
at

at

(000) (#40) and Um (040)


2

hex
be

Zachariasen considers that the UAOs structure can derived from


by

agonal UO, referring the trioxide orthohexagonal axes and


to
a

an

choosing axis three times longer than required. The oxygen


004 and |\} (broken-line circles) are missing U.O.s.
in
at

an

short time later Gronvold reported independent structure de


A

of

termination UAOs [45]. The cell was identified orthorhombic


b as
Å,

is,
Å,

Å,

with b=3.977 and c=4.144 (Zachariasen)


that
=

6.716
a

=3b (Gronvold). The measured density (8.34 g/cc) led Gronvold


suggest ReO, but with deformed (UOc) oc
to

in to

structure related
*-
a

tahedra with anion vacancies substantial agreement with Fig. 6.15.

Gronvold and Haraldsen reported the composition limits


of

the U.O.
phase UO, so-UO, [20].
be
to

or
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 255

@-O-O-º

9 @ 9 @

§ URANIUM
O oxygen

FIGURE 6.15. Structure of a UOs and U2O, According to Zachariasen; Broken


line Circles Represent Oxygen Atoms Removed from Structure in Converting
a UO, to U2O, [44].

Hering and Perio found a lower composition limit of UO2.sº for


the orthorhombic phase [11]. Dimensions were given for UAOs
(a=6.721 Å, b=3.988 Á, c=4.149 Å) and UO, ss (a =6.76 Å, b=3.96 Å,
c=4.15 Å) but not for the lowest single-phase preparation in this
region, UO2.82.
Hoekstra, et al., investigated the region between UO.s and U.O., by
using room-temperature X-ray studies of quenched samples as well
as high-temperature X-ray techniques [46]. The Rundle U.O. phase,
256 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

related structurally to UAOs, was observed only over the composition


limits UO2 as to UO, Little change

ex.
cell parameters with com

in

to Å,
position was noted. The parameters given were a=6.751 b=31.76
UO,
Å,
phase the for

Å.

of
Oxidation

at
c=8.286 the led

C
500°

a
UAOs (8) having the cell constants
Å, of

of
mation new modification

Å,
The conventional form was formed

Å.
a=7.05 b=11.42 c=8.29

a
by

High-temperature X-ray studies did not

C.
reheating
to
750°

C,
an

locate a-to-3 transformation temperature below 1,000° thus,


suggesting that U.O.,

of
Conversion
metastable modification.

is
8

a
trigonal oxide (probably slightly oxygen-deficient rela
to

UAOs
a

C.
to

at
tive UAOs) was observed 400°
X-ray
In

of
his high-temperature study the UO2—U.Os region,
Gronvold reported homogeneity range between UO, and UAO,

as
a

C; UO, [28].
at

at
25° between and UAOs and 750°

C
as

500°
Crystal parameters measured for the orthorhombic phase were:

Å;
UO,
Å, Å,

Å, Å,

density 8.408 g/cc


= =

ess: 6.735 b=3.966 c=4.144


a

UO, ear: 6.720 b=3.983 c=4.146 Å; density 8.378 g/cc.


a

Blackburn, however, found substantially


an
invariant boundary
UO, between room temperature and 1,150° (Fig. 6.8) [36].
at

C
a

neutron diffraction diagrams U.O.s, Andresen


of

of

On the basis
a

concludes that the structure previously assumed X-ray


on

of
the basis
(Fig. 6.15)
be

data must error [47]. He derives modified unit


in

(Fig. 6.16) the uranium atoms are surrounded by


of
in

cell which two


U-2On
of

six oxygen atoms


of

at

distorted octahedron distances


a

U1–4Olve
A.

The other four uranium atoms are


=

2.07 and 2.18


Å

by

seven oxygen Un-2On-2.07


at

Å.
surrounded atoms distances
Un-O-2.21 Un-2Oly-2.42 The sym
Å,

Å,

Un-2Ov-2.17
Å.

and
group
D.

metry with orthorhombic space


-

conforms the C322.


Whereas Zachariasen's model predicted U" and U" ions per
2
1

molecule, the Andresen model favors U" and U" ions, agree
in
2

the magnetic
of

ment with Haraldsen and Bakken's measurements


Application Zachariasen's relationship between
of

moment [48].
bond length and bond strength, however, results equivalent val
in

(5.33) for all uranium ions the compound [5].


in

ences

Berman reported the preparation U.O, by high


of

of

cubic form
a

pressure techniques [49]. More recently, Wilson produced


a
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 257

§ URANIUM
O OxYGEN

FIGURE 6.16. Structure of U.O, According to Andresen [47]. (Courtesy, Acta


Crystallographica.)

high-pressure,high-temperature phase of UAOs, but the crystal struc


ture of the new UAO, phase was not determined [50]. Preliminary
attempts were made to index the observed X-ray diffraction reflections
without success. These data are insufficient to rule out the possible
extension of the cubic UO., phase to this composition (UO2 gr).

6.5.3 Thermodynamic Properties

Until recently, very little in the way of reliable thermodynamic data


for the UO,-U3Os system has been available. Mixter determined
258 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

heats of formation of UO, U.O.s, and UO, by two methods (combus


tion with molecular oxygen and reaction with sodium peroxide); the
results were in rather poor agreement [51]. Biltz and Mueller made
estimates of the heats of formation of UAOs and UOs from UO, on the
basis of their equilibrium pressure data [18]. Biltz and Fendius
calculated heats of formation of UO, from heat-of-solution measure
ments [52]. Their results, together with Brewer’s “preferred” values
calculated from these publications, are summarized in Table 6.2 [53].
Recent interest in uranium chemistry has resulted in considerable
progress in the field of uranium-oxide thermodynamics. Jones,
Gordon, and Long measured the heat capacity of uranium metal, UOs,
and UO, between 15° and 300° K
[54]. The calculated entropies at
298.16° K were 12.03, 23.57, and 18.63 calories-degree-'-mole", respec
tively. Huber, Holley, and Meierkord determined heats of formation
for UO, and UAOs [55]. Using these data, Brewer arrived at the
thermochemical properties given in Table 6.3 [56].
Coughlin calculated AH and AF values for the formation of UOs,
U3Os, and UO, between 298° K and 1,500° K (Table 6.4), while Glass
ner listed constants for the evaluation of the thermodynamic and
thermochemical properties of the uranium oxides up to 2,500° K
[57, 58]. Wagner reviewed the thermodynamic data for the uranium
oxides, pointed out discrepancies in the calculated data, and sug
gested several experimental techniques to resolve the observed
discrepancies [59].

TABLE 6.2–HEAT OF FORMATION OF URANIUM OXIDES [53]

—AH in kcal/gram-atom uranium

UO2 UO2.75 UO2.87 UOz.s" UO;

Mixter (O2) -- - - - - - - - - - - - - - 256.6 |- - - - - - - - - - - - - - - - 281. 7 287


Mixter (Na2O2) - - - - - - - - - - - 269, 7 - - - - - - - - - - - - - - - - 298. 5 304
Biltz and Fendius---------|--------|--------|--------|------------ 292
Biltz and Mueller - - - - - - - - - X -- ------ --- ---- - X-H 26.3 X-H 32
Brewer------------------- 270 282 297 298. 5 305
LRANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 259

TABLE 6.3—THERMOCHEMICAL PROPERTIES OF THE URANIUM


OXIDES [56]
—AHzes —AFrºs —ASzws
Compound (kcal per gram- | (kcal per gram- (cal per gram
atom uranium) | atom uranium) atom uranium
per "C)

"0,------------------------------ 259. 2 246. 6 43.4


"0.1----------------------------- 270. 0 256. 6 (47)
"Ola----------------------------- 285. 6 269. 9 (55)
"Oº----------------------------- 284. 5 268 55. 4
*UO,---------------------------- 291 273 62. 1

TABLE 6.4—HEAT AND FREE ENERGY OF FORMATION OF URANIUM


OXIDES [57]
|
UO2 UO1.87 UOs
Temp (* K)

—AH(+0.6) —AF (+0.7)" | – AH (+0.5) —AF(+1.2)* —AH (+5)* —AF(+5)*

298. 16 259.2 246.6 284.5 268.0 291.0 272.5


400 258.9 242.3 284.2 262.4 290 ..5 266.0
500 258.6 238.2 283.9 257.0 290.5 260.0
600 258.3 234.1 283.6 251 ..6 290.0 254.0
700 258.0 230. 1 283.3 246.3 290.0 248.0
800 257.8 226. 1 283. 1 241.0 289.5 242.5
900 257.6 222. 1 282.9 235.8 289.5 236.5
935 357.6 220.7 282.9 234.0 289.5 234.5
935 258.3 220.7 283.6 234.0 290.0 234.5
1000 258.2 218. 1 283.4 230.5 290.0 230.5
1045 258. 1 216.3 283.4 228. 1 290.0 228.0
1045 259.2 216.3 284.5 228. 1 291.0 228.0
1100 259.1 214. 1 284.3 225.2 291.0 224.5
1200 258. 8 210. 0 284.0 219.8 290.5 218.5
1300 258.4 206.0 283.7 214.5 290. 5 212.5
1400 258.1 201.9 283.3 209.2 290.0 206. 5
1405 258. 1 201.7 283.3 208.9 |----------|----------
1405 261.3 201.7 286.5 208.9 |- - - - - - - - - - - - - - - - - - - -
1500 261 - 1 197.7 286.3 203.7 ----------|----------
*AH and AF in kcal per gram-atom uranium.

Blackburn, in his effusion studies of the UO,-U3Os system, was able


to calculate partial molar free energies for the solution of oxygen in
oxides over the composition range UO, as to UAOs (Table 6.5) [36].

57.4789 0–61–18
260 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONs

He assumed that the curves of log p versus composition between UO.


and UO2.15 could be represented by

log p=a+ba.

where a and b are temperature-dependent constants and a is given by


UO2.x. An expression for the partial molar free energy of solution
was derived from the dissociation-pressure curves by assuming that
the partial molar heat of solution is constant and that the entropy of
solution is a linear function of the oxygen concentration.
Aronson and Belle obtained the free-energy change of a uranium
oxide cell by the relationship AF= —nº E and were able to obtain
partial molar free energies of solution of oxygen in the solid oxide,
both as a function of temperature and as a function of composition

*
[38]. The partial molar entropy of oxygen solution was calculated
from =–AS, while the partial molar heat of solution was obtained

from the expression AH., AF,4-TAS.,. The cell emf and, conse
quently, AF., were found to be linear functions of temperature in the
UO,... region, while the AS, and AH., were temperature-independent.
The calculated free energies, entropies, and heat contents are given in
Table 6.6.The partial molar entropy increases negatively with in
crease in oxygen content, whereas the partial molar heat content varies
relatively little.

TABLE 6.5—THERMODYNAMIC VALUES FOR THE FORMATION OF


URANIUM OXIDES AT 1,300° K

Blackburn [36] Coughlin [57]


Compound

– AH" –AF" —AS** —AH" —AF* —as”

UO,-----------------|--------|--------|------- 258. 4 206 40. 3


UO2.25--------------- 266. 9 211. 5 42.6 |--------|--|--|--|--|--|--|--|--
UO2.81--------------- 280, 8 216. 6 49.5 |--------|--|--|--|--|--------
UO2 aſ --------------- 282.8 f216, 8 50. 3 283. 7 214. 5 53. 3
UOa----------------- 290. 1 213. 0 59. 3 290. 5 212. 5 60 0

* kcal/g-atom U.
** cal/g-atom U/* C.
* From an assessmentof various thermal data, Ackermann, et al., calculated a value for the free energy
of formation of UiO, at 1,300° K which agreed with that of Blackburn to within approximately
0.5 kcal/mole (60].
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 261

TABLE 6.6—PARTIAL MOLAR FREE ENERGIES, ENTROPIES, AND


ENTHALPIES OF SOLUTION OF OXYGEN IN UO, e, [38]
—F., (kcal/mole) —S TT
2 °,
0U Atomic
ratio *.
(cal/mole/* C) (kcal/mole)
1,150° K 1,250° K 1,350° K

2. 013 54. 8 54. 2 53. 6 6. 3 62


2. 018 53. 8 53. 2 53. 6 6.4 61
2. 044 52. 4 51.2 49. 9 12. 8 67
2. 044 51. 8 51. 1 50. 1 6. 8 60
2. 051 51. 5 50. 3 49. 1 12. 2 66
2. 0.75 50. 9 49. 5 48. 1 13. 6 66
2. 078 50. 7 49.4 48. 1 12. 7 65
2. 087 48. 9 47. 3 45. 6 16. 7 67
2, 112 47. 6 45. 8 44. 0 18. 3 69
2. 128 46. 3 44. 0 4.1. 8 22. 5 72
2. 142 44. 6 42. 1 39. 6 25. 3 74
2. 143 44. 4 42. 0 39. 6 24, 3 72
2. 160 42. 5 39. 8 37. 1 27, 1 74
2. 164 42.8 40. 1 37. 4 27. 0 74
2, 181 40.4 37. 6 34. 7 28. 0 73
2, 184 40. 5 37. 7 34. 9 27. 6 72
2, 198 *(39. 2) 35. 1 32. 3 27. 9 70
2. 203 (38. 7) 35. 2 32. 2 29. 4 72
2. 248 (34.2) (30.9) 6)|------------|------------
(27.
2. 251 (33.9) (31. 3) (27.2)|------------|------------
2. 295 (31. 8) (28. 7) (25.0)|------------|------------
2. 304 (31. 3) (28.9) (25.3)|------------|------------
2. 504 (31.6) (28. 8) (25.2)|------------|------------
2. 505 (32.0) (28.5) (24.7)|------------|------------

"The values in parenthesesrepresent free-energychanges in 2-phaseregions except possibly at the com


Pºsitionsnear UO 2.2s. (Courtesy, American Institute of Physics.)

Shown in Table 6.7 are the calculated partial molar thermodynamic


quantities (Fo, Ho, So,) calculated by Roberts and Walter from their
ºxygen equilibrium pressure measurements. Note that these values
agree with those of Aronson and Belle for compositions above
0/U=2.18, but are lower at lower compositions.
The Roberts and Walter measurements of the partial molar free
energies for the UO., phase did not extend below UO2.1, a composition
range covered by the data of Aronson and Belle. Since the agreement
between these two sets of measurements is sufficiently close, however,
it is permissible to combine the results. Figure 6.17 is a plot of both
for
the

Roberts and Walter values for Fo, 1,396° com


at

and values
K

positions below UO, extrapolated temperature


of

from the data


in

by

Aronson and Belle. The combined data are best represented two
Fig. 6.17.
as

straight lines,
in

shown
Roberts and Walter calculated integral thermodynamic quantities
at

two temperatures (1,300° and 1,396° K) for the limiting composi


K
262 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 6.7—EQUILIBRIA IN THE UO,t, REGION*


O/U —Logio P (atm) —F., (kcal/mole) —H., –So,

1,493° K | 1,600° K | 1,695° K | 1,493° K 1,600° K 1,695° K kcal/mole e.u./mole

2.094- -------------- 5.450 4.965 ||----------- 39.90 38.42 63.0 14.4


*112”-----|----------------------|-----------|-----------|-----------|----------- 69 1s. 3
2.113--- 5.900 5.262 4.760 40.25 38.52 36.90 65.0 16, 5
2.114--- 6.010 5.402 4.925 41.02 39,52 38.20 62.2 14, 2
2.141--- 5.445 4.808 4.317 37.20 35.02 33.48 64.3 1s. 2
2-142"-----|-----------|-----------|-----------|--------------------------------- 74 25, 3
(2.158)----- 5.110 4.420 3.882 34.88 32.35 30.10 69.8 23.4
74 27.0
68.4 24.8
68.5 24- 9
71.0 26.6
71.2 25.4
72 27. 6
68.7 25 s
71.0 7.8
71.8 29 o
72.1 29.4
7.2.1 29 7
7 29. 4
68.7 28, 2
7.2.6 31.6
73.1 32.5
70.0 30. 9
70.5 31- 7
------ 67.7 3d 7
2.237------- 3.112 2.450 1.932 21.25 17.94 14.99 67.4 30. 9
2.237------- 3. 167 2.495 1.965 21,62 18.26 15.24 68.6 3:1.5.
2.239------- 3.138 l-----------|----------- 21,43 |-----------|----------- 72.5 34-3
2.239------- 3.035 2.362 1.840 20.72 17.30 14.27 68.2 31. s

*L. E. J. Roberts and A. J. Walter, “Equilibrium Pressuresand Phase Relations in the Uranium Oxide
System,” report in courseof publication, AERE, Harwell.
**Results taken from Table 6.6.

tions of each phase in the phase diagram from UO, to UAOs. The
lower temperature was chosen for direct comparison with the data of
Blackburn, and the higher temperature was selected because 1,396° K
is the highest temperature at which U.O., can exist, according to
Roberts and Walter. These integral thermodynamic quantities were
calculated from partial molar quantities by using equations such as
the following:

AF=N.
ſ "Fºlz.
Here, AF
represents the change in free energy on oxidizing UO.
(phase 1) with O2 (phase 2) through a nonstoichiometric range to a
composition where N, is the mole fraction of UO, and

Z=\;=} for the composition UO, , ,, where N2 is the mole fraction


1.

of excess oxygen.
Partial molar free energies (F) at 1,396° K were
taken from the plot shown in Fig. 6.17.
\ | URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 263
I I
LEGEND
-
O CALCULATED FROM DATA OF ARONSON AND BELLE
eſ
O

-30 H
O-ROBERTS AND WALTER
•’ -

ºu
-
o
: -
S -40 H
-: O
3.
O
>6
O

-50 H -
.*O o
l |
2.OO 2.IO 2.2O
COMPOSITION, 0/U

FIGURE 6.17. Partial Molar Free Energy of O, in UO,.. at 1,396° K. (L. E. J.


Roberts and A. J. Walter, “Equilibrium Pressures and Phase Relations in the
Uranium Oxide System,” report in course of publication, AERE, Harwell.)

Results of these calculations are shown in Tables 6.8 and 6.9.

Data of Blackburn
at K are included in Table 6.8 for
1,300°
comparison. Values of AF and AH for the overall reaction,
U0,40.3350,-UO,..., based on thermal data (see Table 6.5), are also
given in Tables 6.8 and 6.9. The agreement between the results of
Roberts and Walter and those of Blackburn at 1,300° K is good, con
sidering the different techniques involved. Roberts and Walter con
clude that the agreement between the AF values for the overall

Table 6.8–THERMODYNAMIC FUNCTIONS AT 1,300°K*

AF AH AS AS AFI
Reaction kcal/g kcal/g e.u.ſg e.u./mole kcal/mole
atom U atom U atom U O2 O:

"Orł0.000501-UO, in........---------- —1.45 –14.6 –60


ºr
–4. 09 —5.98
0.01850,-UO, e................. –0. 619 –2. 09 —1.13 –61 —118
ºn 10.007O-UO, e......... - - - - - - -- - –0. 216 –0. 627 –0. 32 –46 –86
–69.
a is-

"010.1250-UO, - –4. 93 –8. 70 –2. 90 —23


5

st-01801-Uo, –6, 71 –37.4 –77


to

—5.06 —13.80
"044003 01-UO,
e.................... —0.23 –2.00 —1.36 —45.3 –66
***0210,-UO, e................... –38 –75

—5.29 15.8 –8. 07


"940.3350-UO, e..................... 10.22 –24. 50 —10.97 –33 –73.
5

||

ºnparison with data


of

Blackburn [36].
Vºtºlz; on-UO, is...... --------------- –5, 50 –8.50 –2. ||-----------
0 7 3

**021 O-Uo...................... —5.30 15.9 –7, |------------|------------



|

"9:40.3350,-UO, e..................... —10.8 –24.4 -10. ---


|_
__
_
__
_
__
_
_
_!

-
--
-
-
"[..

Walter, “Equilibrium Pressuresand Phase Relations


A.
E.

Robertsand the Uranium Oxide


in
J.

of J.

*m,” report course publication, AERE, Harwell.


in
264 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 6.9—THERMODYNAMIC FUNCTIONS AT 1,396°K*

AF AH AS AS AH
Reaction kcal/g kcal/g e.u./g e.u./mole kcal/mole
atom U atom U atom U O2 Oz

UO2+0.122 O2=UO2.244----------------- --- –4.61 —7.5 –2.08 —17.0 –61


UO2.24+0.003O2=UO2.250----------------- –0.071 –0.8 –0. 52 – 173 –270
UO2+0.125O2= UO2.25--------------------- –4, 68 —8.3 –2.60 —20.8 –65. 4
O2=UOn 81--------------------
UO2.25+0.18 –4, 28 —13.8 –6.80 –37.8 –77
–0. 147 –2.0 —1.32 –43 –66
–4. 43 —15.8 –8. 12 –39 –75
UO2+0.335O2=UO2.87--------------------- –9. 11 –24. 1 —10.7 –32 –72

*L. E. J. Roberts and A. J. Walter, “Equilibrium Pressuresand Phase Relations in the Uranium Oxide
System,” report in courseof publication, AERE, Harwell.

reaction calculated from the oxygen pressures with those derived from
thermal data is also close enough to remove serious uncertainty arising
from the extrapolation of the partial molar free energy plots for
UO,..., to UO2.00.
The following values were calculated by Ackermann and Thorn
from their equilibrium oxygen dissociation-pressure measurements
[40]:
UO2.6—-U,Oo., +O. – AH=88.3+2.0 kcal/g-atom U.
U.Oo—-U,O,-, + O. – AH=98.8+3.7 kcal/g-atom U
Belle and Lustman summarized the results obtained by Blackburn and
by Aronson and Belle in a review of the properties of uranium dioxide
[36, 38,61].
Osborne, Westrum, and Lohr determined the low-temperature heat
capacity and calculated the entropy of U.O., from 5° to 300° C [62].
The entropy at 298.16° K is 20.07 calories-degree-" -uranium", giving
a AS*ass=47.0 calories-degree-' -mole' for UO, excellent agree
as,

in

ment with Brewer's estimate [53].


The thermodynamic data provide valuable contribution
an

un
to
a

derstanding the uranium-oxygen phase diagram. However, addi


of

required, particularly
on

tional information the tetragonal phases


is

U.O. Thermodynamic measurements may prove


on

UAO,
of to

near and
the only positive means establish the stability metastability
or
be

to

U.O.

Phase Studies and Crystal Structure Data


of

6.5.4 Correlation

apparent from the investigations summarized


It

above that the


is

by

uranium-oxygen system has been studied intensively


of

means
a

broad variety techniques, including X-ray studies room tempera


of

at

ture and high temperature, kinetic studies, magnetic susceptibility,


at

electrical conductivity, microstructure examination, equilibrium dis


sociation-pressure measurements, and emf methods. The temperature
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 265

range of these investigations is from room temperature to about 1,500°


C. In general, the studies are in qualitative agreement but differ in
details. It seems worthwhile, therefore, to attempt to correlate the
resultsobtained by the various methods in order to obtain a picture in
conformity with the maximum amount of data and to point out the
principal areas of disagreement. The results can be separated con
veniently into high-temperature and low-temperature regions of the
uranium-oxygen system.

High-Temperature Structures and Phase Equilibria


(a)

The relatively simple high-temperature phase relationships the

C in
UO,-U,Os system temperatures up approximately 1,500° ap
at

fairly to
be

well understood. There are disagreements concern


to

pear
U.O.,
ing

the exact placement phase boundaries (extent


of

UO2.x,

of
U.O...), but room temperature UO.
generally agreed that at
it
is

and

and U.O., show little, any, solubility other, that with increas
if

each
in

temperature the UO., composition range expands much more


ing

U.O., and that the lower limit


of

rapidly than that the UAOs-type of


y,

e.g.

structure lies within the limits UO, ºs-UO, Whether this phase
exhibits solubility reversal with increasing
(UO2.so temperature
a

C.,

C) remains
be
at

to

and 750° UO2.s 1,000°


determined.
at

500°

Some investigators have identified two closely related but structurally

distinct phases this composition region (UO.s and UAOs), while


in

the entire range the UAOs-type structure.


in

others class
by

UAOs reported Hoekstra, al., apparently closely


et

The
is
to 8

related the modification [46]. Where UAOs derivable from


a

is
a

UO,
by
the

hexagonal
of

structures decrease the a/b ratio from


an a

0.561, the form shows


to

increase
to

"577 0.618.
8

present the one region UO,-U3Os sys


At

the high-temperature
in

which requires further careful study the vicinity U.O.


of
in
is

'em
phase relationships between UO., UO.s
at

to to

700° and
C

The

are not clearly defined.


particular, one may refer
In
C

1200°

the UO., and


the

on

disagreements among the various investigators


[.0.., phase boundaries and the limits the U.O, phase.
of

U.O,
by

by
of

as

The structure interpreted Anderson and Perio


also open
to

question [25, 26].


be

Uranium dioxide (U.O.) can


is

interpenetrating cubic lattices


as

consisting
of

is of

tonsidered uranium
oxygen atoms which every other uranium atom missing
in

and
(see Chap. 5). Thus, each unit cell UO, contains
of

+4
U

O
+
4

(U.O.). Anderson and Perio assumed that oxidation


of

vacancies
by

UO, occurs the entry oxygen, either statistically dis


of

T0.
to


an

ordered fashion, into one


or

of
in
or

tributed more these vacancies.


in by

This hypothesis was supported density measurements which ruled


by

X-ray
out

of

the existence vacancies the uranium lattice and


266 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

diffraction results which seemed to show a retention of the fluorite


structure with only a slight adjustment in the cell parameter. The
X-ray scattering power of oxygen

is,
however, much less than that

of uranium. Thus, any changes oxygen positions during oxida

in
UO, U.O., would result only minor changes X-ray
of
tion

in

in
to
diffraction patterns.
During the past few years has become increasingly apparent

it
and theoretical grounds that the U.O., structure
on

both experimental
originally appeared.
be

complex argued

be
It
must more than can

it
that unlikely that the interstitial oxygen would enter the cell
is
it

(004) (equivalent positions), since this would neces


or
at

(###)
sitate placing the ninth oxygen atom cube con

of
at
the center

a
taining eight oxygen nearest neighbors. addition, further experi

In
mental work has shown that the diffraction pattern contains addi
tional weak interferences which cannot be accounted for on the basis
of

simple cubic structure [39].


a

Andresen,” cooperation with Frazer the Brookhaven National

at
in

Laboratory, obtained neutron diffraction diagram for U.O. This


a

diagram showed

be
of

to
number extra lines which could not traced
a

impurities and which required unit cell least six times the UO, at
a

the fluorite-type reflections showed that the


of

cell. The intensities

up
Hering and Perio
an
of

extra oxygen being taken


of

in
idea the
UO,
of

supported [11].
be

center the cell could not The observed

of
require rearrangement of
to
as

intensities were such several the


a

oxygen atoms.
low-temperature oxida
on

Andresen and Frazer also observed that


UO, the intensities the fluorite-type reflections varied
of

of

tion
linearly with oxygen content, indicative two-phase region. This
of
a
be

phase
of
UAO. Since the observed intensities
to

new was assumed


the fluorite-type reflections for U.O. showed the same deviations from
UO, did those for U.O., Andresen and Frazer concluded that
as

U.O., and U.O. are structurally very similar and that the extra
by

oxygens are taken up the same mechanism.


by

Perio" reported the U.O., structure,


as

determined neutron
by

X-ray diffraction both powders and


on

on

diffraction powders and


of

single crystals,
be

cubic (four subcells) with


to

cell size 21.7


in be a

angstrom units. The space group His sug


to

believed 143d.
is

gestion for the distribution the excess oxygen the U.O., cell
of

is

shown Fig. 6.18.


in

by

As pointed out Aronson and Belle, the existence diphasic


of
a

region between UO, and U.O., temperatures


of
to

900°
in

excess
argues against any simple relationship
of

the structures these


in

*A. Andresen, personal communication.


F.

Perio, personal communication.


P.
*
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 267
Y Y
º
O
O
|
O
- Ol
O
O
O
x x
O x
000 OOI OO2

FIGURE 6.18. Positions of the Excess Oxygen in the U.O., Cell. (P. Perio,
personal communication.)

two compounds The entropy changes during the various stages


[38].
of the formation of the U.O., phase show clearly that this is a highly
ordered phase compared to the UO., phase. Further neutron dif
fraction work can be expected to resolve this problem, since the neutron
scattering powers of oxygen and uranium are comparable.

Low-Temperature Structures and Phase Equilibria


(b)

phase relationships
At

low temperatures
of of

determination com

is
a
of by

plicated not only the larger number reported structures, but by


poor crystallinity oxide samples, apparent
of
slow attainment
a

equilibrium, and the corollary problem metastability. Anderson,


of

Roberts, and Harper have shown that bulk oxidation inactive UO,
of
as
at

temperatures low C.;


is at

at
as

occurs measurable rate 70°


a

lower temperatures oxidation essentially surface phenomenon


a

(Inactive UO, dioxide powder with surface


as

[27]. defined here


is

is,

Poorly crystalline however, oxi


or

area ×2m+/g.) active oxide


C,

its

largely because
to

of
at

dized considerable extent 25°


a

extremely high surface area [31, 33]. Present evidence indicates


that the surface oxidation product amorphous UO, [43].
is

disagreement exists concerning the effects


of

of

certain amount
A

low-temperature oxidation inactive UO. generally conceded


of

It
is

by

that the crystallinity impaired,


of

as

the oxide evidenced the


is

X-ray unanimity
no

development pattern, but


of

in

diffuse lines the


opinion evidenced concerning the remaining details. Results
of

is

indicate that oxygen pressure, temperature oxidation, temperature


of

the annealing process, sample crystallinity, and per


of

and duration
haps even the previous history the starting material
of

are factors
be considered.
to

An attempt has been made Table 6.10 summarize the low


to
in

temperature oxidation data with respect composition limits and


to

no

the phases involved. There can, however,


be
of

cell dimensions
direct comparison among the observations, since experimental condi
If

are not identical. their


to

at

tions the observations are be taken


value, evident that differences particle size (reactivity),
in
is
it

face
268 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

temperature, etc., can produce widely varying results. Not only


are there large variations in the upper limit of the cubic phase
(UO2.01–UO2.so), but in the cell edge as a function of composition
as well. the extreme case, UO2.so resulting from room-temperature
In
oxidation of active UO, is converted to amorphous UO, rather than
UAOs on being heated in air at 300° C. No evidence is found for the
formation of tetragonal UAO, as an intermediate.

TABLE 6.10—LOW-TEMPERATURE OXIDATION OF URANIUM


DIOXIDE

Cubic phase Tetragonal phase

References o º
Composition Cell edge, A Designation | Composition Cell edge, A
limit limit a c

19------------|--------------|--------------|-------------- 2.33 -----------------


20------------ 2.34 5.41 *------------ 2.40 5.38 5. 55
21------------ *2.24 5.468 || Y------------ *2.34 5.397 5.565
23------------ 2.10 --------------|-------------- 2.50 ------------------
11------------ 2.35 5.427 | Y------------ 2.40 5.38 5.55
24------------ 2.10 -------------- -*------------ 2.32 5.447 5.400
7------------ 2.40 5.375 5.542
26------------ 2.33 5.468 Y’----------- 2.31 c/a = 1.016
7"---------- 2.37 cfa = 1.031
7------------ 2.29 5.44 Y”---------- 2.37 c/a. = 1.031
25------------|--------------|-------------- 71----------- 2.40 5.37 5.54
71----------- 2.33 c/a = 1.016
73----------- 2.33 c/a = 1.010
33------------ **2.50 <5.44 | None-------|-------------- ------------------
31------------ "2.40 l--------------|--------------|--------------|------------------
34------------ 2.07 5.466 |-------------- 2.33 5.408 5. 504
34------------ 2. 13 5.466 || Y------------ 2.35 5.387 5. 549
35------------ 2.06 --------------|-------------- 2.34 || -----------------
37------------ 2.01 5.469 || Y----------- 2.41 5,384 5.556

43------------ <2.04 5.469 | a_----------- 2.33 5.45 5. 40


S.----------- 2, 33 5, 38 5.55
<2.33 c/a = 1.016
<2.33 c/a = 1.012

*Corrected for UO2.04starting material.


**Active oxide.

As many as four tetragonal phases have been identified by X-ray


studies of the UO2.s-UO.s composition range. Two of the tetragonal
phases, a with c/a = 0.991 and 8 with c/a+1.031, are obtained during
the oxidation process, the others during the annealing step. All four
phases have c/a ratios approaching unity and unit-cell or pseudocell
dimensions within a few percent of the parent UO2. This fact, to
gether with their poor crystallinity, has led to difficulty in identifica
tion of the individual phases. This situation holds true especially for
the a phase, which has been identified only under conditions where
maximum resolution was obtained. The ready conversion of a to 8
U3O may help to explain the inability of some investigators to recog.
uRANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 269

its

to
nize existence and report
cubic solid-solution region preceding

a
crystallization

no
phase. appreciable

of
the Kinetic data show
| |

3
change oxidation rate attributable conversion,
be in

to
such but none

a
would expected the oxygen diffusion rates the two phases are

if

in
comparable. While early work indicate some oxygen solu

to
seemed
U.O., more recent report crystallization or
in

investigations

of
tion
thorhombic U.O, phase the composition UO,

as.
Whether the

in at
annealing process results

of
the formation
two distinct structures
UAO, and U.O.,

or
intermediate between whether solid-solution
8

to a
region with continuous variation c/a from 1.031 exists, has

in

1
not been conclusively proved. At present, the data seem

to
indicate
stepwise process.
The exact composition the annealed phases

of

is
a

doubt. Perio and Anderson place them between UO2 and


in

to
UO2.2s, whereas later work would indicate composition between

a
UO2 and UO2.25 [25, 26]. Both, however, agree that they contain
as

less oxygen than the phase from which they are derived.
8

The evidence appears favor the conclusion that the phases ob


to

tained during low-temperature oxidation experiments are metastable.


Perio reported proof for the thermodynamic stability the tetrag

of
he

onal phases, since detect their presence


to

was able mixtures


UO, U.O,
heated for extended periods [25]. This in
of

at

and 300°
C

result can, however, the following two-stage reactions:


be

to

attributed
U3Os—-UAOs-,
+

O2
UO2+O2–U3O,
The more conclusive experiment (heating U.O., and
of

mixture
a
by

U.O.) has been performed Vaughan, al., Blackburn, al., and


et

et
al.

Hoekstra, [31, 37, 43]. No tetragonal phase was detected.


et

pointed out that no reliable structure determination has


on be

should
It

any the tetragonal phases. For convenience, they


of

been made
have been discussed only the parent fluorite structure
to

relation
in

apparent c/a ratio based


an

on

and have been given face-centered


a

reality,
no
In

tetragonal cell. such symmetry exists, and the struc


body-centered cell with anº-are/V2.
be

tures should
to

converted
a

on

Considerable debate has resulted from density measurements the


range UO, UO,
the composition
to in

in to

uranium oxides These data


1.

determine whether oxygen


of

have been used excess stoichio


metric UO, interstitial oxygen, whether uranium vacancies are
is

whether partial substitution


or

created, oxygen for uranium occurs.


of

now generally agreed that U,O, and the tetragonal phases near
is
It

by

oxygen
is of

the addition the fluorite lattice. On


to

U.O. are formed


recent experiments apparent, however, that diphasic
of

the basis
it

are involved between UO, and U.O., UAO. Apparently,


or

systems
compounds unique UO, but
to

these have structures which are related


by

not connected solid solution region, and one not technically


is
a
270 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

correct in referring to the excess oxygen as interstitial oxygen. The


high-temperature UO., solid-solution region can, of course, be dis
cussed in terms of interstitial oxygen ions or uranium ion vacancies.
The data of Gronvold and Schaner on lattice parameter versus com
position indicate that nonstoichiometry is due to interstitial oxy
gen [28, 29]. The single exception to a general agreement on
UAO, density data pertains to a U.O. Perio reports a density in
agreement with uranium vacancies, while Hoekstra, et al., find com
parable densities for a and 8 U.O., both indicative of oxygen addition
to the fluorite lattice [24, 25,43].
The electrical conductivity data of Willardson, et al., are puzzling
in that they do not agree with X-ray results, and the phase diagram
suggested by these measurements does not agree with other work [35].
The first portion of the oxidation process is straightforward: UO2., is
formed and conductivity increases. With further oxidation, U.O.
begins to appear and a decrease in conductivity occurs. Extrapola
tion to U.O., indicates that little solid solution exists in this phase (in
agreement with Gronvold) [28].
The sudden shift from p-type to n-type with increased conductivity
is difficult to explain, since X-ray results indicate only a continued con
version of UO., to U.O. The conductivity data suggest that above
UO, as a second type of U,O, structure is formed, a structure which is
inherently a better conductor, and closely related to the UAO, con
figuration, since no U.O., metal-deficit region is observed. These
anomalous results further emphasize the need for complete struc
ture determination on U.O., and the U.O. phases.
Note, that the electrical conductivity measurements of
however,
Aronson 7 (see Chap. 5) are in good agreement with other determina
tions of the UO,-U,O, phase diagram. In Aronson's work, the phase
boundary of the UO., region was determined by observing the tem
peratures at which abrupt changes occurred in the electrical conduc
tivity for a number of nonstoichiometric compositions. The phase
boundary data obtained from the conductivity measurements agree
with the metallographic data of Schaner at temperatures below 900°C
[29]. At higher temperatures, the conductivity results deviate from
the metallographic observations but are in good agreement with the
phase diagram obtained from measurements of the oxygen pressures
in equilibrium with uranium oxides [38, 41]. The existence of a
large range of nonstoichiometry in the U.O., region which was found
in the metallographic study is not confirmed by the electrical conduc
tivity measurements.
In conclusion, it can be said that in spite of extensive research activ
ity on the UO-U,O, system since 1940, many details remain to be
* S. Aronson, personal communication, Bettis Atomic Power Laboratory.
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 271

clarified before a complete understanding of this complex system is


realized.

6.6 BINARY OXIDE SYSTEMS


H. R. Hoekstra and H. S. Parker

Systems composed of uranium oxide mixed with another metal oxide


have been investigated in much less detail than has the uranium oxide
system itself. Only infrequently has there been an attempt to make a
detailed study under carefully controlled conditions with a large num
ber of samples covering the entire composition range. These mixed
oxides have generally been studied as two component systems
(MO-HUO, or MO-FU,Os) rather than as three component systems
(MO-H-UO,--O or MO-HUO, +U,O.). As a result, our information
is restricted to one or two tie lines within the phase diagram. Ther
modynamic data on these are virtually nonexistent. In general, the
stability relationships of uranium oxidation states in the binary oxides
are similar to those encountered in the uranium oxide system itself.
No binary oxides of trivalent uranium are known, and those contain
ing tetravalent uranium must be prepared under controlled conditions
(in vacuo, or in an inert or reducing atmosphere). One noteworthy
difference in the binary oxides is a stabilization of U (VI) through
the formation of uranates (M.UO, M.U.O., M.UAO, o, etc.) with
monovalent or divalent metal oxides. Since the present discussion is
limited to systems containing U (IV) and U (V), the uranates will
not be considered here. Air-ignited mixtures containing Group III
or IV metal oxides are classed under U (V), even though the uranium
valence may be somewhat in excess of this value.
Direct comparison between mixed-oxide systems can best be made
in terms of mole percentages rather than weight percentages. For
this reason, all data in the figures and text of this chapter are given in
mole percentages. To aid further comparison between various sys
tems, oxide percentages are calculated from formulae containing one
gram-atomic weight of the metal rather than from the conventional
whole-number formula; thus, UO, ºr rather than UAOs, UO, as rather
than U, O, La O.'s rather than La2O, etc.

6.6.1 Binary Systems with Group I Metal Oxides


(a) Alkali Metal Oaſides

Very little information is available on systems containing UO, or


U.O., with Group I metal oxides. Heumann, Salmon, and Wilk, in
studying the system formed by a slurry of UO, in liquid NaK, found
some indication for the formation of a compound Na, UO, but the
272 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

evidence is insufficient at the present time to place much reliance on


the suggested formula [63].
Rudorff and Leutner were unable to find evidence for compound
formation in the Li2O-UO, system, but succeeded in preparing Nau(),
and LiDOs by the reaction
M.UO, +UO,->2MUO, [64].
Attempts to prepare the corresponding potassium compound were
NaDO, has a density of 7.02 g/cc;

its
unsuccessful. structure

is
LiLO, has rhombic monoclinic structure;

or
deformed perovskite.
a

a
its density 8.26 g/cc.
is

(b) Copper Oa'ide

Lang, Knudsen, Fillmore, and Roth investigated

of
the behavior
UO,-CuO
is at

system helium atmosphere [32].

in
to
the 700° 900°

C
to a
They found that CuO part the cuprous oxide and
in
reduced

in
by

part

in of
Mixtures uranium dioxide
to

the metal uranium dioxide.


and cuprous oxide, heated for helium, indicated
at

700° hours
C

by 6
that the cuprous oxide was being reduced the uranium dioxide, with
subsequent oxidation Cuprous oxide, heated
of

the uranium dioxide.


alone under the same conditions, showed only slight decrease oxy

in
a
gen content.

6.6.2 Binary Systems with Group Metal Oxides


II

(a) Beryllium Oaside


or

No solid solution compound formation with BeO was observed


by

Lang, [32]. The liquidus curve


al.

between 800° and 1,800°


et
C

region composition be
to
of

was found
in

obtained the the eutectic


good agreement with Epstein and Howland's theoretical calculations in
(Fig.6.19) [65]. Budnikov, al., reported the eutectic temperature
et

and composition
to

2,170+20° mole percent BeO [66].


68
be

and
C

(b) Magnesium. Owide

Epstein and Howland also calculated simple eutectic for the


a

UO,-MgO system. Lambertson and Mueller, however, reported the


presence two eutectics (see Fig. 6.20) and two immiscible liquid
of

phases [67]. They observed partial oxidation UO, by MgO


of
a

above 1,200° (as evidenced by change UO, cell parameter) and


in
C

rapid vaporization MgO above 2,000°


or
C.

No compounds
of

meas
urable solid solutions were detected. Lang,
no

al., agreed that


et

compound solid solution was formed [32].


or

as

oxygen addition
of

Anderson and Johnson studied the effect


binary UO,-MgO
as

well the simple system. No compound formation


URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 273
3OOO i l i I i I I I i
º-
- -- LANG, et al. H

--
28OO H. Nº. EPSTEIN AND HOWLAND –
S. * -
(THEORETICAL)
N. SS BUDNIKOV, et al.
O
•.
uj 26OO
ar
E
<I

#!
# 2400
u
H.

22OO

2OOO
I l L l l l l l I
O IO 2O 3O 4O 5O 6O 70 8O 90 IOO
UO2 MOLE PER CENT BeO BeO

FIGURE 6.19. UOz—BeO System.

was observed between 300° and 2,350° C, but a few mole percent
solubility of MgO in UO, was reported at the highest temperatures
[68]. The addition of oxygen resulted in a marked increase in MgO
solubility; approximately one molecule of oxide was dissolved for
each atom of oxygen absorbed to give a fluorite-type lattice.
Budnikov, et al., reported a single eutectic in the UO2–MgO system
and confirmed the oxidation of UO, by MgO at high temperatures
[66]. They found that mixtures of UO, and MgO heated in air at
temperatures of 1,600° to 1,750° C resulted in solid-solution formation
(up to 37 mole percent MgO) with preservation of the uranium
dioxide lattice (Fig. 6.21). The fact that the presence of oxygen en
hances solubility of MgO in UO., was confirmed by an experiment in
which a solid solution containing 33 mole percent MgO was heated in
argon at 2,000° C. Magnesium oxide precipitated out of the solid
solution and the lattice parameter increased from 5.28 angstrom
units to 5.462 angstrom units as the MgO content decreased to 3 mole
percent. Hoekstra and Katz studied this system from the opposite
up is,

extreme, that thermal decomposition magnesium uranates


of

at

temperatures Compounds having magnesium


to

1,200° [69].
C

1:1, 1:2, and 1:3 form fluorite-type lattice ex


of

uranium ratios
a

the (Mg,U) The composi


on

O.

tending composition.
of

either side
triangular
of

be

tion limits the fluorite structure can estimated


in
a
by

plot (Fig. 6.22) combining the data obtained these investiga


in

percent solubility MgO UO.


37

of

evident that the


It

tions.
in
is

mole percent
50

C.
to
at

has increased least 1,200°


at
274 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
3OOO T T I i I
I T | I

-
--- LAMBERTSON AND MUELLER
EPs TEIN AND HowLAND

---
W (THEORETICAL)
2800 HA - BUDNIKOV,et al.
N. V
*

26OO H.

- LIQUID
O
\ ſ

\\
º
º
# /
->
: 2400 H 1 –
§

- - - - - -, -* - - - \* - - - - - - - - */ - - --
#
* L.
|- 2280°

/
2259. - N- - - - >34 - - -

— —2330°
22OO H.
- - _\ - -- - - - - -
-
2159° V

——º-------------------
*
<-----
z

2OOO H. -
UO2 + MgO

18OO
1 | I | | | L 1– l
O 2O 4O 6O 8O too
UO2 MgO
MOLE PER CENT MgO

FIGURE 6.20. UO,-MgO System.

The line drawn between UO, and MgCO, represents the stoichio
metric fluorite composition line. The 1:1 magnesium-oxygen solu
tion in UO, reported by Anderson and Johnson falls along this line
[68]. It cannot extend as far as MgCO, since the monouranate has
a rhombic structure. Zachariasen has shown, however, that this phase
may be interpreted as a deformed fluorite lattice [70].
It will
become evident during this review of binary oxide systems
that the fluorite structure is frequently retained as uranium is oxidized
to U (V). This property seems to hold true for the MgO-UO2-UO,
system as indicated (Fig. 6.22) by lines B, C, and D, which give the
composition range observed for as Mg Us Os–
fluorite structure
Mg|U.O.s, MgC.O., MgC.O., and MgCO, ,-MgUOss. The upper
limit of the phase with respect to oxygen has been fairly well estab
I-
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 275
29 OO
T-I I-I T-I I

-
25oo H.
\\ -
/
| \\ /

2 - \\ / -
ſ

/
1900 H.
\
|-
V /
17oo H
O O
-

1500 H.
-
I | 1 | I | l | I
O 2O 40 6O 80 IOO
MgO
U02+x
MOLE PER CENT MgO

FIGURE 6.21. UO2.x-MgO System.

lished;the lower limit is less definite. Anderson and Johnson re


of a single fluorite phase from Mgolos, UO,
oss

at to

ported the existence


UO, (line A) upon oxidation
as,

the initial composition


oz.

of

Mg,
the fluorite lattice below the UO, line
of

[68]. Extension
C

400°
can, perhaps, the closely related
be

attributed
to

(anion vacancies)
rare-earth type structure with which UO, forms extensive solid
C

solutions.

(c) Calcium Oaside

Alberman, Blakey, and Anderson investigated the CaO-UO. phase


[71]. They
in

the temperature range 1,650° 2,300°


to

diagram
C

eutectic temperature (Fig. mole percent


45
of

6.23) 2,080°
at

found
C
a

57.4789 O–61—19
276 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FIGURE 6.22. UO,-UO,-MgO System; Lines A through D Represent Observed


Extent of Fluorite-Type Structure, Ruled Area Outlines Estimated Extent of
the Structure.

of UO. No measurable solubility of UO, in CaO was observed, but


an extensive solid solution of CaO in UO, (ranging from 20 mole
percent CaO at 1,650° C to 47 mole percent CaO at the eutectic tem
perature) was found. Two compounds, Ca.UO, (tetragonal) and
CaLOs (cubic, 10.748 angstrom units), are formed below 1,750° C.
The latter is equivalent to the rare-earth type C structure with Ca”
and U“replacing 2M".
Lang, et al., disagree with Alberman, et al., relative to the com
pounds formed and the structure assigned [32, 71]. In studying
three compositions containing 20, 62, and 83 mole percent CaO,
evidence for only one compound in addition to the solid-solution
fluorite lattice was found. The second structure was identified as an
orthorhombic perovskite-type compound. Since it was observed only
in the 62 and 83 percent CaO mixtures, it was assigned the formula
Ca(U,Ca) O.
Lang and associates also annealed several mixtures (which had pre
viously been heated to 1,800° C) for 360 hours at 1,500° C in a helium
atmosphere. Considerable care was taken to ensure a pure helium
atmosphere surrounding the mixtures. Varying amounts of oxida
tion of the uranium dioxide were noted after this treatment. In the
mixture containing 75 mole percent of calcium oxide, the Ca(U,Ca) O.
H-H-I-I-T-I
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 277
2800


Nss
2600 H.
SS
SSS
LIQUID -
SN
\\\ \ -
N. N.
_2^
24OO H.

/
2^
UO2s.s. *
\\
\
- LIOUID
Q 2^ -
o
#
->
º
22OO -
|-
\
\ \\
\ /
/ CoO+
LIQUID
2O8O3:2O"
-
-
§ U02 s.s.
-
-- - - - - - - - --
# Coo-UO2s.s.
ºf 2000
*-
H 1950°

i8OO H. -

16OO H.
/
Couos-
UO2s.s.
Couos +
Co2UO4
Co2UO4+
CoO -

14OO I | l | l | I | I
o 2O 40 60 8O IOO
MOLE PER CENT CoO

FIGURE 6.23. UOr—CaO System According to Alberman, et al. [71]. (Courtesy,


The Chemical Society, London.)

compound was the major phase, with a small amount of Cal JO, pres
ent. The 1UO. :1CaO mixture containing only CalT0, and the
mixture containing 25 mole percent of calcium oxide showed only the
cubic UO, solid solution. Heating the 1UO, :1CaO mixture at
1,350° C for4 hours in air caused no further change in the X-ray
pattern, although a color change from gray-green to bright yellow
was noted.

The results reported by Lang, et al., are open to question, since sev
eral lines of evidence indicate that their samples were partially
oxidized during preparation [32]:

1. The cell dimensions (5.440 angstrom units) quoted for 20 percent


CaO, 80 percent UO, are lower than the value reported by Alber
UO, has
al.

of

man, et (5.4426 angstrom units) [71]. Oxidation


decrease the unit-cell dimension, and similar phe
to

been shown
a

nomenon has been reported for the UO,-CaO-O system; Hoekstra


and Katz report 5.379 angstrom units for Cau.O., com
of

value
a
278 URANIUM Dioxide: PROPERTIES AND NUCLEAR APPLICATIONS

pared with Alberman's 5.429 angstrom units for the composition of


CaL.O. [69, 71].
2. The fact that the ionic radii of U" (0.93 angstrom unit) and Ca”
(0.94 angstrom unit) are very nearly the same would appear to
make the formation of a perovskite structure doubtful, whereas
oxidation to U” (0.87 angstrom unit) would give a more favorable
radius ratio.
3. Annealing experiments by Lang, et al., at 1,500° C led to the iden
tification of Cal]O, in some of the mixtures [32]. The free-energy
relationships are such that it seems unlikely that the uranate could
be formed by any means except the introduction of additional
oxygen.

(d) Strontium Oaside

Lang, et al., reported the UO,-SrO system to be analogous to the


calcium system [32]. A cubic fluorite-type UO, solid solution and an
orthorhombic, perovskite-type structure were identified. X-ray dif
fraction data gave the parameters a =6.101 Å, b=8.60 Å, and
c=6.17 Å. Here again, the likelihood of oxidation must be considered,
since SrüO, was identified in several annealed samples.

(e) Barium, Oacide

Lang, et al., found the BaO-UO, systems to resemble the calcium


and strontium systems, except that the perovskite-type compound,
probably Bal]Os, is pseudo-cubic (a+4.387 Å) rather than ortho
rhombic [32]. The solubility of BaO in the UO, solid solution struc
ture was estimated to be 25 mole percent at 1,800° C. Some evidence
for solid solubility of barium oxide in Bal]O, was observed. Mixtures
of 1UO, ; 2BaO (mole ratio) contained a single-phase, cubic perov
skite-type solid solution after reaction at 1,800° or 1,900° C for one
half hour in argon, as contrasted with 1UO. : 3BaO mixtures, which
contained free barium oxide.
Furman has also identified the compound BaúO, as one product of
the reaction between barium metal and uranium dioxide [72]. He re
ported the compound as being in the cubic system with a = 4.372=0.002
Å, and concluded that barium oxide and uranium dioxide react stoichi
ometrically to form the true compound Bal IOa. The formation of
BaúOa was first noted after heating a sodium-potassium suspension
(NaK) containing a 1 to 2-mole mixture of barium metal and uranium
dioxide at 77.5°C for 205 hours. Subsequently, BaúO, was also pre
pared by the reduction of Bal IO, with hydrogen at 800° C. Furman
reported that the barium uranium trioxide underwent no detectable
oxidation after 48 hours at 780° C in air [72].
URANIUM-oxygen AND BINARY oxIDE SYSTEMS 279

6.6.3 Binary Systems with Group III Metal Oxides and Rare
Earths

(a) Aluminum Oaside

The small aluminum ionic radius led Epstein and Howland to pre
dict the absence of any solid solution or compound formation in the
UO,-Al2O, system (Fig. 6.24) [65]. This was verified by Lambertson
and Mueller, but a two-liquid immiscibility region was reported be
tween 54 and 75 mole percent Al2O, (70 and 86 mole percent AlO1.5)
[73]. Lang, et al., however, found only a single eutectic at 1,915°C,
in close agreement with the Epstein and Howland prediction [32].
Budnikov, et al., also reported a single eutectic at 1,900+10° C con
taining 85 mole percent (74 mole percent Al2O3) [66]. Mix
AlO.s
tures on either side of the eutectic tend to separate into two layers.
With an excess of alumina the less dense corundum crystals float to

-
the top. Similarly, uranium dioxide sinks to the bottom of mixtures
containing an excess of the dioxide. This liquation phenomenon is
believed to be responsible for Lambertson and Mueller's report of a
two-liquid immiscibility region.

--
i T I I I I I I i

A EPSTEIN AND HOWLAND

- --- -
(THEORETICAL)
B LAMBERTSON AND
3OOO.H.
C LANG, et al. wºuin

28OO LOUID

26OO | -
2 LIQUID
IMMISCIBLE
2400 H.
N
i 2200 H. uo, + Loud
N.
N^
2OOO. H.
N /
1930+IO* >4
1900° 1915+

1800 L
IO
|
2O
|
3O 40

UO., + AIO
| * '
50 6O
1–1–
70 8O 90 IOO
O
MoLE PER CENT Aiols

FIGURE 6.24. UOz-Al O.s System.


280 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Roberts, et al., have reported the formation of a cubic solid solution


with cell constants of 5.42 to 5.43 angstrom units when Al2O, and
UO2, were heated in sealed containers at 1,100° C [41]. Reduction
with hydrogen resulted in an increase to 5.448 angstrom units, as com
pared with a constant of 5.468 angstrom units for pure UO2.

(b) Scandium Oaxide

Wilson reported that a mixture containing 55 weight percent UO2,


45 weight percent Sc.O., (76 mole percent ScQ, s) did not form a
solid solution after ignition in vacuum at 1,750° C [74]. Air ignition
at 1,375° Cconverted the material to a fluorite solid solution with
a=5.13 Å. Hund and Peetz gave a parameter, a = 5.234 Å, for an air
ignited mixture containing 50 mole percent ScQ1.5 [75].

(c) Yttrium, Oaxide

Anderson found almost complete miscibility in the UO2–Y.O.


system after ignition of the mixed oxides at 2,000° C [26]. Some
what later, Anderson, Ferguson, and Roberts reported that carbon
monoxide reduction at 1,200° C, or vacuum reduction at 2,000° C
of coprecipitated yttrium and uranium (VI), did not convert the
uranium to its tetravalent state [76]. Instead, the oxygen-to-metal
(uranium plus yttrium) ratio approached two over a wide com
position range. Heating with uranium metal at 1,000° C was re
quired fully to reduce the uranium oxide to U (IV). Exposure to
air at room temperature resulted in reoxidation to compositions
approaching (U,V) O2. At higher temperatures, further oxidation
occurred to produce a lattice containing interstitial oxygen.
In a detailed study of the crystal structure of the UO2–YO, a sys
tem, Ferguson and Fogg found an extensive solid-solution range
with pseudofluorite structure extending between 0 and 78 mole per
cent YO,... [77]. An immiscibility gap exists between 78 and 96
mole percent YO,..., and a solid-solution region based on the Tl2O,
structure exists between 96 and 100 mole percent YO,...
Hund, Peetz, and Kottenhahn studied the yttrium-uranium oxide
system by coprecipitating Y(III) with U (VI) and igniting the mixed
oxides in air at 1,200° C [78]. A diphasic system (MO...+UO2.s.)
was observed between 0 and 30 mole percent YO,... and a fluorite-type
solid solution from 30 to 65 mole percent YO,.5. At higher yttrium
concentration the type C rare-earth structure is found together with
the fluorite phase. A summary of the data on the uranium-yttrium
oxide system is given in Fig. 6.25.
The specific resistance as a function of temperature was measured
on three of the mixtures (40, 52, and 65 mole percent YO,...). The
mixture containing 52 percent yttria showed the lowest conductivity
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 281
I I I I I i I I i

o Yos
U02 +
-

(U,
5.45
A Y)02
UO2+x YO15

+
M.

N
N.

5.4O H.

eq.
J

o
º
uu
H.

5.35
=

uu
O

5.30
H

IO _l

L
|

20 30 40 50 60 7O 80 90 IOo
O

of

MoLE PER CENT Yos


FIGURE 6.25. UO2–YO.s—Oz System.

and, consequently, the nearest approach perfect fluorite (MOA)


to
a

lattice. The low activation energy for electronic conduction (0.12 ev)
U(IV)a=U(VI) exchange. Lang, al.,
of
to

was attributed
et

the ease
looked for the appearance pyrochlore-type structure (A3'B:”O.)
of
a

the yttria-urania system, but could find


no

evidence for the doubled


in

of

unit cell characteristic this structure [32].

(d) Lanthanum, Oaxide

Wilson has studied the high-temperature properties the uranium


of

lanthanum-oxygen system and stabilization


of

the fluorite structure


valence compensation [74]. Mixed oxide systems
of
by

urania with
calcium, scandium, and thorium oxides were also investigated
in to

some
extent for comparison purposes. Properties investigated these
high temperatures, extent
of

volatilization
at

systems include uranium


uranium oxidation air, crystal structures,
in

and electrical
of

properties.
by

Y.O., fluorite-type
of on

As has
been shown previous studies
a

retained over considerable extent the La2O,-UO, system.


is

lattice
a

Air oxidation leads to retention of the fluorite structure with reduced


lattice parameters, agreement with Hund and Peetz data for U.O.,
in

and La2O, mixtures [75]. Unfortunately, Wilson has plotted his


282 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

results in terms of weight percent rather than in mole percent, thus,


obscuring some of the correlations with previous work on this system
as well as other systems.
A mixture containing 40 weight percent
(28 mole percent) and UO,
60 weight percent La2O, (72 mole percent LaO,...) showed no volatili
zation of uranium when heated in air at 1,550° C. Oxidation of

all
uranium to UO.e. occurs rapidly (as is true for the mixed-oxides
studies). Uranium volatilizes from mixtures containing less La-O,
and from urania-thoria mixtures comparable temperatures.

at
UO2–La2O, mixtures showed decreasing p-type
on
Electrical studies
conductivity with increase

70
to in
lanthanum content. Above mole
percent lanthanum, shift n-type conductivity was observed.
a

the UO, fluorite structure

of
Wilson discussed stabilization terms

in
of

the zone theory


of

of
solids and electron/atom ratio [79]. The
region over which the fluorite structure exists dependent

be
to
shown

is
upon the valence and size
of

the additive ion the oxidation

as

as
well
state of uranium.

(e) Neodymium, Oaside

Lambertson and Mueller studied the UO,-Nd2O, system

at
elevated
temperatures [80]. The phase diagram (Fig. 6.26) indicates fluorite

a
type solid-solution region extending mole percent Nd2O3, anal
64
to

32OO
T
I
I

-
28OO
FSS^ LIQUID
-

* ** ~ * * *
T

**
-

* ** ^
~Y-J
*-
--
-
~
Y

24OO H.
O
NS
Y-SA-e:-
•.

—H
:#

| + |
|

2000
H

or
ul
|

0.
-:
|

ul
H.
|

16OOH -
|

|HEXAGONAL
+
|

12OOH. FACE-CENTERED CUBIC F.C.C. S.S.


|

SOLID SOLUTION
|
|

8OO
|

|
|
l

2O 4O 6O 8O IOO
o

UO2 Ndol.5
MOLE PER CENT Ndols

UO-Nd2O, System. According Lambertson and Mueller [80].


to

FIGURE 6.26.
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 283

ogous to the binary yttrium system. The neodymium-rich phase,


however, crystallizes in the hexagonal, type-A rare-earth structure.
No solid solution of UO, in Nd2O, was detected at 1,600° C.
Lambertson and Mueller also investigated the oxidation resistance
of the solid solutions by reheating the various mixtures in air at 1,000°
C[80]. Compositions containing up to 29 mole percent of neodymium
oxide contained orthorhombic U,Os and a face-centered cubic phase
with a lattice parameter smaller than that of the original solid solution.
In samples containing from 29 to 66 mole percent of neodymium oxide,
only the face-centered cubic phase was present after the air heating;
in samples containing more than 66 mole percent of neodymium oxide,
hexagonal neodymium oxide was present before and after heating in
alr.
Lang, et al., investigated the possible existence of a pyrochlore-type
structure in this system; their results were inconclusive [32].

(f) Dysprosium Oaside

In a oxide ceramics, Ploetz, et al., found single


study of dysprosium
phase face-centered cubic solid solutions of Dy2O, in UO2 to at least
70 weight percent Dy2O, (77.2 mole percent DyO,...) [81]. The lattice
constants (5.448 to 5.434 angstrom units) were found to vary according
toVegard's law up to 50 weight percent Dy2O, (59.2 mole percent
DyO,...). In the region 50 to 70 weight percent Dy:Os, which is still
single-phase, there is a discontinuity, and the lattice constants vary
from 5.344 to 5.340 angstrom units. The DyO,.5–UO, compositions
containing 80 weight percent Dy.O. (85.3 mole percent DyO,...) and
90 weight percent Dy,O, (92.9 mole percent DyO,...) were two-phase
with a body-centered cubic solid solution of Dy2O, as the second phase.
No compound formation was observed in the UO2–Dy2O, system.

(g) Other Rare Earth Oa'ides

Hund and Peetz investigated the solid-solution regions encountered


in some of the rare-earth-oxide, uranium-oxide systems (Fig. 6.27)
[75, 82, 83]. The coprecipitated oxides were ignited in air to
1,200° C for 4 hours and the products studied by X-ray diffraction

and density determinations. At low rare-earth concentrations (0 to 25


mole percent), mixtures of UAO, and a fluorite phase were observed.
A single fluorite-type phase was obtained at rare-earth concentrations
between approximately 25 and 65 mole percent; at higher rare-earth
concentrations, mixtures of the fluorite phase and the rare-earth oxide
were observed. Hund and Peetz were able to show by density measure
ments that the fluorite phase is based on an essentially perfect cation
lattice, with anion vacancies or interstitial oxygen making up the defi
284 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
O/M RATIO
2.4 2.3 2.2 2.I 2.0 1.9 1.8
5.6 H. i I I I I i i -

l I I l l l l I I
O KO 20 30 40 50 60 70 80 90 IOO
Mole PER CENT (RE)ols

FIGURE 6.27. Fluorite Structure in the Rare-Earth-Oxide and Uranium-Oxide


System According to Hund and Peetz [75, 82,83].

cit or excess from the theoretical MO, ratio. It is evident that extra
polation of the fluorite-phase lattice parameters to 100 percent uranium
oxide, on the assumption that Vegard’s law is obeyed, does not lead to
a single O/U ratio for all of the systems. Estimated composition of
the uranium-oxide end member varies between UO, as and UO2.s.
Whether this is due to failure to achieve equilibrium, variations in ex
perimental conditions, or in properties of the respective oxide mixtures
is not known; however, the results on the uranium-yttrium oxide sys
tem referred to above indicate that equilibrium is not attained during
a 4-hour heating.
The oxygen-to-metal (U+R.E.) ratios given at the top of Fig. 6.27
are calculated on the assumption that the rare-earth metal remains
trivalent (true, except for praseodymium) and that the uranium va
lence is +5.2 (UO2.s). This assumption is in agreement with the ob
served composition of uranium oxide in air at 1,200°C, with weight
changes observed by Wilson on mixed oxides ignited in air, and with
the electrical measurements of Hund, et al. [74, 78]. The calculated
limits of the fluorite phase of 1,200° C then become approximately
MO, ss-MO.2s. The upper limit (MO...) is in reasonable agreement
with the corresponding phase boundary for pure uranium oxide. The
lower limit, in contrast to the pure uranium-oxide system, extends well
below the 1:2 ratio.
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 285

6.6.4 Systems with Group IV and Actinide Oxides

(a) Silicon Oaxide

Studies of the UO,-SiO, system at high temperatures have shown


no evidence for solid solution or compound formation (Fig. 6.28)
[32]. The discovery of a tetragonal uranium-silicate mineral iso
morphous with zircon (ZrSiO,) and thorite (ThSiO.) reopened
investigation of this system. Hoekstra and Fuchs succeeded in synthe
sizing USiO, by a hydrothermal technique at 250° C [84]. The
failure to observe this compound in high-temperature preparations
is explained by the fact that disproportionation to UO, and SiO,
occurs about 1,000° C.
2200
I I I I
TV- I T I

2000 H.
\ LIQUID
-
o
\ -
\
|- UO2 + LIQUID
•.
#
P
g isoo H SiO2+ —
u LIQUID
Gl.
-:
* - .*
1650+lo”

i800 H. -
UO2 + SiO2

1400 I I I I l l | | !
O 2O 40 60 8O IOO
U02 MoLE PER CENT side SiO2

FIGURE 6.28. UOz—SiO, System According to Lang, et al. [32].


(b)

Titanium. Owide
by

experiments
C of

series Lambertson and Mueller between 600


A

with the composition U.O, TiO, gave


no

and 1,500° evidence for


solid solution [80]. The small Ti” ion
or

in

compound formation
TiO, could expected show very limited solubility either UO.
be

to

in

U2Os.
or

(c) Zirconium Oaxide

and Mueller found extensive solid-solution regions


in

Lambertson
UOz-ZrO2 system [85]. approximately
at
A

the eutectic was observed


286 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

2,550° C in a composition containing 52.5 mole percent zirconia.


UO, will dissolve more than 40 per

C.,
Between 1,000° and 1,800°

20
cent ZrO2, while the tetragonal ZrO2 will dissolve about percent

C,
UO2. Above 1,800° however, solubilities increase rapidly and
the diphasic region nearly,
not entirely, eliminated.

It
if
was

is
suggested that ZrO2 changes structure another polymorphic modi

to
this temperature.
at
fication
Wolten also studied the UO,-ZrO2 system elevated temperatures

at
(Fig. 6.29) [86]. The absence miscibility gap

of
temperatures

at
a
cubic solid solution was found

in
above 1,900°

A
was verified.
C

all compositions containing mole percent

54

of
zirconium dioxide.

to
0
In mixtures containing slightly more than zir

of
mole percent

54
conium dioxide, the solid solution becomes tetragonal, with c/a

a
ratio very close unity. The tetragonal character solu

of
the solid
to

tion increases with increasing zirconium dioxide content.


3OOO

I
I

I
LIQUID
-

-
26 OO H

22 OO H. SOLID SOLUTION —
CUBIC TO TETRAGONAL
H.
uſ $’

1800
T.
E9 z
<–

º
-> Two-PHASE 33
REGION 38
º

400 :*
%#

FACE- 55
|

ul CENTERED-
H
5

H.
|

CUBIC SOLID ºn
SOLUTION
|

|
|

|OOO H.
|

|
|
|

600 H.
|
|

|
|

|
|

|
|

200 H. MONOCLINIC
SOLID SOLUTION
|

l
|
l
O

2O 4O 6O 8O IOO
O

MoLE PER CENT zro,

FIGURE 6.29. UOr-ZrO2 System According Lambertson and Mueller and Wol
to

ten [85, 86]. (Courtesy, American Chemical Society and The American Ce
ramic Society, Inc.)
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 287

A continuous solid solution between two compounds having dif


ferent structures is relatively unusual; however, Wolten points out
that both UO, and ZrO, have a fluorite-type structure [86]. The
zirconia lattice is slightly distorted to give a tetragonal axial ratio
of 1.018: 1. Since the U* and Zr” ionic radii are not too dissimilar,
one can expect complete miscibility of the oxides at elevated
temperatures.
The normal room-temperature structure of zirconium dioxide is
monoclinic. During heating, a reversible monoclinic-tetragonal
inversion occurs at about 1,000° C, with an accompanying large
dimensional change. It has been well established that the addition
of certain oxides, such as calcium oxide, magnesium oxide, and
yttrium oxide, to the zirconium dioxide results in the formation of
solid solutions having a cubic structure without inversions [87, 88].
In general, about 15 weight percent of other oxide is required to form
the stable structure. Wolten reported that 15 mole percent of UO2
was also sufficient to lower the inversion in zirconia to below room
temperature.
Hayter and coworkers reported that a tetragonal UO2–ZrO2 solid
solution containing only 10 mole percent of UO, could be prepared by
reaction of ZrO, and UO, at temperatures as low as 900° C if a very
intimate mixture of the reactants was obtained [89]. Coprecipita
tion of Zr(OH), and (NH,), U.O., with NH,OH from solutions of
the nitrates with subsequent drying, milling, and heating resulted in
a very intimate mixture of the oxides when the salts decomposed.
Reaction was complete after heating at 900° C in hydrogen for 4
hours, and no uranium dioxide lines could be detected in the X-ray
patterns.
The observations of Roberts, et al., are in general agreement with
is,

the results of the previous investigations, that mixtures contain


ing percent UO, monoclinic, those containing
10

15
to

to

mole are
5

per UO, tetragonal, containing


55

65
to

mole cent and those 100 mole


percent UO,
cubic [41]. They report that preparations containing
percent UO, tendency separate into two phases
30

55
to

to

mole show
a
C.
at

when annealed 1,100°


An X-ray study the UO,-ZrO, system by Voronov, al., reveals
of

et

several differences with the earlier work of Lambertson and Mueller


[85, 90]. Comparison Figs. 6.29 and 6.30 shows the following:
of

(1) The high-temperature solid solution decomposes not 1,900°C


at

1,675° C; (2) the UO,-ZrO2 face-centered cubic solid solu


at

but
produced the decomposition
of

of

tion which the initial


as

result
is

solid solution contains only about mole percent ZrO2; (3)


13

of

will UO,
or

tetragonal ZrO2 dissolve about 13.5 mole percent almost


by

Note,
of

two-thirds that indicated Lambertson and Mueller [85].


288 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
H-I-T-I-T-T—T-T—T-T—H-T—H
28 OO

26 OO

- | -
- | -
22 OO H. | 2 -
|
H FACE-CENTERED TETRAGONAL
|- CUBIC SOLID SOLUTION | SOLID SOLUTION
|

|
91800 H
- -

-
[-
#
#
:
5
-
FACE-CENTERED
+ TETRAGONAL
SOLID SOLUTION
SOLID SOLUTION -
#1400 H.
|- -
>
Liu
H. - -
- MONOCLINIC SOLID SOLUTION
+ TETRAGONAL SOLID SOLUTION
| OOO H. N
|-
MONOCLINIC SOLID SOLUTION

6 OO H. FACE-CENTERED SOLID SOLUTION


- + MONOCLINIC SOLID SOLUTION

—1%.
2OO H.

|-- H^1–1–
O 2O 40 60 80
MOLE PER CENT
ZrO2

FIGURE 6.30. UO2–ZrO, System. According to Woronov, et al. [90].

however, the reasonably close agreement between the work of Wolten


and Voronov, et al., for the high ZrO2 region [86, 90].
Voronov, et al., suggest that the differences between their work and
that of Lambertson and Mueller may be due to the fact that the latter
authors underestimated the long times required to achieve equilibrium.
In still another report on the UO,-ZrO., system, Evans concluded
the following: (1) At temperatures just below the eutectic (2,550° C)
about 45 percent ZrO2 is soluble in UO, and vice versa; (2) at 1,350°C
there is a mutual solubility of not less than 15 mole percent; (3) the
tetragonal form of zirconia cannot be stabilized by the addition of
UO, ; and (4) the tetragonal to monoclinic phase change of the zir
conia-rich solid solution takes place between 940° C and room tem
perature [91].
Ochs studied the UAOs–ZrO, system by coprecipitating U (VI) and
Zr(IV) from aqueous solution and igniting to 1,025° C [92]. No
new compounds were observed, nor was there any evidence for a
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 289

change in the ZrO, structure. Changes were observed in the UAOs


cell dimensions, i.e.,

U3Os: a=6.711 Å, b=3.980 Å, c=4.146 Å


U3O,--ZrO. : a = 6.761 Å, b=3.905 Å, c=4.125 Å.

This effect was ascribed to an increase in the equilibrium oxygen


pressure over U.O, caused by the presence of ZrO, to give a uranium
oxide phase of composition between UO, as and UO2.ss. Since the ob
served cell parameters were independent of zirconia content, the
changes were not attributed to solution of zirconia in the UAOs phase;
however, only three compositions were investigated (1:5, 1:1, 5:1).
If the solubility of ZrO2 in UAOs (or UAO,...) is less than 17 percent, no
change in lattice parameter should be expected. Extension of the
orthorhombic structure to UO, is has never been observed in pure
uranium oxides. Thus, it seems likely that at least a portion of the
change in orthorhombic cell dimensions is due to the addition of a
small amount of zirconia to the UAO, phase.
Roberts and his associates found that the cubic oxide containing
75 mole percent UO, oxidized to give cubic phases at first, but other
structures were obtained if oxidation was continued [41]. The limit
ing cubic phase had a composition of MO2.2s. The unit cell contracted
(5.395 to 5.339 angstrom units) and the density increased 2.8 per
cent during oxidation. A tetragonal phase containing 15 mole percent
UO, was oxidized to a limiting uranium oxidation number of 4.41 with
increase in density. The results were interpreted as indicative of in
terstitial oxygen in the oxidized samples. They suggest that oxida
tion in the UO,-ZrO2 system is less than in the UO,-ThC), system
because of the smaller size of the UO,-ZrO, unit cell. Further work
is necessary to resolve the discrepancy between this investigation and
the results reported by Ochs [92].

(d) Tin Oaxide

Lang, et al., heated mixtures of UO, and SnO2 in an inert atmos


phere and found no evidence for compound formation or solid solu
tion [32]. The rapid volatilization of SnO, above 1,400° C prevented
investigation of this system at higher temperatures.

(e) Cerium. Oride

Both CeO, and UO, crystallize in the fluorite structure and would be
expected to show extensive solid solution. Results of several investiga
tions indicate that a continuous solid solution of the oxides can be pre
pared at temperatures as low as 1,000° C in vacuum (Fig. 6.31)
[93, 94, 95]. Whether the observed deviation from Vegard's law is
290 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
547 I I I I i i I I i

- VEGARD'S LAW

546 H.
W

\
D- -
O ---O MAGNELl 8, KIHLBORG
O HUND,
H - et ol - -
\ m--s
-
fºº!?!?!

'N
C -- (SAMPLE HEATED
O LANG, et al.
IN AIR)

5.45

oq.
+
: 5.44
;
ul
92

:E
—l
5:43

542

5.4 |

5.40
l l l l I l I L 1—
O 2O 40 6O 8O IOO
UO2 MOLE PER CENT CeO2 CeO2

FIGURE 6.31. UO,-CeO, and UO,-CeO-O, Systems.

real or occasioned by traces of interstitial oxygen has not been deter


mined. Hund, Wagner, and Peetz found incomplete solubility in the
U3Os–CeO2 system; between 0 and 37 mole percent CeO2, a diphasic
region composed of UAOs and the fluorite phase is obtained [96].

(f) Lead Oaside

Hoekstra and Siegel found that PbO is partially reduced by heat


ing it with UO, above 600° C in an evacuated system. The remain
ing lead oxide is incorporated into a binary oxide containing partially
oxidized uranium in a fluorite-type lattice. The extent of the solid
solutions has not been determined, but a sample having the composi
tion Pb, U.O. has a cell edge of 5.61 angstrom units.

(g) Thorium, Oaxide

Research on the UO,-Tho, system showed that it is monophasic


over the entire composition range (Fig. 6.32) [97, 98, 99]. Mixtures

* H. R. Hoekstra and S. Siegel, unpublished work.


URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 291
5.600 i | H i I i I I

D-
- --
--
vEGARD'S LAW
DSLOWINSK AND ELLIOTT

O-O

al.
| O © LAMBERTSON, et
HUND AND NIESSEN

IN
(SAMPLES HEATED AIR)

5.450
|
I

4O 6O 8O IOO
UO2 Thoz
MOLE PER CENT Thoz

FIGURE 6.32. UOr-Tho, and UOr-Thor-O, Systems.

and thorium hydroxide form solid solutions


at of

ammonium diuranate
Lambertson, Muel
as

as

temperatures low hydrogen.


in

1,000°
C

Gunzel measured the liquidus and solidus temperatures for the


ler, and
UO,-ThC), system and found only minor variations from the theoreti
curves (Fig. 6.33) [99].
cal

Hund and Niessen found partial solution range (44 100 percent
to
a

the UAOs—Tho, system after ignition


at

ThO.) air
in

1,200°
in

The uranium-rich oxide compositions contained UAO, addi


in

[100].
the fluorite phase.
to

tion
Handwerk and associates prepared solid solutions similar
to

those
by

by

heating
of

obtained Hund and Niessen various mixtures uranium


U3Os) and thorium dioxide air
at

to
as

oxide (added 1,700°


in

[100–102]. X-ray examinations revealed that fluorite solid


C

1,750°
up

mixtures containing 59.5 mole percent


in

to

solutions were formed


phase, probably UAOs, was
of

uranium oxide. Evidence second


of

containing ura
of

compositions percent
in

found more than 59.5 mole


nium oxide.
by

Anderson,
on

Controlled oxidation experiments UOs–Tho.al.,


et

showing inter
as

interpreted oxygen into


of

mixtures were addition


positions the original lattice [103]. Low-temperature oxida
in

stitial
C)

phases very similar


of

(~200° the formation


to

to

tion leads those


57.4789 O–61–20
292. URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I I i I

33OO H.
--- ExPERIMENTAL
THEORETICAL

32OO

FACE-CENTERED CUBIC SOLID SOLUTION

28OO I l l I
O 2O 4O 6O 8O |OO
UO2 Thoz
MOLE PER CENT THORIA

FIGURE 6.33. UOz-Tho, System. According to Lambertson, et al. [99]. (Courtesy,


The American Ceramic Society, Inc.)

encountered in the uranium-oxygen system, i.e., U.O. Mixtures con


taining more than 50 mole percent thoria crystallize only in a cubic
structure even after oxidation under pressure; 22 mole percent thoria
is sufficient to maintain the cubic structure at any temperature under
atmospheric pressure. This minimum thoria concentration (22 per
cent) is considerably less than the limit observed by other investigators
(40 and 44 percent). Anderson, et al., attribute the difference to the
absence of any UAOs in his samples to nucleate the phase [103]. The
lattice parameter was observed to decrease during oxidation until the
uranium valence reached --5; further oxidation led to expansion of
the lattice. The results are interpreted as indicating that isolated
uranium cations with more than + 5 charges cannot exist in these crys
tals. Roberts, et al., have determined values of TIo, the partial molal
heat of solution of oxygen in solid solutions containing uranium at a
mean oxidation number of 4.35 (see Table 6.11) [41]. Aronson and
Clayton have obtained thermodynamic information on the solution of
oxygen in urania-thoria solid solutions with the electrochemical tech
nique that was employed to obtain information on uranium dioxide
(see Table 6.6) [104]. Partial molar free energies, entropies, and en
thalpies of solution of oxygen in the solid were calculated and are
shown in Table 6.12. It was found that the partial molar free energy
decreases negatively with increasing thorium content for thorium con
entrations greater than 30 atomic percent. The partial molar entropy
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 293

increases negatively with increasing oxygen content or with increas


ing thorium content. Aronson and Clayton discuss their results and
the results on uranium dioxide in connection with a mechanism of - in
terstitial solution of oxygen ions in the fluorite lattice [38].
The magnetic susceptibility of partially oxidized UO,-ThC), mix
tures was measured by Dawson and Roberts [105]. The rapid decrease
in susceptibility during the early stages of oxidation (Fig. 6.34) is
believed to result from oxidation to U(V) rather than U(VI).

TABLE 6.11—PARTIAL MOLAL HEATS OF SOLUTION OF OXYGEN IN


UO,.ms—Tho, MIXTURES [41]

Mole-ºu 100 25 5.97 3.3 1.87 1.2 0.76 0. 53


* - Hom(kcal/mole) 69.8 61 44 39 33 25 17.5 (14.5)

TABLE 6.12—PARTIALMOLAR FREE ENERGIES, ENTROPIES, AND


ENTHALPIES OF SOLUTION OF OXYGEN IN U, Thi-,O,4, [104]
y x —Fos(kcal/mole) —Son(e.u./mole) —Hou(kcal/mole)
(1,250°K)

0.90 0.042 51.8 8 62


0.90 0.042 52.4 13 69
0.90 0.076 48. 7 17 70
0.90 0.081 48. 4 17 70
0. 90 0. 123 44. 7 23 73
0. 90 0. 126 44. 7 21 71
0.90 0. 157 40. 7 29 77
0.71 0.046 50. 9 14 68
0 71 0.049 50. 9 14 68
0.71 0.084 47. 2 21 74
0.71 0. 092 46. 9 28 82
0.71 0. 121 44. 0 34 85
0 71 0. 133 42. 9 29 80
0.71 0. 152 40. 3 34 83
0.71 0. 154 40. 2 35 84
0. 52 0.044 47. 5 20 72
0. 52 0.045 46. 8 22 74
0. 52 0.079 43.4 22 72
0. 52 0.087 44. 4 31 83
0. 52 0. 119 39.8 35 84
0. 52 0. 122 39. 5 34 82
0. 52 0. 152 35. 7 40 85
0. 52 0. 152 35. 2 35 78
0.29 0.019 45.4 22 72
0.29 0.027 43. 2 19 67
0.29 0.046 37. 3 24 67
0.29 0.053 36. 2 26 69
0.29 0.053 36. 8 30 74
0.29 0. 069 33. 5 30 71
0.29 0.073 33. 1 34 76
294 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

3OOO

2OOO

|OOO

9 tºo 75. 50 25 O (A)


2 2.25 2.5 2.75 3. O ( 8)
4 4.5 5 5.5 6 (C)
FIGURE 6.34. Magnetic Susceptibility in the UOz–Thor-O, System According to
Dawson and Roberts (Scale A, Percent UO, ; Scale B, x in UO, ; Scale C,
Average Uranium Valency in UOz-Tho-O, System Containing 14.8 Percent
UO,) [105]. (Courtesy, The Chemical Society, London.)

(h) Neptunium, Oaside

Solid solutions of neptunium and uranium oxides were prepared by


Roberts, et al., by coprecipitation of the hydroxides and ignition in
air [41]. Oxidation beyond UO2–NpO, was evidenced by the fact that
º- URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 295

the cubic product was always smaller than that which


the

of
cell edge
expected from Vegard's-law plot between the two dioxides.
be

would

a
evidence was found for reaction between neptunium and the addi
No

oxygen, since the maximum oxygen incorporated

of
tional amount
Npo.-UO, solid solution was less than the oxygen
|

of
an

into content

a a
-
corresponding ThC), UO, solid solution. Attempts prepare

to
UAOs–Np,Os have

on
mixed-oxide phase based been unsuccessful;
amorphous products are obtained.
|

Plutonium
(i)

Oaside

Mulford and Ellinger report that continuous solid solution also


a
the UO,-PuC), system [106]. The lattice parameter varies
in

exists
linearly with composition between UO, (5.4700 angstrom units) and

by
Pu(), (5.3960 angstrom units).
The oxide mixtures were prepared
Pu (IV) and U(VI) ammoniacal solution, drying
in

toprecipitating
70°C, then firing for hydrogen. The
to

to

1,000° hours
in
C
at

preliminary drying was found necessary


to

obtain single-phase
product. a
UO,-PuO,
oxides ignited

at at
Russell observed that mixed argon
in
usually contain two fluorite phases, while samples ignited
C

1400°
1,700° contain single solid-solution fluorite structure.”
to

1800°
a

Evidence for oxidation was noted after ignition Arc


C.
air
in

at

500°
melted samples yielded multiphase products, one
of

which was cubic


a

large cell constant which decreased sharply


on
tell

with oxidation.
a

solid solution containing


as

The results were interpreted


of

indicative
a

Pu:0s. Structure types and lattice dimensions obtained from some


*

the

arc-melted mixtures are given Table 6.13.


in
ºf

TABLE 6.13—LATTICE PARAMETERS OF PHASES IN ARC-MELTED


UO,-PuC), MIXTURES*

Composition(mole %) Lattice parameters (Å)

Puſ), UO2 Pu2O3 PuC): Solid solution UO:


a

100 00 10. 978 5.391 ||------------|----------


80 20 10.391 5.391 ||------------|----------
70 30 11.048 5. 393 ||------------ 5.465

*
60 40 10.990 5. 424
|_

|- |- |-
_
_
_
_
_
_
_
_
_
_
_

- - -
- - -
- - -
- - -
- - -
- - -
- - -

- - -
- - -

40 60 ------------|------------ 5. 420
20 ------------------------
i

80 5.445
------------|------------|------------ 5.457
0

100

Russell,personal communication.
*
B.

"L.E. Russell, personal communication.


296 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

6.6.5 Systems with Group V Metal Oxides

(a) Vanadium Oaside

King and Suber heated mixtures of vanadium dioxide and uranium


dioxide in the composition range UO, : VO, to 1 UO, :4 VO, at 1,650°
C in argon for 2 hours [107]. Cubic solid solutions were obtained
in each instance, but free vanadium dioxide was also present in each
of the mixtures. Microscopic examination indicated a solid solution
similar to uranium dioxide. Melting temperatures of the solid solu
tions were greater than 1,650° C. Ignition of a UAOs–V.O. mixture in
airled to the formation of a compound whose composition was sug
gested to be 3V.O.2UAOs.
Lang, et al., briefly studied the UO,-V.O., system at temperatures
up to 900° C [32]. Excessive volatilization of the vanadium oxide
prevented investigation at higher temperatures. Reduction of V.O.
to VO, by UO, was observed, and two unknown phases, possibly
WO, UO, and 3VO,-UO, were found.

(b) Tantalum, Oaxide

Gasperin has investigated the UO,-Ta2O, system and found that


two compounds were formed at 1,200° C (in air?): U.Ta2O, and
UTa2O, [108, 109]. The former is obtained as a white powder with
a face-centered cubic structure (a = 10.37 Å). The postulated formula
is assigned by analogy with the pyrochlore M.Ta2O, (OH, F). The
suggested composition is probably incorrect, since colorless uranium
compounds are virtually unknown, and the valence requirements for
the metals (3+ and 4+, or 2+ and 5+) are improbable. The forma
tion of only trace amounts of the cubic oxide suggests the presence
of impurities in the original material.
The yellow oxide is reported to be hexagonal (a = 6.41+0.05 Å.
c=3.95+0.05 Å) with a Ta/U ratio of 1:9. The assigned formula
(UTa2Os) would indicate a stabilization of U (VI) by tantalum, since
the binary oxide is stable at 1,200° C. No conclusive evidence is pre
sented for the existence of a binary oxide containing U (IV) or U (V).

6.6.6 Summary

Melting points of binary mixtures of UO, and various oxides were


recently determined [110]. Table 6.14 summarizes the results of this
work. From the survey of the binary oxide systems of uranium it
becomes apparent that, with few exceptions, only the broad outlines
of their phase relationships, structures, etc., have been determined.
Table 6.15 summarizes the data on these systems. Ionic radii are
listed for all of the elements, although they have smaller significance
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 297

for the less electropositive metals [1, 2]. Several generalizations


can be made from the available information: (1) Oxides of Group
IV and Group III
elements having fluorite or related types of struc
ture show extensive solid-solution regions with uranium oxides. The
heavier elements of Group II
also form solid solutions, but to a lesser
extent. Where information on density and electrical properties is
available it indicates that the solid solutions are based upon an essen
tially perfect cation lattice. Within each group, elements with ionic
radii closest to that of uranium exhibit the greatest solubility in the
uranium oxides. (2) Oxides of Group and Group Ielements tend II
to form compounds rather than solid solutions with uranium oxides.
The compounds can be readily oxidized to uranates on ignition in air.
(3) The lighter elements, with small ionic radii, do not form com
pounds or solid solutions with uranium oxides; the apparent exception
in the silica system is due to the formation of the orthosilicate ion.
(4) Easily reducible oxides are converted to the metal or a lower oxide
with simultaneous oxidation of uranium.
Many investigations of the binary oxide systems of uranium have
been initiated in an attempt to stabilize tetravalent uranium towards
air oxidation. All have been unsuccessful; uranium is invariably
oxidized to a degree comparable with the pure uranium oxides

TABLE 6.14—MELTING POINTS OF BINARY MIXTURES OF UO,


AND WARIOUS OXIDES [110]

Composition (mole %) U/M ratio


from Melting point
chemical §§, Mode of melting
UO, Other oxide analysis

50 BeO------| 1/1.5------- 2,200+50 || Melts with difficulty, droplets.


1,900+50 | Droplets form with difficulty.
1,750+50 | Droplet formation.
1,850+30 | Droplet formation.
2,000+50 | Droplet formation.
1,940+30 || Melts with droplet formation.
1,770–£30 Surface melting.
1,480+30 || Melts easily.
2, 600+50 | Same as HeO.
------------------ Does not melt in arc.
<800 Melted during preliminary roasting.
900–1,000 Same as V2O5.
- 1,100–1,200 Same as V:Os.
1,850+30 || Melts easily.
1,800+50 | Melts little, evaporatesstrongly during heating.
1,850+50 | Same as BizO3.
------------------ Does not melt in arc, strong evaporation.
2,050+100|| Droplets form with difficulty.
1,370+30 | Melts easily.
1,650+30 || Melts easily.
2,700–2,750+100| Difficult to melt.

Reprinted from Nucleonics, Vol. 18, No. 7, July 1960, Copyright 1960, McGraw-Hill Publishing
Company, Inc.
298 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

(i.e., UO2.s). It has been possible to stabilize the fluorite structure in


the binary oxides, but only because the combined metal-to-oxygen
ratio is compatible with the fluorite-structure composition limits.
Solid-solution regions involving the fluorite structure have been of
greater interest and, consequently, have been studied in more detail

TABLE 6.15—BINARY OXIDE SYSTEMS OF URANIUM

Uranium oxidation state ***


Group Element Ionic radius (Å)"
IV V VI

I------- Li(I)-------------- 0. 68 N U
Na--------------- 0. 98 || C7 C U
K---------------- 1. 33 || N N U
Rb--------------- 1. 48 l----------|---------- U
Cs---------------- 1. 67 |----------|---------- U
Cu(II)------------ (0.70) || R. ----------
II- -- -- - Be(II)------------- 0.30 | N |_ _ _ _ _ _ _ _ _ _
Mg--------------- 0.65 | N C?, S U
Ca--------------- 0.94 C, S C?, S U
Sr---------------- 1. 10 | S S U
Ba--------------- 1. 29 || S S U
Zn---------------- **0. 74 |- - - - - - - - --|- - - - - - - - - - U
Cd--------------- **0. 97 |_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ U

III - - - - - Al(III)------- ----- 0.45 | N S


Sc.---------------- 0. 68 |- - - - - -- - -- S
Y---------------- 0.88 || S S
R.E.-------------- 1. 04–0, 85 | S S

IV------ Si(IV) - - - - - - - - - - - - 0.38 (C) |- - - - - - - - - -


Ti---------------- 0.60 |- - - - - - - - - - N
Zr---------------- 0.77 S S
Ce--------------- 0.92 || S S
Th--------------- 0. 99 || S S
NP--------------- 0.92 || S S
Pu--------------- 0. 90 | S S
Sn---------------- **0. 71 || N |_ _ _ _ _ _ _ _ _ _
Pb(II) ----- - -- ---- **1. 20 | R C, SY U

V------ V(V) ------------- **0, 59 || R. |_ _ _ _ _ _ _ _ _ _


Ta--------------- **0. 68 || C7 |_ _ _ _ _ _ _ _ _ _

*From Zachariasenunless otherwise noted [2].


**From Pauling [1].
“Uranium ionic radii: U(IV) 0.97A[1], 0.93Å, U(v) 0.87Å, U(vi) 0.83Å.
N = No compound or solid solution.
C=Compound formed.
S=Solid-solution region formed.
U= Uranate formed.
R= Element reduced by U(VI).
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 299

than those of orthorhombic UAOs. On the basis of very meager infor


mation, solution of metal oxides in the orthorhombic structure ap
pears to be less extensive than in the fluorite structure of UO2. This
lack of solubility in the UAOs structure may be due to the uranyl-type
bonding, which has more than one metal-oxygen bond distance, or to
the inherent inability of the orthorhombic structure to accommodate
less than about 2.6 anions per cation.
The fluorite structure is more susceptible to solid-solution formation
and is retained from an O/M ratio of 2 to approximately 2.25 at
1,000°C in the uranium-oxygen system. The upper limit is virtually
unchanged in the binary oxides, but the lower limit extends approxi
mately to 1.85 as divalent and trivalent metal oxides are added to
UO2. Thus, the lower limit established in the uranium-oxygen system
appears to represent a property of the uranium ion itself rather than
inability of the fluorite structure to exist as a stable lattice with less
than two oxygen atoms per metal atom.
In general, the binary oxide systems appear to be much less complex
than the simple uranium-oxygen system or other multivalent oxide
systems, such as praseodymium or titanium. This simplicity may,
however, be only apparent. Further study may reveal the presence
of comparable structures in the binary oxide systems as well, for ex
ample, the MO2.25 and MO, as of uranium and the MO, sa and MO,.11
compositions of praseodymium. The rare-earth type C structure has
already been found in CalVO, and may be discovered in other mixed
oxides as well. Perovskite (ABOs) and pyrochlore (A.B.O.) struc
tures have been tentatively identified in several instances. Further
investigation may lead to the discovery of additional examples of
these structure types as well as to several additional types.
Are the extensive solid-solution regions actually single phase re
gions, or will further study show the presence of several closely related
structures? Most studies on binary oxide systems have been made at
high temperatures to achieve equilibrium rapidly. The samples are
then quenched or cooled rather rapidly for investigation. Would an
nealing result in crystallization of ordered structures from the solid
solutions? These are some of the questions that will be answered by
future studies.

REFERENCES

1. L. PAULING, “The Nature of the Chemical Bond,” 3d ed., Cornell University


Press, Ithaca, New York, 1960.
2. W. H. ZACHARIAseN, “Crystal Chemistry of the 5f Elements” in “The Actinide
Elements,” G. T. Seaborg and J. J. Katz, eds., NNES Division IV, Vol. 14A,
McGraw-Hill Book Co., New York, 1954.
3. B. G. CHILDs, “The Cohesive Energy of Uranium Dioxide and Thorium Diox
ide,” CR Met—788, Aug. 1958.
300 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

4. H. W. CRANDALL, “The Formula of Uranyl Ion,” J. Chem. Phys. 17, 602–6ſº


(1949).
5. W. H. ZACHARIAsen, “Crystal Chemistry of Uranyl Compounds,” Acta Cryst.
1, 795–799 (1954).
6. J. J. KAtz and E. RABINow ITCH, “The Chemistry of Uranium,” NNES Div.
VIII, Vol. 5, pp. 246—247, McGraw-Hill Book Co., New York, 1951.
7. R. E. RUNDLE, N. C. BAENZIGER, A. S. WILSON, and R. A. McDon ALD, “Struc
tures of the Carbides, Nitrates and Oxides of Uranium,” J. A.m. Chem. Soc.
70, 99–105 (1948).
8. R. F. DICKERson, A. F. GERDs, and D. A. WAUGHAN, “Metallographic Identifi.
cation of Nonmetallic Inclusions in Uranium,” J. Metals 8, 456–460 (1956).
9. D. A. WAUGHAN, C. W. MELTON, and A. F. GERDs, “Experiments on the Prepara
tion of UOs-, and UO,” BMI–1175, Mar. 6, 1957.
10. W. H. ZACHARIAseN, “X-Ray Diffraction Results,” N–1973, Apr. 24, 1945.
11. H. HERING and P. PERIo, “Equilibria of Uranium Oxides between UOs and
Bull soc. chim. France 19, 351–357 (1952).
U3Os,”
12. S. ANDERson, J. O. SAwYER, H. W. WoRNER, G. M. WILLIs, and M. J.
J.
BANNISTER, “Decomposition of Uranium Dioxide at its Melting Point,"
Nature 185, 915–916 (1960).
13. R. J. ACKERMANN, “The High-Temperature, High-Vacuum Vaporization and
Thermodynamic Properties of Uranium Dioxide,” ANL–5482, Sept. 14, 1955.
14. J. S. ANDERson and J. O. SAwYER, “The Stability of Uranium Dioxide in Hy
drogen at High Temperature,” Proc. Chem. Soc., 145–146 (1960).
15. L. G. WISNYI and S. W. PIJANowski, “The Thermal Stability of Uranium Di
oxide,” KAPL–1702, Nov. 1, 1957.
16. H. R. HoekstEA and S. SIEGEL, “The Uranium-Oxygen System: UAOs—UO,” in
“Proceedings of the Second United Nations International Conference on
the Peaceful Uses of Atomic Energy, Geneva, 1958,” Vol. 28, pp. 231–234,
United Nations, Geneva, 1958.
17. G. W. WAtt, S. L. AcHoRN, and J. L. MARLEY, “Some Chemical and Physical
Properties of Uranium Peroxide,” J. Am. Chem. Soc. 72, 3341–3343 (1950).
18. W. BILtz and H. MUELLER, “On Uranium Oxide,” Z. anorg u. allgem. Chem.
163, 257–296 (1927).
19. P. Jolibois, “A New Oxide of Uranium, UAO,” Compt. rend. 224, 1395–1396
(1947).
20. F. GRoNvold and H. HARALDSEN, “Oxidation of UO,” Nature 162, 69–70
(1948).
21. K. B. ALBERMAN and J. S. ANDERsoN, “The Oxides of Uranium,” J. Chem. Soc.
Suppl., 5303–5311 (1949).
22. E. ZINTL and U. CROAtto, “Fluorite Lattice with Anion Vacancies,” Z, anoru.
u. allgem. Chem. 242, 79–86 (1939).
23. A. BotſI.LÉ, R. JARY, and M. Do MIN É BfRGīs, “An X-ray Study of the Ap
pearance of Grain in Uranium Oxides,” Compt. rend. 229, 214–216 (1949):
“The Oxides of Uranium of Variable Composition,” 223, 1281–84 (1951).
24. P. PERIo, “Oxidation of Uranic Oxide at Low Temperatures,” Bull, soc. chim.
France 20, 256–263 (1953).
25. P. PERIO, “Observations on Uranium Oxides formed between UO, and U.O.,"
Bull, soc. chim. France 20, 840–841 (1953).
26. J. S. ANDERson, “Recent Work on the Chemistry of Uranium Oxides.” Bull.
soc. chim. France 20, 781–788(1953).
27. J. S. ANDERSON, L. E. J.
Roberts, and E. A. HARPER, “The Oxides of Urani
um VII, The Oxidation of Uranium Dioxide,” J. Chem. Soc., 3946–3959
(1955).
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 301

F. GRoNvold, “High Temperature X-ray Study of Uranium Oxides in the


UO-U,0s Region,” J. Inorg. Nuclear Chem. 1, 357–370 (1955).
B. E. SchANER, “Metallographic Determination of the UOz-U.O., Phase
Diagram,” J. Nuclear Materials 2, 110–120 (1960).
A. ARRoTT and J. E. GoLDMAN, “Magnetic Analysis of the Uranium-Oxygen
System,” Phys. Rev. 108, 948–953 (1957).
31. D. A. WAUGHAN, J. R. BRIDGE, and C. W. SchwArtz, “Comparison of Active
and Inactive Uranium Dioxide-Oxygen Systems,” BMI-1241, Dec. 10, 1957.
S. M. LANG, F. P. KNUDseN, C. L. FILLMoRE, and R. S. Roth, “High-Tem
perature Reactions of Uranium Dioxide with Warious Metal Oxides,”
NBS—568, Feb. 20, 1956.
H. R. Hoekstra and S. SIEGEL, “Recent Developments in the Chemistry of
the Uranium-Oxygen System” in “Proceedings of the International Con
erence on the Peaceful Uses of Atomic Energy, Geneva, 1955,” Vol. 7, pp.
394–400, United Nations, New York, 1956.
S. ARoN son, R. B. Roof, JR., and J. BELLE, “Kinetic Study of the Oxidation
of Uranium Dioxide,” J. Chem. Phys. 27, 137–144 (1957).
R. K. WILLARDsoN, J. W. Moody, and H. L. GoFRING, “The Electrical Prop
erties of Uranium Oxides,” J. Inorg. Nuclear Chem. 6, 19–33 (1958).
P. E. BLACKBURN, “Oxygen Dissociation Pressures over Uranium Oxides,”
J. Phys. Chem. 63, 897–902 (1959).
P. E. BLAckBURN, J. WEIssBART, and E. A. GULBRANSEN, “Oxidation of Ura
nium Dioxide,” J. Phys. Chem. 62, 902–908 (1958).
S. AroN son and J. BELLE, “Nonstoichiometry in Uranium Dioxide,” J. Chem.
Phys. 29, 151–158 (1958).
S. SIEGEL, “Crystal Symmetry in the Phase Region UO-U.Oo,” presented at
the American Crystallographic Association meeting, Milwaukee, June
1958.
R. J. Acker MANN and R. J. THoRN, “Partial Pressure of Oxygen in Equili
brium with Uranium Oxides,” presented at the American Chemical So
ciety meeting, San Francisco, Apr. 1958.
. L. E. J. Roberts, A. G. Adwick, M. H. RAND, L. E. RUSSELL, and A. J.
WALTER, “The Actinide Oxides” in “Proceedings of the Second United
Nations International Conference on the Peaceful Uses of Atomic Energy,
Geneva, 1958,” Vol. 28, pp. 215–221, United Nations, Geneva, 1958.
J. BELLE, “Properties of Uranium Dioxide" in “Proceedings of the Second
United Nations International Conference on the Peaceful Uses of Atomic
Energy, Geneva, 1958,” Vol. 6, United Nations, Geneva, 1958.
43. H. R. HoekstEA, A. SANToRA, and S. SIEGEL, “The Uranium-Oxygen Sys
tem: UO-U-Or,” presented at the American Chemical Society meeting,
Atlantic City, Sept. 13–18, 1959.
W. H. ZACHARIASEN, “Crystal Structure Section” in “Report for Month
Ending Jan. 15, 1945,” CK–2667 and CC–2768 Mar. 12, 1945.
. F. GRoNvold, “Crystal Structure of Uranium Oxide (U.O.),” Nature 162,
70 (1948).
. H. R. HoF.KsTRA, S. SIEGEL, L. H. Fuchs, and J. J. KAtz, “The Uranium
Oxygen System : UO2.5 to UAOs,” J. Phys. Chem. 59, 136–138 (1955).
. A. F. ANDRESEN, “The Structure of U20s Determined by Neutron Diffrac
tion,” Acta Cryst. 11, 612–614 (1958).
. H. HARALDSEN and R. BAKKEN, “Magnetic Properties of Uranium Oxides,”
Natururiss. 28, 127 (1940).
. R. M. BERMAN, “The Role of Lead and Excess Oxygen in Uraninite,”
Am. Mineralogist 42, 705–731 (1957).
302 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

50. W. B. WILsoN, “Application of Ultrahigh-Pressure High-Temperature


Equipment to Study of UO, Reactions,” BMI-1328, Mar. 18, 1959.
51. W. G. MIxtER, “Heat of Formation of Oxides of Vanadium and Uranium,"
Z. anorg. u. allgem. Chem. 78, 221–238 (1912).
52. W. BILTz and C. FENDIUs, “Systematic Affinity Principle XLVIII, Heats
of Formation of UCl4, UCl and UOs,” Z. anorg. u. allgem. Chem. 176, 49–63
(1928).
53. L. BREweR, L. A. BromlEY, P. W. GILLEs, and N. L. LofgREN, “The
Thermodynamic Properties and Equilibrium at High Temperatures of
Uranium Halides, Oxides, Nitrates, and Carbides,” MDDC 1543, Sept.
20, 1945.
54. W. M. Jones, J. GoRDON, and E. A. LoNg, “The Heat Capacities of Uranium,
Uranium Trioxide and Uranium Dioxide from 15° K to 300° K,” J. Chem.
Phys. 20, 695–699 (1952).
55. E. J. HUBER, JR., C. E. Holley, J.R., and E. H. MEIERKoRD, “Heats of Com
bustion of Thorium and Uranium,” J. Am. Chem. Soc. 74, 3406–3408
(1952).
56. L. BREweR, “The Thermodynamic Properties of the Oxides and Their
Vaporization Processes,” Chem. Revs. 52, 1–75 (1953).
57. J. P. Cough LIN, “Heats and Free Energy of Formation of Inorganic
Oxides,” U.S. Bureau of Mines Bull. 542 (1954).
58. A. GLAssNER, “The Thermochemical Properties of Oxides, Fluorides and
Chlorides to 2500° K,” ANL 5750, 1957.
59. C. WAGNER, “Thermodynamics of the System Uranium-Oxygen,” WAPD
144, July 23, 1955.
60. J. ACKERMANN, R. J. THoRN, C. ALExANDER, and M. TETENBAUM, “Free
R.
Energies of Formation of Gaseous Uranium, Molybdenum, and Tungsten
Trioxides,” J. Phys. Chem. 64, 350–355 (1960).
61. J. BELLE and B. Lust MAN, “Properties of UO,” in “Fuel Elements Con
ference, Paris,” TID–7546, Book 2, pp. 442–515, Mar. 1958.
62. D. W. Osbor NE, E. F. WESTRUM, JR., and H. R. Lohr, “Heat Capacity and
Thermodynamic Functions of U,O, from 5 to 310° K,” J. A. m. Chem. Soc.
79, 529–530 (1957).
63. F. K. HEUMANN, O. N. SALMoN, and E. A. WILK, “Reactor Fuels Suspended
in Liquid Metals. II. Fuel Preparation and Capsule Studies of Liquid
Metal-UO, Slurries,” KAPL–1672, June 12, 1957.
64. W. Rudorff and H. LEUTNER, “On Lithium and Sodium Uranate (V),” Z.
amorg. u. allgem. Chem. 292, 193–196 (1957).
65. L. F. EPstEIN and W. H. How LAND, “Binary Systems of UO, and Other
Oxides,” J. Am. Ceram. Soc. 36, 334–335 (1953).
66. R. P. BUDNIkov, S. G. TREsvatsky, and V. I. KUSHAkovsky, “Binary Phase
Diagrams UO-Al. Oa, UO-BeO, UO-MgO" in “Proceedings of the Second
United Nations International Conference on the Peaceful Uses of Atomic
Energy, Geneva, 1958,” Vol. 6, pp. 124–131, United Nations, Geneva, 1958.
67. W. A. LAMBERTson and M. H. MUELLER, “Uranium Oxide Phase Equilibrium
Systems. II. UO-MgO,” J. Am. Ceram. Soc. 36, 332–334 (1953).
J. Joh Nso III. The
I.).

N,

S. ANDERSox and K.
of

“The Oxides Uranium.


B.

68.
System UOz–MgO—O,” Soc., 1731–1737 (1953).
J.

Chem.
KAtz, “Studies
H.

the Alkaline Earth Diuranates,”


on

Hoekstra and
H. A. R.

69.
J.
J.

m. Chem. Soc. 74, 1683–1690 (1952).


J.

ZACHARIASEN, “Crystal
of
of

70. W. Chemical Studies the 5f-Series Elements.


XXI. The Crystal Magnesium Orthouranate,” Cryst.
of

Structure Acta
7,

788–791 (1954).
URANIUM-OXYGEN AND BINARY OXIDE SYSTEMS 303

71. K. B. ALBERMAN, R. C. BLAKEY, and J. S. ANDERSON, “The Oxides of Ura


nium. Part II. The Binary System UO2–CaO,” J. Chem. Soc., 1352–1356
(1951).
72. S. C. FURMAN, “A Study of the Interaction of Barium Metal and Barium
Oxide with Uranium Dioxide,” KAPL–1664, Jan. 1957.
73.

LAMBERTson and M. H. MUELLER, “Uranium Oxide Phase Equilibrium


A.

W.
Systems. UO-Al2Oa,” Am. Ceram. Soc. 36, 329–331 (1953).

J.
I.
74.

WILsoN, “Stabilization UO, by Valence Compensation,” BMI-1318,

of
B.

W.
4,

Feb. 1959.
HUND and U. PEEtz, “Further Fluorite Phases the Mixed Oxides of

in
F.

75.
the Rare Earths with Uranium. Investigations the Systems LazOs,

u. of
Nd2O, Sm,0s, Yb2Os, Sc.Oa, with UAOs,” anorg. allgem. Chem. 271,

Z.
6–16 (1953).
76.

ANDERson,
FERGUsoN, and Roberts, “Anionic Vacancies

J.
F. in S.

E.
F.

L.
J.

I.

Fluorite-type
Oxides,” Inorg. Nucl. Chem.

1,
340–341 (1955).
J.

G. T. Fogg, “The Oxides Part VIII. The

of
FERGUsoN and Uranium.
P.

77.
I.

System Uranium Dioxide-Yttria,” Chem. Soc., 3679–3681 (1957).


J.
78.

HUND, PEETz, and KotteNHAHN, “The System. Yttrium Oxide-Ura


u. G.
U.
F.

nium Oxide,” anorg. allgem. Chem. 278, 184–191 (1955).


Z.

H. Cottrell, “Theoretical Structural Metallurgy,”


79.

52, Edward Arnold,


p.
A.

Ltd., London, 1955.


80.

LAMBERTson and M. H. MUELLER, “Uranium Oxide Phase Equilibrium


A.

W.
Systems: UO-Nd2Oa; VI. UO,Os-MgO VII. U20s–TiO2,” ANL-5312,
V.

Sept. 14, 1954.


81.

PLoetz, MUccIGRosso, M. OsikA, and W. JAcoby, “Dyspros


T.

R.
A.
G.
L.

L.

ium Oxide Ceramics,” Am. Ceram. Soc. 43, 154–159 (1960).


J.
82.

PEETz, “The Fluorite Phases the System Praseodymium


U.

HUND and
in
F.

Oxide,”
Z.

Oxide-Uranium Elektrochem. 56, 223–228 (1952).


83.

PEETz, “The System Uranium Oxide-Erbium Oxide,”


U.

HUND and Z.
F.

anorg. allgem. Chem. 267, 189–197 (1952).


u.

Fuchs, “Synthesis Coffinite—USiO,”


of
H.

Hoekstra and
R.

L.
*4

Science
H.

123, 105 (1956).


85.

MUELLER, “Uranium Oxide Phase Equilibrium


H.
A.

W. LAMBERTson and M.
System. III. UO-ZrO2,” m. Ceram. Soc. 36, 365–368
J.

(1953).
A.
86.

WolteN, “Solid Phase Transitions


M.

the UO,-ZrO., System,”


J.
G.

Am.
in

Chem. Soc. 80, 4772–4775 (1958).


87.

YAvorsky, “Effects
on
of

GELLER and Some Oxide Additions the


P.
F.

J.
R.

Thermal Length Changes Zirconia,” Research Nat. Bur. Standards


of

J.

35, 87–110 (1945).


MAUER, and M. SchwARTz, “Observa
8.

H.
C.

A.

A.

WEBER, GARRETT,
F.
J.
B.

tions on the Stabilization of Zirconia,” m. Ceram. Soc. 39, 197–207


J.
A.

(1956). -
89.

HAYTER, SMITH, BADMAN,


H.
C.
R.

R.

R.
A.

BARD, DUM Rose,


J.

E.

W.
S.

BERTINo, and KIRCHER, “Development Batch Process for


A.

of
P.
J.

J.

ZrO2–UO, Solid Solution,” LA-2138, June 1957.


of

the Production
A. a

Voitekhova, and DANILIN, “Phase Equilibrium


M.

Vokonov,
*)

A.
E.
N.

S.

Diagrams the UO-ZrO, and Tho.-ZrO, Systems” “Proceedings


of

of
in

the Second United Nations International Conference on the Peaceful Uses


Atomic Energy, Geneva, 1958,” Vol. United Nations,
6,

pp. 221–225,
of

Geneva, 1958.
91.

Evans, “The System UO-ZrOs,” mer. Ceram. Soc. 43, 443–447


J.
E.
P.

(1960).
92.

Ochs, “The U.O.-ZrO, System,”


L.

Naturforsch. 128, 215–222 (1957).


Z.
304 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

93. A. MAGNELI and L. KIHLBORG, “Cerium Dioxide-Uranium Dioxide System


and Uranium Cerium Blue,” Acta Chem. Scand. 5, 578–580 (1951).
94. W. RUDORFF and G. VALET, “The Constitution of Cerium-Uranium Blue and
the Mixed Crystals of the System CeO,-UOI-U3Os,” Z. anorg. u. allgem.
Chem. 271, 257–272 (1953).
95. G. BRAUER and R. TIEssl.ER, “The Chemical Composition of the Cerium
Uranium Blue Precipitated from Aqueous Solutions,” Z. amorg. u. allgem.
Chem. 271, 273–280 (1953).
96. F. HUND, R. WAGNER, and U. PEETz, “Anomalous Mixed Crystals in the
System Cerium Dioxide-Uranium Oxide,” Z. Elektrochem. 56, 61–65
(1952).
97. W. TREzeBIAtowski and P. W. SELwood, “Magnetic Susceptibilities of
Urania-Thoria Solid Solutions,” J. Am. Chem. Soc. 72, 4504–4506 (1950).
98. E. Slow INSKI and N. ELLIott, “Lattice Constants and Magnetic Suscepti
bilities of Solid Solutions of Uranium and Thorium Dioxides,” Acta
Cryst. 5, 768–770 (1952).
. W. A. LAMBERTson, M. H. MUELLER, and F. H. GUNzEL, Jr., “Uranium Oxide
Phase Equilibrium Systems: IV. UOz-Thoi,” J. Am. Ceram. Soc. 36,
397–399 (1953).
100. F. HUND and G. NIEssen, “Anomalous Solid Solution in the System Thorium
Oxide-Uranium Oxide,” Z. Electrochem. 56, 972–979 (1952).
101. J.H. HANDwerk, L. L. ABERNATHY, and R. A. BACH, “Thoria and Urania
Bodies,” Am. Ceram. Soc. Bull. 36, 99–100 (1957).
102. J. H. HANDwerk, C. L. HoeNIG, and R. C. LIED, “Manufacture of the
Thor-UO, Ceramic Fuel Pellets for Bor Ax–IV,” ANI, 5678, Aug. 1957.
103. J. S. ANDERson, D. N. EDGINgtoN, L. E. J. Roberts, and E. WAIT, “Oxides of
Uranium : IV. System UOz-Tho-O,” J. Chem. Soc., 3324–3331 (1954).
104. S. ARONSoN and J. C. CLAYtoN, “Thermodynamic Properties of Nonstoichio
metric Urania-Thoria Solid Solutions.” J. Chem. Phys. 32, 1749–1754
(1960).
105. J. K. DAwson and L. E. J. Roberts, “Magnetochemistry of the Heaviest
Elements. Part IX. The System UO,-Tho,0,” J. Chem. Soc., 78–80
(1956).
106. R. N. R. MULFoRD and F. H. ELLINGER, “UO,-Pu(), Solid Solutions,” J. A. m.
Chem. Soc. 80, 2023 (1958).
107. B. W. KING and L. L. SUBER, “Some Properties of the Oxides of Vanadium
and Their Compounds,” J. A. m. Ceram. Soc. 38, 306–311 (1955).
108. M. GASPERIN, “Synthesis and Identification of a Double Oxide of Tantalum
and Uranium,” Compt. rend. 243, 1534–1536 (1956).
109. M. GASPERIN, “Crystallographic Study of a New Double Oxide of Tantalum
and Uranium,” Compt. rend. 244, 1225–1226 (1957).
110. S. G. TrESVYATsKIY and V. I. KUSHARovskIY, Atomnaya Emergiya 8, No. 1.56
(1960); Nucleonics 18, No. 7, 101 (1960).
Chapter 7

SOLID STATE REACTIONS OF URANIUM


DIOXIDE
J. BELLE, Editor

7.1 INTRODUCTION

Included in this chapter are a number of topics which exemplify


rate processes in the solid state: diffusion, sintering, and metal-oxide
reactions. Self-diffusion (both cation and anion) in uranium dioxide
is discussed first, followed by a review of the sintering behavior of
U0. Finally, information on the solid state reactions of uranium
oxides with zirconium and aluminum is summarized.

7.2 SELF-DIFFUSION KINETICS IN URANIUM DIOXIDE


J.Belle and A. B. Auskern

7.2.1 Introduction

Although a large mass of data exists for self-diffusion in metals


and certain ionic crystals (principally halides), self-diffusion in
oxides has only recently been investigated [1–21].
There are two independent methods for determining diffusion coef
ficients of constituent ions in oxides: (1) indirectly by electrical con
ductivity and transference number measurements; (2) directly by
isotopic tracer diffusion measurements. Where electrical conduction
in the oxide is essentially ionic, as for example in ZrO2–CaO, use of
the

first method has been successful [19, 22]. For oxide systems
where electrical conductivity primarily due electron
or
to

electron
is

conduction, the first method generally inapplicable. Measure


is

hole

its electrical properties have shown that UO,


an
of

intrinsic
is

ments

semiconductor (see Chap. 5).


Better success with oxide systems has been obtained through use
of

given parti
is,

self-diffusion, that
In

of

isotopic tracers. the diffusion

"The related problem fission gas diffusion Chap.


of

9.

uranium dioxide discussed


in

in
is

305
306 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

cles in a medium containing the same particles, the diffusing particles


and the diffusion medium cannot be distinguished. Nevertheless, a
very close approximation can be made by the use of different isotopes
of the same element where the differences in atomic mass or radio
activity are used to trace the displacements of the chemical species
that the tracers accompany.
The methods ordinarily used in the study of self-diffusion in solids
with the aid of isotope tracers are of two kinds: The first is based on
the phenomena of diffusion that are produced between solids in direct
contact; the second is an application of the heterogeneous isotopic ex
change reaction. The various techniques (sectioning, surface count
ing, and autoradiography) for the former method usually depend
upon the use of a radioactive tracer; for the isotope exchange tech
nique, a stable isotope can be used [23]. Both these basic methods
have been used in self-diffusion measurements in uranium dioxide.
Results of these measurements are presented and discussed in Sects.
7.2.2 and 7.2.3.

7.2.2 Uranium Ion Self-Diffusion in UO,

(a) Earperimental Methods and Theory

Auskern and Belle have measured self-diffusion coefficients for the


uranium ion (U") in essentially stoichiometric UO, by using two ex
perimental techniques: the methods of sectioning and surface activity
decrease [20, 21].
In both methods, the instantaneous source and semi-infinite solid
geometry were approximated by depositing, through evaporation in
vacuo, a thin layer of highly enriched U*O, onto the plane surface
of a sintered, high-density natural UO, pellet. In the sectioning tech
nique, diffusion coefficients were determined by analysis of the U"
concentration gradient after diffusion anneals in a hydrogen atmos
phere. The gradient was determined by grinding off thin layers from
the compact surface and analyzing the material for U*—U* content.
The weakness in this method was the inability to obtain well-defined
concentration gradients because of the difficulty in removing uniform
and thin enough layers from the compact surface by grinding.
The solution to the diffusion equation for the semi-infinite solid with
an instantaneous source of diffusing material is well known [1]. The
equation can be given as

C=
sº exp (–a"/4Dt) Eq. (7.1)

where c is the concentration of the diffusing specie and c, is a constant


representing the product of the tracer layer thickness times the concen
SOLID STATE REACTIONS OF URANIUM DIOXIDE 307

tration of tracer at time t-0


and a = 0. Equation (7.1) in the form
log c-0.1086 a.”/Dt-H log k shows that a plot of log concentration
against diffusion distance squared (log c vs a “) is a straight line.
In the surface activity decrease method, advantage is taken of the
difference in alpha activity” between natural UO, and UO, enriched
in U*. From Eq. (7.1) the concentration at the surface (a =0)
as a function of time is simply c=c./V-Dt; however, the variation of
activity at the surface as a function of time is a more complicated ex
pression. Since alpha particles have an effective range in matter, some
alphas that originate from within the solid will be registered. Thus,
the amount of alpha radiation registered at the surface will be greater
than the activity caused by the concentration of diffusing material at
the surface. This problem was first considered by von Hevesy and
Seith and is discussed by Wahl and Bonner [24, 25]. The solution, the
derivation of which may be found in Wahl and Bonner, is

Ao
A=Ao erf (£)— (1—e-t”) Eq. (7.2)
tyr

where A is the activity at the surface after a diffusion anneal and Ao

is the initial activity [25]. The function & is defined as


aft , where

R is the effective alpha particle range in the material. The range R


of an alpha particle of a particular energy can be calculated by the
method discussed in Friedlander and Kennedy [26]. Although the
range for the 4.8 Mev alpha from U* in dense UO, is about 10 mi
crons, the effective range of the alpha particle depends on the particu
counting technique used. the experiments carried out by
In
lar

Auskern and Belle, scintillation technique employing ZnS


as
a
a

the UO, couple was separated


of

Scintillator was used [21]. The face


by

by

from the ZnS small air space and thin aluminized Mylar
a

Thus, alpha particle must


an

window over the crystal.


be
to

counted
through
of

have certain residual energy after passing thickness


a

U0, an air space, and the aluminized Mylar. Auskern and Belle
estimated that the effective range the 4.8 MeV alpha UO, was
of

in

3.75+ 0.25 microns [21]. Radiation originating from distance


a

greater than the effective range cannot


be

counted.
Diffusion coefficients were calculated from the initial surface activ
ity, the final surface activity, the effective alpha range UO2, the
in

anneal time, and the functional relationship given Eq. (7.2).


in

Diffusion anneals were carried out for times varying from


a 48
to
5

flowing hydrogen. Because UO, has


an

atmosphere
of
in

hours
significant vapor pressure (see Chap. the annealing temperatures
5)
at

*The principal alpha radiation the 4.8 Mew alpha from U24.
is

57.4789 O—61–21
308 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

(1,450° to 1,785° C), loss of the active UO, layer by evaporation must
be prevented. To this end, diffusion anneals were made with the active
faces of two diffusion couples face-to-face. In an arrangement of this
type it is assumed that there may be a transfer of material from one
face to the other, but that there is no overall loss of material from the
edge of the interface. It has been shown that if material transport
takes place through evaporation across the couple interface the results
can be analyzed by using the combined activities of the two couples
|27].

(b) Results and Discussion

Experimental results of uranium ion self-diffusion coefficients in


UO, obtained by the surface activity decrease method are shown in
Table 7.1 and are plotted in Fig. 7.1. A least squares analysis of
these data resulted in the diffusion coefficient equation,
D=4.3×10−4 exp (-88,000/RT);
the activation energy was reported to be 88+11 kcal/mole [21].

TEMPERATURE, *c
2OOO 18OO 16OO 14OO 1300
|O 12 T-I-T-I I T I

ſ
:
URANIUM SELF-DIFFUsion IN UO2

| D=4.3 ×10^* ExP (–88,OOO/RT)


-
| -

10-13

- O -
: -

CM
to-" —

- O
| O

to -15 1–1–1–1–11 l I l
4O 45 5O 5.5 6.O
104/T (ok)

FIGURE 7.1. Uranium Self-Diffusion in UO, [21a). (Courtesy, Journal of Nu


clear Materials.)
SOLID STATE REACTIONS OF URANIUM DIOXIDE 309

TABLE 7.1—URANIUM ION (U") SELF-DIFFUSION COEFFICIENTS IN


UO, [21]

Temperature Diffusion coefficient


(° C) (cm2/secY101")
1,450 0. 52
1,485 0.71
1, 545 0. 39
1, 570 1. 1
1, 605 3. 8
1, 645 3. 3
1, 680 5. 4
1, 700 6. 1
1, 705 9. 6
1, 710 14. 0
1,740 23. 0
1, 785 15. 0

Lindner, Matzke, and Schmitz reported the diffusion coefficient


equation for uranium in UO, to be D=1.0 exp (-108,000/RT) over
the approximate temperature range 1,300–1,550° C (see Table 7.2).
These results were obtained using the surface activity decrease method
also, but with U* as the isotopic tracer instead of U* used by
Auskern and Belle [21]. Since a combined diffusion couple was not
used, there is the possibility that evaporation losses may have oc
curred. Diffusion experiments carried out at lower temperatures
(900–1,150° C) in argon resulted in higher diffusion coefficients and
a lower activation energy.
Carter and Richardson have emphasized that diffusion coefficient
measurements by the decrease in surface activity method can be incor
rect because of bound tracer at the surface [10]. Point contact, rather
than total contact, between the tracer layer and the bulk material arises
as a result of discrete crystal formation by the tracer layer. This
bound tracer results in erroneously low diffusion coefficients. Another
consequence of this phenomenon is an observed decrease of the diffu
sion coefficient with time. The decrease with time of the diffusion
coefficient has also been attributed to grain boundary diffusion.
Mahmoud and Kamel, for example, observed a constant diffusion
coefficientwith time for single crystals of cadmium but a decreasing
diffusion coefficient for polycrystalline cadmium [28]. On the other
hand, the possibility of loss of surface activity because of evaporation
would lead to high diffusion coefficients.
These effects have been cited as limitations of the surface activity
decrease method which can be eliminated by sectioning techniques.
The real advantage of the surface activity method, however, is that it
permits measurement of small diffusion coefficients. For example,
for diffusion coefficients on the order of 10-4 cm.”/sec and times of
about 5 × 10° seconds, the total penetration is on the order of 21. Tech
TABLE 7.2–CATION SELF-DIFFUSION IN OXIDES


|
Q
Oxide Structure M.P. C)* Temperature range Do DM.p. Reference


C) (kcal/mole)
6.
×

1,
- - - - - -- - -- Cubic (NaCl type) a=4.213A 2,800 400–1, 600 0.249 79. 11 10-7 14
|| ||

||
- -- -- - -- - -- Cubic (NaCl type) a-4. 812A 2,570 850–1, 600 0.4 81. 2.42X 10-7
X

=
5.
a
- -- -- - - - - -- Cubic (NaCl type) 542A 925 1,075–1, 250 102° 250. 51 104
578

1. 1.
||
- - - -- - - -- -- Cubic (NaCl type) a-4. 292A 370 700–1,000 0.118 29. 34x 10-5
×

2.
- - --- - - - --- Cubic (NaCl type) a-4. 25Å. 800 1,000–1, 350 15X10-2 34. 5.02 10-7 10
- - --- - - -- -- Cubic (NaCl type) a-4. 1768A 955 1,000–1, 400 4.4×10-4 44. 2.06)× 10-8 11
12

| || || |
- - - - -- - - - -- Cubic (NaCl type) a-4, 1768A 955 740–1, 400 1.7×10-2 56. 56X 10-8
5. 2.

- - - -- - - - -- Cubic a-4. 261A 230 800–1,050 4.36X 10-2 36. 50X 10-7
38

1, 1, 1, 1, 1, 1, 1,
||
Fes0,---------- Cubic (Spinel) a-8. 374A 590 750–1,000 5.2 55. 1.88X 10-0
- - - - - -- - - -- Fluorite a=5. 4704A 2,760 450–1, 750 4.3×10-4 88, 15X 10-10 21
2. 1.
| |

2,
1, 1,
- - - -- - - -- -- Fluorite a=5. 4704A 760 300–1, 550
| 1.0 108. 67x 10-8
9

|| || || || ||
000 7 520 10000

--- - -- - - - -- Tetragonal 886 550–625 105 66. 3.74}{10-8


a=3. 947

4.

C
= 988
6
||

1,

2,
- -- -- - - - - -- Hexagonal 520 570–1, 730 5.56× 10^ 111. 1.06X 10-8 17
a=2. 693A
c=4, 370Å
4,
||
1

1,

2,
Hexagonal 520 730–1,934 6.14×10-2 66. 20X 10-7 17
a=2. 693A
c=4, 370A
×

||
0

||

1,

-- -- - --- - - - Hexagonal 975 900–1,025 4.8 73. 3.9 10-7 13


a=3.235Å
c=5. 209A

--
3
||

1.

1,
7
9.
1
Hexagonal 975 800–1, 370 73. × 10-8 4

a
=3.235A
c=5. 209A

|0
6
8

1.

1,
||

-
---
---

o,
Fe2O3- Hexagonal 545 950–1, 300 4.0 103 112. 4X 10-8 and
a=5.025Å
c=13. 735A

of
of

*
Melting points (exceptfor UO2) taken from “Data on Chemicals for Ceramic Use,” Bulletin the National ResearchCouncil, No. 118,published by University Pittsburgh,
June 1949.

**
Z.

R.
F.
in

J.
Lindner, H. Matzke, and Schmitz, “Fission Product Diffusion and Self-Diffusion High Temperature Fuel Materials,” Elektrochem.64,1042–1045
(1960).
:
312 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

niques which permit a good concentration profile to be obtained over


such a small distance are rare. Complex precision grinders have
been described for removing layers of about 0.2a, but these instruments
are not readily available [29]. The residual surface activity which
results from a Dt of about 5 × 10−" and an effective alpha range of
3.75p is about 0.8, an amount readily detected by standard counting
techniques. The equivalent penetration distance, however, was found
to be too small to give reliable sectioning information [20].
As discussed by Auskern and Belle, the surface activity decrease
measurements were free of major systematic errors (recrystallized
tracer layer, finite thickness of the evaporated layer *) except for
possible errors in the effective alpha particle range [21a].
Borisov, et al., have analyzed the kinetics of the decrease in sur
face activity when grain boundary and volume diffusion are both
present [31]. Their results indicate that, for certain values of the
diffusion coefficient, the absorption coefficient (for the case of 3.)
emitters), and the annealing time, a decrease with time in the apparent
volume diffusion coefficient can occur when the observed surface
activity decreases are considered to be due to volume diffusion effects
alone. Nevertheless, the effect of grain boundary diffusion on the
activity decrease kinetics is small at low temperatures and can be
negligible at high temperatures. For the temperature range investi
gated (1,450° to 1,785° C) for uranium diffusion in UO.,
since the
results are consistent with one activation energy, any grain boundary
effect may be minor.
Belle, et al., suggested that these diffusion results represent true
bulk diffusion of the uranium ion (U") in UO, [21b.j. The funda
mental mechanism probably involves the movement
(see Sect. 7.2.4)
of a uranium atom into a vacant adjacent lattice site. The mobility of
the uranium ion is controlled by the vacancy concentration which de
pends exponentially upon temperature.
A number of empirical correlations have been proposed for self-dif
fusion in both metals and oxides. For example, if the self-diffusion
coefficients for a number of cubic metals are plotted on a logarithmic
scale against the reciprocal of the reduced temperature where(T/T,
7 m is the melting point of the metal), it is observed that the lines have
nearly the same slope; this indicates that the ratio Q (activation en
ergy)/7'm does not vary greatly. Nachtrieb and Handler suggested
that since the latent heat of melting, I.,
divided by the melting tem
perature is nearly constant for most metals, the activation energy for

* The average error introduced by the finite thickness of the evaporated layer is about
10 percent and causes the calculated diffusion coefficient to be too large. The use of a
tracer technique, however, results in a measured diffusion coefficient for a face-centered
cubic lattice that is about 10 percent lower than the true cation diffusion coefficient [30 j.
SOLID STATE REACTIONS OF URANIUM DIOXIDE 313

self-diffusion correlates closely with L, for many cubic metals, accord


ing to Q=16.5 Ln [32].
Dienes found a correlation between the frequency factor, Do, and
the activation energy in the form of a quasilinear plot of ln Do/vd”
vs Q/RT, where v is the Debye frequency and d is the interatomic
distance [33].
For oxides, however, Lindner found that cation self-diffusion co
efficients in oxides at the melting point usually have values of -log D
equal to approximately seven [9]. It can be seen from the data pre
sented in Table 7.2 that this correlation is followed; some notable ex
ceptions are BaO and FeO. Note that the diffusion data for UO, of
Lindner, et al., are in better agreement with the empirical correlation
than are the results of Auskern and Belle. Since oxides, in general,
vary greatly in effects of stoichiometry, defect concentrations, and
ionic or covalent bonding characteristics, it would be surprising if all
self-diffusion phenomena were so simply correlated.

7.2.3 Oxygen Ion Self-Diffusion

(a) Introduction

The experimental approaches to the specific problem of determin


ing oxygen ion self-diffusion coefficients in uranium dioxide were first
discussed by Wagner [34]. He suggested two methods: One method,
analogous to that for uranium ion diffusion discussed in Sect. 7.2.2,
consisted of preparing two kinds of UO, pellets, one enriched in O”
and one of normal O" concentration. The pellets are pressed together
and heated in an inert atmosphere. The transfer of heavy oxygen
between the pellets can then be analyzed and diffusion coefficients
determined. was tried but not used because the analyti
This approach
cal methods available for Oº were not satisfactory [20].
The second method attempted and subsequently used was the
method of heterogeneous isotopic exchange [15, 21b). This method of
isotopic exchange with gases has a number of advantages: (1) It is
simpler, experimentally, than the sectioning technique for determining
the time dependence of the concentration distribution in the solid; (2)
the measurements can be carried out on individual crystals, so that
the method is essentially free of the uncertainties due to grain bound
ary effects as they occur in sintered bodies;(3) the method is suitable
for measurement of anion diffusion, which has been little investigated
compared with cation diffusion; and (4) through variation of the
volume ratio between gas and solid and through variation of the
particle size of the solid, diffusion measurements can be carried out
at various temperatures with comparable accuracy.
314 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

For oxygen diffusion in UO, O" in either the gas or solid phase
exchanges with the normal oxygen of the other phase. This general
technique had been used to study catalyst surface reactions as well
as diffusion processes [35–37]. It was found very suitable not only
for oxygen diffusion in uranium dioxide, but also for zinc ion diffusion
in zinc oxide, oxygen ion diffusion in cuprous oxide, and oxygen ion
diffusion in cadmium oxide [13, 15, 16, 18, 21b].
The reaction observed is an exchange reaction between isotopes
of oxygen at the interface of a gas and solid phase, and it may be
considered as occurring in three steps, any one of which may be rate
controlling. These steps are: (1) diffusion of the exchanging specie
through the gas phase to the solid surface; (2) reaction at the inter
face where exchange takes place; and (3) diffusion of the exchanged
specie into the solid phase. To ensure that the measured phenomenon
will be diffusion through the lattice, steps (1) and (2) must be rapid.
This can be realized if the gas is kept well stirred and the diffusion
path sufficiently long.
Solutions to the diffusion equation for diffusion from a solution of
limited volume into a solid have been derived by Carman and Haul
and by Zimen [38, 39]. The experimental conditions which must
be met for the particular solution to be applicable are as follows: (1)
The concentration of exchanging specie is uniform at all times; (2)
the concentration at the surface of the solid is in equilibrium with the
concentration in the gas phase; (3) the amount of gas remains con
be is,

stant throughout the experiment, that gas withdrawn


of

the amount
from the system for analysis must small compared the total
to

gas.
of

amount
The appropriate diffusion equation solutions and the experimental
applied UO, have been dis
as

conditions oxygen self-diffusion


in
to

cussed [15]. For completeness these are repeated here:

1–WW-->
total amount exchanged
[sº] exp (-q,”Dt/a”) Eq. (7.3)
W

where time
in
= =

W. total amount exchanged t=


in

o
in is,

A=“total effective volume ratio,” that num


of

the ratio the


to in

exchanging specie the gas


of

of

that
to

ber moles the


gas relative
of

of

solid
(a

measure the amount the


of

amount solid)
qn= nonzero positive roots the equation
of

tan q=37 (3-H Aq*)


D= diffusion coefficient (cm3/sec)
t=time (sec)
a=average particle radius (cm) for assumed spherical par
ticles.
SOLID STATE REACTIONS OF URANIUM DIOXIDE 315

Equation (7.3) is applicable for small values of 1-W/W. (for long


times); for shorter times Eq. (7.4) can be used:

-
cric-º]-, Eq.(7.4)
^, TV,
1–W/W.– 1 +x|...}. erfeº,7+ 71–H Y2
e

erfc =exp
arº
where e erfc

a.
7–4|Gº)"

|
72–71–1

r= Dt/a”
Walues of the functions erfe
and erfc are found in Carman and
a,

a
Haul and Carslaw and Jaeger [38, 40].
in

To calculate diffusion coefficients, the exchange rate curve and the


particular for the experiment are necessary. For the particular
A

value, 1–W/W.
vs
of

theoretical curve constructed. From


is
a
A

r
curve, corresponding particular
of

to
this values are obtained the
r

1–W/W. value found experimentally. The values are then plotted


be r

against Since r=Dt/a”, this plot should straight line through


t.

origin slope Knowing average particle radius,

a,
of

the D/a". the

D
-
be

may determined.
The directly from the gas anal
of

extent the reaction calculated


is

ysis. The relationship –W/W. =#. used, where the


P
is

is
1

P,

isotopic concentration the isotopic concen


at

the gas
t,

time
in

is

tration initially the gas t=0, and P. the equilibrium isotopic


at
in

is
at

concentration the gas t=co. the experiment has not reached


If
in

equilibrium, P. can determined from P. =XP,--Pi/1+\, where


be

Pi

the initial isotopic concentration


in

the solid.
is

The
presumed diffusion model and the necessary experimental con
ways. That the exchange re
be

of

ditions can tested number


in
a

controlling
be

determined from the


at

action the surface not rate can


is

exchange rate curve. plot log W/W.


of

For first order kinetics


vs
a
fit

first order kinetics together


of

linear. Failure
to

time the data


is

with the linear nature of the curve shows that the diffusion
vs
r

rate controlling. Another important test


of

mechanism the mecha


is

determine diffusion coefficients for different size particles.


to

nism
is

particle size; this


be

independent
of

The diffusion coefficient should


has been observed experimentally and discussed below.
is

(b) Ezperimental Results

The first experiments reported by Auskern and Belle were made


with uranium dioxide powders nearly stoichiometric and nonstoi
of
316 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONs

chiometric compositions prepared by oxidizing uranium metal in


steam at 2,000 psig and 340° C [15a, b]. For these powders, since
the water used for oxidation was enriched in O", the tracer isotope
was incorporated directly in the UO, lattice; the increase in the O*
content of the gas phase was followed kinetically. In subsequent
work, experiments were conducted in which the tracer isotope was
in the gas phase [15c, 21b]. In this case the decrease of Oºº in the
gas phase was measured.
In both investigations carbon dioxide was used as the exchanging
gas, since oxygen could not be employed. As a result of the presence
of CO2 there was some low oxygen pressure in the system due to the
2CO2=2CO+O, equilibrium. Since this pressure is much higher than
the oxygen equilibrium pressure over UO, oxidation of the UO,
could theoretically take place during an exchange reaction (see
Chap. 6). It was found, however, that the uranium dioxide samples
did not change in composition as a result of the exchange reactions."
Moreover, an exchange measurement conducted in the presence of a
considerable hydrogen pressure resulted in a diffusion coefficient es
sentially identical with that obtained in the absence of hydrogen.
There are real advantages to performing the exchange experiments
with the tracer in the gas phase. No measurements need be made of
the isotopic concentration of the solid phase (the concentration of Oº
in the UO, is considered normal). With the tracer initially in the
solid, it is necessary to determine the actual tracer concentration.
This measurement can be performed by equilibrating the O* enriched
solid with a very small amount
of normal CO2 gas. The gas is then
analyzed for isotopic
concentration. Because of the very small
amount of gas compared to solid, at equilibrium the isotopic concen
trations of both phases are essentially equal. This is a much more
involved procedure than a direct gas measurement. Another com
plication arises when exchange measurements are attempted with
oxidized uranium dioxide powders; analyses must be made on the
oxidized powders since, as the O* enriched powders are oxidized, the
O” content becomes diluted.
The materials used and their preparation and characteristics are
shown in Table 7.3. The average spherical radii of the powders were
determined by assuming that the powders consist of spheres of uni
form size. The spherical radius can then be expressed in terms of the
sample surface area and density. The relationship is a =3/Sal, where
a is the spherical radius in cm, S, the surface area in cm"/g (de
termined by BET methods, see Chap. 3) and d, the density in g/ce.

* This result does not preclude some small amount of oxidation at the surface; however,
the analysis of the bulk material did not show a change in oxygen content that could be
attributed to the presence of CO2.
SOLID STATE REACTIONS OF URANIUM DIOXIDE 317

Table 7.3—CHARACTERISTICS OF THE URANIUM DIOXIDE POW


DERS USED FOR OXY GEN ION DIFFUSION COEFFICIENT
MEASUREMENTS [15]

Surface Spherical
Powder Source Treatment O/U" | area” radius
(In’/gm) (microns)

E0H-UO,-------- High pressure steam | Minus 100mesh powder, H. 2.004 0.65 0. 5


oxidation of U metal anneal, 12 hours, 800° C.
(2,000psig, 340° C).
FOH-T0tº ------. High pressure steam Above treatment plus vac- 2.063 || 0.22 1.3
oxidation of U metal uum anneal, 10 days,
(2.000psig, 340°C). 800°C.
MCW-U0;(1):. . . . . Obtainedfrom Mallinck- Minus 100mesh powder, H2 2.002 0.1 2.65
rodt Chemical Works. anneal, 62 hours, 1,725°
C, plus 16hoursat 1,400°C.

H.,
MCW-CO's)f......] Obtained from Mallinck- Minus 100mesh powder, |-------. 0.45 0.6
rodt Chemical Works. anneal, 1,200°C, 24 hours.

fo.
as

MCW-UOis. Obtained from Mallinck- Same MCW-UO2(1) plus 2.057 2.65

1
.
.
..
.
.

10
C,
rodt Chemical Works. vacuum anneal, 700°
days.
2.65
as

MCW-UOis. Obtained from Mallinck- Same MCW-UO2(1) plus 2.092 0.1


||
.
..
.
.

rodt Chemical Works. vacuum anneal, 750° C,


10 days.
by

'Determined polarographic analysis for (VI).


U
by

(s) by

adsorption
N.

"Measured Modified Innes surfacearea method [41].


largeparticle size, particle

to

to

ſelers small size.


lºw temperature UO, these oxygen contents does not increasethe surface area.
to
of

oxidation

isotopic exchange diffusion experiments are shown


A of

The results
plot the logarithm the diffusion coefficients ver
of

of

Table 74.
in

Fig. 7.2.
the

reciprocal absolute temperature


is

It

shown can
in
is

the two source powders, EOH and MCW, falls


of

seen that each



into

all

two groups. For the MCW powders, the high temperature


ºrogen-reduced and annealed powders fall one straight line;
on
the

particle radius. Addition


is

independent
of

diffusion coefficient
oxygen the hydrogen-reduced powders signifi
to
ºf

stoichiometric
"ntly increases the rate oxygen diffusion; subsequent removal
of

to of
by

*\gen hydrogen reduction decreases the diffusion coefficient


"sinitial, unoxidized value.
diffusion coefficient with temperature (Fig. 7.2) for
of

The variation
by

"stoichiometric oxide was represented the equation, D=1.2 10°


*"Tºo/RT); the activation
×

energy was reported


an be
to

65.3+5.0
of

*/mole. Because the results, only


of

*""
the scatter estimate
the activation energy for the oxidized MCW mate
of
be

made
an

The data indicated activation energy between


22

32

and
kal/mole.
318 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TEMPERATURE, *c
7OO 6OO 5OO 400
lo - 800
I I C. I
A
A

-
ExP(-29,700/RT)
-

—’
MCW UO2.092
q =2.65/1
K)
|
-
| O 12
MCW UO2.057
EOH UO2.063
-
q =2.65/1 Ö
o =1.3/4
©
A Ö

o NA ©

N
| O-3 w
A
# W. ANA A
~ rº A
* EOH UO2.004
: o =0.5/1
317°C.--

Nº.
X o “exe-zeroon?

| O-14 A

O
MCW UO2ooz
\.
o =2.65m.

D-1.2 x 10°ExP(-65,300/RT)

| O-15
\
ºn Hz PRESENT
O SMALL PARTICLE SIZE, ozo.6A
D REDUCED UO2 ost
O

| Orie I I l
O.9 I.O 1.1 1.2 1.3 1.4 1.5
lo”/T ("k)

FIGURE 7.2. Self-Diffusion of Oxygen in Uranium Oxide [15c]. (Courtesy,


Journal of Nuclear Materials.)

EOH powders was rep


Oxygen diffusion in the nonstoichiometric
resented" by the equations: Duo, o, 7.0×10−"exp (-29,700/KT);
Duo, nº-2.06X 10-" exp (-29,700/RT). The activation energy was
reported to be 29.7+2.3 kcal/mole.
It appears that oxygen diffusion in UO, is structure sensitive; the
magnitude of the measured diffusion coefficient may be affected not

* Subsequent to the appearance of these data, the particle radii that were used were
found to be in error; however, only the pre-exponential factor had to be revised [15a, b].
SOLID STATE REACTIONS OF URANIUM DIOXIDE 319

TABLE 7.4—SELF-DIFFUSION COEFFICIENTS OF OXYGEN IN VARIOUS


URANIUM DIOXIDES [15]

Temperature Diffusion coefficient


(° C) (cm2/sec)
EOH-UO,------------------------------------- 445 5. 3X 10-15
465 1. 0X 10-14
497 4.5 × 10-14
517 4.5 × 10-14
522 6, 8X 10-14
535 3. 5X 10-14
540 4. 8X 10-14
571 1. 4 X 10-13
600 1.7 X 10-13
602 3. 5 × 10-13

EOH-UO, os----------------------------------- 317 1.9 × 10-14


360 2. 1 × 10-13
386 4. 6X 10-13
405 4 × 10-12
406 6. 0x10-13
421 4 × 10-13
442
465
482
i 2X
4.8X
7X 10-12
10-12
10-12
484 8. 9X 10-12

MCW-UO,------------------------------------ 550 1.9 × 10-16


-
591 4.9 × 10-15
625 1. 3 × 10-14
647 5. 3 × 10-14
666 1. 8X 10-13
712 3.0.X 10-13
748
780 i
1. 5.X 10-12
1 × 10-12

MCW-UO,(s)---------------------------------- 593 . 4 × 10-15


635 . 5X 10-14

MCW-UO2.06---------------------------------- 415 2X 10-13


503 4 × 10-13
550 1 × 10-12
600 : X 10-11

MCW-UO.00---------------------------------- 470 5X10-13


551 . 4 X 10-12
601 : 6 × 10-12

MCW-UO2.06 (reduced) - - - - - - - - - - - - - - - - - - - - - - - - - 655 8.0 × 10-14

0x
(1)

MCW-UOz (H2 present)


4.

647 10-14
-
-
-

-
-
-
-
-
-
-

-
-
-
-
-
-
-

-
-
320 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

only by traces of oxygen but also by surface effects such as crushing


or grinding of particles."
A few exchange experiments carried out at high temperatures
(725 to 850° C) on a coarse UO, powder prepared by crushing sin
tered, dense UO, plates indicated absolute values for diffusion coeffi
cients about three times larger than those for the nonstoichiometric
MCW powder. (The activation energy, however, was in agreement
with the value of 65 kcal/mole.)

7.2.4 Mechanisms for Oxygen and Uranium Ion Diffusion in UO,

The strong dependence of the oxygen ion self-diffusion rates on the


amount of excess, that is,
interstitial oxygen the uranium dioxide

in
lattice, suggests that the diffusion current carriers are interstitial
oxygen anions. diffusion model was assumed which the equili

in
A

oxygen isotopic concentration gradient


an
of

bration nonstoichio

in
by
brought about

of
metric uranium dioxide movement the excess
is

interstitial oxygen [15c, 21b). The activation energy for diffusion


uranium dioxide, kcal/mole, 22

32
consist
in

to
nonstoichiometric

a is
ent with that found from oxidation studies of UO, where similar
type implied (see Chaps. [42, 43].
be
of

8)
motion can and
6

an
The diffusion step was considered
be

interstitial
to

one which
in

ion and lattice ion move combination, the lattice ion moving into
in
a

available neighboring interstitial site and the interstitial ion moving


an

into the evacuated lattice site. Thus, equilibration and diffusion both as
occur the same combined step. mechanism such this, which
in

has been called interstitialcy diffusion, seems preferable from steric


considerations alone than one where motion directly between inter
is

stitial positions and equilibration results from direct interstitial lat


a

tice exchange [44,45].


The much higher activation energy observed for the reduced, stoi
chiometric MCW-UO, powder (65.3 kcal/mole) suggests more
it a

complex diffusion process. For the stoichiometric oxide, also


is
if

assumed that the diffusion current carriers are interstitial oxygen ions
and the mechanism the interstitialcy mechanism, account must be
is

Although the possibility


of

made for the creation current carriers.


the presence of
all of

of

some small surface oxidation


as

exists result
a

CO2, experiments designed detect this have been unsuccessful


to

[15c].
The formation of diffusion current carriers can be accounted for in
stoichiometric UO, Frenkel-type defect pro
A In

another manner.
is
a

duced thermally. lattice oxygen moves into vacant interstitial


a
an

position vacancy and interstitial oxygen ion.


to

create lattice
a

similar effect has been noted experiments on fission gas diffusion UOs (see
in

in
A
*

Chap. 9).
SOLID STATE REACTIONS OF URANIUM DIOXIDE 321

Diffusion and equilibration is accomplished by the movement of the


interstitial ion; the mobility of the vacancy is considered small com
pared to that of the interstitial oxygen ion. The occurrence of this
type of defect in UO, has been discussed by Belle [20].
By using the relationship between W, the energy to form a Frenkel
defect, U, the energy barrier for defect motion, and E, the activation
energy for diffusion (E = U+1% W), the energy to form a Frenkel
defect in UO,
was estimated to be about 71 kcal/mole [15c, 21b, 46].
The activation energy for uranium self-diffusion in stoichiometric
UO, is higher than that for oxygen self-diffusion, viz., 88 vs 65
kcal/mole. At 1,600°C, the extrapolated oxygen diffusion coefficient
is about 10° times greater than that for uranium.
It is interesting
to note that for cuprous oxide the self-diffusion
coefficients for
both copper and oxygen are approximately the same
[16]. It has been suggested that both copper and oxygen ions diffuse
via vacancies in the copper lattice; diffusion via oxygen vacancies was
excluded because the oxygen diffusion coefficient increases with
increasing oxygen pressure [16].
The large difference between oxygen and uranium diffusion coef
ficients in UO, implies that a direct coupling of the diffusion mecha
nisms does not exist in UO. Diffusion of oxygen in stoichiometric
UO2 is considered to arise from the creation and mobility of intersti
tial oxygens and oxygen ion vacancies. If it is assumed that in stoichi
ometric UO, the predominant thermal disorder in the cation lattice is
also due to Frenkel-type defects, then the measured activation energy
for uranium ion diffusion is composed of two parts: the energy re
quired to create a uranium vacancy plus a uranium interstitial and the
energy required to move the diffusion current carrier. It has been
suggested that the diffusion current carrier is the uranium vacancy
[21b). Measurements on the effect of nonstoichiometry (both oxygen
ion interstitials and oxygen ion vacancies) may clarify this problem.

7.3 SINTERING OF URANIUM DIOXIDE


J. R. Johnson and H. G. Sowman

7.3.1 Introduction

Sintering may be defined as a process in which powdered materials


are consolidated into a coherent mass under the influence of heat at
temperatures below the fusion point.
A number of mechanisms for
sintering nonmetallic materials have been proposed, some of which
have been discussed for UO2. These include:
1. Diffusion (movement of ions).
a. Volume (movement through crystal lattices).
b. Surface (movement on or near surfaces of crystallites).
322 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

2. Macroscopic Flow (movement of groups of atoms).


a. Wiscous (nonelastic movement where deformation is propor
tional to stress). -

b. Plastic (nonelastic movement where deformation requires a


minimum yield stress).
3. Vapor Movement (vaporization-condensation).
No single mechanism can account for all the observed sintering be
havior of UO,. It is likely that all these mechanisms contribute to sin
tering in various stages of densification or, at least, are operative to
some degree. Factors which affect the sintering of UO, include crys
tallite size, particle size, shape and density, surface area, oxygen
uranium ratio, fabrication procedures, sintering atmospheres and
impurities. Some observations on the sintering of UO, are discussed
briefly, followed by a discussion of their probable meaning toward
sintering mechanisms [47–76].

7.3.2 Experimental Observations

(a) Physical Nature of Presintered Powders

There is a general indication that grains (or crystallites) having


higher surface areas sinter more readily, which is expected inasmuch
as free surface energy is the driving force. This observation must be
considered, however, as invariably related to the chemical nature of
the surfaces as well. Comminution has been cited as increasing the
sintering rate that leads to higher ultimate densities [47, 54, 55, 65].
This is probably a result of the decreased particle size and increased
surface area, but it may also provide higher surface energy on the
freshly cleaved surfaces. Correlations have been made between sin
tered density and both particle diameter and BET surface area [56,
66]. Particle shape as well as size was found to be important [72].
Dense grains of carefully selected particle size will yield a green
pressed compact of higher initial density, so that more intimate grain
contact may lead to observations of ready sinterability. It should be
noted that the effects of the physical nature of UO, powders are
common to sinterable materials in general.

(b) Chemical Nature of Presintered Powders

The oxygen-uranium ratio in the initial powder may affect the sin
tering rate, presumably as a result of either or both the defect struc
tures created or the increased plasticity of nonstoichiometric UO.,
[76, 77]. It was first observed that in neutral atmospheres oxide pow
ders having high oxygen-uranium ratios (2.14) sintered more readily
[47]. It was later pointed out that the differences observed in the
SOLID STATE REACTIONS OF URANIUM DIOXIDE 323

powder used could be attributed solely to variations in particle size


[76]. A similar observation 7 of the effect of O/U ratio was made with
MCW powders prepared by H, reduction of UO, where preoxidation
to UO2.15 increased sinterability in hydrogen atmospheres. Other re
ported results are conflicting, indicating similar effects, no effect, and
an opposite effect of actually lowering final density [54, 55, 68]. It
is commonly observed that UO, powders prepared from ammonium
diuranate have high O/U ratios as well as high surface areas and sinter
most readily. Furthermore, powders having high O/U ratios, for ex
ample, peroxide precipitates, sinter more readily than those of lower
O/U ratios but higher surface area, for example, hydrogenated steam
oxidized materials.” It has also been found that the influence of ex
cess oxygen on sintered density is fairly small for O/U ratios between
UO, or and UO, [67].
is
As discussed in Chap. 3, UO, powders can be characterized by
differential thermal analysis-gravimetric measurements [78]. Oxides
of fine particle size and high surface area react in air to form UAO,
at relatively low temperatures. Figure 3.16 shows the pattern of
ADU powder heated in air. The oxidation to UAO, is revealed by the
exothermic DTA peak and a corresponding gain in weight. The
second reaction, U.O.-U3Os, is distinctly separated from the first,
and a sharp DTA peak is observed. A larger grained UO, (fused)
is shown (Fig. 3.17) at the other extreme in particle size and surface
area. The two DTA peaks are nearly merged into one, and the
UO,->U.O. reaction has taken place at a higher temperature for the
same rate of heating. The most easily sintered uranium oxides are
characterized by rapid reactions to higher oxidation states at lower
temperatures. The DTA peak separation has been considered to be
the best technique to evaluate sinterability of UO, powder prepara
tions [66].

(c) Effects of Fabrication

UO,
all

shapes have been fabricated by conventional ceramic tech


Chap. With proper selection
as

niques
by of

materials
in

discussed
4.

and processes, high density UO, shapes have been made all these
fabrication methods.
correlation exists between sintered density and green density
A

for pressed bodies, although very sinterable powders,


of
in

the case
for example, ammonium diuranate precipitated materials, this
is

only sintering temperatures. the proper organic


If

lower
at

revealed

Johnson and Curtis, Oak Ridge National Laboratory, unpublished data,


C.
R.

E.
J.
*

1954.
CO2 prepared by oxidation hydrided uranium chips (see Chap. 3).
of
*

57.4789 O–61–22
324 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

binder-lubricant is used with ceramic oxides, isostatic pressing above


about 30,000 psi on simple shapes has little more effect in obtaining
maximum sintered density. Where length to diameter ratios are
maintained of the order of 1, the same is essentially true for steel die
pressing. At very high pressures such as 100,000 psi, fracturing
of the grains is observed [50]. This may increase mobility of surface
ions on the cleaved surfaces. Essentially, however, pressing serves to
provide intimate contact of the particles prior to sintering.
Excessive pressure in fabricating green pellets of ceramic materials
commonly causes such defects as end-cracking and laminations which
affect apparent density adversely. Slip casting and extrusion are
more sensitive to the physical and chemical nature of the powders
and require highly sinterable UO, if high densities are to be obtained.

(d) Effects of Furnace Atmosphere

Sintering of UO, is very sensitive to the presence of oxidizing gases


in the furnace atmosphere. This sensitivity makes it difficult to
evaluate the various observations of sintering in steam, hydrogen,
vacuum, and inert gases, and to separate the effects of atmosphere
from the effects of composition.
Sintering to high density may be accomplished (at 1,400° C) if
the UO,is heated in steam and cooled in hydrogen to room tempera
ture [61]. Ifthe preheating and cooling are done in argon, a presoak
in hydrogen at 1,200° C is necessary to accomplish similar results [63].
Removal of excess oxygen from nonstoichiometric UO, by reduction
in hydrogen prior to sintering in a steam atmosphere appeared to have
no effect on sintered density [66]. On the other hand, introduction
of excess oxygen into UO, by controlled oxidation, prior to sintering
in argon, was found to be beneficial in attaining high sintered density
at 1,400° C [66]. It was also reported that at the same temperature
lower densities were achieved in an argon atmosphere than were ob
tained in a dry hydrogen atmosphere [67]. Vacuum sintering at a
temperature as low as 1,200° C resulted in high densities (95 percent
of theoretical), but only if nonstoichiometric powders were used [69].
The type of atmosphere also appears to have an effect on the grain
size in the sintered body. It was found, for example, that steam
sintering leads to a larger grain size than does either moist or dry
hydrogen sintering at the same temperature [67]. Use of an argon
atmosphere at 1,200° C for large surface area powders produced a
more coarse-grained product (5 to 10p) than did a hydrogen atmos
phere (111) [70]. Comparison of the effect of atmosphere on grain
size is valid, however, only if densities are equal in each case.
SOLID STATE REACTIONS OF URANIUM DIOXIDE 325
(e)

of
Effects Additives

foreign materials, whether present impurities

as
of
Small amounts
intentionally added, are known influence the sintering behavior

to
of or

powdered materials. This influence has been observed with com

by

of
powders
pacted
of uranium dioxide number different

a
investigators.
most striking, where only per
of

of
The effect TiO2 fraction
is

a
0.5) lowers the required sintering temperature several
to

cent (0.1
hundred degrees centigrade [53, 65, 66, 74, 79]. CaO-TiO, and Nb.O.
have similar effects [65, 74]. In
hydrogen-sintered UO2,

to
contrast
the additives tested resulted in increased densities for steam
of

none

sintered UO, [74].


Other additives which have been investigated include: CaO, ZrO2,
SiO, Fe2O3, Al2O3, NiO, V.O., Cr2O, MnO, MoC), CeO2, SiC, and
In

alkalihalides [66,74, 80].

of
some cases indications beneficial effect
have been found for Al2O3, CaO, Cr. O, MnO, V.O., and CeO2, with
Ce0, having the most pronounced effect. Uranium oxide materials
di
by

being produced the present time, especially


at

the ammonium
precipitation techniques, are sufficiently fine
or

uranate peroxide
permit achieving high densities without the use
to

grained and active


additives, however, has possible advantages:
of

additives. The use


of

may make possible relax the process control necessary


(1)

to

to
It

it

UO, since the


or
of

of

produce uniform batches active sinterable lots


relatively greater poorly sinterable materials
on

additives
of

is

effect

than with highly sinterable ones; (2)


of
may permit the attainment
it

lower temperatures than normally required.


at

desired densities
UO,
of

The microstructure sectioned pellets which contain TiO2


grains with
an

show large well-shaped apparent interstitial phase be


tween the grains (see 66, Fig. 7.3) [53, 74]. The grains are rounded
giving the appearance large grains liquid.
the

of

It
in

edges, has
at

been assumed (though


no

analysis has been made) that the interstitial


phase titanium rich.
is

reducing agents mixed with the UO, pronounced.


of

The effect
is

metal powders markedly warm


of

of

Additions lower the densities


C) UO, [81].
It

pressed (800° was concluded that the metals were


the UO, thus, destroying
up
of by

using the excess oxygen


of

ºxidized
the sintering mechanism. Carbon has
the

defect lattice
in

effect
a
a

effect, but this case that voids left


in

was concluded the


in
it

similar
*ly
by

the reacted carbon caused the lower density [73].

Microstructural Aspects Sintering


of
ºf

*-
"

Microstructural differences among UO, powders and pressed com


widely different sintering- behavior [20, 55, 72,75].
in

Pacts result Rate


R.

'T. Padden and Belle.


J.
326 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

/ º,
- -
*
FIGURE 7.3. Peroxide-Precipitated UO, with 1.0 w/o TiO2 Additive, Sintered in
Hydrogen at 1,800° C for 3 Hours; X 250 [53].

IOO
T I I T T
MA 55 PER CENT

95 MA 65 PER CENT -

9O MCW 65 PER cENT

85

MCW 55 PER CENT

6O

55 | _l l l I I l l
8 I6 24 32 40 48 56 64 72
SINTERING TIME IN HOURs

FIGURE 7.4. Sintering Rates of UO, Powders [55].


===-
SOLID STATE REACTIONS OF URANIUM DIOXIDE 327
I I I I I I I T I I I I I I I I I I I

- Residual open Porosity PER cent compact vouumE

too - MA 66 PER CENT


33 -

_-
5.
39s H
MA 55 PER CENT -
2
*
: 96 -
-

_-F
3
: 94H Mcw 65 PER CENT O.O -
a.
° 32 – -
*
:
5 so - -
:
Mcw 58.5 PER CENT 7.8
#as L
> to-2
- 86 |
Ž PER cent
: -
84
15.1
82 -
Mcw 52 PER cent
80 -

78 -

70 l I l L l l I l
l l l l l l I l l l l 2
7o so. 9o too lio 120 130 140 150 16O 17O 18O 190 200 210 22O 230 240
sinTERING TIME IN HOURS

FIGURE 7.5. Effect of As-pressed Density on Sintered Density for Long-Time


Sintering (1,700°C, H,) [55].

of sintering, density attained after prolonged sintering, amount of


residual open porosity, and dependence of sintered density on pressed
by
all

pronounced degree
to

density are affected the characteristic


a

both the UO, powder preparation and the pressed


of

microstructure
compact. Sintered compacts made from various UO, preparations
differ considerably grains and
of

both size and distribution


in

also

residual pores.
of

Padden made study the microstructural changes that occur


a

during sintering with two UO, preparations: as-received and wet ball
milled MCW oxide [20, 55, 75]. The results, summarized here, typify
diverse behavior that has been observed for other UO, powders.
the

Sintering curves Figs. 7.4 and 7.5 show two pronounced differ
in

ences between the milled and the unmilled MCW oxide. The ball
MCW (MA) oxide compacts sintered
to

density percent
94
of

milled
a

fast rate and reached densities greater than per


99
at

theoretical
of

theoretical after prolonged sintering (Fig. 7.4). Residual open


of

cent
on

porosity was absent and the pressed density had little effect either
final density (Fig. 7.5). On the other hand, the unmilled com
or

rate
density only about
of
at

pacts sintered much slower rates and reached


a
an

percent (for as-pressed density percent theo


65
of

of
91

theoretical
retical) after long-time sintering (Fig. 7.4). The pressed density
determined both the density attained with prolonged sintering and the
residual open porosity (Fig. 7.5).
328 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONs

A. Wet Ball-Milled MCW UO2, Trans- B. Wet Ball-Milled MCW UOs, Sur
mitted Light Micrograph, X250, Re- face Replica Electron Micrograph,
11,000X, Reduction Factor,
duction Factor, 35. 35.

D. As-Received MCW UO2, Surface

-
Replica, Electron
-
Micrograph,
--
6,400X, Reduction Factor, 35.
C. As-Received MCW UO2, Polished
Section through Particles, Reflected
Light, 250X, Reduction Factor, 35.

FIGURE 7.6. Comparison of UOs Particle Size and Structure, Reduction Factor 35.
SOLID STATE REACTIONS OF URANIUM DIOXIDE 329

The observed differences in the sintering behavior of these two oxide


powders can be correlated with the microstructures shown in Figs.
particle size and particle struc
7.6

in
to

7.9. Pronounced differences


ture between the two oxides are seen (Fig.
7.6). The unmilled oxide
spherical particles
of

or
consists more less which are porous and range
up (Figs. and D). Particle structure, which var
to

size 7.6,
ies in

2001

C
particle

an
little with size, aggregate

of of
smaller units

to in
to
15p

is

1
size. The milled oxide, however, consists fragments from the

1
15A units; the electron micrograph shown Fig. 7.6D represents

of in
of

one these discrete units. Polished sections pressed compacts show


Fig. milled and unmilled MCW oxide
of
(see 7.7) the structures
pressed percent percent density. Micro
65

87 55
to

and theoretical
structures after sintering percent theoretical density are shown
to

Fig. 7.8 and after prolonged sintering Fig. 7.9.

in
in

evident from these micrographs that the ball-milled oxide


It
is

compacts have homogeneous microstructures. The as-pressed com


pacts are uniform appearance and have homogeneous distri
in

a
After sintering
of

small pores. density percent

87
of
to

bution
a

theoretical, the microstructures (Fig. 7.8) are still uniform, and there
no appreciable difference due pressed density; grains are small
to
is

and uniformly distributed, and the pores are primarily the grain
on
boundaries. With prolonged sintering, the grain size, though larger,
still uniform, the residual porosity

no
closed, and the pores are
is
is

longer preferentially located grain Again, there


on

the boundaries.
no

pressed density.
of

marked effect
is

The microstructures the unmilled MCW compacts are much dif


of

ferent. The as-pressed microstructures (see Fig. 7.7) are heterogene


High density areas containing small pores are set matrix
in

ous.
a

density and larger pores. The mean pore size much larger
of

lower
is

and the pore size distribution much broader than the milled oxide
in

compacts. This heterogeneity persists throughout sintering and


affects sintering characteristics.
Graphic data porosity density show that the
of

of

function
as
a

open and closed porosity ratios follow relatively consistent pattern


a

and that open porosity rapidly decreases, approaching zero the


in
of

of

range percent
theoretical density [55, 66]. For all
90

95
to

practical purposes, sintering complete when most the remaining


of
is

closed pores are trapped within grains. Grain size increases only
up

percent theoretical and


97

minor extent
to

at

to

densities about
a

then increases rapidly densification proceeds.


as

Typical results changes porosity and grain size for one UO,
of

in

powder (wet ball-milled MCW oxide) studied


in

detail are shown


in

Fig. 7.10 [55]. The associated changes pore morphology are illus
in

trated Figs. 7.11


in

to

7.17.
330 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

A. Wet Ball-Milled MCW UO2 Pressed B. Wet Ball-Milled MCW UO2 Pressed
to 55 Percent T.D., X250, Reduc- to 65 Percent T.D., X250, Reduc
tion Factor, 36. tion Factor, 36.

C. As-Received MCW UO2 Pressed to D. As-Received MCW UO, Pressed to


55 Percent T.D., X250, Reduction 65 Percent T.D., X250, Reduction
Factor, 35. Factor, 36.

FIGURE 7.7. Comparison of Pressed Compact Microstructures, Polished Sections.


SOLID STATE REACTIONS OF URANIUM DIOXIDE 331

A. Wet Ball-Milled MCW UO, Pressed B. Wet Ball-Milled MCW UO, Pressed
to 55 Percent T.D., Sintered to 87 to 65 Percent T.D., Sintered to 87
Percent T.D., X500, Reduction Fac- Percent T.D., X500, Reduction Fac
tor, 35. tor, 36.

C. As-Received MCW UO, Pressed to D. As-Received MCW UO2 Pressed to


55 Percent T.D., Sintered to 87 65 Percent T.D., Sintered to 87
Percent T.D., X250, Reduction Fac- Percent T.D., X250, Reduction Fac
tor, 35. tor, 35.

FIGURE 7.8. Comparison of Sintered Compact Microstructure at 87 Percent


T.D., Polished and Etched Sections.
332 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

-
*

A. Wet Ball-Milled MCW UOs Pressed B. Wet Ball-milled MCW UO, Pressed
to 55 Percent T.D., Sintered to 99.5 to 65 Percent T.D., Sintered to 99.7
Percent T.D., 250X, Reduction Fac- Percent T.D., 250X, Reduction Fac
tor, 35. tor, 35.

º:
**
**.

;:

C. As-received MOW UO2 Pressed to I). As-received MCW UO, Pressed to


55 Percent T.D., Sintered to 87.0 65 Percent T.D., Sintered to 93.9
Percent T.D., 250X, Reduction Fac- Percent T.D., 250X, Reduction Fac
tor, 36. tor, 35.

FIGURE 7.9. Comparison of Microstructures after Prolonged Sintering (240


Hours, 1,700° C), Polished and Etched Sections.
SOLID STATE REACTIONS OF URANIUM DIOXIDE 333
i i i I i I T- T I I i T I
wet BALL-MILLED MCW UO2 PRESSED TO
65%, TD AND SINTERED IN HYDROGEN
w
s —H40
O CLOSED POROSITY, 1500°C %
3
> © CLOSED POROSITY, 1700°C 2
35
D OPEN POROSITY, 15OO"C A
5 #
s A OPEN POROSITY, 1700°C I -
A + 30
#
O *.
GRAIN SIZE , 1500°C §
A GRAIN SIZE , 17oo "c tº

k-
--
Vil f- * A
25

O
º:
O ~
|
a? – 20 -
-I
u
a. W 5
– 15
~
«
u
-E

i
85 86 87 88 89 90 9. 92 93 94 95 96 97 98 99 IOO
DENSITY, PER CENT THEORETICAL

FIGURE 7.10. Porosity and Grain Size as a Function of Density [55].


--
º

FIGURE 7.11. Rough Porous Structure at Low Sintered Density, 74 Percent T.D.,
Fractured Surface, Window Replica, Unshadowed; 15,000X, Reduction Fac
tor, 12. (Note large irregular cavities and patches of small pores.)
334 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FIGURE 7.12. Fractured Surface Structure at 81.4 Percent T.D. Sintered Den
sity, Window Replica, Shadowed; 15,000X, Reduction Factor, 14. (Most of
the large irregularly shaped cavities are open pores at this density; porosities
measured independently were 15.6 percent of compact volume open porosity
and 2.9 closed.)

Figure 7.6 B is an electron micrograph made from a window replica


of the starting material (see Appendix B). The rough surface and
convoluted particle structure account for the difference between the
particle size measured microscopically (up to 5p) and the equiva
lent particle size (0.3p) calculated from BET surface area and
density measurements (see Chap. 3). At a sintered density of 74
percent theoretical, the pores (see Fig. 7.11) are observed as two types:
large irregular voids and small regular-shaped pores occurring in
patches. (In some areas the symmetry of these small pores is
apparent.)
The fractured surface of a compact sintered to 81.4 percent theo
retical density (Fig. 7.12) is less rough, and the two classes of pores
are more apparent. From the open and closed porosities of this com
pact, 15.6 percent and 2.9 percent, respectively, and the relative con
centrations of the two pore classes observed microscopically, it can be
SOLID STATE REACTIONS OF URANIUM DIOXIDE 335

Fractured Surface Structure at 91 Percent T.D. Sintered Density,


FIGURE 7.13.
Window Replica, Shadowed; 15,000X, Reduction Factor, 14. (Little grain
growth has occurred ; most of large pores are on grain boundaries; most of
small pores are inside grains; small pores have regular shapes and are
bounded by planes.)

inferred that most of the large irregular pore class must be the open
pores at this density. This irregular shape is that expected for the
open pores during the early stages of sintering, since the open pores
originate at the interstice between powder particles in the pressed
compact. -

Although considerable densification has occurred by the time a


density of about 91 percent theoretical is achieved, little grain growth
has occurred (Figs. 7.10 and 7.13). Pore structure is similar to that
observed at 81.4 percent theoretical density except that the large
pores are less irregular in shape at the higher density. Most of the
large pores are located on grain boundaries, and patches of small
pores are located inside the grains.
At a density of 94 percent theoretical (Fig. 7.14) the large pores,
now all closed and existing as isolated cavities, have begun to spheroi
dize, tending to reduce their area to volume ratios. Well defined
planes are observed on some of the pores of the large size class. Grain
336 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

--- *

-
-
- Q. --
ºs -o' --
º ºr
i.
-
FIGURE 7.14. Fractured Surface Structure at 94 Percent T.D. Sintered Density,
Window Replica, Shadowed; 15,000X, Reduction Factor, 14. (Grain size
increased somewhat ; some large pores now closed have well-developed planes
and are inside grains.)

size has increased somewhat, and although most of the large pores
are still located on grain boundaries, some are now within the grains.
Considerable grain growth has occurred by the time a density of
97.6 percent theoretical is reached (Fig. 7.15), and a larger fraction
of the large pores are located inside the grains. Patches of small pores
are still present, but end abruptly at the grain boundaries in some
areas. The regular shapes of the pores are well defined; the small
pores are bounded by planes only, and the large pores are bounded
by both planes and curved surfaces. The symmetry of both pore
classes is that of the truncated octahedron, and the orientation is the
same within one grain.
Above a density of 97 percent theoretical, grain size increases rapidly
with increasing density (see Fig. 7.10). More of the large pores are
located inside the grains, some patches of small pores still exist (Fig.
7.16), and the large pores also have well developed symmetry but are
still bounded by both curved surfaces and planes (Fig. 7.17). The
SOLID STATE REACTIONS OF URANIUM DIOXIDE 337

FIGURE 7.15. Fractured Surface Structure at 97.6 Percent T.D. Sintered Den
sity, Window Replica, Shadowed; 15,000X, Reduction Factor, 14. (Grain
size has increased considerably ; most of large pores are inside grains; patches
of small pores still exist and end abruptly at grain boundaries in some cases;
both pore classes have truncated octahedron symmetry.)

ratio of curved surface to plane surface appears to be related to size


evenafter prolonged sintering; the larger the pore, the greater the
amount of curved surface.
The angles between the planes on the pores shown in Fig. 7.17 were
measured, and the truncated octahedral symmetry was verified. The
uppermost plane and the three planes on the perimeter were identified
as {111} The three planes intermediate in height were identi
planes.
fied as {100) planes, and the centersof the concentric growth steps are
along the [110] direction. Assuming that the growth steps conform
to the surface of the sphere the difference in height between steps is
about 50 angstrom units. These identified planes are the same planes
that bound growing UO, crystals and crystals of other compounds that
have the CaF2 structure. For crystals, the solid is on the inside of the
boundary surfaces, whereas for the pore the solid is on the outside.
These pores could be described as negative crystals.
The rate of grain growth for UO, sintered in hydrogen at 1,500°
and 1,700° C was reported [55] to be related to time by the expression
338 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
|

º-
-
--

FIGURE 7.16. Patch of Small Pores Remaining after Prolonged Sintering to


98.8 Percent T.D., Fractured Surface, Window Replica, Unshadowed; 30,000X, .
Reduction Factor, 4. (Small pores are bounded by (100) and {111} planes;
shape can be described as negative crystals.)

FIGURE 7.17. Isolated Pore Remaining after Prolonged Sintering to 98.7 Percent
T.D., Fractured Surface, Window Replica, Shadowed; 75,000X, Reduction
Factor, 14. (Pore bounded by curved surface as well as the {100) and {111}
planes; centers of concentric growth steps along the [110] directions.)
SOLID STATE REACTIONS OF URANIUM DIOXIDE 339

D=Kt", where D=mean grain diameter, K=temperature-dependent


constant, t-time, and n=an exponent characteristic of the material.
This equation, applicable for grain diameters up to 10 microns, de
scribes grain growth during the isothermal sintering (1,500° and
1,700° C) of UO2. It appears that deviation from this equation oc
curs at the point in time when the open porosity becomes or approaches
zero.”
As with the pore size distribution, the grain size distribution also
shows a remarkable variation with different UO, powders. Here,
again, the ball-milled powder with its more uniform starting particle
size has the most uniform grain size distribution. The as-received ma
terials when fabricated in pellet form and sintered show a varied grain
size across any section.

(g) Miscellaneous Observations

In sintering UO2, it is commonly observed that material has vapor


ized and condensed on the furnace walls. Where UO, powders have
been heated at 1,700° C in hydrogen, sufficient UO, is reported to
have been lost so as to contaminate furnace insulations [78]. At
1,700° C the vapor pressure of UO, has been measured to be about
0.001 mm Hg (see Chap. 5) [82].

7.3.3 Conclusions

(a) Sintering Mechanisms

A somewhat artificial division of sintering stages may be assigned


for convenience. The initial stage includes the densification from
pressed density to approximately 90 percent of theoretical density.
Near 90 percent density, closed pores become predominant over open
pores or channels, so that volume diffusion of matter and vacancies is of
necessity the primary mechanism remaining for further sintering.
Where very large pores are closed in by regional sintering of agglom
erates, the first stage may be essentially complete at as low as 75 per
cent of theoretical density. In such cases, high final sintered densities
are difficult to achieve.

The variation of grain size distribution across sintered pellets indi


cates that the agglomerated particles which still exist in the pressed
pellets begin sintering first and pull away from the surrounding
medium as the particle shrinks. This, then, probably causes the ap

*In an investigation of grain growth in UO, of 94 percent theoretical density area the
temperature range of 1,555° to 2,440° C. J. R. MacEwan (“Grain Growth in Sintered
Uranium Dioxide,” CRFD-999, January 1961) reported the equation:
D1–D = k104 exp (-87,000/RT),
where D is the mean grain diameter after annealing for t hours at a temperature T (*K),
De is the initial grain size, and ko is a constant.
57.4789 O–61–23
SOLID STATE REACTIONS OF URANIUM DIOXIDE 341

tions,liquid phases, compounds, or the addition or removal of electrons


from the parent material will affection mobility, both on surfaces and
in the interior of particles. Titanium dioxide, Nb.O., and CeO2, it
may be noted, can be reduced to a lower valent oxide. In reducing at
mospheres, it would be expected that TiO2 would be reduced, possibly
to TiO. This would contribute some additional oxygen to the uranium
oxide system either as oxygen ions or possibly in combination with hy
drogen when sintering in hydrogen atmospheres. Not to be overlooked
is the possibility of limited solid solubility, even though only at the
surface of some This would cause some lattice distortion
particles.
with a consequent effect on ionic mobility. The formation of solid
solution, however, does not necessarily mean an improvement in sin
tering. For example, ZrO, which forms some solid solutions with
UO, does not improve sintering (see Chap. 6) [53, 66, 74]. The nega
tive effect on density observed on addition of metal ions which reduce
nonstoichiometric UO., is to be expected on the basis of these conclu
sions. The apparent liquid phase around each UO. grain in the TiO2
additive case suggests very high surface mobility in this system [53,
66,74].
The complex oxidation behavior of UO, makes it readily susceptible
to external oxidizing agents during sintering. It is quite possible that
during the first stage of sintering, material transfer involves move
ment of layers 10 to 20 unit cells deep.” Relatively few oxygen ions
would be needed to maintain a highly mobile defect structure in these
layers. Considerable additional controlled experimentation will be re
quired to permit resolution of these processes.
Diffusion of uranium and oxygen ions on surfaces or in defect lat
tices, where uranium valence may be +5 or +6 rather than +4, or
the bonds may be very directional in character, will be different from
volume diffusion. Thus, sintering mechanisms involving surface ma
terial in motion (surface diffusion and plastic flow which must de
crease to zero as the shear stress goes below a critical value) will be
markedly influenced by surface defects [83].
The activation energy for H, sintering of UO, has been reported
to be 76,000 cal/mole which is similar to the values calculated for
macroscopic flow mechanisms in other refractory oxides [55]. How
ever, in view of the possible complex nature of UO, sintering and
its

low sintering temperature relative melting point,


no
its
to

serious
comparison justified. The activation energies cannot given
be
is

direct physical meaning.


concluded that diffusion, macroscopic flow, and vapor move
It
is

ments all participate sintering UO, the latter probably small


to
in

a
In

degree. the first stage, surface diffusion and later plastic flow

Johnson, “Sintering Tho,” Conference,


R.

of

Gordon Solid State Studies


J.

in
*

Ceramics, unpublished, 1954.


342 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

are believed to predominate. Volume diffusion is considered to be


the primary mechanism in the final stage. Sintering of UO, is in
fluenced considerably by external phenomena which tend to alter the
valence of uranium, creating defect structures. These effects are
believed to be most significant at surfaces.

7.4 ZIRCONIUM-URANIUM DIOXIDE REACTION


M. W. Mallett

7.4.1 Introduction

Fuel elements composed of UO, clad with zirconium or zirconium


alloys are being used for some reactor applications, particularly
water-cooled reactors. Of major importance is the knowledge of the
extent of the reaction between UO, and zirconium and its effect on
the useful life of such fuel elements.
When zirconium and UO, components are heated to elevated tem
peratures while in contact, they will react to form a series of inter
mediate diffusion layers. The reaction that occurs is a reduction of
UO, by zirconium accompanied by interdiffusion of the basic com
ponent elements, zirconium, uranium, and oxygen. The reaction lay
ers that form are detrimental to the quality of the fuel element for
two reasons (see Chap. 9, Sect. 9.8). The uranium-rich diffusion lay
ers are highly corrodible and, in the event of a cladding defect, will be
rapidly oxidized by the primary coolant water with the release of
hydrogen which is absorbed in part by the Zircaloy cladding. The
other detrimental effect of the reaction layers is the embrittlement
of the cladding by the oxygen released by the fuel-clad reaction.
This section is a discussion of the nature of the zirconium-uranium
dioxide reaction. A summary of investigations of methods for elim
inating or minimizing the reaction is also presented.
The literature contains correlative data on the zirconium-uranium
oxygen system, the effect of oxygen on the stability of the epsilon
uranium-zirconium phase, the binary uranium-zirconium system, the
zirconium-oxygen system, the effects of oxygen on the mechanical
properties of zirconium, and the complicated uranium-oxygen phase
relationships (see Chap. 6) [85–90].
One possible mode of interaction between UO, and zirconium is

UO2 +Zr—ZrO2+ U Eq. (7.5)

If the reactants and products are considered to be in their standard


states, or nearly so, the change in free energy at 100° to 1,500° F cal
culated from literature values ranges from about -1.0 to +1.0
kcal/mole with an uncertainty of +1 kcal/mole [91]. It is apparent
SOLID STATE REACTIONS OF URANIUM DIOXIDE 343

that no reaction will take place near 1,500°F. At lower temperatures,


reaction might be possible although the kinetics certainly would be
unfavorable. However, it is known that there is a high solubility of
oxygen in zirconium (6.75 weight percent) [88]. Considering this,
reaction becomes thermodynamically possible in the form,

UO2 +Zr—-U (with Zrin solution) +2O (in solution in Zr)


Eq. (7.6)

To check this hypothesis, Mallett, et al., carried out an experimental


study of the UO,-zirconium reaction at temperatures between 750°
and 2,000° F [91]. A discussion based largely on a paper describing
this work is presented below [92].

7.4.2 Experimental Procedure

The experiments were made with welded sandwich-type elements.


Contact areas were 1.5 inches and 0.3 inch in diameter. Plane sur
faces of the UO, cores (0.25 inch thick) were in intimate contact with
plane surfaces of the end plates of close-fitting zirconium or Zircaloy-2
jackets.
Uranium dioxide cores were prepared from MCW oxide powder
pressed into pellets in steel dies and sintered in hydrogen at about
3,200° F to a density of 75 percent of theoretical. The hydrogen
removed oxygen to a content equal to UO2004, as determined by a
gravimetric method [93].
Elements heated in the temperature range of 1,450° to 2,000° F
were hot pressed in graphite dies in an argon atmosphere. About
3,000 psi pressure was applied to the ends of the capsules with a
lever-arm device. heating, the die containing the elements was
After
lifted out of the furnace tube and quenched in cold water. Other ele
ments were heated in autoclaves at 750°, 950°, 1,100°, 1,200° and 1,300°
F under 5,000-psi helium pressure. After heating, the autoclaves
were removed from the furnaces and cooled with a fan. Element tem
peratures could be dropped from 1,300° F to about 300° F in 1 hour
by this technique.
Small elements were sectioned longitudinally and their reaction
zones examined metallographically. The phases formed were identi
fied by X-ray diffraction techniques. No evidence for diffusion of
zirconium into UO, was observed.
The reacted elements were machined into layers, each 0.5 or 1.0 mil
thick, beginning at the zirconium-UO, interface. Each layer then
was divided into two portions. One was analyzed for oxygen by the
dry-crucible, vacuum-fusion method (extraction at 2,100° C) and the
other for uranium by the fluorescence method [94, 95].
344 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

7.4.3 Experimental Results

(a) Zirconium-UO,

The microstructures of zirconium-UO, elements heated at 1,300° to


2,000° F were all somewhat similar. They contained a two-phase
region consisting of alpha zirconium and alpha uranium adjacent to
the UO, which appeared gray on photomicrographs. Next to this
region and deeper in the zirconium was a stained material which con
tained alpha uranium in elements heated at 2,000° F, the epsilon
zirconium-uranium intermetallic phase in elements heated at 1,300°F,
and mixtures ofthe two phases in elements heated at intermediate
temperatures. Deeper in the zirconium than the stained phase, the
light-etching, oxygen-saturated alpha zirconium was found, and
beyond this the relatively unreacted zirconium structure.
Figure 7.18 shows the microstructure at the interface of an element
heated at 2,000° F for 2.3 days. The dark-etching band was found
by X-ray diffraction techniques to have the alpha-uranium structure.
Chemical analysis showed about 80 weight percent uranium and 20
weight percent zirconium. Elements heated at 2,000° F for other
times showed the same relative distribution of phases that is shown
in Fig. 7.18.
Figure 7.19 shows the microstructure at the interface of an element
heated at 1,600° F for 21 days. The stained band in this element
gave a strong X-ray diffraction pattern of the epsilon-intermetallic
phase of the uranium-zirconium system as well as some evidence for
alpha uranium. The dark areas at the final UO2-metal interface are
evidence of voids produced because the products of the reaction have
higher densities than the reactants.
Figure 7.20 is a typical plot of the thicknesses of the several dis
crete phase regions as a function of time (tº/*). The regions appear
to broaden directly as the square root of time. This behavior was
noted for the range 1,100° to 2,000° F. For the 1,600°F reactions,
the alpha-zirconium (containing dissolved oxygen) region was nar
rower than expected for times longer than 10 days. This apparently
was caused by the formation of the voids at the UO2-metal interface.
The uranium and oxygen gradients in an element heated at 1,600°
F for four days are shown graphically in Fig. 7.21. The data can
be related to the regions in the microstructure shown in Fig. 7.19.
For example, the uranium concentration reaches a maximum in the
uranium-rich phase. At this point, the oxygen content is relatively
low. As the uranium concentration decreases sharply, the oxygen
increases and reaches a peak at about 5 to 7 mils from the interface.
The uranium content decreasesvery rapidly through this region and
appears to level off below 0.05 weight percent at 7 mils from the
--
SOLID STATE REACTIONS OF URANIUM DIOXIDE 345

*º ºf
--- ---
- º 5. - -
------ º
- -- T-----------
º **
- --~~~~~
-- * & - ---
- -

--
-
T.
* ..". - Nº.

FIGURE 7.18. Reaction Zone of a Zirconium-UO2 Element Heated at 2,000° F


for 2.3 Days in Argon; Etchant, 0.5 Volume Percent HF on Polishing Wheel;
+100 [92].

interface. The oxygen decreases much more slowly, reaching a value


of about 0.08 weight percent at 24 to 25 mils from the interface.
Figure 7.22 shows the microstructure at the interface of an element
heated at 1,450° F for 22 days. Numerous small voids are seen at the
UO,-metal interface. The stained uranium-rich phase penetrates
quite deeply by way of grain boundaries into the adjacent zirconium
oxygen solid solution regions.
For the reaction at 1,300°F, the stained uranium-rich material was
zirconium grain boundaries and
lie

be
in

to

found to less continuous


higher temperatures Fig. 7.23. This mate
of as
at

be

in

than can seen


the epsilon-intermetallic phase
of

rial has the lattice structure the


uranium-zirconium phase system. The lighter-etching phase which
346 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

-
-

º
- -*-*.
---
-- -- " -
" -- .
- --
-- -- - -
-

FIGURE 7.19. Reaction Zone of a Zirconium-UO2 Element Heated at 1,600° F


for 21 Days in Argon; Etchant, 0.5 Volume Percent HF on Polishing Wheel;
X100 [92].

appears in grain boundaries deeper in the zirconium and also as a

needle-like phase is precipitated zirconium hydride.


Zirconium reacted with UO, at 750° to 1,200° F showed an almost
complete absence of the uranium-rich phase found in the specimens
reacted at higher temperatures. Figure 7.24 shows the microstruc
ture of an element reacted at 1,200° F for 216 days. A two-phase
alpha uranium-alpha zirconium region exists adjacent to the UO.
The several larger particles found in the region containing alpha zir.
conium with dissolved oxygen may be evidence of either the epsilon
intermetallic phase or of zirconium hydride. The large crack exists
in the oxygen-embrittled alpha zirconium region. Equiaxed alpha
zirconium grains containing precipitated hydrides appear deeper in
the zirconium.
SOLID STATE REACTIONS OF URANIUM DIOXIDE 347
TIME, HR

O.O.8
O
|--|--|-71–1–1–1–
I6 64 IOO .44 196 256 324 400 484 576 784

g
O.of H. / -
RELATIVELY /
UNREACTED /
ZIRCONIUM /
/o -
O.O6 H / O O
ºE /
w /
§u /
ºr / -

2
oos H jo
an /
o /
->
-E
/
-2 o'o.4 H / -
O /
O
ºr /
R. / ...”

s / 2^
ºr O.O3 H. / 2^ -
u OxYGEN DISSOLVED IN
/ ALPHA ZIRCONIUM O

-- --"-
#
s / STAINED zonE
wo /
° 0.02 – /
O (EPSILON
PHASE) .
.”
/ O
/ O O

O.O H.
/
/ -
/ ALPHA URAN ium PLUs
/ ALPHA ZIRCONIUM (WITH
/ OxYGEN IN SOLUTION)
I I l l | l l I l l 1 I l
O
O 2 4 6 8 IO |2 ||4 ||6 18 20 22 24 26 28
TIME, HR#

FIGURE 7.20. Estimation of Thicknesses (on Basis of Microstructure) of Wari


ous Zones in Zirconium, Measured from the Zirconium-UO, Interface, for the
Reaction at 1,600° F [92].

The reaction at 1,100° F produced a much shallower reaction zone


than at 1,200°F, as can be seen in Fig. 7.25. The two phases, alpha
uranium and alpha zirconium, were present in a very narrow region
adjacent to the UO2. The few larger inclusions in the alpha zirconi
um region containing oxygen in solution are believed to be precipi
tated zirconium hydride as are the long needles in the unaffected zir
conium structure.
At 950° and 750° F, the reaction between zirconium and UO, was
very slow. For example, Fig. 7.26 shows an element heated for 503
348 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

4.O T-I-T-I-T-I-T-I-I-I-I-I-I 80

3.8 H — 76

3.6 H

3.4 H
j —H72

— 68

3.2 H. — 64

3.O. H. O OxYGEN — 60
[] URANIUM
2.8 H —H56

2.6 H —H52

2.4 H – 48
S
Sº 2.2 H. —H44 3
> -
;o 2.0 H. – 4o 5
>
z
- «
|.8 H —H36
g
3.
1.6 H. — 32

1.4 H. —H28

1.2 H. – 24

I.O. H. – 20

O.8 H —H16

O.6 H —H12

O. 4 H. – 8

O.2 H. — 4

O 1–1–1R-1–1–1–1–1–1–1–1–1 ––
O 2 4 6 8 IO I2 |4 I6 (8 20 22 24 26 28
DISTANCE FROM INTERFACE, MILS

FIGURE 7.21. Composition of Zirconium Sample after Heating in Contact with


UO, at 1,600° F for 4 Days [92].
SOLID STATE REACTIONS OF URANIUM DIOXIDE 349

FIGURE 7.22. Reaction Zone of an Element Heated at 1,450° F for 22 Days in


Argon; Etchant, 0.5 Volume Percent HF on Polishing Wheel; X100 [92].
350 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FIGURE 7.23. Reaction Zone of an Element Heated at 1,300° F for 32 Days


in Helium; Etchant, 0.5 Volume Percent HF on Polishing Wheel; X100 [92].
SOLID STATE REACTIONS OF URANIUM DIOXIDE 351

-
- - -
* .
--
*
º |- -
º
- | *
-
-
-
- º -
-
---
.

--
--> - " - A- --

--
*- , -

º---
-
-
º -

**- ->
-

--
- -
, , ,
--
-
- - - ->
- -

- - /
FIGURE 7.24. Reaction Zone of Zirconium-UO, Element Heated at 1,200° F for
216 Days; Etchant, 0.5 Volume Percent HF on Polishing Wheel; X100 [92].
352 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

*** * *********, * w - -- - - v '...' . . , ,,

| w

- *
2^
\
t /

-
2^
*
--
\
/ \ >
-
- - T
-*

_Tº `
S

* r -
__
*
~~
~~
º
\ \
-
~.
•* -
`--~ N > `s
º `. > *~ N

2
FIGURE 7.25. Reaction Zone of Element Heated at 1,100° F for 394 Days:
Etchant, 0.5 Volume Percent HF on Polishing Wheel; X 100 [92].

days at 950° F. No evidence of reaction is seen. The precipitated


phase in the zirconium is zirconium hydride.
The element that had been heated at 1,300° F for 32 days was ex
posed to water at 650° F for one hour in an autoclave. The field
shown in Fig. 7.23 before corrosion was examined. The uranium
at the interface had been completely removed by corrosion. The zir.
conium phase as well as the epsilon-intermetallic phase and the pre
cipitated zirconium hydride were little affected.

(b) Zirca/oy-2–I ().

Zircaloy-2 appears to react with UO, less rapidly than zirconium.


The microstructure in an element reacted at 1,300° F for 51 days is
shown in Fig. 7.27. A two-phase region exists adjacent to the UO.
This region is somewhat similar to the alpha uranium-alpha zirconium
SOLID STATE REACTIONS OF URANIUM DIOXIDE 353

/-
º y -
º º
º

.*:, .
- &
-
-- -
-
-
-
-
1. - -
- - - *U

º
J -
-

FIGURE 7.26. Reaction Zone of Element Heated at 950° F for 503 Days;
Etchant, 0.5 Volume Percent HF on Polishing Wheel; X100 [92].

region found in the unalloyed zirconium elements. Next to this


region is a band of what also appears to be a two-phase material. An
occasional tongue of one of these phases penetrates more deeply into
the Zircaloy-2, apparently by way of grain boundaries. A bulky
phase, similar to the phase identified as epsilon in the unalloyed ele
ments, lies still deeper. Beyond the bulky phase, a rather narrow
region containing oxygen in solid solution is noted. The width of
this solid-solution region is much narrower and the quantity of the
bulky epsilon phase is much less than that found in the unalloyed
zirconium.
(c)

Uranium Diffusion Barrier


as
a

The presence the uranium-rich phase


of

the reaction zone caused


in
on

speculation its effect on the diffusion that must occur


to

sustain
354 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FIGURE 7.27. Zircaloy-2–UO, Element Heated at 1,300° F for 51 Days and


Slowly Cooled; Etchant, 0.5 Volume Percent HG on Polishing Wheel; X250
[92].

the reaction. In order to study the effect of uranium as a diffusion


barrier, experiments were set up in which uranium foils, 1 to 2 mils
thick, were placed at the original interface prior to welding and heat
ing of the elements. A typical result for an element heated at 1,450°
F for about 22 days is shown in Fig. 7.28. The element heated in
the same experiment but without the foil appears in Fig. 7.22.
In general, the reaction layers occur in the same sequence and rep
resent the same phases regardless of the initial presence of the uranium
foil. The two-phase region adjacent to the UO2, consisting of alpha
uranium and alpha zirconium, appears slightly thinner in the end of
the element in which the foil was placed. Conversely, the stained
uranium-rich region is thicker and more continuous and penetrates
quite deeply into the zirconium by way of grain boundaries.
SOLID STATE REACTIONS OF URANHUM DIOXIDE 355

-*- *-i--* º
A \, g
- tº

FIGURE 7.28. Reaction Zone of Zirconium-UO, Element Containing a 1-Mil-Thick


Uranium Foil at the Initial Interface after Heating at 1,450° F for 22 Days
in Argon; X100 [92].

7.4.4 Analysis of Results

The experimental results were interpreted in terms of total amounts


of UO, reacting, volume changes during the reaction, growth of
reaction layers, and diffusion constants. Such an interpretation
makes possible a comparison with literature values for diffusion of
oxygen and also permits extrapolation of the results to
in zirconium
various times and temperatures. Choice of some of the factors re
quired in the computations was of necessity somewhat arbitrary.
The rate data that follow are in the form, K=a:/t/*, where a is the
amount of reactant (UO2) consumed and t is time. For layer growth
or penetration, a is distance. The reaction or penetration coefficient,

57.4789–61—24
356 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
TEMPERATURE,”F
16OO
5.0 [. I- 15OO
I-
14OO
r- |3OO
I
12OO
I-
I IOO
||

40 P- -

3 O H.

2.O.H.

O.3 H

O.2 H.

o ll 1– | | 1– l | 1|
9 Ox IO-4 9 Ox1O-4 IO KIO-4 10.5 x 10−4 || xIO-4 11.5x 10-4
T (*K)

FIGURE 7.29. Total Amount of UO, Reacting [92].

K, is proportional to the square root of the chemical diffusion co


efficient, D. From this one obtains the relation

K=Ko exp (-Q/2RT),


where Q is the apparent activation energy in calories per mole for the
diffusion process and A o is a constant.

(a) Quantity of UO, Reacted

Since zirconium did not migrate into the UO, the simplest evalua
tion of the overall reaction can be made in terms of the consumption
of UO, by the metal phase. The amounts of UO, reacted, as shown
SOLID STATE REACTIONS OF URANIUM DIOXIDE 357

in Table 7.5, were determined from chemical analysis for oxygen and
uranium in the zirconium. These elements crossed the UO,-zirconium
interface at a parabolic rate. The rate data, plotted in Fig. 7.29,
are summarized by the equation

K=131 exp (-46,800/2RT) Eq. (7.7)

where K is in units of g of UO, reacting per mmº per hr/*.

(b) Volume Change

As the reactions proceeded there was a tendency for voids to form


at the UO2-zirconium interface. In some cases void formation ulti
mately slowed or stopped the reaction. The voids formed because
the reaction products were more dense than the reactants. The effect
was intensified in the experiments because the density of the UO, was
less than (75 percent of) theoretical. The volume-change data are
given in Table 7.6 and plotted as shrinkage rates in Fig. 7.30. The
rate constants for shrinkage are expressed by the equation

s=9.64 exp (-46,300/2RT) Eq. (7.8)

where Ks is in units of mm per hr/*.

TABLE 7.5—AMOUNT OF UO, REACTING

Temperature Time Oxygen Uranium UO2 Wt.-1/2


(°F) (hr) (ſmolesłmmº]10-0|(Ig atoms/mm3|10-0|| (gºmmºx10-) ([g(mm)[hrº]}10-)

1,100- - -- 9, 456 0.0058 0.0054 1.45 0.015


1,300 - --- 888 0. 0112 0.0090 2.44 0.082
1,450- --- 495 0. 0121 0.0152 4. 11 0. 185
1,525- --- 227 0.0189 0, 0175 4. 72 0. 313
1,600- --- 95 0.0152 0. 0151 4. 08 0. 418

TABLE 7.6—VOLUME CHANGE ON REACTION

Total weight Volume of reactants Final vol


(mg/cm3) (mm3/cm3) ume of Shrinkage
Temperature products
(°F)
UOz Zr UO2 Zr Total (mm3/ (mm) (mils)
cm3)

1,100-- - - - 14.5 211. 3 | 1. 77 32. 31 34. 08 33. 04 || 0, 0104 0.41


1,300----- 24.4 247. 7 || 2.98 37. 87 40. 85 39. 37 0.0148 0. 58
1,450- - - - - 41. 1 449. 4 || 5. 03 68. 72 73. 75 71. 12 || 0, 0.263 1. 03
1,525- - - - - 47. 2 431. 3 || 5. 77 65. 95 71. 72 68. 58 0. 0314 1. 24
1,600- - - - - 40. 8 383. 6 || 4.99 58. 65 63. 64 60. 96 || 0.0268 1. 06
358 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
TEMPERATURE, *F
16OO 1500 14OO 13OO I2OO I IOO
3.O I I I I

O.6 H |

O.4 H

9xIO-4 9.5x1O-4 loxo-4 10.5×10−4 Ilklo-" 11.5xlo'


T(*K)

FIGURE 7.30. Calculated Shrinkage, Based on 75 Percent of Maximum Density


for UO, [92].

For UO, of density 10.9 g/cc, the equation is

Ks=7.38 exp (-49,000/2RT). Eq. (7.9)

The calculated shrinkage would be about 1 percent lower if the more


accurate theoretical density of 10.96 were used.

(c) Penetration of Diffused Elements into the Zirconium Phase

The thicknesses of the several layers formed during reaction were


estimated by microscopic means. Figure 7.20 shows the type of data
obtained. The slope of a line in such a plot may be termed a growth
constant. The logarithm of growth constants for the various zones
are plotted in Fig. 7.31.
the

(1) URANIUM. The lowest curve represents the thickness of


two-phase alpha uranium-alpha zirconium layer. Since the bound.
SOLID STATE REACTIONS OF URANIUM DIOXIDE 359
TEMPERATURE, *F
I loo I2OO 13OO 1450 1525 16OO 2OOO

2.
1.8 I I i I i T
#
2.2 H. -
2.2 H. -
II -

2.
L^
2.4 H

2.6 |- -
---
-

-2
2.8 H

# 3.0 H. z -
z
:a. 2. 2%
zz -
= 3.2 H.
- z
*-
> 3.4 H.

#.
o
§ 38 H z
z
/ 4..
&
2^ A OxYGEN CONCENTRATION
=O.25wo (ANALYSIS OF
-
-
i A: LARGE ELEMENTS)

3.8 H
O ExTENT OF Oxygen PENE
TRATION (VISUAL ESTIMATION)
-
D END OF STAINED ZONE
-
4.O.H.
O
(VISUAL ESTIMATION)
BEGINNINGOF STAINED
-
º
/
ZONE
4.2 H. -
4.4 H. -
4.6 H -
4.8
12.O II.O IO.O 9.O 8.O 7.o
lo"/ T(*k)

FIGURE 7.31. Variation of Zone Thickness with Temperature [92].

ary between this region and the adjacent uranium-rich phase was,
in most cases, well defined, there is relatively little scatter about this
line in Fig. 7.31. The results provide good evidence for a break in
the the

fit

Above this temperature the data


F.
at

curve about 1,300°


equation
Ku-1.58 exp (-33,700/2RT) Eq. (7.10)

Kr
to of

where the unit cm/hrº/*.


is

From 1,100° 1,300° the equation


is
F

Kr-64,600 exp (-75,000/2RT). Eq. (7.11)


of

The presence spikes the uranium-rich phase extending into


of
the

zirconium, especially temperatures, made the


at

the intermediate
360 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS |
estimation of the thickness of this region difficult. This difficulty is
reflected in the scatter of points around the second line of Fig. 7.31.
The points also reflect the more compact nature of this region at
higher temperatures. Since the uranium-rich phase did not occur sig
nificantly at temperatures below 1,300°F, the line extends only from
*
1,300° F upward to 2,000°F.
(2) Oxygen. The uppermost series of points in Fig. 7.31 repre
sent the extent of penetration of oxygen into the zirconium. Since
results of visual observations of this boundary were erratic, the line
was based on the oxygen analyses. These points represent the distance
from the interface to an oxygen concentration of 0.25 weight percent
chosen to represent the moving oxygen penetration boundary. The
line below 1,300° F
was not well established. From 1,300° to
F,

fit
2,000° the rate constants for oxygen penetration the equation

Ko–16.6 exp (-37,000/2RT) Eq. (7.12)

be
where Ko Similar equations can
of
in

units cm/hrº/*. derived


is

for other concentration levels.

(d) Diffusion Coefficients: Oaygen


in

Zirconium.
be

on
Oxygen penetration moving

of
can evaluated the basis

a
boundary oxygen concentration, above,
of

or
as

of
constant discussed
concentration gradients. The latter yields diffusion coefficient.
a

Figure 7.21 typical example


of

the variation concentration in


is
a
of

oxygen and uranium with distance from the interface. The maxi
the uranium curve corresponds the oxygen
in

to in
to

mum the minimum


The solubility
be
oxygen low,
of

uranium known
in

as
curve.
is

zirconium. For computations these solubilities


of

uranium
in

that
is

by
all

nil, and oxygen found analysis was assumed


be
in to

were assumed
present the zirconium phase. By use the analytical value
20 be

of
of to

weight percent for zirconium the uranium-rich phase, smooth


in

curves for oxygen zirconium were derived (Fig. 7.32). The ap


in

parent diffusion constants calculated graphical manner from


in
a

Fig. The straight line, cor


in

these curves are shown 7.33 [96].


responding the equation
to

D=9.4 exp [(–51,780+220)/RT] Eq. (7.13)

cm3/sec, was taken from Pemsler [97].


of

where units
in
1)

It
is

oxygen
of

fits his data for the diffusion zirconium from 400°


to
in

Mallett, Albrecht, and Wilson for the dif


as

of
as

585° well those


C

oxygen Zircaloy-2
of

fusion [98].
at

to

1,000° 1,500°
in

C
SOLID STATE REACTIONS OF URANIUM DIOXIDE 361
6 I I —r I

O 5 IO 15 2O 25
DISTANCE FROM EFFECTIVE INTERFACE, MILS

FIGURE 7.32. Computed Diffusion Gradient of Oxygen in Zirconium after 4 Days


at 1,600°F [92].

(e) Activation Energies

A rough evaluation of the several reaction equations given above


can be made by comparison of apparent activation energies with those
of binary systems. The activation energy reported for the diffusion
of oxygen in zirconium is 51.8 and in Zircaloy-2 is 41.0 kcal/mole
[97, 98]. Mash and Disselhorst found 44.3 kcal/mole for the dif
fusion of uranium in zirconium [99]. From the work of Adda and
Philibert the equation for the average D for the uranium-zirconium
system at 20 weight percent zirconium shows an activation energy of
27.4 kcal/mole [100]. The value increases to 47.0 at 88 weight per
cent zirconium. The values of 33.7 to 46.8 from Eqs. (7.7), (7.8),
362 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
TEMPERATURE, *F
16OO I5OO |400 I3OO I2OO | OO

N
9.O I I I I I

9.5 H. O
-
O

PEMSLER’s LINE
IO.O. H. -

O
O
IO.5 H. -
u!
vo
or.
ul
0.

*Q II.O H.
-
c.

3
| 11.5 – -

12.O.H.
N

12.5 H. Ol

13.O
8.6 9.O 9.5 IO.O IO.5 | I.O i I.5

IO*/T (*K)

FIGURE 7.33. Apparent Diffusion Constants for Oxygen in Zirconium.

(7.10), and (7.12) for the complex oxygen-uranium-zirconium system


compare favorably with the activation energies of the related binary
systems. The value of high and Eq. (7.11) probably
75.0 appears too
could be eliminated on the basis of experimental error. A break in
diffusion behavior at about 1,100° F (600° C) has been reported for
uranium in zirconium, but this was accompanied by a lowering of the
activation energy to about 8.9 kcal/mole [99].

7.4.5 Reaction Mechanism

On the basis of theoretical considerations results, and experimental


a reaction mechanism can be proposed. It is evident that the oxygen
from the UO2 diffuses away rapidly into the zirconium while the re
sulting reduced uranium acts as a solvent for zirconium. In the pres
SOLID STATE REACTIONS OF URANIUM DIOXIDE 363

ence of it appears that uranium is only very slightly soluble


oxygen,
in zirconium. Since oxygen has an extremely limited solubility in
uranium, the uranium-rich phase contains dissolved zirconium but very
little dissolved oxygen. Thus, two solid solutions are formed: one a
solution of oxygen in zirconium, the other a solution of zirconium in
uranium. The uranium-rich phase did not prevent the diffusion of
oxygen.
Zirconium diffuses through the uranium-rich layer in the direction
of the reaction zone. The driving force for this diffusion of zirconium
is presumed to be related to the lower activity of zirconium which is
caused by the higher concentration of oxygen at points closer to the
source of the oxygen.
As proceeds, the uranium formed at the reaction
the reaction
interface is transported through the intervening zirconium phase to
the uranium-rich band several mils away. This band is continuous
at the higher temperatures. The uranium transport through the first
few mils takes place by diffusion along grain boundaries of the zir
conium matrix.
It should be reiterated that the reaction proceeds at a maximum
only when good contact is maintained between the zirconium and UO2.
Under pressure, zirconium deforms and flows into contact with the
UO2. This occurs more readily at high temperatures and low oxygen
concentrations. As reaction proceeds, dissolved oxygen makes the
zirconium phase less plastic and the maintenance of contact more
difficult. The formation of voids becomes competitive with plastic
flow and eventually causes the reaction to stop.

7.4.6 Effect of Barrier

Various barrier materials investigated as possible means for pre


venting the UO2-zirconium reaction included CaO, MgO, graphite,
and carbon applied as coatings on the UO, and metallic barriers which
were applied by inserting thin foils between the fuel and clad [101,
102].
Ceramic oxide coatings of CaO and MgO were applied to UO, by
spraying with aqueous dispersions of the oxide powders to a thickness
of 0.001 inch; these did not inhibit reaction. Barriers, 0.3 to 1.5
mil, of Cr, W, Ta, Mo, and Pt were also tried. In general, the metallic
layers preventedthe fuel-clad reaction, but the interdiffusion of the
barrier and the clad resulted in a layer or zone of low corrosion
resistance.
Graphite coatings were initially applied by brushing or spraying
an aqueous suspension of graphite onto the UO, [102]. Pyrolytic
carbon coatings produced by thermal decomposition of methane were
364 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

later used [101]. Although barriers showed reaction areas,


both
for the pyrolytic carbon coating.
reaction was localized and less severe
Moreover, the latter method is preferred because the pyrolytic coat
ings do not spall when heated to high temperatures (about 1,000° C).

7.5 ALUMINUM-URANIUM OXIDE REACTION


R. C. Waugh, J. E. Cunningham, and R. J. Beaver
7.5.1 Introduction

The first major use of a UO, dispersion fuel in an aluminum matrix


(see Chap. 4) was the manufacture of the core loading for the Geneva
Conference Display Reactor by Oak Ridge National Laboratory in
1955. The fabrication procedure used was a modification of that
developed at ORNL for producing MTR-type fuel elements [103].
During manufacture it was established that UO, was chemically
incompatible with aluminum at 600° C. Reaction was accompanied
by significant growth of the fuel plates. During the final brazing
operation, this growth caused many fuel elements to be rejected be
cause of unacceptable fuel plate spacing.
Triuranium octoxide will also react with aluminum at 600° C. Re
action in this case, however, does not create a similar fabrication
problem because fuel plate growth is greatly delayed and is ultimately
of a much smaller magnitude relative to UO2.
The reactions of UO, and UAOs with aluminum have been investi
gated at the Oak Ridge National Laboratory [104, 105]. Further
studies on the chemical reactivity of aluminum with UO, prepared by
various methods have been performed at Sylvania-Corning [106].
Results of these investigations are summarized below.

7.5.2 UO,-Aluminum Reaction

The reaction of UO2 with aluminum at 600° C occurs as shown in


the unbalanced chemical equation below [104]:

Al-H UO2—-UAl2 + UAla-HUAl, + Al2O, Eq. (7.14)

UAl2 and UAl, are intermediates in the formation of the equilibrium


product UAl... Unfortunately, the lack of data on the U–Al inter
metallic compounds prevents thermodynamic analysis of the reaction.
Chemical compatibility studies at 600° C have been made on full
size, MTR-type fuel plates containing 52.3 weight percent of several
types and particle sizes of UO, in an aluminum matrix [104,105]. The
cores were prepared by cold pressing the blended mixture at 33 tsi.
The plates were hot rolled at 590° C utilizing the conventional picture
SOLID STATE REACTIONS OF URANIUM DIOXIDE 365
24 I I I I I I I I T

2O |-

F-e -

r
2°-e O
O

16 -

12
| O
-

• || -

4 H-
ſ -

O l I | I | I | l l
o 20 40 60 8O IOO I2O 140 I6O 18O 2OO
HEAT TREATMENT AT 600 °C (HR)

FIGURE 7.34. Growth Characteristics of 52.3 w/o UO,-Aluminum Fuel Plates,


Minus 100, Plus 325 Mesh UO2 Reduced from UOa. H2O in Hydrogen [104].

frame technique, flux annealed at 610°C, and cold rolled prior to ex


perimental heat treatment at 600° C. The original volume of each
fuel plate was determined prior to heat treatment by water displace
ment and again, afterwards, to determine the volume increase or
growth which had occurred.
Significant growth was found to occur in all cases. Figure 7.34
shows the growth curve obtained for fuel plates containing minus
100, plus 325 mesh UO, prepared by the hydrogen reduction of
UO, H2O. Growth is seen to increase very rapidly to 20 percent
after 24 hours of heat treatment, decrease slightly, and finally reach
a plateau value of approximately 19 percent. The heat treat control
plate exhibited only slight peripheral reaction as shown in Fig.
7.35. Figure 7.34 would, therefore, depict nearly all of the volume
change associated with a reaction. The reaction products present
in plates heated for 3 and 6 hours were UAlz, UAla, UAl, and Al2O3.
The U—Al intermetallics were identified by X-ray diffraction and the
Al2O, by analytical chemistry. Reaction was complete at 16 hours,
the products being UAl, and Al2O3. The reason for the large growth
is not understood. Calculation of the volume change expected for
the solid X-ray densities and complete densi
state reaction, assuming
fication, indicates that a contraction of approximately 5 percent
should occur. Preliminary data indicate that growth may be due
to the generation of an internal gas pressure during reaction and
366 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

subsequent volume increase by creep; however, this has yet to be


experimentally substantiated.
Figures 7.36 and 7.37 are growth curves for minus 100, plus 325
mesh UO2, prepared by the reduction of UO, H2O in high purity
argon and fused UO2, respectively. These curves are seen to be
similar in shape to those in Fig. 7.34, with the argon-prepared UO,
attaining a maximum growth of 26 percent, and the fused UO, a
maximum growth of 31 percent. Slight reaction had again occurred
in the heat treat control plates of each type. The differences between
the three types of UO, with respect to the time required to reach
maximum growth and the maximum growth attained are not under
stood.
Fuel plates containing minus 100, plus 325 and minus 325 mesh UO,
prepared by steam oxidation were found metallographically to have
undergone extensive reaction during fabrication. Figure 7.38 shows
the severe reaction which occurred with the minus 100, plus 325 mesh
oxide. Heat treatment and volume change determinations on such
plates have no significance.
Pronounced warpage of the fuel plate occurs during heat treatment
because of the growth. Figures 7.39 and 7.40 show the range of warp
age which occurs during heat treatment of fuel plates containing 52.3
weight percent UO, prepared by the reduction of UO, H.O in hydro
gen and argon, respectively.
The compatibility of UO, with aluminum in cores cold pressed at
tsi

approximately
to

percent densification has been studied


95

30.8
up

static vacuum capsule tests [104].


80

600°C for times hours


in
at

to

by

Minus 100, plus 325 mesh UO, prepared the hydrogen reduction

of
UOs: H2O was used weight percent UO, compacts. The
50
in

the
primary reaction products were UAla and Al2O3. Small amounts
UAl, and UAl, were found several cores. The absence UAl, of
on of
in

as

primary product surprising


of
in

view the results reacted


is
a

Al-UO, fuel An average value


plates. percent reaction was
76
of
by

an

determined chemical analysis after 80-hour heat treatment. The


fabricated fuel plates prob
of

cores relative
to

slower reaction rate


is

of

ably related the decreased core densification and the presence


to

as a

the aluminum matrix particles which acts


on

continuous oxide film


diffusion barrier.
a

The chemical reactivity with aluminum UO, prepared by various


of

weight percent UO,-aluminum fuel


10

70
in

to

methods was studied


of

plates and lower temperatures [106]. The method manu


at

600°
C

UO,
on

its reac
pronounced effect
of

facture
to

the was found have


a

by

as

tivity. The reaction rate accelerated fine oxide particles


is

result of the increased surface volume ratio. At 600° visible


to

C
a

occurs within few hours, whereas lower


at

reaction 500° and


a
*º:- -
ſº SOLID STATE REACTIONS OF URANIUM DIOXIDE 367
-

Q
*

-
1.
siſ
I

º
º:

W.

oc

Longitudinal w/o Dispersion Hydrothermal


of

of
FIGURE 7.35. Section 52.3
a

Type UO, of Minus 100, Plus Particle Size


325 Mesh Aluminum Showing
in

only Slight Peripheral Reaction after Hot Rolling


at

590° and 1-Hour


C

a
610° C.; As-polished;
at

Heat Treatment X500 [104].

32 T T T-
T
T

I
I

I
i

i
I

I
i

-r-
24 H.
*e
-
T*-*—e—e—e O O
-
- Q:

O O -
H
#

O
-
; :
I6

w
"||O -
2

-
g= #
8
-

L
l
O

I
I

I
I

40 8O I2O 16O 2OO 24O 28O


O

HEAT TREATMENT AT 600° (HR)

Figure 7.36. 52.3 w/o UO,-Aluminum Fuel Plates,


of

Growth Characteristics
Minus 100, Plus 325 Mesh UO, Reduced from UOa. H2O High Purity Argon
in

[104].
368 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
50 i I I T i i I

40 H -O

O O -
N
3O H. O

—loo + 325 MESH FUSED UO2

2O H. -

IO - —

O
| l —l- l I | I
O 2O 40 60 8O IOO 12O |40 I60

º
HEAT TREATMENT AT 600° C (HR)

FIGURE 7.37. Growth Characteristics of 52.3 w/o UOz-Aluminum Fuel Plates,


JMinus 100, Plus 325 Mesh Fused UO, [104].

º
-----
-
i

º º, -
º º- ºs-º-º-º-º:
- "-

FIGURE 7.38. Longitudinal Cross Section of 52.3 w/o Dispersion of Steam


-º-º:

oxidized UO, in Aluminum Showing Extensive Reaction which Occurred during


Hot Rolling at 590° C and a 1-Hour Heat Treatment at 610° C.; UO, Particle
Size prior to Reduction was Minus 100, Plus 325 Mesh Size; As-polished:
× 500 [104].
SOLID STATE REACTIONS OF URANIUM DIOXIDE 369
6 T-I-T-TTTT I I-T-TTTTI I T-TTTTTT I-T
5 - A - 100+ 325 MESH U30s PREPARED FROM UO3.H2O -
º e - 325 MESH Usos PREPARED FROM Uos. Hzo
2
§ 4 H. -
0. Al, UAl4, Al2O3
º
F 3 H. -
§ Al, UA14, Al2O3

: 2H -
# Z.
Al, UO2, UAl4 A.
:-> | H.
u
-
—-e-T>%/
27

§
O
Al,U3O8,UO2, Al2O3
/ |
Ö
A.Usos.uoz, Al20s -ī’t V. - Al, UO2
-I l L | | | | |1 | | | | | | || L L I LI Lll l L-L
| 2 5 IO 20 50 IOO 200 500 IOOO 5OOO
HEAT TREATMENT AT 600 °C (HR)

FIGURE 7.39. Growth Characteristics of 55.7 w/o Us Os-Aluminum Fuel Plates


[104].

50
-T- I T I
p
2
w
Q
: 40 H
o -
w
3.
z
2
; 30 H O -
<
#
- O
* 20 – O -
C
w

—T _^
: O
2
10H -
º- O

O I I | l
O IOO 2OO 3OO 40O 500
HEAT TREATMENT AT 600°C (HR)

Figure 740. UO, Formation durine deat Treatment of 56.4 w/o U.O.-Aluminum
Fuel Plates at 600° C [104].

temperatures the specimens have to be at temperature for several days


in order to show signs of reaction.

7.5.3 U.O.-Aluminum Reaction

The reaction of U.O, with aluminum at 600° C appears to occur in


two steps as shown below [105]:

4A1+3U Os–90O2+2Al2O, Eq. (7.15)


16A1+3UO2=3UAl, +2Al2O3. Eq. (7.16)
370 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
45 I I T

4O H. -O

35 H. -

O —"
30 H
ALL PLATES RECEIVED TOTAL
OF 144 HOURS AT 600°C
-

I ! l
25
O |O 2O 3O 40
PER CENT COLD REDUCTION AFTER 24 HOURS AT 600° C

FIGURE 7.41. Effect of Intermediate Cold Reduction on UO, Formation in 56.4


w/o Us Os-Balance Aluminum Plate [104].

Reaction products were identified by X-ray diffraction. Although not


detected in limited experiments, UAl, and UAls are undoubtedly in
volved at an intermediate stage of reaction [105].
Compatibility studies have been made at 600° C on fuel plates con
taining 55.7 weight percent UAOs in an aluminum matrix. These
plates were fabricated in a manner similar to the UO,-aluminum
plates previously described and contained an equivalent amount of
uranium. The U3Os, made by calcining UOs: H2O, was studied in the
minus 100, plus 325 and minus 325 mesh particle size ranges. Figure
7.41 shows the growth curve determined. A portion of each curve is
represented as a dotted line to indicate its uncertainty. The identity
of the core components at selected times is noted and illustrates the
two steps of the reaction. The minus 325 mesh UAOs reacts faster, as
expected from surface/volume considerations. Reaction is complete
in both cases after 3,000 hours at 600° C. It is perplexing to note that
at no time during the course of reaction, as shown in Eq. (7.15),
intermetallics detectable by X-ray diffraction, that
is,

were U-Al the


UO, formed by the reduction the UAOs would not react with alu
of

Os
to
of

minum until all the UAOs had been reduced. Relative


U

aluminum, the UAO,-aluminum reaction proceed very


to

seen
is
SOLID STATE REACTIONS OF URANIUM DIOXIDE 371

slowly. Although, ultimately, the same amount of reaction product


UO, is formed in the U.O, plates as that contained initially in the
UO,-aluminum plates previously described, the final growth for the
U.Os case appears to be an order of magnitude less than that for the
UO, case. Calculations based upon theoretical densities for the bal
anced equations indicate that approximately 20 percent volume de
crease occurs during reaction as shown in Eq. (7.15) and 5 percent
decrease during reaction as shown in Eq. (7.16) to give an overall
decrease of 12 percent. The reasons for these differences are not
understood.

REFERENCES

1. R. M. BARREB, “Diffusion in and through Solids,” University Press, Cam


bridge, 1952.
2. W. Jost, “Diffusion in Solids, Liquids, Gases,” Academic Press, Inc., New
York, 1952.
3. W. J. MooRE and B. SELIKson, “The Diffusion of Copper in Cuprous Oxide,”
J.
Chem. Phys. 19, 1539 (1951).
4. R. LINDNER, D. CAMPBELL, and A. AKERSTRoM, “Diffusion of Radioactive Zinc
in Zinc Oxide,” Acta Chem. Scand. 6,457–467 (1952).
R. LINDNER, St. AUSTRUMDAL, and A. AKERSTRoM, “Kinetics in Calcium
5

.
Oxide,” Acta Chem. Scand. 6, 468–474 (1952).
6. R. LINdNER, “Diffusion of Radioactive Iron in Iron (III) Oxide and Zinc
Iron Spinel,” Arkiv Kemi 4, 381–384 (1952).
7. R. W. REDINGToN, “Diffusion of Barium in Barium Oxide,” Phys. Rev. 87,
1066–1073 (1952).
8. L. HIMMEL, R. F. MEHL, and C. E. BIRCHENALL, “Self-Diffusion of Iron in
Iron Oxides and the Wagner Theory of Oxidation,” Trans. AIME 197,
827–843 (1953).
9. R. LINDNER, “A New Investigation on Solid State Reactions” in “Proceedings
of the International Symposium on the Reactivity of Solids,” pp. 195—205,
Gothenburg, 1952.
10. R. E. CARTER and F. D. RICHARDsoN, “An Examination of the Decrease-of
Surface-Activity Method of Measuring Self-Diffusion Coefficients in
Wustite and Cobaltous Oxide,” Trans. AIME 200, 1244–1257 (1954).
11.

M. T. SHIM and W. MooRE, Nickel Nickel Oxide,”


of

J.

“Diffusion
in
J.

Chem. Phys. 26, 802—804 (1957).


AKERSTRoM, “Diffusion
R.

A.

of

LINDNER and Nickel-63 Nickel Oxide


in

12.
(NiO),” Discussions Faraday Soc. 23, 133–136 (1957).
13.

SEcco and W. MooRE, “Diffusion and Exchange Crystalline


A.

of
E.

Zinc
in
J.

Zinc Oxides,” Chem. Phys. 26, 942–948 (1957).


J.
14.

PARFITT, “Diffusion Radioactive Magnesium


G.
D.

in
R.

of

LINdNER and
Magnesium Oxide Crystals,” Chem. Phys. 26, 182–185 (1957).
J.
15.

BELLE, “Self-Diffusion Oxygen


B.
A.

of

(a) UskERN and Uranium


in
J.
A

Dioxide,” Chem. Phys. 28, 171–172 (1958).


J.

AU's KERN, “Oxygen Ion Self-Diffusion


B.
A.

(b) BELLE and Uranium


D. in
J.

High Temperature Processes,” W. Kingery,


of

Dioxide” “Kinetics
in

ed., pp. 44–49, John Wiley and Sons, New York, 1959.
BELLE, “Oxygen Ion Self-Diffusion
B.
A.

J.

(c) AUSKERN and Uranium


in

Dioxide,” submitted for publication Journal


to

of

Nuclear Materials.
57.4789 O—61—25
372 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

16. W. J.
MooRE, Y. EBIsuzAKI, and J. A. SLUss, “Exchange and Diffusion of
Oxygenin Crystalline Cuprous Oxide,” J. Phys. Chem. 62, 1438–1441
(1958).
17. S. B. AUSTERMAN, “Diffusion of Beryllium in Beryllium Oxide,” NAA-SR
3170, Dec. 15, 1958.
18. R. HAUL and D. JUST, “Measurement of Diffusion in Crystals by Isotope
Exchange with Gases,” Z. Elektrochem. 62, 1124–1130 (1958).
19. W. D. KINGERY, J. PAPPIs, M. E. DoTY, and D. C. HILL, “Oxygen Ion Mobil
ity in Cubic Zros. Cao is O.s,” J. Am. Ceram. Soc. 42, 393-398 (1959).
20. J. BELLE, “Properties of Uranium Dioxide” in “Proceedings of the Second
United Nations International Conference on the Peaceful Uses of Atomic
Energy, Geneva, 1958,” Vol. 6, p. 585, United Nations, Geneva, 1958.
21. (a) A. B. AUSKERN and J. BELLE, “Uranium Ion Self-Diffusion in Uranium
Dioxide,” Am. Ceram. Soc. Meeting, Philadelphia, Pa., Apr. 1960, sub
mitted for publication to Journal of Nuclear Materials.
(b) J. BELLE, A. B. AUSKERN, W. A. BosTRoM, and F. S. SUsko, “Diffusion
Kinetics in Uranium Dioxide,” paper presented at Fourth International
Symposium on Reactivity of Solids, Amsterdam, May 30–June 4, 1960.
22. K. KIUKKola and C. WAGNER, “Measurements on Galvanic Cells Involving
Solid Electrolytes,” J. Electrochem. Soc. 104, 379–386 (1957).
23. R. E. HoFFMANN, “Tracer and Other Techniques of Diffusion Measurements"
in “Atom Movements,” pp. 51–68, American Society for Metals, Cleve
land, Ohio, 1951.
24. G. von HEVESY and W. SEITH, “Use of Radioactive Recoil in Diffusion Ex
periments,” Z. Physik. 56,790–801, 1929.
25. A. C. WAHL and N. A. BonneR, eds., “Radioactivity Applied to Chemistry,"
pp. 66–68, John Wiley and Sons, New York, 1957.
26. G. FRIEDLANDER and J. W. KENNEDY, “Nuclear and Radiochemistry,” Chap.
7, John Wiley and Sons, New York, 1955.
27. J. STEIGMAN, W. S.HocKLEY, and F. C. Nix, “The Self-Diffusion of Copper,"
Phys. Rev. 56, 13–21 (1939).
28. K. A. MAHMOUD and R. KAMEL, “Tracer Diffusion of Cadmium 115 in Pure
Cadmium” in “Radioisotopes in Scientific Research,” R. C. Exterman,
ed., Vol. I, pp. 271–278, Pergamon Press, Ltd., London, 1958.
29. H. W. SchAMP, Jr., D. A. OAKEs, and N. M. REED, “Grinder for Sectioning
Solid Diffusion Specimens,” Rev. Sci. Inst. 30, 1028–1031 (1959).
30. J. BARDEEN and C. HERRING, “Diffusion in Alloys and the Kirkendall Effect”
in “Atom Movements,” pp. 87–111, American Society for Metals, Cleve
land, Ohio, 1951.
31. W. T. BORIsov, V. M. GoLIKov, B. Y. LJUBov, and C. W. SHTs HERBEDINsky.
“Study of Grain Boundary Diffusion in Metals” in “Radioisotopes in
Scientific Research,” R. C. Exterman, ed., Wol. I, pp. 212–231, Pergamon
Press, Ltd., London, 1958.
32. N. H. NACHTRIEB and G. S. HANDLER, “A Relaxed Vacancy Model for Diffu
sion in Crystalline Metals,” Acta Met. 2, 797–801 (1954).
33. G. J. DIENES, “Frequency Factor and Activation Energy for the Volume
Diffusion of Metals,” J. Appl. Phys. 21, 1189—1192 (1950).
34. C. WAGNER, “Thermodynamics of the System Uranium-Oxygen,” WAPD
144, July
23, 1955.
35. (a) E. R. S. WINTER, “The Use of O” in Studies of the Reactivity of Solid
Oxides,” Discussions Faraday Soc. 8, 231–237 (1950).
(b) E. R. S. WINTER, “Exchange Reactions of Solid Oxides,” J. Chem. Soc.,
1170–1177 (1950); 1509–1527 (1954).
SOLID STATE REACTIONS OF URANIUM DIOXIDE 373

36. W. C. CAMERoN, A. FARKAs, and L. M. LITz, “Exchange of Isotopic Oxygen


between Vanadium Pentoxide, Gaseous Oxygen, and Water,” J. Phys.
Chem. 57,229–238 (1953).
37. R. A. W. HAUL and L. H. STEIN, “Diffusion in Calcite Crystals on the Basis
of Isotopic Exchange with Carbon Dioxide,” Trans. Faraday Soc. 51, 1280–
1290 (1955).
38. P. C. CARMAN and R. A. HAUL, “Measurement of Diffusion Coefficients,”
Proc. Royal Soc. A222, 109–118 (1954).
39. K. E. ZIMEN, “Kinetics of Heterogeneous Exchange Reactions. Studies of
Solid Reactions by Means of Isotope Exchange,” Kemi, Mineral. Geol.
A20, No. 18, 1–26 (1945).
40. H. S. CARSLAw and J. C. JAEGER, “Conduction of Heat in Solids,” 2nd ed.,
p. 373, Clarendon Press, London, 1948.
41. J. C. CLAYTON and J. E. RULLI, “Measurement of Low Surface Areas of the
Uranium Oxides by the Innes Method,” Chemist-Analyst 47, 62–64 (1958).
42. S. ARONSON, R. B. Roof, Jr., and J. BELLE, “Kinetic Study of the Oxidation
of Uranium Dioxide,” J. Chem. Phys. 27, 137–144 (1957).
43. J. S. ANDERSON, L. E. J. Roberts, and E. A. HARPER, “The Oxides of Uranium.
Part VII, The Oxidation of Uranium Dioxide,” J. Chem. Soc., 3946–3959
(1955).
44. F. SEITz, “The Theory of Diffusion in Metals,” Acta Cryst. 3, 355–363 (1950).
45. E. A. SEcco, “Diffusion and Exchange of Zinc in Crystalline Zinc Sulfide,”
J. Chem. Phys. 29, 406–409 (1958).
46. N. F. MoTT and R. W. GURNEY, “Electronic Processes in Ionic Crystals,”
Clarendon Press, Oxford, 1948.
47. P. MURRAY and R. W. THACKRAY, “The Sintering of Uranium Dioxide,”
AERE M/R 614, Nov. 1950.
48. F. H. GUNzEL, Jr., and W. A. LAMBERTson, “Urania Bodies: Fired Density
vs Particle Size,” ANL–5094, Aug. 1954.
49. A. G. ALLISON and W. H. DUCKworTH, “Ceramic Investigation of Uranium
Dioxide,” BMI–1009, June 16, 1955.
50. J. GLATTER, H. R. Hoge, and B. E. SCHANER, “The Fabrication of Dense
UO2 Cylindrical Compacts by Cold Pressing and Sintering,” WAPD—126,
Oct. 14, 1955.
51. M. D. BURDIck and H. S. PARKER, “Effect of Particle Size on Bulk Density
and Strength Properties of Uranium Dioxide Specimens,” J. Am. Ceram.
Soc. 39, 181–187 (1956).
52. W. A. LAMBERTson and J. H. HANDwerk, “The Fabrication and Physical
Properties of Urania Bodies,” ANL–5053, Feb. 1956.
53. H. G. Sow MAN and G. L. PLoetz, “An Investigation of the Sintering of
Uranium Dioxide,” KAPL–1556, Aug. 17, 1956.
54. N. F. H. BRIGHT, K. V. Gow, and A. T. PRINCE, “Factors Influencing the
Sintering Properties of Uranium Dioxide for Use in Nuclear Reactor
Fuel Elements,” Can. Dep. Mines and Tech. Surveys, Mines Branch,
MD-209, Sept. 18, 1956.
55. J. BELLE and B. LUsTMAN, “Properties of UOs,” WAPD–184, Sept. 1957;
“Fuel Elements Conference, Paris,” TII)—7546, pp. 442–515, Mar. 1958.
Work of T.R. Padden in :
J. BELLE, ed., “Résumé of Uranium Oxide Data–V,” WAPD—PWR-PMM-429
(Del.), Mar. 6, 1956.

J. BELLE and J. Jon Es, eds., “Résumé of Uranium


L. Oxide Data—VI,”
WAPD—PWR-PMM–466 (Del.), June 5, 1956.
374 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

J. BELLE and L. J. Jon Es, eds., “Résumé of Uranium Oxide Data–VII."


WAPD—PWR–PMM-491, Sept. 12, 1956.
J. BELLE and L. J. Jon Es, eds., “Résumé of Uranium Oxide Data-VIII,"
WAPD-PWR–PMM-904, Dec. 3, 1956.
J. BELLE and L. J. Jon Es, eds., “Résumé of Uranium Oxide Data-IX,”
WAPD–TM–44, Mar. 15, 1957.
J. BELLE and L. J. JoNEs, eds., “Résumé of Uranium Oxide Data-X,"
WAPD–TM–73, Aug. 15, 1957.
J. BELLE and L. J. Jones, eds., “Résumé of Uranium Oxide Data-XI,"
WAPD–TM-101, Jan. 15, 1958.
56. D. R. STEN QUIST, B. MASTEL, and R. J. ANICETTI, “Note on Correlation of
Surface Characteristics with the Sintering Behavior of Uranium Dioxide
Powders,” J. Am. Ceram. Soc. 41, 273–274 (1958).
57. L. C. WATsoN, “The Production of UO, for Ceramic Fuels” in “Fuel Ele
ments Conference, Paris,” TID–7546, pp. 384–401, Mar. 1958.
58. D. R. STENQUIST and R. J. ANICETTI, “Fabrication Behavior of Uranium
Dioxide Powders,” HW–51748, Dec. 1, 1957.
59. D. A. WAUGHAN, J. R. BRIDGE, A. G. ALLIsoN, and C. M. SchwARTz, “Process
ing Variables, Reactivity, and Sinterability of Uranium Oxides,” Ind. Eng.
Chem. 49, 1699–1700 (1957).
60. J. WILLIAMs, E. BARNES, R. W. THACKRAY, and P. MURRAY, “The Non
stoichiometric Oxides of Uranium,” Trans. Brit. Ceram. Soc. 56, 608–623
(1958).
61. A. H. WEBSTER and N. F. H. BRIGHT, “The Effects of Furnace Atmospheres
on the Sintering Behavior of Uranium Dioxide,” NP–6667, Feb. 5, 1958;
Can. Dep. Mines and Tech. Surveys, Mines Branch, R–2, 1958.
62 . P. MURRAY, S. F. PUGH, and J. WILLIAMs, “Uranium Dioxide as a Reactor
Fuel” in “Fuel Elements Conference, Paris,” TID–7546, pp. 432–441,
Mar. 1958.
C. A. ARENBERG and P. JAHN, “Steam Sintering of Uranium Dioxide,” J. Am.
Ceram. Soc. 41, 179–183 (1958). -
R. KIEssi,ING and U. RUN Fors, “Sintering of Uranium Dioxide" in “Fuel
Elements Conference, Paris," TID–7546, pp. 402–413, Mar. 1958.
B. E. SchANER, “The Fabrication of High Density UO, Fuel Platelets,” Am.
Ceram. Soc. Bull. 38, 494–498 (1959).
66 . C. H. CHALDER, N. F. H. BRIGHT, D. L. PATERson, and L. C. WATson, “The
Fabrication and Properties of UO, Fuel” in “Proceedings of the Second
United Nations International Conference on the Peaceful Uses of Atomic
Energy, Geneva, 1958,” Vol. 6, pp. 590–611, United Nations, Geneva, 1958.
67. U. RUN Fors, N. SchöNBERG, and R. KIEssli Ng, “The Sintering of Uranium

Dioxide” in “Proceedings of the Second United Nations International


Conference on the Peaceful Uses of Atomic Energy, Geneva, 1958,” Vol. 6,
pp. 605–611, United Nations, Geneva, 1958.
A. BEL and Y. CARTEREt, “Contribution to the Study of the Sintering of
Uranium Dioxide” in “Proceedings of the Second United Nations Inter
national Conference on the Peaceful Uses of Atomic Energy, Geneva, 1958."
Vol. 6, pp. 612–619, United Nations, Geneva, 1958.
69. J. TERRAzA, J. CERRo1.AzA, and E. APARICIo, “Sintering UO, at Medium Tem
peratures” in “Proceedings of the Second United Nations International
Conference on the Peaceful Uses of Atomic Energy, Geneva, 1958,” Vol. 6,
pp. 620–623, United Nations, Geneva, 1958.
70. A. BEL, R. DELMAs, and B. FRAN cois, “Sintering of Uranium Oxide in
Hydrogen at 1,350° C.” J. Nuclear Materials 1, 259–270 (1959).
SOLID STATE REACTIONS OF URANIUM DIOXIDE 375

71. W. G. CARLsoN, H. D. Root, and R. S. SHANE, “Pressure-Temperature-Den


sity Relationships in Several Samples of UO,” Am. Ceram. Soc. Bull. 39,
69–73 (1960).
72. J. C. CLAYTON and L. BERRIN, “A Study of the Sintering Behavior of Some
Uranium Dioxide Powders,” Bettis Technical Review, WAPD—BT-20,
Sept. 1960.
73. N. F. H. BRIGHT, K. V. Gow, and A. H. WEBstER, “The Effect of Carbon on
the Sintering Behavior of Uranium Dioxide,” Can. Dep. Mines and Tech.
Surveys, Mines Branch, MD–213, Nov. 7, 1957.
74. A. H. WEBstER and N. F. BRIGHT, “The Effects of Additives on the Sintering
of Uranium Dioxide,” Can. Dep. Mines and Tech. Surveys, Mines Branch,
MD–223, Nov. 7, 1957.
75. T. R. PADDEN, “An Electron Microscopy Technique for Studying the Shape,
Size, and Distribution of Pores in Sintered UO, Compacts,” paper presented
at the 15th Annual Meeting of the Electron Microscope Society of America,
Boston, 1957.
76. J. WILLIAMs, E. BARNEs, R. Scott, and A. HALL, “Sintering of Uranium
Oxides of Composition UO, to U20, in Various Atmospheres,” J. Nuclear
Materials 1, 28–38 (1959).
77. R. Scott, A. R. HALL, and J. WILLIAMs, “The Plastic Deformation of Ura
nium Oxides above 800° C,” J. Nuclear Materials 1, 39–48 (1959).
78. J. R. Joh Nson, C. E. CURTIs, L. M. DoNEY, S. D. FULKERson, A. J. TAYLoR,
J. M. WARDE, and G. D. WHITE, “The Technology of Uranium Dioxide, A
Reactor Material,” Am. Ceram. Soc. Bull. 36, 112–117 (1957).
J. R. Johnson and C. E. CURTIs, “The Technology of UO, and Tho,” in
“Proceedings of the United Nations International Conference on the
Peaceful Uses of Atomic Energy, Geneva, 1955,” Vol. 9, pp. 169–173,
United Nations, New York, 1956.
79. J. GLATTER, “Fabrication of Bulk Form UO, for Use as Nuclear Reactor
Fuel,” WAPD–T-593, Aug. 1957.
80. Work of B. E. SCHANER, reported in “Résumé of UO, Data-VII,” J. Belle
and L. J. Jones, eds., WAPD—PWR—PMM–491, Sept. 12, 1956.
81. R. Scott and J. WILLIAMs, “The Warm Pressing (800° C) of UO, and
UO,-Metal Mixtures,” Trans. Brit. Ceram. Soc. 57, 199–207 (1958).
82. J. J. KATz and E. RABINow Itch, “The Chemistry of Uranium,” National
Nuclear Energy Series, Div. VIII, Vol. 5, p. 261, McGraw-Hill Book Co.,
New York, 1951.
83. A. R. BLAckburn, T. S. SHEv1.1N, and H. R. Low ERs, “Fundamental Study
and Equipment for Sintering and Testing of Cermet Bodies; I. Wetting
of Alumina by Nickel, Cobalt, Iron, Chromium, and Chromium-Boron,”
J. Am. Ceram. Soc. 32, 81–89 (1949).
84. R. F. WALKER, “Mechanism of Material Transport during Sintering,” J.
Am. Ceram. Soc. 38, 187–197 (1955).
85. H. A. SALLER, F. A. Rough, J. M. FACKELMANN, A. A. BAUER, and J. R.
Doig, “Phase Relations in the Uranium-Zirconium-Oxygen System In
volving Zirconium and Uranium Dioxide,” BM I–1023, July 1955.
86. F. A. Rough, A. E. AustiN, A. A. BAUER, and J. R. Doig, “The Stability and
Existence Range of the Zirconium-Uranium Epsilon Phase,” BMI-1092,
May 28, 1956.

87. D. SUMMERs-SMITH, “TheConstitution of Uranium-zirconium Alloys,” J.


Inst. Metals 83, 277–282(1955).
88. R. F. Dom AGALA and D. J. McPHERson, “System Zirconium-Oxygen,” Trans.
AIME 200, 238–246 (1954).
376 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

89. R. M. TREco, “Some Properties of High-Purity Zirconium and Dilute Alloys


with Oxygen,” Trans. Am. Soc. Metals 45, 872—892 (1953).
90. J. P. Cough LIN, “Contribution to the Data on Theoretical Metallurgy. XII
Heats and Free Energies of Formation of Inorganic Oxides,” U.S. Bur.
Mines Bull. 542 (1954).
91. M. W. MALLETT, J. W. DROEGE, A. F. GERDs, and A. W. LEMMON, JR., “The
Zirconium-Uranium Dioxide Reaction,” BMI-1210, July 22, 1957.
92. A. F. GERDs, J. W. DROEGE, M. W. MALLETT, and A. W. LEMMON, Jr., “The
Zirconium-Uranium Dioxide Reaction,” Preprint 193, Nuclear Engineering
and Science Conference, Chicago, Mar. 17–21, 1958.
93. S. H. ANoNSEN, R. W. BRAGDoN, C. L. FRENCH, and G. L. MARTIN, “A Com
parison of the Gravimetric and Volumetric Determinations of Free UO,
in Uranium Dioxide,” MDDC–1435, June 1947.
94. D. I. WALTER, “Determination of Oxygen in Titanium,” Anal. Chem. 22,
297–303 (1950).
95. E. J. “Topical Report on the Direct Micro-Determination
CENTER, of Ura
nium Using a Modified Fluorophotometer,” AECD–3006, June 30, 1948.
96. M. W. MALLETT, E. M. BARooDY, H. R. NELSON, and C. A. PAPP, “The Diffu
sion and Solubility of Nitrogen in Beta Zirconium,” J. Electrochem. Soc.
100, 103–106 (1953).
97. J. P. PEMSLER, “The Diffusion of Oxygen in Zirconium and Its Relation
to Oxidation and Corrosion,” NMI–1177, May 31, 1957.
98. M. W. MALLETT, W. M. ALBRECHT, and P. R. WILsoN, “The Diffusion of
Oxygen in Alpha and Beta Zircaloy-2 and Zircaloy-3 at High Tempera
tures” in Section B of A. W. Lemmon, Jr., “Studies Relating to the Reac
tion Between Zirconium and Water at High Temperatures,” BMI–1154,
Jan. 3, 1957.
99. D. R. MASH and B. F. DISSELHoRST, “Uranium-Zirconium Diffusion Studies,”
AECD–3701, June 1954.
100. Y. ADDA and J. PHILIBERT, “Diffusion of Y-Uranium-Zirconium,” Compt.
rend. 242, 3081–3083 (1956).
101 S. J. PAPROCK1, E. S. Hodge, D. C. CARMICHAELs, and P. J. GRIPs Hove R,
“Gas-Pressure Bonding of Zircaloy-Clad Flat Plate Uranium Dioxide
Fuel,” BMI-1374, Aug. 28, 1959.
102. R. A. WolfE, F. O. BING MAN, R. L. FISCHER, R. C. LACERTE, and E. F.
Losco, “Eutectic-Diffusion-Bonding of Plate-Type Fuel Elements Contain
ing Ceramic Fuel,” WAPD–211, Aug. 1960.
103. J. E. CUNNINGHAM and E. J. Boy LE, “MTR-Type Elements” in “Proceedings
of the International Conference on the Peaceful Uses of Atomic Energy,
Geneva, 1955,” Vol. 9, pp. 203–207, United Nations, New York, 1956.
104. R. C. WAUGH, “The Reaction and Growth of Uranium Dioxide-Aluminum
Fuel Plates and Compacts,” ORNL–2701, Mar. 23, 1959.
105. R. C. WAUGH and R. J. BEAVER, “Recent Developments in the Powder Metal
lurgy Application of Uranium Oxides to Aluminum Research Reactor Fuel
Elements,” CF–57–9–60, Sept. 16, 1957.
106. A. L. EIss, “Reactivity of Certain Uranium Oxides with Aluminum,” SCNC–
257, Feb. 10, 1958.
Chapter 8

OXIDATION AND CORROSION OF


URANIUM DIOXIDE
S. ARONSON, Editor

8.1 INTRODUCTION

In the first sections of this chapter (Sects. 8.2.1 through 8.2.4),


kinetic studies of the oxidation of UO, by gaseous oxygen and by oxy
gen dissolved in water are discussed. Whereas relatively little work
has been done on the oxidation of UO, by oxygen dissolved in water,
the oxidation of UO, by gaseous oxygen has been studied intensively in
the past decade. While much information has been accumulated, and
plausible oxidation mechanisms have been postulated, several funda
mental aspects of the mechanisms of gaseous oxidation are still
unresolved.
The reactivity of UO, compacts in water is discussed in Sects. 8.3.1
through 8.3.5. The effects of various water conditions such as pH and
oxygen content and of impurities in the UO, compacts on the rates
of corrosion and erosion of UO, compacts are considered. Data are
presented on the stability of UO, compacts in flowing water.
In the concluding sections of the chapter (Sects. 8.4.1 through 8.4.4),
kinetic data obtained on the reduction of UAOs and U.O., to UO, in
gaseous hydrogen are summarized. The reduction of higher uranium
oxides to UO, has been studied less intensively than the oxidation of
UO2. Conclusions concerning the reduction processes can only be
tentative at present. It is evident, however, that the rate laws gov
erning reduction of higher uranium oxides to UO, are different from
those controlling UO, oxidation.

8.2 KINETICS OF OXIDATION OF URANIUM DIOXIDE


S. Aronson

8.2.1. Oxidation of UO, in Air and Oxygen

A discussion of the oxidation of UO, can be conveniently divided


into two temperature ranges in which different oxidation mechanisms
377
378 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

are operative. A process of low temperature oxidation occurs up to a


temperature of about 50° C. Above this temperature, a more exten
sive oxidation process occurs, and different oxidation mechanisms
apply.

8.2.2 Low Temperature Oxidation of UO, in Air and Oxygen

(a) Chemisorption of Oaxygen at -183° C

Roberts, Ferguson and McConnell, and McConnell studied the


chemisorption of oxygen on uranium dioxide and on solid solutions of
uranium dioxide and thorium dioxide [1–4]. Roberts found that the
chemisorption of oxygen on freshly reduced surfaces of UO, at 185° -
C was rapid and approached completion within a few minutes. The
rapidity of the chemisorption indicated a very low activation energy
for the process. The amount chemisorbed was primarily dependent
on the surface area of the sample and relatively independent of the
method of preparation of the UO, powder. It corresponded to a
value of 0.55 to 0.68 for W.ZWin, the ratio of the quantity of oxygen
chemisorbed to the quantity required to form a physically adsorbed
monolayer at –185° C as determined from the BET equation [5].
Since most of the adsorption sites were not vacated on high tempera
ture evacuation, it would appear that the chemisorbed oxygen is firmly
bound to the surface.
In solid solutions of uranium and thorium dioxides, the ratio
Vo/Vin decreased linearly with decreasing mole fraction of uranium
dioxide. This linear relationship was followed even when the solu
tions were dilute in uranium. Since the uranium ions are almost en
tirely surrounded by thorium ions in the dilute solutions, it is ap
parent that oxygen molecules reacted with single U" sites in the
primary act of adsorption. Adsorption probably results from the
transfer of electrons from the uranium ions to the oxygen molecules.
inability
its

Thorium does not adsorb oxygen, presumably because of


higher
to

than +4 and, thereby, an


to

as

attain valence state act


a

electron donor.
Ferguson and McConnell made
of

calorimetric study the heats


a

adsorption
of on

oxygen -183° [3].


of

of

uranium dioxide
at

They found that the limit chemisorption was generally between


V./Vn, Incre
of

Roberts [1].
to

0.25 and 0.35 contrast the data


in

oxygen yielded differential heats adsorption


of

of

mental additions
which decreased as the surface became covered. The initial heat of
adsorption was 55+2 kcal/mole. The authors discussed this de
creasing heat adsorption surface heterogeneity.
of

of
in

terms

See Chap.
3.
*
OXIDATION AND CORROSION OF URANIUM DIOXIDE 379

(b) Oa’idation at Temperatures of -130°C to 50° C.

The oxidation of UO, in oxygen attemperatures of -130°C to 50°C


was studied by Anderson, Roberts, and Harper [6]. In addition to
chemisorption, another oxidation process occurs at these temperatures.
This absorption of oxygen was found to follow a logarithmic rate law
expressed by
do/dt = Ke-ºc Eq. (8.1)
where c is the amount of oxygen absorbed in time t, and K and u are
constants. For c=0 at t=0, the integrated equation can be written
c=1/u ln (Kut--1) Eq. (8.2)
and for Kut- >1
c=2.3/u (logo Ku-H logiot) Eq. (8.3)
Parallel, linear plots of c vs log tº were obtained for a UO, sample at
several oxygen pressures, indicating that the value of u was approxi
mately independent of pressure. The logarithmic rate law was found
to hold for amounts of gas ranging from 0.05 Wim to about 2 Wm. The
oxidation of UO2–ThC), solid solutions also followed a logarithmic
rate law. The rates of absorption decreased sharply when the mixed
crystals were dilute in uranium.
Anderson, Roberts, and Harper found that annealing oxidized
samples at temperatures above 100° C in vacuo permitted additional
oxygen to be absorbed upon further oxidation at the lower tempera
tures [6]. Holding the oxidized sample in vacuo for several hours
at the reaction temperature itself did not permit additional oxygen
to be absorbed upon readmitting oxygen. The results indicated that
the absorbed oxygen migrated away from the original positions into
the interior of the particles at temperatures above 100° C. Addi
tional oxygen could then be absorbed into the original sites.
Some oxidations were performed on UO, powders of very small
particle size (average particle radius equal to about 0.025p). These
preparations were oxidized regularly and slowly by exposing them
to oxygen first at -183°C, then at -78°C, and then by slowly rais
ing the temperature to room temperature. After standing in a des
sicator for one to two weeks at room temperature, the samples were
found to have a composition of UO2 re-UO2.1s. The quantity of
oxygen absorbed by the logarithmic process was about 2.2 Win. Den
sity measurements, made on several of these samples after exposure
to oxygen at -183°C, showed a density decrease; further oxidation
at temperatures up to room temperature caused an increase in den
sity. The X-ray diffraction patterns of the final oxidized products
were those of a slightly distorted UO, structure with a contracted
-
unit cell.
380 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

It was the opinion of the investigators that the low temperature


oxidation process occurred by the filling of interstitial positions in the
UO, fluorite structure with oxygen anions [6]. The contraction in
lattice parameter was similar to that occurring on high temperature
oxidation where an interstitial mechanism was postulated (see Sect.
8.2.3). The increase in density with oxidation at these temperatures
supported an interstitial oxygen mechanism rather than a uranium
vacancy mechanism. The initial decrease in density on oxidation at
-183° C was explained by the formation of a chemisorbed layer of
oxygenof lower density than that of the body of the particle.
It was estimated that the low temperature oxidation process was
limited to a region within 50A of the surface. This suggested the
possibility that the mechanism was similar to that involved in the
formation of thin oxide films on metals [7]. However, the authors
pointed out that an adequate theoretical treatment of the low tempera
ture oxidation of UO, awaits further study.

8.2.3 High Temperature Oxidation of UO, in Air and Oxygen

Several detailed studies have been made of the oxidation of UO.


in air and in oxygen at temperatures above 100° C. In general, UO, is
oxidized to a tetragonal oxide of composition UO, s—UO,... At tem
peratures above 250°C, a second oxidation step occurs in which the
tetragonal oxide is converted to orthorhombic UAOs.

(a) Oa’idation of UO, to Tetragonal Oaxide

Jolibois oxidized UO, in airand noted that oxidation proceeded to


UAO, at about 220° C [8]. Additional oxidation to UAO, commenced
at about 300° C.
Gronvold and Haraldsen studied samples of UO, oxidized in oxygen
at temperatures of 100° to 265° C [9]. They noted that UO, was
oxidized to at least UO, as at 150° C and that the density of the powder
increased from 10.8 g/cc for UO, to 11.05 g/cc for UO, s. At tem
peratures of 200° to 250°C, the oxygen content exceeded UO.s, and
a tetragonal phase was found. The lattice constants of a sample of
composition UO2 as were a=5.38A, c=5.55A, and the c/a ratio was 1.03.
The density of the UO, as was 10.00 g/cc, a much lower value than
that for the UO2. -
Alberman and Anderson studied the oxidation of uranium dioxide
powder in air at temperatures of 100° to 180° C [10]. The com
position of the starting material was UO, os.” The powder had a nomi
nal particle size of 5u, but the primary particles were estimated to be

* The composition was stated to be UO2 on in the paper. The correct value was given
In a latter communication (see Chap. 6) [11].
APPI'lly OXIDATION AND CORROSION OF URANIUM DIOXIDE 381

ººº:
tº 0.2u from electron microscopic de
01
ºw to studies. Reaction rates were
from the weight changes the samples during oxidation.

on
termined
The

parabolic form

of
the rate curves indicated that the reaction was
X-ray patterns were taken the partially oxi

of
diffusion controlled.

dized samples. UO, was retained samples

of
The fluorite structure

in
hºllº

no
compositions up

to
UO2.2. The diffraction patterns showed
of

| displacement, systematic
broadening changes the relative inten

in
or
|| |
compositions between UO, as-UO, tetrag

sa,
At
of

the
sitieslines.

a
mºtiº
º
Onalphase was present. This phase, which was very similar that

to to
ºilº UO, UO2.2,
had c/a ratio which increased from about

at
1.031
of

º
a

1
Arik
The X-ray data are given Table 8.1.

in
at

U02.3a.
|
Anderson, Roberts, and Harper studied the oxidation UO,

of of

in
sigº
temperatures pressures
of
at

to

at
and 500 mm

C
Oxygen 130° 180°
miſſ
Hg [6]. The UO,
Hg

10-4 mm samples were reduced hydrogen

in
to
th:
º

I,
bring the O/U ratio down
to

each run order The


in

to
before 2.00.
alſº
particle diameters
their two powder preparations,
of

deter

as
average

mined from gas adsorption measurements, were 0.41p and 1.0. Many
the

oxidation experiments were continued until the oxidation rate


of

miº
is,

no

of
became zero, that until measurable absorption oxygen oc
24

hº hours. The limiting compositions


in

to
reached were some
º'
occurred
pressure. The compositions and brief descrip
of

extent function
a
a
tion

the X-ray characteristics


of
º:

the final products are given

in
of
A.
0.
+

82.

The sign implies that oxidation had not completely


P
ſº

Table
iſ

stopped. The UO, structure


In

oxida
as

of

described the course


a.
is

diffraction lines broadened, indicating


the

of
tion, contraction the
a

"it cell, until the U.O., structure, designated


8,
to as

was reached.
Oxidation much beyond the composition UO, led tetragonal struc
as
y”

y”

tures and eventually structure (see Chap. 6). The phase


to

the
tetragonal structure with c/a ratio equal
on to

1.031.
*

a
a

Surface area measurements were performed the products


of

the
by

Only samples per


27

11

varied
in

"midation process. more than


4
the

material, indicating that


of

*nt from surface areas


the unoxidized
oxidation. Density
on
if

little,
area change occurred
any, surface
by

*asurements, reported Anderson, showed that the density


of

oxidation, from approxi


of

"nium oxide increased with the degree


mely 10.89 g/cc for UO, about 11.21 g/cc for UO, [14, 15].
of to
os

as

Perio studied the oxidation UO, air temperatures


to
at

of
in

120°
Two powders were used: UO,
C

175° [11, 16]. specific


with
on

UO, powder which had been exposed


1.2

mº/g, and
of

to to
an

surface
*
a

for several years. He found that the UO,


air

...”
.

be
on

could oxidized
"PProximately UO, UO,
ſº

powder UO, Perio deter


to

to.

and the
as

in

"d
by
on

the degree
of

of

oxidation number samples, both weight


a
ºº:

by
º

chemical analysis. oxygen


on

of

oxidation and The contents


by

*P*rtially oxidized samples were lower when determined chemi


382 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 8.1—LATTICE PARAMETERS OF PHASES OCCURRING IN THE


OXIDATION OF UO,

Oxidation Tetragonal phase Cubic phase


Reference O/U ratio temperature
(° C)
a(A) c(A) c/a aſA)

9------------ 2.43 200–250 5. 38 5. 55 1.03 |----------


10------------ *2.08 120 --------|--------|-------- 5. 468
2. 13 180 --------|--------|-------- 5. 467
2. 17 180 --------|--------|-------- 5. 466
2. 23 200 ----------------|-------- 5. 467
2. 27 120 5. 440 5. 493 1.010 |----------
2. 29 180 5.431 5. 536 1.019 |----------
2. 33 175 5. 397 5. 565 1.031 l----------
11**_ _ _ _ _ _ _ _ _ _ 2. 32 120 5.447 5. 400 0.991 |----------
12------------ 2. 10 230 ----------------|-------- 5. 467
2. 25 230 5. 41 5. 50 1.017 ----------
2.25 275 5.43 5. 50 1.013 I----------
2. 27 275 5. 41 5. 517 1.020 ! ----------
2. 30 230 5. 41 5. 55 1.026 l----------
2. 30 275 5.404 5. 520 1. 021 |----------
2. 32 275 5. 398 5. 534 1.025 ----------
2. 34 230 5. 38 5. 55 1.032 ||----------
2. 35 275 5. 387 5. 549 1.030 ----------
13------------ 2. 07 200 ------------------------ 5. 4693
2, 11 200 5. 413 |-------- 1. 016 5. 4680
2. 16 200 5. 409 |- ------- 1. 020 5. 4664
2, 20 200 5. 404 |- - ------ 1. 022 5. 4673
2. 24 200 5. 403 |- -- ----- 1. 026 5. 4670
2. 27 200 5. 405 |- -- --- -- 1. 025 5. 4661
2. 29 200 5. 405 |-------- 1.027 5. 4694
2. 31 200 5. 386 |-------- 1. 031 5. 4680
2. 33 200 5. 380 |- - - - - - - - 1.032 ----------
13f----------- 2. 06 280 5. 394 |-------- 1. 031 5. 4686
2. 13 250 5. 393 |- - - - - - - - 1. 029 5, 4701
2, 19 250 5. 393 |- - - - - - - - 1. 029 5. 4684
2. 23 250 5. 392 |- - - - - - - - 1. 031 5. 4692
2. 26 250 5. 390 - - - - - - - - 1. 031 5. 4697
2. 30 250 5. 395 |- - - - - - - - 1. 030 |----------
2. 32 250 5. 388 |_ _ _ _ _ _ _ _ 1.032 |----------

*The compositions have been corrected by adding 0.04to each value given in Ref. 10. (See footnote
in Sect. 8.2.3(a).)
**Perio found no variation in the parametersof the tetragonal phase with degreeof oxidation.
fThis work was done on pellets.

cal analysisfor U (total) and U (IV) than by the weight changes.


The values approached each other as the oxidation approached com
pletion. Perio suggested that the discrepancies in composition might
be due to adsorbed oxygen which was not detected by chemical
analysis.
OXIDATION AND CORROSION OF URANIUM DIOXIDE 383

TABLE 8.2–FINAL PRODUCTS OF OXIDATION OF UO, [6]

Specimen Tempera- | Pressure | Final com- X-ray characteristics


ture ("C) (CIn) position

h 184 49 2.36 Tending to y”


182 45.7
A pure grade of commercial uranium 0.45 2.28
dioxide; density 10.15g/cc (measured 183 0.48 2.31
with helium); averageparticle diameter 24 >2.32
1.01(gas adsorption). 183 3.1 2.32 || 8, some y”
184 0.60 2.22 | ot-8
182 0.011 2.27 | Distorted 8

or,
182 0.0021 >2. 21 contracting
173 0.011 2.21 ot-H contracted

a
|
C3 154 48 2.38 y”
155 0.6
Prepared by reduction UiOs with CO 48 >2.37
of

700°C: density 10.90g/cc (measured 155 3.2 2.36 Y”


at

|| ||
8, 8,
with helium); averageparticle diameter 154 1.0 2.29 no

Y
0.41m(gas adsorption). 154 1.0 2.26 poorly defined crystal
153 48 structure
2.24
0.
5

8,
155 0.49 2.26 poorly defined crystal

||
structure

6 8,
154 0.0135 2.24 (?)
| ||

Y
154 0.0020 2.27
143 0.0135 2.24

Perio’s X-ray data showed that the tetragonal phase composition


of

UO, parameters did not change


of

had c/a ratio 0.991 and that the


as

UO,
of

spectra
of

with the amount oxidation. He noted that the


to
UO, indicating
as,

phase
of

were similar that the same


to

those was
present. Some spectral anomalies indicated the possible presence
of
a

second phase.
several samples oxidized
of

Surface areas
at

120° were measured.


C

The area increased from 1.2 m”/g for UO, 1.7 m"/g for UO2.24,
to
on

mº/g for UO2.so, and m”/g


as.

of

for UO2 The densities


to
to

1.8 2.1
the UO.
at

several samples oxidized


of

were less than those


C

120°
the starting material, 10.5 10.7 g/cc
as

as

compared
to

to

used
11.0 g/cc.
Hering UO, powders studied
of

and Perio observed that one the


up

cubic structure compositions UO, during oxida


at

to

retained
as
a

[17]. The lattice parameter decreased linearly with


at

tion
C

120°
oxidation, going from value for UO,
of

of

to

extent 5.468 5.431


A

Å
a

on

for UO2.sa. The structure did not change annealing the UO2
as

sample for 240 hours. Perio also observed the cubic struc
at

120°
samples having composition higher than UO2 [16]. Data
in

as

ture
a

the particle size these UO, powders were not given. possible,
on

of

It
is

however, that these powders were very fine and that this might have
accounted for the fact that the tetragonal structure was not formed.
Vaughan, Bridge, and Schwartz prepared active oxide which re
an
384 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

tained cubic symmetry with a contracted lattice up to UO2.0 upon


spontaneous oxidation in air at room temperature [18]. Hoekstra
and Siegel reported that UO, prepared by thermal decomposition of
UO.C.O., at 300° C in a closed system will oxidize to UO2.5 (cubic) at
room temperature [19].
Aronson, Roof, and Belle studied the oxidation of UO2 in dried air
and in oxygen at temperatures of 160° to 350° C and at an oxygen
pressure of approximately 150 mm Hg [12]. In agreement with the
other studies, parabolic curves, suggestive of a diffusion mechanism,
were obtained. The compositions of the two powder preparations used
were UO, on and UO.o., and the respective surface areas were 0.60
m*/g and 2.4 m”/g, corresponding to average particle radii of 0.5
and 0.13p. X-ray studies of partially oxidized samples were made at
230°C and 275° C. Only the cubic UO, structure was present up to
compositions of UO, 1a–UO2.1s.
At higher compositions, an additional phase appeared, a tetragonal
phase with an increasing c/a ratio. The c/a ratio approached a value
of 1.031 at about UO, a, , oa, the limit of oxidation throughout the
range of temperatures studied. Lattice parameter measurements are
given in Table 8.1. Surface area measurements were made on samples
of UO, or which were fully oxidized at 230° C and 270° C. The new
surface area values were 0.72 m”/g and 0.61 m”/g, respectively,
indicating that the oxidation did not cause appreciable- particle
breakdown.
Blackburn, Weissbart, and Gulbransen investigated the oxidation
of sintered UO, pellets as well as UO, powder at temperatures of 100°
to 280° C in oxygen [13]. The powder had a surface area of 0.57
m*/g, corresponding to an average particle diameter of 0.48p. This
was the same type of powder used by Aronson, et
al.

(surface area
0.60 m.”/g) [12]. the pellet sections used was
of

The surface area


about 0.030 mº/g. The UO, samples were reduced hydrogen prior
in

to oxidation.
Blackburn, al., found that the limiting composition over the tem
et

perature range studied was UO, saroo, [13]. In their X-ray studies,
oxidizing the pellets
on

new phase appeared


of

No
to

lines UOzos.
a

change occurred the lattice parameters either the UO, phase


of
in

or
of

of

the tetragonal
phase the pellets
on

continued oxidation
at

250°
tetragonal phase pellets
of

The c/a ratio


C.
to

at

280° the 250°


C

UO, powder, the


of
In
be

the oxidation
to

was found about 1.031.


tetragonal phase appeared UO, oxidation temperatures
of
at

at
in

the cubic UO, phase was


C.

The lattice parameter


of

150° and 200°


again independent the degree oxidation. However, the parame
of

of

limiting c/a ratio


of

the tetragonal phase varied up


of

to

ters 1.019
a

for UO2 and up 1.032 for UO.s,


of

cſa ratio
at

to

at
as

150°
C

a
OXIDATION AND CORROSION OF URANIUM DIOXIDE 385

200° C. Some of the X-ray data are presented in Table 8.1. Although
the lattice parameters of the tetragonal phase varied with composi
tion in the case of the powder samples, the cell volume, calculated from
the X-ray parameters, remained constant.
Blackburn measured the densities of UO, powder samples oxidized
at 200° C. The density increased from 10.17 g/cc at UO, o, to 10.50
g/cc at UO, The change density with composition
as.

in
shown

is
Table 8.3.
in

TABLE 8.3—CHANGES IN DENSITY ON OXIDATION OF UO,

Reference O/U ratio Ox1dation tem- Density (g/cc)


perature (°C)

9--------------------------------- ------------ 10, 80

2. 2
34 150 11. 05
2.43 200–250 10. 00
16-------------------------------- ------------ 11.

7 0
2

not given 120 10. 5–10.


15-------------------------------- 2.03 |------------ 10. 89
10. 95
2. 2. 2. 2 2. 2. 2. 2.

14 120
19 140 11. 05
27 140(?) 11. 13
34 150–165 11. 21
13-------------------------------- ------------ 10.

1
04 200 10. 17
08 200 10. 24
12 200 10. 23
2.17 200 10. 38
10. 36
2. 2. 2.

21 200
25 200 10. 54
28 200 10. 50

The oxidation UO, tetragonal oxide appears


be
of

diffusion
to

to

controlled process, since parabolic oxidation rates are observed. The


the diffusion process uncertain, however. Two simple
of

nature
is

by

diffusion mechanisms are possible. Oxidation could occur the


growth surface layer
of

tetragonal oxide definite stoichiometry


of

of
a

the UO, particle. The rate


of

into the body


be
of

oxidation would
by

oxygen through the tetragonal


of

of

controlled the rate diffusion


UO, Oxidation might also result from the
to

oxide the interface.


oxygen into the body the UO, particles without the
of

of

diffusion
by

new phase. Oxidation could also occur


of

formation some com


a

bination of the two mechanisms.


by

of
be

assumed that oxidation occurs the inward diffusion


It

can
oxygen rather than by the outward diffusion This as
of

uranium.
in
by

sumption supported the density measurements. An


of

most
is
386 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

crease in density with degree of oxidation would be expected if oxy


gen were diffusing into the solid, whereas diffusion of uranium out
ward would cause a decrease in density. The measurements of Ander
son and of Blackburn, et al., on uranium dioxide and those of Ander
son, et al., on uranium dioxide-thorium dioxide solid solutions show an
increase in density with extent of oxidation [13, 15, 201. Anderson
observed that solid solutions of uranium dioxide and thorium dioxide
retain their fluorite structure on oxidation and that a contraction of
the lattice occurs [20]. However, some of the measurements of Perio
and of Gronvold and Haraldsen show a decrease in density on oxida
tion [9, 11, 16]. The density measurements are listed in Tables 8.3 and
8.4. The assumption of oxygen diffusion is also supported by evi
dence from studies of the rates of self-diffusion of oxygen and uranium
in UO, which indicate that oxygen is more mobile than uranium in the
UO, lattice (see Chap. 7) [21, 22].

TABLE 8.4—DENSITIES OF THE FLUORITE PHASES, U, Thu-, O2+, [20]

Reduced phase Oxidized phase

y Density” (gſcc) x* Density” (gſec)

0.146----------------------------- 10. 00 0. 10 10. 03


0.244----------------------------- 9. 94 0. 16 10. 00
0.244----------------------------- 9. 94 0. 16 10. 13
0.244----------------------------- 9. 68 0. 16 9. S2
0.498----------------------------- 10. 27 0. 30 10. 50
0.783----------------------------- 10. 52 0.33 10, 81

...
"Samples were oxidized in air at 500°C
helium or toluene was used as the displacing fluid.

Magnetic susceptibility measurements have been made on samples


of uranium oxide of composition between UO, and UAOs [23].
The authors concluded that oxygen enters the UO, lattice interstitially
and deduced information concerning phase boundaries in the UO
UAOs region (see Chaps. 5 and 6). Their conclusions, however, are to a
large degree invalidated by the lack of good sample characterization.
Very little X-ray information on the phases present in the samples was
-
presented.
The two oxidation mechanisms were first considered for application
to UO, by Alberman and Anderson [10]. For quantitative evaluation
of the data it was assumed that UO, powder is composed of uniform
spherical particles which are exposed to a constant pressure of oxygen.
It was also assumed that the surface of each particle becomes covered
instantaneously with a continuous layer of product.
OXIDATION AND CORROSION OF URANIUM DIOXIDE 387

If diffusion through a product phase is rate controlling, the follow


ing rate equation is applicable [24]:
2DHM,t/da?– r/(r-1)– (1–c)*/*[–1/(r-1)][1+ (r-1).c]*/*
Eq. (8.4)
where c is the fraction of transformed material; a is the radius of the
particles; D is the diffusion coefficient; r is the ratio Mido/Mod, ;
d...M.,d, and M, are the densities and formula weights of the reactants
and products, respectively; and II is the difference in the activity of
the diffusing species at the particle surface and at the product-reactant
interface. If the molecular weights and densities of reactant and
product phases are approximately equal, the equation reduces to
2DHMot/doa”-1-(1-c)*/*— (2/3c). Eq. (8.5)

If the diffusion of oxygen occurs in the UO, phase itself, with the par
ticle surface instantaneously saturated with oxygen to the limiting
composition, then an equation analogous to one used in heat conduction
can be used [25]. A concentration gradient occurs along the radius
of the particle, and the integrated rate equation is

c=1–6/"> 1/n exp(-ºn-Dt/a") Eq. (8.6)

The form of the rate curves obtained from Eqs. (8.4) or (8.5)
is,

is similar to that obtained from Eq. (8.6). It therefore, diffi


distinguish between the two mechanisms
on

of
cult
to

the basis the rate


might hoped that one could discriminate between the two
be

data.
It

on

new phases formed on


of

of

mechanisms the basis structural studies


oxidation. However, such evidence from X-ray diffraction studies
has not been conclusive. Alberman and Anderson, Anderson, al.,
et
by

and Aronson, al., believe that oxidation occurs


of

the diffusion
et

oxygen into interstitial positions the UO, fluorite structure [6, 10,
in
by

12, 15]. This opinion supported their X-ray data which show
is

second phase does not appear until the reaction per


30

70

that
to
is
a

cent complete. The tetragonal phase which does appear similar


to
is

UO2 and might the UO, cubic oxide


be

of

distortion
as

considered
a

the region near the surface where the oxygen content high. The
in

is

this phase presumably does not interfere with the kinet


of

formation
ics, since break the rate curves does not occur when the second
in
a

by

phase observed. Additional evidence reported Anderson, al.,


is

et

that oxygen near the surface after low temperature oxidation (see
is

Sect. 8.3) diffuses into the body the UO, upon annealing tempera
of

at

tures above 100°C, indicating that uranium dioxide can support in


terstitial oxygen [6]. Anderson, al., have found
of

studies solid
in
et
U,

Thi-, O2, that for


of

of

solutions >0.5 the limit oxidation


is
y

r=0.32–0.34, the fluorite structure [20]. The sat


of

with retention

57.4789 0–61—26
388 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

uration values for a of found in the oxidation of UO.


0.30 to 0.40
appear to constitute a natural limit on the amount of excess oxygen
that the uranium dioxide lattice can support. Anderson has studied
the oxidation of monocrystalline particles of UO, by following changes
in the profile of X-ray diffraction lines [15]. For oxidation at low
temperatures (less than 180° C) line broadening rather than line
splitting occurs as oxygen is added. The results are in accord with
a diffusion process in which oxygen diffuses along a concentration
gradient from the surface to the center of each particle; they are
not in accord with a mechanism in which oxidation occurs by the
inward growth of a stoichiometric surface layer.
Blackburn, et al., on the other hand, believe that their X-ray evi
dence supports the conclusion that a new stoichiometric phase is formed
[13]. They find lines of a new phase appearing at a composition of
UO, os during the oxidation of sections of sintered UO2 pellets. This
new tetragonal phase has a constant unit cell volume, although the
lattice parameters may vary with the degree of oxidation. Blackburn,
et al., also find that the lattice parameter of the UO, phase remains
constant during the course of oxidation. They contend that a linear
lattice contraction, according to Vegard’s law, would occur if oxygen
were dissolving in the UO, lattice. Such a contraction occurs when
UO, containing excess oxygen is annealed at high temperatures and
also in some cases of low temperature oxidation (see Sect. 8.2.2(b))
[26].
Other evidence, used to support one side or the other, are changes in
surface area and density on oxidation. Anderson, et al., and Aronson,
et al., have found that the surface area does not change significantly
on oxidation [6, 12]. It might thereby be deduced that a new phase
does not form on oxidation, since the formation of a new crystalline
material might strain the lattice and cause particle fracturing. How
ever, the unit cell volumes of the UO, and tetragonal phases are
similar, and one could not, a priori, expect particle fracturing. Be
sides, Perio has found that the surface area increases by a factor of
two on oxidation of UO, at 120° C [11]. Blackburn, et al., have
compared their experimental results of the increase of density with
the degree of oxidation with calculated increases based on either oxy
gen solution or the formation of a new phase [13]. They find that
the best agreement is obtained when the model is based on the forma
tion of a new phase; however, agreement between the experimental
data and the theoretical model is not very good. The authors attribute
the discrepancies to changes in the nature of the voids resulting from
oxidation and formation of the U.O. phase.
Blackburn, et al., proposedan explanation consistent with their
mechanism for the nonappearance of the UAO, tetragonal phase in
OXIDATION AND CORROSION OF URANIUM DIOXIDE 389

the pellet samples below a composition of UO, on and in the powder


samples below a composition of UO2.11 [13]. In the pellet sections,
for which an average particle diameter of 2.7, was calculated, based
on the assumption of uniform spheres, the composition UO2.0, cor
responds to a layer of UAO, 0.2, thick, whereas a composition of UO2.11
for the 0.48, powder samples corresponds to a 0.09, layer. They
suggest that a layer about 0.1, thick may be necessary before the new
phase appears in X-ray diffraction patterns. Therefore, small par
ticle size might account for the late appearance of the UAO, phase
in the work of Alberman and Anderson, and Anderson, et

al.
[6, 10].
Perio, general X-ray
on

on
the other hand, believes that

of
the basis
experience layer 0.009, (90 Å)

be
thick should sufficient for the
a

appearance the new phase [11]. He, thereby, explained the ap


of

UO, for samples having

an
of

pearance the new phase average


at

on

cubic crystallite size


of

0.15p.
study UO, pellets by Smith indicates that
of

of

the oxidation
A

*
of by

oxygen through layer


of
the oxidation the diffusion
controlled
is

a
of

UAO. sintered pellets


of
The oxidation uranium dioxide

in
oxygen pressures 800 mm Hg and temperatures
of

of
450°C
at

to

to
3

0
the first stages
Smith found that of
in

was examined. the reaction


by

an initial rapid oxidation rate was followed constant oxidation


a

of an

rate. The initial rate was explained on the basis adsorption


of

of
oxygen on previously deposited chemisorbed layer oxygen. The
a

rate constant for this process can


be

written
(— 12.8+2.8)
k1=5.5×10−7 exp
P

RT
k,

where the oxygen


in

the rate constant moles O/cm3-sec and


in is

is
P

pressure mm Hg; the activation energy 12.8 kcal/mole. The


is

on

constant reaction rate which followed was explained


of

the basis
oxygen through UAO, (or U.O.) layer
of

of

the diffusion constant


a

thickness (224-1-74A). The diffusion coefficient for this second step


can be written

(–21.9-E 1.4).
ex
=

D(cm°/sec) 5.5X10^* RT
To explain why the UAO, layer does not grow during the period
of

constant oxidation rate, Smith postulated that oxygen diffuses more


rapidly UO, than U.O. Growth
of

the U.O., layer does not


in
in

occur because the oxygen moves away rapidly from the UAO-UO.
interface. The UAO, layer begins grow only when the oxygen
to

concentration the UO, lattice sufficiently high


in

to

becomes reduce
its rate of diffusion.

T. Smith, “Kinetics and Mechanism


of

1, of

the Oxidation Uranium Dioxide and Uranium


*

Dioxide Plus Fissia Sintered Pellets,” NAA—SR—4677, Nov. 1960.


390 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Perio has modified his original position that the oxidation pro
ceeds purely by the formation of a new stoichiometric tetragonal
phase [27]. He believes that his X-ray data are not in agreement with
either of the mechanisms discussed above. Thus, unlike Anderson, he
does not obtain a gradually broadened X-ray profile of the UO2 phase
as oxidation proceeds [15]. Such a broadening would be expected if
oxygen enters the UO, lattice interstitially along a diffusion gradient,
causing a lattice contraction in the surface region. On the other hand,
the tetragonal phase does not appear until relatively late in the reac
tion. Perio proposes that oxidation does occur by the diffusion of
oxygen into the UO, lattice. There is a slight contraction of the
lattice in the region of the particle surface where the oxygen content
is highest. The resulting stresses are relieved by plastic flow, causing
the formation of intragranular boundaries. The diffusion is stopped

is,
at such a boundary until the gradient across it is high, that when
the external layer almost complete. The reaction con
of

oxidation
is

tinues such cycles until the whole particle oxidized.


in

is

is,
The problem determining the mechanism thus,
of

of
oxidation
not yet resolved. Quantitative evaluation
of
their data has been made
by

on

the various investigators


of

of
the basis the two mechan
either
Alberman and Anderson, Anderson, al.,

et
isms discussed above.
Perio, and Aronson, al., applied Eq. (8.6) the data and cal
et

to
culated diffusion coefficients and activation energies for the diffusion
oxygen into UO, [6, 10, 12, 27]. Blackburn, al., calculated rate
of

et
on

constants and activation energies the assumption that stoichio


a
metric UAO, layer formed [13].
is

by

Some typical rate curves obtained Aronson, al., are shown


et

in
Fig. 8.1. The theoretical curves calculated from Eq. (8.6) are also
Fig. 8.1 for purposes plot log
on

comparison. 1/T
vs
of

of
A

entered
D

for the data Aronson, al., Fig. 8.2 for air and
of

shown
in
et

is

oxygen oxidation UO, powders: MCW UO, (average particle


of

two
radius 0.5u), and second powder (average particle radius 0.13a) [12].
a

The expression for the diffusion coefficient obtained from least


a

squares plot, cm3/sec, D=5.5 10-8 exp (-26,300+1500/RT).


in

of is

comparison
of

some the diffusion coefficients and activation


A

by

energies obtained the various investigators Table


in

listed
is

8.5. Since the powders used are not composed uniform spheres,
of

as

was assumed, but are composed particles with distribution


of

of

sizes
a

and shapes, the agreement among the values


of

the diffusion coef


ficients fair.
is

Anderson, Roberts, and Harper studied the pressure dependence


of
of

the oxidation reaction [6]. Some their results are listed Table
in
of

8.6. The authors suggest that the pressure variation the calculated
diffusion coefficient may result from the fact that, although the sur
OXIDATION AND CORROSION OF URANIUM DIOXIDE 391

:
O ExPERIMENTAL POINTS
OTHEORETICAL POINTS

o 200 400
|
600
I l
800
I l
IOOO
I I
12OO
I
14OO
l l |
16OO
l |-l
1800 2OOO

TIME IN MINUTES

FIGURE 8.1. Comparison of Experimental and Theoretical Curves for Oxidation


of UO, [12]. (Courtesy, American Institute of Physics.)

T*(c)
350° 300° 250° 2009 150°
o i I I

-
TöNo
loſiz— O E=26.3 t 1.5kcal/mole -

10°– O -
&
O
s O Alr
E
s |-
O OxYGEN
e Oxygen AND HELIUM
O
-
- O POWDER B -
O &
-4
IOTH-
O
-

- O -

<! -
l
1.6
l
1.7
I
1.8
|
1.9
|
2.O
|
2.
l
2.2
*e
2.3
to 3/T (*K)

FIGURE 8.2. Oxidation of UO2 to UO2.s.o.ca, Variation of Diffusion Coefficient


with Temperature [12]. (Courtesy, American Institute of Physics.)
392 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

face concentration of mobile oxygen quickly comes to a value charac


teristic of the pressure at any temperature, this concentration is not
equal to the limiting composition finally reached. The values of c
used in Eq. (8.6) are, therefore, too low, resulting in small diffusion
coefficients.

TABLE 8.5—COMPARISON OF ACTIVATION ENERGIES AND


DIFFUSION COEFFICIENTS IN THE OXIDATION OF UO,
Activation | Diffusion coefficient
Reference Temperature range (°C) Oxygen pressure (cm) energy (cm2/sec)
(kcal/mole)

10------------------------------ 100–175 Air, atmospheric------------ 27 || 4.1×10-15(175°C)


6---------- Specimen H-------155–185 48--------------------------- 18 8.9x10-is (184°C)
!---------------------------- 20
0.25------------------------- 21
0.013------------------------ 24
Specimen C3------ 130–165|| 48--------------------------- 27
12------------------------ 25
!---------------------------- 22
0.013------------------------ 14–18
Fine powder------- 120–175 Air, atmospheric-----------
27--- - - - - - - 29 | 1.6×10-1s(175°C)
Coarse powder----120–150 Air, atmospheric------------ 25–30
12----------|-------------------- 160–350|| Air, atmospheric; oxygen, 26.3+1.5 2.1x10−1(182°C)
15cm

TABLE 8.6—VARIATION OF DIFFUSION COEFFICIENT WITH OXYGEN


PRESSURE [6]

Temperature Pressure (cm) 101*D(cms/sec)


(° C)

Specimen C3*---------------------
- 154 48 2. 50
155 3.2 1. 52
154 1. 0 0.83
154 1. 0 0.83
155 0.49 1. 67
154 0.0135 0.37
154 0.0020 0.15
Specimen H"- - - - - - - - - - - - - - - - - - - - - - 184 49 8.9
183 3. 1 5.2
184 0, 60 7.7
183 0.48 4, 8

182 0.011 1.9


173 0.0135 1.2

*See Table 8.2for description of powder specimens.

Blackburn, et al., used a more convenient equation than Eq.


(8.4) to evaluate their data by making the assumption that in spherical
particles the increase in thickness of the oxide layer with time is
OXIDATION AND CORROSION OF URANIUM DIOXIDE 393

inversely proportional to the thickness of the layer [13]. They


obtained the formula

[3m/dA |[1-(1-c) /*]= (kt)” Eq. (8.7)

where c is the degree of conversion of UO, to U.O., k is a constant, t

its
is time, m is the sample weight, d is its density, and A is surface
For small values expresson for spheres reduces
of

to
area. the the

c
equation for plates,
mc/dA (kt)*/*. Eq. (8.8)

=
Using these equations, Blackburn, al., obtained the following ex

et
pression for the rate constant:
k=21,670/RT-9473-1-0.15 Eq. (8.9)
ln

cm”/sec and the activation energy kcal/mole


in

where 21.7
is

is
k

by
of

the tetragonal phase given


of

The rate advance the

is
[13].
equation
:

a”-7.69x10-'t exp (-21,670/RT) Eq. (8.10)

the UAO, layer formed and

in
of

where the thickness the time


is

is
a

t
plot log Fig.
vs
of

1/T
in
A

seconds. shown 8.3.


is
k

re

on
Blackburn, al., also studied the effect oxygen pressure
of
et

Their data are shown Table 8.7.


in

It

action rate [13]. was observed


significant change
no

at

that there was the oxidation rate pressures


in

of
atmosphere oxygen. pressure
of

between 0.2 and Below 0.007


a
1

the speed the reaction decreased with decreasing


of

atmosphere,
pressure. The authors suggested that the latter effect may have
oxygen saturate the outer layer
of

resulted from lack


to

sufficient
U.O., causing
of

diminution
in

the rate reaction.


of

TABLE 8.7—VARIATION IN OXIDATION


RATE WITH OXY GEN PRESSURE AT
C

200° [13]
Orygenpressure(atmospheres) 101*k(cm2/sec)
121
0. 0.

97
60 118
0.40 119
0. 0.

20 121
10 68.
0

0.014 59.
5

0074
0.

58.
7

0.0027 31.
5

0.0007
6.

78

The thermodynamic stability the phases formed


on
of

oxidation
is

briefly. For oxidized UO, samples compositions


of

now considered
UO, and UO,
as,

annealing the samples


of
in

between the absence


temperatures disproportion
of
at

oxygen 200° and above causes


C
394 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I I I
O
is H -

$
->14 H -
&
E
S.
.x:
|- -
o
o
T 15 – O-PELLET -
O-POWDER

16 H -
| | l l | l | l
1.8 2.O 2.2 2.4 2.6
lo”/T ("k)

FIGURE 8.3. Logarithm of Reaction Constant vs Reciprocal of Temperature for


Oxidation of UO, in 0.1 Atmosphere O, [13b).

ation of the oxidized structures into UO., and U.O.). The amount of
oxygen dissolved in the UO., phase, denoted by x, increases with
temperature [26, 28]. Very little oxygen is soluble in UO, at tempera
tures below about 300° C [26, 28]. The thermodynamic stability of
the tetragonal phases in the composition range above UO2 as has not
been definitely established. Two tetragonal structures have been noted
by Anderson in the range UO, a-UO, having c/a ratios of 1.016 and
1.031 [14]. Perio has observed an additional structure having a cºa
ratio of 1,010 [16]. The relationships between these structures and
their conversion to other phases have been studied by several investiga
tors (see Chap. 6 for a more detailed discussion).

(b) Oa'idation of Tetragonal Oaside to U,0,

Uranium dioxide is oxidized to UAOs in air or oxygen at tempera


tures above 250° C. Although oxidation to UO, is thermodynamically
feasible, oxidation beyond U.O. occurs only under special conditions
(high oxygen pressures, presence of water) [29, 30].
Aronson, Roof, and Belle studied the oxidation of UOes, roos in
air and in oxygen at temperatures of 260° to 350° C [12]. The two
stage nature of the oxidation of UO, to UAOs is shown in Fig. 84.
The S-shaped form of the second reaction step implies that the proc
esses of nucleation and growth are both important in the kinetics of
the reaction.
OXIDATION AND CORROSION OF URANIUM DIOXIDE 395
1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–H

o 20 4o 60 80 IOO I2O |40 16O. 18O 2OO


TIME IN MINUTES

FIGURE 8.4. Experimental Rate Curves for Oxidation of UO, [12]. (Courtesy,
American Institute of Physics.)

X-ray diffraction
patterns were taken of samples of MCW uranium
dioxide oxidized to compositions between UO, as and UAOs at 300° C.
This composition range appeared to be a normal two-phase region. As
the O/U ratio increased, increasing amounts of UAOs appeared at the
expense of the UO.s, loos present. No variations in lattice parameter
were observed in either the orthorhombic U.O. phase or the tetragonal
UO2.3.−0.03 phase.
Surface area measurements were made on samples of MCW oxide
(surface area 0.60 mº/g) which were oxidized to U.O.s at tempera
tures of 300° and 326°C. The specific surfaces obtained were 2.8 and
2.4 mº/g, respectively. Thus, the oxidation of tetragonal oxide to
U.Os caused a fourfold increase in surface area. The difference be
tween the densities of the two structures probably accounts for this
particle breakdown; the density of the tetragonal phases is approxi
mately 11.4 g/cc whereas that of UAOs is 8.35 g/cc.
The rate data obtained with MCW UO, and powder B (surface
area 2.4 m”/g) are shown in Table 8.8. The data are expressed in
terms of
the ratios of the time required to reach a certain fraction of
transformed material to the time required for the reaction to be half
complete. Some of these data are plotted in Fig. 8.5. It is seen that
the form of the rate curves obtained with MCW powder does not
change significantly with temperature, whereas there is a systematic
variation with powder B.
The data were compared with theoretical calculations given by
Johnson and Mehl for a quantitative treatment of surface nucleation
and growth [31]. The following qualitative physical picture can be
396 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
I I i T- T

O.9 |
O.8 -
O.7 -

O6 K

O.5 -
O Powder B, 278°C

O.4 wº D Powder B, 297°C -

A POWDERB, 325°C

O MALLINCKROOT POWDER,300°C

I l —l —l l l I l I l
O O2 O4 O.6 O-3 I.O 1.2 1.4 1.6 1.8 2.O 2.2 2-4 26
RATIO OF REACTION TIME TO TIME OF HALF COMPLETON

FIGURE 8.5. Experimental Rate Curves for Oxidation of UO, sea.m.to U.Os [12].
(Courtesy, American Institute of Physics.)

given for such a reaction in terms of a process of nucleation and


growth. When UO, is exposed to oxygen, the composition of the
surface of each powder particle reaches UO.s, almost instantaneously.
Nuclei of UAOs start forming immediately on the surface and begin
to grow. The data obtained with powder B can be explained qualita
tively if it is assumed that at the lowest temperatures studied approxi
mately one nucleus is formed on each particle. The reaction ap
proaches first order kinetics with nucleation as the rate controlling
process. If
it is assumed that the activation energy for nucleation is
greater than that for growth, the rate of nucleation increases more
rapidly with increasing temperature than does the rate of growth.
The growth process, therefore, becomes more important as a rate
controlling step. The shape of the rate curves departs from a first
order curve as the ratio of the rate of nucleation to that of growth
increases. At the higher temperature, the form of the curves ap
proaches that of the larger particle size MCW powder. The activa
tion energy for growth of the UAO, phase into the UO2.s. phase has
been estimated from the temperature dependence of the reaction rate
to be about 35 kcal/mole.
OXIDATION AND CORROSION OF URANIUM DIOXIDE 397

TABLE 8.8–COMPARISON OF RATE DATA AT DIFFERENT


TEMPERATURES [12]*
Temperature* C to.5(min) to.1/to..s to.za/to..s to.7s/to..s to.sºſto.;

MCW OxIDE

260---------------- 1,400 0 71 0.77 1. 30 1. 51


200---------------- 1, 330 0.62 0.78 1. 32 1. 56
272---------------- 535 0.68 0. 82 1. 24 1. 43
275---------------- 442 0. 70 0. 81 1. 29 1. 58
276---------------- 348 0.64 0.80 1. 25 1.45
277---------------- 351 0.67 0.82 1. 24 1. 40
299---------------- 98 0. 64 0.82 1. 22 1. 39
300---------------- 96 0.66 0.83 1. 25 1. 41
304---------------- 107 0. 58 0. 74 1. 36 1. 6.3
321---------------- 39 0.63 0.77 1. 28 1. 49
324---------------- 25 0.66 0.86 1. 24 1. 48
326---------------- 29 0. 59 0.79 1. 35 1. 59
348---------------- 9 0. 54 0.72 1. 57 1. 82
350---------------- 11 0. 64 0.82 1. 34 1. 68

Powd ER B

278---------------- 642 0.26 0. 52 1. 89 2. 52


281---------------- 768 0.30 0. 56 1. 75 2. 30
297---------------- 224 0.36 0. 59 1. 70 2. 23
305------ - - --- -- - -- 110 0.48 0.65 1. 59 2. 04
325---------------- 17 0. 53 0. 74 1. 53 1. 94
325---------------- 18 0. 56 0.72 1. 44 1. 78

*Courtesy, American Institute of Physics.

DeMarco and coworkers have suggested that the oxidation of UAO,


toU.O., may consist of three processes. These are (1) the addition of
oxygen to the particle surface until the composition UO2.so is attained,
with of the tetragonal structure; (2) nucleation of an
retention
Orthorhombic phase of the same composition; and (3) addition of
oxygen to this structure until the composition UAOs is attained [32].

82.4 Oxidation of UO, in Water Containing Oxygen

Few studies have been made of the reaction of UO, in water-oxygen


atmospheres. Vier observed that UO, slurries are stable in water at
temperatures of 185° C and 300° C [33]. Kuhlman passed oxygen
free water vapor over UO,
powder samples at 450° C and 600° C for
Several hours and observed no significant change in weight or in
chemical analysis [34]. Work on sintering of UO, compacts in steam
at temperatures above 1,000° C indicates that compositions up to
l
398 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

2.90
H–H–H–H-I-I-I-I-I-I-I-I-I
o First Run 9/177 °C
2.80 H D SECONDRUN

- I
2.7OH
DEVATION
152°C W 127°C
-
2.60H
T
T
-
-
º ſoa-c
-
T

o
H.
2.5o H |

ar - -
-
ſt
*>
in
S 2.40H ſ [
|- I -
2.3OH -
87°C
-
|- I J [In
-
2.2OH. I -
- T
t -
5-3. -
2 ºr
2.IO
WATER
—n.
I-T—º—HT-T-Y-T-I-T-I-T-I-L-L-L
—º- —º- 177°C

2.00
~o 2 4 6 8 IO i2 14 16 18 20
DAYS

FIGURE 8.6. Oxidation of UO, in Water Containing Oxygen [30].

about UO, a can be obtained (see Chaps. 6 and 7). It has been noted
that small amounts of water vapor accelerate the oxidation of UO, in
air or oxygen at temperatures of 113°C and 180° C [18, 35].
Aronson studied the reaction of MCW UO, powder in degassed
water and water containing 25 cc of O2/kg at temperatures of 87 to
177° C [30]. The experiments were made in stainless steel contain
ers having sintered, porous bottoms. A very thin layer of oxide
powder was placed in each container. Water was forced upward
through the bottom of the container by pumping. Fairly homo
geneous oxidation was obtained by this procedure. The water was
checked periodically for constant temperature, oxygen content, pH,
and resistivity. Samples of oxidized powder were analyzed chemi
cally and by X-ray diffraction.
The rate data are shown in Fig. 8.6. Each curve represents the
results from an analysis of three samples. The average
obtained
deviationsare shown by the vertical lines through the points corre
sponding to average values of the composition. The squares in the
data at 104° C and 177° C correspond to data obtained during a
second run at these temperatures.
OXIDATION AND CORROSION OF URANIUM DIOXIDE 399

X-ray analysis of the partially oxidized samples indicates increas


ing amounts of a new phase with increase in oxidation. The new
phase was identified as UO, 0.8H2O by comparing sin”0 values ob
tained from the diffraction patterns with values given for known UOs
hydrates by Dawson, et study

al.

of
In
(see Table 8.9) [36]. the

a
UO, on 180–320°C, Dawson,
of

temperatures al.,

of

et
at
effect water
present agreement with the

be
found UOs. 0.8H2O

in
to

at
180°

C
data presented here. Vier also studied the stability UO, hydrates

at of
water and found that the structure stable in water 60° to 325°C
in

approximated monohydrate [33]. The increase

in
water content
a

on
with degree samples
of

oxidation (shown Table 8.10) measured

in
Hg) overnight
in

dried also indicates that


at
vacuo (0.1 mm

is C
50°
UO,
of

the molecular ratio


to

water about 0.8.


Fig. 8.6 that UO, powder exposed
It

degassed water
in

to
is

seen
O2/kg)
or
hus, water itself
of

(<0.05 did not oxidize


cc

tydrate.
does not react with UO2.
The data indicate that oxidation occurs by the combined reaction
water and oxygen with UO, and that UO, 0.8H2O
of

the direct

is
product. The data are insufficient postulate
to

reaction mechanism.
a

The temperature dependence


of

the reaction rate indicates that the


activation energy low,
10

kcal/mole. The results show that


to
is

TABLE 8.9—COMPARISON OF SIN20 VALUES OBTAINED IN THE OXI


DATION OF UO, IN WATER WITH CALCULATED SINºe VALUES
OF DAWSON AND COWORKERS ON UO, 0.8H2O

Index (hkl) UO, 0.8H2Oſ36] 104°C 127° 1779


C
C

200------------------ 0.0227 0.0241 0.0241 0.0239


020- 0.0500
-
-
-
-
-
-
-

-
-
-
-
-
-
-
-
-
-

0.0514 0.0521 0.0521


111------------------ 0.0506
}

220------------------ 0.0727 0.0751 0.0751 0.0741


400 0.0909 0934 0934 0.0934
0.

0.
-
-

-
-

-
-
-
-

-
-
-
-
-
-
-
-

311------------------ 0.0961 0.0976 0.0986 0976


0. 0. 0. 0. 0. 0. 0. 0.

002------------------
0. 0. 0. 0. 0. 0. 0.

0.

1297 1308 1308


0.

1308
420------------------ 1409 1428
0.

1440
0.

131------------------ 1506
1440
1502
}

202------------------ 1539
0. 0. 0. 0.

0. 0. 0. 0.

1524 1527 1539


511------------------ 1869 1799 1813 1826
331------------------ 1960 1894 1908 1894
040- 2000 2005 1991 2005
-

-
-
-
-
-
-
-
-
-
-
-
-
-
-
-
-

222------------------ 0.2024
||

0.2075 2089 0.2075


0.

600-
0. 0.

2045
-
-
-
-
-
-

-
-
-
-

-
-
-
-
-
-

402------------------ 2206
0.2219
0.

2233
0.

22.19
240- 0.222
-

-
-
-
-
-
-
-
-
-
-
-
-
-
-

620- 0.2545 2576


0.
0.

0.

2576 2591
-
-
-
-

-
-
-
-

-
-
-
-
-
-
-
-
-
400 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 8.10–INCREASE IN WATER CONTENT WITH DEGREE OF


OXIDATION [50]
Temperature (°C) Composition Exposure time (days) Percent water H2O/UO,

2.02 || 0---------------------- 0.05 |_ _ _ _ _ _ _ _ _


177 - - - - - - - - - - - - 2.02 || 4 (degassed water)------- 0.09 |_ _ _ _ _ _ _ _ _ _
177 - - - - - - - - - - - - 2.42 2---------------------- 2. 1 0. S0
177- - - - - - - - - - - - 2.57 || 4---------------------- 2.8 0.80
177- - - - - - - - - - - - 2. 67 || 6- - - - - - - - - - - - - - - - - - - - - - 3. 2 0. 78
177 - - - - - - - - - - - - 2. 85 | 8-- - - - - - - - - - - - - - - - - - - - - 3. 6 0.70
127- - - - - - - - - - - - 2, 17 | 4---------------------- 0.80 0.73
127------------ 2.25 | 8---------------------- 1. 3 0.83
127------------ 2.40 | 12- - - - - - - - - - - - - - - - - - - - - 2. 2 0. 90
127------------ 2. 52 | 16- - - - - - - - - - - - - - - - - - - - - 2.4 0.75

the oxidation of UO, in water, although requiring the presence of


dissolved oxygen, proceeds in a completely different manner from
oxidation in air or gaseous oxygen.

8.3 CORROSION OF COMPACTED UO, IN WATER AND STEAM


W. T. Lindsay, Jr.

8.3.1 Introduction

The corrosion behavior of UO, compacts in static and flowing high


temperature water is discussed and the observations of the effects of
various water chemistry environments and impurities in the compacts
are described in Sects. 8.3.2 through 8.3.5. It should be noted that
corrosion behavior of UO, compacts will depend, to some extent, on
the method of fabrication. Most of the information contained in these
sections is applicable to compacts prepared by the cold-pressing and
sintering method used in the manufacture of the first PWR core (see
Chap. 4) [37]. Many of the compacts used in the corrosion tests
were, therefore, production-run pellets fabricated by this process.
Others were prepared by generally similar processes run on a labora
tory scale. Specific mention is made where compacts were prepared
by hot-pressing or by other techniques differing significantly from
the cold-pressing, sintering process.

8.3.2 Corrosion of UO, Compacts in Static Tests

The corrosion information applicable to static conditions can be


divided into three general categories: (1) long-term tests of com
pacts in degassed, neutral, and high pH water at 650° F and degassed
steam at 750° F; (2) a very large number of short-term corrosion
OXIDATION AND CORROSION OF URANIUM DIOXIDE 401

tests of production pellets with inspection


carried out in connection
procedures for PWR Core application; (3) a series of tests
1 blanket
of intermediate length on compacts doped with various impurities
(either desirable or accidentally possible) for the purpose of deter
mining the effect of additives on the corrosion behavior of the
compacts.
The effect of oxygenated water on the corrosion behavior of com
pacts was not intensively investigated in any series of static tests.
However, it was a common observation in static tests that the pellets
would show signs of some attack on the surface if proper precautions
were not taken for the exclusion of oxygen during a test under nomi
nally degassed conditions. For instance, at the beginning of the se
ries of production-run tests, the pellets occasionally showed signs of
surface attack, as indicated by dulling of the shiny surface produced
during the final grinding operation. Such effects disappeared when
a standard procedure was developed for degassing the autoclaves. In
a 17-day test at 650° F with water containing 25 cc O2/kg H2O, a
greenish-yellow scale was observed on the compacts, indicating oxi
dation to higher uranium oxides. The cores of the pellets remained
unaffected. These observations are consistent with the results of
studies of the effect of oxygenated water on compacts exposed to
flowing water as described in Sects. 8.3.3 and 8.3.4.

(a) Long-Term Corrosion. Behavior

Compacts prepared by cold-pressing and sintering are extremely


stable in high-temperature, degassed water at neutral or high pH.
Hot-pressed compacts do not appear to be quite as stable, but the
slight difference is quite probably due to differences in fabrication.
Aronson conducted long-term tests in which bare UO, compacts
were exposed for nearly a year to 650° F water (both neutral and
high pH) and to 750° F steam [38]. The results are summarized in
Table 8.11. The percentage weight changes were quite negligible for
the cold-pressed and sintered compacts, despite some surface dulling.
Somewhat higher weight losses were found for the hot-pressed com
pacts. Some of the hot-pressed compacts fractured during the course
of the experiment, mostly during handling. Data for the fractured
compacts are not included in the table. Chemical analyses for hex
avalent uranium showed little evidence of internal oxidation during
the year's exposure. Analyses also showed only slight pickup of wa
ter. A number of exposed compacts were examined microscopically.
Comparison with unexposed compacts did not disclose any change in
the internal structure caused by any of the test conditions.
402 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 8.11—LONG-TERM STATIC CORROSION TESTS OF UO, COM


PACTS

Average TvT, water


Specimen type Test conditions Number of weight cºntent content
specimens ! change after test after test
percent, percent,' 'Percent."
.
Cold-pressed and sin- 343 days in refreshed, de- 14 +0.03 0.08 0.06
tered gassed, neutral water at 0.06 o, US
650°C 0.25 tº tº
Hot-pressed. . . . . . . . . . . . . 343 days in refreshed, de- 5 –0. 20 0.31 0.10
gassed, neutral water at
650° F
Cold-pressed and sin- 316 days in refreshed, de- 11 +0.02 0.11 0.06
tered gassed water at pH 10.5 0.20 0.03
with LiOH 0.25 0.07
Hot-pressed . . . . . . . . . . . . . 316 days in refreshed, de- 6 –0. 28 0.05 to dº
gassed water at pH 10.5
with LiOH
Cold-pressed and sin- 302 days in degassedsteam 14 –0. 04 0.27 0 01
tered at 750°F, 2,000psig” 0.64 0.09
0. 50 0- or,
Hot-pressed - - - -- -- - - - - - 302 days in degassedsteam 4 –0. 42 0.43 0 ºr
at 750°F, 2,000psig”

* Analyses carried out on randomly selectedcompacts.


** Includes 128days in a completely static test and 174days in a refreshedsteam system. Hydrogen
buildup from corrosion of the metal vesselwas eliminated by the refreshing.
R EMAPKs.-Slight dulling of surfaceswas observedfor all specimens. No qualitative changein internal
microstructure wasobservedfor any test condition. Some hot-pressedcompacts fractured during handling
as a result of poor fabrication; data are not included for these.

It is apparent that cold-pressed and sintered compacts, and even


hot-pressed compacts, are remarkably inert in degassed, high-tem
perature water and steam. The slight surface dulling, while insig
nificant from a practical viewpoint, may have resulted in part from
traces of oxygen not completely removed from the test apparatus dur
ing the numerous shutdowns when the compacts were removed for
interim examination. The insignificant weight changes, despite con
tinuous refreshment of the autoclave vessels during the tests, appear
to indicate that the solubility of dense, compacted UO, in high-tem
perature water and steam is extremely low. Thus, oxidation and
solution corrosion processes do not play a significant role in affecting
the behavior of UO, compacts in degassed water.

(b) Corrosion of Production Compacts

The reproducibility of the corrosion resistance of compacts pro


duced by the cold-pressing and sintering process is illustrated by an
extensive series of short-term tests. A 4-percent random sample of
the fuel compacts produced for the first core of the Shippingport re
actor was exposed to degassed 750° F steam at 2,000 psig for 20
hours. The total number of pellets tested was 166,000. No visual
OXIDATION AND CORROSION OF URANIUM DIOXIDE 403

evidence of corrosion or other unusual deterioration was noted, and


the original shiny surface produced by grinding was retained [39].
Since dulling is the first noticeable evidence of surface attack, and
since this can occur without significant weight change, it would appear

that the visual observations gave sensitive and satisfactory evidence


that no corrosion had occurred during the 20-hour period. These
results and those of the long-term static tests demonstrate that cold
pressed and sintered UO, compacts have two highly desirable prop
erties: extreme reproducibility of corrosion resistance and excellent
long-term stability in the coolant environment of pressurized water
reactors.

Impurities and Additives


(c)

of

Effect

determining the effect

of
Considerable effort has been devoted to
UO,
on

impurities and additives compacts

of
the corrosion resistance
Early
the Bettis Atomic Power Laboratory development
in

[40].
on

program UO2, corrosion failures were encountered which were


the presence carbon; hence, this impurity was most
of
to

ascribed
intensively studied. Other impurities evaluated were SiO, and Fe.
Numerous additives have been proposed for special purposes, pri

on
marily densification and strengthening, and the effect these
of
several scattered series of
in

corrosion resistance has been evaluated


tests. Among the additives for which some corrosion data are avail
able are TiO2, Cr,0s, CaF2, Al,Os, Ce(NO3), UO, (NOA), UAOs,
ZrO., SiO, CeO2, CaO, BeO, polyvinyl alcohol (PVA), and Sterotex.
With the exception carbon, minor amounts these impurities and
of

of

UO,
do

not appear
of
to

additives affect the corrosion resistance


compacts.

(1) CARBON. Early sintering procedures employed graphite fur


an

and argon atmosphere. Compacts prepared this way were


in

naces
or

disintegrate
to

powder high-temperature water


to

chunks
in

found
This occurrence,
termed desintering, was shown
be
to

to

tests. related
by

which UO,
the

the compacts
of

of

carbon content series tests


to in
a

powder from three different sources was cold pressed form com
two different ways: graphite fur
in

pacts which were sintered


in
a
an
or

argon atmosphere
an

with alundum furnace with


in

nace
a

hydrogen atmosphere. Chemical analyses showed that compacts made


all

powder contained 350 ppm carbon


50
of

to

from three sources


the graphite furnace.
in

when sintered
hydrogen contained
at 10
in

Those sintered the alundum furnace


in

In
or

two corrosion test series, one days


of

Ppm carbon
17

less.
which hydrogen had been added and one
to

650°F degassed water


in

days degassed water without hydrogen addition,


in
at
of
18

650°
F

57.4789 O–61–27
404 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS |

all the high-carbon pellets disintegrated, while the other pellets had no
significant change in weight or appearance. It was speculated that
the water gas reaction
C+ H2O->CO+H,
was responsible for attack on carbon at the points of cementation of
the powder grains, thus, causing the desintering. However, attempts
to reproduce this effect by deliberately doping UO, powder with
carbon were unsuccessful. In one test, 650 ppm carbon was added
to the UO, powder, which was then cold-pressed to form pellets which
were sintered at 1,450° C for 2 hours in either argon or hydrogen.
It was found that argon sintering did not affect the carbon content,
whereas hydrogen sintering reduced the carbon to about 20 ppm.
A subsequent 21-day corrosion test at 650° F gave negative results:
none of the compacts desintered. It was then postulated that a higher
temperature or longer time of sintering would be required to disperse
segregated carbon and to cause it to migrate to the grain boundaries.
In additional tests, carbon was added at concentrations up to 860
ppm, and sintering was done at 1,750° C in hydrogen for 8 hours
or at 1,550° C in argon for 10 hours. Chemical analyses showed that
hydrogen sintering had reduced the carbon to about 20 ppm and
argon sintering had reduced the carbon to a maximum of 50 ppm.
Again, no disintegration occurred in corrosion tests. Finally, carbon
doped pellets were sintered in argon for 10 hours at 1,550° C while
surrounded by graphite. Analyses showed that 220 ppm of segre
gated carbon remained after sintering. A 650° F degassed water
corrosion test produced no effect, and a 20-day test at 750° F in
refreshed, degassed steam likewise caused no significant weight change
and no disintegration.
Carbon-bearing impurities in the UO, powder used as a starting
material for fabrication are not likely to cause desintering in hot
water or steam. Hydrogen sintering may reduce the carbon content
significantly; the carbon which remains, apparently, does not readily
cause failure. PVA and Sterotex, carbon-bearing additives, have been
successfully employed in large scale pellet manufacture [37]. It ap
pears, however, that argon sintering of UO, in a graphite furnace has
produced compacts with poor stability in high-temperature water.
This effect is most likely related to carbon pickup during the sinter
ing, although nitrogen pickup has been suggested as an alternate
explanation.
(2) IRON AND SILICA. The effect of tramp iron was investigated by
mixing iron with UO, powder before pressing and by adding iron to
the center of the compact during pressing. Corrosion tests gave some
visual evidence of iron oxide formation; no pellet disintegration oc
curred, and weight changes were insignificant.
OXIDATION AND CORROSION OF URANIUM DIOXIDE 405

The effect of silica as an impurity was investigated by adding


amounts up to 1,600 ppm to the UO, powder before fabrication.
Sintering in hydrogen was found to lower the SiO, content, whereas
sintering in argon or nitrogen had no effect on the silica content. Cor
rosion tests showed that no pellets disintegrated and that no significant
weight changes occurred. However, higher concentrations of silica can
cause lower corrosion integrity. Bowers, Allison, and Duckworth
reported that sintered UO, bodies containing 10 weight percent SiO,
showed higher weight losses in 650° F and 750° F steam than did
unmodified UO, compacts [41].
(3) OTHER ADDITIVEs. Titania additions up to 1 percent are with
out effect on the corrosion resistance of either extrusions or cold
pressed, sintered compacts in degassed water and steam. Small addi
tions of ZrO, (up to 5 percent), CeO, (up to 5 percent), and BeO
(up to 10 percent) have been found not to affect corrosion behavior
of compacts. Minor additions, on the order of 1 percent, of UAOs,
UO2(NO2)2, Ce(NO3), CaF2, Al,Oa, and Cr,0s are likewise without
effect [40b, 41, 42]. However, larger additions of some of these
materials may affect corrosion behavior significantly.
These results show rather clearly that the corrosion resistance of
sintered UO, compacts is not sensitive to the presence of small
amounts of many oxides or other materials. The excellent repro
ducibility of corrosion resistance found in large-scale production of
UO. pellets is undoubtedly related, at least in part, to this fact.

8.3.3 Erosion and Corrosion of UO, Compacts and Defected UO,


Fuel Rods in Dynamic Tests

The information given in Sect. 8.3.2 indicates that cold-pressed


and sintered UO, compacts of the Shippingport reactor plant type
are inert in static high-temperature water and steam in the absence
of oxygen. The effect ofdeviations of reactor coolant from the static,
nonoxygenated conditions of the tests remains to be discussed. Al
though radiation-induced recombination normally keeps the oxygen
concentration of a deliberately hydrogenated water coolant below
detection, accidental addition of excess air may occur in pressurized
water reactors. In such a case, the coolant pH would drop as a re
sult of nitric acid formation, causing any fuel exposed to the coolant
to be placed in an acid, oxygenated-water environment. Incomplete
neutralization of a base additive present in the coolant could result
in a temporary environment that would be basic and oxygenated.
Similarly, a neutral, oxygenated condition is plausible. Further
more, the erosion of any exposed fuel by the moving coolant may
affect the amount of fuel released in an accident involving loss of
cladding.
406 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

These and other accidental possibilities, such as the presence of


residual oxidizing decontamination reagents, provided the incentive
for a series of dynamic-loop corrosion and erosion tests at Bettis
Atomic Power Laboratory on UO, compacts. The tests were di
vided into three categories: (1) erosion of bare compacts in oxygen
free, flowing water; (2) corrosion of bare compacts in oxygenated
water; (3) corrosion of compacts partially protected by defected clad

all
ding when exposed to oxygenated water. In cases, except where
specifically stated, PWR production-run pellets

or
pellets fabricated

by
the laboratory similar process were used for these tests.
in

a
UO, Compacts
of

(a) Erosion

bare, sintered UO, compacts flowing, oxygen


of

The erosion

in
tests [43, 44]. The

of
Bettis

in
at

free water was studied three series


first of these was conducted while neutral water flowed over the
specimen velocity low temperature
of

28.7 ft/sec and


at

at

of
a

a
150°F. The pellets used during these tests were much smaller than
the production-type, cylindrical compact, but were prepared by the
same techniques. They were mounted plastic holder

of
the side in

a
leading edge about mm high
as

way expose
in

to to

to
such the

2
a

allow the water flow over one exposed surface.


to

water stream and


In

one test, the surface the compact was scored


of

to
simulate the
rough, fractured area that might

as
exposed
be

of
result
to

coolant

a
cladding loss from fuel rods. Photomicrographs the exposed sur
of

the scored specimen are shown Fig. 8.7. The gradual


of

face
in

easily visible. Weight loss data for these


of

erosion the surface


is

tests are plotted Fig. 8.8. The fractured surface specimen un


in

derstandably suffered somewhat higher weight losses than did smooth


specimens exposed
of to

the same conditions. F,


of at
A

second series tests was conducted neutral water


in

525°
20

with water velocities over the exposed compact surface


24
to

ft/sec. different mounting scheme was used for the high-tempera


A

ture experiments. Here, standard-size compacts were prepared with


on
so

that they could rods and suspended


be

center holes mounted


axially flowing stream. Similar test sections also were used for
in
a

the oxidation-corrosion test series discussed below. Weight change


data are shown Fig. Weight losses were much less than for
in

8.8.
the low-temperature erosion tests, presumably because
of

the lower
density and viscosity high-temperature water, the somewhat lower
of

leading
no

velocity over the exposed surface, and the fact that edge
was exposed the concentric mounting arrangement.
in

F,
of

third series
at

erosion tests was conducted also 525° but


A

pH 9.5 LiOH, with


at of

with water with


at

to

10 15

10.5 water velocities


ft/sec, and with O./kg H.O. Hydrogen
23
to

cc

less than 0.1


OXIDATION AND CORROSION OF URANIUM DIOXIDE 407

INITIAL —fººt— I68 HR

4OI HR 785 HR

orienTATION OF PELLET WAS REVERSED AFTER 168 HR

FIGURE 8.7. Erosion of Fractured Wafer in 150° F Water; X7, Reduction


Factor, 14.
408 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I | I I-I-I I I T

525 ° F 8 600 °F
pH IO LIOH

*
105 H
10cc O2 /kg H2O
18- 20 FT/SEC -

600° F NEUTRAL
10cc O2/kg H2O
-
I8 20 TFT/SEC
}.
525 ° F NEUTRAL
10 cc O2 /kg H2O
18–20 FT / SEC
2
104
525° F pH 4-5 HNos
-
10cc O2 /kg H2O
18–20 FT/SEC

*e
•”
150° F 28.7 FT/SEC, NEUTRAL
No o 2 – FRACTURED surface
º
2% A

a’.150° F 28.7 FT/SEC


O No oz NEUTRAL

103

} EROSION

lo? H 525° F No oz NEUTRAL -


2O - 24 FT/SEC –

525° F No oz
pH to LiOH
I5 - 23 FT/SEC

IO L | 1– I I l l I I
IOO 200 3OO 400 500 700 IOOO 2OOO 3OOO 4000
ExPOSURE TIME IN HOURs

FIGURE 8.8. Corrosion and Erosion of Sintered UO, Compacts in High


Temperature Flowing Water.

to 35 cc./kg H2O was added during both 525° F erosion test series.
The results are not shown in Fig. 8.8 because the weight changes
were completely insignificant. Some slight weight gains were re
corded for the longer exposure times.
These tests confirm the inertness of sintered UO, compacts in high
temperature water in the absence of oxygen and establish the order of
magnitude of weight loss rates to be expected from mechanical erosion
OXIDATION AND CORROSION OF URANIUM DIOXIDE 409

processes. It was concluded that even the highest observed erosion


rates gave no cause for concern in the Shippingport plant in view of
the limited bare fuel surface area that would be exposed to the coolant.

(b) Corrosion of Bare UO, Compacts in Oaxygenated Water

Similar tests were conducted with added oxygen at a level of about


10 cc O2/kg H2O, controlled by intermittent bomb addition of oxy
gen to an operating test loop [44]. Neutral pH (6 to 8), high pH
(9 to 11), and low pH (4 to 5) conditions were studied. . The pH was
controlled by continuous circulation of the water through a demin
eralizer or through a mixed-bed ion exchanger in the LiOH or the
HNOs form, as appropriate. The results of the tests are shown in
Fig. 8.8. Figure 8.9 shows the progressive change in appearance
of the specimens during exposure.
Dimensional changes of the compacts agreed qualitatively with the
weight changes, but attempts at quantitative correlation were unsuc
cessful. For the case of pellets exposed to the low pH condition, it ap
peared as if the density were being decreased, as the decrease in diame
ter was always less than would correspond to the weight loss. Micro
scopic examination of polished specimens indicated the presence of
considerable internal porosity of pellets exposed to this condition and
also revealed the presence of a thin, relatively nonporous scale on the
surface about one mil thick.The pellets exposed to neutral water at
both 525° F and 600° F had a scale of a different type; loose, flaky,
and colored yellow, green, and red, as compared with the dark green,
adherent scale of the low pH specimens. No scale was visible on the
highly attacked compacts exposed to the high pH, oxygenated con
dition. It would appear that attack on the specimens exposed to the
low pH, oxygenated water was retarded by the resistance of the ad
herent scale to erosion, but that some internal attack proceeded be
neath the scale nevertheless. Apparently, the loose, insoluble scale
on the specimens exposed to neutral water was subject to erosion and
so did not offer as much protection. It is possible that a more soluble
material was formed as a result of oxidative attack at high pH that
permitted no scale formation and allowed corrosion to proceed unhin
dered. The greatly accelerated attack by oxygenated water at high
pH, which nearly consumed the exposed compacts in 320 hours, indi
cates some such explanation.
Some additional tests were conducted on bare compacts to determine
the effect of residual traces of oxidizing decontamination reagents.
Small amounts of reagents, including hydrogen peroxide, sulfamic
acid, and ethylenediaminetetraacetic (EDTA) acid, were added to a
test loop after the temperature had been established at 525° F. No
410 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
TYPICAL NEW COMPACT

PH 6-8 (NEUTRAL) PH 4-5 (HNOs) PH IO-II (LiOH)

198 HR 16i HR

460 HR 580 HR 320 HR

FIGURE 8.9. Appearance of Compacts after Exposure to Oxygenated


Water; x7, Reduction Factor, 35.

attempt was made to replenish the chemicals by subsequent additions.


Weight changes of the compacts were insignificant, and the surfaces
retained their original shine. Presumably, the reagents decomposed
rapidly. Analyses showed that the oxygen concentration was down
to 0.07 cc O2/kg H2O within 24 hours in one run. Thus, the oxygen
produced by decomposition of the added hydrogen peroxide was rap
idly consumed as a result of corrosion processes in the test loop and
did not affect the specimens.

8.3.4. Behavior of Defected UO, Fuel Rods in High pH, Oxygenated


Water
A series of tests was undertaken at Bettis Atomic Power Labora
tory to evaluate the protection afforded by cladding perforated by
defect holes, since many pressurized water reactors operate with the
coolant deliberately controlled at high pH [45]. These tests were im
portant because bare UO, compacts have a high corrosion rate in
oxygenated, high pH water.
RTIES AND NUCLEAR APPLICAT OXIDATION AND CORROSION OF URANIUM DIOXIDE 411

- NEW COMPACT

BAREPELLETs
AFTER corrosion
TESTING

spaceR

FIGURE 8.10. Assembly of Fuel Elements [45].

Short fuel rods were manufactured by the regular procedures, with


the cladding
the

exception that 5-mil and 20-mil holes were drilled

in
before the pellets were loaded. Three such rods and two bare pellets
The rods were assembled form train, such
in

to
were used each test.

a
asis shown Fig. The rods were sectioned after exposure for
in

8.10.
with LiOH, with added
137

water, pH
10
to

154 hours 600°


in
to

cc./kg H2O.
10
of to

oxygen controlled about


The appearance typical rod after test shown Fig. 8.11.

in
is
a

This rod contained two 5-mil holes, one over the first (upstream) pellet
and one over the last pellet. The last pellet every rod was fractured
in

during welding Pits


be
of

the final closure and could not examined.


were found pellets located beneath all upstream defect holes and
in

pellets beneath the downstream defect holes wherever examina


in

also

tion was possible. Pits also were found beneath the holes rods
in

which contained only one hole. For the rods with two holes, the
and dulling, visible Fig. 8.11,
of

pattern surface discoloration


in

water through the an


of

suggests that there was some circulation


nular gap within the rod. The weight changes the compacts were
of

great and did not appear correlate with position


to

the rod.
in

not

was concluded that little, any, fuel material was actually carried
if
It

from the rods during the tests.


microscopic cross section through the pit formed beneath 20-mil
A

Fig. porous
in

defect hole shown band visible the


in

8.12.
is

is

vicinity the pit, penetrating deeper the pit than


of

of

either side
at
pit
the

This band
as

itself. then becomes less deep the distance from


in
º
412 URANIUM
DIOXIDE:

d
|

13 13 roN
PROPERTIES AND NUCLEAR APPLICATIONS

H
1

*ºº suno
wciuio posodx{Hsq.)
ºg 1·

løn {
6
01

g*1
į

on
a
e

puº.O-ioſionu uopų upu jºpici


ow

…ooº

į I·
annoiji tºs
on
}o nu
„№i

C1pººlponovojº waa aan


\- AN
-
OXIDATION AND CORROSION OF URANIUM DIOXIDE 413

2O-MIL DEFECT IN FUEL ROD


NOTE POROUS BAND.

I loo Ll

CROSS SECTION THROUGH PIT FORMED BENEATH 20-MIL DEFECT

FIGURE 8.12. Pit Formed beneath 20-Mil Defect in Fuel Rod [45].
414 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

the pit increases, disappearing on the parts of the surface which show
no discoloration, dulling, or other indications of attack. As mentioned
in Sect. 8.3.3, the development of internal porosity may be related
to some protection against loss of surface oxidation products by ero
sion or solubilization. In this case, the cladding provides protection
except in the immediate vicinity of a hole, where the formation of the
pit, with its shallower band of porosity, is evidence of eddy diffusion
and mixing of bulk water with that immediately within and below the
hole.
The effects of irradiation on UO, are covered in Chap. 9. How
ever, it is not out of place to note here that any ephemeral oxygen
produced by radiation-induced water or steam decomposition does not
appear to cause visible UO, corrosion in defected fuel rods." Remote
examination of compacts from defected rods tested in in-pile pressur
ized water loops has only very occasionally indicated UO, corrosion.
In the few cases where some yellow or greenish surface coloration was
visible, improper loop operation could not be ruled out as a possible
source of the oxygen required for the oxidation. Deliberate oxygen
addition during testing of a one-hole defected fuel specimen in a
Bettis loop in the Materials Testing Reactor did not result in any
evidence of increased release of gaseous or soluble fission products
from the defect. There were no indications of gross release of fuel
or long-lived, leach-resistant fission products during the period of
operation with oxygen [46]. A preirradiated, clad UO, pellet, with
a clad defect, was found to be resistant to release of long-lived fission
products in an out-of-pile test with oxygenated water [47].

8.3.5 Summary and Conclusions

Sintered compacts of UO, have been found to be extremely resistant


to corrosive or erosive attack in high-temperature water and steam in
the absence of oxygen. This resistance to attack is insensitive to the
presence of many plausible impurities or desirable additives. Only
the pickup of carbon on sintering in a graphite furnace with an argon
atmosphere has been found to cause a loss of fuel stability.
In the presence of oxygen, surface attack can occur in neutral, low
pH, or high pH water at high temperature. The attack is most rapid
at high pH and is somewhat retarded at the neutral and low pH con
ditions, probably as a result of insoluble scale formation. When the
UO, compact cladding is perforated to simulate cladding defects, the
corrosion resulting from the most severe condition (oxygenated, high
pH water) is limited to slight pitting and insignificant surface attack.

• Under certain irradiation conditions, however, UO, may exhibit corrosion behavior in
water (see Chap. 9).
OXIDATION AND CORROSION OF URANIUM DIOXIDE 415

It appears to be highly unlikely that oxidative corrosion or erosion


processes could result in significant loss of fuel from rods containing
sintered UO, compacts.

8.4 KINETICS OF REDUCTION OF U.O., AND U,O, TO UO, IN


HYDROGEN
S. Aronson

8.4.1 Introduction

Uranium dioxide can be prepared from higher uranium oxides by


reduction with hydrogen, carbon monoxide, and ammonia (see Chap.
2). The discussion in Sects. 8.4.2 through 8.4.4 is concerned pri
marily with studies which clarify the fundamental mechanisms in
volved in the gaseous reduction process. After a brief discussion of
the adsorption of hydrogen on UO, the reduction of U.O., to UO,
in hydrogen is treated; discussion of the more complex reduction of
U.Os in hydrogen follows. The reduction studies have, generally,
been made in hydrogen.

8.4.2 Chemisorption of Hydrogen on UO,

Roberts found that hydrogen is not appreciably chemisorbed on UO,


powders below 400° C [48]. At temperatures of 500° to 700° C an
activated adsorption occurs, and high surface coverages are attained.
The rate of hydrogen adsorption increases with increasing tempera- .

ture and, on one UO, powder, reached a final value in about 40 minutes
at 615°C and in about 4 minutes at 700° C. An approximate activa
tion energy of 35 kcal/mole was calculated by comparing the rates of
adsorption at half the saturation adsorption at 615° and 700° C. The
maximum amounts of H. adsorption as a function of pressure are
plotted in Fig. 8.13 for temperatures of 615°, 670°, and 700° C. The
maximum amounts of hydrogen chemisorbed were about 1.0 to 1.6
V., where Vm was the quantity of oxygen required to form a physically
absorbed monolayer at -183° C according to the BET equation.
From these values, Roberts estimated that approximately one mono
layer of hydrogen formed on the uranium oxide surface.

8.4.3 Reduction of U,O, to UO, in Hydrogen

Aronson and Clayton studied the reduction of U.O., powders in gas


eous hydrogen at temperatures of 400° to 600° C and at pressures of 20
to 500 mm Hg [49]. They used three U.O., powders that were prepared
from UO2 powders by oxidation in oxygen below 200° C to UO2 as
and annealing either at 200° C or at 800° C. The surface areas of the
416 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I- I T I

615°C

670 °C

O.5 H.
| | I |
O |O 2O 3O 40
PRESSURE IN cm Hg

FIGURE 8.13. Maximum Adsorption of Hydrogen on UO, as a Function of


Pressure [48]. (Courtesy, The Chemical Society, London.)

three powders, designated powder A, B, and C, were 0.73, 1.75, and


0.45 m”/g, respectively.
Typical rate curves for the reduction of the three U.O., powders at
constant temperature and pressure are shown in Fig. 8.14. The
shapes of the rate curves obtained for the three powders were fairly
similar. The curves were approximately linear until the reaction was
about 75 percent complete in the case of powder B and up to about 50
percent complete in the case of powders A and C.
Surface area measurements were made on samples of powders A and
C reduced at 410°C. The values obtained on the two reduced powders
were 0.80 and 0.45 mº/g, respectively. Thus, little particle breakdown
occurred during reduction. X-ray diffraction patterns were taken at
room temperature of samples of powders A and C that were partially
reduced at 470° Similar results were obtained on both powders.
C.
A mixture of U.O., and UO., was observed on partially reduced Sam
ples of composition above UO,
10.

ar,

The subscript
of

the value which


indicate the possibility that the uranium
to

less than 0.10, used


is

is

dioxide phase was not stoichiometric.


of

The relative amount the


UO., phase increased with the degree
of of

The amount
of

reduction.
U.O, phase decreased with the degree reduction and was not de
composition below UO, Thus,
no.

samples appears that


of

tected
in

it

the reduction process initially converted U.O., into nonstoichiometric


30.10). The oxygen content the latter phase was
of

UO2,...
(a

then
composition UO.o.o.o. was attained.
of

reduced until
a
OXIDATION AND CORROSION OF URANIUM DIOXIDE 417

:
O5

O5

O.4.

O3

o2

Oil
- - l
o ro 2O so 4O 5o 6O 7o 80 90 IOO
TIME IN MINUTES

FIGURE 8.14. Typical Rate Curves for Reduction of U.O., in Hydrogen [49].
(Reprinted with permission from S. Aronson and J. C. Clayton, “Kinetics of
the Reduction of U.O., in Hydrogen,” Pergamon Press, Inc., 1958.)

A reduction mechanism comprising three steps was postulated for


the reduction of U.O., : (1) the reaction between hydrogen and lattice
oxygen at the particle surface with the formation of water; (2) the
transformation of U,O, to UO., at the interface between the phases;
and (3) the diffusion of oxygen, released at this interface, through the
UO,... phase to the particle surface.
The surface reaction was chosen as the slowest and, hence, rate
controlling step on the basis of the following evidence: (1) the rate
curves were not parabolic, eliminating step 3; (2) step 2 as the rate
controlling process was inconsistent with the fact that the reaction was
found to vary strongly with hydrogen pressure; and (3) the accumula
tion of oxygen in the UO, phase to form the nonstoichiometric initial
product UO., indicated that the rate processes in the solid were more
rapid than those at the surface.
Since the surface area and gas pressure remained constant during
reduction, the simplest mechanism would give a constant reaction rate
in the case of a powder of uniform particle size. Possible reasons
given for the deviations from linearity in the latter stages of the reac
tion were the following: (1) the U.O., powder particles had a range
of sizes and shapes, and the smaller particles were consumed earlier,
causing a decrease in reacting surface and, hence, in reduction rate;
(2) after the disappearance of the U.O., phase, the rate of reduction
of UO2., may have decreased with the decreasing value of a and
(3) the accumulation of adsorbed water formed on reduction could
possibly have impeded the reaction.
418 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

-loo H H-T-I-T-I-T-I-T-I-T-I-T-I-T-I-T-I
- 10.2 H. D 500 mm Hg
O 100 mm Hg
- 10.4 A 20 mm Hg -
- IO.6

.º - IO.8
- I.O

- I 1.2
: - I 1.4

- I 1.6
-n a 1––––––––––––––––A–1–1––––––.
1.16 1.2O 1.24 1.28 I.32 I.36 I.40 l.44 I.48 (.52 .56 I.6O L64

loº/T ("K)

FIGURE 8.15. Rate Constant as a Function of Temperature for Powder C [49].


(Reprinted with permission from S. Aronson and J. C. Clayton, “Kinetics of
the Reduction of U.O., in Hydrogen,” Pergamon Press, Inc., 1958.)

Rate constants were determined from the slopes of the initial linear
portions of the plots of weight loss versus time. Rate constants were
calculated in terms of the number of moles of oxygen released per
square centimeter of surface per second. A plot of log rate constant,
k, versus 1/7' for powder C at hydrogen pressures of 20, 100, and 500
mm Hg is given in Fig. 8.15. Similar plots were obtained for the
other two powders. The calculated activation energy for the reduc
tion of U, O, was 25+3 kcal/mole. The variation of reduction rate
with hydrogen pressure for powders A and C at temperatures of 470°
C and 500°C, respectively, is shown in Fig. 8.16. The dependence
of reaction rate on hydrogen pressure was calculated to be about 0.7
from least square plots; however, the pressure dependence of the
reaction rate decreased with increasing pressure.
The rate constant approximately fits an equation of the form
} (T,p)=A(f(p)e-º/”, where f(p) is an undetermined function of
pressure and is approximately equal to p", E is the activation energy,
and A is a proportionality constant.
The reaction at the surface, although possibly occurring at a con
stant rate, may be quite complex. It probably involves the adsorption
and possibly the decomposition of hydrogen molecules into atoms on
the solid surface, the reaction of this hydrogen with lattice oxygen, and
the desorption of water from the solid. The following qualitative
OXIDATION AND CORROSION OF URANIUM DIOXIDE 419
-IO.O I | I I I I i I-I-I-I-I-I- T T- T
T2"
,” -
2?
-IO. H.

-IO.2 H. *2^ -
& -IO.3 H. >~~
D ºf
-
E *
.*
-
2^_c
,
* -104 H
- * ,”
** ,
:
~
-10.5 H.
B -
dº -IO.6 H .*
e.
o,” -
:
# -107 H ,”"
.*
2^ z22 -
...’
E
~ -10.8 H
- 2^ -
*
3 -109 H
-1
*
2^
,2 D POWDER A, 470° C
-

,’
-II.O H. (LEAST SQUARES SLOPE, O.65)T
.* 2Eſ 2^
- I.I. H. 2^ O Powder C, 500° C
-
- 1.2 H..* *.* -6O (LEAST SQUARES SLOPE, O.72) -
-i I.3 H–1–1–1–1–1–1–1–1–1–1–1–1–1–1–1–
I.O. 1.2 1.4 1.6 1.8 2.O 2.2 2.4 2.6 2.8 3.O.
LOG PH2, mm Hg

FIGURE 8.16. Rate Constant as a Function of Hydrogen Pressure [49]. (Re


printed with permission from S. Aronson and J. C. Clayton, “Kinetics of the
Reduction of U.O., in Hydrogen,” Pergamon Press, Inc., 1958.)

reactionscheme for the reduction process was given by Aronson and


Clayton [49]:

H.(gas)#2H(a)

HG)+O(1) + o-H'
O-H-I-H(a) tº H.O(a)
H.O(a)*H.O(gas).
The symbol (a) represents the adsorbed state, O(1) represents lat
of the composition UO, at the particle sur
tice oxygen ions in excess
face, and O-Hº represents an oxygen-hydrogen complex of unknown
stability.
The assumption that hydrogen decomposes into atoms on the sur
face of the solid was suggested by the facts that the dependence of
the reduction rate on pressure was a fractional power less than 1 and
that the dependence approached 0.5 at the higher pressures. This
fractional dependence could be accounted for by assuming hydrogen
decomposition. The experimental activation energy would probably
be a composite activation energy involving more than one reaction
step. A different explanation for a similar pressure dependence was
offered by Tanford, et al., in their study of the reduction of U.O, and
CO, (see Sect. 84.4) [50].

57.4789 0–61–28
420 URANIUM DIOXIDE:
575°C 540°C
PROPERTIES AND NUCLEAR APPLICATIONS
-
T

z
9
*
o
<r
tal
º:
ul
o
3
*
tal
&
-: *~~
o
O
-
#
F.
O
>
<r
{

O #5 40
l
60
TIME
I
8O
IN MINUTEs
I
IOO
—1–
I2O l
FIGURE 8.17. Hydrogen Reduction of U.O. [52].

8.4.4 Reduction of Us Os to UO2 in Hydrogen

have studied the reduction of UO, and U.O. to


Many investigators
UO, [22a, 50–57]. In general, the reduction of UOs to UO, is
more complicated and less reproducible than the reduction of U.O.
because of the more variable nature of the UO, preparations and be
cause of the greater structural changes that occur in the reduction of
UOs. Detailed discussion is given of several kinetic studies of the
reduction of UAOs to UO. Only occasional reference is made to the
reduction of UOs.
Kuhlman studied the reduction of UAOs to UO, in hydrogen at
atmospheric pressure at temperatures of 400° to 600° C [52]. The
kinetics were almost identical with those found for the reduction of
the UO, parent material [51]. The UAOs was prepared by heating
the UO, powder in air for an hour at 700° C. The rate curves ob
tained by Kuhlman (shown in Fig. 8.17) are in good agreement with a
first order rate law. The activation energy for the reaction was cal
culated to be 34.2+2.3 kcal/mole from the temperature variation of
the first order rate constant. The activation energy for the UOs reduc
tion was 35.1+0.6 kcal/mole (see Chap. 2). Kuhlman studied the
dependence of the rate of reduction of UO, on hydrogen pressure by
reducing the oxide at atmospheric pressure in mixtures of hydrogen
TIES AND NUCLEAR APPLICATION: OXIDATION AND CORROSION OF URANIUM DIOXIDE 421
i I I T i I T I I i I
[] soo-c
2
% soo’c*
547 h-----—–––––––––––– — UO2
— — ————— —— — — — — ——

| *? -
- 546-
ºº-
#
°5.45}- 2. Z
º
% -
ºº EZ.
*<t
*544––––––––––––– —— — — — — — — — — — — — — — —--— — ——
U409

2.00
—l 1–
2.IO
—l 1
22O
I l
2.3O
I l
24O
I- _l
2.50
I l
2.60
O/U ATOM IC RATIO

FIGURE 8.18. Lattice Constants of Products Formed during Reduction of UsOs


in Hydrogen [22a].

and nitrogen [53]. The rate was directly proportional to hydrogen


pressure. The variation of the first order rate constant with hydro
gen partial pressure is shown in Table 8.12.

TABLE 8.12—VARIATION OF UOs REDUCTION RATE WITH HYDROGEN


PARTIAL PRESSURE [53]
Temperature(° C) Pressure (atmospheres) Rate constant(min-1)
450 0. 20 0.0053
0. 50 0. 0.139
0. 75 0.0231
500 0. 10 0.0193
0. 20 0. 0315
0. 50 0.0770

Although Kuhlman's data empirically obey a first order rate law,


it does not appear that a simple first order mechanism can be postu
lated for the reduction of UAOs or UOs. X-ray data obtained by
Aronson and Roof showed that reduction of UAOs occurs by the forma
tion of a new phase rather than by a decrease in the oxygen content of
the U.O.s phase [22a). Aronson and Roof obtained X-ray patterns
of U.O. samples partially reduced in hydrogen at 500° and 600° C.
The UAOs was prepared from MCW UO, by oxidation in air at 800° C.
The X-ray pattern showed a mixture of the orthorhombic UAOs phase
and a cubic phase to be present during most of the reduction. The
relative amount of UAO, decreased as the reduction proceeded. Lattice
parameter measurements on the reduction product (Fig. 8.18) showed
that there was a change in the lattice parameter of the product phase
422 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

with extent of reduction. The lattice parameter varied from a value


close to that for U.O., in the first stages of the reaction to a value
corresponding to the lattice of UO, at the end of the reaction. It was
inferred from the data that the reduction of UAOs occurs in two stages:
UAOs is first reduced to U.O., which is then reduced to UO,. It was
not clear to the investigators why a gradual transition of lattice pa
rameter with average composition occurred in the middle portion of
the reaction nor why a mixture of U, O, and UO, such as occurs on
reduction of U,O, was not observed.
Tanford, Tichenor, and Larson studied the reduction of UsOs in
hydrogen at temperatures of 450° to 850° C [50]. The UAOs powder
was prepared from UO, by calcination at 800° C. Since the reaction
rates were followed by pressure changes in the system, the hydrogen
pressures decreased during the course of the reductions. This makes
it somewhat difficult to interpret the rate curves. However, the de
pendence of the rate on the extent of reduction appears to be relatively
small. The authors inferred from this, as well as from the nature of
the pressure dependence, that the reaction was surface controlled.
Table 8.13 shows the dependence of the initial UAOs reaction rate on
the hydrogen pressure at 520° C. Although the rate increased with
increasing pressure, the pressure dependency became smaller in a
manner similar to that occurring in surface adsorption, e.g., the
adsorption of hydrogen on copper powder at 250° C [58a].

TABLE 8.13–DEPENDENCE OF U3Os REDUCTION RATE


ON HYDROGEN PRESSURE [50]
Rate
Pressure (mm Hg) (Moles H./moles U-min)
32 0.0029
110 0.0075
225 0.0102
330
0 ().

0131
430 136
0

580 0.0160

Figure the reaction rate with tempera


of

variation
8.19 shows the
hydrogen pressure, 400 mm Hg.
of

calculation
at

ture constant
A

the
experimental activation energy was made from the steepest portion
the log the rate against reciprocal temperature.
of

of

of

the curve
This yielded
of

the value 30.7 kcal/mole.


equilibrium volume hydrogen was
at an

of

The authors assumed that


adsorbed on the oxide surface all times and that the rate of reaction
was proportional
as

hydrogen adsorbed, Vaas,


of
to

the volume well


Ae-**
which represents the rate deter
of
to
as

factor the form


a

mining step the solid. The rate equation took the form
in

rate=Vaa, Ae-P/RT.
by

with temperature the adsorption hydrogen


of

of

The variation
OXIDATION AND CORROSION OF URANIUM DIOXIDE 423

•l -T-I-T-I-H -
O.048 -
O.O40 - -
i
ir OO-52H-
ºn
-
uu
—l - -
-
O
-:
uj 0.024H-
&
ar |- -
5
# oolsH
o
-
«
u |- -

OOO8H —

H–1–1–1–1–1–1–
4OO 500 6OO 700 8OO
TEMPERATURE, *c

FIGURE 8.19. Effect of Temperature on U.O, Reduction [50].

uranium oxide was assumed to be similar to the adsorption of


hydrogen on a MnO-Cr2O, catalytic surface [58b]. Tanford, et al.,
showed that under special conditions it is possible to obtain a curve
of rate versus temperature similar to the experimental curve in Fig.
8.19 by properly combining the curves obtained on plotting Vad, and
Ae-** separately as functions of temperature.
DeHollander studied the reduction of UAOs to UO, in hydrogen
using activated UAO, powder obtained by repeated cycles of reduction
to UO2 in hydrogen and oxidation to UAOs in air [55]. In this man
ner he was able to obtain U.O, powders which followed the same re
duction kinetics upon reoxidation and reduction again.
An important reason for the variability in the reduction kinetics
of UO, and UAOs in hydrogen is the surface area change which may
occur on reduction. DeHollander measured the changes in surface
area occurring on reduction of UO. The variation of area with ex
tent of reduction is shown in Table 8.14 for reduction at 500° C. It is
seen that the area increased by a factor close to 10 during the course of
is,

The breakup of the oxide particles least, partially


at

reduction.
changes the crystal structure
of
to

the material.
in

due
An kinetic spectrum for reduction
of

idealized activated UAOs


is

Fig. 8.20. Except for small perturbation


at

shown the start and


in

short tail, the reaction rate approximately con


of

the existence
is
a
424 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

PHASE CHANGES

WEIGHT--

SURFACE
CONSTANT

<-400 SECONDS—º
TIME

FIGURE 8.20. Kinetic Spectrum of Active U.O, at 513° C [55].

TABLE 8.14—SURFACE AREA OF UO, AS A FUNCTION OF EXTENT


OF REDUCTION (55)
Estimatedpercentreduction Arerage area (m2/g)
0.68
1–2 1. 33
15–25 1. 84
25–40 2. 08
40–60 3. 60
100 5. 48

stant. By varying the flow rate of hydrogen mixed with nitrogen, the
reaction was found to be first order with regard to hydrogen pressure.
DeHollander has explained the nature of the reduction curve in
the following manner. The slow initial rate of weight loss is due to
the time required for hydrogen to displace helium in the reaction
chamber and sample. The brief period of rapid loss which then
occurs probably results from the ease with which oxygen can be
extracted from the UAOs lattice. Then follows a phase change to a
pre-UO, lattice. During the final adjustment to the pre-UO, struc
ture, the rate slows as though the oxygen were more tightly bound.
Once the pre-UO, lattice is attained, reaction proceeds at a constant
rate. The decrease in rate towards the end of the reaction probably
results from the distribution of particle sizes in the powder.
Zero order rate constants were obtained from the slopes of the linear
portions of the rate curves. An activation energy of 27.7+0.3
kcal/mole was calculated from a plot of log rate constant versus re
ciprocal temperature.
OXIDATION AND CORROSION OF URANIUM DIOXIDE 425

On the basis of the linear dependence of rate on hydrogen pressure,


DeHollander assumed that the rate-controlling step was the adsorp
tion of hydrogen molecules on the oxide surface. By comparing the
experimental reaction rates with a number of theoretical adsorption
models, he chose the following reaction as the rate determining step in
the reduction:
H
-
t
Hz2 (gas)
gas +S-O == Yo S

H^

where the notation S—O refers to an oxygen atom chemisorbed on


auranium atom on the surface and

t
HQ

represents an activated complex from which water is subsequently


desorbed. The rate equation derived for this model is
- *
#–7.65× 108
º: (—º exp

where F is the fraction remaining to be reduced at time t (seconds),


P is the hydrogen partial pressure in atmospheres, and A* is the
specific surface area in m”/g which is effective in the reduction.
The value of A * may depend on the nature of the crystal faces ex
posed, the extent of exposure of the surfaces to hydrogen, and the
relative number and size of particles which have already been com
pletely converted to UO2.
DeMarco and Mendel and Notz and Mendel have recently studied
the reduction of uranium trioxide in hydrogen at partial pressures of
14 to 1 atmosphere [56, 57]. I)eMarco and Mendel investigated
the reduction of high surface area UO, (26.3 mº/g) at temperatures
of 300° C to 400° C. The reduction appeared to be a two-step
process corresponding to the UOa-H Hº->UO, as
reactions and
UO, Hº->UO2.
ºs-H The rate curves for both steps were linear, indi
cating that the reaction rates are surface controlled. Activation
energies of 26.6 and 39.1 kcal/mole were calculated for the reduction
of UO, and UO, Exponential hydrogen pressure
so,

respectively.
dependencies for the UO, and UO,
of

0.62 and 0.43 were determined


so

reductions.
UO, samples having sur
of

Notz and Mendel studied the reduction


the range mº/g temperatures
of

of
in

to

to

face areas
at

450°
2

5
426 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONs

550°C. They interpreted the data in terms of three consecutive reac


tions: UOs—-UAOs.,U3Os, >UAOs and U.Os —-UO2. U-Os, and UAOs
refer to the upper and lower limits of the UAO, phase. X-ray pat
terns of a number of partially reduced samples showed only the
presence of the UOs, U.O.s, and UO. phases. A U.O., intermediate
phase was not observed. The rate of reduction of UO, was found
to be proportional to the surface area of the UO, powder. Activa
tion energies of 25.2 and 30.6 kcal/mole were calculated for the
reduction of UO, and UAOs , respectively. An exponential pres
sure dependence on hydrogen of about 0.8 was determined for the
two reaction steps.
The studies reviewed above indicate that the reduction of U.O.,
and UAOs to UO, is probably controlled by the rate of reaction at the
surface of the oxide particles. It is difficult, however, to correlate
the results of the various studies. There are no obvious relationships
between the rates and extent of adsorption of hydrogen on UO, and
the rates of reduction of U,O,
and UAOs. One might infer from the
similar values of the activation energies obtained in the reduction of
U.O, (25+3 kcal/mole) and UAOs (27.7, 30.7, 30.6, and 39 kcal/
mole) that the reduction processes may be similar [49, 52, 55, 56].
The pressure dependence of the reduction of U.O., was found to be
complex and similar to the pressure dependence obtained by Tanford,
et al., for the reduction of UAOs. The results of Kuhlman and DeHol
lander indicate that the reduction rate of Us Os is linearly dependent
on the hydrogen pressure. Notz and Mendel find an 0.8 dependence
on hydrogen pressure for the reduction of UAOs. DeMarco and
Mendel observed a 0.43 dependence on hydrogen pressure for the
reduction of UO, From their pressure dependence, Aronson and
sc.

Clayton infer that hydrogen dissociates into atoms when adsorbed


DeHollander assumes molecular adsorption
on

on

the surface. the


pressure appears that considerable kinetic and
of

It

basis his data.


structural work will required before the basic reduction mecha
be

nisms are clarified.

REFERENCES

Roberts, Chemisorption Oxygen on Ura


V.
of

of
J.
E.

“Oxides Uranium.
L.
1.

on

nium Dioxide and Uranium Dioxide-Thorium Dioxide Solid Solutions."


Chem. Soc., 3332–3339 (1954).
J.

Roberts, “The Surface Chemistry


of
E.

D. M. McCon NELL and


J.

L.

J.
2.

Uranium and Thorium Oxides” “Chemisorption: Proceedings of


in

Symposium held at the University College North Staffordshire,


of

Keele.
Staffordshire, by the Chemical Society,” W. Garner, ed., Academic
E.

Press, New York, 1957.


OXIDATION AND CORROSION OF URANIUM DIOXIDE 427

3. I. F. FERGUson and J. D. M. McCoNNELL, “Heat of Adsorption of Oxygen


on Uranium Dioxide at –183°,” Proc. Roy. Soc. A241, 67–79 (1957).
4. J. D. M. McCoNNELL, “Heats of Adsorption of Oxygen on Solid Solutions of
Uranium Dioxide in Thorium Dioxide at –183°,” J. Chem. Soc., 947–950
(1958).
5. S. BRUNAUER, P. H. EMMETT, and E. TELLER, “Adsorption of Gases in Multi
molecular Layers,” J.
Am. Chem. Soc. 60, 309–319 (1938).
6. J. S. ANDERSON, L. E. J. Roberts, and E. A. HARPER, “The Oxides of Ura
nium. VII. The Oxidation of Uranium Dioxide,” J. Chem. Soc., 3946–3959
(1955).
7. N. CABRERA and N. F. Mott, “Theory of Oxidation of Metals,” Repts. on
Progress in Physics 12, 163–184 (1949).
8. P. Jolibors, “A New Oxide of Uranium, U.O.,” Compt. rend. 224, 1395–1396
(1947).
9. F. GRoNvold and H. HARALDSEN, “Oxidation of Uranium Dioxide,” Nature
162, 69–70 (1948).
10. K. B. ALBERMAN and J. S. ANDERSON, “The Oxides of Uranium,” J. Chem.
Soc., S303–S311 (1949).
11. P. PERIo, “The Oxidation of Uranic Oxide at Low Temperatures,” Bull. soc.
chim. France, 256–263 (1953).
12. S. AroNsoN, R. B. Roof, Jr., and J. BELLE, “Kinetic Study of the Oxidation of
Uranium Dioxide,” J. Chem. Phys. 27, 137–144 (1957).
13. (a) P. E. BLACKBURN, J. WEISSBART, and E. A. GULBRANSEN, “Oxidation of
Uranium Dioxide,” J. Phys. Chem. 62, 902–908 (1958).
(b) P. E. BLACKBURN, J. WEISSBART, and E. A. GULBRANSEN, “Oxidation of
Uranium Dioxide,” Westinghouse Research Report 100 FF 942–R2, July
10, 1956.

14. J. S. ANDERson, “Recent Work on the Chemistry of Uranium Oxides,” Bull.


soc. chim. France, 781–788 (1953).
15. J. S. ANDERson, “The Oxidation of Particles of Uranium Dioxide” in “Aus
tralian Atomic Energy Symposium, 1958,” pp. 588–599, Melbourne Univer
sity Press, Australia, 1958.
16. P. PERIO, “Observations on Uranium Oxides Formed between UO2 and U.O.s,"
Bull. soc. chim. France, 840–841 (1953).
17. H. HERING and P. PERIo, “The Equilibrium of Uranium Oxides between UO2
and U.Os,” Bull. 80c. chim. France, 351–357 (1952).
18. D. A. WAUGHAN, J. R. BRIDGE, and C. M. SchwARTz, “Comparison of Active
and Inactive Uranium Dioxide-Oxygen Systems,” BMI-1241, Dec. 10, 1957.
19. H. R. HoEKSTRA and S. SIEGEL., “Recent Developments in the Chemistry of
the Uranium-Oxygen System” in “Proceedings of the International Con
ference on the Peaceful Uses of Atomic Energy, Geneva, 1955,” Vol. 7, pp.
United Nations, New York, 1956.
394–400,

20. J. S. ANDERSON, D. N. EDGINGTON, L. E. J. Roberts, and E. WAIT, “Oxides of


Uranium. IV. System UOz–Thor-O,” J. Chem. Soc., 3324–3331 (1954).
21. (a) A. B. AUSKERN and J. BELLE, “Self-Diffusion of Oxygen in Uranium Di
oxide,” J.
Chem. Phys. 28, 171–172 (1958).
(b) J. BELLE and A. B. AUskERN, “Oxygen Ion Self-Diffusion in UO." in
“Kinetics of High-Temperature Processes,” W. D. Kingery, ed., pp. 44–49,
John Wiley & Sons, Inc., New York, 1959.
428 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

(c) A. B. AUSKERN and J. BELLE, “Oxygen Ion Self-Diffusion in UOs,” sub


mitted for publication to Journal of Nuclear Materials.
22. (a) J.
BELLE and B. LUsTMAN, “Properties of UOs,” “Fuel Elements Confer
Paris,” TID–7546, pp. 442–515, Mar. 1958.
ence,
(b) A. B. AU's KERN and J. BELLE, “Uranium Ion Self-Diffusion in UOs,” sub
mitted for publication to Journal of Nuclear Materials.
23. A. ARROTT and J. E. GoLDMAN, “Magnetic Analysis of the Uranium-Oxygen
System,” Phys. Rev. 108,948–953 (1957).
24. G. VALENSI, “Kinetics of Oxidation of Metallic Spherules and Powders,”
Compt. rend. 202, 309–312 (1936).
25. H. S. CARSLAw and J. C. JAEGER, “Conduction of Heat in Solids,” p. 200, Ox
ford University Press, London, 1947.
26. F. GRoNvold, “High Temperature X-ray Study of Uranium Oxides in the
UOz-U20s Region,” J. Inorg. and Nuclear Chem. 1, 357–370 (1955).
27. P. PERIo, “Contribution to the Crystallography of the Uranium-Oxygen Sys
tem,” Doctoral Dissertation, University of Paris, 1955; CEA-363.
28. B. E. SchANER, “Metallographic Determination of the UOz—U.O., Phase Dia
gram,” J. Nuclear Materials 2, 110–120 (1960).
29. J. J. KATz and E. RABINow ITCH, “The Chemistry of Uranium” in “National
Nuclear Energy Series, Div. VIII,” Vol. 5, McGraw-Hill Book Co., New
York, 1951.

30. S. ARONsoN, “Oxidation of UO, in Water Containing Oxygen,” Bettis Tech


nical Review, WAPD—BT—10, Oct. 1958, p. 93.

31. W. A. Johnson and R. F. MEHL, “Reaction Kinetics and Processes of Nuclea


tion and Growth,” Trans. AIME 135, 416–442 (1939).
32. R. E. DEMARCO, H. A. HELLER, R. C. ABBOTT, and W. BurkHARDt, “Oxidation
of UO, to UAOs,” Am. Cer. Soc. Bull. 38, 360–362 (1959).
33. D. VIER, “The Thermal Stability of Uranium Oxides and Uranium Oxide
Hydrates in Water,” A–1277, May 26, 1944.
34. C. W. KUHLMAN, “Treatment of Uranium Dioxide with Water Vapor at High
Temperatures,” MCW—103, Aug. 13, 1948.
35. S. M. LANG, F. P. KNUDseN, C. L. FILLMoRE, and R. S. Roth, “High
Temperature Reactions of Uranium Dioxide with Various Metal Oxides,”
Nat. Bu. Standards Circ. 568, 32 pp. (1956).
36. J. K. DAwson, E. WAIT, K. ALcock, and D. R. CHILtoN, “The System
Uranium Trioxide-Water,” J. Chem. Soc., 3531-3540 (1956).
37. T. J. BURRE, J. GLATTER, H. R. Hoge, and B. E. SchANER, “Fabrication of
High Density Uranium Dioxide Fuel Components for the First Pressurized
Water Reactor Core" in “Nuclear Metallurgy,” vol. 4, pp. 135–143, AIME,
New York, 1957.
38. J. BELLE and L. J. Jon Es, eds., “Résumé of Uranium Oxide Data-X,” WAPD
TM-73, Aug. 1957, p. 15.

39. J. GLATTER, Letter of Transmittal for Report WAPD—NCE–3242, dated


Mar. 19, 1957.
40. (a) Z. M. SHAPIRo, ed., “Résumé of Uranium Oxide—II,” WAPD—PMM-167,
June 1955, p. 8.
(b) Z. M. SHAPIRo, ed., “Résumé of Uranium Oxide Data—III,” WAPD—PMM
197, Sept. 1955, pp. 6, 12.
OXIDATION AND CORROSION OF URANIUM DIOXIDE 429

(c) Z. M. SHAPIRo, ed., “Résumé of Uranium Oxide Data–IV,” WAPD—PWR—


PMM-417, Dec. 1955, p. 6.
(d) J. BELLE, ed., “Résumé of Uranium Oxide Data—V,” WAPD—PWR–PMM
429 (Del.), Mar. 1956, p. 6.
(e) J. BELLE and L. J. Jon Es, eds., “Résumé of Uranium Oxide Data—VI,”
WAPD—PWR–PMM-466 (Del.), June 1956, p. 11.
41. (a) D. J. Bowers, A. G. ALLIsox, and W. H. DUckworth, “Ceramic Studies”
in “Résumé of Uranium Oxide Data—VI,” J. Belle and L. J. Jones, eds.,
WAPD—PWR-PMM-466 (Del.), June 1956.
(b) D. J. Bowers, A. G. ALLIsoN, and W. H. DUckworth, “Ceramic Studies”
in “Résumé of Uranium Oxide Data—VII,” J. Belle and L. J. Jones, eds.,
WAPD—PWR-PMM-491, Sept. 1956.
(c) D. J. BowFRs, W. A. HEDDEN, M. J. SNYDER, and W. H. DUckworth,
“Effect of Ceramic or Metal Additives in High—UO, Bodies,” BMI-1117,
July 24, 1956.

2. A. P. “Studies Concerned with Uranium


BEARD, Dioxide at the Knolls
Atomic Power Laboratory" in “Résumé of Uranium Oxide Data-XI,”
WAPD–TM-101, Jan. 1958.
. L. A. WALDMAN and J. M. LojFK, “Erosion of UO, Fuel Material,” WAPD
PWR-CP—1814, Feb. 15, 1956.
. J. M. Loj Ek, W. T. LINDSAY, Jr., and P. Cohen, “Corrosion and Erosion of
Sintered UO, Compacts in High Temperature Water,” Bettis Technical
Review, WAPD—BT-7, Mar. 1958, p. 16.
45. J. M. Lojek and W. T. LINDs.AY, Jr., “The Effect of Oxygenated Water on
Clad-and-Defected UO, Fuel Specimens,” Bettis Technical Review,
WAPD—BT-7, Mar. 1958, p. 33.
. J. D. Eichenberg, P. W. FRANK. T. J. Kisiel, B. LustMAN, and K. H.
Vogel, “Effects of Irradiation on Bulk Uranium Dioxide” in “Fuel Ele
ments Conference, Paris,” pp. 616–716, Mar. 1958.
. L. A. WALDMAN and W. T. LINDs.AY, Jr., “Out-of-Pile Dynamic Loop Tests
of Irradiated Fuel Materials,” Bettis Technical Review, WAPD—BT-7,
Mar. 1958, p. 1.
. L. E. J. Roberts, “The Oxides of Uranium. Part VI. The Chemisorption of
Reducing Gases on Uranium and Thorium Dioxides,” J. Chem. Soc.,
3939–3946 (1955).
S. ARONSON and J. C. CLAYTON, “Kinetics of the Reduction of U.Oa in
Hydrogen,” J. Inorg. and Nuclear Chem. 7, 384–391 (1958).
. C. TANFORD, R. L. TICHENOR, and C. E. LARso N, “Tuballoy Oxides in The
Reduction of TOs,” AECD–3961, July 16, 1945.
. C. W. KUHLMAN, Jr., “Reduction of Uranium Trioxide to Uranium Dioxide
with Hydrogen-Reaction Rates at Various Temperatures,” MCW-142,
Oct. 1948.
52. C. W. KUHLMAN, Jr., “Reducation of UAO, with Hydrogen-Reaction Rate,”
MCW-210, Aug. 19, 1949.
53. C. W. KUHLMAN, Jr., “Reduction of Uranium Trioxide with Hydrogen
Nitrogen Mixtures,” MCW-215, Sept. 2, 1949.
54. R. H. MooRE, “Factors Affecting the Reactivity of Uranium Trioxide,”
HW-31670, Apr. 29, 1954.
430 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

55. W. R. DEHoLLANDER, “A Kinetic Study of the Reduction of Uranium Oxides


with Hydrogen,” HW–46685, Nov. 8, 1956.
56. R. E. DEMARco and M. G. MENDEL, “The Reduction of High Surface Area
Uranium Trioxide,” J. Phys. Chem. 64, 132–133 (1960).
57. K. J. Notz and M. G. MENDEL, “X-ray and Kinetic Study of the Hydrogen
Reduction of y—UOs,” J. Inorg. and Nuclear Chem. 14, 55–64 (1960).
58. (a) S. BERKMAN, J. C. MoRRELL, and G. EGLOFF, “Catalysis,” p. 74, Reinhold
Publishing Corp., New York, 1940.
(b) Ibid., p. 78.
Chapter 9

IRRADIATION EFFECTS IN URANIUM DIOXIDE


B. LUSTMAN

9.1 INTRODUCTION

The widespread acceptance of UO, as a reactor fuel material during


the past five years has sparked an extensive investigation of its prop
erties. In particular, the factors affecting application to its
fuel ele

in
ments and its performance when subjected reactor irradiation have
Whereas the incentive for use of UO,

as
received detailed attention.

a
reactor fuel arose from its chemical stability high temperature
to in

water, and the major applications present are water-cooled reac


at

its irradiation stability and ability sur


its
tors, demonstration
of

of

to
vive high fissioning depletions elevated temperatures has extended
at
in its

the range utility even reactors utilizing inert gas liquid or


of

to

chemical stability
to its

minor importance.
of

metal coolants which


is

this chapter correlate critically the extensive literature


of

The aim
is

on irradiation effects UO, and isolate and define those factors


to
in
of its

as

which control performance fuel.


a

The theory radiation damage ionically bonded solids first


to

is

reviewed. Experimental data which relate both structural and prop


erty changes metal oxides under both fast neutron and alpha par
in

ticle bombardment are then compared with theoretical predictions,


by

and the damage suffered sensitively affected


be
to

shown the
is

In

composition and structure the subsequent section,


of

the oxide.
by

damage fission fragment bombardment


to

discussed and shown


is

more severe than that predicted from comparison with fast neutron
or be

alpha decay exposure. The ability UO, ex


of

to

withstand such
posure without extensive and permanent property de
or

structural
unique and favorable behavior ma
be

of

terioration found this


of to
is

fission products from irradiated


of

terial. The kinetics the release


UO, then discussed; the rate volatile products con
of

of

release
is

is
by

trolled primarily
of

diffusional mechanism and the effects oxide


a

on

of

structure and physical characteristics are correlated the basis


fission products both within fuel
of

such mechanism. The release


a

431
432 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

element claddings and, through cladding defects, into the fuel element
environment is described. The ability of UO, to contain fission prod
ucts within its lattice is shown to be high, but limited, and the manner
of breakdown when the containment limits are exceeded is described.
Some preliminary experiments on the chemical behavior of bulk UO.
in aggressive environments as modified by irradiation simultaneous
with the exposure are then discussed. In a subsequent section, the
factors affecting the ability to transfer heat generated in UO, fuel to
its metallic cladding are discussed; the thermal performance of UO.
elements is shown to be a sensitive function of the gap between fuel
and cladding. In a final section, the application of the preceding
data to the design of oxide fuel elements is described. Three principal
types of oxide fuel elements have been utilized: bulk oxide fuel ele
ments, pin-type fuel elements, and dispersion fuel elements. The
fields of application of each of these types are described.

9.2 RADIATION DAMAGE IN REFRACTORY OXIDES

As discussed in Chap. 6, UO, is most properly considered to be an


ionically bonded solid. The observations and theories of radiation
damage to ionic solids have been the subject of a number of reviews
[1–4]. In general, the same theory of damage has been applied to
such solids as to metallic materials with the following distinctions:
1. Atomic displacements by ionization processes appear capable of
yielding a large fraction of the total displacements suffered by
ionic materials during charged particle bombardment, and many
of the remanent physical effects of such exposure arise from ion
ization.
2. Displacement of metal or oxygen ions into wrong lattice sites in
the case of the ordered oxide lattices can produce lattice strains and,
hence, damage which is unimportant in the case of single compo
nent materials or random solid solutions.
3. Because of the refractory nature of most metal oxides, the dis
placements produced are more permanent and less easily removed
by recovery and annealing processes than those in less refractory
materials.

9.2.1 Theory of Displacement Reactions

For the several bombarding particles of interest in this chapter,


fission fragments, fast neutrons, helium atoms from alpha-decay, etc.,
it is possible to estimate the number of displacements produced per
incident particle. Such estimates are listed in Table 9.1 for the several
oxides of interest here following the treatment outlined by Kinchin
and Pease [1].
IRRADIATION EFFECTS IN URANIUM DIOXIDE 433

TABLE 9.1—ATOMIC DISPLACEMENTS PRODUCED BY BOMBARDMENT IN


VARIOUS OXII) ES

Bombarding particle

Natural decay of U2′ nucleus || Integrated single U235nucleus


neutron
flux of
|_ Fission of
Material Displacements produced by 101*,2 Mew Displacements produced by
single decay neutrons fission fragment
per cm3

Atomic Atomic
Fraction No.—40 No.—55
a-Particle | U234recoil || Total of atoms atomic atomic Total
E=4 Mev | nucleus displaced mass—95 mass–137
energy— energy—
92 Mew 63 Mew

33 1370 1403 0.009 57,000 111,500 168,500


35 1370 1405 0.017 42,000 72,000 114,000
32 1370 1402 0.0085 88,000 160,000 248,000
24 1370 1394 0.004 58,500 119,000 177,500
35 1370 1405 0.011 18,500 42,000 60,500
39 1370 1409 0.013 I------------|------------|----------
35 1370 1405l------------

This treatment estimates the number of atoms or primary knockons


displaced by direct collisions with the bombarding radiation as well
as the number of displacements subsequently produced by these pri

maries. A moving particle of mass M1 and atomic number Z. pro


duces hard sphere collisions with stationary atoms of mass and atomic
number Ma and Z2, respectively, at energies of the bombarding par
-
ticle up to the value La given by

La _2EEZ.Z.(Zº-HZ.”)*(M-FM)
M, Eq. (9.1)

where ER is the Rydberg energy (13.6 ev). The maximum energy,


Emas, which can be imparted to stationary atoms by moving particles
of energy E by hard sphere collisions is
4MM,
Pºw-tº E. Eq. (9.2)

Since in such collisions displacement energies of zero to Ema, are


ºually probable, the average energy of the displaced atoms, E, is
*rnax
T2- and the number of displacements, Na, is

_{*mas Eq. (9.3)


Na T4E,
its

where Ea is the energy required to displace an atom from lattice


generally equal for
25
be

ev
to

to

site and assumed both anions


is

and cations.
434 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

For energies greater than L4, the bombarding particles can pene
trate the electron screening about the nuclei and a minimum energy
Emin can be transferred, which is given by

E.-ºr-º-º:
_4E, ZAZ.”(Z” +Z,”)M. Eq. (9.4)

The cross-section on for such collisions is

tra:
Eq.
***ZEATZ. (9.5)

where ao is the Bohr radius and traž, -8.8×10−" cm”. The mean
energy E of the primary knockons is then

- Emax
Eq.
E=Emin ln
(#.) (9.6)

and the number of displacements per primary knockon is again

E
N=#.
The above expressions hold to a value of Emin up to an energy Le
given by

-
L.-è-º-º-º-º:
_4E. Z.” Z.”(Z2/34–2,218). M,
Eq. (9.7)

At incident particle energies greater than LR, the cross section for
primary collisions is approximately

_4M, Z2 Z.” E.” tra;

the mean energy of the primaries is

F=E. ln (º) d
Eq. (9.9)

and the number of displacements per primary is

Emax
IRRADIATION EFFECTS IN URANIUM DIOXIDE 435

Still a third energy must be defined, Lc, at values of E above which

by
all
it is assumed that energy lost
electronic excitation and none

is
below. Dienes and Vineyard indicate that this energy approxi

is
mately equal weight kev [3]. Thus,

of

in

in
to
the atomic the atom
energies primary Eqs. 9.6 and

of
which the
in

the knockons

in
E
cases
Lc, estimating

of Lc
of

in
9.9 exceed the indicated value substituted

is
Ma. For fast neutron bombardment light elements, this substi

by
tution results the replacement the following expression for
in

Eq. 9.3:
2–4;

).
Emax Eq. (9.11)
Nº- 4E,

Using these relations,


the atomic displacements produced various

in
U*
by

alpha decay nuclei, fast neutron bombardment, and


of

oxides
Table 9.1. The various
in
U*** fission have been estimated and are listed
types differ widely their efficiency
of

irradiation are observed


to

in
for the production displacements, but, for the same type
of

atomic
bombarding particle, only minor differences among the various
of

is,

be
therefore, clear that the information

to
oxides are revealed.
It

discussed later which shows marked differences the structural and


in

property changes observed such oxides upon irradiation cannot


in

explained
on

such displacement reactions alone but


be

of

the basis
rather requires also detailed consideration the specific
of

material
properties.

Radiation Damage
of

(a) Saturation

The considerations outlined above predict, general, many more


in

displacements than actually are observed. The concentration dis


of
by
be

placements for given bombardment flux can decreased recovery


a

processes, which interstitially displaced atoms recombine with


in

by

thermally activated diffusion. the irra


of

lattice vacancies Most


diations performed the refractory metal oxides have been conducted
on

far below the melting point and well below the Tammann temperature
half the melting point. Furthermore, irradiated oxides showing
of

extensive damage are resistant annealing such damage quite


of
to

at

in is,

elevated temperatures, far above the irradiation temperature.


It

therefore, probable that recovery processes play but minor role


a

refractory relatively
of

the irradiation low temperatures


at

the oxides
and that observations of differences behavior between various oxides
in in

their recovery rates.


be

cannot Kinchin
to

ascribed differences

57.4789 O—61–29
436 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

and Pease indicate the existence of a geometrical limit to the number


of displacements a crystalline body may sustain, thus, leading to a
saturation of damage [1]. Saturation is postulated to occur at a
maximum concentration of interstitially displaced atoms such that
each interstitial atom can be surrounded by occupied lattice sites.
If Q is the number of lattice sites surrounding an interstitial position,
- - - --- - - In "
the saturation concentration of interstitial atoms is given by
Q
this concentration is attained when the fraction of atoms displaced to
interstitial sites is approximately . . As a result of diffusional proc

esses, combination of vacancies and interstitials to form multiple


defects, etc., Q is expected to be temperature dependent; but in any
case the maximum interstitial content is estimated by Kinchin and
Pease to be less than about five percent in crystalline materials.

(b) Replacement Collisions

While these considerations serve to mitigate the severity of radiation


damage, the particular structural mechanism whereby the displace
ment reactions occur can, particularly in ordered compounds such as
the refractory oxides, intensify the damage. One such mechanism,
which is of consequence only in ordered compounds, is termed re
placement collisions [1]. These occur when a moving atom is left,
after a collision with a stationary atom of different mass, with an
energy less than the displacement energy (Ea) and also less than the
energy imparted to the struck atom. From this definition, it can
be inferred from Eq. 9.2 that the ratios of the masses of the
atoms in the compound must be less than about six for replacement
collisions to occur. Thus, for the oxides of interest here, only those
with the metals of the first four periods are likely to suffer extensive
damage by this mechanism. In cases where such replacements are
possible and, in particular, for energy ranges in which hard-sphere
collisions occur, replacement collisions may well outnumber displace
ment collisions. Varley has suggested another mechanism for damage
in ionic solids involving multiple ionization of anions by inelastic
collisions of charged particles with atoms [4]. Anions thus ionized
to the extent of becoming temporarily positively charged can be
repelled by neighboring positively charged cations to interstitial
positions. This effect may, particularly in the case of fission fragment
IRRADIATION EFFECTS IN URANIUM DIOXIDE 437

bombardment, produce as many displacements as those produced by


direct momentum transfer.

(c) Thermal and Displacement Spikes

The effects of radiation with energetic particles can be described


not only in terms of individual displacements but also in terms of the
cumulative effects produced by successive collisions. Thus, Seitz pro
posed the model of thermal spikes in which a moving particle is postu

its
lated to heat the material surrounding track through the solid [5].
Vineyard temperature radially

of
Dienes and describe the variation
particle passage the particle
of

of
from the axis track time after
at
a

t
by

the expression
Q

T= T,4- 4TCd (-p”/4Dt) Eq. (9.12)


1

exp
D7
To

the ambient temperature the material, temperature


of

where

is
T
is

the track,
at

time and radial distance from the axis of thermal

D
is
p
t

diffusivity-#. where thermal conductivity, heat capacity,


K

C'
is

density, and energy released per unit length is

of
and track [3].
is

Q
is
d

Thus, for fission fragment energy 100 Mev traversing uranium


of
a

with recoil length cm, assuming uniform dissipation


of

of
4×10−1
a

energy along the recoil track, seconds after the recoil, temper
10

a
of

angstrom units
10

ature about 1,100° over ambient attained


C

is
C,

from the spike axis and 900° 100 angstrom units from the axis.
For UO, assumed thermal conductivity
an
of

of

0.017 w/cm-*
at C

(0.0041 cal/cm-sec-* C), the corresponding figures are 18,000°


C
10

angstrom units and 470° 100 angstrom units. Thus, the lower
at
of C

the thermal conductivity the medium, the more intense and the
longer the spikes. The effects such spikes are,
of

of

the duration
therefore, likely the refractory oxides than
be

in
to

more marked
in

ex
as

metal lattices and have indeed been invoked,


discussed below,
to

plain SiO. Another


of

mechanism,
of

some the effects noted that


in

of by

the displacement spike, has been postulated Brinkman which dis


in

placed atoms struck by primary knockons kev energies are driven


interstitially into shell material, leaving
of

the base void which


a

collapses rapidly under the pressure the shell [6]. This mechanism
of

which can affect regions 100 angstrom units linear dimension par
in

is

ticularly useful explaining homogenization microscopic concen


of
in

tration gradients and may explaining the rapid loss


be

of

useful
in

crystallographic structure some oxides upon irradiation [7]. Such


in
438 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

a mechanism is particularly significant in the case of oxides of heavy


metals in which a large fraction of the displacement energy is dissi
pated in secondary displacements. Finally, the phenomenon of focus
ing collisions, as discussed by Silsbee, in which displacements are
transmitted along close-packed directions in the crystal allows prop
agation of energy for relatively large distances from the primary site
of damage, particularly in case of atoms with large ion cores [8].
It has been assumed thus far that in order to produce damage the
bombarding particles must be sufficiently energetic directly to displace
atoms or to cause fissioning reactions, thus, producing energetic parti
cles. Foderaro and Burnham have pointed out that recoil energies
from the emission of capture gammas, resulting from the absorption
of neutrons of insufficient energy directly to displace atoms, may be
sufficiently high to cause displacement not only of the capturing atom
but also of neighboring atoms [9]. The energy E of the recoil atom
derived from emission of a photon of energy E, is

Eq. (9.13)

where M is the mass of the recoil atom and c is the velocity of light.
Thus, the capture of a neutron by a U* nucleus results in emission of
about 4.9 Mev gamma energy, imparting to the recoil atom an energy
Such ef
of its

of about 50 ev which is sufficient to cause displacement.


fects are likely significant only exposure
be

of

of
materials
to

in

case
relatively high thermal neutron capture cross section predominantly
to

thermal neutron fluxes. They may, however, explain observation

of
extensive damage heavy metal oxides exposed
to

such environments.
in

Experimental Observations
of

9.2.2 Fast Neutron Damage

Table 9.2 are many the significant data with respect


of

Collected
in

property changes metal oxides upon fast


of

the structural and


to

neutron irradiation exposure. the damage ob


of
It

characteristic
in is

served that, general, larger magnitude than that encountered


in

is
it

of of by

metal alloy specimens.


be

This observation may explained


in

two
factors; first, the sensitivity property change disordering
of

to

the
regular oxide compound structure and, second, the occurrence sev
eral mechanisms for producing such disorder, such replacement
as

collisions which are unimportant for more random structures.


A
by

inspection
of

second characteristic revealed Table 9.2 the marked


is
by

susceptibility damage the various oxides


to

differences exhibited
in

investigated, ranging, for example, from complete loss


of

structure
of

crystalline quartz little apparent effect spinel.


of
in

to

in

case case
IRRADIATION EFFECTS IN URANIUM DIOXIDE 439

TABLE 9.2—EFFECTS OF FAST NEUTRON IRRADIATION ON NONFISSION


ABLE OXII) ES

Changes on irradiation
Material Crystal structure | Fast neutron
dosage-nºvt
Property before Property after Reference

ZrO: Monoclinic to 1.5x101sat Monoclinic Monoclinic


º -
1,100°C 100°C a0= 5.143A a0= 5.123A
Tetragonal 1,100 be- 5204 Å bo- 5.141A 10
to 1,900°C cº- 5.311Å co- 5.208A
Cubic (?) 1,900to 3 =80.75° 8 =81.3°
2,700°C

(Seefurther discus- 6x1019at Monoclinic Cubic


sion in Sect. 9.3) 100°C Cubic reverts to 11, 12
monoclinic on
annealing at 800°C

BaTiO, Tetragonal to 2x1019 Tetragonal Tetragonal


120°C, cubic as-4.0349 A as-4.001 A
to melting- co-3.9023 Å cº-4,037 Å
point

1.8x1010 Cubic
aq-4.0822 Å.

Anneal 1/2hr at 11, 13


500°C, as–4.062 Å

Anneal 1/2hr at
1,000°C, an=4.034Å

BeO Hexagonal 1 to 4x1019 No detectable X-ray


at 40°C structural change
0.03to 0.12% 14
decreasein density.
Compressive Compressive
strength 169,000 strength 165,000 to
to 175,000psi 200,000psi

0.036%increasea- 15
axis
0.31%increase c-axis;
parameter change
anneals at 1,400C
Comp. strength Comp. strength
(p=2.839g/cc)
219,000psi 139,000
psi
(p=3.006g/cc)
310,000psi 100,000
psi

Thermal conduc- See Fig. 9.3.


tivity 0.01.1%increasea-axis
0.095%increasec-axis
19% decreasein mod
ulus of elasticity at 16
p=2.62g/cc
4% decreasein mod
ulus of elasticity at
p=2.74g/cc, more
than 10%decrease
in tensile strength
440 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
|
TABLE 9.2—EFFECTS OF FAST NEUTRON IRRADIATION ON NONFISSION.
ABLE OXII) ES—Continued
Changes on irradiation
Material Crystal structure | Fast neutron
dosage-nſwt
Property before Property after Referentº

BeO Hexagonal 1 to 4x1019 82–84%reduction in 16


at 40°C tensile strength
Elastic modulus. | Elastic modulus
(p=2.704g/cc)
38.3X10 psi 19.5x100psi
(p=2.902g/cc)
49.5x10°psi ~17.7x10%psi
Radiation effect
anneals at 1,300°C

7x1019 K=0.25 w/cm—"C | K-0.168 w/cm—"C 11


2 to 3x1020 p=2.84g/cc p=2.85g/cc
Lattice expansion 15
parallel to c-axis
still greater than at
lower exposures;
sintered samples
disintegrate to
powder

Al2O: Hexagonal 0.7 to 1x1021 0.01to 1% density 17


decrease
0.03%expansion a
axis
0.09to 0.3% expan
Sion c-axis

1.2x1019 Lattice expands


parallel to c-axis
0.0023%/101* nvt
Lattice expands
parallel to a-axis
0.0019%/101* nvt ls
Density decrease
0.12%at 1.2×101*
nWt
Expansion anneals
out in two steps at
300°C and at
550°C

3x1019 K=0.168wſcm—*C K=0.096 wicm—“C


p=3, 559g/cc p=3.553g/cc
(sintered Al2O3) (–0.16%)
4x1020 K=0.038 w/cm—“C 11
p=3.8 g/cc
(+7%).
6x1019 K=0.25 w/cm—“C K=0. 126w/cm—“C
p=3.983g/cc p=3.969 g/cc
(sapphire) (–0.35%).
6x1020 K=0.084 w/cm—"C
p=3.944 g/cc
(–1.0%)

5x1018 K-watts/cm—“C
40° K 5.7 3
100° K 4,8 2.2 (see Fig. 9.5) 19

5x1020 No detectable X-ray 20


change.
IRRADIATION EFFECTS IN URANIUM DIOxIDE 441

TABLE 9.2—EFFECTS OF FAST NEUTRON IRRADIATION ON NONFISSION


ABLE OXIDES-Continued

Changes on irradiation
Material Crystal structure | Fast neutron
dosage-n/vt
Property before Property after
——
| Reference

SiO, Hexagonal (or- 0 to 3x1019 a-axis expansion


quartz) to 870°C greater than c-axis
Tridymite—870° expansion (Fig. 9.1,
to 1,470°C Fig. 9.2); X-ray and
Cristobalite— hydrostatic density 21
1,470°to 1,910°C increase together;
little line broad
ening.

K at 10° K–w/cm-* C (a-quartz) 22

10 6 at 3x101*nwt
0.5 at 101*nwt
0.2 at 2.4x101*nwt.
0.02at 1.9x101*nwt

K at 10° K–w/cm—" C (fused silica)

0.0052 0.0061at 1.7x1019


0.007at 4.1x1019
(5° K)

3x101°to cº-5.393 A c.–541 Å 21, 23


6.6x1019 as-4.903 Å (6.6x101*nvt)
ac-501 Å
X-ray density de
creasesfaster than
hydrostatic den
sity; 20 reflections
greater than 90°are
smearedout. In 4
hours at 640°C,
higher order reflec
tions return, in 12
hours at 900°C
revert to normal
a-quartz structure.

6.6x1019to Quartz, tridy- All long range order 20,21, 24


2x1020 mite, cristobal- disappears (at 1.2x
ite structures, 1020nwt.)
density about p=2.26g/cc; anneal
2.6g/cc 16hours at 930°C,
polycrystalline
a-quartz structure
forms.

Coesite (synthetic | No structure or den- 21


silica) sity change at 2x1010
p=3.01g/cc nvt

Vitreous silica; p=2.26g/cc. Recover


p=2.20g/cc to density of unir- 25
radiated vitreous
silica at 900°to
1,000°C.
442 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 9.2 EFFECTS OF FAST NEUTRON IRRADIATION ON NONFISSION:


ABLE OXIDES–Continued
Changes on irradiation
Material Crystal structure | Fast neutron
dosage-n/vt
Property before Property after Reference

7×1019 Silica glass p=2.255g/cc


p=2.204g/cc
4x1010 p=2.23g/cc 11
3×1019 Plate glass
p=2.504g/cc p=2.53g/cc
6x1020 p=2.515g/cc

Zircon Tetragonal 5x1019 K=0.5 w/cm—"C K=0.0097w/cm—"C


ZrSiO4 p=3.73g/cc p=3.48g/cc 11

3×1020 p=338g/cc

3.6×1020 Density decreases4%;


anisotropic expan
sion; c-axis increases
50% more than 25
a-axis. High angle
reflections disappear.
Annealing does not
occur below 1,000°C;
damage still present
at 1,600°C.

Spinel Cubic 7x1019 K=0.105w/cm-9C K-0.055 wicm—” C


MgO. Al2O3 11
p=3.6g/cc p=3.6 g/cc

4x1020 p=3.6 g/cc

1x1020 Lattice expansion of 27


0.12%

Forsterite Orthorhombic 6x1019 K=0.105 w/cm—"C K=0.0315 w/cm-º C 11


2 MgO. SiO2 p=3.056g/cc p=3.03g/cc

Steatite Monoclinic 7x1019 K=0.0319w/cm—“C K-0.0117 w/cm—" C 11


Mgs (OH)2SiO10 p=2.796g/cc p=2.76g/cc

Corderite Orthorhombic 5X1019 K =0.0307 K=0.0084 11


4 (Mg, Fe) O w/cm—" C w/cm—“C
4 Al2O3
10SiO2

TiO2 Tetragonal 6x1019 K =0.069 K=0.046 w/cm–º C 11


(Rutile) w/cm—" C
p=4.01g/cc p=3.99g/cc

1x1020 No detectablestruc- ...?


tural change.

3x10-0 K =0.0273w/cm—" C 11

p=3.98g/cc

Porcelain 6×1019 K=0.113 whem—"C K =0.054whom-º C 11

p=3.41g/cc p=3.40g/cc
IRRADIATION EFFECTS IN URANIUM DIoxIDE 443

TABLE 9.2—EFFECTS OF FAST NEUTRON IRRADIATION ON NONFISSION


ABLE OXIDES-Continued

Changes on irradiation
Material Crystal structure | Fast neutron
dosage-n/vt
Property before Property after Reference

4x1020 K=0.036 w/cm—" C 11


p=3.39g/cc

Mica 4×1019 K=0.0071 K=0.00505whem—"C 11


w/cm—°C
p=2.845g/cc p=2,738g/cc

2x1020 K=0.0118w/cm—* C 11

p=2.444g/cc

Beryl Hexagonal 1x1020 No detectable struc- 27


BesAl2Si3Ois tural change.

3.6×1030 All coherent X-ray 26


reflections
destroyed.

Chryso- Orthorhombic 3.6×1020 1% expansion in c- 26


beryl BeAl2O4 axis; little lattice
distortion.

Phenacite Hexagonal 3.6×1030 0.7% expansion a- 29


Be2SiO4 axis; little lattice
distortion.

(a) Silicon Dioacide

Considerable experimental effort has been devoted to measurement


of fast neutron irradiation effects on the various polymorphs of
crystalline quartz as well as vitreous silica, both because of the magni
tude of the effects observed as well as the complexity of the changes.
At low neutron dosages (up to 2 to 3 × 10” nºvt) lattice expansions
occur with a greater expansion parallel to the a-axis than in the
c-direction (Fig. 9.1). Furthermore hydrostatic measurements of
density agree with the density decreases calculated from the lattice
expansions, indicating that the damage is uniformly distributed
through the structure (Fig. 9.2) [21]. As the neutron exposure in
8×10” nºvt, the two measures of density change diverge
creases to 7 to
more and more until the lattice expansions can no longer be measured
because of smearing of the diffraction peaks. At this stage, the single
crystal a-quartz pattern can still be restored by high temperature
annealing. With further exposure, all long-range order disappears (at
1.2×10° nºvt), and the density saturates at a value of 2.26 g/cc,
approximately 15 percent less than the initial density of 2.6 g/cc.
This final density is independent of the particular polymorphic form
-
444 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TEMPERATURE *c
O 3OO 6OO
4.2 I I I

--- THERMAL ExPANSION


IRRADIATION ExPANSION
3.6 H

|| TO-o-Axis
3.O. H.

2.4 H.

II TO – o – AXIS

O.6

NEUTRON ExPosURE x Io'9,nvt

FIGURE 9.1. Lattice Expansion of Quartz Caused by Neutron Irradiation and


Temperature [21].

of quartz exposed and, upon annealing at 930°C, a polycrystalline


a-quartz structure forms.
Of particular interest is the sigmoidal form of the density-neutron
dosage curve shown in Fig. 9.2 which differs radically from the prop
erty-irradiation exposure curve characteristic of processes showing
a saturation of damage in which the rate of property change decreases
continuously with dosage. Primak suggested that the sigmoidal
form of the damage curve (Fig. 9.2) was evidence of a thermal spike
mechanism of damage in that the decrease of thermal conductivity
upon irradiation would lead to attainment of ever-increasing tempera
tures in the spike, thus accelerating the rate of damage (Eq.
9.12) [28]. This interpretation was considered untenable by Klemens
less rapid quench
its

in that the longer duration of a spike and, hence,


ing conductivity, may
of
in

matrix lower thermal well lessen rather


a

than increase the disorder frozen into the spike [29]. He explained
by

Fig. postulating that


of

the form 9.2 less dense vitreous regions


IRRADIATION EFFECTS IN URANIUM DIOXIDE 445
15 -T I I T

>
E
co
fc 10 H. -
z
u
o
z
:
O
- —

;
*-

o
x-RAY
HYDROSTATIC

or
uſ
5 H. -
fl.

O l l 1– I
O 5 IO 15 20 25

INTEGRATED NEUTRON FLUx, x 10"?

FIGURE 9.2. Density Change on Neutron Irradiation of Quartz [21].

are formed in the crystalline matrix from the beginning of the


irradiation and that the restraint of the matrix prevents their expan
sion at low concentrations. Expansion and relief of compression of
the vitreous region occur at an increasing rate as the volume of the
latter increases and, hence, as shear stresses sufficient to cause plastic
flow are built up. Wittels, et al., and Kinchin and Pease conclude
from the similarity of the lattice expansions produced by temperature
and by neutron irradiation (Fig. 9.1) that interstitially displaced
into channels parallel anisotropic
an
fit

the c-axis leading


to

to

atoms
expansion, and that the structure becomes unstable further accumu
to

lation of interstitial atoms before saturation concentration can be


a

attained [1, 21, 26]. Wittels cited the insensitivity coesite, syn
of

a
by

thetic, high-density polymorph SiO2, damage neutron irradia


of

to
as

evidence for the important role


of

initial structure
in

tion the
of

accumulation damage [21].


Of interest are the results of irradiation of vitreous silica listed
in
of

Table 9.2. The density silica increases upon irradia


of

this form
percent and reaches
of

tion about value about 10”


at

dose
×
3

6
a

a
by

nºvt equal crystalline silica upon prolonged expo


to

that attained
sure (about 2× 10° nºvt). This result may
be

of

evidence the effect


accelerating diffusion and permitting local
of

neutron irradiation
in
446 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

atomic rearrangements to an internal equilibrium stable in the glass


at the irradiation temperature. In this sense, utilizing the terminol
ogy of glass technology, the irradiation permitted attainment of a
lower fictive temperature [30]. Wittels and Sherrill cited the com
mon density reached by the exposure of both glassy and crystalline
silica as evidence for the attainment of an equivalent structure [24].
However, the observation that the thermal conductivities of the ir
radiated quartz did not decrease to as low values as those of the glass
(Table 9.2) and that, upon annealing, vitreous silica irradiated to a
density of 2.26 g/cc reverted to the density of glassy silica, whereas
a-quartz was formed on annealing crystalline silica irradiated to
the same density as irradiated vitreous silica in Wittels and Sherrill's
experiments, casts some doubt on this interpretation [24, 25]. Further
data on properties, structure, and annealing characteristics of highly
irradiated silica are necessary before it can be concluded that a true
glass is formed by neutron irradiation of crystalline quartz.
The mineral beryl appears to behave similarly to quartz in that

all
coherent X-ray reflections are destroyed

an
exposure

of
3.6×10°

at
The magnitude the density changes mica, about
of

15
nºvt [26].

in
percent, indicates that this material also becomes highly disordered
density decrease percent the tetragonal
to of

[11]. about noted

in
A

is
4
on

an
mineral zircon exposure 3.6 10° nºvt, again with anisotropic
×

expansion
of

The latter material


of to

the c-axis relative the a-axis [26].


highly resistant annealing

no
the structural damage, apparent
to
is

annealing occurring below 1,000° damage still being


of

and traces
C

Note, however, that the single crystal lattice


at

C.

evident 1,600°

is
restored upon annealing damaged zircon crystals, indicating higher
a

damage crystalline
of

order this material than


to

resistance
in

in

quartz. major com


of

Examination Table 9.2 indicates that SiO2


is
a

showing major property


or

ponent structural change


of

those oxides
on fast neutron bombardment.

(b) Polymorphic Transformations Orides


in

by

Another interesting class


of

observations afforded the behavior


is

monoclinic ZrO2 and tetragonal BaTiO, which undergo polymor


of

phic transformations upon neutron irradiation cubic phases unstable


to

the temperature irradiation. The structural changes are con


of
at

by

tinuous with exposure and are again characterized anisotropic


expansion crystallographic ZrO,
of

of

the various axes. The case


of is

particularly interesting that cubic modification formed


in

is
a

larger lattice parameter than that stabilized cubic ZrO2, even though
of

pure
of

of

the evidence for the existence stable cubic modification


a

ZrO2 not clear [11]. (Note, however, this connection Sect. 9.3
in
is

stability the neutron-induced phase trans


of

below.) The thermal


IRRADIATION EFFECTS IN URANIUM DIOXIDE 447

formations in both minerals is high. A temperature of 800° C is


required to revert the cubic to the monoclinic form of ZrO2, whereas,
at annealing temperatures up to 1,000°C, the cubic form of BaTiO3
remains stable upon cooling to room temperature even though the lat
tice parameter of the cubic phase decreases continuously upon anneal
ing at successively higher temperatures. These observations empha
size the conclusions of Wittels concerning the importance of the initial
structure on the damage sustained by crystals and, furthermore, indi
cate that the cubic crystal structures may be inherently least suscepti
ble to damage by nucleon bombardment [21]. The high order of
stability evidenced by the cubic spinel mineral in Table 9.2 supports the
latter postulate [11, 27].

(c) Aluminum Oride and Beryllium. Oride

A further type of behavior is shown by the oxides Al2O, and BeO


which are damaged to a minor extent by fast neutron bombardment
but which, as will be shown in a later section, suffer extensive struc
tural damage as a result of fission fragment bombardment. The con
siderations outlined in Sect. 9.2.1 concerning atomic displacements
in materials by particle bombardment yield little information on
effects of the intensity of the bombardment. Other than the experi
mental observations on Al2O, and BeO, the only information bearing
on this point is that discussed by Crawford and Wittels on the neutron
damage sustained by diamond at two different flux rates but approxi
mately equivalent total exposures [11]. At the higher flux rate, the
ability of diamond to diffract coherently was found to disappear,
whereas at one-half to one-third the flux rate, a saturation decrease
of 4 percent in density with little attendant lattice strain was noted.
The experimental evidence in Table 9.2 for BeO and Al2O, indicates
a small density decrease, but quite appreciable reductions in thermal
conductivity. For the BeO the results of various investigators, plot
ted in Fig. 9.3, show considerable variance in the effect of irradiation
on change in thermal conductivity. Gilbreath and Simpson report
a relatively minor reduction in thermal conductivity at the highest fast
neutron exposure used; reductions of 37 percent were noted after an
exposure of 4 × 10” nºvt [14]. On the other hand, Elston and Caillat
reported reductions of over 80 percent at exposures of 5 to 7 × 10” n.vt,
little greater than those of the other investigators [15]. Both investi
gators found the effects of irradiation to be more marked the higher
the density of the samples. Presumably the differences between the
results of the two investigations, shown in Fig. 9.3, are attributable
to the differences in sample densities. Elston reported that sintered
samples exposed to 2 to 3 × 10° nºvt disintegrated to powder, presum
448 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
I I T I i —T- i I i I

-
TREATMENT DENSITY FAST NEUTRON REFERENCE
2.O H. ExPOSURE -
O O UNIRRADIATED 2.990 g/cc ELSTON AND
--
O © IRRADIATED 5-7 x 1019
nwt. CAILLAT
-- -- -4
A 8 HR AT 520°C
-- in -
A 40 HR AT 600°C
-- -- --

-
Ó 4 HR AT 400°C
-- -- -
- @ 12HR AT 1400°C
g/cc GILBREATH ANDT
D UNIRRADIATED 2.87
E IRRADIATED --
-
- 1x 109mvt SIMPSON
-- ++
V 1.5 HR AT 98.0°C

O |

o I l l l I | l I I l
50 IOO 2OO 3OO 4OO 500 6OC
TEMPERATURE, "c

FIGURE 9.3. Effect of Neutron Irradiation on Thermal Conductivity of BeO.

ably as a result of continuing anisotropic lattice expansions and result


ing strains due to the deformation [16]. Correspondingly, compres
sive and tensile strengths were drastically reduced at lower exposures.
The measurements of Martin on dimensional expansion of Al2O3 in
dicate that saturation has not been attained at exposures of 1.2 × 10”
nºvt (Fig. 9.4) [18]. The thermal conductivity decrease in Al2O, has
not saturated at an exposure of 6 × 10° nºvt, although no detectable
X-ray change has been noted in Al2O, at this exposure level [20].
Measurements of the thermal resistivity introduced in Al2O, by fast
neutron exposure are shown in Fig. 9.5 and, for comparison, the
thermal resistivity introduced by the inclusion of about one atom
percent chromium in Al2O, is also plotted [19]. The greater effec
tiveness of the fast neutron exposure (which, from Table 9.1, intro
duces approximately a number of displacements at an exposure of
5× 10° nºvt, equivalent to one atom percent of chromium) in increas
ing thermal resistance is evident in this plot. From the constancy of
thermal resistance with temperature, Berman, Foster, and Rosenberg
concluded that the damage regions are nonspherical and that the
phonon scattering is due to a long, thin obstacle [19]. Antal and
Goland measured directly the number of defects in neutron-irradiated
a-Al2O, by scattering of long wave-length neutrons; their results,
IRRADIATION EFFECTS IN URANIUM DIOXIDE 449
O. 15 I I I I I

f– AS CALC FROM KINCHIN AND PEASE

d - ANTAL
-
AND GOLAND
O. i O

O.O5 H

s—"T m—-
PARALLEL TO C-Axis

m—T PERPENDICULAR TO C-Axis

| | l |
O 2 4 6 8 IO 12
FAST NEUTRON INTEGRATED ExPosure (nvt), x lo'8

FIGURE 9.4. Density Changes and Defect Production in Fast Neutron Irra
diated o:-Al2O3.

shown in Fig. 9.4, indicate a continuous increase in number of defects


at exposures up to 1.2 × 101* nºvt [17]. The number of defects experi
mentally observed was much smaller than those predicted by the
displacement mechanisms previously discussed, even though tem
peratures of 1,800° C were required to anneal out the damage [1].

9.2.3 The Metamict State

Still another class of phenomena has been extensively investigated,


resulting from the finding that certain natural minerals occurring in
association with the radioactive elements uranium and thorium have
lost their optical birefringence and show little or no trace of coherent
X-ray diffraction. This condition, noted mineralogically as metamict,
has been most recently reviewed by Pabst [31]. Other properties of
the metamict state generally, but not universally, noted are reconstitu
tion of the crystalline structure upon heating (such recrystallization
often yielding a single crystal structure), evolution of heat often at
tending the recrystallization, and usually an increase in density upon
heating, the density increase corresponding to the decrease observed
during metamictization. The heat of transformation of metamict
gadolinite back to the crystalline form has been found by Faessler to
450 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

O.4 I I I

.3 -

O
º 3:7– 5.02xio"n,


202x 10" nºvt

sexio'n
v.

-- ---
lºck
-
Arousſes

1.5x io" nºv


I
l
I
O

2O 40 60 80
O

TEMPERATURE, *K

FIGURE 9.5. Increase Thermal Resistivity of Sapphire upon Neutron Irradia


in

(Courtesy, Institute Physics and The Physical Society.)


of

tion [19].
as

magnitude
be
of

of

of

the same order solidification


to

the heat the


crystalline oxide [32]. The specific heats recrystallized and meta
of

mict gadolinite were, furthermore, noted bear the same relationship


to

crystalline silicates and their glassy equivalents. The vol


as

of

those
ume changes occurring
as

of

minerals result their metamictization


in to in

are reported by Hutton sufficient magnitude cause shattering


be
of

to

which the metamict crystals may be en


of

unaltered minerals
The damage these minerals the radio
to

closed [33]. attributed


or in

is

active decay the crystals. For


of

uranium thorium contained


in

example, years percent the U* originally


15

of

contained
in

in

10°
a

mineral decayed eight separate g-disintegrations [1]. As


to

lead
in

Table 9.1, the major fraction the damage can


be
of

noted attributed
in

the recoil nucleus rather than the emitted a-particle.


to

to
IRRADIATION EFFECTS IN URANIUM DIOXIDE 451

Many mineral types are found in the metamict condition. The list
ing in Table 9.3 is by no means complete, but illustrates the types of
minerals which have been observed to exist in the metamict state. It is
of interest to note that not only minerals high in silica content are ob
served to become metamict, but also those containing columbates and
titanates as constituents. This finding suggests that neutron radiation
damage studies in the latter system would be highly instructive. Ap
parently, as in the case of neutron-damaged crystals, single crystals
are reformed on annealing partially metamict minerals in which a re
sidual matrix may exist on which growth of the crystalline phase can
occur, whereas polycrystalline patterns result on annealing fully meta
mict structures. The annealing of the latter, as exemplified by the
experiments with thorite and zircon, is quite complex, involving first
the precipitation of crystalline Tho, or ZrO2 before the final stable
structure is attained.
Kinchin and Pease have used the normalizing parameter of num
ber of bombarding particles per atom in comparing alpha disintegra
tion and fast neutron damage (in the latter case, the number of
primary displacements produced by neutron flux is utilized) [1]. It
may be noted in Table 9.2 that an integrated flux of about 3× 10”
nºvt (corresponding to about 7 × 10° primaries/atom) expanded the
crystal lattice of ZrSiO, about 4 percent and decreased the total
density about 9.5 percent, indicating that, as in the case of neutron
damaged SiO2, amorphous
regions are produced in addition to the
expanded ZrSiO, lattice. Hurley and Fairbairn report a dose of
3× 10^* particles is required to half-saturate alpha disintegration
damage [34]. It, thus, appears that for the case
in zircon crystals
of zircon, the observations of naturally damaged crystals are in good
agreement with the results of fast neutron damage studies. On the
other hand, quartz, which is so readily damaged by exposure to fast
neutrons, is not reported as occurring naturally in the metamict state.
This observation may serve to confirm the importance of damage rate
discussed above.
It
may be noted from Table 9.3 that a wide variation exists in the
dosages at which minerals are observed to become metamict. Thus,
the siliceous minerals allanite and zircon become metamict at about
3 to 10' alpha particles per atom, whereas thorite requires almost
6 ×
500 times as great a dosage, and the mineral phenacite (Be2SiO,).
with a hexagonal lattice, is apparently unaffected by either neutron
(see Table 9.2) or alpha damage. The particular nature of the cations
in the silicates thus plays an important role in determining suscepti
bility to In agreement with the results of Wittels on the
damage.
resistance of a particular polymorph of SiO2 (coesite) to fast neutron
damage, Pabst found the high temperature monoclinic form of

57.4789 O—61—30
452 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

ThSiO, to be resistant to alpha disintegration damage, illustrating


again the importance of initial crystal structure in determining resist
ance to particle bombardment [21, 31].
Many minerals, in addition to quartz, show no alteration in struc
ture as a result of alpha disintegration. Thus, uraninite (UO2) and
thorianite (Tho.) are structurally unaffected in spite of the severe
dosages to which they have been subjected. It is of interest to record,
however, that weathered uraninite in which the uranium has been
oxidized to a higher valency generally occurs in the metamict condi
tion. Minerals such as xenotime and monazite (R*PO,) also retain
their crystal structures at very high dosages.

9.2.4 Factors Affecting Structural Instability

It, thus, appears clear that structural damage in the metal oxides by
either fast neutron or alpha decay bombardment is a manifestation of
the instability of the crystal lattices to such exposures. It appears
equally evident that the degree of damage for a given dosage is sensi
tively affected both by the sterochemistry of the compound (as ex
emplified by the varying resistance to structural damage exhibited by
the different isomorphs of SiO, and ThSiO,) and by the specific chem
ical nature, ionic radius, bond strength, etc., of the constituent atoms
of the crystal, vide the differing sensitivities to damage of the ortho
silicate minerals ZrSiO, (zircon), ThSiO, (thorite), and Be2SiO,
(phenacite). An interesting example of the interplay of these effects
is afforded by the observation of lack of structural damage in the rare
earth phosphates, such as monazite (CePO.), which are structurally
similar to the high temperature form of ThSiO, (Huttonite), also
resistant to metamictization. However, small variations in the link
age of the MeO, tetrahedra from which such structures are built lead
to the structures of zircon (ZrSiO,) or thorite (ThSiO,) which are
noted to be metamict; phosphates (such as xenotime), vanadates, and
arsenates of the same structure are not reported as metamict. On the
other hand, complex oxides, such as Fergusonite (Table 9.3), of other
elements (Nb, Ta) in the Group V-A, in which the metal ion is
octahedrally rather than tetrahedrally coordinated with oxygen, are
shown to become metamict [35].
Insufficient data under directly comparable irradiation conditions
are available to permit formulation of general rules for crystal in
stability. Thus, metamictization under alpha decay exposures has
been shown in the case of SiO, not to be comparable to structural
damage under fast neutron bombardment, whether through differences
in the type or the time scale of the bombardment. However, various
authors have proposed criteria for predicting susceptibility to damage.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 453

TABLE 9.3—METAMICT MINERALS

Mineral General composition Normal crystal Remarks Reference


structure

Allanite-------------- R!'R''(oh)sion. Monoclinic------ Becomes metamict at about 6


X 10-4of atom.
Fergusonite Y (Ch, Ta) Oa ----- Tetragonal ------ Becomes metamict at about 31,36,37
tantalite) 2 × 10−3 aſatom; releasesheat
Inignite). at 600°to 700°C on heating
metamict crystal, and poly
crystalline crystal structure
reconstituted.
Tetragonal -----. 31

-| ThSiO4- - - - - -- ------ Tetragonal.----- Becomes metamict at 0.13 aſ 36


atom. Monoclinic ThSiO4 31
(Huttonite) is not metamict.
Annealing of metamict tho
rite begins at about 600°C,
forming tetragonal ThSiO4
and Thor; at temperatures
above 935° C, monoclinic
ThSiO4 forms.
Gadolinite- - - - - - Y2Fe3+Be2Si2O10----- Evolves 81to 89 cal/g on heat 37
ing; polycrystalline X-ray 31
pattern produced after heat
ing at 950°C.
Ellsworthite- - - - - - --. CaO.Cbz O3.2H2O.--. Exothermic peaks at 980° C 37
and above.
(Y, Ca)(Ch., Ta,
Ti);Os. Exothermic peaks at 600° to 37
Orthorhombic---
Y(Cb, Ta):Os 650°C.
Ce(Ch., Ti):Os. --
ZrSiO4- ------------- Tetragonal ------ Becomes metamict at 3 X 10-4
a/atom; partly metamict
crystals form single crystals
on heating;fully amorphous
zircon forms ZrO2 and, at
higher temperatures, poly
crystalline ZrSiO4.

Thus, Primak, Pellas, and Crawford and Wittels cited a high degree
of ionicity of bond as a prerequisite for resistance to damage: Hutton
related the metamictization of ZrSiO, to the unfavorable radius ratio
for eight-coordination of ZrO ions; Kinchin and Pease, and Craw
to
ford and Wittels, proposedthat structures such as those of quartz
and zircon are unable to accommodate the number of interstitial
atoms produced by bombardment without disruption of the lattice
[1,11, 26, 27, 33, 36]. None of these suggestions appears to be
uniquely capable of prediction of degree of stability. Indeed, it would
be surprising if some single descriptive property were adequate to
predict the results of the complicated damage and annealing processes
occurring during particle bombardment.
From the observation that crystals when extensively damaged lose
coherent X-ray diffraction,
all

tempting
of

trace
to

assume
is
it

a
454 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

phenomenological relation between formation of the glassy state and


particle bombardment damage and, thus, to utilize the information
derived from glass technology to predict crystal irradiation damage.
In fact, description of particle bombardment processes in terms of
thermal or displacement spikes in which a region of the crystal is
rapidly quenched from a high temperature or a condition of consider
able atomic disorder strengthens the analogy. On the other hand.
sufficient differences, particularly with respect to annealing behavior,
have been discussed above to raise doubts, which can be settled only
by further structural and property studies, whether the structure of
damaged crystals is sufficiently disordered to resemble that of a glass.
As in the case of disordering of crystals by radiation, a number of
criteria have been proposed for glass formation. Zachariasen pro
posed the following set of structural rules [38]:

1. Anoxygen atom is linked to not more than two metallic atoms.


2. The number of oxygen atoms surrounding the metal atom must be
small.
3. The oxygen polyhedra or triangles share only corners and form
three-dimensional networks.
4. At least three corners in each oxygen polyhedron must be shared.
Jones listed a number of other criteria [30]. He observed that the
activation energy for viscous flow was in general much higher for glass
formers than for liquids which do not form glasses. This high acti
vation energy inhibits crystallization of the liquid and is related to
one of the rules of Zachariasen, namely, the formation of a three
dimensional network. These effects are again related to the directed
chemical bonds formed by the glass former and characterized by low
coordination number. Directional bonds can be formed in hydrogen
bonded substances such as water and glycerol or, more importantly in
the present connection, in salts having at least partially homopolar
bonding. Purely ionic-bonded, metallic-bonded, or van der Waals
bonded substances are characterized as nonglass formers and show
high coordination numbers and low activation energies for viscous
flow. Highly charged, small cations which can polarize the surround
ing oxygen ions and, hence, impart a homopolar or directional char
acter to the bond favor glassy oxides. Jones proposed the value
Z - - - - -

a2 .
where Z is cation charge and a is the distance between cation and
oxygen ion centers, as descriptive of the homopolar nature of the
bond. High values of this parameter characterize oxides which are
glass formers; median values, oxides which are intermediates (i.e., can
replace glass forming oxides in high proportions, but cannot them
selves form glasses); and low values, network modifiers (i.e., oxides
IRRADIATION EFFECTS IN URANIUM DIOXIDE 455

which break up the three-dimensional network formed by the glass


forming oxides). Other measures of covalency of bond such as radius
ratio of cation to oxygen ion or the position of the cation on the Paul
ing scale of electronegativity rank the metal oxides in similar se
quences [39].

A useful measure of bond strength was derived by Sun based on the


dissociation energy of the oxide to gaseous atoms divided by the cation
to anion coordination number [40]. Values of the bond strength so
calculated are listed in Table 9.4. Note, since the dissociation energies
increase with increasing valency and since the coordination number
will tend to decrease with decreasing cation size, that this description
of bond strength will tend to be equivalent with that of Jones given
above. -

In a complexmixture of oxides, the average bond strength, and,


thus, a measure of the stability of the lattice, may presumably be
derived by averaging the single-bond strengths in the correct ratio.
Comparison of Tables 9.2 and 9.3 with Table 9.4 shows many qualita
tive agreements between predictions based on bond strength and obser
vations on structural damage. Particularly instructive for the pur
poses of this chapter are the positions of hexavalent and quadrivalent
uranium; the former would be predicted to undergo considerable
structural damage, the latter only moderate damage. It is obvious,
however, that considerably more data are required on damage sus
tained in various oxide crystals before predictions made on the basis
of rankings such as those shown in Table 9.4 can be made with any
degree of confidence.

9.2.5 Summary

This review of structural and property damage in metal oxides by


irradiation has emphasized the important role of crystalline structure
and, hence, of chemical bonding and has revealed a qualitative similar
ity to solidification processes which are also sensitive to chemical bond
ing effects. Derivation of more quantitative relations suitable for
prediction of stability under particle bombardment would appear to
be a fruitful field for investigation.

9.3 STRUCTURE AND PROPERTY CHANGES IN FISSIONABLE


OXIDES UPON IRRADIATION

In the previous section information was summarized on the effects


of neutron and alpha-decay particle bombardment on nonfissionable
oxides, and it was shown that the degree of damage suffered was a
sensitive function of structure and composition of the oxide crystal.
456 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 9.4—CALCULATED BOND STRENGTH OF OXIDES [40]

Dissociation Coordination Singlebond


M in MO, Valence energy per MO. number strength
(kcal) (kcal)

GLAss-ForMERs:
B-------------------- 3 356 3 119
Si--------------------

1ſt
4 424 4
Ge------------------- 431 108

3 4

4
Al------------------- 402–317 101-79

4
U--------------------

89 Ql
725

8
B-------------------- 356

4
3
P-------------------- 442 111–88"

4
5
V-------------------- 449 112-90'

4
5
As------------------- 349 87-70'

4
5
Sb------------------- 339 85–68"

4
5
Zr-------------------

81
485

6
Intermediates:
U--------------------

7; 7? 75 74
593
2 4 2 4 3 2 2 4 4

8
Ti------------------- 435

6
Zn------------------- 144

2
Pb------------------- 145

2
Al------------------- 317–402 53–67

6
Th-------------------

60 61 63 64
516

8
Be------------------- 250

4
Zr------------------- 485

8
Cd------------------- 119

2
Modifiers:
Sc.-------------------

20 20 30 32 32 33 36 36 36 7 39 43 45 46 50 58 60
362
6
3

La------------------- 406
7
3

Y-------------------- 399
3

Sn------------------- 278
4

Ga------------------- 267
3

In------------------- 259
3

Pb------------------- 232
4

Mg------------------ 222
2

Li------------------- 144
1

Pb------------------- 145
2

Zn------------------- 144
2

Ba------------------- 260
2

Ca------------------- 257
2

Sr-------------------- 256
2

4 8

Ci------------------- 119
2

Na------------------- 120
1

Cd------------------- 119
2

K-------------------- 115
1

Rb-------------------
10 1. 12

115 10
1

Hg------------------- 68
2

Cs------------------- 114 12
1

*Double value arisesfrom possibility oxygen being doubly bonded.


to
of

bond
1
IRRADIATION EFFECTS IN URANIUM DIOXIDE 457

Similar effects are noted in the case of oxide crystals subjected to fis
sion fragment damage, with the exception that many oxides which are
damaged to only a minor extent by fast neutrons or alpha decay prod
ucts undergo extensive deterioration when exposed to the fissioning
process. The remarkable and technologically important stability of
UO2 under irradiation is best revealed by comparison of its stability
with the marked changes exhibited by other oxides under fission frag
ment irradiation.

9.3.1 Structural Changes

(a) UO,

Boyko, et al., exposed natural UO, powders to integrated neutron


fluxes up to 4×10° nvt at temperatures less than 100° C and meas
ured changes in the diffraction peak profiles after irradiation [41].
As shown by comparison of Figs. 9.6 and 9.7, diffraction profiles of
the 531 reflection before and after irradiation, respectively, the prin
cipal effect of an exposure to 4×10” nvt (or 3.4×10* fissions/cc) was
broadening of the diffraction maxima. Analysis of the cause of the
line broadening revealed, as shown in Table 9.5, that it resulted primar

.
|
A.

50

|,
40 |

º
É
30 | l

-
O
II3 II2 || ||
2 8 (DEGS)

Figure 9.6. Effect of Irradiation on X-ray Diffraction Peak Profiles of UO, ;


Unirradiated MCW UO, 531 Reflection; Full Line Is Slow Traverse Trace
(5 Min/Deg of 26); Points Represent Statistical Count with 1.2 Percent
Error [41].
458 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

50

4O

3O \
§

2O
Y "

|13 | 12 | li
2 8 (DEGS)

FIGURE 9.7. Effect of Irradiation on X-ray Diffraction Peak Profiles of UO, ;


MCW UO, Irradiated to 3.42X10” Fissions/cc, 531 Reflection; for Neutron
Cxposure IDamage Compare with Fig. 9.6 [41].

ily from lattice strain and that reduction in crystallite size played only
a minor role in broadening the diffraction maxima. Irradiation of
a strained UO, lattice, in which the strain was produced by preparation
of the UO, by low-temperature oxidation of metallic uranium or by
low-temperature oxidation of UO, introduced little additional strain
and, even in the case of the UO, by uranium oxidation,
prepared
caused some relaxation of the initial strain, presumably as a result of
irradiation annealing. The magnitude of the average strain intro
duced in the UO, lattice by the irradiation, about 0.16 percent, is com
parable with that which might be introduced by the solution of about
7 percent Tho, in UO, (were the strain in this case localized about
each thorium ion): consequently, it would be expected that the changes
in properties observed would be minor and of the extent expected from
such additions.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 459

º
TABLE 9.5—EFFECT OF IRRADIATION ON UO, STRUCTURE

Material asonse || ||
*"...lº"
|cºls
+0.002 broadening broadening B, (6) in
(sº ll.9
ſº,
A
o
Rºr
ences
in degrees in degrees| radians
26 at 000 26 at 000

41, 10
0.

0.
MCW UO, (U.O.; pre- 118 00044 670

5.
468 0.100
tı)

||
0
pared by reduction 5.475 ------------|------------|------- ---|---------- 10
of

4.7×1016
UO, Hz, hr 5.477 ------------|------------|-------------------- 10
1. 9.

1x1016
at

790°
in

5.473 |------------|------------|----------|---------- 10
C

4x1017
560 41
5.
1.95X101* 469 0.368 0.141 0.0016
3.2x101* 5.477 ------------ - - -- - - - - - - - ---- - - - -- - - - - -- - - - - - - - 10

0,
5.470 0.378 0.158 001.65 500 41
3.

4x101s

||
6.2×101s 5,475 ------------|------------|----------|---------- 10
------------|------------|----------|---------- 10
5.

1.1.x1021 475
5. 5. 5,

41
0.

Steam-oxidized UO2 467 0.621 0.245 0.0027 320


(UO: 61) prepared by 1.7x101s 471 0.431 0.100 (0.188 790 41

|| ||
0 0
reacting uranium with 2.9x101s 471 0.446 0.110 (0195 720 41
2,200
at

steam 48 hr
psi, 343°C

41
0.

UO1 prepared by oxi- 0.439 0.100 001.92 790


5.

443 ||
is

0
dizing MCW UO, 560 41
0.

5.446 566 0.141 0.0025


at

1.95X101s
|
C.

130°to 160° 0026 560 41


5.

3.4×1018 446 0.600 0.141


0

An additional low temperature UO,


of

effect irradiation in of was


by

by

observed both Boyko and Berman and tabulated Table 9.5,


is

namely, slight lattice expansion which saturates fission density


be at
a

101° fissions/cc [10,41]. anticipated that


in of

It

less than would


×
5

ionic solids both lattice vacancies and interstitially displaced ions


would cause lattice expansion. Mott and Littleton showed that,
in
a

the ionic solid NaCl, nearest neighbors lattice vacancy are dis
of
a
by

placed outward about 0.07 the distance between the neighbors,


X

slightly larger displacement


an

and that inter


|

would occur about


a

stitially displaced ion [42]. Eshelby showed that the fractional vol
is by

singularities in
an

ume change caused


an
of

atomic fraction
in
f

finite elastic medium given by


1-
A

. Cf
V

iH:
a
.

=i--410.3
.

the strength the singularity, Poisson's ratio, and


to of of

where
("
is

is
o

the volume per atom matrix [43]. Assuming the displacement


is
Q

apply for UO, may


be

be

estimated above for NaCl


to

estimated
("
|

Assuming =0.3 and that


cc.
an ".7

10−4 each displacement produces


×

interstitial-vacancy pair,
IA V

d.
=

2.0S

the atom fraction displacements. the lattice expansion


If

where
is
d

101° fissions/cc, and utilizing the


to

then assumed saturate


at

×
is

5
460 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

estimate of 6×10" displacements per fission in UO, tabulated in


Table 9.1, a lattice expansion of 0.015 angstrom unit should be noted
in irradiated UO, in agreement with the magnitude of the effects
noted in Table 9.5. Thus, the structural effects of fission fragment
irradiation of UO, to exposure levels at which each atom would have
been displaced some 100 times by fission fragments are consistent with
the production of a relatively low saturation concentration of point
defects which would be anticipated to affect relatively mildly the
properties of the oxide. A similar behavior was demonstrated by
Berman in the case of the irradiation of UO2-base solid solutions con
taining ZrO2 and CaO which showed a lattice expansion of the same
magnitude as that found for pure UO, [10]. Similarly, Wait, as re
ported in Ref. 44, found the lattice expansion of single crystal
UO, to saturate at a value of 0.004 angstrom unit for 1.9 × 10" fis
sions/cc exposure, with no further change on exposure to 50 times
this burnup. This expansion annealed completely in 18 hours at
200° C.

(b) U,0,

In contrast to the slight structural damage sustained by UO2 from


fissioning exposures, both Berman and Boyko found the X-ray struc
ture of UAOs to be almost completely destroyed by neutron exposure
[10,41]. This finding is not unexpected in view of the observations
recorded in Sect. 9.2.3 of the metamictization of weathered uran
inites. What is surprising is the low level of irradiation exposure at
which all discernible traces of structure disappeared. Berman, after
exposures of 1.9 × 10" and 4.8×10" fissions/cc at room temperature,
found no evidence of the original diffraction maxima, whereas Boyko,
after 1.3 × 10° and 2.1 × 10's fissions/cc at a slightly higher tempera
ture of irradiation (135° C), found only very weak diffraction peaks.
Also, Childs and McGurn, utilizing Debye-Scherrer powder photo
graphs, found no diffraction peaks after 8 × 10° fissions/cc exposure
at 30°C [45]. It may be estimated, using the lowest of the exposures
listed above, that to produce the destruction of the structure noted
each fission event must have affected at least 3.5 × 10" atoms, a factor
of 60 greater than the number of displacements estimated in Sect.
It,

thus, appears likely that fission events U.Os can affect the
in

9.2.1.
considerably larger volume
of

of

structure material than that


be in
a

which direct displacement reactions can occur. The process may


similar principle the disordering
to

to

of

that discussed relation


by in

in

quartz displace
of

neutron irradiation: critical concentration


at
a

ments, the orthorhombic lattice (), may no longer


be
of

able
to
U

by

accommodate the anisotropic strains introduced the displaced atoms


and, hence, long-range disruption the lattice will result.
of
a
IRRADIATION EFFECTS IN URANIUM DIOXIDE 461

Additional evidence for the markedly different behavior of UO.


and UAO, under irradiation is afforded by the stored energy measure
ments of Childs and McGurn [44, 45]. They observed no indication
of energy release in UO, irradiated to a fission density of as much as
8×10" fissions/cc at temperatures up to 750° C. Small energy re
lease peaks in bodies of composition UO, 0 at about 500° C were at
tributed to annealing of damage in the uranium and uranium nitride
inclusions present in this air-fused material. On the other hand,
energy releases saturating at a value of about 25 cal/g at exposures
over 7.5 × 10" fissions/cc were noted in U.O., and the temperature
at which the energy was released increased from 150 to 350° C as
the exposure increased from 7.5 × 10” to 3 × 10" fission/cc. The varia
tion of energy release and temperature of release with neutron ex
posure is shown in Figs. 9.8 and 9.9. The energy release at low ex
posures was ascribed to an annealing mechanism dependent on the rate
of oxygen ion diffusion and is presumed to indicate recrystallization of
damaged regions upon remanent nuclei of undamaged UAOs. As the
exposure exceeds 7 × 10" fissions/cc, no nuclei remain, and annealing
of damage occurs by nucleation and growth of crystalline UAO, nuclei
within an essentially amorphous matrix. In agreement with this
supposition, X-ray examination of the specimens after they had re
leased their stored energy showed reconstitution of the UAOs lattice.
Furthermore, the energy releases after these higher exposures are con
sistent with the estimated latent heat of fusion of UAOs. From these
data, Childs and McGurn estimated that the volume of material dam
aged by one fission fragment is 2.7 × 10-1° cc, or that the number of
atoms affected per fission event is 35 × 10°.

(c) I’.0,
Childs and McGurn reported that, in spite of the crystalline com
plexity of U.O., comparable with U.O.s, little or no damage or storage
of energy was noted upon
its

irradiation which complete


to

at

levels
damage UAO, stability
of

of

structural was noted [44]. The cubic


irradiation thus, again, graphically illustrated. Non
to

lattices
is

stoichiometric irradiated UO, appreciable


an
to

was found
os

release
ir
of

energy 340°C. This result probably indicates


an

amount
at

radiation-induced homogenization (see Ref. the precipitated


of
7)

U.O., phase (see Chap. the UO, matrix and reprecipitation


its
6)
in

upon subsequent heating. This finding, considerable significance


of

in

interpreting the results defected fuel element operation


of

hot water
in

loops (see Sect. 9.4.10), suggests that nonstoichiometric ()., operates


U

in-pile
its
as

the single-phase solid solution and that, furthermore,


precipitation upon cooling can probably prevented only with quite
be

rapid cooling rates.


462 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
I I I I
CAL 9 º'

——
3O H. TOTAL ENERGY -
O
2-? 8.
O O
#:
220 H
O -
o PEAK ENERGY
ë
o:
u!
->
ul
|O H. -

I I l I
105 loſe 10'7 lole lo” n.cm *
7.1x 1013 7.1x 101* 7.1x 1015 7.1x 1016 7.1x 1017 FISSIONS
cc
NEUTRON EXPOSURE, LOG SCALE

FIGURE 9.8. Energy Stored by Irradiated U3O, as a Function of Neutron


Exposure [45].

•C I I I I

400 H. -
O
8-7,
O

3OO H.
Te
-
O
i RATE OF RISE OF TEMPERATURE
Approximately 5x loº-c sec-'
200
T. - calorimETER TEMPERATURE
T - TEMPERATURE
SPECIMEN
of IRRADIATED
(

|OO
| l I l
1015 1016 107 1018 Iolº n.cm 2
7.1x 10'5 7 | x 10" 7.1x ſolº 7.1x 1016 7.1x 107 Fissions
cc
NEUTRON Ex POSURE, LOG SCALE

FIGURE 9.9. Irradiated and Control Specimen Temperatures at Time of Maximum


Rate of Energy Evolution [45].

(d) A7,0,

The severe structural damage sustained by UAOs on exposure to


fissioning might be anticipated from the correlations listed in Table 9.4
in which relative metal-oxygen bond strengths are compared; hexa
bonding characteristics
its

valent uranium is noted to be comparable in


by

with other oxides which are readily disordered irradiation (or


IRRADIATION EFFECTS IN URANIUM DIOXIDE 463

which readily form glasses). On the other hand, the oxide Al2O,
(corundum) was shown to be quite stable under fast neutron bom
bardment and, from Table 9.4, there is no apparent reason why it
should be affected more severely than UO, by fissioning reactions.
Berman found, however, that a sintered mixture of 41 volume percent
Al.0, plus 59 volume percent UO, exposed to 1.5×10" and 3.7×10"
fissions/cc in each case exhibited no coherent X-ray diffraction from
Al2O, phase [10]. This Fig.
the

illustrated which

in

in
9.10

is
the diffraction traces from this material before and after irradiation,
respectively, are compared. Note that after irradiation only the dif
fraction peaks from the UO, phase are visible and that coherently
scattering Al2O3, present all, percent

10
at

to
if

If is reduced less than


originally present.

of
volume assumed that the fraction
of

the it
is
I

LN-T
#
*

LA
C
T| | |

SPURIOUS TRACES
#

3k
|
I

70 60 5o 4O 30 2O
DEGREES
2
8

Figrkº. 9.10. X-ray Spectrometer Traces Fissionable Oxides before and after
41 of

Volume Percent UO, Volume I’ercent Al-O, Unirradiated


59

Irradiation:
A.

Same as A., 1.5×10" Fissions/cc Volume I’ercent ZrSiO,


12
SS
C.

Volume
B.

Percent UO, Unirradiated;


C,
as
D.

Same 2.4×10" Fissions/cc [10].


(Courtesy, North-Holland Publishing Company.)
464 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS |
FISSIONS PER co x ſo-29
2 4 6 8 IO 12 14
25 T

O
2OH.
º –
ſe
|
O
©
G

15 -
|
DENSITY CHANGE CALCULATED
| mwaro so-3 experiment FROM DIMENSIONAL
| MEASUREMENTS

off wapp 29-lo ExPERMENT DENSITY CHANGE CALCULATED T


FROM DIMENSIONAL
| MEASUREMENTS

© WAPD 14-27828 EXPERIMENT DENSITY CHANGE CALCULATED


5 FROM PRE-POST DENSITY —
MEASUREMENTS

o l 1–1–1–1–1–1–1–1–1–1 l
O O.2 O4 O.6 O.8 I.O 1.2
ATOM PER CENT BURNUP OF TOTAL ATOMS

FIGURE 9.11. Decrease in Density of Al2O3-H21 Weight Percent UO, as a Function


of Burnup [10]. (Courtesy, North-Holland Publishing Company.)

the fission track passing through the Al2O3 phase is proportional to


the relative volume of Al2O, in the sample, it may be estimated that
destruction of structure occurred at a fission density of about
6.3×10" fissions/cc, which is equivalent to the dose at which Childs
found the stored energy in U.O., saturated (Fig. 9.8) [45]. Thus, in
spite of the disparate stabilities of the two crystals, structural dam
age is found to occur at equivalent fission dosages. It should be noted
that no damage was observed by Berman in samples of Al2O, ir
radiated along with the UO-Al2O, mixtures to equivalent neutron
flux exposures.
Confirmatory evidence of the marked structural damage resulting
from fissioning in an Al2O, matrix is afforded by the results of Berman,
et al., who observed the marked swelling shown in Fig. 9.11 in
sintered high density platelets of Al2O,-UO, containing nine vol
ume percent (21 weight percent) of UO, in the form of small
particles, in diameter, in a matrix of 95 percent dense a-Al-O.
5p

[10]. The samples, 0.040 inch thick, were clad in Zircaloy-2 and,
in the case of the WAPD–30 and WAPD-29 experiments, were ex
posed in a high temperature water loop at 2,000 psi pressure with
the surfaceof the oxide body at a temperature of about 400° C.
As shown in Fig. 9.11, marked dimensional and density changes were
observed which were apparently saturated at the lowest exposure
level investigated. It may be inferred from the structural damage
IRRADIATION EFFECTS IN URANIUM DIOXIDE 465

noted in Fig. 9.10 that the density changes were probably complete
at a dosage of 10° fissions/cc or less. Since the occurrence of
amorphous or glassy Al,O, has not been noted previously, it is not
possible to relate the magnitude of the density change noted in Fig.
9.11 with the disappearance of coherent diffraction. However, the
noted decrease in density of Al2O3, about 18 percent, agrees in mag
nitude with the density decrease noted in Sect. 9.2 for a-quartz
under fast neutron bombardment, about 15 percent. In the latter
case, the density of the irradiated crystals was observed to decrease
almost to the value for vitreous silica. Thus, the structural and den
sity changes induced in a-Al2O, by fissioning are similar to those
noted in neutron-bombarded crystalline silica.
The extensive damage in a-Al2O3 caused by fissioning may be com
pared with that resulting from fast neutron bombardment reported
by Martin and Antal and Goland and plotted in Fig. 9.4 [17, 18].
The latter authors measured the concentration of defects by measuring
the scattering of long-wave neutrons and were able to correlate the
observed density decrease with the concentration of defects formed.
It is noteworthy that Antal and Goland found the defect concentration
to be experimentally far less than that predicted by theory, even
though the irradiations were conducted at temperatures of 20° to 30°
C [1]. Annealing of the defects was found to begin only at 400°C
(300° C in the work of Martin) and to require temperatures of about
1,800° C for complete annealing [18]. The swelling of Al,O, during
fissioning reported in Fig. 9.11 was observed at irradiation tempera
tures of over 400° C and, thus, represents a relatively stable disturb
ance. If an amount of energy is stored in Al.O., swelled by fissioning
analogous to that found by Childs for UAOs, about 17,000 cal/mole, it
can be estimated that, at the estimated saturation dosage of 6.3×10”
fissions/cc, about 1 percent of the fission energy has been expended in
lattice disruption, a number very close to the maximum energy which
can be expended in displacement reactions. It may, thus, be concluded
that lattice damage by fissioning in Al-O, and UAO, is a far more effi
cient process than that caused by fast neutron bombardment. It is
important to note, however, in Fig. 9.4 that the dimensional changes
under fast neutron bombardment have not reached saturation, and it
therefore, dangerous
is,

conclude that excessive structural damage


to

will not eventually result with continued fast neutron exposure.


The importance the swelling be
of on

temperature
of

by of

irradiation
Al2O,-UO, comparison
of

havior indicated these results with


is

those reported by Simnad and Wallace [46]. The latter investigators


reported no dimensional change specimens Al2O, plus
20
of

volume
to in

!ercent UO, irradiated 1,000° 1,600°F for burnups approaching


at

The low fissioning exposure which instability


of
at

1x 10° fissions/cc.
466 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
|
Al2O3–UO, takes place at low irradiation temperatures may, by anal
ogy with the UAOs system, predict a sensitivity of damage to tempera
ture of exposure. Evidence confirming this sensitivity to temperature
has been obtained from fission-damaged Al2O3–UO2. The specimens
shown in Fig. 9.11 were found to increase markedly in density and to
regain crystallinity of the Al2O, phase upon annealing at 1,000° C.
Berman, et al., noted in the course of their microexamination of the
specimens of Fig. 9.11 that microscopic pores initially present in
the as-sintered platelets appeared to close completely at the lowest
exposure level examined and also that macroscopic voids between the
oxide platelets and the metallic receptacle in which they were con

or It,
tained became completely filled with oxide [10]. thus, appears
that the irradiated material can exhibit liquidlike glasslike me
fact, this observation truly indicates the
In
properties.

if
chanical
body with low viscosity, this would constitute strong
of

formation
a

the X-ray diffraction results indicating the for


of

in
confirmation
glassy condition fission-fragment bombarded Al-Os.
of

in
mation
a

The implications the ability

on
of
of

the formation such structure

a
fuel matrix will be discussed
as

of Al2O3 later section.

in
to

act
a

a
(e) BeO

Another oxide matrix material which has been used to contain


UO, BeO exposed
BeO.
to

fissionable fast neutron bombardment


is

by

of
alone was found expand linear dimension maximum
in
of to

1
percent Changes lattice parameter were
in
at

dose 10° nvt [47].


or a

either nil 0.3 percent (Table 9.3) with

no
of

or of

the order
to

0.03
reported line-broadening change crystallite size [14, 15, 47].
in
In

containing volume percent


of

the case sintered bodies 0.6 and


3

UO, made from 200 mesh U.O, powder BeO matrix, however,
in
a

Gilbreath and Simpson found that neutron irradiation


to

the levels
Fig. 9.12 caused over times the linear expansion noted
10

shown
in

neutron-irradiated BeO,
or

density decrease percent


of
in

to

1.5
2
a

Comparison Fig. Fig.


of

9.11 with 9.12 indicates that


in

[14].
BeO-UO, Al2O3–UO, the major fraction the expansion
as

of
in

very low fissioning exposure levels. Included Fig. 9.12


at

occurs
in

are results obtained from elastic modulus measurements; elastic mod


uli are observed decrease upon irradiation, which would be ex
to

pected from the expansions noted. Again, the variation


of

modulus
change with fissioning indicates major damage early state
an
at

in

the exposure. No X-ray structural measurements have been reported


for this material after irradiation. However, such studies are re
Yeniscavich, Bettis Atomic Laboratory, unpublished
R.

M. Berman and W. Power


*

information.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 467
i.O i i H I i i I i I i T i

un O.9 H.
-
-
->
-
3 O.8 H
o
5 §
O
5 s O.7 H.
; : os
on un LINEAR ExPANSION

* T ".
uz -
ăg
#

ºf
O4 H. -
5
: 33
G
§
O.3 H

——
O
—g- ELASTIC MODULUS
O

*
3
5 oz FTs
O
-
§ O. H.
ExPANSION RANGE FOR NONFUELED BeO ExPOSED TO SAME FLUX
-
O
== C-E-F-----|--|--|--|--|--|--|--|--
O I 2 3 4 5 6 7 8 9 IO II 12 13 14
Fissions PER co, x ſolº

FIGURE 9.12. Expansion and Modulus Changes in Irradiated BeO-UO, [14].

ported at high fission depletions and indicate retention of structure


(Sect. 9.5.3). Furthermore, the density changes which are observed
are small compared with those noted in Al2O3–UO,. These facts
argue against a degree of structural damage in fissioned BeO-UO2 at
these exposure levels similar to that noted in Fig. 9.10 for fissioned
Al,Os-UO2. It is also significant that the structural damage by fast
neutrons in BeO is appreciably less than that suffered by Al2O3,
which again argues for a greater degree of structural stability of the
former matrix. It may, thus, be concluded that BeO is intermediate
in its behavior toward fission fragment damage between UO, and
materials such as Al,O, and UAOs. However, at fissioning depletions
100 times larger than those shown in Figs. 9.11 and 9.12, Yenisca
vich and Bleiberg found complete disappearance of diffraction peaks
in BeO-UO, bodies (at 12× 10° fissions/cc), marked swelling (25
percent or more), and high rates of fission gas release (see Sect. 9.4)
[48]. Thus, BeO must be considered intermediate in stability to fis
sion damage between UO, and Al2O, or UAOs.

(f) Other Orides


Additional of structural damage have been made for
observations
other oxide matrices containing dispersed or dissolved UO2. In
Fig. 9.10, the structure of a ZrSiO,-UO, mixture containing 25
percent natural UO, is shown to revert from that of the unirradiated
material (Fig. 9.10C) to a structure exhibiting no diffraction peaks for

57.4789 O–61–81
468 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
|
the ZrSiO, phase (Fig. 910D) after 2.4×10" fissions/cc exposure
(eBrman, et al.) [10]. In Sect. 9.2.2 it was shown that the ZrSiO,
structure is extensively damaged by exposure to fast neutron bom

is,
bardment. It thus, not surprising that complete structural dam

as
age occurs fissioning.

of
result These same authors found that

a
percent UO, which con

20
weight

of of
sintered solution ZrO2 and
a

sisted sintering
on major tetragonal ZrO2—UO, solid solution

a
phase and minor monoclinic solid solution phase showed, by X-ray
a
diffraction, only the tetragonal phase after exposure

to
7.5 10”

×
fissions/cc: the tetragonal phase after irradiation showed little trace
of

structural damage. The lattice parameters the tetragonal phase

of
before irradiation were ao-5.11A, co–5.21 and, after irradiation,

Å
Å,

of
ao-5.099 co-5.201 The direction and magnitude
Å. the shift
are consistent with the supposition that, upon irradiation, the ZrO.
rich monoclinic solid solution not only reverts the tetragonal form

to
but also dissolves homogeneously the tetragonal matrix [49]. Con
in
firmatory evidence the stability
of

of
the ZrO.-base fluorite structure
(note that the tetragonal ZrO2–UO, solid solution has slightly dis

a
by
torted cubic fluorite lattice) afforded the experiments reported
is
by

Yeniscavich and Bleiberg


on

sintered body ZrO2, weight

of

13
a

percent CaO, and weight percent UO, [48]. This material, con
17

sisting predominantly the tetragonal ZrO2–CaO-UO,


of

solution and

an
the compound CaZrO3, exposure
of

minor amounts

of
withstood
appreciable resulting dimensional
no

with
12

about fissions/cc
×

10%"
change. Similarly, Kittel and Paine could detect dimensional
no

changes the fluorite structure Tho.-base solid solutions with UO.


in

after 1.7 10° fissions/cc exposure [50].


×

Confirmatory evidence for the marked effect fission fragment


of
by

damage monoclinic ZrO2 afforded Wittels and Sherrill who,


in

is

repeating the work crystal


on

of

reversion structure ZrO2 from


in

in

monoclinic cubic form after neutron bombardment of natural


to
a

crystals reportedSect. 9.2.2(b) (Ref. 12), were unable dupli


to
in

cate the effect pure synthetic single crystals and powders even
in

at

fast neutron exposures four times higher than those originally em


ployed [51]. They concluded that their original observations resulted
crystals contaminated with uranium and thorium and,
of

from use
in

subsequent experimentation, found the following rates


of

conversion
fissioning exposure:
of

from monoclinic function


to

as

cubic
a

percent conversion 8X10" fissions/cc


at at at at
5

50 percent conversion 10" fissions/cc


× X
11 7

80 percent conversion 10° fissions/cc


100 percent conversion 34×10" fissions/cc

These observations indicate that about 10" atoms are affected by each
fission fragment, agreement with the results noted above the case
in

in
IRRADIATION EFFECTS IN URANIUM DIOXIDE 469

of U.O.s and Even further complication is introduced by


Al2O3.
Adam and Cox who showed that pure monoclinic ZrO2, irradiated with
fission fragments from UO, separated from the ZrO, by an aluminum
foil (0.71, thick) insufficiently thick to stop the fission fragments but
adequate to prevent evaporation by surface fissioning of UO, into
ZrO2, did not change in structure at exposures such that 100 percent
transformation is noted without the foil [52]. Thus, the phase trans
formation in ZrO2 requires the presence of impurities which can pro
vide nuclei of a cubic or tetragonal phase as well as (or possibly rather
than) exposure to fission fragment damage.

(g) Summary

It may be concludedfrom these observations that oxide matrices


decrease in stability to fission fragment damage in the following order.
Of maximum stability are the oxides with fluorite or fluorite-like struc
ture UO2, Th(O2, and ZrO, (tetragonal). Beryllium oxide is less
stable, Al,O, shows marked instability, at least at low temperatures,
and oxides such as U3Os and ZrSiO, are least resistant to damage.
The further inference may be drawn that fission fragment damage is
far more severe than that resulting from fast neutron bombardment.
The damage from the former occurs at exposures consistent with, or
even less than, theoretical estimates of numbers of displacements,
whereas it is generally observed that far less damage results in the
latter case than is theoretically anticipated [1]. In fact, the value of
the volume of UAO, disturbed by each fission fragment (2.7×10−" cc)
found by Childs and McGurn agrees with the total volume of material
within the fission track volume, indicating that ionization effects in
addition to displacements may cause structural damage in these mate
It,

rials [45]. thus, appears that stability fast neutron bombard


to

judge probable stability toward fission


as

gage
to

ment unsuitable
is

fragment damage.

9.3.2 Property Changes

(a) UO,

appli
all

Of property changes, the most significant with respect


to

conductivity
of

of

cation fissionable oxides the variation thermal


is

with exposure fissioning. An important series


of
to

measurements was
by

reported Ross, who determined the change thermal conductivity


in

UO,
percent dense sintered 60°C after various exposure levels
to 92
up of

at

low irradiation temperatures [53]. His


at

6.75
fissions/cc
×

101°

100 percent theoretical density are plotted Fig.


to

results corrected
in

exposure level. The thermal conductivity Satu


as

of

9.13 function
a
470 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I
|---- - -- --- - -- -- - - - -- - - -- -- - - - " " -
O.O65 I I I I I I i I

_-MEAN value for UNRRApared samples, sºap


# o.oso
-
3
Hº-
§ss
<
G. O.O55H. S$ N -
* w

-
: ooso H
$
§ *~ *
>
; ~3---4 ––8---4------- A-
F
oo<5H
SERIES FACILITY SYMBOLS §:
1958 SELF-SERVE A C
É
# ooºoh
O
1959
1959.
self-serve
J-ROD
a o
v D, v =
-

<. MAxIMUM VALUE OBTAINED taneo-T -
# ooss H MEAN VALUE
#
* MINIMUMvalue obTANED-
§
l l l l l l I l l
o.O3O
LIO* 5.1014LIO-5 5.1015Llo!6 5.1016LIO'7 5.IO'7 Liolò 5.1016Lloº

FISSIONS PER co

FIGURE 9.13. Variation of Thermal Conductivity of UO, at 60° C with


Burnup [53].

rated at a value 30 percent less than that for unirradiated UO, at a


fission density of less than 5×10" fissions/cc and remained unchanged
to exposures of 6.7 × 10° fissions/cc.
By comparison of the irradiation level at which the effects on ther
mal conductivity saturate with the structural damage information
listed in Table 9.5, it is apparent that saturation of both thermal con
ductivity and lattice parameter change has occurred at 5 × 10" fis
sions/cc. The inference may, thus, be drawn that a similar type of
disturbance has produced both effects, and it is tempting to identify
this as point defects produced by atomic displacements. The results
presented by Ross on recovery of the irradiation-induced decrease in
conductivity argue against so simple an interpretation, however. As
shown in Fig. 9.14, the reduction of thermal conductivity induced
by irradiation at exposures to 2×10" fissions/cc is completely recov
ered at annealing temperatures below 1,000° C in 1-hour annealing.
However, with irradiation exposures over 10° fissions/cc, the recovery
is only about 50 percent complete at 1,000° C and below, and further
more, the shape of the curves indicates a tendency to level off at such
values, even though, as shown in Fig. 9.13, the thermal conductivity
was unaffected by continuance of the exposure beyond 10" fissions/cc.
This behavior was interpreted by Ross to indicate the formation, with
continued irradiation, of more complex arrays of defects which have
little additional effect on room temperature conductivity than simple
:
IRRADIATION EFFECTS IN URANIUM DIOXIDE 471
IOO
*--Tº:
^ =
90H
~ 6.25.10°f/cc -

8OH 23.10°t/cc Z -

L 70H
S6 A& 74.107 f/cc -
fO REcovery FACTOR R- I
or
loo x =K_ANNEALED=K_RRADAIED
-
K UNIRRADIATED K IRRADIATED •* -
6O |-
#.
-
o:
A2
-
§ 50H
S|| -
§u- ALI
7:55.0°t/cc , PLUS PREVIOUS
>
ºr 40H Z ANNEALS AT
-
-
B2 6OO*AND 7OO°C
§
§ S7

* 30H – L7
ANNEAL TIME I HOUR
UNLESS OTHERWISE
-
18 SHOWN

2O |
I.O.IO” f/c c
6.8.lo"t/cc -

ºl-T*
SII
30 MINUTE ANNEAL

IO
AT 7OO’C PLUS
PREVIOUS ANNEAL
-
AT 600°C

O L *
-* sII
I L' I I I
400 500 6OO 7OO 800 900 IOOO |IOO
ANNEALING TEMPERATURE, "c

FIGURE 9.14. Recovery of Irradiation-Induced Decrease in Thermal Conductivity


of UO2 upon Annealing [53].

point defects, but which are more resistant to annealing. Thus, at


exposures up to 5×10" fissions/cc defects which anneal in 1 hour at
400° C are predominant, while at higher exposures defects which
anneal at 400° to 1,000° C and at 1,000° C are considered to form.
The continual increase in strain line-broadening in Table 9.5 with
increasing exposure may confirm the presumption of formation of
such complex defect arrays. Note that Antal and Goland in their
study of defect concentrations in fast neutron irradiated Al2O, also
found it necessary to assume the formation of defect clusters to ex
plain their experimental results [17].
Berg, Flinta, and Seltorp found a similar insensitivity of the ther
mal conductivity of UO, to irradiation in an in-pile experiment in
which central and surface temperatures of 95 percent dense sintered
pellets were measured while the pellets were subjected to neutron irra
diation at a surface temperature of about 300°C, and the neutron flux
was estimated from analysis of flux monitors adjacent to the speci
mens [54]. No change in thermal conductivity from
its

initial value
472 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
5.5 I-I-I-T-I-T-I-T-T—T-T—T-T—T-T—H
- - - BEs.TFIT IRRADIATED
- o “Nº
- Paºlºnsolated
- CONDUCTIVITYDATA -
§ K. Hºa (95% DENsity)
K= 43.8
; a SCOTT DATA-UNIRRADATED
H.;
(T+4oo)+344(438)
# so H
i. K. (95%.DENsity) where 344 #F#
#3; - THERMALRESISTANCE
§§§; D wapD 22-11ExPERIMENTAL DUETO RRADATION
É ; DATA(ZERO DIAMETRAL
CLEARANCE,
T=*c in ALL EQUATions
Kr 4 xe ATMOSPHERE,---THERMAL CONDUCTIviTY
* *|º
H- 45
S-,
-
PRODUCTION UO2)MTR RRADATION AFTER ExPOSUREOF
-
*
8.7xlo” Fissions/cc
A 13 ×lo” Fissions, cc -
ſ
tº so H —
^-
*->
H.
| -
co
;: 35 H
-
t>
Hº- - -
§3 so H -
O
-1 - -
#
A -
#
E 25 H - *~*
§
l 68 - A 193 A
:u - Az- A -
- w”
* 2o I i Zia I I I l l l L l 1–1–1–1 l
O IOO 2OO 3OO 400 500 6OO 7OO Boo sco
U02 CENTERTEMPERATURE,
*c

FIGURE 9.15. Average Thermal Conductivity of UO, vs Center Fuel Tem


perature [55].

was noted as exposure continued to 3×10" fissions/cc; the low initial


value of conductivity noted by these authors, 0.02 w/cm-º C, is prob
ably not related to conduction through the solid but to the effect of
clearances between the UO, and its cladding, as will be discussed in
Sect. 9.7.2.
Cohen, et al., measured the thermal conductivity of UO, during
in-pile exposure utilizing a technique whereby the central temperature
of 95 percent dense pellets
is measured and the heat flux is estimated
from the temperature drop and conductivity in a thick-walled stain
less-steel tube surrounding the oxide [55]. This experiment was de
signed primarily to measure the effects of initial fuel-cladding gaps.
internal atmosphere, etc. However, one of the experiments, the re
sults of which are shown in Fig. 9.15, was assembled with no
diametral clearance between the stainless steel sheath and the pellets
to minimize the effects of temperature drops at the interface. The
information is considered applicable in the present connection if it is
borne in mind that some uncertainty, which would act to depress the
calculated thermal conductivity, may exist with respect to the thermal
resistance of the metal-oxide contact. Since the measurement yields
an average conductivity of the oxide between central and surface tem
peratures, the experimental data have been plotted as a function of
central temperature and compared with the measurements of Kingery,
IRRADIATION EFFECTS IN URANIUM DIOXIDE 473

et al., and Scott on unirradiated UO, (see Chap. 5) recalculated for


the same central temperature and heat production as in the in-pile
experiment [56, 57]. It is apparent that over a temperature range up
to 730°C the reductions in conductivity over the unirradiated values
obtained by Kingery, et al., are even less than the 30 percent reduction
found by Ross [53, 56]. Repetition of these measurements after ex
posures of 8.7 × 10* and 1.3×10° fissions/cc showed considerable dis
agreement with the original experimental data at low temperatures, the
average conductivities at 100° C being reduced to 50 percent of their
original value; this effect is presumed to result from cracking of the
pellets and will be discussed in a later section. However, at technically
significant temperatures of 400° C and above, the values for average
conductivity agreed well with the original measurements for both
levels of exposure. Data have been obtained which are similar to those
plotted in Fig. 9.15 for 1 to 3 × 10° fissions/cc to exposures of 5× 10”
fissions/cc, or about seven times the highest level attained by Ross,
with no significant change apparent in the conductivity.
If the effect of irradiation upon thermal conductivity be represented
by the addition of a temperature-independent thermal resistance term
to that of the unirradiated material, the irradiated thermal con
ductivity attemperature T is then
- (KT) unirradiated
(Kr) irradiated T 1 + Wirradiated X
(AT) unirradiated
where Wirradiated is the thermal resistance due to irradiation and the
other terms are obvious [58]. Using the data of Kingery, et al., to
represent the unirradiated values, and matching the curves at 500°C,
Cohen, et al., estimate War, to be 3.44 (w/cm – “C)- (see Fig. 9.15).
A similar estimate for Ross's data in Fig. 9.11 yields a value for
Wºr, −5.6 (w/cm – “C)-'. A number of comparisons can be made to
indicate that such a level of irradiation-induced thermal resistance
is reasonable for the amount and type of damage postulated to exist
in fissioned crystalline UO. Thus, from Fig. 9.13 and Table 9.1,
Wirr is seen to reach its saturation level at 1 × 10" fissions/cc or
9 × 10−" fraction of the atoms displaced. In Al-O, a fast neutron bom
bardment of 5× 101* nvt or, from Fig. 9.4, 5 × 10-' fraction defects
formed, causes (from Fig. 9.5) a thermal resistance of about 0.2
-
(w/cm "C)-'. Thus, 18 times this concentration of defects would be
expected to increase the thermal increase to 3.6 (w/cm –
“C)-', which
is of the correct order of magnitude if the information for Al-O, is at
all directly applicable. It was estimated above, from consideration of
the lattice strain results in Table 9.5, that the reduction in structure
sensitive properties, such as thermal conductivity, may be of the order
of that introduced by the alloying of 7 percent Tho, with UO.
474 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Unfortunately, information on thermal conductivity for such com


positions is not available. However, Kingery has found the thermal
conductivity of Tho, at 100° C (0.20 w/cm-9C) to be reduced
to 0.05 w/cm- “C by the solution of 26 mole percent UO, or a
thermal resistance increase of 15 (w/cm-9C)-" [59]. Again, cor
recting linearly for a 7-mole percent addition and assuming the
thermal resistance increase for Tho, additions to UO, is equal to that
for UO, additions to Tho, a thermal resistance of 4.0 (w/cm-*C)
is calculated, in agreement with the observations. As a final indica
tion, Shapiro and Powers found that the thermal conductivity of
UO, is 33 percent greater than that for UO, to at 422°C, a tempera
ture at which essentially all the oxygen in UO, to should be in solution
(see Chaps. 5 and 6) [60]. Using Kingery, et al., values for the
thermal conductivity of UO2, it is estimated that 3.3 percent addition
of defects in the form of oxygen ions adds a thermal resistance of 5.9
(w/cm-* C)-", Again, correcting for a defect concentration of
9×10−9, a value of 1.6 (w/cm – “C)- is obtained. Thus, these com
parisons, while certainly not proof, lend confidence that the thermal
conductivity and structural damage behavior of irradiated crystalline
UO, at exposures up to at least 5 × 101° fissions/cc are satisfactorily
reconciled by assuming the formation of about 1 percent displaced
atoms, a number which saturates at 1 to 5 × 101° fissions/cc.
Few other property measurements of irradiated UO, have been
reported. Padden has measured the hardening by irradiation of 97
percent dense UO, fissioned to an exposure of 2× 10° fissions/cc at
450° C. Using a Knoop indenter and 500 g load, he found the pre
irradiation hardness of 625 (see Chap. 5) to increase to 830. This type
of measurement may yield valuable information on the effects of an
nealing, burnup, operation temperature, etc.; further observations
using such techniques should be made. This observation of hardening
under irradiation indicates retention of the strong bonding of the
compound and is consistent with the mechanism of damage postulated
for UO, at exposures up to 5 × 10° fissions/cc, namely the formation of
a limited number of point defects saturating at short exposures.

(b) BeO-UO,

A
comprehensive series of measurements was reported by Gilbreath
and Simpson on property changes in BeO containing 2 weight percent
UO. of densities 90 and 92.5 percent of theoretical and also in BeO con
taining 10 weight percent UO, of overall densities 88 and 93 percent of
theoretical The irradiation temperature of the former speci
[14].
mens was estimated to be 250° C, and of the latter, 650° to 700° C.

* T. R. Padden, Bettis Atomic Power Laboratory, unpublished work.


IRRADIATION EFFECTS IN URANIUM DIOXIDE 475

INTEGRATED NEUTRON Exposure x toº nwt


O 2 4 6 8 IO 12 14

.
I I —I I I T

7 H
-
e
BeO-2 % UO2 . . DENSITY, p = 2.85 g/cc (93% T.D.)
2%
- Be O-IO% UO2
i 6
p =3.01 g/cc (93% T.D.)
i
*-*
.º ſ al

_T
x:
>
E
5 H- p=2.78 g/cc (91% T.D.)
-
=
wo

É
A p =2.85 g/cc (88%. T.D.)

-1 A
* FR s
#
5 A
+\
— à
E []
Liu
> - A -
= 3
s k unlrradiATED
FUELED Be O
# K IRRAdiATED

2 H (:
K. UNIRRADIATED
IRRADIATED )
Be O -

_-T
p =2.9 g/cc (9.6% T.D.)

o--e. 2.7 g/cc


thox
8
T.D.)
l
Be O(PLOTTED
|
- ?
vS nw?)
I
8–

O 2 4 6 8 IO 12 |4
Fissions PER ce x lot"

FIGURE 9.16. Effect of Irradiation on Thermal Conductivity of Fueled and


Unfueled BeO [14].

The property changes were compared with those of unfueled BeO


samples of 90 and 96 percent theoretical density exposed to the same
integrated neutron flux. Changes in thermal conductivity after irra
diation were measured using a comparison technique with an unirra
diated sample of the same type as a comparison standard. This
technique measures ratios of thermal diffusivities and, if it is assumed
that density and specific heat are unaffected by the exposure, will, thus,
yield a measure of the relative conductivities before and after irradia
-
tion.
The obtained are plotted in Fig. 9.16 as a function of
results
fissions/cc of material. For comparison, measurements of conduc
tivity changes in unfueled BeO samples are also plotted as a function
of neutron exposure. The thermal conductivity of the fueled samples
was observed to decrease drastically upon exposure to fissioning to
values 15 to 20 percent of that of the unirradiated samples, whereas
that of the unfueled samples decreased, at most, to 60 percent of the
476 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

unirradiated conductivity. It will be noted that the conductivity


change plotted versus fissions/cc exposure appears to be different for
the two fuel loadings, the changes in the 10 percent UO, samples being
less drastic than those in the 2 percent UO, specimens. This effect
Gilbreath and Simpson ascribed to the higher irradiation temperature
of the former. However, it should be noted that the fissioning expo
sures, while greatly different for the two types of samples, were ob
tained at the same fast neutron exposure level and that the conduc
tivity of unfueled BeO changes appreciably as a result of this fast
neutron exposure level. Therefore, if the effect of fast neutron bom
bardment is considered to be additive to that of fission fragment bom
bardment, the effect of the latter alone can be reported by comparing
the thermal conductivities of the fueled BeO with that unfueled at the
same neutron exposure level. This correlation is plotted as the dotted
line in Fig. 9.16 and it is apparent within the scatter of the experi
mental data, that in agreement with the results for UO, plotted in
Figs. 9.13 and 9.15, no effect of fission fragment damage on conduc
tivity of BeO can be detected beyond the initial experimental observa
tion at about 1 × 10° fissions/cc exposure, at least to a level of 1.4× 10”
fissions/cc.
Extrapolating the experimental measurements reported by Gilbreath
and Simpson on the thermal conductivity of BeO to 100°C, and utili
zing the information plotted in Fig. 9.16 indicating a 3.8 fold increase
in thermal resistivity upon irradiation due to fission product damage
alone, there may be derived in the same way as indicated above for ir
radiated UO, a thermal conductivity at 100° C of fission-damaged BeO
of 0.26 w/cm=9C compared with 1 w/cm–"C for unirradiated
BeO and for thermal resistivity increase due to irradiation,
Wº’"–2.8 (w/cm—*C)-1 The latter is in sufficiently good agree
[14].
ment with the estimates made previously for UO, and Al2O, to lend
some confidence to the interpretation of fissioning damage previously
presented. While it is not to be expected that fissioning effects in
materials so dissimilar as BeO, Al...O., and UO, should lead quantita
tively to the same property changes, the materials are sufficiently
similar in bond strength (see Table 9.4), melting point, etc., that the
qualitative agreement discussed above may be anticipated so long as
the damage produced is not so extensive as to lead to general lattice
disruption.
Correlation of the data in Fig. 9.16 by canceling out the neutron
bombardment effects on BeO alone implies, as discussed above, that
such effects are additive to those produced by fission fragment bom
bardment and, consequently, that the defect structures produced by the
two types of exposures are different, at least at the fissioning exposures
investigated. A similar conclusion was drawn in the case of UO, with
IRRADIATION EFFECTS IN URANIUM DIOXIDE 477

respect to influence of fission exposure level from observations of re


covery of conductivity during annealing and is substantiated in the
case of BeO by the annealing experiments reported by Gilbreath and
Simpson [14]. Thus, the neutron-bombardment induced suppression
of conductivity in BeO is restored by annealing 1.5 hours at 900°C,
whereas BeO a large fraction of the damage re
in fission-damaged
mained after annealing at temperatures as high as 1200° C.
Gilbreath and Simpson found the compressive strength of fission

its
damaged BeO to be depressed to about 70 percent of unirradiated
value, whereas that neutron-damaged BeO rose about 110 percent
of

to
of

its unirradiated value [14]. These results are difficult

to
reconcile
with the hardening observed UO, and may reflect
on ir

of
irradiation
strength.

of
radiation embrittlement rather than true loss

(c) Summary

Fission fragment damage metal oxides, thus, appears many


in

in
respects conform better predictions than does fast
to

to

theoretical
quantitative explanation

at
neutron damage. the damage level
of
A

which fission fragment bombardment saturates not available. In


is
on

formation the defect structure produced upon prolonged exposures


UO,
be

will ex
of

Sect. 9.5. The behavior at

at
discussed least
in

posures up fissions/cc, appears explicable


on

the basis that


to

10°
×
5

percent displaced atoms

to
of

saturation level
at
about attained
is
5 a

1
10" fissions/cc; both structural and property changes yield results
×

agreement with such postulate. fission fragment


of

The effects
in

damage
of

BeO are similar UO2 that


to

to

those noted the case


in

in

relatively
of

saturation level permanent displacements


at

attained
is

low exposures.

9.4 FISSION PRODUCT RELEASE FROM UO,


UO,
of

of

of

One the characteristic features the irradiation behavior


In

fission products.
of

of

its containment and release the case metal


is

fuels, fission products are, general, released from external surfaces


in

only
to

the extent predicted from recoil from the surface. The fission
products remaining the metal cause embrittlement, growth, and
in

plastic distortion the metal matrix, and, critical temperature


of

at
a

which appears highly structure sensitive and differs from one


be
to

metal fuel another, the volatile fission products agglomerate creating


to

gross voids which can lead exaggerated swelling the body. Re


of
to

extents occurs only


of

volatile fission products appreciable


to

lease
with swelling physical damage
of

concomitant extensive and the


a

body. UO,
In

however, products
be
of

the case fission can retained


478 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

within the UO, structure to concentrations at least up to 3 percent of


the metal atoms fissioned at high temperatures with no observable ten
dency to cause either homogeneous lattice expansion or gross dimen
sional distortion [61]. On the other hand, and possibly as a conse
quence of this ability to retain fission products within the lattice, the
volatile fission products are released from external surfaces at rates
which vary with temperature, relative surface area exposed, etc., in a
definable manner. This section is concerned with the experimental
observations on release of fission products from UO, and with the cor
relations which have been proposed to explain the observations. The
effects of exposure beyond 7 percent of the uranium atoms in UO.
fissioned are described in Sect. 9.5.

9.4.1 Identity of Fission Products

The fission products most important with respect to their effect on


stability of reactor fuel elements are the stable fission products. For
this purpose, the tabulation of Katcoff has been used as reference [62].
In Table 9.6 is tabulated for the case of the thermal-neutron fission of
U* the distribution of yields of stable nuclides or of those whose decay
is so slow that their concentration changes inappreciably during con
ceivable fuel element lifetimes. In their effect on external systems
after their from the oxide body or for use as tracers for more
release
stable isotopes, the short-lived nuclides are most important. The
nuclear reactions in which some of the most important of these are in
volved are also listed in Table 9.6.

TABLE 9.6—YIELD OF STABLE AND LONG-LIVED NUCLIDES FROM


THERMAL NEUTRON FISSION OF U"* [62]

Element Atomic weight of nuclide Percent Element Atomic weight of nuclide Percent
yield yield

Ge--------- 72,73 0.0001.26


|| Sb---------- 121,123 0.0163
Se---------- 77,78,79 0.0853 Te---------- 125,126,128,130 2.441
Br---------- 81 0.14 I----------- 127,129| 1.03
Kr--------- 83,84,85,86 3.857 Xe--------- 131,132,134,136 21.83
Rb--------- 85,87 3.50 Cs---------. 133,135,137| 19,15
Sr---------- 88,90 9.34 Ba--------- 138 5.74
Y---------- 89 4.79 La---------- 139 6.55
Zr---------- 97,92,93,94,96 ||31.05 Ce---------- 140,142 12.39
Mo--------- 95,97,98,100 24,44 Pr---------- 141 6.0
Tc---------- 99 || 6.06 Nd- - - - -- - -- 143,144,145,146,148,
150 21.04
Ru--------. 101,102,104|| 10.9 Sm--------- 147,149,151,152,154|| 4.322
Rh--- - - - - - - 103|| 3.0 Eu. -- - -- - - - 153|| 0.15
Pol -- - - - - - - - 105,106,107 1.47 Gd - - - - - - - - - 155,156,157 0.0538
Ag--------- 109|| 0.030 Col - - - - - - - - 158 || 0.002
Col. -------. 111,112|| 0.029 Th - - - - - - - - 159 0.00107
In---------- 115 0.0104 Dy-------- 161 0.000075
Sn---------- 117| 0, 011
16,

78
ff,
4,
p.

"Reprinted from Nucleonics, Vol. No. April 1958,Copyright 1958,McGraw-Hill Publishing


Co., Inc.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 479
78m Krº-Hºn
0.04 2.
/
15.5S

".
0.96

ºr
Yu
-ºsn Rbº-stable Sr*
3.57

2.8h Krºs-Hn

/ 2.
4.5S
prº 16zYºº m
0.0002
Y.
3.2m Krº–15.4m Rb"—"| 514 Srs,
4.50 4.79
0.
Stable Yºº
3.2m Krs?--n

/ 2.
1.4S

~ Y.
33S Krº->2.7m Rb%—- 28ySr 30 — 64.3hYº->Stable Zrwº
5.0 5.77

51m Ywi m
1.67mRbºl 2. N
/
-
2. 0.6/ N30.015
N
10sKrst/ N9.7h Srºl | NStable Zrºl

//
3.45 N A 5.84
— N
Y.
7' 5.81
0.4Y.
!
14m Ribºl 58dYºl
~5.4

90hNbººm
0.02 |

** º

0.98 l /
7'
Stable Moº
5.27

35dNb%
6h Tcºm

-
0.87_*
30s Zr"—-3m Nbº—' 66hMow | Stable Ruº?
6.06
0.13N. l /
2.1×109 Tcº

`
30h Telºlm 12dXelil"

T -
10.1570.44. N-0.80 0.008 |
| !
3.4m Sntil—-23m Sb.131/

0.85N
24m #
0.992 l
S. Stable


2.9xe"

3.2mSnim—2.1m Sbin—'ſ 77hTeun


4.7
23h
*—sºxer 4.

63mTelzim 2.3dXely, ".


oyº 0.024

/ *,
4.1m shutº - .8h I113 Stable Caiº

Q
0.28N l
2
0.976 S. 5.27dxens
7' 5.59.

2m Telºs 6.62
480 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONs
Stable Xelsº-Hºn

248Iist/
/ 2.

N 2.57m Ba137
m

-
0.92
Yu
3.9m Xen->| 29, csis,
6.15
S.
-
0.08 S.
Stable Ba137
3.9m Xels"—Hºn

5.8S Ilas/
/
N
– ––

32.2mCs138 Stable Ba138
| Fºxen 5,74

lar
aſsin–assºn—
- ºncº sººn-stable
6.55

16S Xel"—66S
3.8
Cs”— 12.8dBaiº |→40.2h Laº—Stable Cel"
6.32 6.44
6.32
(6.44) (6.44)

285dCeia
6.0
— 17.4m Priº—5X101W Ndiº
5.67

As can Table 9.6, the stable fission prod


be

of

from examination
seen
predominantly metal isotopes which would

be
of

ucts consist expected


exist substitutionally interstitially
or

to as
to

in
cations the oxide lattice
or, sufficiently high concentrations, precipitate
at

as
condensed
phases. the products, consisting Se, Te,
of

of
much smaller fraction
A

Br, anion impurities substitutionally


I,

expected
be

to

as

would exist
interstitially, precipitate high concentrations
or

sufficiently
to

at

as
to or

condensed phases compounds


with the matrix cations fission or
or

formed cations, evaporate from the oxide surface. The third


fission products, the most important
ir
of

oxide fuel element


to

class
radiation behavior, consists the noble gases xenon and krypton
of

which are formed amounts corresponding


to

about 0.25 noble gas


in

atom per U* nucleus fissioned. (Note, however, that


as

pointed out
Xe”, this ratio can increase
to by

by

Lewis, because neutron capture


of

0.3 noble gas atoms per U* nucleus fissioned high flux exposures
at

[63].) Their method incorporation


to of

the oxide lattice unknown:


in

is

however, their ability the oxide indicates,


of
at

escape free surfaces


expected, that they are not tightly bound
as

be

would the lattice.


in

crystalline UO, these atoms not precipi


do

The fact that, least


in
by at

tate internally gas bubbles has already been discussed.


of

formation
Thus, two methods postu
be
of

their escape from the oxide lattice can


by

lated. The first involves escape recoil from those atoms fissioned
within the recoil range the solid. This mechanism most import
in

is

or

performed temperatures
of

irradiations
in

ant low
in
in at

case case
fine particles such dispersion fuels. Note that
of

of

irradiation
as
by

this mechanism independent


of

the instantaneous release rate


is
IRRADIATION EFFECTS IN URANIUM DIOXIDE 481

the nature of the isotope or of the irradiation temperature and depends


only on fission yield and recoil range the ex

on
its
the solid and

in
by solid state dif

of
ternal surface area. The second method escape

is
by
fusion; the instantaneous rate

be
escape this mechanism may

of
anticipated

be
the isotope, total

of
to

to
sensitive the chemical nature
the solid, concentration the isotope the solid,

of

of
free surface area

in
etc., and most important for the case bulk oxides operating ele

of

at
is

by
vated temperatures. The systematics

of

of
escape each these
methods reviewed below.
is

9.4.2 Equations for Fission Product Release

The nomenclature the most important symbols


of

Sect. 9.4

in

is
defined below:

N. isotope external
to
of

of
= == = = =

Number atoms solid


i

[N, isotope external

to
of

of

Concentration atoms solid


)

Time (sec)
f

Geometric surface area (cm?)


l., ,l, S

Recoil length isotope


of of

fuel (cm)
in in
,

i i

Recoil length isotope fuel-clad gap medium (cm)


,

Fissioning rate per unit volume (fissions/sec-cm”)


-=
f Y,

Fission yield isotope


of

(atoms/fission)
i

sphere (cm)
of
= = = =

Radius
h a

Clearance between fuel surface and cladding (cm)


N,.. isotope per unit volume t=t (atoms/cc)
at at
of of

of of

of of

Number atoms solid


i

N. Number atoms isotope per unit volume solid t=0 (atoms/cc)


a

- Radioactive decay constant for isotope (secT')


i
}

Diffusion coefficient for isotope (cm3/sec)


i
=

D.
P.

(secº-1)

R, isotope
of

of

Rate (atom/sec)
+

release
i
F,

isotope escaping solid


of

Fraction
= =- - = - -

Escape rate coefficient (sect')


y,

F. fission products released from fuel element


of

Fraction
e, :

D, Composite diffusion coefficient fuel element (cm3/sec)


in
o,

Fission cross section (barns)


Perturbed neutron flux (n/cm2-sec)
q

!",
of of

Volume fuel (cc)


Volume system (cc)
W

= =

equivalent time full power irra


at

to
of

of

Ratio total time from start


t,
r

B - diation
system (sect')
of

Purification and leak rate constant


A, Specific isotopic loop activity =X,[N.] (dis/sec-ml)
in
-

by

(a) Factors Affecting Fission Product Release Recoils

Frank, Ref. 64, derived the following expression for the


in

(dN.)
number isotope i(N) escaping per second by re
of

of

atoms
a
di

coil mechanism body geometric surface area


of

from cm” whose


S
a
482 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

thickness dimension is large compared with the recoil length of the


isotope in the fuel,

ly,
cm:

dN_f YıSlf. Eq. (9.14A)


dt

Y,
where the fissioning rate per unit volume, fissions/sec-ce, and
is
f
the fission yield the isotope, atoms/fission.

a of
is

For the case small spherical particle where the recoil range

of
the particle radius (this par
an

may appreciable fraction


be

of

is
case
ticularly important for UO, used matrix),

as
dispersed phase

in
a

a
the above expression becomes

º–fy. (1,4-#

3
f
Eq. (9.14B)
3.
#,

for the sphere [65].


of
where the radius
is
a

a
In

the usual applicationfuel element, the oxide body con


of

is
a

by

or
tained within sheath from which separated gap clearance
is
it
a

a
which the recoil length
If

this gap
of of

filled with medium


in
cm.
is
h

the isotope h/l., 21, which recoiled fis


at
if

7.4 cm and the rate


is

sion fragments come the gap given by:


to

rest
in

is

*}=ſ** (M.,(2-hl.).
Eq.
(8.14c,

UO,
as

The range fission fragments plotted


of

of
function
in

is

Fig. Thus, given geometry the system, the con


of
in

mass 9.17. the


fission products
of

of

tribution this effect


to

release calculable.
is

IO
i
I

I
I

H

8

ºw - -
-
# SºE

H
6

-
§

z
|

-
<ſ
or
H
5_
4

O
ul
ar
- -
H -
2

-

!

I
O

I
l

l
I

70 80 90 |OO |||O I2O I30 14O


MASS OF FISSION FRAGMENT

Figure Recoil Range (), vs Mass Fragments [64].


of

Fission
in

9.17.
U
IRRADIATION EFFECTS IN URANIUM DIOXIDE 483

Note that these relations are derived purely from geometric con
siderations and do not take into account effects which might arise from
passage of the recoil fragment through the oxide lattice. From con
sideration of the path of the fission fragment as a thermal spike the
momentary existence of high temperatures in an appreciable volume
about the path of the fission fragment can be calculated, and evapora
tion of appreciable quantities of volatile fission fragments or even of
oxide molecules from the heated volume of those fission spikes which
intersect the free surface of the body is to be anticipated [5]. In fact,
Laptev and Ershler have measured the evaporation of 24 U" atoms
per fission in the recoil range of oxidized metallic U* samples [66].
It would, thus, be expected that at least that number of volatile fission
fragment atoms associated with each evaporated metal atom would
also evaporate. As will be discussed in greater detail later, sintered
UOz bodies possess measurable open porosity even when sintered to
almost theoretical densities. While it would be anticipated that the
geometrical surface area of the body is the correct quantity to be uti
lized in Eqs. 9.14 A, B, and C to account for the disposition of
the recoiled fission fragment, the total free surface would be active in
yielding fission fragments volatilized by recoils. Thus, if it is assumed
that those volatile atoms associated with 24 uranium atoms are evapo
rated at all the open surfaces, the rate of evaporation will be given by

dN_24Zf Sºl,

where Z is the number of noble gas atoms per uranium atom and S'
is total surface area. In a sintered UO, body of 90 percent theoreti
cal density in which, as will be shown in Sect. 9.4.3, the actual open
surface area may be 1,000 times as large as the geometrical area, it
may readily be shown that, at a burnup of 1 percent of the uranium
atoms, the contribution to release of volatile fission fragments by evap
oration from the recoil path may be at least 250 times as large as
that from direct recoils. Furthermore, because of the high tempera
tures of operation of UO, fuel elements and the high vapor pressure
of UO2 at its melting point, the amount of evaporation should increase
rapidly with operating temperature and, hence, a pronounced tem
perature sensitivity should be noted.
Walton considered, in attempting to reconcile continuous measure
ments of fission gas release from uranium in-pile with the much
lower values obtained in out-of-pile experiments, not only the evap
oration effect discussed above but also the driving force of thermal
gradient diffusion within the thermal spike volume [67]. This effect,
which should also be operable in UO2, should again tend to increase the
effectiveness of recoils at the surface in enhancing volatile fission prod

57.4789 O–61—32
484 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

uct release. Again, a temperature coefficient of release is predicted


which should vary directly as the volume of material raised to a tem
perature above the melting point in the spike or inversely as the square
of the difference between ambient and melting temperature.
Finally, there will be discussed further in Sect. 9.4.9(c) the interest
ing concept of Lewis that fission fragment recoils can serve to limit the
pressure of released fission gas developed within the confines of a fuel
element sheath by collisions between the fragments and void-filling
gases which can inject the latter back into the lattice of the oxide [63].
Thus, many ramifications of the recoil process are conceivable. Until
more accurate experimental data are available to compare total release
rates with those arising from diffusion phenomena alone, the import
ance of these concepts is difficult to assess. Fortunately, as shown
below, measurements of the relative release rates of isotopes of differ
ent half-lives permit distinguishing between recoil and diffusion
mechanisms in in-pile experiments, and postirradiation measurements
are less confounded by the complications introduced by fission recoils.
It is of interest to estimate the relative contributions of recoil and
diffusion phenomena on the rate of release of a stable fission product
isotope. From Eq. 9.14A the release of a stable isotope of, for
example, xenon from the geometric surface of a 0.357-inch diameter,
95 percent dense cylindrical UO, pellet by recoil amounts to about
0.085 percent of that generated. Such a pellet operated at a central
temperature of 2,200° C and a surface temperature of 400° C would
release by diffusion in 10,000 hours about 8 percent of the xenon
generated. Thus, release by recoil from the geometric surface of UO.
becomes relatively more important the higher the density of the fuel
and the lower its operating temperature.

(b) Mathematical Description of Fission Fragment Escape by


Diffusion,

The diffusion phenomena of interest in the present connection differ


in two major details from those ordinarily investigated: (1) in many
cases of interest the isotope under consideration is decaying radio
actively at a rate comparable to the rate of diffusion; (2) for the
case of greatest technological interest, that of in-pile irradiation, the
isotope is being generated within the body at a rate which, in the
following discussion, is considered to be constant with time. Solu
tions to the diffusion equation applicable to this situation have been
given by various authors and are collected here for reference.

(1) DIFFUSION DURING IRRADIATION. The basic diffusion equation


to be solved is
ON,
Ot
=DV*N,+f Y-X, N. Eq.(9.16)


v
~~ __ ==--~~~~

__
p.485) (Face 574789–61
IRRADIATION EFFECTS IN URANIUM DIOXIDE 485

where N, is the number of atoms of isotope i per unit volume,

-
atoms/cc, Ai is the decay constant for the isotope i, sec−", D, is the
diffusion constant for the isotope i, cm”/sec, and the other terms
have been defined above. The applicable initial and boundary con
ditions are: M, (a = a, t = t) =0, M. (w = a, t = 0) = 0. For the case
of a sphere of radius a, Frank showed that

- sinh./* -- ( X, -

--
#5; cosh r
i 1—e
Y ºt i

...)
N=4ra”f
–-tºn w
M
4
v

J
t
i

M
r

£y 1– e-nº-'Dº
2

Eq. (9.17)
A

*
tt

-za m*(nºr’Dſ--X)
e

N, isotope sphere
of

the total number

to
of

where atoms external


is

-
D=; sect',
D
-

,
at

radius cm time sec, and the other terms are


of

this function have been derived by


of

defined above [64]. Values


Table 9.7 (68]. The rate isotope
of

of
Beck and are tabulated release
in

by
R.
as

defined (atoms/sec), given


is
i,

TN, ,-º-ºp.
...

}-º-2- "nº-fi,
f

R,-4ra‘f
*—.
Y,

coth >] Eq. (9.18)


º

i
t

The corresponding flat plate


of

solution for thickness cm and


a

surface area cm” was developed by Litz [69]:


S

N=
*Nº annºy;-2,
G-cº tanh.
(1—e D, ())".
D.)
2a.
\;

"
')
M

1—e Eq. (9.19)



486 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

This equation was simplified for the following cases to the form:

N-ſºº,
fy, S JD,
P.

|})
Or At-10,
,
a X Eq.

q.
For Ait 2 (9.19A
D,

)
tanh(.

‘T
X,

V
X.
For
av}<0. and Att- 0.01; Nº fy.S. D.

M
M
f

X,

X,
‘(;)
N3/2
1–6e-Mi') -A Eq.
hºvº- (9.19B

q.
TM") tan .1667):(10-2
4.1667):(
*\D)

i
t
2V D,
[.

)
|
N-ºvº.

X.
For 10 and 10; Eq. (9.19C)

>
Art
-
D,

a
N

i
- -

-
For the steady Eq. 9.16
*}=0 (cor

of
which

in
state solution
responding solution 9.19A for flat plate), Booth derived for
to

a
sphere the steady state number atoms N, external

of

to
of

the case
a

the sphere 170]:

Y,
W;
X, D,

– 4Taºf
At
Eq.
i ,

...)
(9.17A)
(coth D.Ta
|
M

for
Y, D.
(1.
}>10, Nº- 4traºf Eq. (9.17B)
At

M
i

or, defining
F,

escaping the
as

of

the fraction the final concentration


sphere,
_3 D. Eq. (9.17C)
=
F

V
\t

Eq.
an

Booth further devised approximation valid for


to

9.17
-
1

times less than


rºDºº

“rī,
Y,

D.

a?
t

Sira’f -X;t #1/2


*ſ

t!/?dt - Eq.
,

(9.17D)
TIZEN

.
.
.
V

)
(

decay constant
of

of A,

He also considered the which nuclide


in

case
a

3,

as

grows from precursor decay constant the formation


of

such
a

Xe” I* (Table the isotope decay


of

of

from 9.6). Here the fraction

-ºi
the sphere (F)
is:

the sphere relative


X,

that
to

to

constant external
in

- (3,-HX) 6\,(3,--X) |D
'ſ t
=
F

—e-Bitº)a (1—e-Bi") (e-Nºtºdt).


(1

Eq. (9.17E)
*
IRRADIATION EFFECTS IN URANIUM DIOXIDE 487

A quantity of considerable importance in consideration of system


contamination is the escape rate coefficient v, (sect') defined as [64]:

dN.
# =v, N. Eq. (9.17F)

The quantity v is directly measurable and does not require detailed


consideration of processes occurring in the fuel element. Thus, it is
of considerable practical significance, since it can be used to relate
the measured escape rate of isotopes to operating conditions of the
fuel element without reference to temperature or physical conditions
within the element. Its relation to physically meaningful quantities
is readily derived as follows for the case of a body operating at constant
temperature and uniform fissioning rate:

N. p
*}=, N.
—At Nº

*—fx-NN-N.
At steady state, these expressions are equated to zero and since, for
the short-lived isotopes which are of greatest interest for system
contamination problems, v,34, the following expression can be
derived:
vi-X, F, Eq. (9.17G)

For the caseof recoils, F, can be determined from Eq. 9.14A, B, or


C and for diffusion from Eq. 9.17C, yielding the expression:

-
vt=3 VDX. Eq. (9.17H)
— —

In cases where diffusion coefficients are quite small, the impoverish


ment of the surface of the body by recoils can seriously affect the
amount of isotope diffusing out. Flugge and Zimen have solved
Eq. 9.16, considering this correction for the steady-state diffusion
from a sphere for the case where 21,...}~ a [65]:

F– “. 1_1 sinh y(1–2) —1


\z
-
‘T2y’ z sinh y

-- ----
3 a: 1 1 cosh y (1–3) Eq.
;

-
tºy H.; (9.1
...

(9.18)
1–3 cothy sinh
y

where

–!º wºo. lºt


*=-j- via VI,
488 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

A number of approximations have been derived for limiting cases:

y(1–2)×1, y-1
!-- 12-wl –1
#{o-oſt;-#.
3
F, 3
| }*} Eq. (9.18A)

2:31

F=} Eq. (9.18B)


!/ ſooth v- Q/

(Compare with Eq. 9.17C.)

y}-3

F-; 1–) Eq. (9.18C)

y>3
F-3/y Eq. (9.18D)

The analogous solution for a plate of thickness (b+2la) cm was:

––
|D.

D.
{P

**30+3,...) Nily.
At

-V}+.
j, Vº
(b+l, sinh
ly,

cosh

}
b
V}-cosh
.)

P–
1.

i.
t
-

-
M

X
-

sinh (b+l,.
#-sinh lf.
)

Eq. (9.19)

For thick specimens where large, this expression simplifies


is
}
b

-
i

to:

-—“– \*#
\t

— e-lt.:
+
1+1

VD,
Eq.
|

F, (9.19A)
!

"2NN(b+21,0) |N.
f,l,

D,
i,

(2) DIFFUsion AFTER IRRADIATION. The experimentally impor


tant case volatile fission products previously
of

of

release from
a

irradiated specimen, i.e., the case which the term f\", Eq.
in
in

9.16 zero, has been solved for various geometries by Inthoff and
is

F,

Zimen Note that the following, the fractional release


in

[71].
is
f

by F=\;t, N, isotope
of

defined where the number atoms


is
i,

i
o
IRRADIATION EFFECTS IN URANIUM DIOXIDE 489

within the solid at the beginning of the postirradiation experi


ments (at time t-0) and N.,
is the number of isotope i atoms external

-
to the solid after t seconds of diffusion.
For a sphere:
– Dit,”,”
Eq.

...
F=1-4-º' *H
(9.20A)

"
cylinder height
of

For cm and radius cm:


a

a
*
–D,t(2m+1)2+2
- D,tr;


co ord
32
F-1 — ‘’t
1

1
— 2-\# m2 a?
1

*
6

6
r;
(2m+1)*
**
*

Eq. (9.20B)
r,

n” For
of

where the zero value the zeroth order Bessel function.


is

b,

parallelepiped edge lengths


a,
of

and c,
a

–– D,t (21+1)2+2

q2
1_512 e-At

,-a
=
F

Tr" f=# (21+1)*

—D,t(2m+1)?r?
oo
1

ſº-ſo (2m+1)*
–D,t(2n+1)2+2
op
1

c2
*

@nii). Eq. (9.20C)


For very long times, high diffusion rates,


or

small specimens, these


expressions reduce to:
for sphere
a

,-(+)
++
F-1–. Eq. (9.20D)

where
1.

Dºrº
n.T
a”

for cylinder,
a

32
1

—t(x+...) Eq. (9.20E)


***-(2,41s).
z:
6

where

|-p.ſ.º.º. T2 (2.404S)?
1

and for parallelepiped,


a

512
—t(x+.nz Q. (9. 20F
F=1-#
E

()
e

)
490 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

where
1 1 : 1 - 1

p a?

- - - -- - t
These approximations are accurate to within several percent for ;2 1.
- - - 1 -
Zimen and Schmeling point out that may
be considered a rate

constant for diffusion analogous to A, the rate constant for radioactive


decay [72].
- - . D
For short time periods or such that the quantity #~! , where a

is the resultant linear dimension or radius of curvature of the body,


the above expressions can be approximated by

F=2e-A ſº ; Eq. (9.20G)

where S is surface area in cm” and V is volume in cm”. Inthoff and


Zimen have also considered this case for the condition that the surface
of the body is impoverished by recoils during irradiation [71].
(3) DIFFUsion of STABLE Isotopes. Finally, of interest is the case
of formation of a stable isotope in which the term AN, in Eq.
9.16 is zero. Booth derived for the case of a sphere in which diffusion
occurs after a prior irradiation the expression [73]:
*H º',

6 –”:'P4
q2
nº 1

Eq.
F=1–. (9.21)
m

=

This expression simplifies to:

*>. *-i-.
_m^*Dif
2

for Eq. (9.21A)



e

Pºp!
2

for
rºle. F=}
Eq. (9.21B)
V

For the analogous irradiation occurring simultaneously with


of

case

º
the diffusion,


º,
6aº

"Tºrtrij, ºn
6a’
Eq.
1

(9.22)
*
2

which, for simplified


be

to

can

F– !-º!. Eq. (9.22A)


V
IRRADIATION EFFECTS IN URANIUM DIOXIDE 491

Thus, by means of the expressions developed above, experimental


measurements of fission product release rates can be related to basic
material parameters. The results of such investigations are reviewed
below.

9.4.3 Surface Area Measurements of Sintered UO2

(a) Surface Area-Density Relationship

The amount of release by diffusion can be seen to depend not only on


diffusion rate in the solid but also on the smallest dimension to a free
surface. Therefore, adequate correlations of experimental release data
will not be derivable unless the latter quantity is adequately specified.
In a completely dense body, such specification is obtainable from
dimensional measurements. However, since UO, bodies of interest
technologically are fabricated in the great majority of cases by sinter
ing (either in-pile or prior to irradiation), residual porosity will
remain in the body in the form either of closed porosity, i.e., not com
municating to a free surface, or of open porosity. The relation
between these forms of porosity and sintered density are discussed in
Ref. 74 and Chap. 7. In Ref. 64, the diffusion path to the free surface
was estimated by measuring total surface area by gas adsorption
techniques and idealizing the structure of the body as an assemblage
of equivalent spheres of radius a such that the surface area of the
spheres equaled that of the sintered body. A plot of surface area of
UO2 (measured by krypton, butane, or ethane adsorption) as a function
of sintered density is shown in Fig. 9.18, utilizing data in Ref. 64 as
well as the more recent data of Aronson, et al. [75]. The parameter
plotted, S in sº, can be related to the equivalent sphere of radius a

by the relation
- : (fraction theoretical density). Eq. (9.23)

Alternatively, the measured surface area parameter could be used


directly in expressions such as Eq. 9.20G. It is probable that the
actual diffusion paths in a body vary over a range about a mean length.
However, the degree of inaccuracy and nonreproducibility in experi
mental measurements both of surface area and of release rates are
presently such as to render refinement of the treatment unnecessary for
correlation of fission product release with oxide characteristics. Ex
perimental observation of release of fission products both from speci
mens operated in-pile as well as from postirradiation heating confirms,
at least qualitatively, the relationship of surface area and, hence, of
492 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

+
5
T- I I i I

104

i
IOOO

IOO

o l l l l l
70 75 8O 85 90 95 too
PER CENT THEORETICAL DENSITY

FIGURE 9.18. Measured Surface Area (BET) as a Function of Percent


Theoretical Density for Sintered UO,.

diffusion coefficient, with density shown in Fig. 9.18, as will be dis


cussed in Sect. 9.4.4.

(b) Effect of Closed Porosity

Closed porosity in sintered compacts may contribute appreciably


to the total surface area contained in the body. In fact, as discussed
in Ref. 64, in a 95 percent dense UO, compact, the surface area of
the closed pores will approximately equal the total surface area
communicating to the exterior of the compact and, thus, in a body of
such density, it would be anticipated that as much volatile fission
products would collect in closed pores as will be driven from the ex
terior of the sample. Experimental verification of this expectation
is discussed in Sect. 9.4.9b. Drainage of dissolved fission products
into such closed pores may introduce errors in calculation of diffusion
coefficients. However, so long as small fractional releases are of con
cern, for which equations such as 9.20C are applicable, as is generally
the case, little or no error is introduced by the presence of closed pores,
and such errors as are introduced in estimations of release rates are
in a conservative direction.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 493

9.4.4 Experimental Measurements of Diffusional Release of


Fission Gases from UO,

Almost all the experimental measurements of fission product dif


fusion rates from irradiated UO, have been performed on the rare
gases Xe and Kr. The release of these gases is of most interest tech
nologically in their effect on fuel element behavior because of the
effect of buildup of pressure within fuel element cladding by collec
tion of these noncondensible gases, because of possible thermal effects
of their release to gaps or voids within the sheath, and because of the
measures necessary to assure their radiologically safe collection and
disposal in reactor plants containing oxide fuel elements.

(a) Applicability of Diffusion. Model

The most careful study of the applicability of the diffusion kinetics


outlined above to the release of fission gases from UO, has been that
of Booth and Rymer [76]. These authors irradiated 20- to 50-mg
samples of natural UO, in air, wrapped in aluminum foil, in a
2 × 101* n/cm3-sec flux for 1 to 10 hours (7 to 70 × 101° fissions/cc).
Specimens were then cooled for 3 days so that Xe” was the only radio
active rare gas atom remaining in appreciable quantity. The speci
mens were then heated for various times at different temperatures in
a circulating, purified He atmosphere and the release of Xe” meas
ured. In a few parallel experiments, the authors detected no signi
ficant difference between the release rates of krypton and xenon
isotopes. Neither was any difference detected between the release
rates of Xe” and Xe", the latter isotope being generated by decay
of I*. From this observation they concluded that I diffused at the
same rate or slower than Xe.
The results of Booth and Rymer’s work are presented graphically
in Figs. 9.19 to 9.26. Initial experiments to confirm the validity
of the diffusional mechanism of fission product release were performed
on samples of fused UO, crushed to various particle sizes. As can be
seen from Eq. 9.21B, at small fractional releases the fraction re
leased should be inversely proportional to particle radius. This rela
tionship is excellently confirmed in Fig. 9.19. Also, as can be seen
from this same equation, the diffusional mechanism predicts that frac
tional release should be proportional to (f)”. Satisfactory agreement
for heating times up to four hours is shown in Fig. 9.20, although it
should be noted that the data show an intercept at zero time and a
slight tendency toward curvature. Finally, in Fig. 9.21, the varia
tion in diffusion coefficient for Xe” in fused U(), is plotted as a func
tion of temperature, and the variation with temperature expected for
a diffusional process is obtained:

D=1.5× 10-s e-º-º/RT cm3/sec.


494 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

4.5 I I I I I I

4.O. H. -

3.5 H. H— — “—— —
PASSING 325
MESH screen

l I I I l
O |OO 2OO 300 400 5OO 6OO 7oo

# Powder RADIUS, o, IN cm

FIGURE 9.19. Release of Xe” from Fused UO2 as a Function of Powder


Radius [76].

(b) Applicability of Diffusion Model to Release from Sintered U0,

Having demonstrated the validity of the diffusional mechanism


under these experimental conditions, the authors then considered re
lease from sintered rather than fused UO. In Fig. 9.22 the frac
tional release from a sintered body of about 85 percent theoretical
density is measured as a function of the particle size of the crushed
fragments. However, as opposed to the observations in Fig. 9.19, no
effect of fragment size on fractional release was noted. This is ex
pected from the information in Fig. 9.18, since the particle radius of
an 85 percent dense pellet, from the surface area measurements, should
be about 2.5 × 10 cm, and the fragmentation through 325 mesh should
add inappreciably to the actual surface area exposed. The marked
effect of density on the fractional release is graphically shown in Fig.
9.23 which, at least qualitatively, confirms the conclusions drawn above.
/
IRRADIATION EFFECTS IN URANIUM DIOXIDE 495

2.5 -I —I- I

2.O. H. -

º
ul
un
<r
Lil
-
1.5

º;
H

ck
ro
re
+

2:
c
2
Lu
;

×
* -
Lo
H
:

Q->
ar
Liu
a.

O.5 H. —
5 1
O

I
I

|O 15
O

VTIME, MINUTES

Xe” from Fused UO2 (Powder Radius


of

FIGURE 9.20. Release 0.0028 cm) as


Heating
at

1,400°
of

of

Function Time [76].


C
a

Note that, again, the variation fractional release proportional


to
in

is

(f)”; note, addition, that the intercept


at

zero time even more


in

is

the fused oxide (Fig. 9.20).


of

marked than was the case The


in
it

true particle size sintered compacts illus


of

marked effect also


in

is

Fig. (see Eq. 9.21B) plotted


in

trated which the quantity}


in

is

9.24

gross density for various sources


of

of

function
as

sintered oxide.
a

particle independent density, should, course,


of

of

Were true size D/a”


plot. plot
be

constant this The also illustrates that even


in

at

the
by

same density, material prepared different fabrication techniques


from different initial powders can vary widely gas release prop
or

in

erties. Using the diffusion coefficients obtained from fused UO2


in
H

-||
i
T
T
I
i

i
T
I

I
I
I
i
SC)

T
I

i
T
I
T
I
i
CS

o
c;

- A.
46,
ooo

--
x

-
12H. *; sooth and rºwer, Fuseduozia--2ea.0-1.5 lo-" =# emº/sec :

oe
-

as
**; LINDNER
ANDMATzke 315,UozPowder -
%

D.
-
a
LINDNER
ANDMATzkE, -oo.25-oxisº.uoz,D-2xio-"e- **. cm"/sec *
2Ausºn
ºveooth

, ,
and Rºwen

-
\. -I
LINDNER

*** 85
ANDMATzXE

>
H.

Y Y. \
:

co,
AuskERN. o.22,.uo, Powder,0-4s, lo-" ---tº em”,sec

“"
,:
a. or
AuskERN,
t5

28, uozoo,crushed PELLETs;D-2ss, lo-"---º em”asec

-
14H.
)

x• x• x• Kr k, K. x•
;
(-
**; BosTRow, 97%DENsePLATE D-6.6 lo-"exp cm2/SEC º -

i
x a
Kr" susko 93% DENSESINTERED PELLET (FRomFig.9.16)-32,
:3

- AA v v O
a
,,
Krºº
-
-> susko 97.4%DENSESINTERED PELLET (FROMFIG.916) 1461, --

l
Lindner And Matzke
MATzk
-

3
H
x
ww3

=
8.1 10-7e-*7,oooºººcm2/SEC it.
- -18

3\
:0

5
i i
i
i

w
=

xe
LoNG SINTERED
PELLETs 58%.DENSE 0.12/1
3

N §
q=

64%.DENSE: 0.14/1

&
-1 H tr;

:
-
& ©o
65% DENse ore, -0

2
7
D.
: q g-a-

exio-, -ºº-emº/sec 77.5%dense oss, H


93%.DENSE 0-26,
s

Long

-
17H 93%.DENSE a-sion wº

i
Kr", susko ANDMARKowitz sinterEDPELLETs 80%.DENSE 1.7M.
(FROMFIG.9.16) -
w

/
91%.DENse 12.8,
:

K
TRom
susko,BosTRO 91.7%DENse;a-15,
=

t;
D
92.5%DENSE 7-"a

-
ie. H
eQ d [E D B G G

-
: ; : : q=: a-; a: :
a-

95% DENSE 43,


Z

-
19H.
-
t
5g

:0

w

-2O.H.
*

-
*t

t+
x
to

4.7 -22

I
l
I

1
I

O
l
l
l
l
§
l
l
l
l

—1– —l
>

5
I7
to li

l 6.
12 13 14

l 9.
is 16 17 te 19 20 21 22 23 24 25 26
H

10-yr(“K
co
- º
ar.

FIGURE, 9.21. Diffusion Constant. IX, for Fission Gases in UOa.


–"
IRRADIATION EFFECTS IN URANIUM DIOXIDE 497

--
-
50H.
O

.I | |
|
2o -

PASSING325
MESH

l l I I 1– I
t 5 IO 50 IOO 500 IOOO
# (a • RADIUSof FRAGMENTs.IN cm)

FIGURE 9.22. Release of Xe” from Sintered UO, of Density 9.35 g/cc as a
Function of Size of Pellet Fragments; True Particle Radius from Figure 9.18
for 85 Percent Dense Oxide e
2.5X10^* cm (1/a-4000 cm") [76].

Fig. 9.21, Booth and Rymer calculated the true particle radius which
is plotted as a function of density in Fig. 9.25. Also replotted in
Fig. is the curve of particle radius versus sintered density
9.25
derived from the surface area measurements in Fig. 9.18. It may
be noted that the agreement of particle radii derived from surface

area measurements with those from fission gas release measurements

is quite good at corresponding densities except for the points labeled

Mines Branch (Standard). Data on surface areas in Fig. 9.18


were derived from samples sintered 4 to 12 hours at 1,700° to 1,750°C,
and, thus, a stable condition of pore shape would be anticipated at
each density level. The Mines Branch (Standard) samples were sin
tered one-half hour at 1,700° C and, presumably, have not attained
a stable pore shape or distribution even though they may have reached
for full

is,

therefore, not surprising that


It

density. values calculated


a

samples appear represent different population than the


to

these
a

represented Fig. 9.25. Thus, may


be

ºther samples concluded


in

it
the

comparison plotted Fig. 9.25 indicates the applicability


in

that
"fout-of-pile surface area measurements for determination
of

suitable
sin
of

fission gas release properties


of

Values for characterization


*

as,

"red UO, and further indicates the utility Fig. 9.18 least,
of

at

*qualitative guide for highly sintered compacts.


the widely variant gas release properties
as
of
In

spite function
a
498 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I - DENs. 9.35, (85% T.D.) |400°C, MEAN OF IO


ao L 2 -- SAME, SINGLE, RAN, 33
9.85, (90% T.D.)MIN.INIEºs
1400°C, MEAN
F- -
3 DENS. OF 2
4 – DENs. IO-50, (96% T.D.) 1400°C, MEAN OF 2
5 – DENs. 9.35, (85% T.D.) IOOO"C., MEAN OF 4
6 – DENs. IO-63, (97% T.D.) 1400°C MEAN OF 2

-VTIME, MINUTES

FIGURE 9.23. Release of Xe” from Sintered UO, as a Function of Time of


Heating at 1,400° C [76].

of density, Booth and Rymer found that the activation energy ob

tained from plots of logº vs 1/7' for sintered compacts (Fig. 9.26)

was quite similar to that observed in the case of fused UO, (Fig. 9.21).
On this basis, Booth proposed characterizing sintered UO, compacts
by irradiating for short periods, measurement of gas release after
heating three hours at 1,400° C, calculation of p'-'. from these

data, and use of values of 1) plotted in Fig. 9.21 to calculate the ap


plicable value of a [73]. Alternatively, the measured D’ values may
be used directly and extrapolated to temperatures of interest using
the measured activation energy of 45,000 cal/mole.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 499

PER CENT THEORETICAL DENSITY


7o 75 8O 85 90 95 IOO
Io-6 I I I i T I

O O

10-7 – -

O
5
o
3 to
in-8 – -
&

I. MINES BRANCH, STANDARD METHOD

to-e L 2. M.BR., Long sintering


3. C.G.E.
TIME -

N
4. B. M.I.
*
5. W.A.P.D.
2N. *
**
&\ ** W

10-10 l I | l l l
8.O 8.5 9.0 9.5 IO.O IO.5 | I.O
DENSITY OF PELLET, g/cc

FIGURE 9.24. The Diffusion Constant of Xe”, D' (D/a”) at 1,400° C, as a


Function of Density of Sintered UO, [76].

(c) Other Measurement of Fission Gas Diffusion. Rates

Additional information on release of Xe” from irradiated UO2


was obtained by Lindner and Matzke [77–79]. These investigators
used various chemically prepared powders of small particle size and,
thus, were able to extend measurements to temperatures below 700° C.
The results of the investigations are plotted in Fig. 9.21. The
burnups in each case were of the order of 10° fissions/cc: surface
areas and particle sizes were measured by BET gas adsorption and
electron microscope measurements. The initial results are considered
high because of the possibility of oxygen contamination of the UO.

57.4789 O–61—33
500 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

200 T I I I I I

IOO H. -
90 H
ſ -
80 H. I
7o H
I -
-
, 2"
60 H.

50 H. BATTELLE -
4O H. -
3O H. -

on 20 H
-
->
co
ar.
9
->
uj
ºr:
#
#
on
Io
9 H.
H --
§ 8 H
-
g 7 H
M. B.R.
(LONG SINTER NG) -
g 6 H M. BR.
-
§
<I 5 – (STANDARD)
º:
4 H -

3 H -

,/
2 H. Z —
O O

VARIATION OF o witH DENSITY


FROM FIG.9.18

*z
| | > | 1 l l —l
7.5 8.0 8.5 9.0 9.5 IO.O. IO-5 II.O.

DENSITY OF PELLET, g/cc

FIGURE 9.25. The Radius, a, of the Hypothetical Diffusion Sphere as a Function


of Density of Sintered UO, [76].

[77]. Later results, however, showed remarkably good agreement


with those of Booth and Rymer, considering the wide difference in
particle sizes employed [78]. The authors also measured the diffusion
coefficients for I*
and, although the results were not reported, stated
that diffusion constants and activation energies were similar to those
observed for Xe”. It is important to note that these authors ap
parently used heating times up to one hour in duration. In later
IRRADIATION EFFECTS IN URANIUM DIOXIDE 501
O
IO-6 Q

to 7

;
-ch-
108

IO-9

O-M. B.R.PELLET, FROM x-2-f LOOP TEST


O-M. BR. ExPERIMENTAL PELLET *Iy DENs. 9.35

10-ſo l l l l I I
6.O 65 70 75 8.0 85 9O 95

RECIPROCAL OF ABSOLUTE TEMPERATURE x IO-4

! FIGURE 9.26. The Diffusion Constant D' (D/a”) for Xenon in Sintered UO, of
Approximate Density 9.4 g/cc as a Function of Temperature [76].

work, shown also in Fig. 9.21, the same authors found the diffusion
coefficient in these powders for Kr" to be slightly higher than that
for Xe” [79].
Auskern performed experiments on the release of Kr" from UO,
powders irradiated to about 4× 10° fissions/cc in quartz vials under a
helium atmosphere [80]. Two types of powder were used, one pre
pared by hydrogen reduction of UO, with the powder subsequently
ball-milled to reduce a to 0.22 microns, the other made by crushing
sintered 92 percent dense pellets to minus 200, plus 325 mesh (a =2.8).
Surface areas were measured by BET gas adsorption prior to irradi
ation. Heating times at temperature varied from 8 to 24 hours; the
isotope Kr" was detected. The results of this work, plotted in Fig.
502 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

9.21, deviate significantly from those of Booth and Rymer and of Lind
ner and Matzke, particularly with respect to the magnitude of the ac
tivation energy for diffusion. Auskern finds this value to be 65,500 to
73,800 cal/mole, while the other authors obtained values of 46,000 to
48,900 cal/mole. The high values of D obtained with the crushed pel
lets as compared with hydrogen-reduced powder may be related to
strain produced in the particles during the crushing operation. A
similar effect of crushing on increasing the self-diffusion rate of oxy
gen ions in UO, has been detected (see Chap. 7, Sect. 7.2.3).
The results of Long, plotted in Fig. 9.21, were obtained from com
pacts sintered to various densities, and diffusion path lengths were
determined for each compact from BET surface area measurements.”
The agreement with the lower of Auskern's curves is remarkably good,
considering that the escape of Xe, rather than Kr, was measured. Not
only were similar activation energies calculated, but also the magnitude
of the diffusion coefficients at each temperature agree quite well. Note
that the values obtained at low temperatures deviate appreciably from
the D values extrapolated from the higher temperature points.
Also plotted in Fig. 9.21 are the data of Susko which differ from
the other measurements in that they were performed on sintered speci
mens irradiated to some appreciable burnup [64, 81]. The data out
lined in Ref. 81 were obtained at a burnup of about 1.3×10”
fissions/cc,whereas certain of the specimens described in Ref. 64
were irradiated to as much as about 2×10° fissions/cc. Again, for the
samples treated at elevated temperatures, the activation energies ob
tained agree well with those measured by Long and by Auskern [81].
Likewise, the low temperature data reported by Susko in Ref. 64
show the same break in the plot of log D vs 1/T as was found by Long
at a temperature of about 800° C. However, it must be noted that the
measurements of Susko at elevated temperatures yield absolute values
of the diffusion coefficient about one-thirtieth of those reported by
Auskern and about one-fifth to one-tenth of those measured by Long.
The data of Bostrom were obtained from high density UO, platelets
subjected to a short irradiation, and the isotope Xe” was measured
[82]. His data show good agreement with those of Long and of Sus
ko and lie intermediate between the two latter sets of data. In fact,
the agreement between the data of Long and those of Bostrom is re
markably close. The latter, working in a temperature range of 800°
to 1300°C, found for Xelas that

D=6.6×10−" exp (-71,700/RT)

whereas Long extended the temperature range from 800° to 1,800°C


and reported
D=7.8×10−" exp (-71,000/RT).
* G. Long, Atomic Energy Research Establishment, Harwell, unpublished data.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 503

(d) Low Temperature Diffusion Rate Measurements

A final significant set of measurements was that of Markowitz re


ported in Ref. 64. Here samples of varying density were exposed
to fissioning in a circulating stream of helium in-pile and the diffusion
rates of the short-lived species Kr" and Xe” were measured as a func
tion of temperature by determining the equilibrium rates of release -
of each isotope and applying Eqs. 9.17B, 9.17H, 9.18D, etc. It
will be noted in these equations that the evolution rate (or escape rate
coefficient, e.g., 9.17F) is inversely proportional to a or (Eq.
9.14A) directly proportional to true surface area. As shown in Fig.
9:27, in spite of the scatter of the data, the effect of increasing density

soft-H-I-T-I-T-I
- DENSITY-9. THEORETICAL
91.8
-
5
§ 5 94.2

4O H.

30 -
KRYPTON-87

Kr 87 xelss
O D- IOO" C -
e --350°C
A A-550°C
2O

IO [] -
A -

Le
O
-
HA -
O
O
G I-1–1–1–1–1–1–1–1–
5OO |OOO |500
TOTAL SURFACE AREA, cm3/cc

FIGURE 9.27. Effect of UO, Density (Surface Area) on Emanation Rate


In-Pile [64].
504 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

in decreasing emanation rate may readily Thus,

the
be recognized.
decreasing density
increasing gas

of
marked effect fission

in
release
postirradiation annealing and

on

on
has been demonstrated both con
tinuous exposure in-pile. Fig. 9.27, little

no
In

or
trend the varia.

in
emanation rate with temperature can

be
of
tion noted between 100°
This observation bears out the trend the data plotted

of
C.
and 550°
Fig. that changes diffusion coefficient with temperature
in

in

in
9.21
higher tempera

at
below 800° are much less than those observed
C
Note also that the differences

in
tures. escape rate coefficient between
krypton and xenon are the decay con

be
attributed

in
to

to
differences
stants for the two isotopes rather than

of
true difference diffusion

a
coefficient.

(e) Summary

The experimental results presented this section have, thus, dem

in
certainty fission product

of
onstrated with reasonable that release
by

gases from UO, occurs diffusional mechanism and that effects

of
a

density variation sintered compacts are consistent with such


in

a
mechanism. However, the results have posed serious question

of in in
a
that, six independent investigations, three have yielded results
of

agreement with each other but discordant with the measurements


three other investigators whose results, again, were mutually compati
explain correlate these discrepant

re
or

ble. Before attempting


to

sults, additional related information will reviewed which may aid


be

resolution of this dilemma.


in

of

9.4.5 Diffusion Noble Gases and Self-Diffusion UO.


in

(a) Radon

The Hahn emanation method has been used for many years
to

detect
or

phase changes other phenomena solid materials [83].


In
in

this method, the isotope Thº" (radio-thorium) form


to

mixed
is

solution with the material under investigation, and the escape


of

the
second a-decay daughter product, the gas thoron (radon–220), meas
is
of by

this technique was employed


of

ured. variation Lindner and


A

on

Matzke who adsorbed radium-226 the surface powders and


by

injected Rn” a-recoil into the particle surface [79]. Flugge and
Zimen developed the theory reviewed Sect. 9.4.2 relating the emana
in
of

diffusion rate the solid [65]. The diffusion rate


to

of

tion rate Rn
in

of

radon, the heaviest pertinent


of

to

the noble gases, discussion the


is

the fission gases xenon and krypton.


of

diffusion rates
Gregory and Moorbath measured the diffusion rate
of

thoron
in

Al2O, and also reviewed the results obtained


on

other oxides [S4].


a
IRRADIATION EFFECTS IN URANIUM DIOXIDE 505

– 12 H -- \ I

U IN UO2
OF
w
WN
\ \
T WTI
\
\
r–
T

DIFFusion of o-IN
STOICHIOMETRIC UO2.o
I

Ç-~ *
\ Diffusion of
Rn **9 IN ALzos
".
N.
N. Rn 220 IN Fezos
N.
SSS
9. IoW
\ Ts

\N ^, \ S.
W
W Rn 220 IN
N A creos S
N Rn 220 INW

N TiO2 W

Nº W
N \
N \
N W

BOOTH 8 RYMER
*
\ DIFFUsion
He IN UO2
OF
N \W
\ \
-

SUSKO
LONG
\\
\\
DIFFUSION OF FISSION

- if H
GASES IN UO2
O

+ UNPUBLISHED DATA, J.C. CLAYTON, BAPL

- 18 1 I l l I I _l
O.5 O.6 O.7 O.8 O.9 I.O I. I 1.2 1.3
Io-3/T ("k)
FIGURE 9.28. Diffusion of Noble Gases in Metal Oxides.

The data on these materials 84 are plotted


from Ref. in Fig.
9.28. Several may be noted in these data;
common characteristics
from the variation of diffusion coefficient with temperature, a high
activation energy for the diffusion of thoron is observed at high tem
peratures, and at low temperatures a much lower energy. The tem
506 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

perature at which the activation energy changes has been identified


with the Tammann temperature, namely, that temperature at which
self-diffusion rates are sufficiently rapid that displacements over dis
tances of the lattice parameter can occur during the time of measure
ment, 1 to 10° sec. Thus, the Tammann temperature has been taken
to indicate the temperature at which the diffusion coefficient is 10-" to
10-” cm”/sec, or a temperature equal to 0.51 to 0.52 times the melting
temperature, etc. In any case, it has been assumed to signify the tem
perature below which the concentration of lattice defects is fixed or
frozen in, and above which the equilibrium number of lattice defects
forms at each temperature. Thus, activation energies above the Tam
mann temperature, T., should consist of the energy of disorder E, plus
U, the energy barrier for atomic motion, whereas at T3Ts, Q, the
activation energy, should consist primarily of the energy term U.
From the definition of T, in terms of diffusion coefficient, Flugge and
Zimen postulated that the term Q/T's should vary between 60 to 80
cal/mole—” K [65]. The data for the oxides Al,Os, Cr,0s, Fe2Os, and
TiO2 are not in obvious agreement with this postulate, as shown in
Fig. 9.28. In any case, however, the deviations of the diffusion
coefficient versus temperature plot for fission gases in UO, noted at low
temperatures in UO, by Susko, Markowitz, and Long (Fig. 9.21) are
confirmed in the related case of diffusion of thoron through oxide
lattices and are shown to have some theoretical justification.
Measurements of the diffusion of thoron through UO, have been
reported by Anderson, et al., for powder and by Clayton for sintered
platelets 80 percent dense [85]. Their data are plotted in Fig. 9.28.
The agreement between the two sets of measurements is quite good at
the lower temperatures, but the data diverge by a factor of 10 at tem
peratures above 1,300° C. The data of Anderson, et al., indicate a
gradually increasing activation energy for thoron diffusion from about
20 kcal/mole at 500°C to over 80 kcal/mole at 1,300° C [85]. An
derson, et al., plotted their data to show three ranges of activation
energy, 81 kcal/mole above 1,155°C, 38 kcal/mole between 800° and
1,155° C, and about 20 kcal/mole below 800° C [85]. The data of
Clayton yielded an activation energy of about 45,000 cal/mole, in
excellent agreement with the data of Booth and Rymer and Lindner
and Matzke on fission gas release, whereas the data of Anderson, et al.,
appear to confirm the results of Bostrom, Long, Susko, and Auskern.
Furthermore, consideration of the significance of the Tammann tem
perature of 800 to 900° C shown in the plots of Fig. 9.21 and applica
tion of the factor (9/7's=62 found for UO, in Fig. 9.28 yield an
activation energy of 65 to 72 kcal/mole, again in agreement with the
latter group of observations. Thus, consideration of radon diffusion
* J. C. Clayton, Bettis Atomic Power Laboratory, unpublished data.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 507

cannot serve to differentiate between the equivocal fission gas diffusion


results.

(b) Helium.

Bostrom measured an apparent solubility and diffusivity of helium


in UO, powders 0.16 to 0.44, in diameter [81, 82]. Bostrom's results
on diffusion rates of He in UO, are plotted in Fig. 9.28 and are seen
to be in excellent agreement with Clayton's higher temperature thoron
emanation measurements. He also reported the solubility of helium
in UO2 at one atmosphere helium pressure to be 2.5 × 10-4 cc He per cc
UO, at 700° C. Bostrom's observation of He solubility and diffusion
in UO, may be considered surprising in view of the results of other
investigators who have been unable to detect helium permeabilities at
20°C and 415°C through a wide variety of ionic crystals including
the mineral fluorite [86]. However, the maximum reported sensi
tivity of measurement in Ref. 86 was about 6X 10-11 cc/hr/cm”/mm.
thickness/atm pressure. Applying Bostrom's measurement of
2.5 × 10−" cc He per cc UO, solubility to the reported maximum tem
perature of measurement of permeability in Ref. 86 of 415°C,
it can be shownthat the measurement technique must possess a sensi
tivity of about10-4” cc/hr/cm2/mm/atm to detect permeabilities in
agreement with the diffusion coefficients of Bostrom plotted in Fig.
928. The application of measurement techniques of sensitivity equal
to those used by Bostrom permitted Gentner and Trendelenburg to
measure solubility and diffusivity of He in NaCl [81, 87]. In glasses
in which permeability and solubilities of rare gases may readily be
measured, diffusivities 14 orders of magnitude greater than those
plotted in Fig. 9.21 are encountered. Thus, there is no reason to sus
pect that true solid state solution and diffusion processes are not
occurring in the case of the helium diffusion experiments reported by
Bostrom. If this, indeed, be true, the He diffusion data would tend to
support the validity of Booth and Rymer's and Lindner and Matzke's
data for Xe–Kr diffusion.

(c) Uranium and Oaygen

Additional light is shed on the diffusion rates of fission gases by


consideration of the self-diffusion rates of oxygen and uranium ions in
UO, (see Chap. 7). The self-diffusion rate of the former has been
shown to vary markedly with oxygen content, the activation energy
for oxygen diffusion in nonstoichiometric UO, being independent of
oxygen content and having a value of about 30 kcal/mole [81, S2].
In stoichiometric UO, on the other hand, as shown in Fig. 9.28, the
activation energy for diffusion of lattice oxygen was found to be about
508 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

65 kcal/mole [82]. Similarly results of uranium ion self-diffusion, as


shown in Fig. 9.28, agree not only in magnitude of activation energy
but also in value of diffusion coefficient with the Xe-Kr data of Long,
Susko, and Bostrom [82]. This agreement is probably more than
fortuitous. As discussed above, diffusion of thoron or other noble
gases probably occurs by a vacancy mechanism involving motion from
occupied to unoccupied lattice positions. In metal lattices, LeClaire
and Rowe have shown the diffusion rate of rare gases to approximate
that of metallic self-diffusion rates [88]. Thus, depending on the
occupancy by the fission fragments of anion or cation sites in ionic
lattices, diffusion rates would be anticipated which would be similar in
magnitude to those noted for anion or cation self-diffusion. Since the
absolute magnitude of each of the measurements of fission gas diffu
sion rates corresponds with that of cation self-diffusion in UO2 (Fig.
9.28), it is logical to assume that fission gases (as well as radon and
helium) diffuse via cation sites. Thus, these self-diffusion data would
lend support to the validity of the high activation energy set of data.

9.4.6 Factors Affecting Fission Gas Diffusion in UO,

(a) Activation. Entropy

Considering the internal consistency of each of the sets of fission gas


diffusion data and the support offered each set by the supplementary
observations discussed above, it is impossible to select one set of data
over another; rather it is likely that each set of data is in itself correct
and that unappreciated and uncontrolled experimental or material
variables yield the differences which are observed. Reconciliation of
the data may be possible by considering the implications of the nu
merical values obtained by the various investigators. Zener has postu
lated for diffusion in metal systems that the activation entropy for dif
fusion should be a small positive number and that large negative num
bers indicate the availability of short-circuiting paths for the diffus
ing species [89]. Applying this same reasoning to ionic solids, such
as UO, the entropy of activation (AS) may be estimated from the pre
1),

Figs. 9.21 and 9.28


by

exponential term the Arrhenius plots


of
in

the relation

p-yº. Eq. (9.24)

constant equal one for face-centered cubic arrays,


to

where
is

is
y

a
a

U-U distance UO2, 3.86S


be

in
to

the interatomic distance (assumed


Å), and
an

the frequency
of

vibration calculation from assumed


is
v

UO, 6–873° Values of AS for


Debye temperature for K–% the

various experimental observations are listed Table As would


be
in

9.8.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 509

expected, the data indicating low activation energies for diffusion


yield large negative values for the entropy of activation. It appears
reasonable, therefore, to hypothesize that even at temperatures above
the Tammann temperature, T's, nonequilibrium structures can exist in
the oxide which prevent observation of true volume diffusion.

TABLE 9.8—ENTROPIES OF ACTIVATION FOR DIFFUSION IN UO2

Diffusing species Do(cm2/sec) S(e.u./mole) References


Xe”-- -- -- ---- -- - - 1. 5 × 10-8 – 28. 7 76
Xe”- - - - - - - - - - - - - - 6. 6X 10-6 — 16. 6 82
Xel”-- -- ------- - - - 7, 8X 10-6 — 16. 3 >k

Kr"------- - - - - - - - - 4. 9 × 10-4 – 6. 03 80
Kr”--------------- 8, 1 × 10-7 –20. 7 81

Rnz”, He –24.
--- ---- - - 1. 15X10-7 7
{:
Rn??0 --- ---- ------ 5. 89 × 10-2 1. 54 85
O----------------- 1200 21. 3 82
U----------------- 4. 3 × 10-4 –8. 3 82

*G. Long, Atomic En Research Establishment, Harwell, unpublished data.


**J. C. Clayton, Bettis Atomic Power Laboratory, unpublished data.

(b) Anomalous Intercepts

If nonequilibrium structures indeed exist at high temperatures,


it would be expected that anomalies would appear in measurements
at such temperatures. Thus, anomalous intercepts were commented
on in connection with the isothermal experiments of Booth and Rymer
(Figs. 9.20 and 9.23); similar intercepts were observed by Lindner
and Matzke [77, 78]. Many investigators have commented on burst
phenomena noted upon changing temperature during fission gas re
lease measurements. Felix observed on heating irradiated UAO,
powders an initial rapid evolution of xenon followed by a much slower
release over a period of a number of days, as shown in Fig. 9.29 [90].
Long-time heating experiments have been performed by Bostrom *
who irradiated 95 percent dense UO, pellets to slight burnups and
observed their rate of Xe evolution at 1,300° and 1,400°C over long
time periods up to 7 days (Fig. 9.30). In contrast to the results of
short-time (3 to 4 hours) exposures reported by Booth and Rymer,
plotted in Fig. 9.21 and in agreement with the findings of Felix in
Fig. 9.29, after an initial period of
8 to 20 hours, divergence of the
fractional release from a parabolic relationship with time is noted
and becomes very marked after three to four days. The diffusion
coefficients calculated from the final data points of Bostrom in Fig.
9.30 are plotted in Fig. 9.28 as the half-filled circle points and are
seen to agree best with the data of Long and Susko.

* W. A. Bostrom, Bettis Atomic Power Laboratory, unpublished work.


510 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

o.7 I I i

*—
O.O.
56
--
O. -
4

O. -
3
|

-—7OO"
O. H. C

-— 500 °C —
12 I6 20
4

8
O

TIME IN DAYS

Figure 9.29. Diffusion of Xe” from UAO, Powder [90].


(Courtesy, Springer-Verlag.)
.

necessary about the method plotting experi


of

of
A

word caution
is

obtain valid values for diffusion coefficients when finite


to

mental data
Figs. 9.20, 9.23,
as

intercept zero time are encountered,


at

values
in

will yield valid


of

9.29, and 9.30. Plots fraction released versus tº/*


diffusion coefficients from the slopes sufficiently long-time
of

at

values
periods for cases which isotopes are being released parallel from
in

in

sites exhibiting different diffusion rates. If, however, volume diffu


by

an
of

sion from the bulk oxide inhibited the presence altered


is

skin through which diffusion must occur for some time before true
a

gradient established, Inthoff and Zimen indicate that values


of

diffu
is

sion coefficient can more correctly obtained by plotting, Eq.


be

in

9.20G, versus time and measuring the slope, rather than


(Feºt')
*

plot
In

determining the slope Fig.


of

of

Flexi' versus tº/* [71].


a

20

9.30, the former procedure yields values


as

large
of

almost times
D

those noted with the second procedure: even


of

the rela
in
as

the case
tively minor intercept noted Fig.
difference 9.20, 20-percent
in

is
a

noted between the two methods. Since the latter condition appears
physically more reasonable for fission gas diffusion UO, use
of
F.
in

plots
of

versus recommended for calculation diffusion coefficients.


is
f

9.4.7 Summary

fission gases from bulk UO, has been demonstrated


of

to

Release fol.
an

low the kinetics and temperature dependence


of

activated solid
IRRADIATION EFFECTS IN URANIUM DIOXIDE 511

O.OOI O.O2 I i i i I i

- o -
D=1.68x10-15 cm
cm3/SEC
e-e—"
-
|400°C (FULL
•*
SCALE =O.O2)
4-w-"T
2
ºw

,-
1300 °C (READ FULL SCALE =O.OOI)
e
ºv
I -

/
|-
I gº
I
I
-
O.OOO5 O.O. H.
if
O
J

I l L I l I
o I 2 3 4 5 6 7
TIME IN DAYS

FIGURE 9.30. Diffusion of Xe” from Sintered Dense (94.8 Percent Theoretical
Density) UO2.

statediffusional process. The effect of decreasing density in increasing


available surface area and of decreasing diffusion path, thus, accelerat
ing the rate of diffusional release, has been quantitatively demon
strated. The relation between diffusion of fission gases in UO, and
other diffusional processes in UO., particularly that of uranium ions,
has been demonstrated. Quantitative agreement on values of diffusion
rates between different investigators has not yet been achieved. These
differencesmay be real and may demonstrate real differences between
experimental materials which are not controlled or experimental
reasons for the differences may become clear as further data are ac
512 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

cumulated. For example, segregation of insoluble fission products at


particle surfaces as discussed in Sect. 9.4.10, is a phenomenon, the
significance of which in relation to fission gas release has not as yet
been assessed. The available information on fission gas release from
UO, at high burnups will be discussed in Sect. 9.5; the data in Sect.
9.4 are considered applicable to fission depletions in UO2 up to
about 10× 10° fissions/cc. Requirement for further work and for
definitive experiments in this field is obvious from the preceding
discussion. In the absence of such, it is recommended that, for appli
cation to fuel element design, a diffusion coefficient for fission gases in
UO, be utilized at temperatures above 800° C given by the expression,
D=6.6×10−" exp (-71,700/RT) cm3/sec. At temperatures below
800°C, it is recommended that a temperature-independent value of
D=10-19 cm2/sec be used. º
9.4.8 Diffusion of Fission Gases in Other Oxides

(a) UO,...,
Other than the thoron emanation information discussed above,
little data have been reported on the diffusion of noble gases in oxides
other than UO2. A valuable investigation has been reported by
Lindner and Matzke on the effect of oxygen content on fission gas
diffusion and their results are plotted in Figs. 9.31 and 9.32 [77, 78].
Oxygen in solution in UO, is observed to increase the diffusion rate
of Xe” and Kr", but not, however, to affect the activation energy for
its diffusion. In fact, as shown in Fig. 9.32, the logarithm of the pre
exponential term, Do, in the equation D = Doe-9/* is observed to in
crease linearly with excess oxygen content. A similar variation of D.
with oxygen content for the diffusion of interstitial oxygen in UO.,
was demonstrated by Belle, who also found the activation energy for
diffusion of interstitial oxygen ions to be independent of oxygen con
tent [81]. The behavior shown in Figs. 9.31 and 9.32 is at first glance
rather surprising, since the lattice parameter of UO., decreases as
oxygen content increases (see Chaps. 5 and 6), and it might be antici
pated that the pre-exponential term and, hence, diffusion rates would
accordingly tend to decrease. However, the change in D, with oxygen
content in Fig. 9.32 is in the direction of rendering more positive the
entropy of activation for Xe diffusion; as discussed in Ref. 89, the
loosening of the lattice with increased oxygen content as evidenced by
the decrease in elastic modulus should make the entropy of activation
more positive [91]. The increased plasticity of UO, with increasing
oxygen content (see Chap. 5) is further evidence for a decrease in lat
tice binding energy [91]. More recently Auskern confirmed the in
crease of Krº diffusion rates in nonstoichiometric powder [92]. How
IRRADIATION EFFECTS IN URANIUM DIOXIDE 513

ever, he found the activation energy for diffusion in the UO, powders
reduced in hydrogen at 1,400° C prior to irradiation to agree with that
reported by Bostrom, Long, and Susko in Fig. 9.21, whereas the
subsequently oxidized powders showed activation energies approach
ing those reported by Lindner and Matzke and by Booth and Rymer.
Thus, the analogy with the case of diffusion of oxygen in UO2, both
stoichiometric and nonstoichiometric, is quite striking and probably
not fortuitous. These results would suggest that stoichiometry of
the oxide must be carefully controlled to ensure obtaining proper
diffusion data.
It will be noted in Fig. 9.31 that the oxide UAOs releases fission
gases at a much faster rate than even nonstoichiometric UO, and that,
furthermore, the activation energy for diffusion in this oxide has
dropped to about 19.7 kcal/mole. Thus, while compositional change
alone caused a relatively minor change in fission gas diffusion rates
in UO2, the crystal structure change on going from the cubic fluorite
structure to the orthorhombic lattice of UAO, profoundly changed the
nature of fission gas diffusion. It should be noted, as discussed in
—I I T T I I

O xe 133
- 12 H. G Kr 85
-
O D Rn222

-13 H. O -

-16 -

-17 H.

-18 H- -
UO2.0 (xe-133)
l —l l l L I
6 7 9

8 IO !I 12 13
loº/T (*K)

FIGURE 9.31. Temperature Variation of Diffusion Constant for Noble Gases in


UO,.. [77, 78]. (Courtesy, Zeitschrift für Naturforschung.)
514 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I I I I I I T I

-
:

1– | l L | | I 1a
2. Ol 2. to

Log Hº-jāśi
MouEs OxYoen
in Uo 2 + x

FIGURE 9.32. Variation of Do for Xe” diffusion in UO2... with Oxygen Con
tent [78]. (Courtesy, Zeitschrift für Naturforschung.)

Sect. 9.3.1 (b), that the exposure levels utilized to activate the U.O.
samples, about 7 × 101° fissions/cc., are sufficient to disorder the struc
ture so that, unless crystallization has occurred, the diffusion curves
in Fig. 9.31 would be characteristic of those for an almost amorphous
oxide. From the annealing results of Childs and McGurn, discussed
in Sect. 9.3.1, it is probable that recrystallization has occurred during
the diffusion anneals [45]. However, the structure formed on an
nealing was probably highly imperfect, which may also serve to
explain the low activation energy as well as the low Do value observed,
about 2.5 × 10−10 cm3/sec. At any rate, the results plotted in Fig.
9.31 illustrate the observation reported by other investigators of
greatly increased fission gas release rates as a result of specimen oxida
tion and emphasize the necessity for avoiding oxidizing conditions if
accurate diffusion coefficients are desired for the stoichiometric oxide.
It is of interest to note in Fig. 9.31, that the relative rates of diffusion
of Kr, Xe, and Rn in UAOs decrease in the order named, while the
activation energies increase, particularly that of Rn compared with
Xe. Lindner and Matzke related these observations to the differences
ſº IRRADIATION EFFECTS IN URANIUM DIOXIDE 515
I I- T T- T I

O – 2:U2- Uo.2- Cao


•- Al2O3-uoz
-
16.0 A- U02 -

o-
12

so -

i_{
O l l l l l l I
o - 80 160 24o 32O 4oo 480 56,o
TEMPERATURE,
*c

FIGURE 9.33. Emanation of Kr" from Various Oxide Fuel Materials [64].

in atom radius of the three gases (Kr-1.89 Å, Xe=2.08 A, Rn=


2.14 Å) [79]. On the other hand, it should be noted that the measure
ments on Rn were obtained in material undamaged by irradiation
and, thus, may not be directly comparable with the other data.

(b) Al,0,—UO, ZrO2–0a0–UO,

Markowitz reported in Ref. 64 the relative emanation rates,


shown in Fig. 9.33, as measured continuously in-pile as a function
of temperature from 100° to 550° C for samples of pure UO2,
Al,0–21 weight percent UO, (containing small UO, particles and
sintered to about 95 percent density), and ZrO2–13 weight percent
CaO, 17 weight percent UO, (consisting predominately of a ZrO2-rich
tetragonal phase and sintered to over 95 percent density). It will
first be noted that, in addition to demonstrating differences in the
temperature coefficient of release rate, the rate of release increases
order UO, Al.O,-UO, and ZrO,-CaO-UO, while release
the

in

rates purely by recoil mechanism should increase almost inversely


to a

proportionally density. further apparent that,


It

value
is

if
a

UO,
of

10-19 be assigned
of

the diffusion coefficient


to

at

550° (see
C

Fig. 9.21), the corresponding values for Al-O,-UO, and ZrO,-CaO


UO., from Fig. 9.33, would
be

about 2.2×10+" and 14×10−"


cm/sec, respectively. Note also, however,
that the maximum tem
which one would expect,
of

still below that


at

perature measurement
is

Fig. true volume diffusion effects predominate and


to

from 9.28,

that the results plotted Fig. may represent structure


in

9.33
a

574789 0–61—34
516 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

sensitive characteristic of the particular samples investigated. Re


cently, results have been obtained for highly dense platelets of ZrO.
containing 25 weight percent UO,. On annealing at 1,000° C after
slight irradiation, the release of Xe” was found to occur with a
diffusion coefficient about 250 times as large as that for UO, at
1,000° C."

(c) MgO-UO,

Stubbs, Silver, and Webster measured the rate of release of Xe”


from samples containing 71 weight percent MgO and 29 weight percent
UO, os and UO2.0 [93]. In the former case, the bodies were made with
0.1p UO, os powder, and in the latter with 100, UOzo powder. The
bodies were sintered to a density of about 75 percent of theoretical.
The fractional release of Xe” was found to remain constant at about
1 percent from 100° to 600° C. Above this temperature, the amounts
released increased rapidly and, at 800°C, almost all the Xe generated
in the UOzos–MgO sample was released, whereas about 20 to 25 percent
was released by the UO,-MgO samples. The authors, as well as Mur
ray and Williams, commenting on these data, concluded that the diffu
sion rate in MgO was exceptionally high and that essentially all the
fission products recoiled out of the UO, into the MgO rapidly diffused
away [94]. The authors also concluded that a different type of diffu
sional process occurred at 600°C to cause the observed acceleration in
release rate. These results are quite explicable also on the basis of
reasonable assumptions concerning the true particle size of the sinters.
If the curve of Fig. 9.18 be applied to estimation of the diffusion
path, a particle radius of 10-14 cm would be calculated. Such a par
ticle radius would predict the observed almost complete release of
fission product gases from MgO at 800° C if the diffusion coefficient at
800° C were of the order of 3 × 10-14 cm”/sec. As can be seen from
Fig. 9.28, this value is by no means inconsistent with other measures
of noble gas diffusion rates in metal oxides. The marked rate increase
at 600° C might indicate reaching the Tammann temperature, al
7",

C.

though a oxide melting 2,800°


an
of

low for
at

600°
is
C

(d) A7,0,-base Porcelains Confaining U(),

Stubbs, al., also measured the rate body consist


of

release from
et

ing percent UO, percent Al...O., percent SiO2, 1.3 percent


31

60
of

MgO, and 0.7 percent CaO sintered temperature such that liquid
at
a

high and,
In

phase formed. this way, density bodies were formed


as

expected, the release rates temperatures up


be

of at

to

would 800° were


quite low and beyond the sensitivity the experimental technique for
"J. Clayton, Laboratory, unpublished
C.

Bettis Atomic Power work.


IRRADIATION EFFECTS IN URANIUM DIOXIDE 517

separation of recoil and diffusional contributions. Postirradiation


measurements of release of Xe” at 800° C from these samples indi
cated values of
No. Xe” atoms released
=v c 1.4 X 10-13/sec.
No. Xe” atoms in sample-sec

of Thus, for
which corresponds to a value
# =9.3×10-1° cm”/sec.

large values of a, corresponding to the high specimen density, results in


reasonable accord with the data in Fig. 9.28 for Al2O3 can be derived.

(e) Diffusion in Glasses

It is finally
considered pertinent to compare the magnitude of the
fission gas diffusion rates in the crystalline solids of Fig. 9.28 with
those of noble gases in amorphous or glassy oxides. At temperatures
of 500°C, corresponding to the range of temperature at which noble
gas (He) diffusion has been measured in quartz and glass, Fig. 9.28
shows that in crystalline solids, D values are in the range 10° to
10-" cm°/sec, whereas in glasses the corresponding values range from
10-7 to 10-11 cm2/sec [86]. Thus, it is anticipated that in case of loss
of crystalline structure, release rates of fission product gases will accel
erate and, thus, limit fuel material usefulness. In any case, the consid
erations of structural stability in Sect. 9.3.1 must be considered to
have implications on fuel material behavior over and beyond dimen
sional stability alone. The connotation of such behavior with respect
to high burnup behavior of UO, will be discussed in Sect. 9.5.

(f) Summary
It may, thus, be concluded that, with respect to use of metal oxide
matrices other than UO, for containment of fission product gases, the
conclusions concerning effects of density and fission gas mobility de
rived above for UO, are probably valid for other oxides. Differences
of fission gas diffusion rates between various oxide bodies are, as in
UO2, a sensitive function of density, and conclusions relating to dif
fusivity are likely to be erroneous unless close control over density is
maintained. Orders of magnitude and relative rankings of various
oxides with respect to their fission gas diffusion characteristics may be
estimated by comparing thoron emanation (Fig. 9.28) data, cation
self-diffusion rates, etc., at a fixed temperature. Thus, at a tempera
ture of 1,000°C, diffusion rates of fission gases may be expected to in
crease in the following order:

UO. : 3.5 × 10−" cm2/sec (Fission gas diffusion)


ZrO2: 8.5 × 10−" cm”/sec (Fission gas diffusion)
518 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

MgO : 8.65× 10−1 cm2/sec (Cation self-diffusion)


Al2O. : 1.6×10−1 (Thoron diffusion)
cm−/sec
BeO: (Cation self-diffusion)
3.4 × 10-12 cm3/sec
TiO. : 1.2×10−14 cm3/sec (Thoron diffusion)
Fe O*: 2.5 × 10-14 cm2/sec (Thoron diffusion)
The limitations of these estimations should be obvious from the fore
going discussions.

9.4.9 Release of Fission Gases from Fuel Elements

(a) Relation between Fission Gas Release and Thermal Rating

The release of fission product gases from UO, fuel elements is com
plicated by the presence of steep thermal gradients existing in the
fuel as a result of its poor thermal conductivity. Thus, the major
proportion of the total fission gas released in an operating UO2 fuel
element will originate from the central, high temperature portion of
the fuel. Prediction of amounts of gas released and comparison with
available data, thus, require at least as accurate a knowledge of tem
perature conditions existing within a fuel element as of the fission gas
release characteristics of the fuel. As will be shown below, the thermal
performance of fuel elements can be characterized by the parameter

ſ
*center
Kd T
* surface

where Kis the thermal conductivity of the fuel. To illustrate the


effect of
temperature gradients, from the irradiated thermal conduc
tivity data of Fig. 9.15, using assumed values of Teenter=2,750° C
and Tournace =400° C and, therefore, a value of Kd7 =54.5 w/cm,ſ
the gas release curves of Fig. 9.34 were calculated for a cylindrical
UO, pellet utilizing two different assumed exposure times, 1,000 and
10,000 hours, and two sets of fission gas diffusion data from Fig. 9.21,
those of Susko and those of Booth and Rymer. In Fig. 9.34, for
each of these conditions the fractional release from each unit volume is
plotted as a function of radial position; also, for each radial position,
the cumulative fraction of the total amount of gas released within the
fuel element from the surface to the radial position is plotted. These
curves illustrate a number of important and characteristic features.
It is immediately apparent that 50 percent of the total gas released in
the fuel element is released by less than 20 percent of the fuel located
at the pellet center, and 90 percent by less than 40 percent. It is
further evident that the fraction of gas released varies over very wide
limits across the cross section of the pellet and that large composition
IRRADIATION EFFECTS IN URANIUM DIOXIDE 519

RADIUS
PELLET RADIUS
O O2 O3 O4 O.5 O.6 O.7 O.8 O9 I.O
T I
cEnter" i I
' surface

-
l,OOO 10,000
HOURS HOURS
FRACTION RELEASED AT EACH
UNIT VOLUME

–O- - - - - CUMULATIVE
RELEASED
FRACTION —H O.2

DATA ON "D"FROM FIG. 9.2


B AND R-BOOTH AND RYMER DATA
S–SUSKO DATA
DENsity, 95% ("d"- 120p.(FIG. ))
2750°C

J/ kat-54.5
4oo” C
WATTS/cm (FIG. 9. 15)

O O.2 O.4 O.6 * O.8 I.O


RADIUS
( PELLET RADIUS y
FIGURE 9.34. Variation in Stable Fission Gas Release Rate with Distance from
Center of Cylindrical Pellet.

gradients of volatile species are to be expected in fuel elements of high


thermal ratings. Likewise, it may be noted, in spite of the large
differences between the diffusion coefficients measured by Susko and
by Booth and Rymer, that the total amount of gas released and the
distribution of gas release through the pellet are not significantly
affected by the particular choice of measurements. Of considerably
greater effect are the temperature of exposure and the operating life.
It will be noted from Fig. 9.34, for the case of uniform fissioning
rate through the cross section of the cylindrical pellet, that the frac
520 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

tion of fission gas released at each location as well as the cumulative


fraction released in the total cross section is independent of the abso
lute value of the pellet radius for a fixed value of the parameter

Teen r - -
"KTä. Since this parameter is proportional to fissioning
J.T surface
rate

per unit volume and to the square of pellet radius, it is also true that,
for a cylindrical pellet, the number of gas atoms released is independ
ent of pellet radius for a constant length and constant value of the
parameter ſ
Kd7". This relation does not hold for fuel elements in the
form of plates, in which the number of gas atoms released at constant
values of the parameter ſ
Kd7 increases inversely proportional to the
however, in

is,
plate thickness; the cumulative fractional release
dependent
of

thickness.
As these relations, possible
of

define the fractional

to
result
it
is
a

fuel element samples the thermal rating param

of
in

release terms
in

eter Kd7". Robertson, al., have calculated the fractional release


et
ſ

cylindrical fuel element for various fuel element thermal


of

Xe
in
a

(in watts per cm length)


as
parameters

of
function sec

in
(t
time
a

1,400°C for particular sintered oxide,


of

at

onds) and the value


#

by
assuming the variation diffusion rate with temperature given
in

Booth and Rymer [76, 95]. The results


of

these calculations are


plotted Fig. 9.35 and may predict fission gas release
be
in

to

used
rates in such fuel elements.
The procedure utilized the calculations for Fig. 9.34 which
in
in

the temperature, diffusion coefficient, and fractional release for each


volume element are separately computed can yield accurate estimates
of

the fractional release from fuel element. The procedure quite


is
a

laborious, and Booth has suggested


an

alternate procedure which


in

fractional release estimated using the Simpson three-point approxi


is

mation rule [73]. He, thus, suggests that the fractional release from
by

the entire element, Fror, given


be
1

Fror=; (Fr.--4Fr.--Fr.) Eq. (9.25A)

where Fr., Fr, Fr, are the fractional releases the temperature
at

and
(T),
(T-":") and surface (T,) the fuel,
of

of

the center median

respectively. For the particular case plotted Fig. 9.34, the two
of in

methods give the following comparative values fractional release:


IRRADIATION EFFECTS IN URANIUM DIOXIDE 521

Volume summation
Fror Booth (73) (Fig. 9.34)
SUsko DATA
10,000 hr 0. 269 0.213
1,000 hr 0. 144 0. 111
Booth AND RYMER DATA
10,000 hr 0.381 0.254
1,000 hr 0. 174 0. 128

While the agreement between the two techniques is not strikingly good,
the accuracy of the approximation is probably adequate in view of
the uncertainties in temperature estimates and fission gas release data.
Equation 9.25A may be used to define the average diffusion coefficient
for the fuel element, Dror, as follows:

Eq.
VDºr- (WD-4-4WD-4AD.) (9.25B)

estimates have been reported comparing calculated


of
A

number
of

values fractional gas release from fuel elements with those observed

IO

40
|

50

|O'

alº, 10° ºſ
`-
§§

surface TEMPERATURE OF
oxide Assumed 400°
C

S$’
03
o°g
i

DIFFUSION RATE cm2/SEC


=
D

AT |400°
C

*
5

10° -

105 —l —l
!

IO 2O 3O 40 50 6O
O

F.P. XENON RELEASED, PER CENT

Fractional Release of Xe Cylindrical UO, Fuel Element as


in

FIGURE 9.35.
a

Function of Thermal Rating, Time, and Diffusion Rate [95]. (Courtesy,


Mining, Metallurgical, and Petroleum Engineers, Inc.)
of

American Institute
522 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

o I i
ExPOSURE TIME DENsiTY
76 DAYs 237 DAYS º THEORETICAL
O O 86
D 89
A A 92
- H V w 95
O O 96
© & 96 (O/U=2.12)
ge BB 9 I (O/U=2.12)

-2 |

-3 -

I I l l
-5
-5 -4 -3 -2 -I O
LOG FRACTIONAL RELEASE, CALCULATED

FIGURE 9,36. Calculated and Observed Fractional Fission Gas Release (Kr”).

experimentally upon puncturing fuel elements after their in-pile ex


posure. A particularly suitable experiment for making such a com
parison is that initially reported in Ref. 64 and described in
Ref. 96. In this experiment, a number of fuel element rod
samples containing sintered oxide varying in density from 86 to 96
percent of theoretical, exposed at heat fluxes varying from about
40,000 to over 300,000 Btu/hr-ft” for two in-pile exposure periods
(76 and 237 days), and
for burnups between 1,000 and 13,000
MWD/Ton UO, were punctured
after irradiation and the amount
of gas released measured. In Fig. 9.36, these measurements have
been compared with those calculated utilizing the gas diffusion data
of Susko (Fig. 9.21) and estimates of surface area from Fig. 9.18.
The agreement between the observations and calculations is considered
to be satisfactory. However, this agreement is not cited to validate
one set of diffusion measurements over another; as discussed above,
es: the

estimates of fractional amounts of release are quite insensitive to


by

particular values chosen and, fact, are much more affected


in
IRRADIATION EFFECTS IN URANIUM DIOXIDE 523
22
I-I-T-I-T-I-TI
U02 DENSITY, 10.64 g/cc
H + (97%. T.D.) ANNEALED AT 1400 °C FOR -
3 HOURS BEFORE GRINDING
LOSS DURING ANNEAL •O.2*

14
II
12
NO ANNEAL BEFORE
GRINDING
-
IO

-
i
CURVE I-CURVE II
ASSUMED TO BE RELEASE
6 OF xe” DIFFUSED TO -
CLOSED POROSITY DURING
4. ANNEAL

o l l l I l I l l
O 40 80 12O 16O 200 240 280 300
GRINDING TIME IN MINUTES

FIGURE 9.37. Release of Xe” on Grinding Irradiated Sintered UO,.

timates of
center temperatures attained. Rather, Fig. 9.36 illus
trates the predictability, to a high degree of confidence, of the behavior
of UO2 with respect to its fission gas release characteristics in op
erating fuel elements, at least at burnups up to 25,000 MWD/Ton.

(b) Effect of Closed Porosity

A number of factors may contribute to departure from the relatively


simple diffusional behavior outlined above for release of fission prod
ucts. Possibly the best demonstrated ofthese effects is the diffusion
of fission products to closed pores formed in sintered bodies. From
the dimensions of closed pores formed in highly sintered UO, bodies
(about 2p in diameter), and the amount of closed porosity measured
in such sinters, it may be estimated that, in high density compacts,
the surface area of the closed pores may be as much as 1,000 cm”/cc
[74]. This, from Fig. 9.18, equals the externally communicating
surface area of a 90 to 92 percent dense compact. Thus, it may be
anticipated that the fraction of gas diffusing to closed pores will be
comparable to, or even greater than, that released to the external sur
face of a high density compact. This effect was well illustrated by
Stevens in Fig. 9.37 in which the amount of gas release upon grind
ing 97 percent dense irradiated UO, both before and after annealing
3 hours at C is plotted as a function of grinding time." About
1,400°
8 percent more of the Xe” in the oxide was released during grinding
after the anneal than prior to annealing. If this release truly repre
* W. H. Stevens, Chalk River, unpublished data.
524 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

sents diffusion to closed pores, then the magnitude of the latter is a


factor of 40 greater than that (0.2 percent) released externally
during the anneal. Since the fractions released should, by a diffu
sional mechanism, be proportional to surface area, the measurements
would predict, based on a closed pore surface area of 1,000 cm”/cc,
that the external surface area of the 97 percent dense compactis about
25 cm”/cc, a value not inconsistent with that observed in Fig. 9.18.
Gas absorption in closed porosity should decrease the amount released
external to the sintered body and should be more pronounced the
r”Dit
higher the density. For fractional gas releases such that
Q?
s1
(see Eq. 9.22), or less than 10 percent of that formed is released,
the closed pores can be expected to be sufficiently separated from the
external surface that the rates of diffusion to the two types of surface
are independent of each other.

(c) Influence of Fission Fragment Recoils

Lewis considered the consequences of other mechanisms for fission


product release and mobility [63]. These are related to the effects
of fission product recoil processes near free surfaces of the fuel body.
For example, many of the fission-born nuclides which later decay to
form the noble gases originate as a different chemical species which,
if at rest in free or void spaces in the fuel body, can adsorb at the fuel
surface. Subsequently, the process of 3- or y-emission accompanying
decay can drive the daughter product into the fuel matrix similarly
to the process described in Sect. 9.4.5(a). Such nuclides, tenuously
trapped at the free surfaces of the body, may be responsible for the
observations of the initial bursts or rapid rates of release observed
upon heating preirradiated samples for fission gas release studies
(see Figs. 9.23, 9.29, and 9.30).
Another phenomenon is that of fission product knockout by recoil
phenomena in the solid near free surfaces. Laptev and Ershler have
measured, as discussed in Sect. 9.4.2(a), about 24 uranium atoms
evaporated from the surface of oxidized U* for each fission event
within recoil range of the surface [66]. Lewis estimated that this
factor may be as high as 45 UO, molecules for each fission event [63].
The fission product atoms associated with this volume of material
would be expected to vaporize, enter free space, and, thus, contribute
to the amount of gas external to the solid body. It is not anticipated
that such contributions would be large in the case of volatile species,
since the volume within which recoils could cause knockouts to the
open void volume would already be depleted by diffusion. Such
processes may, however, amplify the recoil release of nonvolatile or
insoluble species.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 525

Still another process has been postulated by Lewis, namely, that of


return of fission gas released to open void volumes by knockons be
tween recoiling fission fragments and gases. He has estimated that
the number of fission gas atoms left embedded in the oxide per fission
will be
PA 273\ V,
GN
22,400 Ps (#) V.
where p, is the density of the solid oxide, V, and V, are the volumes of
open porosity and solid oxide, respectively, P is the pressure in atmos
pheres of gas of mean atomic mass A and temperature T in * K filling
the open porosity, N is the number of atoms knocked on with sufficient
energy to be embedded were the fission track entirely in the gas space,
and G is the fraction of these actually becoming embedded. From this
relation, Lewis has estimated that if as few as four fission gas atoms
per fission become embedded when the pressure of fission gas reaches
100 atmospheres, the rate of return by knockons will exceed the rate
of release by diffusion at temperaturesbelow 1,200° C. The signifi
cance of this effect is that it predicts
a limiting fission gas pressure
which cannot be exceeded during steady-state operation and further
predicts a redistribution of volatile species to a higher concentration
than anticipated in Fig. 9.34 in the cooler portions of the oxide.
Lewis estimated that the actual value of GM, the number of fission gas
atoms embedded per pair of fission tracks, was 100 atoms. Such limi
tations on fission gas pressure developed would be more important the
its

lower the density of the oxide and, hence, the higher


of

rate fission
by

gas release diffusion and also the weaker the cladding and the less
capable resisting high internal pressures. Proof this model re
of

of

quires the performance sophisticated and difficult experiments.


of

is by
by

quoted undoubtedly indicate entrapment Xe


of

Results Lewis
of
of at

recoil phenomena low exposures.


One the unsolved questions
recoiled atoms compared with the
of

the relative rate embedment


of

their re-emission appreciable concentra


at

rate the recoil zone


in

Experiments performed pumping


on

noble gases
of

tions. the ionic


by

metallic surfaces have revealed that the pumping rate (under im


pumping
of

as

pressed energies about 100 electron volts) decreases


proceeds spontaneous previously pumped
of

of
as

result re-emission
a

gas [97]. This limitation pumping can arise either from re-emis
to

resulting sputtering
or

the target material the trapped


of

sion from
the sites suitable for trapping gas. The
or

gas
of

from saturation
ionic pumping con
of of

of

saturation
in

at

case metal substrates occurs


The ef
or

trapped gas equivalent monolayer


to

centrations less.
a

fect of such saturation or re-emission would be the


in

decrease
a

a
of

the equation above.


in

value
G
526 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

(d) Influence of Pore Migration

Another physical method for securing fission gas release from sin
tered UO, bodies is discussed by Lewis. This is the removal of gas
contained within closed pores by transporting the pores to surfaces
of open porosity under the action of thermal gradients. It has been
demonstrated that under such gradients UO, may evaporate from
a hot pore surface and deposit at a cold surface, thus, resulting in a
gross movement of the pore toward high temperature regions. The
result of such a process would be an enhanced rate of release from UO.
operated in a temperature gradient as compared with isothermal
measurementS.
It is difficult to evaluate the significance of the physical effects
discussed in Sects. 9.4.9(b) to (d) primarily because of the lack
of unequivocal critical experiments and also because of the poor preci
It

is,
sion of the experimental data presently available. thus, deemed
advisable, until confirmatory data become available, utilize the

to
diffusional treatment for prediction fission product release rates
from fuel elements discussed Sect. 9.4.9(a). of
in

Nonvolatile Fission Product Isotopes


of

9.4.10 Release

Mo”
of

(a) Release

The only direct measurements


of

of

the release nonvolatile fission


by

product species from UO, are those reported Shiriaeva and Tol
machev [98]. These authors heated irradiated powders UO, and of
of

preparation the powders nor their


of

UAOs (neither the method


physical characteristics were specified) atmospheres oxygen,
of
in

hydrogen, argon, and vacuum and measured the evolution Mo"


of

from the oxide matrix. Libby has shown that insoluble atoms
in

ionic lattices will migrate free surfaces [99]. The driving force
to

by

for the migration the energy opposed


of

solution which the


is

is

entropy migration given by


of

lº. Eq. (9.26)


=
P
N

4
*
*

where M, surface sites, M, the number


of
of

the number lattice sites.


is

Thus, the rate and amount migration


or of

of

insoluble atoms favored


is

by fine particle size. Little


no

cognizance
of

this effect has hitherto


analysis fission products from UOs.
of

of

been taken the release


in

Its significance determining the effects contained porosity, grain


of
in

size, etc.,
of

oxide fuel bodies merits further attention.


The Mo" segregated the oxide particles was meas
of

the surface
at
by

an

ured exchanging with ammonium molybdate solution which


in

the powders were swirled for several hours. That vaporized during
!
IRRADIATION EFFECTS IN URANIUM DIOXIDE 527
ATMOSPHERE OF .
-
O HYDROGEN
-
O ARGON
A - VACUUM
^ - OxYGEN
O-PER CENT REMOVED BY LEACHING
D - PER CENT VOLATIZED
- - TOTAL PER CENT REJECTED

TIME IN HOURS, VACUUM AT 1200°C


O I 2 3 4
I

4OO 5OO 6OO 7OO 800 900 |OOO |IOO I2OO


ANNEALING TEMPERATURE, *C

FIGURE 9.38. Percent Evolution of Mo” from Irradiated UO, Powder after 1.5
Hours Annealing at Various Temperatures and after Annealing Various Times
in Vacuum at 1,200° C [98].

heating was measured by collection of the distillate at a cooled end of


the tube in which the heating occurred. The results obtained are
summarized in Figs. 9.38 and 9.39. The former plot shows the vari
ation in amount of evolution of Mo” from UO, at various tempera
tures after 1.5-hour anneals in the various atmospheres. As would be
anticipated, minimum evolution from the matrix occurs in hydrogen,
somewhat more in argon, and almost complete separation is noted in
vacuum and oxygen atmospheres after the 1,200°C anneal. Also
plotted in the same figure is the distribution of the Mo" between that
segregated at the UO, surface and that volatilized as a function of
heating time in vacuum at 1,200° C. After only a few minutes of
heating at 1,200°C, almost 80 percent of the molybdenum has segre
gated at the particle surface from which it is volatilized at a much
slower rate. The ability to volatilize this isotope implies that, under
the vacuum conditions employed, sufficient free oxygen was present to
528 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

20 I I I i i
Atmosphere of:
o HYDROGEN
A VACUUM
...tºr-
1.6
c D oxygen
Lu
2.
c
>
tau
• 2 H
on
o
-:

2 88OO
§ os P
or
uu
a.
3 O
O
-O.
-1
O4 H.
400
-

O L l I _l L l l I
6 7 8 * 9 Io il 12 is 14 15
(ABsoLUTE TEMPERATURE -*k)' x 10°

FIGURE 9.39. Variation of Percent Evolution of Mo” from UO, with Temperature;
Numbers on Curves Are Activation Energies (kcal/mole) [98].

permit evaporation of a volatile oxide of molybdenum. This is borne


out by the small amount of separation noted upon annealing in hydro
gen in which formation and subsequent evaporation of volatile oxides
of molybdenum would be prevented. In Fig. 9.39, in which the
logarithm of the fraction evolved is plotted versus the inverse of
absolute temperature of annealing, it is notable that a change in activa
tion energy for diffusion occurs at a temperature quite close to that
noted for the similar change in case of rare gas diffusion from UO.
(Fig. 9.21), about 1,000 to 1,100° C. However, the very low activation
energies for Mo evolution at temperatures above 1,000° to 1,100° C are
inconsistent with those anticipated for diffusional processes in UO.
and argue for a considerable degree of imperfection in the oxide prepa
rations employed or for the operation of other than diffusional proc
eSSes.

(b) Release of Fission Products into Coolant from Defected Fuel


Elements

Other measurements of the release of the nonvolatile fission product


isotopes from irradiated UO, have been obtained from observation of
the radioactive contamination of hot-water loops during the operation
of defected fuel elements. Measurement in-pile of the distribution and
amount of release of radioactive species to the coolant from fuel ele
ment samples exposed to the coolant is of considerable technical sig
IRRADIATION EFFECTS IN URANIUM DIOXIDE 529

nificance, since such measurements permit assessment of the degree of


contamination to be expected in reactor plant systems from operation
with defective fuel elements and, furthermore, permit detection and
isolation of fuel element assemblies containing such defects [100].
Furthermore, as will be shown below, some information can be gleaned
from such tests concerning the mobility of nongaseous nuclides.
Doubts arising from lack of reproducibility and perturbations aris
ing from operation of various physical phenomena other than solid
phase diffusion, which were seen above to plague measurements of
noble gas mobility, can be expected to be magnified manyfold in case of
measurements of release of fission products into the coolant from
defected fuel element samples. The use of a fuel element in which the
temperature of the fuel may vary over wide limits requires the use of
approximations such as Eq. 9.25 A and B to set a suitable average
release rate. Since measurements of release rates have, in general,
been made from clad samples exposed to the coolant only through a
small hole in the cladding, it is essential to assure that the release rates
are controlled by mobility in the solid phase rather than in the steam
or water-filled interstices between fuel and cladding. All the measure
ments hitherto reported have been obtained from defected samples
operated in high temperature water loops. Therefore, phenomena
such as change in composition of the UO, by reaction with water (see
Sects. 9.4.8(a) and 9.7.2(c)) and sweeping out of the volatile compo
nents within the cladding during pressure or temperature changes can
seriously affect the interpretation of the results. Finally, since a rela
tively dense medium such as water or high temperature steam sur
rounds the fuel within a defected fuel element under the conditions of
these tests, it is necessary to determine the mechanism of fission prod
uct release from the fuel, whether by recoil or diffusional phenomena.
(1) MATHEMATICAL DEscRIPTION of RELEASE FROM DEFECTED
FUEL ELEMENTs. The activity of the coolant has been related to
the rate of release from fuel by a number of investigators and is
discussed below, following the treatment of Ehrenreich, Frank, and
Vogel [100, 101]. The total amount of isotopes in a fuel specimen,
N. (atoms/specimen), is given by

*.*.ſvry-º'-(2,4-x)N. Eq. (9.27)

where f= a, q [U”), in which fis fissioning rate, a , is fission cross


section, q is perturbed neutron flux at full power, and [U”] is the
concentration of fissionable isotope in atoms/cc; V, is the volume of
fuel in specimen, ce; r is the ratio of equivalent time at full power to t,
total time in seconds from start of irradiation; oa is total cross section
of U2°; Y, is the fractional yield of isotope i, v, is the escape rate
530 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
.
coefficient (Eq. 9.17F); and A is the decay constant, secT". The
solution of this equation is

N-- v,
fV,ry,
-HX-0.4%r
| e-oaert–e-(vit\) ] +Nºe-(º-tºot
i

where N' is the number of nuclides i in the specimen at t=0. Neg


lecting the fuel depletion and reduced power factor, r, important only
for long half-life isotopes,

fW,Y,
N=#; [1–e-(º-tºº--Nºe-tºot Eq. (9.27A)

The isotopic concentration in the loop water, [N], atoms/ml, in a


system of total volume V, ml, possessing a combined purification and
leakage rate constant 3, sect', is

dIN'ſ
di TV
PºW
(3+\)[NAE] Eq. (9.27B)

j
Substituting 9.27A in 9.27B and solving:

'l- "if V.Y. - fV, YA


[N]-voiº-Vºx)| v. (v-HA,)t
: Fx-Nº ||
wif
+[IN tWüíxīā-y)TVää)]"
7/10 V.Y, - v. Nº –(8+\, t

,-º
for steady-state conditions,

'1– wif|V, Y,

Or

º
(since [N] v,V(3+A).<f V, Y.)

Since X. [N,'] is the specific isotopic loop activity, A., dissec-ml,


and assuming the purification and leakage rate constant has been
allowed for,

Eq. (9.27D)

Or
Y,
_” i fW,
4-, *f;
To determine the relative characteristics of the loop activities ob
served from recoil or diffusional phenomena, v, is related to each of
these mechanisms as follows. For a recoil mechanism, Eq. 9.27 be
(*Onles
IRRADIATION EFFECTS IN URANIUM DIOXIDE 531

N Y, V

using the expression for recoil escape, in which K is the ratio of


geometric surface area to the volume of fuel V, in the specimen. At
steady state, and neglecting the second term in comparison with the
first,

N, _f V.Y.
M

Ry comparison of analogous terms in Eqs. 9.27 and 9.28, vi for recoils


then becomes,

-*[.(2-4)]
v=-1 l. ls.
2 Eq. (9.28A)

–ſºſ
and
h __h
A=***** T., (2 ls. )]
Eq. (9.28B)

Similarly, for a diffusional mechanism, it was shown previously that


for cases where v.3t (i.e., for short-lived isotopes):

1N,
Eq.
--ºw a”
(9.17B)

Therefore,

A—ºyº Eq. (9.28C)

Thus, as shown by Robertson and Allison, these expressions afford


the opportunity of determining, from the relative activity of various
isotopes released to the coolant, the mechanism of the release
process [102]. For recoil processes, from Eq. 9.28B, the ratio of
activities for two isotopes of equal recoil lengths should be propor
tional to their fission yields,
A. Y.
A,TY,
whereas, for diffusional processes, assuming that the isotopes com
pared have the same diffusion coefficients in the solid oxide, the
relative decay constants, as well as fission yields, will determine their
relative activities,
A. Y. X,
A,TY, WX,

Itis apparent that diffusional release favors the escape of long-lived


species as compared with escape by recoil.
57.4789 O—61—35
532 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

In Table 9.9, ratios of activities of various isotopes noted in opera


tion of defected fuel elements are compared with the ratios anticipated
from recoil and from diffusional processes. It may be noted that,
for the isotopes listed, in most cases the results agree best with the
ratios predicted for diffusional release.

TABLE 9.9—RATIOS OF FISSION PRODUCT ACTIVITIES IN SEVERAL

º,
DEFECT TESTS OF UO, IN HIGH TEMPERATURE WATER LOOPS

Fxperimental observations
Ratio for
Nuclides compared recoils Y: ſ\;
Yi Yi Y; W A Ratio Test Reference

Xel33/Krºs . . . . . . . 1. 8 12 5 X–2—f 102


13 X–2–f 102

Krºs/Cs138 ----- --- 0. 6 2–1. 4 2. 5 X-2–f 102


1. 5 | X-2–f 102
1, 2 | X-2–i 102
2. 1 || WAPD–29–1 103
1. 2 | WAPD–29–2 103
4. 2 | X-1–1 103

II*/I133 0. 45 1. 4 1. 5 | X-2–f 102


2. 5 | X-2–f 102
0. 1 || WAPD—29–1 103
1. 0 | X-1–1 103
Brº/Brº - . .. . 0. 57 1, 2 4. 6 | X-1–1 I (14

(2) EFFECT of Fission RATE. By reference to Eqs. 9.28A and


9.17B, it is apparent that the escape rate coefficient v should be unaf
fected by fissioning rate in case of recoil processes, but should be sen
sitively influenced in case of diffusional release through the effect of
operating temperature on diffusion rates. This effect was best shown
by Ehrenreich in two tests (WAPD–29–1 and WAPD—29–2) contain
ing, respectively, 93 and 97 percent dense defected UO, cylindrical
fuel samples 0.413-inch clad diameter, 0.357-inch fuel diameter [103).
In each test the samples were operated for a number of days at a clad
surface heat flux of 230,000 to 290,000 Btu/hr-ft- and for an
approximately equal length of time at a lower flux, 21,000 to 27,000
Btu/hr-ft". The times were, thus, sufficiently long to establish a
steady state rate of isotope release. The values of v for Krºs and
Cs” decreased by factors of 13 and 20, respectively. in the first test
when the flux was lowered and by factors of 5 and 8 in the second test.
To determine if such reductions are reasonable, it should be noted
that the factor v in Eq. 9.17B is proportional to VD, in which
the diffusion coefficient is a mean value for the fuel element given by
an approximation such as that in Eq. 9.25B, i.e.,
IRRADIATION EFFECTS

(T.
- -
IN URANIUM DIOXIDE 533

DE)
1
V v=g V.H.,
-— Dr. |Dr.

ſº
V%t

,
+4
=5 WDr.

ſ
If, for purposes

all
approximation, of the brack

in
but the first term
ets are neglected, then for the particular test under discussion, the
escape rate coefficients should vary shown below:

as
D
:-yDr.,
V2
‘’’, Eq. (9.28D)

the particular conditions


of

For the WAPD–29–1 and WAPD-29–2


in which diametral clearance of 0.003 0.005 inch was utilized

to
tests
a

between fuel and cladding, and the water coolant temperature was
550°F, estimated that center temperatures would
be

be
can attained
it

exposure 3,300° (1,535° K)


of

of
under the two heat flux conditions

F
and 770° (685° K). Thus, utilizing, from Sect. 9.4.7, Disas
F

=4.3 10−" and Dass=10-1” cm3/sec, reduction escape rate coeffi


in
×

a
by

expected from the change


20

be
of

factor

in
cient about would
a
by

fair agreement with the experimental


of

flux factor
in

heat 10,
a

observations.
DENsity.
of

of

(3) EFFECT Another characteristic diffusional


sensitivity density
be

which has
to

to

escape has been shown oxide


the diffusional path (Fig. 9.18). Again the experi
to

been related
a

ments WAPD—29–1 and WAPD–29–2, which defected and


in

of

sintered samples density percent


of

undefected 93.5 and 97.4


density, respectively, were exposed comparable heat
at

theoretical
flux levels, best serve detect the operation this effect. From Fig.
of
to

9.18, the two different sinters should yield difference the factor
in
a

Kr" from the undefected samples


of

of

about 10. Release this


in
a

punctured after this exposure, was found 0.3 percent and


be
to

test,
percent, respectively, that formed, agreement
of

with the
in

0.02
density difference [64]. the escape rate coef
From Eq. 9.17B,
ficient for various isotopes the defected samples should also vary
in
by

this same factor. For the two tests question, values


of

escape
as in

by

rate coefficient are listed Table determined Ehrenreich


in

9.10
103]. While the results are fair agreement with the prediction for
in
|

those species which are volatile relatively soluble high tempera


or

in

ture water (Kr, Cs, I), trend toward higher escape ratio with
no

Ce

Zr, indicat
or

decreasing density detectable with species such


as
is

by

ing that their release may


be

influenced factors other than diffusion.


534 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 9.10—ESCAPE RATE COEFFICIENTS FROM OXII).ES OF WARIOUS


DENSITY IN WAPD–29–1 AND WAPD–29–2 TESTS [103]

Test WAPD-29–1 WAPD-23–2

Fuel density (percent theoretical) -- --- --- -- --- 93.4 97. 4


Estimated central temp (° C)---- - - - - - - - - - - - - 1, 260 1, 530
Fissioning rate of defect specimens (fissions/
see).------------------------------------- 1. 78 × 1014 2, 20 × 101"
Escape rate coefficient (vs, sec−1) for isotope:
(from loop water activities)
Kr”----------------------------------- 2.92× 10-7 0. 465 × 10-7
!"------------------------------------ 1.9 × 10−9 0.293 × 10−3
Cs”---------------------------------- 2. 16X 10-7 0. S65 × 10-7
Cs”---------------------------------- 2, 67 × 10-8 2.89 × 10-4
(from purification resin analysis)
Sr”----------------------------------- 6. 62 × 10-11 2. 32 × 10-in
Ba”-------------------------------- - - 2. 24X 10-12 9. 4 × 10-1:
Ce"---------------------------------- 3. 0.4 × 10-12 8. 4 × 10-1:
Żr"----------------------------------- 2. 2 X 10-12 26. 5 × 10-11

These same effects were correlated graphically by Robertson, et al.,


who plotted the experimentally observed value of escape rate coeffi
cient as a function of UO, density (Fig. 9.40) [95]. For each test,
the thermal rating of the fuel element sample (in each case a cylindri.
cal rod) is shown in watts per centimeter of length beside the plotted
point. Since, as shown previously, escape by diffusion occurs predom
inantly at the center of the fuel, v should increase with increasing
values of this factor, which is proportional to center fuel temperature
attained. As may be observed from the plot, the value of v for Krº
does, indeed, show a slight trend toward decreasing values of v with
increasing density and with decreasing values of center temperature.

ſ
"center
or thermal rating, KdT, thus, confirming the diffusional mech
surface
anism of release for this fission product species.

(4) CoM PARIsoN with DIFFUsioS MEASUREMENTs. Utilizing the


relation of Eq. 9.17B, it is possible to make a quantitative compar
ison between the diffusion rates calculated from activity release from
a defected sample and those observed experimentally in diffusion rate
measurements such as those reported in Fig. 9.21. If this be done
for a gaseous fission product specie, relative values of diffusion rates
for nongaseous nuclides may then be readily derived. In Table 9.11,
diffusion rates have been calculated from escape rate coefficients for
various noble gas species for three tests, WAPD-29–1 and 29–2 and
IRRADIATION EFFECTS IN URANIUM DIOXIDE 535

E-i i I I i I I T T
Ç O © x-i-c (28)
- x-2-f-3 (13) -
T ~ -
-
*~
Y~*
x-1-f(37) O
-
*^
HC X-2-f-4 (12) S< -
S < ^. x-1-n (35) O

--
*~
16' H >< x-I-hſ 33)
|- O x-2-m (30) Y ~ *- O
H -
I x-2-Q (48) 0 Y~
*~
*~
-
I >< I
|-
x-I-q(33) O –
|- X-1-1(24) e x-1-4137) e –

-
->
to
; 16"|-
o P- I
ul -
92 H -
>
ió”


- Rºgºi
figures parentheses
I
Hear
OFin

Fuel ELEMENT-ſkøj
-
surfacewatts /cm

RATING
T

EEC-8 (25) O
- -

—1–
1
1

18°
I

I
l

9.6 9.7 9.8 9.9 IOO IO. IO.2 IO.3 IO.4


I

U02 DENSITY, g/cc

Escape Rate Coefficient with UO, Density [95].


of

FIGURE 9.40. Variation


(Courtesy, American Mining, Metallurgical,
of

Institute and Petroleum


Engineers, Inc.)

X–1–1, and the data compared with the value for diffusion coefficient
by

Kr" measured Susko and plotted Fig. 9.21 [64, 81, 103].
It
in
of

noted that the agreement, while not strikingly good, lies within
be

may
results obtained from the various experiments plotted
of

in

the scatter
Fig. 9.21. Thus, this agreement may additional proof
be

as

taken
noble gases from defected
of

the diffusion mechanism for the release


of

specimens.
(5) DIFFUsion RATE FROM EscAPE RATE MEASUREMENTs. From
Eq. 9.17B, diffusional escape, comparison
of

of

the case the


in
536 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 9. 11–IDERIVATION OF DIFFUSION COEFFICIENT FOR Xe AND


Kr FROM ESCAPE RATE COEFFICIENT DATA (64, 95)

Test WAPD–29-1 (103) WAPD- X-1–1 [54]


29–2(103)

Isotope-------------------------- Krºs Xeiss Krss Krs? xeim


\(secº')-------------------------- 6.88x10-3 6.8x10-4 6.88×10-5 6.88x10-5 1.52x10-4
" (cm)--------------------------- 4. 44x 10−3 4. 44x10-3 7×10-2 4. 44x10-4 4. 44x10-?
*(secT')-------------------------- 2.92×10-7 1.78×10-7 0.465x10-? 2.1.x10-5 1.7×10-8
D(cm2/sec)----------------------- 2.76×10-15 1.04×10-15 1.71×10-14 1.42×10-17 4, 22x10-is
Center temperature (°C).------- 1,260 1,260 1,530 1,700 1, 700
D for Krº" (cm2/sec)------------- 1.1×10-16 1.1×10-18 4. x10-15 1.3×10-14 1.3×10-14
Burnup M W D/T-UO2 --------- 860 -------------- 750--------- 13,600 --------------

*Ref. 81.

2
values of
* for a given fission product specie with that for a noble
gas should yield an estimate of the diffusion coefficient of the specie
relative to that of the noble gas unaffected by differences in tempera
ture or diffusional path (density). This comparison is presented for
several in-pile defect tests in Table 9.12. Several interesting infer
ences may be drawn from this table. It all, apparent that is,

of
first
gases highest diffusion rates, while Cs, Te, and

I,
the rare exhibited the
Br diffused only slightly rates, Group still
at

II
slower the elements

at
IV, and VI III,

V,
slower rates, and the metallic elements Groups
of

was shown that oxygen ions dif


Fig.
In
at

the lowest rates. 9.28,


it

UO, many magnitude greater than that


of
in

fuse rate orders


at
a

do
of

gases, whereas the latter diffuse

as
fission about the same rate
at

would, thus, anticipated that the anionic fission


be

uranium ions.
It

products, Br, and Te would diffuse


I,

at

rates much faster than that


do
of

That they not may


be

attributable
to
the noble fission gases.
be by

their diffusion rate being controlled other species dissolved


of

that
the UO. Such constituent may fission product Cs which would
be in

be a

anticipated tightly associated with the halides and whose dif


to

of

expected comparable other ca


be

be

fusion rate would that


to

to

the UO, lattice. Thus, may inferred that the Group and
be

tions
in

it

Groups VI and VII elements are closely associated when solution


in
by

UO, and that their release from solution governed the diffusion
in

is

less apparent why the other


of

rate the least mobile element.


It
is

Table 9.12 diffuse


at

metallic elements listed rates much slower than


in
or

anticipated
be
of

that fission gases uranium ions. would not from


It

diffusion theory that cationic diffusion rates would


be

observed six
or

magnitude less than that


of

of

seven orders uranium ion self-diffusion


UO. most likely that the release Sr, Ba,
of

as
in

It

elements such
is

by

Ce, Zr, and Mo from defected fuel elements not governed dif
is

by

fusion rates the solid oxide but rather controlled other factors
in

is

solubility steam surrounding the UO,


as

in
or

such the water defect


in
IRRADIATION EFFECTS IN URANIUM DIOXIDE 537

tests. The insensitivity of the magnitude of the escape rate coefficient


of elements such as Ba, Ce, and Zr to oxide density was cited pre
viously as further evidence for control of escape of these species by
mechanisms other than diffusion.

TABLE 9.12—VALUES OF DIFFUSION COEFFICIENT RELATIVE TO


THAT FOR Kr 88

Test WAPD-29-1 (103) WAPD-29–2 (103)


Density (% T.D.)

-
93.4 97.4
Center Temp., 1,260 1,530
•C

e e D;
Isotope vi vi Di wº **
Ai Dr.” Mi DKrºº

Krº- - - - - - - - - -----| 2.92x10"? 1.24x10"? 1 0.465×10−7 3.14x10T11 1

4.65×10-11 3.75X10^* !--------------|--------------|--------------


11:1 1.90×10−9 3.63x10T1? 2.93x10-3 0.293x10-9 8.63x10-13 2.75x10-2

2. 16x10-7 1.33x10T10 1.07x10-1 0.865×10-7 2.13×10-11 6.8 x10-i


2.67x10-s 5.9 ×10-13 4.76x10-4 2.89 x10-8 6.9 ×10T13 2, 2 ×10-2
6.62x10-11 2.8 ×10-14 2.3 ×10-5 2,32 x10T11 3.43×10-15 1.09×10-4
2.31x10-11 6.85x10-13 5.5 ×10" |--------------|--------------|--------------
2, 24x10-12 8 x10-is 6.45x10−9 9.4 x10-12 1.41x10−16 4.5 x10-8
3.04×10-12 3.3 ×10-16 2, 66x10"? 8.5 x10T1? 2.58×10-15 8.2 ×10-8
2.2 ×10-12 3.93X10T1? 3.17x10-s 2.65 ×10-11 5.7 ×10-15 1.82×10-4

pºliº T.D.) x-ºr x-ºf


(64)
Center Temp., 1,700 >2,750

Isotope ** w? D; ** v? I),
X, DR,” X, DK,”

Krºs. - - - - - - - -- - - - - 2.1×10-5 8.4 x 10-12 1 3.5x10-7 1.78×10−9 1


Xe” - - - - - - ------- 1.7×10−3 1.9 x 10-10 22.6 9.7×10-8 6.2 x 10-9 3.48
Xelº. - - - - - - - - - - - - 3.0x10-8 1.32x10-11 1.57x10-1
*--------------- 2.0x10-8 3.26×10−9 389
4.9x10-8 2.6 x 10-10 31 .
1.0x10-8 1.25x10-11 1.49x10-1
2.9X10-9 2.32x10-14 2.76x10-3
7.0x10-0 6.7 x 10-8 8,000 5.5x10- 4.12×10−6 2,320
5.5×10-s 8.65×10-11 1 3.7×10-8 3.91x 10-1: 2.19x10-3
5.4x10-8 2.4 x 10-12 2.86X10' l--------------|--------------|--------------
4.9x10-11 1.53x10-14 1.82×10−3 4.2x10-11 1.12x10-14 6.29%.10-5
-------------------------------------------- 2.7x10-10 9.36x10-11 5.26)×10-2
8.6×10-1: 3.8 x 10-15 4.53X1C" |--------------|--------------|--------------
5.1×10-12 4.16x10-17 4.95X10-8 1.7×10−10 4.62×10-14 2.59x10-5
6.4x10-12 1.46×10-1s 8.2 × 10-7
7.6×10-12 4.7 ×10-16 2.64x10-7
4.0x10-9 5.51×10-12 3.1 x 10-4
2.7×10-8 2.93×10−10 1.64x 10-1

"The highest reported value of , reported during the test is used in theseestimates.
538 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

(6) OTHER FACTORs AFFECTING Fission PRODUCT RELEASE FROM


DEFECTED FUEL ELEMENTs. The data presented above indicate a rea

the
sonably well-verified correlation with a diffusional mechanism of

or
soluble fission products from fuel element samples

of
escape volatile
operated defected in-pile hot water loops. Such experiments,

in

in
which small (0.005-inch diameter) holes are drilled through the metal

by
lic cladding exposing the fuel within

no
the coolant, are

to
means
ideal for establishing such correlations. factors may

of of

be
number

A
visualized which could becloud the true mechanism escape from the
oxide, and, fact, surprising that reasonable correlations have
in

is
been derived considering the number it

of
effects which can confound
The more important these effects which will

be
of
the results. further
discussed below are the following:

the cladding.
of

Escape fission products from defect

in
3. 2. 1.

a
Waterlogging during reactor startups

or
effects shutdowns.
Changes composition during hot water exposure.
of

the oxide
in
of

Escape
Fission Products through Defects. The emission rate
the cladding
of

fission products

of
from hole metal-sheathed
in
a

a
by

the fission products

of
diffusion rate
as

oxide fuel element affected


steam-filled spaces between the fuel and cladding has
or
in

the water-
by

been analyzed Helstrom [105]. The diffusion equation applicable


to this case
is

*}=s+Dv'N.—NN, Eq. (9.29A)


V,

particles the isotope per


of

of

of to of
which
cc
the number
in

the
is

of i

the diffusion space,


of

material
in

the rate escape atoms


D is

the
S

diffusion space (atoms/cc-sec), the diffusion coefficient the


is

the diffusion space (assumed steam), and


be
as of

to

atom the material


in

At

the other terms are previously defined. points far from the
hole, the steady-state solution equation
of

this
is

N=; Eq. (9.29B)


by

steam may approximated


be

of

The diffusion coefficient use


in

the
following equation:
kT(M--M.)
3

Eq.
Tsnai; VT MM, (9.20C)
27

J/1 hydrogen atom (1.67 10-4 g), M,


of

where the mass the


is

is
×
= a

fission particle J/u, J/,


of

of

mass 100 the mass water molecule


is

J/1, constant, the absolute temperature


18

Boltzmann's
is is
=

is
T
k

(49.3°K), water molecules per


of

cc

the number (3.92× 102” cm.").


n

colliding molecules
of

in of

the distance between center water and


is
2
a

fission product Substituting these values the equa


-

10-5 cm.
×
4

tion, found that cm2/sec.


-
D
is

10-2
it
IRRADIATION EFFECTS IN URANIUM DIOXIDE 539

In the steady state, the solution of Eq. 9.29A is

§-N–. v2 N, Eq. (9.29D)

Thus, M, is a function of
º and the distance in which significant

changes in Mi can occur are of the order of For particles of half


V%
lives of the order of one second, !-01 cm, whereas for half-lives of

hour, Thus, short-lived isotopes, such as the delayed


an
w!’-6 cm.

neutron emitters, which are important for location and detection of


failed fuel elements, are emitted primarily in the vicinity of the defect,
whereas those important for the system contamination are emitted
from the bulk of the fuel element [106].
The following solutions of Eq. 9.29D were derived for a num
ber of geometrical arrangements:
1. Long slit in a plane surface
Eq. (9.29E)
n-2s W.
where h is the thickness of the fuel-clad gap and n is the number of
atoms per unit length of slit per second.
2. Long slit in a cylindrical surface

Eq.
n-2s. Vº (9.29F)

for a cylinder of circumference


*Vº and

n=2|Sb/, Eq. (9.29G)

for a cylinder of circumference


*Vº
a plane surface (or cylindrical surface with small hole)
-
Hole in
-
3.

- N -1
S
!-0%) nºvº
iv

Eq. (9.29H)
(in

n >2m
ºh
;

where the radius of the hole.


is
a

To estimate the magnitude cylindrical rod contain


of

this effect
in
a

an

as

ing hole 0.005-inch diameter, the rate isotope such


of

of

release
a

V,

Cs” (A=3.5 10−" secº) from the oxide, (atom/sec) 0.357-inch


in
×

a
540 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

diameter, 9.1-inches long oxide element is compared to the rate


of escape, n, (atoms/sec) through the hole. Since

S
N
Trx0.357539.TX(2.54)3Xi'

N_TX0.357x9.1 × (2.54)2X.h>{3.5×10-4
ºn. 2TX107*Xh

wisvº2 10-2 -

(in
0.572)-li
Thus, the necessity for diffusing the isotope through the gap between

of
sheath and fuel the defect hole reduces the apparent rate
to

release
by

Cs”. Because the scatter and


an

as
isotope such
of

factor 1.7 for


a

nonreproducibility the data between various experiments far exceed


in

is,
factor,
no

this correction has been made for this effect. however,

It
high
of

significance note that

of to
considerable technical concentrations
be up

can build the diffusion space defected fuel elements; this


in

effect will hydriding

of
defected fuel ele
to
discussed relation
in

ment cladding
in

Sect. 9.8.1.
Waterlogging Effects. Robertson and Allison have shown that,

if
by
its
escape from governed

of
defected rod rate escape from the
is
a

isotope activity the loop test water upon


of

solid oxide, the decay


in

of
reactor shutdown should correspond the decay constant the iso
to

tope question [102]. This was, indeed, found the case for Krº
be
in

to

Xe”.However, and Ba”, Eichenberg, I*


of
in

and the case al.,

et
Fig. 9.41 for Ba”,

by
found, isotope activity
an
in
as

shown increase
or in

upon flux change (either increase decrease) by fac


10
of

factor
a

a
8.5 and, furthermore, that decay activity from this peak was
of

of

tor
by

much slower than that given the radioactive decay and loop puri
fication constants [64]. No satisfactory explanation
of

this effect has


been presented. Walton has proposed,
of
in

of

the case observations


*

fission product release during temperature changes, that the time


required establish the concentration gradient characteristic
of
to

the
new temperature delay attaining the
of

observation introduces
in
a

steady-state activity release rate.


In

any case, such phenomena can


significantly fission products from fuel
of
to

contribute the release


elements subjected cyclic operation.
to

The further observation that, particularly on reactor startups,


a
by

peaking activity release rate occurs has been presented


of

number
a

experimenters [64, 102, 107]. This effect has been attributed


of

re to

the expulsion water, trapped within the defected sheath during


of

of

actor shutdown, the defect hole by the formation


of

out steam
as

Walton, Energy Establishment, Harwell, personal


G.

N.

Atomic Research communi


*

cation.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 541

10-10-

.
5

9 --
o',
#
> 8
7
6
5
\ D
4 – \-— RADioactive DECA, -
\ PLUS PURIFICATION
3H \ -
\ D TIME FROM WITHDRAWAL
2 \\ O TIME FROM INSERTION

\
\
10-12 | —l- \l ! | l
O 2 4 6 8 IO 12 14
TIME IN HOURS

FIGURE 9.41. Escape Rate Coefficient, v, for Ba” vs Time from Flux Change,
WAPD 25–2 [64].

the fuel increases in temperature carrying with it dissolved activity.


This phenomenon has been termed waterlogging.” With sufficiently
large defects, a similar peaking may occur on shutdown, caused by
entry of water into the interior of the sheath during cooldown, the
water flashing to steam on encountering hot oxide surfaces and ex
pelling some of the contaminated water within the sheath during this
process. The phenomenon affords a most convincing demonstration
of the presence of defected oxide fuel elements, since similar effects are
not noted with any other source of fission product release, such as
surface contamination of cladding with fissionable elements. This ef
fect has been utilized to prove the existence of a defected fuel element
in the first core of the Shippingport PWR. In Fig. 9.42, the delayed
neutron activity changes during plant startups are compared in an
assembly of fuel elements suspected of containing a defect with a

* This term should not be confused with waterlogging failures of fuel elements which
will be discussed in Sect. 9.8.1b, although the genesis of the two phenomena is probably
the same.
542 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

*H----|--|--|--|--|--|
- TTTTTTT
6 MiN

lo” H

s E

- DEFECTED
o Assembu-Y
# .22
3 to E
o -
5E

- UNDEFECTED
-
Assembu-Y

Io
--~<!
st----1––––––––––––––––––––––––
T STARTUP
REACTOR
POWER ->
*
22 Mw
*
37 Mw
T
-
O606 O618 O630 O642 o654 ofog of is of 30
Time in Hours

FIGURE 9.42. Delayed Neutron Activity in Shippingport (PWR) following Re


actor Startup of September 22, 1958 [104].

similar undefected assembly [104]. The steady rise of activity from


the latter, contrasted with the violent peaking observed during power
changes in the former, is strikingly revealed in this plot.
Change in Composition of UO,. The presence of water in the inte
rior of the cladding of a defected fuel element introduces adjacent to
the fuel an oxidizing medium which can react chemically with UO.
and, thus, change the composition. In a defected cylindrical fuel
element of diameter D and gap d between fuel and clad, the mole
fraction of interstitial oxygen which will be introduced in the fuel by
complete and uniform reaction with the water is about ; if the ele

ment is filled with water, or 0.07


# if the element is filled with Satu

rated steam at a temperature of 220° C.


a value of D=0.357 inch, For
UO,
would change to UO.
ors

in the former case the compositionfor


gap, and UO, ow, the gap considered initially
of

to

each 0.001 inch


is
if

filled with steam. Such oxidation can change the release character
by

two ways: first, altering the diffusion rate fission prod


of
in

istics
by

ucts (see Fig. 9.32) different higher oxide


or

surface oxidation
to
a

of

which has been observed cause almost complete release


to

dissolved
by

volatile fission products from the oxidized layer; secondly, decreas.


ing the thermal conductivity
be

oxide, will
of

as

the discussed
in
a

later section, thus increasing the temperature operation and the


of

consequent fission product release rate. Continued diffusion


of

steam
the fuel element through the defect hole could then
of

into the interior


IRRADIATION EFFECTS IN URANIUM DIOXIDE 543

eventually oxidize the UO, to a composition of UO2.16, the composition


of UO, in equilibrium with water or steam (see Chap. 6).
Such excess oxygen contents should be observable by metallographic
examination of defected oxide samples after in-pile loop exposure, since
the excess oxygen precipitates on cooling as a U.O., phase (see Chap.
6). No such metallographic indications have been reported. This
observation is particularly important in the light of the observation
of Childs and McGurn of U.O., formation in irradiated UO, os during
heating to a temperature of 340°C [44]. Also, analyses of defected
oxide samples have, in general, failed to reveal an increase of oxygen
content, although one analysis indicating a composition of UO2.05 has
been reported [108]. Failure to observe the anticipated increase in
oxygen content may be attributable to the formation of hydrogen dur
ing the reaction of steam with UO, or the interior cladding surface,
and the difficulty of diffusing the hydrogen out of the interior of the
fuel element through the small interstices between fuel and cladding
or through the cracks in the fuel. If a defected cylindrical rod fuel
element is approximated by a plane circular slab of radius b with a
hole of radius a in the center of the slab and diffusion space of thick
ness h between fuel and cladding, the diffusion rate of hydrogen out
of the hole will be given by solution of the relation

Eq. (9.30)
Dvºw–3–0

where Q is the rate of formation of H., (or of reaction of H2O) in


moles/cmº-Sec, and W is the concentration of steam in moles/cc."
For the boundary conditions W-
W., at r=a, and d W/dr=0 at

r= b,
—w
W=W. —" ”,
"," " (a_a:
Fiji, (r”—a”) Eq. (9.30A)
2Dh

If it is assumed that W-0 (i.e., the gas atmosphere is entirely hydro


gen) at r=b, and substituting the values b = 4 inches, a = 0.0025 inch

W.Dh
º–842
From Eq. 9.29, it can be estimated that the diffusion coefficient of
H. in steam is about 0.02 cm3/sec and W. (at 750° F, 2,000 psi) is
0.0032 mole/cc. For h=0.001 inch, a reaction rate of 4.68 × 10−"
moles H2O per cm-sec is necessary to maintain a hydrogen content
of the indicated amount 4 inches from the defect hole. In a 0.357-inch
diameter specimen, about 8.2 × 10° moles UO, are present per square

* R. W. Dayton, Battelle Memorial Institute, personal communication.


544 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
H-H-I-T-I-T-I-T-I-T-I-T-I-T-I-T-I-T-I-H-I-T-I-T-I
EQUILIBRIUM PRESSURE over water AT 2000 psi PREssure
O H. -
14/1 RATIos of P(H,0) to P(H,)

3&i

|-l—l-l-l-l-l-l-l-l-l-l-l-l-l-l-L-L-L-L-L-L-N-Nº!
4.
3

7
5

* 6

9
10

/T
Equilibrium Pressure Oxygen with Various Uranium Oxides [64].
of

FIGURE 9.43.

Thus,
reaction rate yielding

an
cladding.
of

is of
centimeter increase
a

interstitial oxygen UO, 10−8 per second


of

of

mole fraction
in

×
5

atmosphere predominantly
an

hydrogen
of
maintain
to

sufficient
within points distant from the defect. This
at

defected fuel rod


a

by

the possibility
of of

of
amount still further diminished
reaction
is

of

reaction H.O with the interior the metal sheathing. As shown


Fig. 9.43, plot UO., func
as

oxygen content attained


in of
in

in

in a
a

temperature which the partial pressures attained UOs.


of

tion
are compared with those psi
of

various H2/H2O mixtures


at

2,000
H,
an

pressure, atmosphere percent percent H.O will limit


to 50

50
of

the oxygen content UO, about UO, [64]. Thus, contamina


of

or

the UO, within water-cooled reactor appli


of

tion defected sheath


in
a

unlikely affect seriously the rate fission product release.


of
to

cations
is

conductivity
on

the oxide will


be
of

The effect thermal discussed


in

high hydrogen contents within


of

Sect. 9.7.2(c). The attainment


metal-sheathed fuel elements most serious connection with hydro
in
is

gen attack cladding, will also


be
of

further discussed
as

Sect.
in

9.8.1 (b).
fission product release from
of

of

(7) SUMMARY. studies number


A

defected fuel elements have been reported both for in-pile hot water
loop tests operating plants
as

well for [64, 95, 101, 103, 104, 107.


as

109–111]. The most successful parameter for correlating these results


rate coefficient, values
of

of

has been the escape which for number


a
IRRADIATION EFFECTS IN URANIUM DIOXIDE 545

tests are listed in Table 9.12. A list of constants recommended for use
by Cohen and Iltis in table 9.13 and the contribution to
are listed
coolant activity resulting from these parameters in Table 9.14 [112].
The escape rates are highest for elements of Groups I, VI, and VII,
VIII, and lower for the Group II, III, and IV elements.
The absorp
tion characteristics and resulting system contamination fortunately
follow the opposite trend, being essentially zero for Groups V, VII,
and VIII and highest for Groups II and IV. Tellurium in Group VI
is strongly absorbed and also escapes relatively rapidly. Therefore,
this isotope specie is one of the most serious contaminants emitted
from UO, fuel elements. From consideration of the activity contri
butions from the various elements in Table 9.14, it was concluded that
control of activity introduced in a water coolant from a defective fuel
rod could be best achieved by providing for hold-up in the purifica
tion system to allow decay of the relatively short-lived fission product
gases and demineralization to remove the long-lived soluble fission
products.

TABLE 9.13—ESCAPE RATE COEFFICIENTS FOR A UO2 FUEL


ELEMENT [112]
Elements w, sect'

Cs, I, Xe, Kr, Rb, Br---------------- 1.3 × 10-8


Mo-------------------------------. 2 × 10–9
Te-------------------------------- 1 × 10–9
Sr, Ba---------------------------- 1 × 10-11
Zr, Ce, other rare earths.------------ 1.6 × 10-12

TABLE 9.14—PERCENT GROSS ACTIVITY CONTRIBUTED TO COOLANT


AFTER 3,000 HOURS OF OPERATION WITH DEFECTED FUEL ROD

Elements Percent gross activity

Rare gases-------------------------- 79.15


Halogen ----------------------------- 19.12
Alkali metals ----------------------- 1.38
Molybdenum------------------------- 0.26
Alkaline earths.---------------------- 0.05
Total other-------------------------- 0.02

9.5 CONTAINMENT OF FISSION PRODUCTS IN UO,

In Sect. 9.4, it has been shown that the release of fission products
from UO, can be explained more or less satisfactorily on the basis of
solid state diffusion of the fission products out of solution in the UO.
matrix. This finding infers that the fission products are accommo
dated within the UO, lattice, and it is the intent of the present section
to identify the manner of their accommodation. For this purpose, it
546 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

is necessary to secure resolution of two major problems. These prob


lems, which will be discussed in detail below, are:

1. Identification of fission products and physical and chemical effects


of their solution.
2. Limitations to containment of fission products by UO, and the
structural and property changes accompanying the reaching of
such limitations.

9.5.1 Identification of Fission Products

Table 9.6 presents the yield of stable or long-lived nuclides from the
thermal neutron fission of U*. The rare gases Xe and Kr whose be
havior was discussed in detail above constitute over 25 percent of the
number of fissioned atoms. This table assumes decay of the 9.2-hour
Xe” to Cs”. Lewis pointed out that, due to neutron capture by the
Xe” at high flux exposures forming stable Xe”, the fraction of Xe
formed could increase markedly, depending on the fraction of the 6.3
percent of Xe” formed which captures neutrons [63]. At 80 percent
conversion of Xe” to Xe”, the total amount of Xe formed increases
to 26.87 percent, and the total percentage of noble gases produced per
fission of U*can rise to about 31 percent. Thus, since the most severe
physical changes are anticipated from formation of the noble gases,
the flux or fissioning rate should affect the severity of damage result
ing in fissionable materials.

(a) Fission Product Valency

It is apparent from examination of Table 9.6 that the majority of


the stable nuclides from the thermal fission of U*
would be expected
to behave as cations in solution in UO. The distribution of nuclides
freshly formed on fission is such, however, that their solution in UO,
would lead, if anything, to an ºnion excess, and the change in nature of
the fission products occurs gradually with time as B-decay of the fis
sion-formed nuclides progresses. Robinson, on the assumption of the
maintenance of an average fission product valency of plus two (corre
sponding to the formation of two fission product atoms from U"), cal
culated the change in the chemical nature of the fission products as a
function of irradiation time as shown in Fig. 9.44 [113]. The change
in the chemical nature of the fission products with time is clearly
shown. Thus, in irradiation periods of less than an hour (point A of
Fig. 9.44), the fission products retain the valencies shown in the caption
of Fig. 9.44, whereas during a week's irradiation (point B in Fig. 9.44)
many of the fission products must be reduced to zero valency if charge
neutrality is to be maintained. For irradiations extending to a year
or more, still more of the fission products must be reduced or else
IRRADIATION EFFECTS IN URANIUM DIOXIDE 547
x
I I I T
#
u
a
-- -
_--T --
52.5H (a)

L----T
5 (b)

à ~~~ --tº
- * - *- * -
~~~~

(c)
#,6ET A

B
" ... *
-

*
...
- -
- -
woo 3

---- - -

-
u-

T.
(e)


1.5F
§ §

|
l
I
l

iOS IoW los 105 IoW


2.

loz
|RRADIATION TIME IN SEC

CURVE (q) VALENCIES AS IN TABLE BELOW EXCEPT Se”, Te", Brº.,

1*
CURVE (b) VALENCIES AS IN TABLE BELOW
Ge". As", Rhº, Pd*, Ag", ca", In", Sn", Sb", REMAINDER AS

IN
CURVE (c) (b)
CURVE (d) Ru", REMAINDER AS IN CURVE (e)
CURVE (e) Nb",Tc", Mo", REMAINDER AS IN CURVE (d)
ASSUMED VALENCIES OF FISSION PRODUCT ELEMENTS
|N U02
VALENCY ELEMENTS
–2 (Se), (Te)
(Br),
| O —

1)
(
I

Kr, Xe
Rb, Cs, (Ag)
+

+2 Sr.,Bo, (Cd), (Pd)


RARE EARTHS, (Rh), (As), (Sb), (In)
Y,

+3
+4 Zr, (Mo), (Nb), (Tc),(Ru), (Ge), (Sn)

FIGURE Average Fission Product Valency UO, [113]. (Courtesy,


in

9.44.
American Nuclear Society.)

partial the remaining


Thus, the chemi UO,
as of

reduction must occur.


products
as

be
of

cal nature well the distribution fission can assumed


be functions both of irradiation time and neutron flux."
to

The presence reduced fission product species UO, irradiated


of

in

over long time periods increases the likelihood


of
of

chemical reaction
water with UO, defected fuel elements.
In

Sect. 9.4.10(b) was


in

it

unlikely that appreciable reaction with UO, would


be

considered
to

by

the protection afforded hydrogen buildup. How


of

occur because
in by

be

ever, the substoichiometric oxide produced fissioning would


anticipated much less stable than UO,
be

contact with H2/H2O


to

is to at

mixtures and, hence, chemical reaction defected fuel elements


in

high depletions, least the extent necessary


to
at

to

restore the oxide


stoichiometric state, quite probable. Tentative evidence for this
is
a

presented Sect. 9.5.


in

Barney and Seymour,” considering the free energies


of

formation
the individual stable fission product oxides, arrive
of

rather
at
a

fission product atoms between


of

of

similar estimate the distribution


charged and neutral atoms. Their results are tabulated Table 9.15.
in

secondary effect on fission product distribution exerted by the specific nature


of
A

is
*

the fissionable atom (see Ref. 62).


K.

Barney and .W. Seymour, Knolls Atomic Power Laboratory, personal


E.

W.
*

communication.

57.4789 O–61—36
TABLE 9.15—CALCULATION OF THE FISSION PRODUCT VOLUME IN UO2+

Crystal Univalent Molar or Fission

|
|
Fission product AF""/oxygen Atom percent Mole percent Oxygenatoms cations cations Specific Molecular atomic product

A
A
atoin yield yield required radius, radius, gravity, g/cc weight volume volume,
cc/mole” cc/mole

1.
La2O3- 122 00 00 00 15 39 51 325. 50. 150.

6. 4.
9. 7.
Y203 119 80 40 20 0.93 20 84 225. 46. 112.
Ce2O3- 119 12. 20 10 18. 30 11 27 90 328. 47. 290.

3. 2. 6. 3.

6.
9.
Pr:03-- 118 00 00 00 09 25 88 329. 47. 144.
Nd2O3. 117 22. 34 11. 17 33. 51 08 24 24 336. 46. 521.

2.
1.
3.
Prm 203--- 115 60 30 90 06 21 30 342. 46. 60.
1.

0769587
0000081

Eu2O3- 115 0.03 0.02 0.04 03 18 55 352. 53.


Sm2O3- 113 34 17 51 04 19 43 348, 46.9 54.8

-
Sr0 106.9 60 60 60 13 32 70 103. 22. 211.

2. 9. 5.
1. 1. 1. 1. 1. 1. 1. 1.

8
1. 9. 5.
3. 9. 5.
Ba0--- 101. 80 80 80 35 53 72 153. 26, 155.
1. 1. 1. 1. 1. 1. 1. 1. 1. 1. 1.

ZrO2- 97.7 31. 40 31. 40 62. 80 0.80 09 60 123. 22. 690.

0
SnO2- 69. 0.08 0.08 16 71 0.96 95 150. 21.

0. 0.
0.
1. 0.

Sb2O3----- 45. 0.02 0.01 03 0.63 89 20 291. 56.

030
000760

6. 4. 6. 6. 7. 7. 6. 7. 4. 5. 5. 6. 5. 6.

MoQ2- 41. 18. 57 18. 57 37. 15 63 0.93 47 128. 19. 367.


|-
0807086
--
--
---

0. 0.
1.

- --- - - - - --- - -- -- - ---


--- --- --- --- - -- -- --- --- --- --- --- - --- ---
--- --- --- --- --- --- --- --- --- --- --- --- --- --- ---
UO2- 102.5 l----------|----------|---------- 97 30 10. 96 270. 24.

6,
Mo-----------|---------- 43 43 l----------|----------|---------- 10. 20 96. 9.4 60.
6,
4

Te--------------------- 60 2.60 ----------|----------|---------- 25 127. 20. 53.


9.
()

|-
*

6,
Te-----------|---------- 10 10 ----------|----------|-------- 11.00 99. 54.

6. 2, 6. 3.
Rb-----------|---------- 90 3.90 ----------|----------|---------- 48 85. 57.7 225.
0
1,
-- 5 0 9 0 0

Cs----------- - -- -- ----- 18, 60 17. 22 |----------|----------|---------- 87 132, 71. 223.

0.
()

0
Cô ---------- -- - -- 02 02 ||----------|----------|---------- 64 112. 13. 0.3
-|-
---
---
--
--
--
--
---
--
--
---
---
882860096427500 0605940

1. 1. 8, 4,
6,

--
--
---
---
4
6

Se - 0.40 0.40 82 79. 16,


4.

0.
8
2

---
---
---
6.

|-

13
13
44 212.

4.

25
1.25 51 259. 72.

1. 3.
3.
98 98 ~3. 00 111.
18, 10
11. 38
90
18. 10
11. 38
2.90
~3.
12. 60
12. 40
60 131.
101.
102.
: 650.
92.
24.

2. 1.
1.
20 20 11. 40 106. 11.
ii
-- -|- -|- -|- -- -- -|-
-- -- -- -- -- -- --
-- -- -- -- -- -- --
-- -- -- -- -- -- --
-- -- -- -- -- -- --
0.03 0.03 10. 50 107.
5,
7

198. 80 169. 26 200 00 348.

*W. K. Barney and W. Seymour, Knolls Atomic Power Laboratory, personalcommunication.

E. at
**AF=the free energies 1,366°K.

of
of

is
***The averageatomic volume the fission fragments 27.0cc/mole comparedwith the UO2 molar volume 24.6cc/mole.
#
550 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Again, it is apparent that the rare earths, alkaline earths, zirconium,


etc., must be considered to exist in an oxidized condition within the
|UO, lattice.

(b) Fission Product Volume

Further insight into the nature of the incorporation of fission prod


uct atoms into the UO, lattice may be gained by consideration of the
volume changes which should accompany such incorporation. Bar
ney and Seymour,” considering additivity of molar volumes, calcu
lated, as shown in Table 9.15, the volume of fission products in UO.
to be more than double that of the original UO. Following the treat
ment of Weber, an expansion rate parameter Ry is defined as the frac
tional increase in volume per unit of fuel burnup [114]. For the data
in Table 9.15, therefore,

Rx-º'-12,
53.487–24.6

or an increase in volume of 1.2 percent for each percent of uranium


atoms fissioned. Howe and Weber, by a similar technique, predicted
a value of Ry-1.0 to 1.6 [115]. On the other hand, Eichenberg, et al.,
observed no measurable volume changes in bulk UO, fuel elements ir
3.7

percent burnup UO, [64]. Barney and Wemple


of

radiated to the
by

and Barney,
of

careful lineal analysis fuel element sample


a

containing percent dense UO, which had been irradiated per


63

53
to
cent uranium fissioned, taking into account the amount porosity of

in
the oxide, concluded that the expansion rate, RN, was less than 0.33
[116, 117].
No completely suitable explanation for the divergence between the
predictions and the observations
as

yet been suggested. Weber


has
all

has proposed that the metallic fission products atoms are ionized.
the atoms assuming the valency corresponding
of

with some
of to

of

that
of

suboxides [114]. The smaller volumes ionized than neutral


atoms was, thus, postulated smaller expansion rates. As
to

to

lead
support for this hypothesis, Weber cited the expansion rates calculated
for fissionable metallic
at

(and observed) matrices which are least


twice the values quoted above for UO, and which the fission prod
in

be

ucts are anticipated will


to

as

It

reside atoms. observed Table


in

those fission products


of

of

9.15 that the summation the molar volumes


an

oxidized condition predict expansion rate, Ry,


If be

an
to

assumed
in

the remaining unoxidized fission products are then as


of

0.125.
interstitially located atoms within the lattice,
as

may
to

sumed reside
it

accommodate large amounts such fission products


be

of
to

feasible
without undue lattice distortion. Note this connection that one
in

Barney Seymour, Knolls Atomic Laboratory, personal


K.

E.

W. and W. Power
*

communication.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 551

quarter of in the UO, lattice for


an oxygen ion can be incorporated
each uranium ion without introducing large amounts of lattice strain.
It is also of interest to observe from Table 9.15 that those elements
which escape with greatest facility from UO, during irradiation are in
general those which are listed as being unoxidized in their solution in
UO2. decide the mode of incorporation of fission products in the
To
UO, lattice will require dimensional and property measurements after
irradiation to a precision hitherto unattained in such experiments.
The discussions above of fission product accommodation in UO,
infer that the fission product atoms attain chemical equilibrium with
the UO2. It has already been shown in Sect. 9.4.10 and it will further
be demonstrated in Sect. 9.5.2 that fission products tend to distribute
inhomogeneously. If the oxygen ions freed by fissioning of each
uranium nucleus are redistributed independently of the fission-born
nuclei, the resulting volume changes may balance between expansion
due to fission atoms and contraction from oxygen solution in the UO2.

9.5.2 Burnup Limitations of Bulk UO,

may be anticipated that bulk UO, would reach its burnup limits
It
either by inclusion of so much fission gas in its structure that the gas
can no longer be contained and physical fragmentation of the oxide
occurs or by alloying with fission fragments to an extent that struc its
no

longer stable fission fragment damage. An example of


to

ture the
is

latter effect was cited previously the structural change


of

the case
in
by

fission fragments when suitable alloying additions


in

induced ZrO2
were present the ZrO2.
in

by

further example provided the fast neutron damage SiO,


to
is
A

(Sect. 9.2.2(a)). Here was postulated that damage occurred


at
it

increasing rate displaced atoms were formed


to an

in
as

amounts unable
incorporated the lattice leading
be

of

ever increas
in

to

conversion
an

ing volumes amorphous state. will


its be
of

to

It

the lattice shown


below that both the above effects probably operate UO2 burnup
as
in

limit exceeded.
is

Noble Gas Precipitation


of

(a) Effects
by

An example Barnes and Mazey


of

the former effect was noted


(Tho.)
of

who heated samples and noted disinte


mineral thorianite
gration beginning and occurring violently
at

700° (118].
at

950°
C

This sample, which contained about 11.5 cc/g gas


of

which half was


helium (corresponding about 0.07 mole fraction He), was observed
to

fragment into pieces corresponding flaws present


of
to

the sizes
in to

the original mineral body. Thus, contained fission gas, whether


initially present
or

solution the oxide precipitated flaws, may,


in

in

in
552 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

by precipitating and exerting pressure within flaws, cause gross physi- |-


cal disintegration of the body.
Another type of behavior was noted by Frisby, Bisson, and Caillat
in the case of BeO samples which had been exposed to a fast (>1 MeV)
integrated neutron flux of 5.5 × 10”, thus, forming about 0.00002 mole
fraction of helium [119]. On heating such samples to a temperature
above 1,000°C and examining fractured samples (the samples frac
tured intergranularly) by the electron microscope, precipitation of
gas in crystallographically oriented voids at the grain boundaries was
noted. The size of the bubbles increased with annealing temperature
and attained a diameter of several tenths of a micron at 1,500° C.
Thus, gas precipitation in voids and volume swelling due to such
precipitation can occur in oxides as well as in metals. If such precipi
tation occurs when the body is brittle, fragmentation may occur.
whereas at elevated temperatures where plasticity of the oxide is en
countered, precipitation accompanied by swelling can occur. Elston.
et al., point out that the oxygen released by the Be ions consumed by
nuclear reactions may also be effective in causing property and struc
tural changes [120].

(b) Metallographic Observation

Most important in forecasting the burnup limitations of UO, are


the metallographic findings reported by Barney, et [116, 117, 121]. al.
These observations were obtained from two types specimens:
of
The
one, annularly loaded UO, pins, contained highly enriched UO2 pow
an

der packed percent density annulus adjacent


63

the stain
to

to
in

less-steel tubular cladding; the other contained dispersions


of
enriched
UO, stainless-steel matrices which the UO, (although frag
in

in

mented during fabrication operations) percent dense.


94

96
to

was
In

each case, burnups exceeding


of

percent
50

the uranium atoms were


Considering the first type specimen, Fig. 9.45
of

achieved.
of in

is

sample irradiated percent


of

to

shown the structure the uranium


4
a

atoms fissioned. The fuel relatively


to

is to

observed have sintered


is

a
no

compact mass, but other obvious structural change


In

noted.
Fig. 9.46, which the UO, has been subjected burnup per
of
to to
in

9
a

the uranium atoms, fine porosity begins appear


of

the UO.
in

cent
structure, and Fig. 9.47 (25 percent atoms fissioned) the poros
U
in

ity has increased amount. similar appearance was evident


in

ef in
U A

samples exposed percent atoms fissioned. Corresponding


52
in to

dispersion stainless-steel samples. Figure


in

fects were noted the


the UO, unirradiated sample: the
an
of

9.48 shows the structure


in

quite extensively cracked, although


its

oxide may
be

be
to

observed
original density was Fig. 9.49,
In

percent
95

of

in

theoretical. which
UO, percent atoms, partial heal
of
to
12

the has been fissioned the


U
IRRADIATION EFFECTS IN URANIUM DIOXIDE 553

FIGURE 9.45. Uranium Dioxide Irradiated to 4 Percent Burnup, As-etched,


X500, Reduction Factor, 14. [116].

FIGURE 9.46. Uranium Dioxide Irradiated to 9 Percent Burnup, As-polished,


X100, Reduction Factor, 4. [116].

ing of the cracks, as well as appearance of fine porosity in the UO2,


is noted. In Fig. 9.50, no further cracks are apparent and the poros
ity has increased to a uniform extent through the fuel particles.
A number of other significant observations were made in the course
of this work. In Fig. 9.51A, it was observed that UO, particles
which become exposed at the surface during an irradiation at 1,000 to
1,500° C plastically flowed to the free surface, the flow being evi
denced by the deformation of the pores in the UO. The effects of con
fining pressure are evidenced in Fig. 9.51B, in which UO, particles
exposed at the free surface expanded freely, whereas those embedded
554 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FIGURE 9.47. Uranium Dioxide Irradiated to 25 Percent Burnup, As-polished,


X500, Reduction Factor, 4. [116].

FIGURE 9.48. Unirradiated Uranium Dioxide Particle of 95 Percent Density in


Stainless-Steel Matrix, As-etched, X500, Reduction Factor, 4. [116].

in the matrix showed no apparent tendency to bloat the cladding. That


the deformation is caused by expansion of gas within the pores is evi
denced by the much larger pore sizes in the particles exposed at the
free surface compared with the embedded particles. Apparently, quite
moderate pressures are adequate to restrain the bubble growth.
Thus, these results may be summarized as follows: At uranium
burnups corresponding to 10 percent of the uranium atoms in UO.
fissioned, a structural change consisting of the appearance of pores in
the microstructure appears; accompanying this structural change is
an apparent plasticity of the oxide and a tendency toward swelling
and healing of cracks. The swelling is readily restrained by the im
IRRADIATION EFFECTS IN URANIUM DIOXIDE 555

FIGURE 9.49. Uranium Dioxide Irradiated to 12 Percent Burnup, As-etched,


X500, Reduction Factor, 14. [116].

FIGURE 9.50. Uranium Dioxide Irradiated to 30 Percent Burnup, As-polished,


X500, Reduction Factor, 14. [116].

7s

A. Toward Fractured Sur- B. Irradiated to 21 Percent


face at 1,000 to 1,500° C, Burnup at 200° C, As
As-etched, X250, Reduc- polished, X100, Reduc
tion Factor, 4. [116]. tion Factor, 14. [116].

FIGURE 9.51. Plastic Flow of Uranium Dioxide.


556 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONs

position of external pressure. Significant in the interpretation of this


behavior are the results reported by Cashin on the irradiation of a
uranium-bearing glass containing 45 weight percent SiO2, 8.5 weight
percent Al2O3, 13.5 weight percent TiO2, 4.5 weight percent ZrO2, and
18 weight percent Na2O in addition to 10 percent enriched UAOs [122].
On irradiation of this body at a maximum central temperature of
1,750°C to a uranium depletion of 15 percent, formation of voids was
found to occur at the center of the specimens where temperature was
highest, presumably as a result of precipitation of the fission gases at
temperatures above the softening point of the glass, 1,450° C. The ap
pearance of these voids was quite similar to that noted in Fig. 9.45
to 9.51B, leading to the postulate that the structural change occurring
at 10 percent fissioning of the UO, accompanying appearance of poros
ity is a vitrification of the hitherto crystalline structure. Further in
dications of structural change are provided by the results of Wisnyi,
quoted by Weber, who found a progressive decrease in the temperature
at which rapid fission gas release occurred from samples of UO, ir
radiated at burnup levels of 10 to 50 percent of the uranium
atoms [114]. Such a variation with burnup would not be anticipated
from the curves presented earlier (such as Fig.9.21) unless a structural
change favoring the release of dissolved gases, such as vitrification.
occurred upon extended burnup.
It might be anticipated that continued burnup would render the
fluorite structure of UO, unstable. Pauling has shown that stability
of the fluorite structure obtains when the univalent radius ratio of
the cations to the anions exceeds the value of 0.73 [39]. In pure UOs,
the value of this ratio is about 0.74. In Table 9.15, a number of values
of the univalent radius of fission product ions is listed. It may be
noted that these are, in general, less than that for UO, and, in fact, the
average univalent radius for the fission product atoms likely to be
oxidized is 1.170 as compared with 1.30 for uranium. Thus, as con
tamination of the UO, occurs with progressive burnup, the structure
approaches a dilution at which instability of the fluorite structure may
be anticipated.
The porosity noted in the microstructures reproduced in Figs. 9.45
to 9.51B apparently reflects the insolubility of the fission gases at
the concentration levels attained in UO, at about 10 percent fissioning
of the uranium atoms. It may be anticipated that, as fissioning of the
UO, progresses, other fission products generated will exceed the solu
bility limit and will precipitate solid phases. This expectation is
as
confirmed by results quoted Ref. 123. Dispersed particles of
in
UO, in a stainless-steel matrix, after about 70 percent depletion of the
uranium, were observed to exhibit a second phase at the surface of the
UO. particles, and still a third phase precipitating in the particle in
IRRADIATION EFFECTS IN URANIUM DIOXIDE 557

terior. The outer surface of the UO, may be expected to be enriched in


oxygen as a result of recoil of fission products out of the surface into
surrounding matrix. Thus, the compositional change
the

the rim

in
by
burnup proceeds affected not only fission contamination but
as

is
the particle interior are
by

cation depletion. The precipitates

in
also

result from buildup the solid fission products

of
com
in to

to
presumed
their solubility the UO.
of

in
positions excess

High Burnup
(c)

of

Oa'ide Plates

at
Behavior

metallographic structure

of
The observations described above
by

Barney, al., are affected, Figs. 9.51A and


et

as
shown

in
change
by

very rigid structure.


of

9.51B, the containment the oxide within

a
applicable describing the conditions existing high
to

at
Results more

by
bulk oxide fuel elements are afforded the recent obser
of in

burnups

97
Bleiberg, al., who exposed 0.040-inch thick plates

of
et

vations
percent dense UO, clad Zircaloy sheathing the thickness and con
of
in

Fig. 9.52 high temperature water in-pile loops


in

in

figurations listed
several experiments The changes the plates

of
in

[124]. thickness
in

measured intermittently during the course


of

were the irradiation.


The results are shown Fig. 9.52. Only small changes fuel
in

in
burnup percent the ura
10
of

of
thickness were noted until
to

levels
8
T

T
T
T

I
i

i
I

I
I

I
i

CLAD COMPARTMENT COREHEATFLUx ESTIMATEDCENTER


50H. Thickness DIMENsion BTu/HR-FT TEMPERATURE*F 25
-

320,000 900 -
13

Nxooson.
IN. N.

oozº
in

I

- oois 1230
- 6.N.ANDžinal.5IN... o.4on 630.ooo
× =
N
;

º
I

º O.O2.In IN. 31N.× O.O401N. 305,000 900


“. ºn s º
O

20

:*o DEFECTED
-
HT H H T H

Ti

Nasin AND#IN=31Nzooson.
#

630,000 1230
T z a. 3 : § :

Exposed
N.

to

001° water
A

DURINGIRRADIATION
-H

o
9 B
...

- -
15

530
- - -1
#IN compartMENTs
-

=
-H

u
|-
-uu
. º:z §

2
|-

20
to
* º: wz °

;2
–– –

IN. compartMENTs r
I

44

"".
- defect Ed
C

º *
- -
-
A
L

§
H.

10
+

ºf
5

- - ºl.
o

- -
A

C9
- °a -
1—º-1--1–1–1–?—
A

O
Co -
8

©e O
A
O

O
o
i
o

io 20 30, iO20
o

Fissions/cc

5 1

l
I

15
o

PER CENT FISSIONED


u

Change Fuel Thickness on Irradiation UO, Plates [124].


in

of

FIGURE 9.52.
558 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

|
FIGURE 9.53. Cross Section of 0.040-in. Thick Fuel Plate, after 21.5X10^
Fissions/cc, Increased in Thickness by Fuel Swelling [124].

nium atoms fissioned were reached. At this point, depending sensi


tively on the compartment dimension and clad thickness (i.e., on the
degree of external restraint additional to the 2,000 psi system pres
sure applied to the fuel platelet) more or less abrupt increases in plate
thickness were measured. The burnup level at which thickness
changes apparent increased in the order of the restraint
became
afforded by the cladding, the 14-inch span clad with 0.024-inch-thick
cladding affording appreciably more restraint than, for example, the
4%-inch span clad with 0.021-inch cladding and still more restraint
than is afforded by the defected cladding in which the system pressure
cannot serve to apply pressure on the fuel platelet through the clad
ding. It is further noteworthy that those samples in Fig. 9.52 which
were clad with 0.021-inch- and 0.024-inch-thick Zircaloy sheet were
irradiated at less than half the fissioning rate of the samples clad with
0.015-inch sheet. Thus, in addition to restraint, the rate of fissioning
appears to be important, higher rates leading to greater dimensional
instability. Cross-sectional examination of the fuel after thickening
occurred revealed that the fuel had swollen and that the result
observed was not attributable to the release of fission gases forcing the
cladding away from the fuel. This is shown clearly in Fig. 9.53 in
which the fuel in a cross section of a sample which had swollen a maxi
mum of 46 percent in fuel thickness is observed to be in direct contact
with the cladding and, furthermore, to exhibit the typical crowned
its

appearance of fuel which deforms plastically during swelling. As


of

the plastic fuel, the volume changes


of

of

result flow the fuel


a
IRRADIATION EFFECTS IN URANIUM DIOXIDE 559

plotted in Fig. 9.52 are estimated to be about one-half the thickness


change which is measured at the top of the crown.
Evidence of the onset of a structural change accompanying the fuel
swelling was afforded by metallographic examination [124]. In Fig.
9.54 is shown the microstructure of a sample of UO, after 10 percent of
the Uatoms had fissioned and which had increased in thickness a
maximum of 10 percent. The appearance of porosity in the structure
may be compared with the similar photomicrographs of Figs. 9.45 to
9.51B. Still more pronounced swelling was noted in Fig. 9.55 in the
case of samples which had been defected through their cladding and

had been exposed to the water coolant in the course of their irradiation.
Figure 9.55 shows a montage section of such a
across a transverse
sample, indicating the severe porosity, particularly in the center of the
sample. This photograph may indicate that the swelling in UO, ex
posed in condition during irradiation is temperature sensi
a defected
tive, since the swelling became more severe with higher fuel tempera
ture. In addition to the appearance of porosity, a further significant
observation was the inability to develop by etching a grain structure in
either of these samples using the same etching technique which reveals
grain structure readily at lower exposures.
(1) Fission GAs RELEASE. A further important observation was
the enhanced escape of fission product gases noted on puncturing the
swollen samples after their irradiation. In Table 9.16 are listed some
values of fission gas (Kr") release from typical fuel element samples
of similar geometry to those described in the previous paragraph as a
function of fuel burnup. At fuel burnups up to 8.5 × 10° fissions/cc,
lessthan 0.5 percent release of gas was noted from high density (97
percent theoretical) fuel of the type utilized in the experiments plotted
in Fig. 9.52. However, samples which exhibited the porosity in
Fig. 9.54 after some 10 atomic percent burnup of the fuel together
with a thickness increase of 10 percent showed on puncturing the com
partments after irradiation a discontinuous increase in fission gas re
lease to almost 10 percent of that generated. As final evidence, samples
which had been irradiated to 10 atomic percent burnup were heated to
1,000°C after irradiation for collection of the fission gases. A diffu
sion rate at least six orders of magnitude higher than those plotted in
Fig. 9.21 at this temperature was noted.
The correlation of the diffusion coefficient at 1,000°C and the frac
tion of Kr" released into the fuel-containing compartments, as meas
ured by puncturing the compartments after irradiation, with the
fissioning depletion of UO, is shown in Fig. 9.56. Both the diffusion
coefficient and the fraction evolved increase exponentially with fuel
burnup. The former increases by about six orders of magnitude over
the burnup range investigated, whereas the latter increases about two
of
of
;

FIGURE 9.54. Montage Cross Section Fuel Compartment from Plate Oxide after 21.5x10” Fissions/cc x250,

º/,
to
in

Reduction Factor, [124]. (Fuel was not exposed loop water. The large void areas the center of the

to
compartment are due loss of fuel during sectioning of the plates for metallographic examination.)
of
of
Figure 9.55. Montage Cross Section Fuel Compartment from Plate Oxide Element after 21.5×10”
to
in

;
Fissions/cc X250, Reduction Factor, [124]. (Fuel was exposed loop water interval between

#4,
in

to
17.5 and 21.5× 10° fissions/cc. This compartment was adjacent that Fig. 9.54.)
#.
VJL 378 I’6 NOISSIJI—9 SVO GHSVGHTIGH?HJNOHJH CHLW'IA CHCIIXO [] { "ICH SJLNGHINGHTIGH

H
quaIºneſoo/Suoſss! !lòngxnų ssºux 0ļų4

XI

|
“KnţsuºCIlºngquatuſuºdx{ quaouad
enJ-utų/n1{Ileoņauoøųn quºouo 02-01X‘ssøuxoſum seueqopostalºu ožu
søųou!t_0IX

|-------------------1&-+I |----------------------------top I-------------------------------96 'I0#0’0ļot º[qţăț|}}øN0

~
~

ſa
so 0Z#Z~~~~ #0ºg0 N9 oſqȚĂ![×o
n-rozſy ?.0#0’000€.~~~~to øſq!?!!?øN!
n-ºOºty I----------------------------------|----------------------to ‘i0#0’0ļot øſqţăț|3øNZ
0IZI 10 0 · N9 øſqqāſīža
562 URANIUM DIOXIDE:

0861 ºg0 0 o[qįäſſäøNZ


09 80 0 º[q!?![30N9
010#0’006I N.9 º[qĮ3ȚIŻło

0
00&1 ^90; º[q}}][3øN8
ºg0#0’00Z9 øſq!?!!?øN9
‘º0#0’0099 ſøN0 øſqȚĂȚI
0989 **0 0 øſq13||8øNg
ºg0#0’00€g øſq|3||3øNz
19. º.º.080’0* (3øNz oſqqāſ
'0Z0! º080 o[q}}][3øNZ
£080’0099 N0 ø[q131180
'0989 "I001 øſq!?![3øN6
'0019 ºg001 øſq13113øN8
‘0009 ºg001 øſq13||3øN8
'0009 'Z001 N9 ølglºſſa
’0099 ºg001 º[q!?![3øN8
‘oZ91 o[qț3118øN0'8001
'0681 £001 ølq13||8øN0
L-+--On ---------------OºO+toſ, I-------------------------------lo '0zot º[qĮŽIĻºøN0'8001
ſon ---------------------------~ I-------------------------------go '0zot øĮq181|}}øN0'8001
ſ-c-x ~~~~~~ ||--~~~~
ºd

ton |---------------------------- |-------------------------------16 00 9 '90,0 0 01! quºou


~~IZ-6Z~~~~
~~~~
~~~~~
|~~~~
I----------------------------------|-----------------ton-owo-rouz 0• '^i0 0 gºz1 quoouòd
'010 0
ºd

ſ----------------------*On-Oºſh auſsi 08 ----uoſsuadsip 8 6 quoou

|
~~~~tom-0081 ~~~~ ~~~~--~~~~ ~~~~~ osuvoo |---------------uoisuodsſp '^i0 0:0oly
od

91 quoou
PROPERTIES AND NUCLEAR APPLICATIONS
IRRADIATION EFFECTS IN URANIUM DIOXIDE 563

I I I I I I I Ol

C DIFFUSION RATE OF Kr 85AT 1000°C


–2 H *F o FRACTION Kr"RELEASED FROM uo, FUEL PLATE
-
f O
:c –13 H > 1.0 H
O 8 .
§ 8
o
– 14 H
tºn
= os H -
3. x
r
c
o – 15 H.
; O - -
e
* O
g
: O
-
-* | g +9° E 3
-1 Soooo
O O
- 17 H - I.O. H. -
O

- ie l - 1.5 L l
3
l
6
Ol
9
I
12
I
15
!
18
l
21 24

BURNUP of Uoz IN Fissions/cc x to −20

FIGURE 9.56. Variation of Diffusion Rate and Release of Kr" from Dense UO2.

orders of magnitude. The maximum fuel temperature of these


samples during their irradiation was less than 1,000° C. This lower
fuel temperature plus the fact that fractional release is proportional
to VD probably
account for the lower sensitivity of the latter measure
of fuel damage to burnup than of the former.
(2) X-RAY STRUCTURAL CHANGE. Further evidence of structural
change was obtained by X-ray examination of samples from these tests
after 10 atomic percent burnup of the UO2. In the upper plot of Fig.
9.57 is shown the X-ray diffraction pattern of unirradiated UO2. The
center plot indicates a similar scan in the case of a sample which was
its

exposed to high temperature water during irradiation. The diffrac


tion peaks corresponding the UO, structure are still visible, although
to

highly diffuse
on

greatly diminished intensity, and superimposed


in

background. The bottom plot was made using similar sample which
a

had not been exposed water during irradiation. Although the same
to

background intensity noted for the sample


in

diffuse the central


is

residual UO, diffraction peaks


is,
no

thus,
of

plot,
It

trace visible.
is

concluded that the UO, these burnup levels has vitrified and
of no
at

is

longer capable coherent X-ray diffraction. The appearance


of

the sample exposed water may the oxida


be
of in

to

to

peaks attributed
by

fission products
in
to

tion access water (as discussed Sect.


hence, affording the UO, the ability crystal
its

9.5.1(a)),
to

retain
structure. This observation, properly interpreted, thus, tends
to
if

confirm that the structural change caused by fission product con


is

tamination, and that oxidation the fission products permits retention


of

57.4789 O–61–37
564 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I I I I i I ſtill). TI
UNIRRADIATED UO2
(VERTICAL SCALE NOT COMPARABLE)
(22O)
(2OO)

x-3-L, POSITION 3
(311)
DEFECTED
ExPOSED TO WATER

(200) (111)

x-3-L, POSITION 3
UNDEFECTED
NOT ExPOSED TO WATER

l | I l I l l l
6O 50 40 3O
DEGREES 26, Cu Ka RADNATION

FIGURE 9.57. X-ray Diffraction Profiles of UO, Irradiated to 21.5×10”


Fissions/cc [124].

of the oxide structure to higher burnup levels. The vitrification ob


served in Fig. 9.57 appears to be quite stable thermally. Annealing
for 24 hours at 1,000°C failed to cause appearance of a crystalline
structure.
Additional X-ray examination of the oxide in the samples plotted
in Fig. 9.52 shows a close qualitative correlation between the amount
of swelling, as measured by thickness change of the plates, the rate
of burnup, and the degree of destruction of the coherent diffraction
peaks. Oxide from those samples which showed considerable swell
ing revealed strong background scattering and weak or nonexistent
peaks, whereas those samples showing little volume change revealed
strong diffraction peaks and little background scattering even though
their fissioning depletion may have been the same. Thus, the diffrac
tion plots for samples similar to those plotted in Fig. 9.57 which
showed about 10 percent thickness change (in the undefected condi
tion) but exposed to only about 16 × 10° fissions/cc showed strong
UO, diffraction peaks, and their thickness change correspondingly was
only about 5 percent. Similarly, samples exposed to a fission deple
IRRADIATION EFFECTS IN URANIUM DIOXIDE 565

tion of 21.5×10° fissions/cc at a rate of 12.5×10” fissions/cc-sec


showed the complete absence of X-ray diffraction structure as shown
in the bottom plot of Fig. 9.57. On the other hand, samples exposed
to 30 × 10° fissions/cc at a rate of about 6.3 × 10** fissions/cc-sec
showed an appreciable remanent diffraction pattern similar to that of
the central plot of Fig. 9.57.
It is of interest
to note in Fig. 9.57 that the diffraction peaks which
remain even after very high exposures, although greatly diminished
in relative intensity, do not appear to be appreciably broadened com
pared with the diffraction profile of lightly fissioned UO, shown in
Fig. 9.7, nor is an appreciable line shift indicative of change in ao
noted, even though the gross composition of the UO, is such that about
17 percent of the metal atoms in UO, consist of fission products. It
can be determined from the data in Table 9.15 that the average crystal
radii of the metallic fission products, whose free energy of formation
of oxide favor their combination with the oxygen anions freed by the
destruction of uranium ions (amounting to 61 percent of all the fission
product atoms formed), is 1.244 Å or 28 percent larger than that of
U* ions. If the lattice expansion due to the incorporation of only
these fission product ions is estimated at the fissioning depletion of
21.5 × 10° fissions/cc, a unit cell expansion of 3.4 percent would be
expected in the samples plotted in Fig. 9.57. No shift from the unir
radiated samples greater in extent than that noted in Table 9.5 for
slightly fissioned UO, could be detected.
The following inferences may be drawn from these observations:
1. Fission products are either incorporated in the lattice of highly fis
sioned UO, in such a way as to cause no lattice parameter change
or are segregated at discrete locations rather than being uniformly
distributed in the UO2 lattice.
2. The crystal size of the UO, does not change appreciably with ir
radiation and remains in the size range 700 to 1,000 Å.
3. The destruction of crystallinity occurs by conversion of the rela
tively macroscopic crystalline UO, phase to an amorphous (or at
least nondiffracting) phase rather than by a gradual and continual
uniform deterioration of the crystal structure.

9.5.3 Fission Effects in BeO-UO, and ZrO2–CaO-UO2

Before consideringthe implications of these data with respect to


fuel element performance, the behavior at high fissioning depletions
of oxide bodies with matrices other than UO, is reviewed.
Listed in Table 9.16 are some preliminary results of an experiment
(WAPD–29–21) aimed at determining burnup limitations of fueled
566 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

|
bodies in which highly enriched UO, is diluted with low thermal neu-
tron cross-section oxide additives, such as BeO, and ZrO. stabilized
with additions of CaO [48]. The BeO—UO, bodies contained about
28 weight percent enriched UO, which
was dispersed in the BeO ma
trix as coarse (> 150p) particles and as fine (<10p) particles. It will
be noted from Table 9.16 that, at a burnup of about 12×10" fissions/cc.
a swelling of more than 25 percent in thickness of the fine dispersion
occurred. This swelling was accompanied by a large release of fission
gases, about 19 percent of those generated. The coarse dispersion ex
hibited the anticipated effect of increasing the particle size of the dis
persed phase in minimizing damage to the matrix. However, both the
thickness changes and the gas release are still excessive compared with
the properties of UO, at similar burnup levels. These results may be
compared with those of Gilbreath and Simpson in Fig. 9.12 and indi
cate that the damage which was detected at low burnup exposures pro
ceeded to intolerable levels as burnup was extended a factor of 100 over
that plotted in Fig. 9.12 [14]. The reasons for the damage are quite
well revealed in the photomicrographs of Figs. 9.58 A, B, and C. Fig
ure 9.58A shows the high density UO, particles dispersed in the dense
BeO matrix prior to irradiation. After an irradiation corresponding
to about 60 percent fissioning of the UO, in the dispersed phase, the
UO, particles exhibit the characteristic porosity previously noted in
Figs. 9.48 to 9.51B for UO, dispersed in metal matrices surrounded
by a damage zone in the BeO of extent corresponding to the recoil
range of fission fragments in BeO (Figs. 9.58 B and C). The struc
ture of the BeO matrix itself appears to be disturbed even beyond
the recoil range; X-ray diffraction photographs of the body contain
ing the fine fuel dispersion revealed no coherent diffraction peaks
either from the BeO or from the UO2, whereas the body containing a
coarse UO, dispersant showed a residual BeO pattern (although again
no UO, pattern) corresponding to the larger undamaged matrix vol
ume present in the latter. Figure 9.58C shows a UO, particle at the
cladding fuel interface; as in Fig. 9.51B, the cladding exerted suf
ficient restraint to prevent swelling of the particle out of the matrix.
A characteristic between the structures in Fig. 9.5S
difference
and those in Figs. 9.49 to 9.51B is the appearance of the recoil zones.
In the case of the metal matrices, the recoil zone appears as a sharply
its

defined rim around the particles distinguished by etching behavior


the BeO matrix, however,
of

from the remainder


In

the matrix.
gross, rounded pores are visible the recoil range. These pores
in

arise from the agglomeration


of

are inferred fission gases recoiled


to

into the highly loosened recoil zone


of

the BeO.
the background Fig. 9.58B are apparent
of

Observable cracks
in

the BeO matrix. These samples, exposed fast flux


in
to

excess
in

a
IRRADIATION EFFECTS IN URANIUM DIOXIDE 567

A. Structure of Unirradiated Fuel, UO, Particle Size 150p. ; X400, Reduction


Factor, 14.

C. As in Sample B, at the Cladding-Fuel Interface; X400, Reduction Factor, 1%.

FIGURE 9.58. Microstructures of BeO-30.1 Weight Percent UO, [48].

of 2 × 10” nvt, undoubtedly are exhibiting the physical disintegra


tion the matrix due to fast neutron flux bombardment previously
of
reported by Elston in Sect. 9.2.2(c) [16]. Thus, much of the volume
change exhibited by the coarse dispersion may be attributed to for
mation of matrix cracks beyond the recoil damaged zone, and the
high quantities of gas released in this material may, in part, be at
tributed to the matrix deterioration by fast neutron flux bombard
ulent concurrently with the fissioning damage. In the fine dispersions,
it is likely that the damaging effect of fission fragments far exceeded
that of fast neutron exposure and that the volume changes as well
568 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

as very high gas release rates noted are attributable primarily to


crystal lattice destruction by fission fragment bombardment. Thus,
BeO is shown to be far less stable than UO, with respect to ability
to sustain fission fragment damage and fission product contamina
tion effects.
- -
- -
- - -
*- - - -
*

- -
-
-1

B. After 11.8x10” Fissions/cc : X 400, Reduction Factor, 14.

FIGURE 9.59. Structure of Irradiated ZrO2–13 Weight Percent CaO Plus 17


Weight Percent UO, [48].

The fuel consisting of ZrO2+13 weight percent CaO +17 weight


percent enriched UO, is seen, from Table 9.16, to exhibit good dimen
sional stability at burnup levels of 12× 10° fissions/cc. In Fig.
9.59A is shown the microstructure of the unirradiated fuel. Note
that the structure consists of a tetragonal ZrO2-rich phase (pre
dominant), a twinned phase which is presumed to be CaZrOs, and a
phase which was probably molten at sintering temperatures and
which is suspected to be an Al2O,-rich impurity phase introduced
during fabrication. After irradiation (Fig. 9.57B), the complex
phases present at the ZrO,-UO. grain boundaries are observed to be
greatly damaged and give the appearance of having increased in
volume and to have flowed around the ZrO. grains. X-ray diffraction
examination of these samples still showed good structure retention.
is,

It
as

thus, probable that the ZrO.-base oxide exhibits least good


at

fission fragment damage the UO.-base material. This


to

as

resistance
IRRADIATION EFFECTS IN URANIUM DIOXIDE 569

finding, again, illustrates the importance of initial crystal structure in


determining resistance to irradiation damage and confirms the stability
cf the fluorite-type lattice.

9.5.4 Summary and Implications Concerning Burnup Limitations


of UO, Fuel Elements

The behavior of metal oxides when subjected to ever increasing


amounts of fissioning exposure is construed to follow the progression
summarized below:
1. As discussed in Sect. 9.3, UO, at fissioning exposures up to
about 5×10" fissions/cc forms an equilibrium number of point de
fects (or such associated in more complicated arrays) amounting to
about 1 percent displacements. The dimensional, X-ray, and prop
erty changes are consistent in revealing such limited damage.
Oxides such as Al2O, and UAO, when subjected to similar exposures
at low temperatures exhibit severe damage. However, at more
elevated temperatures (above 350° C for UAOs and above 600°C
for Al2O3) the damage remains restricted in extent.
2. As fissioning proceeds beyond 5×10" to about 10×10" fissions/cc
in the case of UO, this stable condition is retained. The fission
fragments become incorporated in the oxide in some manner, as yet
unknown, which affects neither the microscopic appearance nor the
crystal structure. This range of stability appears to be similar for
ZrO2 but much more limited for BeO in which detectable pro
gressive damage is noted at 1.5× 10” fissions/cc.
3. As fissioning proceeds beyond the level of about 10× 10° fissions/cc,
a number of effects begin to appear. Fission gas diffusion rates and
fission gas release begin to increase in an approximately exponential
manner. Volume increases begin in the fuel and also occur at an
increasing rate the higher the exposure. Plastic flow of the oxide
is noted. The appearance in X-ray diffraction patterns of nonco
herently diffracting material is noted at the expense of the crystal
line UO. Porosity is noted in metallographic sections and becomes
more gross the greater the exposure. These effects may be compared
with those described in Sect. 9.2.2(a) for SiO, under fast neutron
bombardment. The progression of damage in UO, beyond 10×10*"
fissions/cc appears similar to the initial portions of the sigmoidal
damage curve noted there for SiO2. In the latter case, the nature
of the damage curve was postulated to arise from the collection of
displaced atoms in the SiO, lattice until a concentration was reached
at which the lattice was disrupted. I)amage then progressed at an
increasing rate as the volume of amorphous phase increased and
the proportion of crystalline phase which serves to repress the vol
ume changes decreased. Similarly, for UO, it is surmised that
570 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

fission fragments segregate and build up at discrete locations to


concentrations that are no longer containable in the UO, structure.
The structure breaks down in these regions of microscopic extent
which grow in volume as fissioning proceeds. Because of the higher
diffusion rate in the amorphous lattice, fission gas diffusion rates
increase. The plasticity of the amorphous phase permits agglomer
ation of fission gas bubbles and, hence, the appearance of porosity
as well as the healing of cracks noted on irradiation of dispersion
Samples.
The question may be posed whether the volume changes notedin Fig.
9.52 are the result of formation of an amorphous or vitreous UO.
of lower density than the crystalline oxide or result from precipitation
and agglomeration of gases in pores as noted in Figs. 9.54 and 9.55.
As will be discussed in Sect. 9.8, the volume change on melting UO.
is estimated to be less than 3 percent, and, hence, the greater part of
the volume changes in Fig. 9.52 is to be assigned to the formation of
gas-filled pores.
The effect of burnup rate, previously commented upon in connection
with Fig. 9.52, in increasing volume changes the higher the rate, im
plies that, given time to do so, the fission product contaminants may
diffuse to locations less damaging to the UO, crystal lattice. If this
interpretation is correct, higher operating temperatures should favor
retention of crystallinity at higher fissioning depletions. On the other
hand, an optimum operating temperature range may well exist with
respect to overall dimensional stability, dictated by the counteract
ing effect of higher temperatures in favoring diffusion and agglomera
tion of fission product gases.
If the quitereasonable assumption be made that the amorphous UO.
formed upon fissioning shows liquid-like properties with respect to
exhibiting a pressure-dependent solubility for the noble gases, a num
ber of speculations may be made concerning the design of fuel elements
for high fissioning depletions. As the structure becomes amorphous,
permitting precipitation and agglomeration of fission gas bubbles, a
degree of porosity will be attained such that the partial pressure of
fission gases both within the pores and in void spaces between the fuel
and its restraining clad equilibrate at a constant level in equilibrium
with a certain concentration of gases in the oxide. As fissioning pro
ceeds and the concentration of gas within the gap increases, higher
pressures are built up within voids, and a new equilibrium pressure
is established depending on the strength properties of the cladding.
The net effect is to limit the pressure of fission gases over the fuel to
a value between zero and that corresponding to complete release of the
gases, the actual value depending on temperature-solubility-pressure
characteristics of the system.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 571

Although for making quantitative estimates,


data are not available
this concept does explain qualitatively many of the observations re
corded above. The influence of mechanical restraint afforded by clad
ding geometry is well evidenced by the comparison of fuel restrained
in a cladding of 14-inch span and 0.024-inch-thick cladding which
swelled, from Fig. 9.52, much less at corresponding fission deple
tions than fuel restrained in a 12-inch span by 0.021-inch-thick clad
ding. At equal deflections of the cladding, at least in elastic range,
the former design applies 24 times as much pressure to the fuel as
does the latter. Hence, at the same deflection, a correspondingly
larger amount of fission gas can be retained in solution in the amor
phous oxide rather than precipitated in voids and pores. The greater
volume changes noted with defected cladding are explicable on the
basis that fission gases are free to escape from the fuel, an equilibrium
pressure limiting the precipitation of gas no longer can be built up,
and the oxide swells at a rate controlled by the diffusivity of gas in
the amorphous solid, the distance to the closest void, and the ability
of the gas to diffuse from the solid into the coolant. In partial con
firmation, more gross voids are visible at the center of Fig. 9.55 where
temperatures are highest and gas diffusivity correspondingly greater.
Similarly, at the free surfaces in Figs. 9.51A and 9.51B, voids are
coarser where the oxide is exposed, and, again, an equilibrium partial
pressure of fission gases cannot be maintained. The higher strength
of the restraining cladding about the dispersion particles of Fig. 9.50
permits buildup to the equilibrium pressure with less resultant defor
mation of the cladding than in the case of the bulk plate oxides in
Fig. 9.52. The effect of temperature of operation depends on the
sign of the heat of solution. If this is positive, as would be anticipated
for the solution of noble gases in glasses, the solubility in the solid
would increase with temperature at a fixed pressure and less swelling
would be expected.
Thus, operation beyond the fissioning depletion at which severe
structural damage has occurred in UO, is favored by fuel elements in
which the fuel can be tightly restrained. Defecting of fuel element
cladding has been observed to be injurious to the dimensional stability
of the fuel. The greater the burnup rate of the fuel, the less is
its

apparent ability sustain structural damage. Considerable experi


to

highly fissioned UO,


or on

necessary
to

mentation establish the above


is

postulates indicate other explanations. fis


of

The distribution
to

sion products and the crystalline and amorphous phases must


be
of

determined, and the existence pressure-dependent solubility


of

of
a

amorphous UO, confirmed


In

addition
or

the noble gases denied.


in

by

the complications segregation fission-product


of

introduced
to

lattice, precipitation fission gases voids,


of

of

atoms, vitrification the


in
IRRADIATION EFFECTS IN URANIUM DIOXIDE 571

Although for making quantitative estimates,


data are not available
this concept does explain qualitatively many of the observations re
corded above. The influence of mechanical restraint afforded by clad
ding geometry is well evidenced by the comparison of fuel restrained
in a cladding of 14-inch span and 0.024-inch-thick cladding which
swelled, from Fig. 9.52, much less at corresponding fission deple
tions than fuel restrained in a 12-inch span by 0.021-inch-thick clad
ding. At equal deflections of the cladding, at least in elastic range,
the former design applies 24 times as much pressure to the fuel as
does the latter. Hence, at the same deflection, a correspondingly
larger amount of fission gas can be retained in solution in the amor
phous oxide rather than precipitated in voids and pores. The greater
volume changes noted with defected cladding are explicable on the
basis that fission gases are free to escape from the fuel, an equilibrium
pressure limiting the precipitation of gas no longer can be built up,
and the oxide swells at a rate controlled by the diffusivity of gas in
the amorphous solid, the distance to the closest void, and the ability
of the gas to diffuse from the solid into the coolant. In partial con
firmation, more gross voids are visible at the center of Fig. 9.55 where
temperatures are highest and gas diffusivity correspondingly greater.
Similarly, at the free surfaces in Figs. 9.51A and 9.51B, voids are
coarser where the oxide is exposed, and, again, an equilibrium partial
pressure of fission gases cannot be maintained. The higher strength
of the restraining cladding about the dispersion particles of Fig. 9.50
permits buildup to the equilibrium pressure with less resultant defor
mation of the cladding than in the case of the bulk plate oxides in
Fig. 9.52. The effect of temperature of operation depends on the
sign of the heat of solution. If this is positive, as would be anticipated
for the solution of noble gases in glasses, the solubility in the solid
would increase with temperature at a fixed pressure and less swelling
would be expected.
Thus, operation beyond the fissioning depletion at which severe
structural damage has occurred in UO, is favored by fuel elements in
which the fuel can be tightly restrained. Defecting of fuel element
cladding has been observed to be injurious to the dimensional stability
of the fuel. The greater the burnup rate of the fuel, the less is
its

apparent ability sustain structural damage. Considerable experi


to

highly fissioned UO,


or on

necessary
to

mentation establish the above


is

postulates indicate other explanations. fis


of

The distribution
to

sion products and the crystalline and amorphous phases must


be
of

determined, and the existence pressure-dependent solubility


of

of
a

amorphous UO, confirmed


In

addition
or

the noble gases denied.


in

by

the complications segregation fission-product


of

introduced
to

lattice, precipitation fission gases voids,


of

of

atoms, vitrification the


in
572 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

etc., the existence and influence of other phenomena in highly fissioned


UO, must be established. For example, unequivocal detection and
estimation of the presence of oxygen liberated by fissioning in UO is
important in determining the mechanism of damage. Largely
neglected in evaluation of fission damage has been the influence of
nuclides, such as protons and alpha particles, formed by high-energy
neutron spallation reactions. These and similar effects may influence
burnup limitations of UO, fuel.

9.6 CORROSION ATTACK OF UO, UNDER IRRADIATION

A strong incentive for the use of UO, as a fuel material in reactors


appar

its
cooled with water or other aggressive environments has been
Chap.

In
in
to

ent inertness corrosion attack such environments.

8
out-of-pile tests that UO, inert hydrogenated water
in

shown

in
it
is

is
and that thermodynamically, even high temperature exposure, U.O.
should suffer only on
compositional change, corresponding solution

to
a

interstitial oxygen UO2, without accompany


of

of
the cubic lattice
in or in

ing deleterious phase volume changes leading

of
material

to
loss
Only exposure oxygenated water has
of
to

to
the coolant. the case
deleterious corrosion attack been demonstrated, resulting formation

in
hydrated UO, anticipated
of

the nonadherent oxide 0.8H2O. The


UO, hydrogenated water environments has been
by of

inert behavior
in

the exposure
of

demonstrated metal-sheathed fuel rods defected


with 0.005- in-pile water-cooled loops
in
to

0.006-inch-diameter holes
long (up
no

for exposure periods reported uranium


14

months) with
to

loss from the fuel elements or evidence of uranium contamination of


loop water [64, 95, 107]. fact, Ehrenreich was unable
to
In

detect
any increase product activity released from fuel element
in

fission
a

0.005-inch hole when the oxygen content the loop


of

defected with
a

water was increased, for 32-hour period, from 0.1 ppm


to

17

about
a

The absorption excess oxygen UO. does


of

solid solution
in

[103].
in

material loss from UO, exposed aqueous


an

itself lead
in
to

not
in

environment even though Murray, Pugh, and Williams have shown


that temperatures above 1,100° considerable loss weight occurs
in
at

from nonstoichiometric UO., due higher


of

the volatilization
to

uranium, possibly UO, [125]. The volatilization affects re


of

oxide
distribution of fuel within the sheath rather than loss of uranium to
the external system.

Experimental
in

9.6.1 Evidence for Irradiation-Induced Attack


Water

those experiences indicating highly satisfactory cor


In

to

contrast
rosion behavior, three independent experiments have been performed
IRRADIATION EFFECTS IN URANIUM DIOXIDE 573

which indicate that, under proper irradiation conditions, UO, may


exhibit a deleterious corrosion behavior. The first of these experi
ments, the X-2–0 test, involved the exposure in-pile in a high-tem
perature water loop of a right cylindrical pellet of 95.8 percent dense
UO2, 0.905 cm in diameter by 1 cm in height, contained in a stain
less-steel wire basket. The specimen attained a burnup of 630
MWD/Ton (1.77 × 10° fissions/cc) at a surface heat flux of 120
w/cm" [107, 126]. After completion of this test, the pellet was
severely broken up; only 1.5 grams of the original 6.75 grams re
mained within the basket, and considerable activity was transported
to the loop system. The second experiment involved the exposure
in-pile at low temperature of a sealed ampoule containing UO, powder
and water [127]. After exposure, the ampoule contained a quantity
of hydrogen gas, but no corresponding amount of oxygen, indicating
reaction of UO, with oxygen formed during fission fragment decom
position of steam. In the third experiment, three plates, two 0.100
inch thick and one 0.04 inch thick, of Zircaloy-sheathed high-density
UO, (97 percent theoretical), were exposed in a high-temperature
water loop for times ranging from 120 to 160 effective full power days
and burnups ranging from 6.25 to 17.6 × 10° fissions/cc with 0.04 inch
defects drilled into compartments which were 14 inch wide by 6 inches
long (X–3–1 test) [124]. Weight losses of 0.192 to 0.726 grams were
noted after the exposure, accompanied by considerable loop activity
and the presence of uranium in the water and deposited on the system.
An autoradiograph of one of the samples is shown in Fig. 9.60, indi

FIGURE 9.60. Autoradiograph of Specimen Operated in High Temperature Water


Loop for 121 Days with 0.040-inch Defect Exposing Fuel to Water in One
Compartment, Five Other Compartments Undefected; Fuel Thickness, 0.040
14,

Inch; Compartment Dimension,


1.4

0.040 Inch, Reduction Factor, [124].


6
x
x

cating that the loss the vicinity


of

of

uranium localized the defect


in
is

hole. noteworthy UO, uniform around the defect


It

that the loss


is

is

hole, indicating that erosion effects arising from flow water parallel
of

length specimen probably important


of

the axis caus


to

the are not


in

ing the material loss. Pertinent information


on

the first and third


Table 9.17. further significant
to

tests contained observe that


in
is

It
is

similarsample the same dimensions and fabrication technique


of
a

sample X-7N–11 the X-3–l test was exposed


as

the same fission


at
in

ing rate almost the same burnup similar facility (the X-1–0
to

in
a

no

test) except defected with 0.005-inch hole and showed uranium


a
or

loss from the sample into the loop system.


574 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

It appears, as suggested by Harder and Sowden for the second of


the experiments discussed above, that these results may best be ex
plained on the basis of radiolytic decomposition of water by fission
fragment recoils into the water environment. As discussed by Allen,
ionizing radiation decomposes water by formation of the molecules
He and H2O, and the free radicals H and OH [128]. Highly ionizing
radiations such as fission fragments produce primarily the molecular
products of the decomposition. Boyle, et al., have found that 1.83
molecules of H. (and a corresponding number of H2O2 molecules) are
produced for each 100 ev of energy dissipated in water by fission frag
ments [129]. Ordinarily, the amount of such decomposition is limited
by the back reaction of the free radicals with the molecular decomposi
tion products, the back reaction being facilitated by increase in dis
solved hydrogen content in the water and by increase in temperature.
Thus, in the case of fuel rods defected with small holes, high hydrogen
contents build up within the fuel element cladding as the result of
small amounts of reaction, and further reaction is inhibited by the
presence of hydrogen which can only slowly diffuse away through the
small defect holes. However, in case of the X-2–0 test, in which the
hydrogen decomposition product is swept away by the loop water flow.
the hydrogen peroxide or its oxygen decomposition product may, if
rate with UO, sufficiently rapid, cause attack
its

of
reaction the
is

latter before swept away from the UO, surface. To determine


is
it

the possible magnitude this effect, the number H2O, molecules


of

of
react quantitatively with

0.
to

formed was calculated and assumed

U
form UO, hydrate slough from the surface.
to

which would and The


weight loss UO, then given by
of

is

pellet (cm”) recoil range (cm)


of


×

no. fiss/ccsurface area


X

1/4
electron volts energy H2O, molecules UO, mole
Sy

J g

X mole
X

fission recoil water electron volts


in

molecules UO.
X (

175 270
2 '×

106
= X =
gg.

NZ
7.5× 10.5
–4

-
/4×4.4.1
...

*
×

10.5
1

10−4
'
''
'
'
''

605. Ina
"

0.0183

This calculated value, Table 9.17, comparable with


as

shown
in

the
no is

weight loss actually observed, even though correction was made


by

for depression H2O, the presence


of

of

the formation rate dis


in in
a

H2O,
by

solved hydrogen the loop water, for the sweeping away


of

loop water flow before attack UO, occurs,


of

for decrease
or

the
in

the
material exposed decreased. The
of of

of
as

amount attack the volume


similar calculations for the X-3–l specimens shown
as in

results Table
as

9.17 show predicted amounts large


20

40
of

to

attack times those


observed, which may suppression
be

of

attributable
to

the the reaction


by

buildup hydrogen within the fuel-element sample sheath. Char


of
IRRADIATION EFFECTS IN URANIUM DIOXIDE 575

acteristically, of the cladding was noted on the down


a discoloration
stream side of the sheathing of both the X-2–0 and X-3–l samples,
an effect which may be attributable to the release of uranium from the

on
its
fuel and deposition the sheath surface.

TABLE 9.17—UO, LOSS UPON IRRADIATION IN WATER IN X-2–0


AND X-3–1 TESTS [124, 126)

Calcu
Defect hole Effective Total UO: lated

|
Test desig- Specification Fuel dimensions diameter full burnup loss weight
||

nation designation power (grams)

of
(cm) (cIn) fissions/cc UO2
days oxidized
(grams)

X-2–0. ----- Unsheathed. 0.005diameter----- Unsheathed- 15 1.77x1019 5.25 10.5

|
Pellet-- 1.0 height---------
--
-
-
-

X-3–1------ X-2N-10---. 0.635X0.254X15----| 0.10--------- 191 6.8 x1020 0.395 14.74

||
0.635X0.1X15------| 0.10--------- 121 13.6 1020 0.192 11.6

×
X-4B–42-
-]
-
-

X-7N–11 0.635X0.254X15----| 0.10 --- 175 6.25x1020 0.726 13.76


X-1-0------| X-1-B ------ 152 5.8 x 1020

0.
0.
0.635X0.254X14----| 0.013-...-------
by

proving the occur


no

While these results are


of in
means conclusive

by
UO, water, they
an
of

rence irradiation-induced corrosion attack


sufficiently UO,

of
indicative warrant caution the utilization
in
to

are
under conditions such that fuel exposed through large cladding
is

speculate that oxide diluents for fissionable


of

to
It

defects. interest
is

Th(),
ZrO, which are stable the quadrivalent state
or
as

fuels such
in

oxidizing may
be

of

under conditions free such irradiation-induced


corrosion effects.

9.6.2 Irradiation-Induced Attack in Other Media

Such corrosion effects in-pile are apparently not confined applica


to

utilizing Morley Young ex


an

tions water coolant. and performed


a

periment similar the second reported above which UO, powder


in
to

was exposed in-pile 80°C sealed ampoule with CO2 atmos


in
at

to a

phere.” They found quantitative conversion CO, with


of

CO2
a
a

corresponding oxidation As shown Fig. 9.61, the


of

in

the UO2.
conversion rate follows first order kinetics with the rate constant pro
portional the fission energy expended the atmosphere. At tem
to

in

peratures above 400°C, while the initial slope the conversion plot
of

rigorously independent equilibrium concentra


an

temperature,
of
is

percent CO percent
28

32
of

tion and 400° attained.


at

in at

600°
is
C

Further corrosion effects must be considered the reactor utiliza


oxide components which, while not characteristically irradia
of

tion

M. Morley and Young, Atomic Energy Research Establishment, Harwell, personal


D.
A.
*

communication.
576 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

iOO I i i I I I i O
- - 1.8
90 H. - - 1.6
- - 1.4 r
g so H — - 1.2 s
× — -
1.0 9
2 7o H —H – 2.8 -
º + -2.6 :C
3 so H - -23
- -2.2 5
: 50 H
?
- -2.0 *
i. A 4.0 cm PRESSURE
oil- -3° g
§ 40 H O 40.0 cm PRESSURE –3 s 2.
2 D OTHER PRESSURES BETWEEN I.O AND 45 cm - -3.4 is
§ 30 H Ó FIRST ORDER PLOT
-+ -32 g
É -3.0
ºl. 2O - -1

IO

O l I l I l l l I I L I l l I
O 1 2 3 4 5 6 7 8 9 IO II 12 13 14 15

4? x TOTAL KINETIC ENERGY OF FISSION FRAGMENTS, UNITS OF 102*ev

FIGURE 9.61. Conversion of CO2 to CO by Fission Fragments from UTOs.

tion-induced, are encountered under the temperature and flow condi


tions of certain reactor concepts. McKisson discussed the attack of
BeO in water vapor contaminated air resulting in the formation of
volatile Be(OH), [130]. The rate of attack at temperatures below
1,560° F (850° C) was limited to 0.0025 in./year by the rate of dif.
fusion of Be(OH)2 gaseous product through the laminar flow layer
adjacent to the BeO surface. Elston found the rate of attack of BeO
by H.O-saturated air to increase linearly with temperature at tempera
tures above 1,150° C [16]. Thus, utilization of fueled bodies with a
BeO matrix in H2O-containing atmospheres requires consideration of
such effects. The formation of volatile oxides, hydrides, or other hy
drated molecules at elevated temperatures occurs not only in case of
BeO, but also for oxides such as K2O, Na2O, Li,O, La,Os, TiO2, MoC),
WO, Fe0, NiO, and SiO, [131].

9.7 THERMAL PERFORMANCE OF BULK UO, FUEL


ELEMENTS

At the of the development aimed at application of bulk


inception
UO, as a power reactor fuel, an unanticipated difficulty arose in the
prediction of thermal performance capabilities of fuel elements from
the out-of-pile properties of UO, and cladding materials. Only
re.

cently has rationale been developed which promises permit suitable


to
a

explanation and prediction. the following section, there


In

pre
is
IRRADIATION EFFECTS IN URANIUM DIOXIDE 577

sented a description of heat transfer properties of bulk UO, fuel ele


ments, followed by a discussion of experimental observations of the
thermal performance of fuel elements containing both sintered and
powder or swaged UO2.
The heat transfer situation in oxide fuel elements is complicated
by several unique characteristics each of which, as will be shown below,
affects in an important way the thermal capabilities of the fuel ele
ment. The first of these is the poor thermal conductivity of uranium
dioxide and the rapid change of this property with temperature. The
second is the fact that the fuel is unbonded to the cladding and that,
in general, a gap exists upon assembly between fuel and cladding.

its
Estimation of the size of the gap and fuel performance

on
effect
by

complicated cracking
of

of
the brittle fuel result

as
the thermal
is

a
stresses developed during operation.

Mathematical Description
of

Heat Transfer Oxide Fuel

in
9.7.1.
Elements
of
(a)

Use
Parameterſkar
For solid cylindrical rod
of

of
in

the case which conduction heat


a

only the radial direction, the equation


of
in

heat conduction

is
occurs

q().2rrdr––K(T)
2.1%
ſ

ſº ſword-ſ: KTat
Eq. (9.31)

ſº q(ºrdr-
ſ.

K(T)dT
ſ

which, consistent units, q(r)


the volume heat generation rate
in
as in

is

radial position
the rod, the radius,
T, of

function the
in

is

is
a
a

pellet radius,
T.

the temperature radius the temperature


at

r,
is

is

(T) thermal
T.,

the rod center, the surface temperature, and


K
is

is
at

conductivity, temperature, This equation


of

function
T.

illustrates
a

ſ.

the importance the parameter K(T) dT defining the ther


of

in
by

properties fuel temperatures.


of

mal fuel elements limited The


by

its

importance this parameter was first pointed out Lewis, and


of

by

use was further elaborated Robertson for various fuel element


[132, Eq.
be

geometries
It

133]. can seen from 9.31 that the


the equality proportional the total heat gen
in

to

left-hand term
is

known, for example from in-pile


If

the rod. this value


in

eration
is
578 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

calorimetric measurements or postirradiation burnup analysis, the

parameter on the right,


ſ"K (T) dT, can be used as a measure of

temperature attained in the center of the rod. K (T) is known, the If


actual center temperatures attained can be calculated (assuming Ta can
be calculated from power output and ambient coolant temperature con
ditions). Furthermore, if the variation of q(r) with radial position is
known or can be measured as well as K(T), the temperature contour
through the rod can be calculated. Alternatively, knowledge of q(r)

permits

tion, r.
assignment

If
of

some recognizable
a value of

and
J. K(T)
reproducible
dT at any radial posi

structural change
- --- T,
occurs at this position, a value of the parameter
J. K(T)dT can be
assigned to this structural change; e.g., if the at which grain position
growth in UO, occurs is located by postirradiation examination.

T,
a value of
ſ.” K(T) dT, where is the temperature of grain
growth, can be assigned which is a measure of temperature at that
location. Correspondingly, if Too is known from other measure
T... -
ments, a comparison of
T K(T) dT with other measurements of
K (T) integrated over the same temperature range permits selection
It
is,

of an appropriate value for conductivity. thus, apparent that


Eq.
by be

of

valuable information can derived from use 9.31 and the

parameter this equation.


ſkdT defined
Fig.
In

there are plotted


of

function
as

of
the values
T

9.62
a
T

(T).d7" for =0° and 400°C for three out-of-pile meas


T

C
A

T1
thermal conductivity UO, [56, 57, 134]. These
of

of

urements
by

plots are similar presented Runnals [135] except that


to

those
by

Kingery, al., extrapolated the empirical


of

of

the data were


et

use

formula
fl-R)=# (T degrees Rankine)
in

and
(Bruhr
A

-- rºy -
39.9
(T
by

C)
a

=
of

Scott (w/cm
in

those the relation


A

TH-400
C,

C) the temperature w/cm-º beyond


to

at

which =0.017
H
°

data were presumed


of

which all three sets


to

show constant value


a
of A.

on

apparent from these curves that, depending


of

which
It
is

set data was applicable, the heat output from fuel rod with sur
a

face temperature =0° and center temperature =2,750°C


T

T
by

could vary percent. This comparison


25

over somewhat mis


is

an

curves, for
as

leading, since,
of

shown the lower set actual


in
IRRADIATION EFFECTS IN URANIUM DIOXIDE 579

IOO i I i I I I

* 2^ • ,”
º
,”
8O H.
,”
e 2^ A
2^ .*
* e
/ /

al.
ROBERTSON, et
2^
T2">< J/2^:
~~

Z
7o H.
zºe .*
z e ...’

/
z ,” /...’
HEDGE AND .” 2^

/
/"

"
60 H. FIELDHOUSE 2.

KINGERY,
et

al.

- scort
O
5
: O
4

so H

20
H

SOLID LINES IN UPPER SET OF


PLOTS INDICATE TEMPERATURE
RANGE OVER WHICH MEASUREMENTS
-
io
H// WERE PERFORMED
|.
O

|
I

400 8OO 12OO 16OO 2OOO 24OO 28OO


O

TEMPERATURE, T/*c
T
FIGURE 9.62. Variation of for Various Measurements of Thermal
Kat
, l

Conductivity of UO, 95 Percent Dense.

power-producing rod which surface temperatures attained are


in

likely higher, differences between the


or
be
of

of

the order 400°C


to

by

data predict power output capabilities differing


of

three sets less


percent.
15

than

Flua: Suppression
of

(b) Effects

q(r), heat generation rate, are shown


In

Fig. 9.63, variations


of

for two important cases, the one flux suppression from surface
of

to

U* enriched pellet, the other plutonium


of

of

that due
to

center
on a

buildup natural UO, pellet after 10,000 MWD/


of

the surface
a

Ton—UO, exposure. The effects these nonuniform heat generation


of
on

patterns temperature distribution pellet with surface tem


in
a

57.4789 O–61—38
580 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

I.O.2 H.
-I I I I

ſ
A. FLUx VARIATION DUE TO U235 M
1.00 H Loading in 769 wºo U235 / || – 1.0
ENRICHED PELLET W
O.98 H
B. FLUx VARIATION DUE TO Pu Z
0.96 H BUILDUP IN NAT. UO2 PELLET Z | – O.9
ExPoSED To loooo Mwd ſt-Uoz ( |
o.94 – ExPOSURE |
q. co
|
O.92 H
#
| – oa g
-> | g
O O.90 H. O
|
z: | z
o 0.88 H o.7 o
| T
5 |
5
90 O.86 H go
o: | or
H.- H.
go.e4 H MEASURED ! -Ho.6 #
|
à os2 | ă
% | | ?
u- O.80 | – O.5 T.
|
O.78 |
CALCULATED
|

ozzl-Hos
**
O.76
Normalized for sAME i

...’ surface HEA ºf


MEASURED SPECIMEN
Asi
= 0.4

|- *
o.74 H.
|

O O.I O.2 O.3 O.4 O5


FUEL PELLET RADIUS, cm

FIGURE 9.63. Fission Distribution for Two Cases of Nonuniform Heat


Generation.

perature of 400° C are compared in Fig. 9.64 with the temperature


contours developed in the case of uniform heat generation. In each
case, the power output from the pellet was held constant, and the
data of Kingery, et al., in Fig. 9.62, were used to develop the tem
perature patterns. It is apparent that enhanced generation of power
at the pellet surface reduces the central temperatures attained. As
will be shown below, this expedient for reducing central temperatures
has been utilized by packing the fissionable oxide annularly about a
hollow or nonfissionable core.

(c) Effect of Fuel Element Geometry

Equation 9.31 has been derived for the case of a cylindrical pellet
conducting heat radially. Following the treatment of Robertson,
similar expressions may be derived for other important geometries
If,

Eq. 9.31, the heat generation rate


in

[133]. constant
is

throughout volume, Eq. 9.31 may integrated


be

to

the rod yield


582 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

stant power output per unit length of rod, the central temperatures
attained are independent of rod diameter. This property of param
eter
ſkat was discussed in Sect. 9.4.9 and shown to result in yield

of a constant fraction of fission gas release and of the same number of


fission gas atoms external to the oxide for the same - value of the
parameter.
For the case of an infinite slab of half thickness a

J.Ta
K.Tat-g
K(T)dt=} (a”—z”)

Qa a”—z’
2 a.” Eq. (9.32A)

where r is distance from the center of the slab, and

ſ
T.
Ta
K(t)at-". -
Qa
2 Eq. (9.32B)

and for the case of slabs of width d

0.

=Wi.
For a hollow cylinder of outer radius a, inner radius b, cooled at radius
a only,

ſ. K(TAT- [w-r) — 2b^lm


º
Ta-a-
–9
zºſº ”)-””;:
"
Eq.

T2 (9:33A)
and,

ſº-º-own
ſ.

K(TAT-
;

[-º, 2b2
-

–% lm
;

_W 1–?”,
*in4

T4t 72–72 Eq. (9.33B)


h

hollow cylinder
a,

For inner radius cooled both


b,
of

outer radius
a

inside and outside such that the surface temperatures T, are


T.

and
equal, for the annular ring radius R3'3d,
of
IRRADIATION EFFECTS IN URANIUM DIOXIDE 583

J
T,

Ktar-4(*-*-*}
(I 0.
Eq. (9.34A)
Ta
4 R

and for b3r-3P

ſ
T,
(R-Y)-2R-in b
T, Kºtºr-1ſ
4 }
where
a2+-b%
R2=
2
(1+in'.
and

ſ
TR
_ ('TR
KTAT=J.T, KTAT-5
_Q,a
{
- iº
2R2
ºn
a
#}
R
{-ſ}.
b 2R2 ,
Eq. (9.34B)
–% in #}
For the important case in which neutron flux is suppressed in the
interior of a fuel element by a uniform addition of absorbing or fission
ing isotopes, the resulting flux distributions have been discussed by
Robertson and by Adam [133, 136]. At steady state when the scattering
cross section for neutrons is much greater than the absorption cross
section, the neutron flux distribution is given by the diffusion equation

DV*@(r)—X,ó(r)=0 Eq. (9.35)

where 2,
is the macroscopic absorption cross section, b(r) is the neu
tron flux at any point within the fuel, and D is a constant. For a
cylindrical rod, the solution of this equation is

d(r)=d(0) I,(Kr) Eq. (9.35A)

in which p(0) is the neutron flux at the rod axis, I, (Kr) is the Bessel
function of imaginary argument, and the constant K is the reciprocal
diffusion length. Since, for uniform concentration of fissionable
isotopes, the heat generation rate is proportional to the neutron flux,

q(r)=q(0) I,(Kr), Eq. (9.35B)

and substituting in Eq. 9.31

ſ word-40ſ. r I,(Kr)dr=
q(o)rli(Kr)
*——K(T)
dT
rº,
Eq. (9.35C)
584 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

.
which on further integration yields the expression analogous to
9.31A and B
* Tyar Qaſ I.(Ka)-I.(Kr)

12,
Kitar-4

J.
Ka (Ka)

{
1/2

I
_Wſ I.(Ka)—I.(Kr)
–% Eq. (Q 31
(9.31C)
(Y
1/2 Ka (Ka)

II
and
Te
_Qaſ I.(Ka)–1
J. Kºtºr-4( 1/2 Kali (Ka)

_W I,(Ka)—1
Eq. (9.31D)
T4t 1/2 Kaſi (Ka)

U
Similarly, for
an

infinite slab of half thickness

a
T. _Qaſ cosh(Ka)-cosh (Kr) Eq. (9.32C)
J.

K(T)dT–% T1/2 Ka sinh (Ka)

}
T. _Qa
K(T)dT= {i}; (Ka)—1
cosh
Eq. (9.32D)
J.

(Ka) sinh
2

and for hollow, externally cooled cylinder internal diameter


of
a

b
kar-º
J...

(Kb)|Jo (Ka)—Joſkr)]+.J. (Kb) [I,(Ka)—I,(Kr)]


{!

1/2Kaſ (Ka).J., (KD)—I (Kb).J. (Ka)]


II

Eq. (9.33C)

"Kºtºr Kºſº): (Kºſº):


29

\ſ

(Ka).K. (KD)—I, (Kb)K. (Ka)


ſ

(I

T. 1/2K'ab
Eq. (9.33D)

Both Adam and Robertson as well as Howieson reference methods


of
A,

determining the value the diffusion length [133,136, 137]. Rob


of

ertson recommends the approximation cases where the flux suppres


in

sion not large,


is

;)

K*=3XaX, — 0.8
X,
(
X,

are the macroscopic absorption


X,

where and and total cross sec


tions. Appropriate microscopic cross sections for enriched UO, fuel
IRRADIATION EFFECTS IN URANIUM DIOXIDE 585
1.OO
iss- *- sº-lºvsk." I

S SS
- i i I
n (Ko) V o
- -
-
O.95 H. URANIUM OxIDE DIAMETER –
| IN INCHES

is oso F
-||
^
^
-
N.
*: | º
-
-
lº-
|
T. o.85 H.
N
is
7|s
* }=c 0.80 F
NATURAL.
URANIUM &
N.

YS
N.
-
N.
5 lº º
º- N.
- I owoO.75 H. -
&
& N
O.7O H. 2X, -
..a

O.65
O
I
2
I
4
L
6
ºt
2Nººk
8 |O
I
12 14
I
16
I
18 20

L- l l
ENRICHMENT,
l i
wºo
1
U” IN TOTAL U
J l l 1 i
O O.25 O.5 O.75 I.O 1.25 1.5 1.75 2.O 2.25 2.5
-
Ko

FIGURE 9.65. The Dependence of fri (Ka) on UO, Diameter and Enrichment for
95 Percent Dense, Solid Oxide Cylinders [133].

rods in a thermal neutron spectrum corresponding to 40° C moderator


temperature are given below in barns:
Uts, U*
Oa-------------------- 578 2. 33
O't -------------------- 586. 5 10.8

The agreement between measured and calculated flux suppressions


in a UO2 pellet is shown in Fig. 9.63. In Fig. 9.65, a family of
curves attributed to Morison is presented, relating the function on the
right side of Eq. 9.31D to U” enrichment and rod diameter for
the important case of high density (95 percent of theoretical) UO,
rods [133].

(d) Effect of Cladding

The information presented above relates the dimensions, conductiv


ity, and power generation within the oxide to temperature attained
within the oxide. However, in practical fuel elements, the oxide is
contained within a metal sheath, and the temperatures attained in the
oxide are similarly influenced by the conditions of heat transfer be
is,

tween the fuel and the metal sheath. It fact, the development
in

suitable model for describing this heat transfer which


of

of

one
is
a

principal subjects
of

the Sect. 9.7.


(1) GAP CoNDUCTION MoDEL.
In

the model which has received


principal attention the past, has been assumed that the fuel
in

it

within the metal cladding and that transfer


of

remains centered
586 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

heat to the sheath occurs through the gap calculated to exist at


temperature between sheath and fuel, correcting for thermal expan
sion of clad and fuel material. The surface temperature of the
fuel, T., is then related to the clad surface temperature, Tm, for
cylindrical fuel elements by the expression:

ln - In
T.-Tº-Q,c
#FF: c g
Eq. (9.36A)

where Q, is the clad surface heat flux, c is the outer sheath radius, g
is the inner sheath radius, a is the fuel radius, and K. and K, are the
thermal conductivities of clad and gap medium, respectively. For
a plate element,

T.-Tº-Q, Eq. (9.36B)


***"...]
where c-g
is the clad thickness and g-
a the width of gap between

is,
fuel and clad. In these expressions, the most critical item

of
course, the temperature drop associated with transfer

of
heat across
the fuel-clad gap. The importance this temperature drop can
of

appreciated from Fig. An increase surface tempera


be

in
9.62.
by

ture, Ta, decrease the power output capabilities


to

400°
C

seen
by is

one-third, whereas drop


of

fuel element allowable central


in
a

temperature the same amount decreases power output by only


of

percent. This arises from the relatively much higher thermal


10

conductivity low temperature and the


of

ceramic fuel materials


at

1/7' dependence this property.


of

The gap conductivity Eq. 9.36A and


be

can written
in

in
B

the
form

"-ºx
(T.

K.-K...+;

T.T.T.T.TH.
(#):
1)

(iii)
"Lº" ("..")
~K,...,
2

Eq. (9.36C)

which Kea, the conductivity the gaseous medium separating


of
in

E, is

fuel and clad, and E, are the emissivities fuel and clad, respec
of

T,
T,

tively,
or,

the Stefan-Boltzmann constant, and and are the ab


is

solute temperatures the fuel surface and clad interior, respectively.


of

seen from this expression that, even for poorly conducting


be
It

can
a

Kr and Xe, heat transfer through the


as

gas the fuel-clad gap such


by in

by

gap radiation becomes comparable with transfer conduction with


gaps 0.01 cm only fuel surface temperatures above
of

of

the order
at
IRRADIATION EFFECTS IN URANIUM DIOXIDE 587

1,000°C, and with gaps 0.002 cm thick, at fuel surface temperatures


above 1,700° C. Thus, for most applications, this model predicts a
sensitivity to the nature of the gas medium between fuel and clad.
Values of thermal conductivities of gases likely to be encountered in
fuel elements may be obtained from Refs. 138 and 139 and some
typical values are listed in Table 9.18. The marked effect of the nature
of the gas on temperature attained in highly-rated fuel elements with
high surface heat flux can be appreciated from the last column in this
table. A consequence of the model discussed is that, as burnup of the
fuel progresses and the initial gas medium such as helium used in the
fuel-clad gap becomes contaminated with Xe and Kr by release from
the fuel, the permissible thermal ratings of the fuel element deteriorate
significantly as the thermal conductivity of the gas bond decreases.
Ubisch, Hall, and Srivastav point out that the thermal conductivity of
rare gas mixtures can be obtained as the geometric mean of the con
ductivities of the pure constituents [139].

TABLE 9.18—THERMAL CONDUCTIVITIES OF GASES (139)

Temperature
Tempºrature, K., whem—"C Temperature" drop through
Gas C variation—S 0.002cm gap at
150w/cm2–* C

29 0. 00152 0. 72 197
He----------------------- 305 0. 00238 |_ _ _ _ _ _ _ _ _ _ 126
520 0.00308 |_ _ _ _ _ _ _ _ _ _ 97
\e----------------------- 305 0.0007.44 |_ _ _ _ _ _ - - - - 403
-
29 0.000182 0. 67 1, 648
A------------------------ 305 0.000294 |_ _ _ _ _ _ _ _ _ _ 1,020
520 0.000383 || -- ---- - 783
29 0.000097 0.88 3,090
Kr----------------------- 305 0.000162 |- - - - - - - - - - 1, 850
520 0.000224 |_ _ _ _ _ _ _ _ _ _ 1, 340
29 0. 0000596 0. 88 5, 030
Xe----------------------- 305 0 0000965 |_ _ _ _ _ _ _ _ _ _ 3, 110

ºn
520 0.000140 |_ _ _ _ _ _ _ _ _ _ 2, 140

15.8 t Kr, 84.2 per -


nt Xe-- ºf
- -- - -- -- --- -- -
29 || 0 0000623 0. 86 4, 810
520 0.000143 ----- ----- 2, 100

50 percent He, 7.9 percent -


29 0.000394 || - - - - - - - - - 762
Kr, 42.1 percent Xe- - - - - -
520 0 000815 -- - ---- 368

50 percent He, 25 percent


29 || 0 0004 11
-729
Kr, 25 percent Xe- ---- ---
| 350
().

520 000857

Kr,
A,

percent 7.9 percent


50

1, *2,

29 000107 800
0

42.1 percent Xe-


ºnt Ae
.
.
.
.
.
-
-
-
-
-
-
-
-

520 0.00023 320


|

.
.
.
.
.
.
.
.

H.0-------------------- 620
().

400 000483
|_
_
_
_
_
_
_
_
_
_

Conductivity proportional T, (T K) (139].


to

in
*

*
588 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

In estimating
the temperature drops required for the first model, it
is necessary to correct for the change in gap as the oxide temperature
changes. The following expression has been derived to describe the
gap dimension [107]:

(*, *-T)+cºſ(*, *-T)

Eq.
(0–0), (0–0)-awa (93.
-(g

-(g
the gap existing temperature and a),
in

at
a)
which

T, is

is
r
that assembly, assembly temperature, co,
at

the and and

a.
are

is

a
expansion coefficients UO, and sheathing material,

of
the thermal
respectively. Note that this expression accounts for the expansion

of
on

the oxide the basis temperature which


of the average between

is
a
the central and surface temperatures. This result derivable from

is
the temperature gradients given by Eq. 9.31

if
assumed that

is
it
the UO, expands, for example, like

no
sand, with

of
column strain

a
single particles.

of
interaction between hot and cold surfaces De
crease gap dimension with increase fuel temperature compensates,
in

in
some extent, for the poor thermal conduction properties

of
to

the gas
equilibrium gap dimension will
an

bond, and maintained, accord

be
ing the model presented, which will initial clearance,
be

of
function
to

heat flux, oxide temperature, and gas medium conductivity. Under


conditions such that gaps are calculated close,
to

assumed that
is
it
the surface temperature
the oxide, Ta, becomes equal the tempera
of of

to
the cladding (T,).
of

ture the inner surface -

(2) CoNTACT CoNDUCTANCE MoDEL. The model presented above,


which will
as

the gap conduction model,


be

as to

referred the text below


in

has been recognized being physically unrealistic two very im


in

portant respects. does not account for the fact that UO,
It

cracks (as
will stresses imposed
be

further discussed later) under the thermal


during operation and that the cracked segments must certainly relocate
the cladding, thus, dis
of

themselves contact with the inner surface


in

tributing the gap volume through the cross section


of

the element.
Secondly, the fragments which are formed are macroscopic dimen
of

undoubtedly elastically
to

sions and deform least accommodate the


at

uneven thermal strains resulting from the temperature gradients


through the fragments. Thus, more realistic model
of

heat transfer
a

contact,
be

from fuel
to

to

at

sheath should assume the fuel least


in

in

local areas, with the cladding exerting pressures


at

the contact areas


initial gap and operating temperature.
of

which are function Not


a

by

only the average expansion


as

the oxide, given Eq. 9.37,


of

important but also the elastic thermal distortions, since


be

should the
as

as

the pressure
of

latter should affect the size the contact areas well


applied. To evaluate the consequences this latter model, which may
of
IRRADIATION EFFECTS IN URANIUM DIOXIDE 589

be denoted the contact conductance model, consideration must be given


to the factors affecting thermal contact resistance.
Cetinkale and Fishenden derived an expression in terms of dimen
sionless quantities for determining the thermal conductance of two
solid surfaces in contact as a function of surface finish, medium in
which the surfaces are contained, contact pressure, etc., by calculation
of the isotherms at the contacting points of the two solids [140].
Their expression is

U-1+–F#–E–5

Eq.
(Aſ-
(9:38)
K
ºn -1}
which the conductance number,
in

U
is

U --K,'
ſub
Eq. (9.38A
(9.38A)

.
thermal per unit area, K,
conductance thermal conductivity
is

is
of w

the medium which the contacting surfaces are immersed, the


in

is
8
as

mean distance between the surfaces, defined

6=eff. Eq. (9.38B)

constant less than one which depends on the shape of surface


is
a
e

irregularities, and the arithmetic mean distance between the


3.
is

surfaces. The constriction number defined by the expression


is
C

c-à-V; Eq. (9.38C)

contacting surfaces and total surface,


of

which
r,

and are radii


in

respectively, A, and A. total contact,


of

of

and are the areas solid and


respectively. further defined by one the following relations.
of
is
C

For plastic flow,


P

the Meyer hard


M

which the average pressure applied and


P
in

is

to is

ness. Where pressure, Pmax, applied leading plastic flow and


is
a

together pressure under elastic strain,


at

the surfaces are then held


P

C= P.P2/3 Eq. (9.38E)


M
The fluid thickness number given by
is
B

= Eq. (9.38F)
B
+
e
590 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

in which the terms were previously defined; r. may further be related


to measured quantities by the relation

r. =V A. C. T Eq. (9.38G)

in which W and T are constants, and A. is the sum of the wave lengths

of the roughnesses of the two surfaces. The final quantity, the


conductivity number, K, is defined by

_K. Eq. (9.3SH)


s

in which K, neglecting radiation and thermal accommodation at the


surfaces, is the fluid conductivity, and K., the equivalent conductiv
ity of the two solids of conductivity K, and K,
is expressed as

/
1

K.T2
1

(#)
1 .. 1

Eq. (9.3SI)

Thus, in application of Eq. 9.38, U is defined by Eq. 9. 38A and


9.38B, in which, for ground surfaces, e was found to have the
value 0.71, and 3. is found from surface roughness measurements; B
is defined by Eq. 9.38F, 9.38B, and 9.38G in which W is given the
experimental value 0.0048, T the value 5/3, and A. is also found from
roughness measurements; C is defined by Eq. 9.38C, 9.38D, and
9.38E in which M
is obtained from the Meyer hardness of the softer
metal and applied pressure P; and K is obtained from Eq. 9.38H
and 9.38I.
Boeschoten and van der Held assumed a simpler form of Eq.
9.38 as follows [141]:

u- usas-Fu, Eq. (9,39)


in which
_K,
**** Eq. (9.39A)
and
u,-2nr}K, Eq. (9,39B)

in which n is the number of conducting spots per cm3. From the


assumption that the average pressure on the contact spots for a con
tact load P is 0.6 M, n and r are related by the formula

P -
H-0.6M Eq. (9.39C)

or combining Eq. 9.39B and 9.39C

_PK,
use ºf Eq. (9.39D)
592 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Thus, the contact conductance model, like the gap conduction model.
predicts a sensitivity to assembly clearances between fuel and sheath.
Since, however, the thermal conductance at contacts is a sensitive func
tion of pressure, it predicts an effect of contact resistance even beyond
the point at which the gap is calculated to close by differential fuel
sheath thermal expansions. Furthermore, because of the sensitivity
of contact conductance to hardness of the contacting materials, the
mechanical properties of the sheathing material are important in
determining interface temperature drops. The gas atmosphere in the
gap is predicted by both models to affect interface drops. Since, how
ever, the gap in the case of the contact conductance model is always
smaller than that in the gap conduction model, the effects of conduction
may be modified in the former case as the gas gap approaches the mean
operating temperature.
its
free path of the gas at The modification
by such effects K, Eq. 9.38H was

of
introduced the values
in

in
discussed by Cetinkale and Fishenden and, neglecting again radiation
by

effects, may expressed the following relation [140]:


be

K- K. Eq.
#:
(9.38J)
1+(++-);(Y-H 1)öuC,
(11 (12
and, Ln->8,
as

K.--yº”

Eq. (9.38K)

(++-)-L
by

viscosity,
u,

which defined the relation


in

is

u-mo VL. is K.,

the gas density, the average velocity the gas atoms,


V

of
is

is

is
p

the thermal conductivity gas, the mean free path,


of

the Ln
is

a
C,

constant=0.495, specific heats, the specific heat


of

the ratio
is
is
y

constant volume, and


a,

are the thermal accommodation coef


at

as

and
apparent
of

of
at

the gases
It

ficients the surfaces solids and


2.

is
1

from this relation that very low conductivities through the gas gap
may obtain the latter two quantities are small.
of

the values
if

Having reviewed above the heat transfer conditions which may ob


tain metal-sheathed oxide fuel elements, the experimental observa
in

tions on thermal behavior of fuel elements will be discussed. Most


of
on

the work hitherto reported has been obtained densely sintered UO.
fuel rods; some observations samples containing UO, powder
on

be

packed various densities have been reported and will also de


to

the greatly different properties


of

of

scribed below. Because the fuel


these cases, the observations are described separately.
of
in

each
592 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

Thus, the contact conductance model, like the gap conduction model.
predicts a sensitivity to assembly clearances between fuel and sheath.
Since, however, the thermal conductance at contacts is a sensitive func
tion of pressure, it predicts an effect of contact resistance even beyond
the point at which the gap is calculated to close by differential fuel
sheath thermal expansions. Furthermore, because of the sensitivity
of contact conductance to hardness of the contacting materials, the
mechanical properties of the sheathing material are important in
determining interface temperature drops. The gas atmosphere in the
gap is predicted by both models to affect interface drops. Since, how
ever, the gap in the case of the contact conductance model is always
smaller than that in the gap conduction model, the effects of conduction
may be modified in the former case as the gas gap approaches the mean
operating temperature.
its
free path of the gas at The modification
by such effects K, Eq. 9.38H was

of
introduced the values
in

in
discussed by Cetinkale and Fishenden and, neglecting again radiation
by

effects, may expressed the following relation [140]:


be

K- K. Eq.
#:
(9.38J)
1+(++-);(Y-H 1)öuC,
(11 (12
and, Ln->8,
as

K.--yº”

Eq. (9.38K)

(++-)-L
by

viscosity,
u,

which defined the relation


in

is

u-mo VL. is K.,

the gas density, the average velocity the gas atoms,


V

of
is

is

is
p

the thermal conductivity gas, the mean free path,


of

the Ln
is

a
C,

constant=0.495, specific heats, the specific heat


of

the ratio
is
is
y

constant volume, and


a,

are the thermal accommodation coef


at

as

and
apparent
of

of
at

the gases
It

ficients the surfaces solids and


2.

is
1

from this relation that very low conductivities through the gas gap
may obtain the latter two quantities are small.
of

the values
if

Having reviewed above the heat transfer conditions which may ob


tain metal-sheathed oxide fuel elements, the experimental observa
in

tions on thermal behavior of fuel elements will be discussed. Most


of
on

the work hitherto reported has been obtained densely sintered UO.
fuel rods; some observations samples containing UO, powder
on

be

packed various densities have been reported and will also de


to

the greatly different properties


of

of

scribed below. Because the fuel


these cases, the observations are described separately.
of
in

each
IRRADIATION EFFECTS IN URANIUM DIOXIDE 593

9.7.2 Thermal Behavior of Sintered High Density UO, Fuel


Elements

(a) Structural Changes in UO,

A number of experiments have been reported in which sintered high


density UO, fuel pellets have been exposed in-pile in sheaths of vari
ous metals at heat flux ratings such that structural changes consisting
of grain growth or melting of the oxide have occurred. Such observa
tions are reported in Refs. 64, 95, 107, 144, 145, and 146. From
knowledge of the heat output and dimensions of the fuel element and
the radial position at which the structural change occurs, it is possible,
using the relations previously derived, to relate the structural change
T.e.

parameterſ Ta Kd7
to a value of the (where T.e. is the temperature

at which structural changes occur and Ta is the surface temperature of


the fuel) or, of more value to the prediction of fuel element perform
ance, if the value of T.e. is known from other experiments, to draw
inferences concerning the thermal behavior of oxide fuel elements.
Such procedures were first employed by Kisiel [145].
The nature of the structural changes which are observed is illus
trated in Figs. 9.67 and 9.68Figure 9.67 illustrates the
[64].
macro appearance of several cross sections through a fuel rod speci
men which, prior to irradiation, consisted of high density sintered
cylindrical pellets of about 0.357-inch diameter contained within a
Zircaloy-2 sheath with a helium-filled 0.008-inch diametral clearance
between fuel and clad. This specimen was operated at a heat flux
(154 w/cm” at sheath surface) such that temperatures leading to micro
structural changes were attained in the center of the oxide. Disregard
ing the cracking of the oxide, several interesting features may be noted

A. Pellet 1, High Flux B. Pellet 3, High Flux C. Pellet 26, Low Flux
End. Reduction Fac- End, Reduction Fac- End, Reduction Fac
tor,
.º/,

tor, A. tor, 2.
FIGURE 9.67. Molten Zones and Solidification Craters the UO, at Warious
in

the X-1—h—4 Specimen; 6.6X [64].


of

Locations
594 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

in these macrostructures.Progressing inward from the sheath surface.


the unaltered oxide structure is observed to exhibit, first, a grain
coarsening, followed at higher temperatures by the formation of
columnar grains radiating to the pellet center, with a void forming at
the center of the pellet. The structures of these various zones are
shown at higher magnification in Fig. 9.68, A and B. In Fig.
9.68A, the transition from coarsened equiaxed grains adjacent to the
unaltered grain structure of the original pellets to the columnar grains
terminating in the central void is clearly shown. In Fig. 9.68B,
the grain structure in a sample operated at a lower heat flux rating
than the specimen in Fig. 9.68A is shown to exhibit only equiaxed
grain growth as the center of the pellet is approached. Specimens
showing such grain growth do not exhibit the central voids shown in
Fig. 9.67.
If it is assumed that radiation exposure has little effect on the rate
of grain growth of sintered UO, at elevated temperatures, measure
ment of the grain sizes attained after radiation should furnish a meas
ure of the temperatures attained at any location if the kinetics of
grain growth were known as a function of temperature. Only a few
systematic measurements of this phenomenon have been reported.
Runfors, Schönberg, and Kiessling showed the effect of temperature
and atmosphere on UO, grain growth [147]. Eichenberg, et al., de
rived the curve shown in Fig. 9.69 relating temperature and exposure
time to grain diameter for UO, compacts previously sintered at 1,700°
6p.

C to a grain diameter of [64]. Robertson, al., pointed out that


et

grain given temperature exposure


the size attained with time and
of
a
by

may powder Vari


be

of

influenced the nature the oxide used [95].


ous experimenters have selected widely different values tempera
of

ture which grain growth begins specimens examined af


et in
at

to

occur
ter irradiation. Kisiel and Eichenberg, al., selected 1,700° in as
C

the temperature grain growth observed microscopically


at

which
is

specimens prepared from MCW oxide powder presintered for hours


8

[64, 145]. Robertson, al., and Runnals utilized


at

to

1,675° 1,750°
et
C

for the temperature which grain growth (esti


to of

at

value 1,500°
C
a

grain diameter 20p)


be

on
of
at

mated observable observed


is
a

macroexamination [107, 135]. From Fig. 9.69, apparent that


it
of is
T,

the temperature grain growth, exposure time


of

function
is
a
a
,

and the experimental method whereby grain growth


as

detected
is
its

temperature alone. The type preirradiation


of

of

well
as

oxide and
sintering history will also influence grain growth.
The genesis the columnar grain structures noted Figs. 9.67
of

in

and 9.68 has been the subject considerable controversy. Kisiel and
of

Eichenberg, denoting the


as

al., have interpreted these structures


et

molten core during irradiation, extending


of

the loca
to

existence
a
º
IRRADIATION EFFECTS IN URANIUM DIOXIDE 595

FIGURE 9.68A. Left. Uranium Dioxide Columnar Structures in the X-1—h–2


Specimen; Original Magnification X100; Reduction Factor, 3%. (Photo
graphic series are not composite.) [64]

FIGURE 9.68B. Right. Uranium Dioxide Grain Growth in the Low Flux Fnds
of the X-1—h–1 Specimen; Original Magnification X250; Reduction
Factor, 34 [64].

57.4789 0–61–39
IRRADIATION EFFECTS IN URANIUM DIOXIDE 597

porosity from the partially dense as-sintered oxide in the center of the
pellet. Newkirk and Bates have shown that UO, can deposit in a den
dritic form by vapor deposition at temperatures as low as 1,800°C
[149, 150]. The irradiation of powder-filled sheaths to be discussed
below has resulted in the formation of dendrites similar to those ob
served in the out-of-pile vapor deposition studies and columnar grains
are found as a result of such a vapor deposition process. Thus, gross
vaporization-deposition processes may also lead to columnar grain for
mation at temperatures well below 2,760° C. Up to the present,
duplication of the structures noted in Fig. 9.68A at temperatures
below the melting point in sintered high density pellets has not been
reported. It is the
consensus of investigators that the temperature at
which columnar grain growth is observed even in high density pellets,
T. may several hundred degrees below the melting point
g.,

be

of
the
oxide,
C.
as

fact, low
in

as

1,800°

(b) Earperimental
of

Observations Fuel Elements

Observing the precautions detailed above concerning interpreta


of

tion the structural changes, there are tabulated Table 9.19


in

a
observations relating
to

structural changes noted


of

number the

in
exposure cylindrical UO, fuel rods. The irradiations encompass
of

Zircaloy-2, aluminum, and stainless-steel, atmos


of

metallic sheaths
pheres within the sheath helium, argon, and H2O, and relatively
of

initial pellet-sheath clearances and pellet diame


of

wide variations
ters. Using the methods previously discussed, the temperatures of
the pellet surfaces grain
Ta

of

are calculated from radial location


growth
T,

assuming 1,500° and 1,700° and assuming applica


=
,

as et C

or C

bility either the Kingery, al., Scott,


of

Hedge and Fieldhouse


thermal conductivity valuesextrapolated Fig. 9.62 [56, 57,
in

grain growth melting


In

134]. which columnar


or

those cases
in

reported, T., For most


or

2,750°
be

is of
C.
to

are taken 1,800°


is

the cases, the operating gap calculated from Eq. 9.37 and
is

all

the last entry


as

listed the table. Note that for specimens listed


in
no

the table, gap between fuel and clad


in

to

at

calculated exist
of is

operating temperatures. The temperature the inner clad surface,


Ti, the cladding ma
of

calculated from the thermal conductivities


is

terials, the ambient coolant water temperature, and clad-to-water


a

film coefficient assumed have the constant value of 5.7 w/cm?-? C.


to

Sheath-fuel gap conductances are then calculated from the pellet


u

(Ta-T). all cases except the hy


by

In

surface heat flux divided


draulic rabbit tests, the irradiations reported were relatively long
of

duration with the oxide accumulating burnups up 5,000 MWD/Ton


to

[144].
598 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 9.19–INTERFACE THERMAL CONDUCTANCE As

*: & 2.
~ gº
#

#3 =#|
# -
# #-> º ==

C
~

#+
|

-3 |

*
#,

#
33

5
s#

||

||
3–
- $ä
- £3 EE

|
*...* º:

3
tº-

:
: #: - -

:#*→
55: --> #3 35
-> ##

tº *
#3

##
&#

ºf
-

g
#3


#
c 35 *—, S

|
<

-
8-

- cm cm cm

C
w/cm2 w/cm tr/cm

|
X-1–F–5 (56)-------- 147 1.048Zr—2- 0.451 0.0076 He 23.8 |__________ 353

|
|

375
0.
X-1–F–3 (56).------- 163 1.048Zr–2- 446 0.0203 He - - 31.6
|

|
X-1–0–3 (56) 150 1.048__ 0.451 0.0064 He 25.35 ---------- 3f6
_
__
_

|
|
-
-
-
--
-
-

X-1–G–5 (56)------- 135 1.048Zr-2. 0.451 0.0064 He >31.8 -- 355


|

.
X-1-A-4 (56) 154 1.048Zr-2_ 0.446 0.0203 He 26 369
|-
|
-
-
-
-
--
-

--
--
-
-
--
-
|
().

X-1–H-5 (56) 141 1.048Zr-2. 0.450 0127 He ----------| 28.2 359


|
-
-
-
--
-

X-2R-IO (98)------- |----------


A.

80 2.033Zr-2. 0.95 0.011 33.5 309

X-2R-IR (98) 100 2.032Zr–2. 0.95 0.010 33 ---------- 3r


A
||
--
-
-
-
-

CR-V-k E–25 (98) 82 2.554Zr–2 1.21 0.010 39 3.11


--

A
||
|
IRRADIATION EFFECTS IN URANIUM DIOXIDE 599
DEDUCED FROM OBSERVATIONS OF STRUCTURAL CHANGE
T., Surface temperature of pellet, * C, cal- u–Gap conductance (w/cm *-* C) calculated
culated from T. , = 1,700° C and using T.

Tº:
from = 1,700° C and using UO, con-
ductivities of

of
Oz conductivities
U

É
---

ÉÉ
#3

#3

#3
:
#3
:
%#

Sc#

%#

Sc#
Ed

|
-: C

w/cm-* w/cm-º cru

C
C
670 ||------------ ------------

0.
788 760 0.4 0.428 553 0.004

g.
T. 1,500°C

------------ = 0.862 |------------


0.

0.
660 650 560 572 592 0.003
|

Teg. =2,750°C

|-
1,

0.283 ||------------
1,

140 100 1,030 |------------ 0.246 0.260 0.001

740 720 620 0.464 0.49 0.682 0.006

--
--
--
--
--
-
-
T. 1,500°C
s.
=

------------ 205 l------------


1.

620 610 510 0.634 0.713 0.003

<570 <550 <430 >0. 727 >0.8 >2.08 ------------ 0.002


--
--
-
-
--
--
-
-

T. 1,500°C
s.
=

------------ l------------|------------
1.

470 470 340 1.36 36 0.001


|

g.

T. =2,750°
C

0.195 ||------------
1,
1,

0.003
1.

400 310 280 0.173 0.190


--
-
--
-
--
-
--
-

T; =2,750°C
,

180 ------------ 0.199 |------------


1,

1,

0.172 0.194 0.005


1.

310 200

525 510 390 0.396 0.426 1.06 |------------ 0.007


--
-
--
-
--
--
-
-
g.

T. 1,500°C
=

530 420 300 |------------ 0.707 0.771 ------------|------------ 0.006

540 520 400 ------------ 0.488 0.538 1.37 ------------ 0.012


g.

T. 1,500°C
=

450 430 310 0.831 0.986 ------------|------------|----------


--
--
--
--
--
--

420 390 260 0.793 1.095 |------------|------------ 0.013


-
--
--
-
--
--
--

1,500°C
...

T.
=

335 310 180 - - - - - - - - - - - - 3.6 |----------|------------|---------------------


600 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 9.19—INTERFACE THERMAL CONDUCTANCE AS DEDUCED

.< *:::
& P
~ 2
:
C. 55 #8 =E
#
º EE
;:
E5
- 2 : tº 5 : ºź §*s £5
~ :
# 3E f: º
*:
= |
…3

*
#2 É
#5 #5 | #: # #3 |

3-5
8-
$3
:
a:E C3
3*
-*
3.
|
573
Ž5
**
=3
-:
53 &
*
*—,
&
* -

w/cm2 cm cm cºn w/cm •C

CR-V—k E-15 (98)-- 82 .21 0.008 41.5 312

CR-V—m H-2 (98)-- 75 . 205 || 0.015 35


*

CR-V-m G-1 (98)-- 73 . 13 0.020 37.5 343

CR-V-m G-2 (98)-- 73 13 0.025 36.5 343

-
CR-V-m G-3 (98)-- 73 13 0.031 32 348

CR-V-m H-3 (98)-. 75 2. . 205 0.014 35


* |

X-Rod GX (98)----- 57 ||3. .76 0.023 43 51

CR-IV X-2–g (102)- 113.5 1. 0.64 0.008 29.9 341


lºw-ºr-º
IRRADIATION EFFECTS IN URANIUM DIOXIDE 601

FROM OBSERVATIONS OF STRUCTURAL CHANGE-Continued


|
T. Surface temperature of pellet, *, C, cal- u–Gap conductance (w/cm2-* C) calculated

: 5 sº g 5 a 2- g #

i
#

#3
X: || 3
ta
| #
: Sc
#3
S4
|| 3
ta :
| #3 i
Sq
#
C

w/cm-* C w/cm-º C cm

365 335 200 l------------ 0.793 1.095 ------------|------------ 0.012

T. s. =1,500°C

290 270 130 ------------------------|----------|------------------------|----------

500 475 350 |------------ 0.415 0.478 1.93 I------------ 0.013

T. s. -1, 500°C

405 385 270 ------------ 0.827 1.042 l------------|------------|----------

425 420 290 l------------ 0.99 1.05 ------------|------------ 0.005

| T.s.-1,500°C

360 335 210 ------------ 4.81 ----------|------------|------------|----------

465 440 310 ------------ 0.665 0.835 ------------|------------ 0.006

T. s.=1,500°C

375 355 235 | - ---- 2.56 6.82 ------------|------------|----------

* 545 425 ||------------ 0.373 0.411 1.052 ------------ 0.008

T. s. = 1,500°C

465 450 340 ------------ 0.7 0.802 ------------|------------|----------

500 475 350 |------------ 0.415 0.478 1.93 I------------ 0.013


|

T. r. = 1,500°C

405 385 270 ------------ 0.827 1.042 ------------|------------|----------

340 310 170 ------------ 0.212 0.236 0.514 |------------ 0.015

t 1,500°C
...

T.
=

270 245 100 ------------ 0.28 0.316 1.25 ------------|----------

620 600 480 ------------ 0.455 0.49 0.914 |------------ 0.006


602 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 9.19—INTERFACE THERMAL CONDUCTANCE AS DEDUCED

& P gº
* 3 .c.
7. # :3
## ăg -= 5 #
E=
3.3
q2
# - #
:5 E
*3 :: : §* &
3.5
*:::
33
-: º
## ~
s
3. *> #:<
##
*# 3=
#5
Z3 --> # #:
--> :
&
- :
S:
.
##
5- E

:#3 až # 5 : s & & S. >


##
E- C
#
ſº *: 35 *-* | *—, R.

w/cm2 cm. crº cºn w/cm tr/cm • c |

CR-V-E (102)------ 105|| 1.408Zr-2- 0.63 0.008 A 26.8 |--------- 334

DB (143).----------- 182| 1.827Zr-2- 0.845 0.018 A 47.5 56 129|

T,...-2,200°C

DO (143).----------- 165. 1,828 S.S.- 0.8225|| 0.036 A 47.5 64 136

T.... =2, 200°C

DH (143).----------- 152| 1.817S.S.-- 0.815 0.043 A 40 55 lº


T. s.-2,200°C

FZ (143).------------ 164 1.832S.S.-- 0.845 0.008 A 55 !---------- 125


T...-2,200°C
- -

GB (143).------------ 139|| 1.832S.S.-- 0.8425|| 0.018 A 54.5 !----------


T.E.2,200° C– lºs

GC (143).------------ 135| 1.832S.S.-- 0.835 0.030 A 48 |----------


T. r. =2,200° C —
| 106

GD (143).---------- 145| 1.807.------ 0.83 0.043 A 49.5 !---------- | 93


T.s 2,200° C.—
IRRADIATION EFFECTS IN URANIUM DIOXIDE 603

FROM OBSERVATIONS OF STRUCTURAL CHANGE-Continued

T., Surface temperature of pellet, “ C, cal- u–Gap cond ºctance (w/cm?-- C) calculated
culated from T. g.= 1,700° C and using from T s... = 1,700°C. and using U On con- w
UO2 conductivities of ductivities of

-5
#

| :
#2
§
iſ #3
§
3
>.
#3 > §§
g
3.
#
F. Q.
is
>< a; #
E
|
Sc
#3
Sz
| #
ta
†=
E
||
Sc
# &
C

-
w/cm-* C. w/cm-º C cm

T. ...+ 1,500°C
|
510 500 390 ! ------------ 0.75 0.8 2.59 ------------|------ ----
- |
l 705 675 570 ------------ 0.316 0.344 0.498 ||------------ 0.007

T. g.= 1,500°C

585 575 470 ------------ 0.468 0.487 0.863 ||------------|----------

| 270 235 85 ||------------ 1.396 1.815 ------------|------- --------------- -


-
440 395 260 ------------ 0.633 0.74 1.5 -----------------------

T. g.=2,750°C

|
460 415 285 ------------ 0. 594 0.688 1.26 ----------------------

270 235 85 ------------ 1.368 1.85 l------------|------------|----------

4110 395 260 ------------ 0.603 0.708 1.48 ------------|----------

T. =2,750°C
...

-
310 265 120 ------------ 1.052 1.421 ----------------------------------
-
395 365 235 ------------ 0.62 0.695 1.5 --------------------- -

610 --------------- -------


0.

570 445 0.347 0.379 525


------------
=2,750°
C.
...

T.

4x40 -----------
0.

415 310 0.473 525 0.9


-
º
12

170 ------------- ------ g5 ºr, -


3

3.05 255 110 ------------. 0.33- 1.47 ------------ ------------- -


-
.

–- º
ºry - -
tº 2.

175
º'
fi 6

3iº 7.4% 12 -
a

----- --- - - -
-

-
-

zº 22*. * 24*.
*
ºn1
--
-
--
-
--
-
--
-

415, 3×5 250 ------- 22


1

24, * ... ºr:


*
2

ºn 2*, ------------ *** 24 - -


4

:
604 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS |
TABLE 9.19—INTERFACE THERMAL CONDUCTANCE AS DEDUCED

º: & 2. gº
->
c.
.c
->
7:
:- i. E
.c
E :
E =
33
:
#
3.
- 5#E - *3
|## |#
-- -
# s~ &
#3
#:
~ £3
*= #2*, *k- | #3 :
F.
~
Sz :
==
:
E.F. ~ + 3 Sº E
§: # 5: 35 . FE
‘-,
## * . . . .
až 5:
*>
£3 33 = #5 & & & R.
8- P: C a- -: 35 *—
w/cm? cra cm cm w/cm tr/cm •C

GJ (143).------------ 145| 1.755S.S... 0.8425|| 0.018 A 56 ---------- 64


T. s. =2,200°C

GU (143).----------- 123 1.832S.S.-- 0.8425|| 0.015 A 51 ---------- 9s


T.s.-2,200°C

X-1–F–2 (144)------- 155| 1.049Zr–2- 0.455 0.0038 H2O 26.2 30.3


- 369
|

X-1–F–4 (144)------- 161| 1.049Zr–2- 0.446 0.0203 H2O 13 21.5 373

X-1–G–1---. - ...---- 124| 1.049 0.452 0.0064 || H2O 17 22 348

X-1–G–2------------ 143| 1.049 0.451 0.01015 H2O 19.5 23 351


IRRADIATION EFFECTS IN URANIUM DIOXIDE 605

FROM OESERVATIONS OF STRUCTURAL CHANGE–Continued

T., Surface temperature of llet, * C, cal u–Gap conductance (w/cm2-9 C) calculated


culated from Trºs. = 1,
U Oz conductivities of
C and using from T. = 1,700°C and using UO, con-
ductivities oſ-
a
º*
~
> § § P. § É
|| ||i
#

#
|
|
is
#3
2
#
vº :is
#3 fSc is
#3
ld
£
ra
||
:3
75
&
| # &
C
I
w/cm-2 C w/cm-* C cm

16m 110 !------------|------------ 1.57 3.28 ------------|------------ ----------


285 240 90 |------------ 0.684 0.858 5.81 ------------|----------

220 18) 30 ------------ 1.098 1.63 ------------------------|----------


320 270 125 ------------ 0.603 0.778 4.95 ------------|----------

715 690 590 184 0.518 0. 55 0.810 ------------|----------

T. s.-1,500°C

600 585 485 ------------ 0.775 0.829 1.543 ------------|----------

Tes. =2,750°C

1, 225 1, 150 1,080 940 0.209 0.229 0.252 0.288 ----------


|
| 1, 155 1, 110 1,035 940 0.241 0.256 0.285 0.333 |----------

T. s.= 1,500°C
l
990 960 905 740 0.306 0.322 0.355 0.514 |----------

Tes. =2,750°C
|
1,595 1,540 1, 515 1,500 0.154 0.162 0.165 0.168 |----------

1.020 970 895 720 0.213 0.23 0.262 0.385 |---------.

T.s. - 1,500°C

70 845 765 520 0.274 0.288 0.344 0.832 ----------

Tes. =2,750°C

1.5S0. 1,530 1,500 1,480 0.116 0. 121 0.124 0.126 ----------

930 885 800 565 0.292 0.317 0.379 0.84 ||----------

-
T. s. 1,500°C

785 770 690 360 0.392 0.406 0. 505 ------------|----------

Tes. =2,750°C

1. 525 1,465 1,430 1,410 0.142 0.150 0, 155 0.158 ----------


606 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

TABLE 9.19—INTERFACE

5 g;
-
THERMAL CONDUCTANCE
7:
53. z
>.
AS DEDUCED
,
-
#. ă =} 53
55 3 =3 5% F- sº EE
Ž #.
3 £5 3
--

E
: 3 : :-
*- §: ##
##
c:...- --
$ä #5~ || ->
3 -
|##| |#5º: -
-
- - fi#3 -
#
F-
| #
: £;
C
#
* |#
| 3
|#
-:
‘-,5 5‘-,s
# ||

|

w/cm2 cm cra cºm w/cm w/cm • C

25–2 L-1 (145)- - - - - - 144| 1.049Zr-2- 0.451 0.01015 H2O 20 32 352

UO, Pellet rod, Mk 59 ||3.023Al 1.27 0.003 A 31 ---------- º


III (98) 57–S
dio
0

may firstinferred from this table that the gap conduction


be
It

previously described fails explain the interface temperature


to

model
drops which are calculated exist since, for practically all assump
is or to

conductivity Too, and Tea,


of

of

tions oxide assumed values finite


a
drop
no

interface temperature observed, whereas gap calculated


is

exist between pellet surface and cladding. Furthermore, the ob


to

be

servation that the interface conductances appear independent


of
to

argon, even
is or

whether the atmosphere within the sheath helium


is

though from Table 9.18 the conductivity


of

the former ten times the


latter, further argues rather conclusively against applicability
of

the
gap conduction model.
Concerning the applicability the thermal conductivity values, the
of

Hedge and Fieldhouse data appear quite definitely incompatible


be
to

with the assumption that Two. 1,500° and fail explain the
to
=

certain cases (CR-V-k, CR-V-m, G-1), even


if

observations
C. in

T,
be

1,700° No inferences can drawn from these data alone on


=
a

the applicability the Kingery, al., vis-a-vis the Scott measure


of

= et

(CR
T,

ments. The assumption that 1,500° one case listed


in
C
,

of of
In

V—k, E–15) with the measurements.


irreconcilable the case
is

the hydraulic rabbit tests which the in-pile exposure times were
in

T,

minutes, inferred from Fig.


of

of

the order value 2,200°


is
C
=
a

a
IRRADIATION EFFECTS IN URANIUM DIOXIDE 607

FROM OBSERVATIONS OF STRUCTURAL CHANGE–Continued

T., Surface temperature of pellet, °C, cal- u–Gap conductance (w/cm2° C) calculated
culºted from Tr... = 1,700° C and using -
from T.,., 1,700°C and using UO; con- a
U O: conductivities of ductivities of 5º:
k
~ cº ~ rº †

tº:
> 5 ~ º: >, 5

# 3.
§: #3; c

-
*

a
#=

Sº T =
#= #3 #3 ##

i
=

||

3.
3

#3


an

ld
º:

:
:

w/cm-°C w/cm-°C crº


-
905 855 775 520 0.308 0.34 0.405 1.06 ||----------
|

T:... 1,500°C

=
553 ------------|----------

0.
755 745 665 320 0.426 0.438

Tes.=2,750°C

0.314 l----------
|

1,090
1,

135 1.005 895 0.216 0.23 0.26


|

------------ 0.173 ------------|----------


0.

600 575 460 0.129 135

Tº 1,500°C
=
s

----------------------
0,

360 |------------ 167 0.23


0.

490 475 1615

[144]. As will noted from the table, this value gives much
be

9.69
better agreement between the melting and grain growth observations
T,

than does assumption


of

lower value for


in

these tests.
n.
a

In order correlate the observations obtained from fuel rod Sam


to

ples
of

various diameters, was found necessary


to

normalize the
it

data by use
of

the relation
assembled diametral clearance
-

diameter pellet
Such normalization appears quite plausible since, for the same sur
a

by

face and central temperatures, the relative amounts which initial


by

gaps are closed thermal expansion are proportional this ratio.


to

Therefore, Fig. 9.70A, the thermal conductance values Table


in
in

both from T,
C,

C,

9.19 calculated 1,700° 2,750° and



,
=

Too.
Kingery, al., UO, conductivity
of

the values and also from


et

T., -1,800°
T,

1,500°C,
as

are plotted
of
=

function the ratio


C

a
a

assembled diametral clearance


pellet diameter
rise precipitously
to

These values are seen


as

the ratios decrease below


the former case and 0.004 the latter and fall
of in

to

to

0.003 constant
in

a
C,

to as

value about 0.18 w/cm3–9 and 0.46 w/cm3–9 respectively,


C

the ratio exceeds 0.01. Furthermore, the gap conductance values,


608 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS
Go3.6
º 25s
ATTX I

TEST ATMOS CLAD STRUCTURE


CURVE I - EQUIAxED GRAIN • Loop Hez, GRAIN GRowTH (700°C)
2 H. GROWTHAT 1700°C, A LOOP He Zºe GRAN growthusood.
COLUMNAR AT 2750°C L LOOP He 22 COLUMNAR
GRAIN GROWTH(2750°C)
CURVE 2-EQUIAxED GRAIN
o LOOP Arzº, GRAIN GRowTH(1700°C)
GROWTHAT 1500°C, A LOOP Ar zz GRAIN GRowTH(soo-c)
COLUMNAR AT 1800°C D LOOPHe Zrz ColumnaR
GRAIN GROWTH(18Ooºc)
| THERMAL CONDUCTIVITYKINGERY

O
8 gº Al sheath- s—=
INC IN DIAMAFTER TEST
l I
O O.O. o.O2 O.O3
ASSEMBLED DIAMETRAL CLEARANCE
PELLET DIAMETER

FIGURE 9.70A. Effect of Clearance on Gap Conductance.

within the limits of accuracy of the data, are seen to be independent


of atmosphere (argon or helium). It is apparent that each set of as
sumptions concerning the temperature for structural change fits the
data equally well, and no choice of conductances value is possible from
these data alone.
Plotted in Fig. 9.70B are the results reported by Bain and Robert
son from hydraulic rabbit tests in which stainless sheathed fuel ele
ment samples were exposed in-pile for short time periods in low
temperature water [144]. While the same trends of decreasing
thermal conductance with increasing fuel-clad gap are noted as in the
loop test specimens, the former are observed to yield higher conduct
ance values. The rabbit data are observed to be in much better agree
ment with assumption of lower temperatures for the structural changes
(1,500°C for grain growth, 1,800° C for columnar growth) than with
the postulate of the higher temperatures.
Robertson, et al., relied heavily on the melting observations in the
hydraulic rabbit tests (as well as grain growth observations in loop
tests based on T, , = 1,500° C) to derive the curve shown in Fig.
9.62 for the variation of
ſkit with temperature [95]. A constant
interface gap conductance of 2 w/cm”-8 C was used in deriving this
IRRADIATION EFFECTS IN URANIUM DIOXIDE 609

—T- I i i i I
Test structural.
TYPE atmos. cladding CHANGE
-
He or A ATMOSPHERE-Loop
tests a Loop COLuºnAR
"20 ºr -2 drain growth (isoo-c)
>
- - curve 1- Fig. 9.70A
urve 2- Fig. 9.706
O
O
Loop
Loop
Hzo
Hzo
2
Zr-2
Zr-2
GRAINGrowth(15oo"c)
GRAINGrowth(17Oo-c)
* º Loop Hao zr-2 MELTING (27so-c)
S. a RAB8IT A s.s. GRAINGrowth(22Oo-c)
f A RAB8IT A s.s MELTING (2750°C)
* ...O.
H.
THERMALconductivity


al.
KINGERY,
et

i:
# -3 8

-
H

º:
- ATMOSPHEre
A

(RAB8It rests) A
--
-
H.

os

*
5

*. stEA
o Atmosphere COLUMNAR 18Oo-c
rº- GRAINGrowth 2750°c
{
l

{:
A
:

EQUIAxºD
|-
#s

GRAINGrowthU18oo” \o


C-)
-

-
I

I
o

I
I

I
o O.or ooz o.os oo<!
Assembled Diamétral clearance
PELLET diameTER

Clearance on Gap Conductance.


of

FIGURE 9.70B. Effect

curve, and the differences shape between the curve proposed by Rob
in

ertson and the other thermal conductivity data presented were ascribed
the thermal conductivity
of

the oxide under irradia


in
to

differences
tion [95]. The thermal conductivity curve derived from differentia
Fig. 9.62 indicates
of

tion drop from


in

Robertson's curve value


a

a
C,

0°C
of

0.04 w/cm–9 the con


at

0.024 w/cm3–9
to

at
C

800°
C, up

ductivity remaining constant the melting point.


at

this value
to

Consequently, temperatures above 1,400°


at

Robertson's assumed
conductivity data exceed those extrapolated 1/7' relationship
on
as a

from Kingery's measurements, the highest values yet reported, im


plying that mechanisms
of

heat conduction other than lattice conduc


tivity may operate UO, elevated temperatures. The possibility
in

at

Thus, Jamieson
of

be
of

such modes heat transfer cannot excluded.


by

forma
at of

and Lawson have theorized that transfer heat exciton


high temperatures and annihilation
at

tion lower temperatures can


appreciable temperatures
in

at

become lattice conductors above


by

1,000° Likewise, transfer


of

[151]. heat radiation has been


C

postulated, MgO, explain deviations from the 1/7'


of
in

to

the case
relationship thermal conductivity with temperature [152]. Thus,
of

former
of

have estimated that the transfer


at

the authors 1,800°


C
610 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

heat in MgO by excitons is comparable with lattice conduction.


whereas radiative transfer is about 25 percent of the latter. Simi
larly, Whitmore discusses the contribution to thermal conduction in
ceramic materials by other excitation processes such as by diffusion of
electron-hole pairs and by dissociated gas molecules [153]. However,
as will be discussed below, it is considered unnecessary to invoke the
operation of such phenomena to explain most of the observations of
the thermal behavior of UO, fuel elements.

(c) Operation of Defected Fuel Elements

Included in Table 9.19 and plotted in Fig. 9.70B are some observa
tions on fuel element samples operated in hot water loops with 0.005
inch intentional defects in the Zircaloy-2 sheaths. The values of gap

lie
thermal conductance derived from these data appreciably below
those obtained from fuel rods operated with rare gas atmospheres.
The following possibilities are presented explain the results.

to
UO,
To
of

(1) Oxidation NoNstoich.IoMETRIC CoMPosition. As


A

UO,

its
Fig. 9,43, exposure oxidation,
of

in
shown steam results to
in

equilibrium, UO,

or
composition above; lower oxygen
of
at

to

is
a

compositions are attained equilibrium with H2-H2O mixtures. As


in

Chap. (Sect. 5.3.1 (b), Table 5.4, nonstoichiometric


in

discussed
5

UO., exhibits much lower thermal conductivity than does the stoichi
oxide, the conductivity having
in of at

to

of
ometric 33° 85° value
C

a
UO, Furthermore,
1a.

as
w/cm–9 composition
at

0.034 shown
C

a
by

Runfors, al., grain growth UO, temperature


et

at

can occur
a

steam atmosphere than reducing atmos


in

almost 400° lower


of in
C

phere [147]. Both these effects will cooperate produce much


to

more marked structural effects stoichio


in

nonstoichiometric than
in

metric oxide operated similar heat ratings and will, thus, serve
at

to
explain the reduced thermal capabilities shown Fig. 9.70B. This
in

important fuel elements operated with


be
to

effect not considered


in
is

sheath, although undoubtedly applicable


in

small defects the


to
is
it

fuel elements operated with large defects which equilibration with


in

be

steam can occur. As discussed Sects. 9.4.10 and 9.6.1, can


in

it

diffusivity gaseous hydrogen


of

shown from the the interstices


of in
as

as

within defected sheath, well from the absence corrosion attack


a

the UO,
on

sheaths with small defects, that the atmosphere within


in

such fuel rod probably hydrogen-rich and maintains the oxide


in
is
a

Furthermore, the absence metallographic


of

stoichiometric condition.
O,

U,
precipitates within the oxide
of

of

evidence defected fuel ele


(Fig.
9.68) argues against the presence appreciable oxygen
of

ments
The difficulty oxygen analysis
of

to of

contamination. irradiated fuel


elements prevents obtaining more definitive answer this possibility.
a
IRRADIATION EFFECTS IN URANIUM DIOXIDE 611

(2) DECREASE IN GAP THERMAL CoNDUCTANCE. The reduction of


the gap thermal conductance by the substitution of poorly conducting
steam for helium in the clad-fuel gap is a possibility. This is dis
counted since, as shown in Fig. 9.70A, use of argon with even a poorer
conductivity than steam (Table 9.18) has an inappreciable effect on
thermal conductance as compared with helium-filled samples. It is
hypothesized more reasonably that the poor gap thermal conduction
can be accounted for by oxidation of the interior of the sheath by
reaction of the cladding metal with steam or water during pre-exposure
testing or during operation. As shown previously, oxidation of con
tacting metal surfaces reduces thermal conductance by factors of 4
to 5. Such an effect of oxidizing the contacting surfaces would be
anticipated to become more marked the better the initial contact or,
as shown in Fig. 9.70B, the smaller the initial clearance. Independ
ent verification of this hypothesis by exposure of undefected samples
with preoxidized inner cladding surfaces would be highly desirable.
In summary, observations of structural changes in fuel element
samples containing dense, sintered UO, pellets and operated in high
temperature water loops under simulated reactor operating conditions
can be satisfactorily correlated at least to burnups of 5,000 MWD/Ton
UO, by use of out-of-pile thermal conductivity measurements, reason
able assumptions concerning temperatures at which observed struc
tural changes occur, and gap thermal conductance values sensitive to
initial assembled gap but not to atmosphere within the sheath as shown
in Fig. 9.70, A and B. The performance of fuel elements in defected
sheaths exposed to the water coolant is less firmly established but,
again, appears explicable on the basis of a gap thermal conductance
still further reduced by oxidation of the contacting surface as also
shown in Fig. 9.70B. These conclusions involve a number of assump
tions as outlined above. However, some experiments described below
have been performed, the analysis of which permits substantiation of
the above postulations.

(d) In-Pile Measurements of Thermal Performance

The first of these experiments are those described by Eichenberg and


al.

further refined and carried on by Cohen, et [55, 154]. these ex


In

periments, oxide pellets are contained within heavy-walled stainless


a

which the temperature gradients can measured by


be
in

steel sheath
thermocouples placed several radial and azimuthal positions. From
in

these gradients, knowledge the thermal conductivity


of

of

the stainless
steel, and measurement the gamma heating capsule adjacent
of

be to
in
a

the sheath, the heat flux and heat generation rate the fuel can
in

thermocouple the fuel pellets records the


of
in
A

measured. the center


57.4789 O—61—40
612 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

central temperatures attained. Thus, each of the quantities necessary


Te - -
for measurement of Kd7" are directly determined. By perform
Tt
ance of successive measurements as the test reactor power is increased

J
T
stepwise, the variation of Kd7 with temperature can be deter.
T.
mined and compared with the various out-of-pile measurements of this
quantity. Furthermore, by variation of fuel-clad gap and atmosphere
within the sheath, the effects of these variables may be assessed. The
second type of experiment attributed to Hawkings and described by
Robertson, et al., consists of measurement of the center temperature
attained in an aluminum-clad cylindrical oxide fuel element, the heat
generation rate in the oxide being estimated from the measured heat
output of adjacent fuel rods [95]. Two such experiments, Mark III
and Mark IV,were reported by Robertson.
(1) MEASUREMENTs of HAwKINGs (Robertson). Robertson as
sumed in the interpretation of the latter experiments a fuel-to-clad con

ſ
T
ductance of 2 w/cm?-? C, and the values of KalT obtained at
0.

various central temperatures are plotted in Fig. 9.71. In these ex


periments, the assembled fuel-clad clearance is very small and, in fact,

50 I
—-v’
I i I i i I i i I
2.
KINGERY
COHEN
SUCCESSIVE START
UPS: 1-O, 2- © –
3-D, 4--, 5-A,
6-K), 7-4)

3.O -

:
AssumEd INTERFACE CONDUCTANCE
OF 2waſ Ts/cm2, *c

l l l l I l L L
o 2OO 4OO 600 800 IOOO 1200
T, “c

T
orſ

FIGURE 9.71. Variation Kd with Temperature.


T
o
IRRADIATION EFFECTS IN URANIUM DIOXIDE 613

in the case of pellet rod test Mark III, the aluminum (57S) sheath was
found to be increased in diameter after testing. The data of Cohen,
et al., for a series of test reactor startups in the case of a specimen in
which the oxide was press-fit within the sheath are plotted in the same
manner for comparison, again assuming a sheath-fuel conductance of
2 w/cm?-? C [55]. The latter data are seen to lie appreciably
above those of Hawkings and, in fact, for the first startup, agree quite
closely with the thermal conductivity measurements of Kingery, et

al.
The decrease thermal performance upon subsequent startups
to in
{56].
may
be

an
attributable irradiation-induced decrease thermal con

in
ductivity considered more probable, however, that
of

an or byIt
the oxide. is
cracking and

or
changes

of
the are caused deformation the sheath
explanation preferred,
of

relocation the oxide both. The latter

is
since irradiation-induced change conductivity

in
not clear how
it
is

could produce the erratic types changes thermal performance


of

in
characteristics upon successive startups (cf. startup with and

3
5), whereas gross physical changes such cracking and fuel relo as
expected give
be

cation would the erratic measurements noted. The


to
by

low values reported Hawkings (Robertson) are considered

to
arise

be
an

from the gap thermal conductance. As can


of

overestimate
seen from Table 9.19 and Fig. 9.70A, the calculated contact conduct
ance for this experiment from measurement grain
of

of
the location
growth after irradiation was,
at

most, one-tenth
of

the value assumed


by Robertson, and use this value for contact conductance brings the
of

pellet rod data better line with the experiments Cohen, al., and
of
in

et

Kingery,
al.

[55, 56,95].
of

the measurements
et

apparent by comparing the data


It

Cohen for zero clearance


of
is

with the out-of-pile measurements Kingery,


of
al., and those
of

et

Hedge and Fieldhouse plotted Fig. 9.62 that only


of

Scott and
in

the first are adequate explain his observations, since the latter
to

two measurements lie below his measurements and since irradiation


expected and has been observed only
be

would
to

decrease thermal
conductivity. On the basis this observation, the Kingery,
of

al.,
et

data alone have been used Fig. 9.70 for correlating the measure
in

ments from fuel rod operation and, the discussion below, for
in

reconciling the experimental measurements Cohen and Eichenberg.


of

by
be

Cohen that cer


of

Mention should made here the observation


tain sintered bodies prepared from laboratory-prepared oxide indi
cate thermal conductivities even higher than those which
he

had
bodies prepared from commercially
in

observed obtained oxide.


Thus, the possibility exists securing improved thermal perform
by of

UO, purified oxide.


of

ance from fuel bodies the use


Cohen. As discussed above, Cohen, Lust
of

(2) MEASUREMENTs
man, and Eichenberg assumed the applicability the Kingery out
of
614 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

2OO I I I I I I

9O ----
i-
CALCULATIONS BY FORMULA OF CET INKALE

f AND Fishenden
© FIRST STARTUP
Assuming Ki-o,
re - 150/1
o SECOND STARTUP
- —

2O H.

|
3 |

5
M=MEYERS HARDNESS
-
-

SECOND STARTUP -

O. I l | l l l I
O 5 iO 15 2O 25 30 35
psi

P, contact PREssure, lo”


,

FIGURE 9.72. Contact Conductance vs Contact Pressure at the Stainless Steel


UO, Interface Experiment Designed
to

the WAPD–22–11 Measure the


in

Effective Thermal Conductivity UO, during Irradiation at Materials Testing


of

Reactor (Zero Clearance, Kr-H3Xe Atmosphere, Production UO.) [55].

of-pile thermal conductivity data and utilized the information ob


calculate sheath-fuel contact conductance [55].
In
to

tained one

the experiments, WAPD–22–11 plotted Fig. 9.72,


of

in

in

which
the pellets were inserted the sheath with press-fit, was pos.
in

it
a

estimate the pressures developed


to

sible between fuel and sheath


IRRADIATION EFFECTS IN URANIUM DIOXIDE 615

by calculating the relative thermal expansions of fuel and cladding


and applying the thick-walled pressure vessel formulas. The re
sults obtained in several test reactor startups are plotted in Fig.
9.72 and indicate, as anticipated, a marked dependence of contact
conductance on pressure. The variations between the various start
ups are attributed to cracking of the fuel and to shifting and reloca
tion of the fuel particles in successive startups. The decrease in
thermal conductance after the first startup may be attributable to
plastic deformation of the stainless-steel sheath during the initial
startup with a consequent increase in cold clearance or to an irradia
tion-induced reduction of conductivity; the calculations of pressure
were based on the premise that the dimensions and conductivity of

|
fuel and sheath were unaltered by the exposure.
To relate these observations to theory, Eq.
9.38, derived by
Cetinkale and Fishenden was utilized, with the assumption that
Kr, thermal conductivity of the medium within the sheath, was
zero [140]. Under these conditions, Eq. 9.38 may be written

u––Pi—T
CK.
tan T'
Eq. (9.38L)
re
{}- }
the terms which have been previously defined. The value of
in
C was calculated using Eq. 9.38D in which M, the Meyer hard
ness number, 150 kg/mm2 at room temperature, was assumed to vary
with temperature proportionately to the yield point of stainless steel.
Calculations were also made assuming M=3 times yield strength,
ro,

as suggested by Wheeler [142]. For one-half the mean distance


between points
of

contact, was necessary


of

to

assume value
it

obtain agreement with the experimental The results


to

1500 results.
Fig. 9.72. apparent
of

these calculations are also plotted


It
in

is

by inspection Eq. 9.38L that qualitatively


of

the theoretical
relation agrees with the experimental observations, since, the pres
as

sure approaches the value M, sharp increase very large


of

to
in
u
a

Furthermore, the values calculated using


be

values should noted.


Meyer hardness for show reasonable agreement with the experi
M

mentally derived results, low contact pressures.


at

at

least
applicable here because
It

postulated that value


of

of

150),
r,

is
is

the fracturing the oxide and the thermal distortion caused by


of

of

the temperature gradients through the fuel particles. Such gross


distortions would tend increase the pressure the contact points
to

at

number, thus, decreasing (Eq.


of

but would lessen their the value


C

9.38C). The sharp average pressures


at

rise conductance noted


in

far less than the value may, thus, result from decrease
of

in
M

the
616 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

value of r, as the applied pressure forces the fuel particles to conform


to the sheath by cracking or elastic deformation.
The results of a number of Cohen's experiments are plotted in
Fig. 9.73 as
T.
Kd T
T.
T-T’
which Eichenberg and Cohen have designated as the smeared con
ductivity averaged between the center temperature T. and the inner
sheath wall temperature T. The data for zero clearance are those
replotted in Figs. 9.71 and 9.72. It may first be noted that a con
tinual drop in smeared conductivity occurs as the assembled clearance
increases from zero to 3.5 mils. Beyond this value of assembly clear
ance to a value of 0.013 inch, the conductivity remains unaffected.
It may be further noted, comparing the results of WAPD–22–2, 22–3,
and 22–4 with initial vacuum, helium, and Kr-H3Xe atmospheres,
respectively (at 14.7 psi assembly pressure in the latter two cases),
and 22–1 and 22–15 with helium and Xe-H Kr atmospheres, respec
tively, that atmosphere within the sheath has no apparent effect.
is,

The latter finding indeed, surprising result, since even the


a

model predicts
as an

of which

in
contact conductance effect medium
postulated, suggested by
Cetinkale and
It

contact made.
is

is

Fishenden Eq. 9.38K, that this insensitivity gaseous medium


in

to

results from low values of the thermal accommodation coefficients

an
on

or

the gases
of

of

either the two solid surfaces [140].


as

and both
Smoluchowski has shown that the conductivity
+ an
of

gas annular
in
a

al
space between two cylinders
of

of

thickness radii and is


is a

a
8

on 3

tered from its value K, when the accommodation perfect both


surfaces
a-Hö
Qa ln TaT
K-–37–
K,

by

given
to

value
a

º
a-H (I,

[alºrº-(#)]
6
...

f— AT

which -
in

5–%."

=; +;
1

1
IRRADIATION EFFECTS IN URANIUM DIOXIDE 617

I T I I I I I + T I T T
PELLET DIAMETER-O.357in.
SHAFT MATERIAL-3o4 stalNLESS stEEL

a KINGERYDATA-UNIRRADIATEDUO2 d WAPD22-3
(95%.DENSE) 8 O MIL DIAMETRAL CLEARANCE,
© WAPD22- . . HELIUMATMOSPHERE,PRODUCTION
O DLAMETRALCLEARANCE,I Krt UO2
3 xe ATMOSPHERE,PRODUCTION O wand 22-2
UO2PELLETS 80 MIL DIAMETRALCLEARANCE,
a wand 22-14 vacuuM,PRODUCTION Uoz
2.5 MIL DIAMETRALCLEARANCE, w wapd 22-9
| Kr + 3 xe ATMOSPHERE, 8.OMIL DIAMETRAL CLEARANCE,
PRODUCTIONUO2 | Krt 3 xe ATMOSPHERE,
PRODUCTIONUoz
09 H wapd 22-6
O MIL DIAMETRAL CLEARANCE,
Kr – 3 xe ATMOSPHERE,uoz is
PELLETS
wapd 22-15
08 H 3.5MIL DIAMETRAL CLEARANCE
IKr 43 x• AIMOSPHERE, Uoz
PELLETs (LO2%,ENRICHED)
wapd 22-7
13MIL DIAMETRALCLEARANCE,
0.7H. IKr 3 x• ATMOSPHERE,Uoz
PRODUCTIONPELLETs
FIRSTSTARTUP(SCRAMMEDAT 30 Mw)
SEcond starTUP

06 –
: wand 22
3.5MIL DIAMETRALCLEARANCE,
HELIUMATMOSPHERE, Mcw Uoz
EICHENBERG

: D waſ D 22-4
1.5MIL DIAMETRAL CLEARANCE,
HELIUMATMOSPHERE,PRODuc-
0.5+
Tion Uo. PELLETs

l
too 2OO 3OO 4oo ºoo soo 700 800 900 todo too 1zºo
uo cent RAL TEMPERATURE.*c

Figths 9.73. Smeared Thermal Conductivity of UO, wap L-22 Experiment


(Instrumented Capsule Irradiation Test at Materials Testing Reactor) [35].

and other terms are defined in Eq. 9.2-K (155].


the The ratio
of conductivities at the same temperature drop, or more informative,
temperature drops the same gas conductivity then,
of
the

is
at

ratio

AT" *L-(+,-1)
#=1 -2
a

lin
a
618 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

|O I I I I T
t
#- 1 vs a FOR VARIOUS GAP WIDTHS, cms

-
-

NG 8 =O.OO3

- i
| -
8 = O.O.

NG
O.OI

8 = O.O3

f
-

- 8 =O.I

O.OOl
l | l l —l
O.2 O.4 O.6O.8 I.O.
ACCOMMODATION coBFFICIENT, a

FIGURE 9.74. Temperature Drop Correction for Wall Discontinuity.

Assuming the gas medium is helium at a pressure of 1 atmosphere


and average temperature 500° K, the effect of variation in a on the

drop
temperature
correction.( #–1) is plotted for various gaps in

Fig. 9.74.” It is apparent that if g is as small as 0.02 for gaps of 30,


the temperature drop through the gas medium is more than 10 times
that which would be noted if the accommodations were perfect.
Values as low as 0.06 for the accommodation coefficient of He on
platinum and 0.017 for the accommodation coefficient of helium on

* J. M. Markowitz, Bettis Atomic Power Laboratory, personal communication.


IRRADIATION EFFECTS IN URANIUM DIOXIDE 619

clean tungsten have been reported, while a value of 0.027 for O2 on


U.O. was estimated by Blackburn " [156, 157]. Whatever the ex
planation for the lack of conductivity through the gas phase, its
consequences are quite important. Thus, the use of helium to serve
as a gas bond between UO, and the cladding would, from Cohen's
results, appear to be unnecessary. Furthermore, deterioration in
the thermal performance of sintered fuel elements by dilution of
the atmosphere within the sheath with released fission gases need not
be taken into account in oxide fuel element design. Rather, as shown
in Figs. 9.72 and 9.73, improved fabrication procedures to minimize
fuel-clad clearances are indicated by the results of Cohen as the most
effective measures which may be taken to improve high density oxide
fuel element thermal performance [55].
Application of the Cetinkale and Fishenden equation as modified
by neglecting gas conduction (Eq. 9.38L) predicts that, as contact
pressure decreases to zero, contact conductance also decreases essen
tially to zero. In most of these cases plotted in Fig. 9.73, it can be
shown that expansion of the UO, will be insufficient to cause contact
with the cladding; yet, a quite finite conductance is found even at the
largest gap sizes investigated(0.0065 inch). It is probable that con
tact of fuel with cladding
occurs in all these cases through cambering
and thermal distortion of the fuel pellets and, at appropriate heat
ratings, through cracking of the fuel. In fact, the apparent increase
in conductivity noted in Fig. 9.73 at central temperatures of 600° to
700°F in the case of the 22–7 and 22–15 experiments is not inconsistent,
as will be discussed in Sect. 9.8.1(a), with cracking of the fuel induced
by thermal gradient strains, thus, permitting relocation and better
contact of the fuel particles with the sheathing. The improvement of
conductivity in the 22–7 experiment upon the second startup may, in
particular, be cited as evidence for such improved fuel relocation after
the fuel cracked during the first reactor startup.
(3) CoMPARISON witH FUEL ELEMENT BEHAvior. The question
may well be raised that the results plotted in Fig. 9.73 are peculiar
to the experimental conditions employed in that thermal expansion
was insufficient to close the fuel-clad gap, whereas in fuel elements of
high thermal rating such assembly clearances are in general closed
when the fuel element reaches its designed operating conditions.
Thus, it is necessary to compare the results of Fig. 9.73 with those
plotted in Fig. 9.70, and tabulated in Table 9.19. In the latter case,
all

it will be remembered that the fuel element samples listed operated


with no gap calculated To make this
to

exist between fuel and clad.


Blackburn, calculated from effusion measurements,
of

The value more mechanical


is
*

than thermal property; however, the mechanisms are probably similar and the magni
a

tudes of the accommodation coefficients should be of the same order.


620 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
|
1.5 go I I I I I I

|- IRANGE of contact conductances Nored BY cohen -


$’ ovaLUE OF CONDUCTANCEAT Ti - 300°C
w
E
S.
$2 to L CURVE I - FIG 9.70A -
; CURVE 2 - FIG 97OA
uj
O
2
<I
5
- —
3
2
o
O
-]
#os H -
º
Lu
--
H.

O I I | l I l I
o O.O. O.O2 Oo3
ASSEMBLED DIAMETRAL CLEARANCE
PELLET DIAMETER

FIGURE 9.75. Comparison of Contact Conductance Calculated from Data of


Cohen, et al., with Observations from Fuel Element Samples [55].

comparison, the validity of the Kingery, et al., conductivity values was


assumed for all the experiments listed in Fig. 9.73 and for all gap
contact conductances calculated and plotted as in Fig. 9.70. In Fig.
9.75, the range of contact conductances calculated for the experiments
noted in Fig. 9.73 is plotted as a function of

assembled diametral clearance.


pellet diameter

The large range for each point plotted in Fig. 9.75 is rather mis
leading, since examination of the data indicates that, as the pellet
attains reasonably high temperatures and, in particular, undergoes
several reactor startups, the conductance values stabilize at a quite
constant value characteristic of the assembly clearance used. For
this reason, as well as to make the comparison at experimental condi
tions approximating those attained in the loop tests, there is marked
on each curve the value of contact conductance at an inner wall tem
perature of 300° C. Superimposed on these data points are the curves
of Fig. 9.70 A and B, used to correlate the data from undefected loop
test specimens. The agreement between the data of Cohen and those
IRRADIATION EFFECTS IN URANIUM DIOXIDE 621

of Fig. –
Too. 1,700° C and Tea =2,750° C, is noted
9.70, assuming
to satisfactory. Thus, these data seem to reinforce the con
be quite
clusions of Fig. 9.70 A and B, that the assembly clearance rather than
the operating clearance is important in determining thermal perform
ance and lend validity to the use of the curve shown in Fig. 9.75 to
predict fuel element performance. The divergence between the
data Cohen and those of Fig. 9.70A calculated on the basis of
of
T, , = 1,500° Cand T.,
= 1,800° C may be attributable to the lack of

al.
contact between fuel and sheath in the experiments of Cohen, et
[55]. However, some contact must the latter experi

be
encountered

in
explain the residual conductance about 0.18 w/cm?-?

of
to

ments

of C
large gaps. Note from Fig. 9.72 that gap conductance
at

even

a
0.18 w/cm?-? would imply contact pressure about 200 psi.

of
C

To test the applicability


of

the modified Cetinkale and Fishenden


equation (9.38L) the fuel element samples listed
to

Table 9.19, values

in
the experiments (using
of

of
conductance were calculated for several
150p) and are tabulated Table 9.20. While the calculated values
in
=
r

show, relatively, the predicted variation with assembly clearance,


quantitative agreement with the measured values, whether calculated
T,

1,700°C, unsatisfactory. The reasons


or

assuming 1,500°
=

is
,

for these differences are not known, and for this reason recom
it
is
mended that the experimentally derived curve shown Fig. 9.75

be
in

predict fuel element performance.


to

used

TABLE9.20–COMPARISON OF CALCULATED AND OBSERVED CON


TACT CONDUCTANCE WALUES FOR ZIRCALOY-2 CLAD LOOP
TEST SPECIMENS, 2,000 PSI EXTERNAL PRESSURE

Contact conductance—w/cm2–º
C

Diametral
-

Specimen assembly Grain growth temperature 1,700° Grain growth temperature 1,500°C
C

clearance
(cm)
Pressure Cetinkale- Table 9.19 Pressure Cetinkale- Table 9.19
||
|

(psi) Fishenden (psi) Fishenden

0.0076 13,400
1.

55 0.4 10,570 1.41 0.572


0.0203 4,850 0.0 0.246 ------------------------------------
0.0064 19,100 2.2 0.464 10,570 1.46 0.634
0.0054 7,710 1.21 0.727 4,850 0.97 1.36
0.0203 10,570 0.885 0.173 ------------------------------------
16,230 0.172 ------------------------------------
1. 1.

0.01.27 325
6,400 11 5,770
1. 1. 1.

0.011 0.396 13 0.707


7, 9,
C.

010 510 1.45 0.488 7,960 38 0.831


5, 4. 5, 5, 6,
1.

0.010 140 31 0.7 230 31 3.6


6,800 -->
C.

008 1.35
1.

566 880 1.33


7.

290 0.827
1.

0.015 24 0.415 250 1.09


6,380
4.

0.020 1.22 0.90 720 1.08 81


1,
7,

0.025 240 1.28 0.665 150 12 2.56


8,960 6,880 0.7
1. 1. 1.

0.031 35 0.373 1.23


8, 5,
7.

0.014 290 24 0.415 260 1.09 0.827


0.008 11,230 53 0.455 920 1.42 0.7
12.980 10,770 0.468
1.

0.008 1.6 0.316 52


IRRADIATION EFFECTS IN URANIUM DIOXIDE 623

erable interest has been aroused in recent years and much experimen
tation has been performed on fuel elements so fabricated. In the
present section, the factors affecting thermal performance of such
fuel elements are reviewed.

(a) Thermal Conductivity of UO, Powder

and Deissler measured the thermal conductivity of UO, pow


Eian
ders in air, helium, and argon over the temperature range 135° to
1,455° F and the gas pressure range 89 to 164 psia. [159]. They
found that above a pressure at which the mean free path exceeded
the characteristic dimension of the gas space in the gas-oxide mixture,
the effective conductivity of the mixture became independent of gas
pressure. This critical or breakaway pressure was given by the
relation

P,-1.770X10-24
T Eq.
Sºl, (9.40)

where P, is the breakaway pressure in psi, T is the gas temperature


in º R, S is the molecular diameter of the gas in feet (He, 6.23 × 10−";
air, 9.9 × 10−"; A, 9.45×10−"), and 7, is the mean sieve opening in
feet to retain the particles. Deissler and Eian derived relations be
tween the effective thermal conductivity of powder mixtures contain
ing various volume fractions of oxide and the conductivities of solid
and gas by interpolating between analytically derived expressions for
these relationships for various assumed geometrical arrays of solid
and gas [160] . By use of the chart shown in Fig. 9.76, these rela
tionships may be established. Boegli and Deissler obtained the meas
urements shown in Fig. 9.77 of the conductivity of 59.5 percent
dense UO,powder compacts in various gaseous media [161]. The
marked effect of gas conductivity in powder-gas mixtures on the con
ductivity of the composite is well illustrated by this work. Berg,
Flinta, and Seltorp and Flinta utilized the following expression for
determining the conductivity of powder-gas mixture:

* == 1
+!.
K (1—r)!/? V2K. K, X Eq. (9.41)
(I-ryū ºſi-ryūtºcontact
in which K is the conductivity of the gas-solid
r is the gasmixture,
volume fraction and K, K.,
and Kontact are the conductivities of
gas, solid, and solid-solid contacts (0.004 w/cm-º C), respectively
[54, 162]. In Fig. 9.77, the applicability of Eq. 9.41 as well
as the graphical correlations of Diessler and Eian in Fig. 9.76 to
624 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS
I.O
I-I-I-HTTH I-I-I-I-III- T-T-T-TTTTI

O.8 - -
- –
O.6 – —

O.4 –
SOLID CONDUCTIVITY, K,
GRS CONDUCTIVF. K. I
O.2 H. —

|-| 3 IO 3O IOO 3OO IOOO -


o I L | | | N 1NI Dº! I L L Lºſ
I 2 3 4 6 8 Io 2O 3O 4O 6O 80 IOO 2OO 3OO 6oo looo
EFFECTIVE CONDUCTIVITY
- K
GAS CONDUCTIVITY Kg

FIGURE 9.76. Relation between Solid Conductivity (K.), Gas Conductivity (Kr),
Effective Thermal Conductivity (K.), and Volume Fraction Void Space (r)
in Gas-Powder Mixtures [160] .

O.O| 7 I
I I I I T T I

O.O| 6 H

O.O||5 H. -
A Q AA A |OO"), HELIUM
O.O||4 H. —
59.3 TO 136.9 psi
O.O 3 H. FLINTA (159) -
loo". He (FIG. 9.74)
O.O||2 | -
V 65.0%. He-35.0% A
ool H
V v - 47.3 to 84.3 psi
O.O.I.OH. -

2- -
A A_48.8%.
O.OO9 H
A
ZN
He-51.2% A
39.4% TO 96.3 psi -

i O.OO8 H

><A T--A-A-A-Lº-Nilº
25%, He-75.0%. A
46.3 TO 84.3 psi
-

-
O.OO7 H. A A-IQQ24 NIIROGEN T
49.3 TO 83.3 psi
O.OO6 H -
O.OO5 H
O ° loo". ARGON -
O O44.3 TO
OO4. H
O.OO4. - O 3. e*sºnon-nº.
ſe-e-o-º-º-º-2-º-º-
_-e-w-HTTE;
O 94.4 psi
-
O.OO3 H. O 18.78 To 74.3 psi
KRYPron
-
O.OO2 H 83%, Xe, 17%. Kr (FIG.9.76) -
O.OO| H. -
O.OOO | l l l l l l I
O IOO 200 300 400 500 600 700 800
TEMPERATURE, "c

FIGURE 9.77. Ke, Effective Thermal Conductivity of UO, Powder in Various


Gases and Gas Mixtures [161].
IRRADIATION EFFECTS IN URANIUM DIOXIDE 625

the data of Boegli and Diessler are illustrated for a helium and a
Xe–Kr atmosphere. Equation 9.41 is seen to yield close agreement
with the measurements.

(b) Thermal Conductivity of Swaged UO,

This equation, while yielding good correlation


with experimental
measurements for powder mixtures, appears to be inapplicable to the
case of UO, powders consolidated by swaging. A measurement of
thermal conductivity at 200° F for a compact swaged to 84 percent
theoretical is reported by Kane, who found the conductivity
density
to be 0.064 w/cm–9 C [163]. The conductivity calculated by Eq.
9.41 (even assuming a helium atmosphere within the compact) is
0.018 w/cm-2 C, whereas that derived from correcting the data
of Kingery, et al., to a density of 84 percent is 0.083 w/cm-º C
[56]. This indicates that the thermal properties of swaged UO2
approach those of sintered UO, and are dissimilar to those of pow
der compacts. This expectation will be confirmed by the irradiation
behavior of fuel element samples discussed below.

(c) In-Pile Behavior of Powder UO, Fuel Elements

In Table 9.21 are listed the results of in-pile exposure at various


thermal ratings of oxide powder compacts ranging in density from
40 to 70 percent as reported by Eichenberg, et al., Robertson, et al.,
Bates and Roake, and Bain and Robertson [64, 95, 96, 107, 144, 164,
165]. The behavior of these compacts is highly sensitive to thermal
rating, and the following general types of behavior have been noted
as a function of in-pile thermal rating:

Up to
parameterſ.t Kd7–5.4
1. values of the w/cm, uniform

fuel densification and shrinkage occurs with no macroscopically


visible change in structure. The densification occurs by decrease
in pellet dimensions.
2. At values of
ſ. Kd7"=16 to 19 w/cm, central

and the fuel densifies in contact with the cladding by forming an


voids develop

annular cylinder of oxide. The oxide assumes the appearance


of coarse columnar grains radiating from the central void to the
cladding surface.

ſ
Te - ---
3. When
T.
Kd7" approaches 54 w/cm, in addition to the for
mation of central voids and columnar grains, the fuel relocates to
one end of the sheath to an overall density of over 60 percent.
When the initial powder density is over 60 percent (samples FS,
TABLE 9.21—BEHAVIOR OF POWDER FUEL ELEMENTS ON IRRADIATION

Initial fuel T. T.

T
Heat Percent conductivity Kd Kd T

|
°
C
|

T, T,

|
Sample Ref. Sample description flux Sample appearanceafter irradiation fission w/cm—" T, initial
o
C

No. w/cm2 gasrelease (He atmo- initial final T.


sphere) w/cm w/cm

8
8

in
0.
3.
3.

|
9.5 0146 86

21
15.0 Fuel sintered,decreased volume 345

64
60
163 percent dense MCW-UO: He

75
96 atmosphere—Zirc-2clad 1.05 cm percent: final density about per

X
Old 0.912cm ID cent 940MWD/Ton burnup

in
0.

in
9.7 0146

as
164 64 As #163-------------------------- 17.0 Fuelsintered #163,3,360
MWD/Ton 4.3 4.3 88 385
96 burnup

||
of

in
241 64 As #163-------------------------- 76.5 Formation dense,columnar grains 13.4 0.0146 19.4 12.5 337 1,665
96 with central void 0.353cm diameter;

71
final density about percent 4,900
MWD/Ton burnup
1

in
in
242 64 As #163-------------------------- 66.4 As #241;13,430MWD/Ton burnup- 13. 0.01.46 16.8 10.5 298 1,440
96 central void 0.419 inch diameter;

79
final density about percent

in

in
|
171 20.7 Fuel sintered and shrunk volume; 6.2 0.0163 5.4 5.4 113 445

64
70
As #163,exceptoxide percent
96 dense 1,150MWD/Ton

in
241 64 As in #171 85.3 As #241,central void 0.338 cm 12.0 0.0163 22.4 15.9 377 1,750
96 diameter, 4,760 MWD/Ton; final
density about 80
percent

in
252 64 As in #171 72.8 As #251,central void 0.399 cm 11.6 0.0163 19.1 12.4 333 1,505
96 diameter, 13,150MWD/Ton; final
85

density about percent


in

in
0.0163 18.4 310

||
71.0 As #251,central void 0.257 cm 14.3

71
BB, BC 107 percent dense powder Zirc-2 1,440
sheath 1,051 cm ODX0,915 diameter, 3,300 MWD/Ton; final
77

cm ID density about percent

-
o
o

||
0

||
in
19. 0.016.3 18.

o.
B.A. 95 72 percentdense powder Zire-2 74. Only grain growth reported to diam- 18. 310 1,410
sheath 1.051 cm ODX0.915 eter 0.49 cm, no central void;
cm ID

r.
T.
Kd T=13 w/cm

ſ
T.

in
4,

251–C 0.0113 54.2 78.5

40
164 percent dense UO2 powder 152.0 Central voids and columnar growth 155 955

w
1.43 cm ODX1.28 cm ID over entire length; fuel redistrib

of

to
Zirc-2 can

36
uted upper can, void about
0.480 cm diameter; final density

70
about percent
5

in

|
4,

251-A 0.01.15 54.2 78. 155

43
164 percentdenseUOz-PuC2 powder 152.0 As 251-C, exceptfuel sinueredwith ---------- 955

in
of
251-B in can as in #251—C fewer cracks than core UOz
powder
0

of

to

in
|
6,

Ia 165 36.5 percent dense UO packed 220.0 Fuel redistributed upper half can |---------- 0.011 ~70.0 140. 202 560

to
85
Zirc-2 can 2.54 cm (final) OD percent density; central, high

X
2.39cm ID, He atmosphere density core (Inolten) 0.835cm di
ameter; Surrounded by columnar

of
zone to diameter 2.01 cm with

of
outer annulus sintered UO2

Teer.
KdT=123 w/cm
T,
To...
Kd7–41 w/cm
T,

Ia
in

Ia
in
in
Ib 165 As above, except sealed 220.0 As above except atmosphere |----------|-------------- ~70.0 140.0 202 l----------

in
vacuum after puncturing rich Hz and fuel
redistributed to bottom half to

Ia
in
in

|
IIa 165 As above; powder contained 220.0 As exceptfuel relocated bottom ----------|-------------- ~70.0 140.0 202 ----------

Ia of
adsorbed CO2 half sheath
§
TABLE 9.21—BEHAVIOR OF POWDER FUEL ELEMENTS ON IRRADIATION.—Continued

Initial fuel T. T.

T
Kd RºdT

|
Heat Percent conductivity

C

C
flux fission w/cm-º T, T, Te initial
Ref. Sample description Sample appearanceafter irradiation
o
C

Sample initial final T,


N w/cm2 gasrelease (He atmo-
sphere) w/cm w/cm

202 l----------

!
Ia
165 As in IIa -------------------------- 220.0 As in ----------------------------------------|-------------- ~70.0 140.0
IIb

0.016 61.0 l------------ 139 3,955

to

in
FS 145.0 Columnar grains 1.24cm diameter; ----------

69
144 percent dense UO2 powder

1
0
stainless steel sheath 1.837 cm central void to cm diameter

1
O.D.x1.705 cm I.D. min T.s.
irradiation Kd T-39 wicm

-
J.
0.016 59.5 l------------ 135 3,855

to

in
140.0 Columnar grains 1.42 cm dia- ----------

68
Fu 144 As FS, exceptpowder percent

0
dense, min irradiation meter;central void to 0.89cm
T.e.
Kd T-27 w/cm
T.

0.016 58.5 l------------ 130 3,790

in
Fx 144 As FS, except powder67 percent 138.0 Columnar grains 1.34cm diameter: |----------

0
to to
dense,40 min irradiation central void 0.58cm diameter
T.,
Kd T=27 w/cm

ſ -
IRRADIATION EFFECTS IN URANIUM DIOXIDE 629

FA, and FX), the gross relocation does not seem to occur, although
it is not clear that such redistribution would not be noted with
sufficiently long irradiation times.

4. At values of
ſ. Kd7 equal to about 70 w/cm, the redistribu

tion of powder becomes even more extensive, overall densities of


85 percent being attained from initial powder packings of 36 per
cent densities. The central voids noted at lower heat ratings are
replaced with cores filled with high density oxide, and columnar
grains are again observed to radiate from the central core.
Considering first the results obtained at very low thermal ratings,
it is indeed surprising that the densification observed could occur
at such low estimated temperatures (samples 163, 164, 171). A
possible explanation is offered by Bates and Roake, who observed, in
postirradiation analysis of the gases contained in samples Ia, Ib, IIa,
and IIb, large amounts of N., and CO, which were adsorbed on the
powder during room temperature storage [165]. It can be calcu
lated that sufficient nitrogen is present, if adsorbed as a monomolec
ular film on the powder of specimens 163 to 252, to dilute the helium
atmosphere with two volumes of nitrogen for each volume of helium.
Thus, the initial conductivity of the powder mixtures can be esti
mated to be about 0.0056 w/cm – 9 C rather than the values listed
in Table 9.21, and the oxide central temperatures of the specimen in
question would be raised 400° to 500° C above those listed in the
table. At such temperatures, densification to the extent noted could
reasonably be expected to occur during the 3-month to 1-year in-pile
exposure of the specimens. Data are unavailable to determine if the
shrinkage effects observed are explicable on the basis of temperature
alone. or if some irradiation effect on sinterability must be invoked.
The formation of central voids and columnar grains at heat outputs
of 240 w/cm clearly is of different origin than that noted in the case
of sintered oxide (Fig. 9.67). In the latter instance, the onset of co
lumnar grain growth was observed to occur at locations in the fuel
within several hundred degrees of the melting point. In most of the
cases listed in Table 9.21, the columnar grains extended either to the
cladding or occupied most of the fuel cross section. This is best seen
in Fig. 9.78, in which the columnar grains of specimen 251A of Ta
ble 9.21 are seen to extend to the pellet surface. Bates and Roake sug
gest that the columnar grain growth arises from vaporization and dep
osition of UO, during operation [165]. The gross relocation of oxide
to the ends of the fuel element cans noted in Table 9.21 supports this
suggestion. Furthermore, the distribution of voids within the colum
nar grains, transverse to the grain axes, indicates the formation of
columnar grains by the growth of vapor-deposited dendrites of oxide.
630 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FIGURE 9.78. Fractured and Polished Surfaces of Irradiated Plutonium


14,

Enriched UO, ; Rod 251A, X3 and X10, Reduction Factor, [164]. (Courtesy,
American Institute Mining, Metallurgical, and Petroleum Engineers, Inc.)
of

the columnar grain structures


of

The
in

such structures
in

absence
presumed originate from solidification
to

sintered oxide which were


from void migration has previously discussed.
of

or

the melt been


specimen 251-A can
be
of

The formation
in

in

transverse voids seen


Fig. their probable origin from intersection secondary
of

9.79, and
dendrite arms may inferred from the dendritic deposits observed
be

radial columnar grains Fig.


as
of

in
at

the outer terminus shown


Again, the dilution the atmosphere within the powder-filled
of

9.80.
by

the powder during storage adequate


of on

fuel rods gas adsorbed


is

explain the development temperatures high enough cause UO,


to

to

speci
of
In

this connection, the results


of

volatilization. irradiation
are instructive; although this sample was exposed
at

men BA heat
flux ratings much higher than those which caused void formation and
columnar grain growth similar samples, only equiaxed grain growth
in
IRRADIATION EFFECTS IN URANIUM DIOXIDE 631

-
* † -- -
T
- - : -
º

- - - - - -
-

FIGURE 9.79. Void Formation Transverse to the Axis of Columnar Grains, Re


#4,

duction Factor, (Courtesy, American Mining, Metal


of

[164]. Institute
lurgical, and Petroleum Engineers, Inc.)

reported for this sample. inferred that this case the sur
in
It
is

was

the powder was adequately purged gaseous impurities prior


of
of

face
op
It,

fabrication. thus, appears essential order


to

fuel element
to in
to

powder UO, fuel high-power ratings observe suit


at

erate elements
precautions during fabrication adsorption gases
of
to

able minimize
conductivity
on

poor thermal the oxide surfaces.


of

UO, Fuel Element


(d)

Swaged
of

In-Pile Behavior
on

Relatively few results, listed Table 9.22, have been obtained


in

the
swaged UO, fuel elements. The results avail
of

irradiation behavior
able indicate, however, that structural changes are much less marked

the powder samples listed


of

Table 9.22 operated


at

the case
in
in

than
632 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

FIGURE 9.80. Dendrites of UO, at Outer Terminus of Radial Columnar Grains,


Reduction Factor, 14 [164]. (Courtesy, American Institute of Mining, Metal
lurgical, and Petroleum Engineers, Inc.)

similar heat fluxes. Whether this improved thermal performance


arises from better packing of the powders alone, or whether it is caused
by a specific effect of the mechanical working on the nature of interpar
ticle contacts, is not known. To settle this point, the behavior of pow
der fuel elements tamp-packed to densities comparable to those listed
in Table 9.22 would be of considerable interest.
Although structural changes are less prevalent in swaged than
in powder oxide samples, grain growth may be noted to occur at

values of
ſ”T.
KdT appreciably lower than those listed in Table

9.19 for oxide samples, about 25 w/cm for the former and
sintered
30 to 40 w/cm (depending on fuel-clad clearance) in the latter.
These structural changes occur at lower thermal ratings, even though
the fuel-clad clearance in swaged samples is obviously negligible.
This observation checks the measurements of Kane, discussed above,
who found the thermal conductivity of swaged samples to be 25
percent less than that of sintered UO, of comparable density [163].
Either the finer grain size of the swaged samples, or the thermal
resistance offered by interparticle contacts, may explain the reduced
thermal conductivity of this type of material as compared with that
of sintered UO2.
IRRADIATION EFFECTS IN URANIUM DIOXIDE 633

TABLE 9.22—BEHAVIOR OF SWAGED FUEL ELEMENTS ON


IRRADIATION

Refer-
Sheath-
water T.
, kar
T.

tº-temper.|
-
Spec.
no. ence Sample description Appearance after RºdT 3:
oC
heat flux
w/cm2 irradiation J.(w/cm)
w
"ature of
structure
change
(w/cm)

BG,BH, 107| Swaged to 89.5 per- 105|| Grain growth to 27,1 25.3 340
BF cent density in diameter 0.21
Zirc—2sheath 1.077 cm in specimen
cm O.D. x 0.915 BG
cm I.D.

BF------- 95 | Swaged to 89.5 per- 107|| Grain growth to 30 20.0 340


cent density in diameter 0.52
Zirc-2 sheath 1.077 cm, operated
cm O.D. x 0.915 defected in hot
cm I.D. water loop

EH------- 144 Swaged to 86 percent 111|| Sintered to dia- 48 21.4 124


density in stainless- meter 1.3cm (to sintering)
steel sheath 1.857
cm O.D. x 1.7 cm
I.D.; 1minute irra
diation.

El-------- 144| Swaged to 86 percent 142| Sintered to dia- 66 22.7 139


density in stainless meter 1.37 cm; (to sintering)
steel sheath 1.857 grain growth to
cm O.D. x 1.715 diameter 0.635
cm I.D.; 1 minute cm
irradiation

9.8 APPLICATION OF URANIUM DIOXIDE IN FUEL


ELEMENTS

In preceding sections, various considerations relating to the be


havior of UO, under irradiation have been discussed, and an attempt
has been made to relate irradiation-induced effects to experimentally
measurable quantities in order to permit prediction of the ranges of
applicability of UO, fuel. In this section, experience with the utili
zation of UO, fuel in fuel elements will be discussed and effects
arising from the interaction of cladding, fuel coolant, and irradiation
conditions described. Uranium dioxide has been applied as fuel
material in three types of fuel elements:

1. Bulk oxide fuel elements in which adequate performance requires


reproducible and controllable behavior of the fuel.
2. Pin oxide fuel elements in which a cladding of sufficient strength
is employed to contain the fuel in the event of severe dimensional
and property changes.
634 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS l
3. Dispersion fuel elements in which the fuel is distributed as dis
crete particles,generally in a metallic matrix, and in which the
matrix properties are utilized to minimize radiation damage effects.
As
may be inferred from these definitions, the functional differ
ence between these fuel element types is primarily in the amount of
structural material utilized to contain the oxide.

9.8.1 Bulk Oxide Fuel Elements

(a) Effects Related to Oaside

In previous sections, there has been reviewed in detail a number


of the unique propertiesof uranium dioxide which affect its use as a
fuel. Two additionally important characteristics which strongly affect
the manner of its application in fuel elements are its brittleness
and consequent tendency to fracture under stresses imposed during
operation or fabrication and its chemical properties which prevent
metallurgical bonding to metallic sheathing. The former property
requires that fuel elements employing oxide fuel utilize cladding
which is either self-supporting under operation or, at least, imposes
only compressional stresses on the fuel. The latter characteristic
leads to susceptibility of oxide fuel elements to waterlogging effects,
increases coolant contamination over that from a corrosion-resistant
bonded fuel in case of cladding defects, and requires provision in fuel
element design to accommodate differential fuel-clad expansions.
(1) OxIDE CRACKING. Eichenberg, et al., have shown for the case
of a cylindrical pellet exposed at a uniform volumetric heat genera
tion rate, q, that the maximum thermal stresses developed at the pellet
surface in the axial (gz) and tangential (or) directions are given by
[64]:
_/ E \oga" Eo. _ Eo. Te
ofT =
•-(+. # =3diº (T-T)=5kº ſ. Kat
Eq. (9.42)

From the modulus of rupture data in Ref. 74, or=13,000 psi, modul
lus of elasticity E=25×10" psi, Poisson's ratio u-0.3, and expan
sivity a -10°/* C, it can be calculated that surface cracking of the
pellets should occur at a temperature difference between center and
surface of 73° C or a value of ſ
Kd7 of about 2.5w/cm. Similar
conclusions were reached by Ellington [167]. Extensive cracking
through the pellet will require, of course, larger temperature differ
of

high
is,

It however, apparent that oxide fuel elements


be in

ences.
power ratings, extensive cracking will observed. This cracking was
being responsible for the improvement conductivity noted
as

cited
in

central temperatures the WAPD–22–7


of

in
at

about 600° and


C
IRRADIATION EFFECTS IN URANIUM DIOXIDE 635

-15 experiments of Fig. 9.70B in which values of T.-T, of 130°


to 160° C are observed at these temperatures, in qualitative agreement
with the estimates from Eq. 9.42. As anticipated from the direc
tion of the principal stresses, cracking occurs predominantly radially
and perpendicular to the pellet axis, although in fuel element samples
operated at high thermal ratings circumferential cracks (usually
bordering regions showing grain growth) are often observed.
From Eq. 9.42, it is apparent that cracking can be restrained
by improving rupture strength or by increasing thermal conductivity.
An apparent improvement in cracking tendencies attributable to one
or both of these effects was reported by Neimark and Kittel [168].
Pellets of Tho,-UO, showed extensive cracking when exposed at a
value of
ſ. Kd7 of 43 w/cm, while similar pellets containing
weight percent of molybdenum fibers, which improve conductivity as
10

well as modulus of rupture, were reported to show no cracks at an ex


posure of
ſ: Kar-w w/cm.
While some improvement in cracking tendencies can anticipated
be
by use of measures such it is
as those described above, unlikely that
such improvements will be sufficient to permit designs requiring
mechanical integrity of the pellets. However, the cracking which is
observed occurs on a gross, macroscopic scale and does not appear to
be progressive with fuel burnup or fuel element thermal cycling. Ap
parently, after the initial cycles of thermal stressing of the pellets they
become reduced to a fragment size which is stable to the thermal
gradient stresses.
(2) RATCHETTING. As a result of cracking of the oxide, it may be
conceived that fragments of oxide could lodge between fuel and clad
or between pieces of oxide, cause plastic strain of the cladding during
power operation, wedge more tightly during shutdown, and, by a series
of such events, cause sufficient clad deformation to lead to overstrain
and rupture. This phenomenon has been referred to as ratchetting.
That relocation of fuel fragments probably occurs during operation
has been demonstrated by the results shown in Fig. 9.71 in which
changes of fuel conductivity were noted on successive startups. A
number of tests are reported in Ref. 64 in which, in addition to
the normal number of reactor startups, fuel element rod samples were
cycled in and out of reactor flux for as ma".y as 9,500 cycles. No di
mensional changes were noted as a result of such cycling, nor was the
structure of the fuel different from that of samples subjected to fewer
is,

thermal cycles. It thus, concluded that spite the fuel cracking,


of
in

ratchetting not probable failure mechanism sintered oxide fuel


in
is

The macroscopic the fragments


of

elements. size formed and the


636 URANIUM DIOxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

resistance of the oxide to progressive fragmentation provide adequate


safeguards against ratchetting failures. Such cycling experiments
were performed with fuel elements of 9.5-inch active fuel length. It
would be expected that the susceptibility to damage by this effect
would increase with fuel element length. Data are unavailable to
assess whether such effects, indeed, are aggravated with increasing
length.
(3) AssBMBLY CLEARANCE. While a metallurgical bond or even a
tight mechanical bond between cladding and fuel would undoubtedly
improve thermal behavior and would minimize effects arising from
fuel cracking, no suitable means of securing such bonds to bulk UO.
fuel has been conceived. The utilization of a helium heat transfer
bond between fuel and clad has been shown in Sect. 9.7 to be of
questionable value, and certainly far less effective than minimization
of fuel-clad clearance. Assembly clearance may, of course, be reduced
by utilizing powder UO, as fuel and swaging or otherwise reducing
the cladding onto the oxide. In case of sintered fuel, the cladding
may be brought in contact with the fuel by a postassembly sinking
operation; in the fabrication of plate oxide fuel elements intimate
contact of fuel with clad is obtained in the course of the fabrication
process [169]. Another means of minimizing clearance is described
by Kittel, Handwerk, and Neimark who tested samples in which the
fuel-clad annuli were filled with lead or, in a few experiments, sodium
potassium alloy [168, 170, 171]. While it would be anticipated that
such a bond would provide very high gap conductance, photographs
of the cross section of a lead-bonded Tho, UO, fueled specimen in
Ref. 168 indicate formation of central voids and columnar grains
at thermal ratings comparable to those at which the same effects are
noted in the unbonded samples of Table 9.19. It was further shown
that the lead bond prevented relocation of fuel fragments, whereas
similar unbonded samples formed cladding cracks attributed to the
extrusion of a fragment of fuel through the cladding.
A factor to be assessed in the evaluation of fuel elements with a tight
mechanical bond to the cladding is the transmission of differential
thermal expansion strains from the fuel to the clad. Such transmis
sion of strain may lead to strain cycling of the cladding as a result of
reactor startups and shutdowns and to its premature failure ifthe
magnitude of the strain is large.

(b) Effects Related to Cladding

(1) CLAD DEFORMATION. The proper performance of bulk oxide


fuel elements is intimately related to the characteristics and nature of
the cladding. For ease of assembly and to permit use of generous
clad and fuel dimensional tolerances, it is desirable to use large diame
IRRADIATION EFFECTS IN URANIUM DIOXIDE 637

tral Such clearances are permissible at considerable sacri


clearances.
fice to thermal performance if the cladding thicknesses and strength
properties are such as to resist external pressure. If, however, the
reactor design requires minimization of structural materials and,
thus, use of claddings which are not self-supporting, undesirable de
formation of the cladding may occur. A detailed study of the de
formations encountered in such claddings was reported by Robertson,
et al., and Runnals [95, 135, 172]. When the diametral clearance
between the fuel and Zircaloy–2 cladding, 1.0 inch OD by 0.025 inch
thick, exceeded 0.006 inch, plastic buckling of the cladding upon the
fuel occurred at temperatures of 300° to 400° C and pressures of 70
to 100 atmospheres, resulting in the formation of a single longitudinal
ridge, the ridge height increasing with increasing assembly clearance
beyond 0.006 inch. Thus, the maximum diametral assembly clearance
is limited by plastic buckling instability of tubing under external
pressures. Another deformation effect was reported by Robertson,
associated with rod-type fuel elements operated at high thermal rat
ings with small assembly clearance, consisting of circumferential
ridges about 0.002 inch high which appeared at the interface between
the stacked pellets [95]. The origin of these deformations is as yet
unknown. Concern was expressed that they may disappear as a result
of axial expansion of the pellets during operation and reform on
reactor shutdown, thus, leading to strain cycling of the cladding.
A potentially more severe problem is that of accommodating the
axial expansion of the fuel. This problem is aggravated the greater
the length of the fuel; thus, a stack of pellets 8 feet long at a tempera
ture 1,000° C above that of the cladding will expand more than 1
inch relative to the cladding. In self-supporting cladding, i.e., of
dimensions capable of resisting external pressure, it is feasible to
leave axial expansion space above the fuel to accommodate this move
ment. However, it seems difficult to guarantee, with such large rela
tive movements, that wedging of fuel fragments will not occur, leading
to clad fracture by an axial ratchetting mechanism. To counteract
this possibility, shorter fuel elements may be employed, as in the
Shippingport PWR blanket fuel elements, or spacers may be em
ployed to distribute the expansion along the length of the fuel ele
ment. In case of nonself-supporting claddings, large amounts of
axial void volume cannot be provided. Thus, in case of the deformable
cladding previously described, deformation into axial voids greater
than 0.15 inch long occurs under external pressure, the deformation
assuming a characteristic four-leaf clover pattern. Fuel element fail
ure due to collapse of axial void space under external pressure was
reported in aluminum-clad tube-in-plate fuel elements utilized in the
Borax IV reactor [168]. The Canadian investigators have proposed
638 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

an ingenious solution to the problem of providing axial expansion


space, consisting of dishing the ends of each pellet an amount adequate
to accommodate the thermal expansion at each radial location [95].
The results of the irradiation testing of such pellets apparently indi
cate this expedient is successful in providing axial expansion space.
(2) WATERLOGGING. An important consequence of the utilization
of unbonded fuel elements in water-cooled reactors is their suscepti
bility to waterlogging failures. Such failure may occur in elements
with defective cladding and arises from entry of the coolant into the
sheath interior during reactor shutdown periods through the defects
and the flashing of the coolant to steam when the reactor is brought
to power. If the defects are sufficiently small, the excessive pressures
developed within the sheath are relieved by plastic deformation or
rupture of the sheath. Such waterlogging failures were discussed
by Eichenberg, et al., and by Mooradian [64, 173].
Two types of waterlogging failures may be distinguished in oxide
fuel elements. The one type occurs in the presence of considerable
open porosity in the fuel such that water can permeate the pores of
the oxide during low power exposure. On approaching operating
temperatures, the vaporization of water in the pores may fragment
the oxide and project fragments with considerable force against the
interior of the cladding. Examples of failures resulting from such
effects are the X-1–b and X-1—c tests discussed by Eichenberg, et al.:
in the former test, in which low density hot-pressed oxide was em
ployed, a gross cladding rupture occurred, whereas in the latter, in
which the fuel was of high density but was rendered permeable by
the presence of corrodible carbide impurities, the cladding bulged
plastically [64]. Such failures can apparently be avoided by the
use of high density sintered fuel with very small amounts of inter
connected porosity. The second type of waterlogging occurs as a
result of permeation of clearance and void space with water. Be
cause of the lower temperatures and, hence, lower pressures developed
by the steam formed in such locations as compared with the first
type, rod-type fuel elements show less susceptibility to failure by
this mechanism than by the first type. As evidence of this, many
tests of rod fuel elements utilizing high density oxide are reported
in Refs. 64, 95, and 107 in which successful long-time operation was
noted even in cases in which the clad contained defects consisting
of holes 0.005inch in diameter. However, in the case of plate-type
oxide fuel elements, severe failures resulting in clad bulging and
fracture have been noted both in out-of-pile and in-pile tests with
defects of the same size [169, 174]. The bulging caused by water
logging of a plate oxide fuel element is shown in Fig. 9.81 in which
the behavior of a defected, 14-inch-wide compartment is compared
-
*

FIGURE 9.81.
-
IRRADIATION EFFECTS IN URANIUM DIOXIDE

Cross Section of Plate Fuel Element Showing Extent


-
639

of Bulging
Caused by Waterlogging [124].

with an adjacent nondefected compartment. Such bulging may


be compared with that experienced by fuel elements in which
clad deformation occurs by fuel swelling (Fig. 9.53) and in which
the fuel is shown to be in contact with the cladding. Waterlogging
failures of this type are sensitively interrelated functions of defect
size, strength of sheath, and internal void volume. Thus, whereas
0.005-inch defects in cylindrical sheaths of dimensions 0.413 inch
ODX 0.027 inch thick do not result in cladding waterlogging failures,
0.040-inch defects are required to prevent waterlogging in flat plate
fuel elements with 0.020-inch claddings and internal bonded ribs or
cladding supports spaced 14 inch apart across the width of the plate.
Furthermore, whereas with internal void volumes of 5 percent in
flat plate fuel elements waterlogging failures are quite prevalent with
0.005-inch holes, fuel elements of the same type but with very small
amounts of void volume have operated with defects of such size for
over six months with no evidence of waterlogging [174]. The risks
associated with waterlogging can be minimized by reducing the
amount of fuel exposed by a defect (i.e., by compartmentation of the
fuel) and by providing internal supports to minimize cladding de
flection prior to rupture and to prevent blocking of coolant channels
by excessive sheath deformation. Thus, the design described in Ref.
169 in which the fuel is contained in separate, sealed compart
ments 14 inch wide by 11% to 6 inches long is considered to be re
quired in plate-type fuel elements to minimize risks attending water
logging failures. A similar design is also considered necessary in
fuel elements such as internally-cooled annular elements in which in
ternal pressures developed by waterlogging can cause collapse of the
internal sheath. Waterlogging failures are peculiar to fuel elements
operated in water or liquid media in which the internal temperatures
approach or exceed the critical point of the coolant and would not
be anticipated in gas or liquid metal cooled reactors.
640 URANIUM DIoxIDE: PROPERTIES AND NUCLEAR APPLICATIONS

(3) HYDRIDING of CLADDING. Another cladding effect, peculiar


to the use of zirconium-alloy cladding in water-cooled reactors, is the
phenomenon of cladding hydriding, described by Eichenberg, et

al.
has been found that defected fuel rods, operated water

in
It
[64].

40
greater, form addi

or
cooled loops

of

of
Kd7" w/cm

at
values

ſ
tional cladding failures discrete, isolated points the sheathing

in
at

of in
several days after reactor startup. Examination

as

as
times short
the failures indicated that high concentration precipitated ZrHis

of
a
preferentially

of
segregated

at
was the outer (cooler) surface the
cladding and that cracks initiated the brittle, high hydrogen con

in
tent zone under the thermal strains induced by the clad temperature
gradients and propagated through the sheath. From the observation
that such high hydrogen content regions were not observed

in
fuel
rods with undefected sheaths operated the same test strings,

in
was

to of it
concluded that hydrogen was absorbed the inner surface

at
the
sheaths and segregated under the clad temperature gradients the
the sheaths. Such thermal gradient diffusion

of

of
outer surface
hydrogen was reproduced out-of-pile by Markowitz who found the
hydrogen temperature T1, Ni,
of

of
ratio mole fraction zirconium

in

at
(T2-T,), by
be
to

the relation
to

at
related that T's, W.

*exp|# (#-4)]

Eq.
(848)

which Q”, the heat transport, was found

to
of

have the value 4.2


in

kcal/mole [175]. The factors affecting hydriding failures are yet

as
incompletely understood.

of
Failure times are sensitive function

a
power rating, ranging from

of

in
few days the case specimens
in the
a

X-1-n test (J. Kd7'-46 w/cm)


no

14

of
failure months
to

over

in
operation (ſ
Kd7–25 w/cm) [108]. Fuel-clad
of

the X-1–1 test

no
clearance also appears affect the failure times, since failures were
to

noted the X-1—f test, with 0.008-inch diametral clearance on assem


in

bly, whereas specimens with smaller clearance the X-1—h and X-1-
in

ºr
tests failed similar exposure time periods [64]. Finally, physic
in

'
the sheath interior appear affect hydride failu
of

to

characteristics
characteristics, since plate elements have been operated high surface
at

heat fluxes with defected zirconium alloy claddings and


no

indication
hydriding attack [174]. has been postulated that hydrogen
of

It

the cladding occurs


as

absorption
of

the interior surface result


at

the buildup hydrogen-rich atmospheres within the sheath,


of

of

as

discussed Sect. 9.4.10(b), and the subsequent gettering of the


in

hydrogen through defects the internal oxide film. Measures


of by in

to

mitigate hydriding failures compositional changes the Zircaloy


of

cladding can limit the time exposure defective fuel elements in


of
IRRADIATION EFFECTS IN URANIUM DIOXIDE 641

pile in water-cooled reactor systems. Discussion of such measures is


beyond the scope of this book. Fuel elements utilizing claddings such
as stainless-steel which are not subject to attack by hydrogen are not
limited by such failures.
(4) FUEL ELEMENT AssFMBLY. The utilization of unbonded oxide
fuel with poor thermal conductivity poses a number of problems with
respect to assembly of fuel elements. If, through bowing of the fuel
element during assembly or operation, one side of the element is in
sufficiently cooled, inability to remove the heat developed through the
opposite side of the element because of the poor conductivity of the
oxide may cause an increase of cladding temperature at the inade
quately cooled section which can aggravate the initial amount of bow
ing. This effect has been tested in the Shippingport PWR type of
fuel element assemblies by deliberately bowing a fuel element prior
to irradiation, so that it contacted an adjacent fuel element [64]. Op
eration of this assembly revealed no adverse effects resulting from such
distortions. Thus, in water-cooled systems, it appears improbable that
unequal temperatures can be developed of sufficient magnitude to
create a serious problem. However, in case of less efficient heat trans
fer media, such as gases, such effects can be much more marked, and
adequate means of fuel element support must be provided to minimize
fuel element distortions. A number of methods of fuel element sup
port have been proposed involving welding of fuel elements to tube.
sheets, utilization of wire wraps or other intermediate supports to
space the elements, or metal grid supports spaced at suitable intervals
[135, 169, 176].
(5) CoMPOSITION OF CLADDING. A number of cladding materials
have been utilized or proposed for utilization with UO, fuel elements,
and tests have been reported which utilized zirconium alloy, stainless
steel, and aluminum alloy claddings. No pronounced effects, other
than those previously discussed, on fuel element performance have
been reported as a result of utilization of these various cladding ma
terials. A material of high thermal conductivity, such as an aluminum
alloy, would be expected to improve the thermal output capabilities
of the oxide, since at similar coolant and central oxide temper
atures the surface temperature of the oxide should decrease, thus, per
mitting use of higher values of ſ Kd7, as shown in Fig. 9.62. This
effect may be counteracted by the apparent poorer gap thermal con
ductance obtained in the case of use of aluminum alloy claddings, some
evidence for which may be inferred from several of the values plotted
in Fig. 9.70A. This effect, if real, may arise from the lower contact
pressures obtainable with aluminum alloy sheaths than with stronger
zirconium or stainless-steel sheaths and may also serve to explain
why use of a lead-filled annulus does not appear markedly to improve
642 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

thermal behavior of oxide fuel elements as previously discussed.


However, such speculations aside, the application of oxide fuel appears
to be compatible with various cladding materials, and selection of a
proper cladding material rests on criteria other than fuel irradiation
performance.

(c) Effects of Fuel Element Shape

The first application of oxide fuel in power reactors utilized metal


sheathed cylindrical rod fuel elements, and the great majority of the
irradiation tests of oxide fuel have been performed with fuel elements
of this shape. The performance of 100,000 such fuel elements in the
Shippingport PWR station which, in the most highly rated core loca
tions, have accumulated almost (as of July 1, 1961) 20,000 MWD/
Ton—UO, burnup with no apparent injurious effect illustrates the
reliability of this fuel element concept. Most of the experience
with this fuel element has been accumulated with the fuel utilized
as high density sintered pellets; with the limitations discussed
in previous sections, elements of high performance capabilities
can be designed and fabricated. For reasons discussed in Sect.
9.7.3, considerable interest has been evidenced in the utilization
of UO, as a powder swaged to high density within metallic
sheaths. In addition to the tests of swaged oxide described in
Sect. 9.7.3, a large number of irradiation tests are in progress
sponsored by the Hanford Laboratories for application in the Plu
tonium Recycle Test Reactor [177]. Swaged fuel elements 0.56 inch
in diameter have been examined after an exposure of 3,000 MWD/Ton
at a maximum value of ſKd7–78 w/cm and exhibited the antici
pated formation of a central void and columnar grain growth at a

ſ:
T.... - - -
value of Kd7–32 w/cm, similar to that observed in Table 9.19

for sintered oxide fuel. It was further reported that swaged arc
fused powder UO, releases 10 percent of the fission gases, whereas
swaged sintered powder UO, releases 30 percent of the fission gases
upon irradiation. Whether this effect arises from differences in the
internal temperatures developed and, hence, of the thermal conduc
tivity of the two products or whether it is characteristic of the dif
ferences in the particle sizes of the two powders was not reported.
Thus, it may be anticipated that, with the limitation of somewhat
higher fission product release rates both within the cladding and
through defects into the coolant, and of a somewhat reduced thermal
capability, this type of fuel element will also find extensive power
reactor application. Utilization of such fuel elements requires,
as pointed out by Runnals, development of inspection and fabrication
methods which will guarantee the absence of cladding cracks formed
during the swaging operation [135].
IRRADIATION EFFECTS IN URANIUM DIOXIDE 643

The rod-type fuel element is particularly adaptable to utilization


with loose or tamp-packed powder fuel. This application is suitable
for plutonium-enriched oxide fuels in which it is desirable to minimize
handling of the fuel. The thermal behavior of powder-packed fuel
elements has been discussed in Sects. 9.7.3. (c) in which it was shown
that the behavior is highly sensitive to filling gas composition. In ad
dition to the data presented in Sect. 9.7.3(c) Stosek and Weidenbaum
reported the results of examination of stainless rods 0.250 inch
ODX 0.028 inch thick containing UO, and UO, +0.5 weight percent
TiO2 powder vibratory-packed to 53 to 65 percent density and exposed
to a peak burnup of about 20,000 MWD/Ton at a peak value of
T
Kd7–24 w/cm and average value about 14 w/cm [178]. Fis
T.
sion gas releases of 20 to 57 percent were noted, and the fuel exhibited
the typical void formation and columnar growth described in Sect.
9.7.3. Local segregations of oxide ascribed to gross relocation of the
fuel were also noted, particularly in samples containing the pure
UO2 powders. These results, obtained on 281/2-inch long samples, in
dicate the possibility of use of the powder-filled metal-sheathed rod
fuel element to high burnup exposures. Hanford experimenters con
firmed the importance of filling gas atmosphere in powder-filled rods
by irradiating 0.570-inch OD fuel rods tamp-packed to 77.5 percent
theoretical density and sealed in helium as well as in vacuum [177].
The former samples released 2.3 percent of the fission gas generated
and the latter 14.5 percent. In addition, the UO, in the helium-filled
rods remained unsintered, while central voids and grain growth across
three-fourths of the fuel diameter were noted in the vacuum-welded
samples.
A variation of the rod fuel element is that employing an annular
fuel pellet rather than a solid cylindrical pellet. It can be shown
from the equations of Sect. 9.7.1 that, for equal central tempera
tures, the power output per unit length of rod (or heat flux for equal
fuel diameters) for an annularly fueled rod, Q', exceeds that for a
solid fueled rod, Q, by the ratio

*——
QT, Eq.
2r"

(9.44)
ºr
1

1-H+ in;
pellet diameter. Thus,
of

which the ratio annulus


to

central
in

is

a
r

the pellet increases the power output


of

hole 0.9 times the diameter


by

percent),
of

of

factor (with
in
10

81

decrease fuel volume


a

while hole 0.2 times the outer diameter allows percent more power
15
a

output
at

percent fuel volume. Opposed


of

of

to

sacrifice these
4
a

57.4789 O–61–42
644 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

favorable effects, a number of adverse factors may be cited. It ap


pears difficult to guarantee that fuel fragments will not fall into
the central cavity as a result of fuel breakup under thermal stresses,
thus, exposing fuel to higher temperatures than those selected for
proper operation. On the other hand, exposure of annularly fueled
rod elements has indicated that fuel, even after cracking, is ade
quately keyed to neighboring fragments to prevent such relocation,
presumably because of the predominantly radial nature of the crack
ing [95]. Use of annular fuel allows a greater void volume to
accommodate fission gases released from the fuel and, thus, prevents
buildup of excessive pressures. On the other hand, the presence of
void volume in fuel elements used in water reactors may increase
the tendency for waterlogging in case of defected fuel elements.
A more speculative disadvantage is that the annular fuel may be
unable to exert cladding pressures suitable for attainment of low
gap thermal conductances. Many of these disadvantages may be
avoided if an insert of an inactive oxide is used in the central annulus.
Considerable test experience has been accumulated with the use
of UO, in plate form [174]. As outlined above, the plate fuel ele
ment applied in water-cooled reactors (or the related internally and
externally cooled tubular fuel element) requires internal (or possibly
external) rib supports to minimize waterlogging damage. Plate fuel
elements are likely to find application in systems in which power
density requirements are so high that rod fuel element diameters
become unattractively small and numbers of rod fuel elements uneco
nomically large. Use of oxide fuel elements in plate form requires
maintenance of low rates of fission product gas release from the oxide
to prevent clad bulging. Thus, plate fuel elements require utiliza
tion of high density fuel from which, as shown in Sect. 9.4.4 mini
mum fission gas release rates can be obtained. The results of the test
ing indicate that gas release can be maintained at adequately low
levels to permit the use of this fuel element concept to high fuel
burnups [174]. An additional advantage of the plate fuel element
ability
its

is accommodate volume changes the fuel without


to

in

cladding failure. Because the biaxial loading applied under in


of

ternal pressure stressing, zirconium alloy tubing has been found to


crack percent diametral expansion corresponding to
at

less than
5

percent fuel expansion. On the other hand, plate fuel ele


10

less than
ments have resisted the swelling UO, fuel percent
50
of

to
as

as

much
thickness change before cracking because the ability plate clad
of

of
by

by

dings accommodate volume changes bending rather than clad


to

thinning [124]. Nevertheless, the advantages high degree of


of
a

clad restraint restricting fuel swelling discussed Sect. 9.5-2


of in

in

favor the use higher strength fuel element shapes. Thus, plate
IRRADIATION EFFECTS IN URANIUM DIOXIDE 645

oxide fuel elements appear particularly suitable for use in high power
density reactor applications.

(d) Effects of Fuel Composition

Because of the ease of fabrication of high density oxide pellets


employing nonstoichiometric UO,..., a number of irradiation experi
ments have been performed using such oxide [107]. As would be
anticipated from the reduced thermal conductivity of nonstoichio
metric oxide (Fig. 9.73), thermal performance capabilities of such
oxide are less than those of the stoichiometric oxide. Bain and
Robertson showed the occurrence of equiaxed grain growth in fuel
elements containing UO2.14, whereas similar elements containing
UO2.0 at the same thermal rating, ſ/CdT=25 w/cm, showed, as
expected, no structural change [179]. Murray, Pugh, and Williams
point out two additional respects in which nonstoichiometric UO,
is expected to differ from the stoichiometric oxide [94, 166]. A higher
oxide of uranium, possibly UOs, is volatile from UO., at elevated
temperatures. Such a higher oxide may evaporate from hotter zones
and decompose upon condensing at colder locations in the fuel. Thus,
a cyclic time-dependent process resulting in removal of oxide from
the center of the fuel and depositing it elsewhere may be set up. This
process was used to explain the formation of a central void in ir
radiated UO., pellets. Secondly, as a result of the enhanced sinter
ability of UO2.x over that of UO, it was postulated that radial cracks
formed during startup may heal at temperature and continually re
form and reheal as the fuel elements heat and cool. This process
was used to explain increased fission gas release rates from the
nonstoichiometric material, to 80 percent (see Ref. 107, test
62
CR-IV-X-2-f) as well as the formation of a fibrous columnar struc
ture through the entire pellet cross section. In view of the sharply
reduced thermal conductivity of UO, is and the large sheath-fuel

clearances used in assembly of the test specimens, quite high tempera


tures were developed during this test which undoubtedly contributed
to the severe structural changes observed. Furthermore, the results
of Lindner and Matzke predict a somewhat higher fission gas release
rate from UO... than from the stoichiometric compound [78]. At
any rate, utilization of nonstoichiometric oxide possesses no apparent
irradiation advantage over that of UO, and application highly
its

in

rated fuel elements not recommended.


is

Experimentation proceeding the incorporation


on

of

metal fibers
is

into oxide fuel pellets improve their thermal conductivity. Prog


to

by

Kane who discussed the incorporation


in

ress this work reported


is

molybdenum and niobium fibers swaged UO, and Tho.-UO,


of

in
646 URANIUM DIOXIDE: PROPERTIES AND NUCLEAR APPLICATIONS

fuel elements [163]. Kane points out that considerable improvement


in thermal conductivity is anticipated, particularly at elevated tem
peratures. The only irradiation experiments on such fuels are re
ported by Neimark and Kittel who exposed Thoa-base pellets contain
ing 10, 30, and 50 weight percent UO, with 85 to 96 percent density
containing 10 percent Nb or Mo fibers in capsule irradiations [168].
They observed that, as compared with unfibered fuel specimens, much
less cracking and structural change occurred which was in agreement
with the anticipated improvement in thermal conductivity. Insuffi
cient work has been performed to indicate quantitatively whether the
improvement in conductivity sufficiently compensates for the fuel
dilution and increase in structural material content of the fuel element
by the metal fiber additions to warrant their use.
A number of other oxide fuels have been investigated for power
reactor fuel element application, although none in the detail devoted
to the exploration of UO2. An extensive application of Tho, fuel
containing 6.36 weight percent enriched UO, has been made to the
fuel elements of the Borax IV reactor [168, 180]. About 2,800 lead
bonded, aluminum-clad fuel elements containing 97 percent dense
pellets were operated for a year in this reactor at a peak heat rating

J. "Kd7–18 w/cm without


surface
incident. Experimental samples have

been examined after 21,000 MWD/Ton burnup at


ſº
T,
KalT-25

w/cm and showed only a 2-percent volume increase. Additional


irradiation tests are reported by Kittel and coworkers [170, 180].
The formation of central voids and columnar grains in a sample
containing 38-inch diameter, 10 percent UO, in Tho, pellets irradiated

at
ſ " Kd7–41
T.
w /cm with grain growth occurring at
-
a value

ſ
Te...
Kd7–22 w/cm indicates by comparison with the values listed

in Table 9.19 that this fuel material possesses no thermal perform


ance advantage over UO2, in agreement with the results of ther
mal conductivity measurements. Robertson, et al., reported irradia
tion in a high temperature water loop of a number of samples
containing hydrogen-sintered pellets, 93 percent dense, 9 weight per
cent UO, in Tho, and air-sintered pellets, 86 percent dense, 8.8 weight
percent UO, in Tho..., 0.357-inch diameter in Zircaloy-2 tubes at

heat ratings

MWD/Ton–Tho,
of
ſ " AdT-20
T.
to 30 w/cm to burnups of about 5,000

[95]. No structural change was observed in the

You might also like