Вы находитесь на странице: 1из 9

Nuclear Engineering and Design 238 (2008) 2792–2800

Contents lists available at ScienceDirect

Nuclear Engineering and Design


journal homepage: www.elsevier.com/locate/nucengdes

On the nature of aerosols produced during a severe accident of


a water-cooled nuclear reactor
M.P. Kissane ∗
Institut de Radioprotection et de Sûreté Nucléaire (IRSN), DPAM/SEMIC, B.P. 3, 13115 Saint-Paul-lez-Durance Cedex, France

a r t i c l e i n f o a b s t r a c t

Article history: Particle behaviour depends strongly on classic characteristics, e.g., size, and less macroscopic ones involv-
Received 30 May 2007 ing structure and composition these being especially important in situations of strong differential forces
Received in revised form 22 May 2008 on a particle, i.e., surface impact or intensely-shearing flows. The former situation may lead to parti-
Accepted 7 June 2008
cle deposition or break-up and re-entrainment (with potential accident-management implications). This
paper reviews information on aerosols from prototypical experiments identifying common features and
typical variations. It emerges that a particle comprising one-third metal, one-third metal oxide and one-
third a mixture of fission-product species would not be out of place in any potential reactor-accident
sequence. Particle shapes appear relatively compact without branching chain-like structures. On size and
structure, aerosols in the upstream part of the primary circuit would comprise a near-lognormal popu-
lation with AMMD no more than 2 ␮m and geometric standard deviation around 2, particles comprising
agglomerates of highly-coordinated clusters as small as 0.1 ␮m. In the containment, aerosols can typi-
cally be represented by primary-circuit particles and their agglomerates though particular circumstances
(core–concrete interaction, hot-leg accident sequence) can alter this simple picture.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction easily-trapped supra-micron particles fragmenting to produce sub-


micron ones for which trapping is less efficient. Such behaviour is
The behaviour of an aerosol particle can depend strongly not just of interest in assessments of the potential consequences of hypo-
on classic properties such as aerodynamic size but on properties thetical severe accidents such as those performed by the French
that are less macroscopic involving internal structure and compo- Institut de Radioprotection et de Sûreté Nucléaire—the reader inter-
sition (where these are not necessarily independent parameters) ested in severe accidents is referred to IRSN (2007) which deals
in the general, realistic case of agglomerated, compound particles. with pressurized-water reactors (PWRs) since background infor-
A particular concern is that the role of structure and composition mation on this complex field will not be provided here. Another
is especially pertinent to situations of strong differential forces on issue arises concerning experiments: given that particle composi-
the particle such as during impact with a surface or in flows involv- tion and structure can have safety implications, an experiment not
ing intense shear. Particle deposition can be highly sensitive in the using prototypical particles and involving high impact velocities
former situation by leading to, e.g., capture of the particle on the and/or intense shear may not produce aerosol behaviour repre-
surface or its break-up and re-entrainment in the form of a num- sentative of what would happen in an accident. Hence, this paper
ber of fragments. With respect to the latter situation, evidence of can also serve as an initial step in the process of assessing the use
particles breaking up due to intense turbulent shear and passage of non-prototypical particles in nominally representative experi-
through a shock wave was observed in high-velocity flow tests of ments performed in severe-accident conditions.
the ARTIST programme (Güntay et al. (2004)) then confirmed and In this paper, information on aerosol characteristics from pro-
studied in detail in associated work, Ammar (2008). This is not an totypical experiments (e.g., those that, for one thing, produced
isolated example of break-up in a shock wave, e.g., see Brandt et al. aerosols from over-heated irradiated nuclear fuel) is reviewed in
(1987) and Strecker and Roth (1994). Evidently, particle break-up order to identify common features and typical variations. To the
may have significant safety implications where it is a question of author’s knowledge, no such review is currently to be found in
the public domain. For completeness, both aerosols in the reac-
tor coolant system (RCS) and the reactor containment are dealt
∗ Tel.: +33 4 42 19 95 72. with though, it must be said, information on these latter particles
E-mail address: martin.kissane@irsn.fr. turns out to be quite limited. Subsequent effort will need to be

0029-5493/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.nucengdes.2008.06.003
M.P. Kissane / Nuclear Engineering and Design 238 (2008) 2792–2800 2793

devoted to detailed consideration of whether interparticle bond- constant and very similar in mass, U was relatively minor in the
ing forces within prototypical and non-prototypical agglomerated first hour at 1860 K evolving to be the main contributor in the third
particles are comparable. This aspect could be resolved relatively (very approximately: 42% U, 26% Sn, 33% Cs). Particle composition
simply by calculating Van der Waal’s forces between primary par- as a function of size and statistical size-distribution characteristics
ticles, e.g., Hamaker (1937), and estimating the bond structure (the were not measured.
so-called internal co-ordination) of the agglomerates from images
of prototypical particles and the non-prototypical ones used in 2.1.2. Severe fuel damage programme (SFD) in the power burst
experiments. However, given the complex composition of core- facility (PBF)
melt-produced aerosols (see below), major difficulties arise from Information is available from analyses of aerosols produced dur-
the heterogeneity of primary-particle composition and, potentially, ing the most prototypical test of the large-scale PBF programme,
chemical reactions and sintering at points of contact. test SFD 1-4, Petti et al. (1989), Petti et al. (1994), which simulated
a small-break loss-of-coolant accident (LOCA). The test comprised
2. Review of available information a bundle of 28 Zircaloy-clad fuel rods (of which 26 with an aver-
age burn-up of 36 GWd/tU) with 4 stainless-steel-clad Ag–In–Cd
A search for literature providing information on aerosols control rods. The presence of 4 control rods leads, according to
produced in prototypical, severe-accident conditions shows public- the authors’ own figures, to a higher-than-normal proportion of
domain information to be rather meagre. Three principal sources control-rod alloy, e.g., a ratio of 3.5 higher than was the case for
have been found for particles relevant to the RCS and only one with the TMI-2 core. The Zircaloy content was also somewhat high,
respect to the containment. around 6 kg as opposed to nearly 15 kg of UO2 . With pressure fixed
at 6.9 MPa and the bundle immersed in water, a flow of water at
2.1. Aerosols in the RCS 0.6 g s−1 was injected while bundle nuclear power was increased.
Nowhere is it stated that the initial or injected water was borated
The information reviewed excludes experiments with signif- and it must be presumed that it was not. During a 1-h transient,
icant non-prototypical features, these being typically a flawed dry-out and extensive degradation occurred with peak tempera-
source (e.g., low temperature, incomplete or uncertain inventory) ture thought to have reached nearly 3100 K. Vapours and aerosols
and injection of only inert carrier gas. The following experiments were transported along a line initially heated to 800 K descending
constitute the best information available where the focus here is not to ∼600 K in the aerosol sampling zone (no steam condensation).
on particles close to their point of formation (for which, in any case, SEM was used on deposits on a specific deposition structure over
there is virtually no information) but more in relation to hot-leg a 2-m distance starting about 0.3 m above the bundle. This analysis,
conditions, i.e., after some amount of conditioning (cooling, further concerning only (seemingly) individual particles, gives no statis-
condensation of vapours, agglomeration—the phenomena that can tics on the size distribution and, being post-test, can be misleading
occur are described in IRSN (2007)) where particle evolution is less given that particles may have agglomerated upon deposition. The
rapid. SEM analyses showed some very large particles, 25–250 ␮m, on
horizontal surfaces. More interestingly, while impactors were not
2.1.1. AECL experiments used (and, consequently, no possibility of quantitative information
Mulpuru et al. (1992) performed small-scale experiments on on composition as a function of aerosol size), large variation in
Zircaloy-clad CANDU fuel samples heating them to and holding deposit composition is seen at the same axial location depending
them at 1860 K in a flowing steam-rich atmosphere at ambient on whether the deposition surface (inserted coupons) was vertical,
pressure.1 It is recalled that CANDU fuel has a typical end-of-life i.e., in the direction of the flow, or horizontal allowing (primar-
burn-up of ≤8 GWd/tU, much lower than for LWRs. In addition, ily) settling. In terms of the structural elements and volatile fission
apart from the cladding, no structural materials relevant to LWRs products (FPs) at the 0.6 m elevation, Table 1 provides the details.
were present in the furnace. Thermal-hydraulic boundary condi- Since thermophoretic deposition is relatively insensitive to size if
tions are not well-known between the specimen and the exit of the particles are not very small, and since gas-to-coupon temperature
furnace tube where temperature (of the gas, presumably) dropped difference is seen to be between 20 and 70 ◦ C at this elevation dur-
to 380 K and, schematically, is indicated to have remained at this ing the period of release (Fig. 38 of Petti et al. (1989)), it can be
level up to and including the aerosol sampling zone. In this zone argued that the vertical-surface deposit occurred mainly due to
aerosols were collected during four distinct phases: during heat up thermophoresis and that its composition can be considered rep-
then during three 1-h periods at the hold temperature of 1860 K. resentative of that of the overall suspended aerosol at this location.
The four sequential systems comprised thin platinum wires for Furthermore, horizontal-surface deposits should primarily repre-
individual SEM2 (size) and WDX3 (elemental composition) analysis sent the larger size fraction by being biased towards settling and
with the outflow from these proceeding downstream to a common inertial deposition due to geometry changes. It can be inferred
filter. that, over the duration of the test in this zone, significant variation
The experimenters conclude that spherical particles of around
0.1–0.3 ␮m formed (though their composition was not established)
Table 1
then these agglomerated giving rise to a mixture of compact parti- Aerosol composition (mass %) in terms of the measured elements (NB: data for In
cles between 0.1 and 3.0 ␮m in size at the point of measurement. are not shown due to being below detection limits)
The composition of the particles was found to be dominated by
Location (m) Position/time Ag Cd Sn Zr Volatile FPs
Cs, Sn and U; while the Cs and Sn mass contributions remained
Vertical surface 0.2 8.3 44.2 13.3 34.0
0.6
Horizontal surface 0.5 75.1 0.3 0.2 24.0

1 2040 s (filter 2) 1.9 6.2 75.9 3.9 12.0


It is noted that these high humidity and relatively moderate temperature condi-
2100 s (filter 3) 1.4 27.9 39.5 2.6 28.6
tions are appropriate to CANDU reactors; the reader is referred to Dick et al. (1990)
13 2385 s (filter 4) 1.6 24.2 36.3 0 37.9
for a description of CANDU severe accidents.
2 2850 s (filter 5) 1.2 13.3 39.9 0 45.6
Scanning electron microscopy.
3 3060 s (filter 6) 3.9 12.5 29.7 2.4 51.5
Wavelength dispersive X-ray.
2794 M.P. Kissane / Nuclear Engineering and Design 238 (2008) 2792–2800

in aerosol composition occurred where larger particles contained For a representative view of the suspended aerosols relevant to
a very high cadmium fraction whereas smaller particles were far the hot leg of a PWR we need to look at deposits or samples not too
more heterogeneous in composition. close to the bundle in a zone where vapours have stabilized at near-
Further interesting measurements for purposes here were six equilibrium values. This is characteristic of the upper part of the
isokinetic, sequential, filtered samples located about 13 m from so-called vertical hot line and downstream of this location. Data on
the bundle outlet. These were used to follow the evolution of the deposits (predominantly thermophoresis) in the vertical hot line do
aerosol composition and to examine particle size (SEM). Based on not provide a complete coverage of elements (e.g., data for uranium,
these analyses the authors state that particle geometrical-mean known to be a major contributor, appears to be lacking). However,
diameter varied over the range 0.29–0.56 ␮m (elimination of the it is clear that the volatile FPs Cs and Te are significant contribu-
first filter due to it being early with respect to the main transient tors as is Ag. Note that Cd is predominantly in the vapour phase
gives the range 0.32–0.56 ␮m) while standard deviation fluctuated at the hot-line temperature and so is negligible in deposits until
between 1.6 and 2.06. In the images of filter deposits needle-like lower temperatures are reached. More quantitative data is avail-
forms are seen but, generally, particles appear fairly spherical. Turn- able from aerosol sampling performed downstream using filters
ing to composition, if the first filter sample is eliminated and “below and impactors both in a zone at 700 ◦ C and another at 150 ◦ C. Table 2
detection limit” is taken as zero, for the structural components and summarizes data on aerosol composition where it is worth noting
volatile FPs we have in terms of percentages the values given in that different phases of the bundle transient have been covered, viz.
Table 1.4 the bundle oxidation runaway (peak at around 11,300 s), the stabi-
Lastly, it should be noted that no data are provided on uranium lization period (roughly 11,500–14,600 s), the secondary oxidation
release; though a very low release in fractional terms, such a release phase (roughly 15,400–17,040 s), the advanced degradation phase
is potentially very significant in terms of absolute mass. It might be (roughly 16,000–17,040 s).5
concluded that the release was below detection limits since the Concerning particle size, impactors were used to take samples
range of analysis techniques applied – which included alpha analy- at both 700 and 150 ◦ C. These samples are, of course, punctual and
sis – ought to have measured any significant U in deposits (though may only be indicative of a particular phase of the transient. The
the authors do not state this explicitly). Furthermore, the test condi- results indicate an aerosol population at 150 ◦ C that is fairly lognor-
tions at high temperature comprising the low water-injection rate mal with an AMMD6 around 3 ␮m and a standard deviation of about
and the metal-rich bundle led to full reduction of the steam dur- 2. The population at 700 ◦ C is less clearly lognormal and somewhat
ing most of the transient as measured downstream of the bundle smaller in mean size. However, these results must be treated with
(Section 5.2.1 of Petti et al. (1989)). This total starvation, probably some caution as the impactor plates were heavily overloaded and
occurring from the hottest point in the bundle upwards, would have sampling flow was not steady: the impactors were not always func-
been unfavourable to uranium release from the bundle. tioning in their range of calibration. Furthermore, in the opinion of
the author, this size information is incompatible with the absence
2.1.3. The Phébus FP experiments of enhanced deposition by impaction in bends indicating particles
For the Phébus tests a great deal of information is available on with a smaller mean size. SEM analyses of the impactor plates for
aerosols that can either be inferred from deposits or arises from FPT0 and FPT1 as well as of filter samples show particles to com-
direct measurements. However, at the present time, only two tests prise agglomerates of particles in the size range 0.1–0.5 ␮m—see
are considered “open” (allowing results to be freely discussed): Fig. 2.
tests FPT0 and FPT1, Clément et al. (2003) and Jacquemain et al. Finally, there is some information on aerosol structure pertain-
(2000). Furthermore, though two other tests, FPT2 and FPT3, are ing to the sample shown in Table 2 taken at the 700 ◦ C location
of great interest here due to the presence of boron (boric acid in late in the transient. This concerns an XPS7 sputtering analysis of
the injected steam for FPT2, boron carbide in the control rod of the the deposit on the filter concerned and so cannot be interpreted as
bundle for FPT3), the final, self-consistent data have not yet been statistically significant. Nevertheless, a depth profile is seen with
issued for these tests and partial results would have to be treated an indium-rich surface layer (also carbon- and oxygen-rich, both
with caution. Hence, presentation of results from the FPT1 test will results to be treated with caution since the former is very prob-
be the focus here. It is worth noting that FPT1 complements the ably contamination while the latter is probably partly the result
PBF-SFD 1-4 test rather well by being at low pressure (<0.3 MPa) of post-test oxidation) from which Cs and Sn are notably absent.
and comprising highly oxidizing conditions apart from a very brief Progressing deeper, Cs and Sn appear below the surface while con-
phase where molar hydrogen concentration peaked at about 50%. centrations of uranium (thought to be probably U3 O8 , i.e., UO2
For purposes here, FPT1 (and FPT0) can be described as com- oxidized post-test8 ) and rhenium increase monotonically. Ag, Cl,
parable to PBF-SFD 1-4 with a slightly smaller bundle in terms of O (once below the surface) and Ni (from the Inconel filter) con-
the number of fuel rods where these were of lower burn-up, viz. centrations appear relatively uniform as a function of depth. It is
23 GWd/tU (trace irradiated for FPT0). Fig. 1 describes schemat- also interesting to note that these results with respect to Cs are cor-
ically the Phébus FPT0 and FPT1 apparatus. A single steel-clad roborated by progressive-dissolution analyses (water, alkali, acid,
Ag–In–Cd control rod was placed centrally in the bundle. There strong acid). The most striking results of these analyses concern the
were about 3.5 kg of Zircaloy for just over 10 kg of UO2 : this is a very limited water-solubility of Cs and Rb in two samples widely
somewhat lower Zrly–UO2 ratio than in SFD 1-4. Steam was injected separated in time taken at 700 ◦ C. It was found necessary to apply
at a rate typically around 1.9 g s−1 —whence the oxidizing conditions
with respect to SFD 1-4. The line between the bundle and the first
release measurement station was maintained at around 970 K (i.e., 5
Compilation of this table used Fig. 5.2-79 of Jacquemain et al. which provides
higher than in PBF). results of ICPMS/OES analyses of filter and impactor samples. SEM/EDX analyses
are also available for three of the five samples showing significant differences, e.g.,
greater Cd content, different (either higher or lower) Sn content and some Zr in the
sample at 700 ◦ C.
4 6
Compilation of this table relies on Table VI of Petti et al. (1994) which provides Aerodynamic mass-median diameter.
7
the volatile-FP content. Data in that table have been taken as mass percentages since X-ray photoelectron spectroscopy.
8
molar percentages would imply even higher contributions of volatile FPs given that Note that evidence of post-test oxidation of samples was also seen in FPT0 where
the other principal contributions arise from Cd and Sn. indium in aerosol samples was found in the form In2 O3 .
M.P. Kissane / Nuclear Engineering and Design 238 (2008) 2792–2800 2795

Fig. 1. Schematic diagram of the Phébus apparatus, FPT0 and FPT1 configuration.

Table 2
FPT1 aerosol composition (mass %) in terms of the measured elements (NB: data adjusted by subtracting Re, W and any other non-prototypical contributions)

Location (m) Position (time) Ag In Cd Sn U Mo Cs Others



13 700 C (17,034 s) 29 0 0 14 18 16 9 13

150 ◦ C (11,051 s) 32 22 9 18 0 4 9 6
150 ◦ C (13,810 s) 30 8 3 8 6 17 24 3
25
150 ◦ C (16,473 s) 52 3 0 3 27 5 3 6
150 ◦ C (17,034 s) 46 10 4 6 13 8 6 8

strong acid in order to dissolve the remaining 70% of these two FPs (as already said) but the potential migration behaviour of Cs and
after only 30% dissolved in water, alkali and acid washes. Either Rb may not have occurred on the time scale of their transport to
they migrated chemically into a protective aerosol substrate after the sampling point but over a much longer period; the relevance of
condensing on its surface (probably non-uniformly) or they were migration to consideration of the nature of aerosols as they reach
“encased” by another/other species condensing subsequently. Te the break in the primary circuit is debatable.
and Mo, both elements that may have been chemically associated To summarize what can be concluded from the above informa-
with Cs and Rb in the vapour phase, are similarly insoluble with tion, and to some extent paraphrase the experimenters (executive
respect to the early sample but mostly soluble in weak acid in the summary of Jacquemain et al.), condensed material was trans-
late sample. W (released from thermocouples), Sn and Cd display ported in the hot leg and cold leg as mixed aerosols, the bulk
similar behaviour in both samples. Low-temperature samples do of which were dominated by control rod (mainly Ag), structural
not show this behaviour where Cs and Rb both exhibit high solubil- (mainly Re, Sn) and fuel material (U). Low-volatility FPs were prob-
ity in water. A cautionary note must be raised with respect to these ably associated with this core of low-volatility materials upon
results since not only is post-test oxidation of the sample apparent which more volatile compounds (control rod In and Cd, Sn from
clad, Cs) condensed. For Cs and Rb, diffusion and chemical reac-
tion with the low-volatility substrate probably occurred. Aerosol
composition depended on the stage of degradation (i.e., condition-
ing release) where highest aerosol concentrations occurred during
bundle oxidation when Ag, In, Cd and Sn were dominant. This sub-
sequently became Ag, Re, Cs and Mo then Ag, Re and U during the
final degradation phase with formation of a molten pool. Impactor
data show the composition to be fairly independent of particle size
(though, in the opinion of the author, this result must be treated
with caution—see above remarks related to aerosol size).

2.1.4. Other sources of information


A piece of information from the European 4th Framework
project OPSA is worthy of note here. One phase of this project
involved realistic tests investigating the consequences of air ingress
for an unirradiated 9-rod fuel bundle where aerosol measurements
were included, Csordás et al. (2000). The absence of FPs limits the
value of the results in the present context but one fact is intrigu-
ing: the smallest particles measured were uranium-rich, about
Fig. 2. SEM images of aerosols from an impactor plate in the circuit of the Phébus
0.1–0.5 ␮m in size and fairly compact (rectangular).
FPT0 test showing an agglomerated structure of particles typically in the range Finally, it is noted that there is a significant amount of infor-
0.1–0.5 ␮m. (Fig. 20 of Clément et al. (2003)). mation on fuel particles produced during the Chernobyl accident.
2796 M.P. Kissane / Nuclear Engineering and Design 238 (2008) 2792–2800

Fig. 3. SEM images of a filter deposit in the containment of the Phébus FPT1 test showing a fairly uniform structure of typically sub-micron particles (Fig. A9-8 of Jacquemain
et al. (2000)).

This source of data can be discounted in the present context since will be covered here since information on aerosols from this source
the aerosols differ significantly from those produced by a LOCA. is relatively adequate.
The vast majority of such aerosols represent whole-fuel micro-
fragments produced by the initial shattering of the fuel resulting 2.2.1. The Phébus FP experiments
from the explosion; subsequent emissions produced aerosols via Information from the Phébus FP programme is derived mainly
the evaporation–condensation–agglomeration route characteristic from the following two references on the two open tests of this
of LOCAs but the conditions (graphite fire, formation and pro- programme: FPT0, Clément et al. (2003), and FPT1, Jacquemain et
longed ageing in highly oxidizing conditions, high dilution in al. (2000). We recall that FPT1 is the more representative of the two
the gas phase) mean that these cannot be considered relevant tests since FPT0 used only trace-irradiated fuel. Fig. 1 describes the
here. Phébus FP apparatus.
Firstly, with respect to aerosol composition, it must be noted that
2.2. Aerosols in the containment the relative humidity in these tests was never more than about 60%
in FPT0 and 85% (at the end of the aerosol production phase) in
The identification of typical features for aerosols in the con- FPT1 so any steam condensation on aerosols would have required
tainment is more problematic than for aerosols in the RCS. Not a hygroscopic effect. Given that analyses do not predict formation
only does a wide diversity of accident sequences exist with vary- of strongly hygroscopic substances prior to release into the con-
ing pre-conditioning of the source in the RCS before release to the tainment, e.g., Kissane and Drosik (2006), a hygroscopic effect is
containment, but also information for (more or less) prototypi- difficult to justify except, possibly, for a short time and to a limited
cal particles is dependent on just one experimental programme, extent in the FPT1 test in relation to the presence of caesium iodide.
i.e., Phébus PF. In addition, the timescale of evolution of aerosols Hence, with negligible water contribution, regarding the elemen-
in the containment – about 1 day – leaves room for transfor- tary actinide, FP and structural content, aerosol composition was
mations due to radiolysis, oxidation, formation of (bi)carbonates, identical to the average composition of aerosols leaving the circuit
etc., to occur; this area is virtually unstudied beyond some (c.f. Table 2 for FPT1 at the 25 m location), viz., in decreasing order:
of the consequences for iodine species, e.g., Dickinson et al. silver, tin, indium, uranium and cadmium for the structural ele-
(2003). ments where the tin, indium and uranium contributions are similar.
Further complications arise in the containment from the (poten- Of the FPs, molybdenum and caesium were significant contributors
tial) occurrence of major secondary sources of aerosol material, (data from FPT1, c.f. Table 2).
i.e., other than the primary source generated by a degrading core.9 The aerosol size distributions were fairly lognormal with an
These are, in particular, pressurized ejection of molten corium average size (AMMD) in FPT0 of 2.4 ␮m at the end of the 5-h bundle-
(high-pressure sequences), hydrogen deflagrations and molten- degradation phase growing to 3.5 ␮m before stabilizing at 3.35 ␮m;
core–concrete interaction (MCCI). Only the latter of these, MCCI, aerosol size in FPT1 was slightly larger at between 3.5 and 4.0 ␮m.
Geometric-mean diameter (d50 ) of particles in FPT1 was seen to be
between 0.5 and 0.65 ␮m; a SEM image of a deposit is shown in
9
It is noted that mechanical resuspension of aerosols deposited in the RCS due to Fig. 3. In both tests the geometric standard deviation of the lognor-
steam explosion(s) resulting from core collapse into residual coolant in the vessel mal distribution was fairly constant at a value of around 2.0. There
can be bundled in with the primary source from the core rather than be treated as was clear evidence that aerosol composition varied very little as
a major secondary source of aerosols. This is because it can be shown that the short
delay between peak core emissions and this subsequent steam-explosion-driven
a function of particle size except for the late settling phase of the
source leads within a few hours to insignificant differences for the population of FPT1 test: during this period, the smallest particles were found to
aerosols suspended in the containment. be caesium-rich. In terms of chemical speciation, X-ray techniques
M.P. Kissane / Nuclear Engineering and Design 238 (2008) 2792–2800 2797

were used on some deposits and there also exist many data on the 50% or more of the aerosol mass. Releases of tellurium and neutron-
solubilities of the different elements in numerous deposits giving absorber materials – silver, indium and boron (from boron carbide)
a clue as to the potential forms of some of the elements. However, – were high. Releases of uranium and low-volatility FP elements
post-test oxidation of samples cannot be excluded since storage were small in all tests. During ablation of the concrete, aerosol
times were long (months) and the value of speculating on poten- composition remained fairly stable and particles were compact but
tial speciation on the basis of the available information is debatable. varied considerable in size the majority being typically micron-
Nevertheless, there is clear evidence that some elements reached sized (geometric diameter) but with some considerably larger sizes.
higher states of oxidation in the containment when compared to In summary, it can be said that MCCI is likely to add a large
their chemical form in the circuit. amount of non-active aerosol material to the containment atmo-
sphere in the size range of the existing aerosols thus promoting
2.2.2. Experiments on molten-core–concrete interaction (MCCI) agglomeration and diversification of the aerosol composition.
The composition of aerosols produced by interaction of molten
corium with concrete is different from that of aerosols generated 2.2.3. Other sources of information
by a degrading core not least due to the addition of compounds As for RCS aerosols, it is worth noting that information from the
of sodium, potassium, magnesium and calcium arising from con- Chernobyl accident is not relevant here (see Section 2.1.4).
crete erosion. The phenomenology leading to aerosol generation
is quite complex and beyond the scope of the present work; in-
3. Discussion
depth coverage is provided in Powers et al. (1985). Here we indicate
only that the source to the containment atmosphere is produced
3.1. Uranium contribution
by evaporation into bubbles (hydrates, hydroxides and carbonates
within the concrete are thermally decomposed producing steam,
The Mulpuru et al. experiments are interesting in that they show
carbon monoxide and carbon dioxide) and at the surface of the
a progression of the uranium content which can only be due to the
corium–concrete mixture as well as by mechanical generation of
progressive oxidation of the UO2 ; however, its dominant contri-
droplets due to sparging.
bution at the later times is not confirmed by the Phébus results
Concerning experiments on MCCI, no tests exist in which
though, here too, a progressive contribution is observed. While the
irradiated fuel was used but some tests used simulant corium
fractional uranium release is very low its dominance in the total
(i.e., a non-radioactive chemically-analogous mixture) including FP
inventory makes this release a major contributor to total aerosol in
species. In addition, for purposes here, few experiments involved
later stages of the transient unless conditions remain reducing (c.f.
satisfactory measurements of the released materials. Early tests of
the contrast with the PBF-SFD 1-4 results).
interest are the NSS series and two tests of the Transient Urania
Reacting with Concrete series (TURC-2 and TURC-3) both per-
formed at Sandia NL where these and others were reviewed by 3.2. Tin contribution
Brockmann (1987). However, neither series included sustained
heating of the corium simulating decay power so the composi- The tin contribution (as oxide from Zircaloy oxidation) to aerosol
tion of the generated aerosols was potentially biased towards a mass is always present. It is not very significant in Phébus FPT0
higher-than-average fraction of the more volatile species. Concern- (results not shown here) and FPT1 outside the two main oxida-
ing the aerosol size distribution, results from the two series are not tion phases in the bundle, i.e., the relatively short (with respect
dissimilar: mean size in TURC is generally small, micron-sized aero- to the total transient) oxidation runaway and the much later and
dynamic diameter with a geometric standard deviation of about 2, gentler secondary oxidation phase. In the reactor case, this variabil-
but with a larger mode of particles (typically several microns) also ity would be smoothed out (see Section 3.8 below). Tin comprises a
observed; NSS produced particles with geometric mass mean diam- very significant contribution in the other experiments, probably too
eter between 1.2 and 3.2 ␮m with a wide spread of the size range significant in PBF-SFD 1-4 due to the high Zircaloy fraction in the
(geometric standard deviation between 2 and 2.8). In addition, the bundle. In summary, tin has a reliable presence whatever the con-
NSS tests suggested a somewhat heterogeneous aerosol composi- ditions. It must be noted that this observation applies to Zircaloy 4
tion with higher volatility components such as and Cs and Sn being whereas newer cladding materials and VVER cladding have a lower
more concentrated in the smaller aerosols. or zero tin content.
It is worth noting an interesting series of intermediate-scale
experiments that indicated generation of small aerosols (sub- 3.3. Silver contribution
micron AMMD) and a larger contribution to aerosol composition
from uranium mainly in the larger particles, Spencer et al. (1988). The silver contribution to aerosol mass in Phébus FPT0 (results
We also mention the large-scale Sustained Urania Reacting with not shown here) and FPT1 is much higher than in PBF-SFD 1-4
Concrete (SURC) tests at Sandia NL that addressed aerosol release despite there being proportionally less silver in the Phébus bun-
(as well as thermal-hydraulic phenomena in the cavity) associated dle. It might be thought that this can be accounted for by purely
with prototypical core-melt materials in various types of concrete an over-pressure effect in Phébus where the lower system pressure
crucible, Burson et al. (1989), but few results are readily available led to mechanical formation of droplets when the control rod burst.
for this programme. However, Ag is found in all samples spread over time so clearly it is
Finally, the richest source of information comes from the evaporating as degradation progresses. Perhaps it is important that
seven large-scale experiments of the International Advanced Con- the control rod in Phébus was centrally placed and experienced the
tainment Experiments (ACE) programme on melt behaviour and highest bundle temperatures. Or that the transient in PBF was more
aerosol release during MCCI, Fink et al. (1995). These addressed than twice as fast as that of FPT1 and more rapid relocation of the
four types of concrete (siliceous, limestone/sand, serpentine and molten Ag–In alloy occurred (Cd being rapidly released). In the end
limestone) and a range of metal oxidations for both boiling-water it may not matter much in the RCS because it seems that either Cd
reactor and PWR core debris. The released aerosols contained and/or Ag contributes in a major way and both exist predominantly
mainly constituents of the concrete. In the tests with metal and in the metallic phase in the aerosols; in the containment it is a dif-
limestone/sand siliceous concrete, silicon compounds comprised ferent matter since silver and not cadmium has a major impact on
2798 M.P. Kissane / Nuclear Engineering and Design 238 (2008) 2792–2800

the behaviour of iodine, Clément et al. (2003) and Dickinson et al. It is worth noting that the information available for contain-
(2003). ment aerosols does not take into account potential effects from
hot-leg sequences. As has been seen in the Phébus FP tests, many
3.4. Cadmium contribution FP species have significant fractions in the vapour-phase at tem-
peratures above 900 K. This means that, for a hot-leg sequence,
Cadmium is seen to be a very significant fraction of the PBF-SFD such vapours would be released into the containment and either
1-4 aerosols whereas in Phébus its contribution is minor, some- nucleate creating a population of small particles rich in FPs and/or
times negligible. A major influence in this difference is, of course, condense onto existing aerosols (conforming to the assumptions
the different temperature of the line between the bundle and the of current analyses). As has been noted in the past, CSNI (1993),
measurement point. Up to the first measurement station in Phébus, the former situation creates a bimodal source to the containment
point C, temperatures have not dropped below about 1000 K and the where, according to models, agglomeration between the two pop-
cadmium is very largely in the vapour phase as Cd(g). In PBF-SFD ulations can be weak. It is important to take into account here,
1-4 the temperature of the equivalent line descends very quickly to therefore, the potential for a small-sized aerosol rich in volatile FPs
800 K then to around 600 K and the cadmium in this case condenses (notably I and Cs) to exist in the containment and be relatively resis-
contributing to the aerosol population. tant to agglomeration and deposition—its importance with respect
to accident management is also clear.
3.5. Indium contribution
3.8. Variability
It is seen in Phébus FPT0 (results not shown) and FPT1 that
indium contributes significantly to the overall aerosol composition Variability of aerosols as a function of time is to a large extent
whereas in PBF-SFD 1-4 its contribution is not mentioned. This con- dependent on scale being very evident in the smallest-scale tests
trast is consistent with thermodynamic calculations indicating that (Mulpuru et al.) where conditions are virtually homogeneous for
Phébus FPT0 and FPT1 conditions (predominantly near-pure steam the whole fuel sample. The more heterogeneous conditions (extent
for most of the transient) are close to optimal for the volatilization of oxidation, fuel burn-up, temperature field, etc.) inherent at core
of indium via the formation of In2 O(g) and InOH(g), Taylor (1998). scale would lead to significant smoothing of the different releases as
In reducing conditions, such as those of PBF-SFD 1-4, the dominant seen in the primary circuit, and, hence, give rise to a more uniform
evaporating species is the much less volatile metal, In(g). particle composition over the majority of the release-from-core
phase. What would have more impact is the accident scenario, i.e.,
3.6. Fission-product contribution a high-pressure sequence relative to a low-pressure one may well
lead to more variation in the aerosol composition than is seen over
Total aerosol mass is seen in the tests reviewed to be typically the duration of a single, given sequence. This may be particularly
between 10 and 40% FPs (though, in the LWR context here, it must true of control-rod and uranium contributions to overall aerosol
be recalled that the AECL tests give too much weight to FPs by not mass where both of these may be expected to decrease at high pres-
including structural components other than cladding). sure: no bursting behaviour and lower Cd volatility for the former
and greater dissolution by Zrly and less oxidation and volatilization
3.7. Size, shape and structure for the latter.
A final consideration is that of the multi-component nature of
There is no evidence that particles are not compact, i.e., some- the aerosols: how uniform a composition do they have? Does com-
what spheroidal, unlike the branched structures that one finds position vary not only with size, but for a given size are there
for soot, for example. Perhaps once initial clusters have agglomer- particles of significantly different composition? Concerning com-
ated then, as conditions cool, successive condensations of vapours position as a function size, results from VERCORS tests without
on agglomerates inhibit any branching trend or induce collapse control-rod materials provide information on aerosols that have
of any branching (i.e., the liquid surface tension provides suffi- formed close to their point of release: a bimodal population is
cient compacting force). As for size, it is difficult on the basis of observed with smaller, sub-micron particles rich in volatile FPs and
the information reviewed to conclude on a typical size; perhaps larger micron-sized particles rich in less volatile FPs such as Ba,
a near-lognormal distribution with an AMMD not exceeding 2 ␮m Leveque and Boulaud (1994). However, the possibility of agglom-
and a standard deviation of around 2 would seem reasonable in eration was perhaps limited in these tests. Phébus FP implies a
the hot leg. The aerosols comprise agglomerates of clusters with significantly homogeneous composition in the primary circuit and,
a typical size in the range 0.1–0.5 ␮m. This cluster size is much more especially, in the containment. This is an aspect that probably
too large to represent true primary particles since these are typi- merits further investigation with respect to primary circuit aerosols
cally ≤10 nm. It is likely these particles represent very compact – and containment aerosols in the event of a hot-leg break (see the
highly-coordinated – clusters of primary particles mainly of ura- discussion of the potential bimodal distribution in Section 3.7).
nia upon which different vapours may have condensed. It is worth
noting that these clusters are themselves probably very resistant to
3.9. Other considerations
break-up and, hence, constitute the smallest possible size of aerosol
fragments post-break-up. Such a phenomenon has been observed
There is a potentially significant contribution to the aerosol
for titanium dioxide particles where agglomerates of 0.1 ␮m-sized
during an accident arising from the presence of large quantities
clusters of 3 nm-sized primary particles could be broken apart by
of boron. This has not been covered in the experiments reviewed
impact at high velocity onto a surface but the clusters themselves
proved resistant to break-up at impact speeds up to 120 m s−1 ,
Froeschke et al. (2003),10 see Fig. 4.
primary-particle bonding force (with qualitative success). The technique developed
is ideal for the problem of particle representativity in high-flow-velocity experi-
ments if only representative particles were available for comparison with those used
10
This interesting work measured break-up of primary-particle agglomerates as in such out-of-pile tests: their break-up behaviour upon impact could be used as a
a function of impact velocity and attempted to discover a relationship with the measure of representativity.
M.P. Kissane / Nuclear Engineering and Design 238 (2008) 2792–2800 2799

Fig. 4. Tunnelling electron microscope images of agglomerates of titanium dioxide nanoparticles: before impact (left); after impact at a velocity of 45 m s−1 (centre) and
71 m s−1 (right). The primary-particle clusters resist fragmentation (Fig. 3(c) of Froeschke et al. (2003)).

above but useful data will be forthcoming from the Phébus FPT2 and consist for the most part of agglomerates of compact clusters as
FPT3 tests. Boron is present as boron carbide in BWRs, French 1300 small as 0.1 ␮m. The evidence from Phébus implies an “onion-skin”
and 1450 MWe PWRs and VVER-1000s; the quantities involved are type of structure where the kernel of the particle is rich in highly
significant, around 330 kg of B4C in the French PWRs, Adroguer et refractory materials. Vapours of more volatile species containing
al. (1998), and around 270 kg in a VVER-1000, Belovsky (1995). In a caesium and rubidium that have condensed on these refractory ker-
BWR, this amount may exceed one tonne, Kissane et al. (1994). As nels may migrate into and interact chemically with the substrate.
for the borated coolant, it has been estimated that, at the time core In the containment, particles are typically larger representing the
uncovery starts, typically about 46 kg of boron (as boric acid) will particles formed in the primary circuit and agglomerates of these.
be present in the water, Kissane et al. (1994). While these quantities A smaller population of FP-rich particles may form at the breach,
are significantly less than the approximately 3 tonnes of Ag–In–Cd i.e., creating a bimodal aerosol population in the containment, in
alloy in a typical PWR, the boron as boric acid (as added to the the event of a hot-leg sequence.
coolant or as a reaction product of steam and boron carbide) is very In the safety assessment context, it would seem helpful to know
volatile: boron cannot be neglected as a potentially significant con- the potential for representative agglomerates to break-up into sub-
tributor to aerosol mass. It is also noted that borated species such micron component clusters in accident conditions since this could
as CsBO2 have a hygroscopic potential comparable to that of CsI. influence aerosol retention. If this potential turned out to be signif-
icant then further characterization of prototypical particles would
4. Conclusion be valuable (where the complications arising from working with
radioactive aerosols need no enumeration here). Effort could be
Literature on aerosols produced from over-heated urania fuel devoted to characterizing surface bonding forces where a technique
has been reviewed and discussed. such as atomic force microscopy could be useful; indeed, it would
The typical aerosol composition for PWRs is seen to be a mix- be helpful to create a benchmark characterization of prototypical
ture of metals (Ag and/or Cd, 15–40%) with a similar metal-oxide particles against which non-prototypically created particles pro-
content (tin oxide and, potentially, uranium dioxide and zirconium duced for experiments could be compared. This would be useful
dioxide) and FPs covering a diversity of compounds (from metal to with respect to experiments involving high-velocity flows where
oxides to salts, e.g., RbI, to ternary compounds). It would appear that using aerosols that simply reproduce appropriate size and shape
a particle composed of one-third metal, one-third metal oxide and would be inadequate (if the representative break-up potential
one-third a mixture of mainly FP species (salts, ternaries, oxides, turned out to be non-negligible). Helpful information could also
hydroxides) would not be out of place in any of the potential LWR result from theoretical analysis based on estimating the internal
accident sequences. Secondary sources of aerosol material in the co-ordination of agglomerates from images of prototypical parti-
containment can add significant masses of aerosol to those of the cles though, as already said, chemical reactions and sintering at
primary source where, in the case of MCCI, significant diversifica- points of contact introduce considerable complexity.
tion of composition occurs due to agglomeration with the largely
concrete-derived contribution. It must be noted that it has not been Acknowledgements
possible to take account of boron (from boron carbide in BWR con-
trol blades and some French PWRs and VVER-1000 control rods or This work was performed partly as a contribution from IRSN
boric acid in the coolant) in this assessment. to the international ARTIST programme (Paul Scherrer Institut,
Concerning particle shape, relatively compact particles without Switzerland) and partly as a contribution to the activities of the
branching chain-like structures appear to be typical in the RCS. If OECD/NEA/CSNI working group on accident management. IRSN’s
steam condensation on particles occurs (e.g., in the cold leg in a activities are supported by the French Ministry for the Environment
cold-leg sequence, or in the containment) then, due to the surface and Sustainable Development.
tension effect, compaction will occur to the limit of the particle
becoming a spherical droplet.
On size and structure, information is less reliable but it would References
seem realistic for aerosols in the hot leg to comprise a near-
Adroguer, B., Poletiko, C., Autrusson, B., Aouri Mercier, M., Cenerino, G., 1998. Impact
lognormal population of particles with AMMD around 1–2 ␮m and du B4C sur un accident grave: incertitudes et besoins d’essais. IRSN Report (IPSN
geometric standard deviation around 2. The larger particles would Note Technique DRS/SEMAR 98/65, June 1998).
2800 M.P. Kissane / Nuclear Engineering and Design 238 (2008) 2792–2800

Ammar, Y., 2008. Agglomeration and break-up of aerosols in turbulent flows. Ph.D. Hamaker, H.C., 1937. The London–Van der Waals attraction between spherical par-
Thesis, University of Newcastle upon Tyne, UK. ticles. Physica 4, 1058–1072.
Belovsky, L., 1995. Heat release from B4C oxidation in steam and air. In: Proceedings IRSN, 2007. Research and development with regard to severe accidents in pressurised
IAEA-TCM, Dimitrovgrad, Russia, October 1995. water reactors: Summary and outlook. Rapport IRSN-2007/83 & Rapport CEA-
Brandt, O., Rajathurai, A.M., Roth, P., 1987. First observations on break-up of particle 2007/351.
agglomerates in shock waves. Exp. Fluids 5 (2), 86–94. Jacquemain, D., Bourdon, S., De Bremaecker, A., Barrachin, M., 2000. FPT1 final report.
Brockmann, J.E., 1987. Ex-vessel releases: aerosol source terms in reactor accidents. IRSN Report (IPSN PH-PF IP/00/479, December 2000).
Prog. Nucl. Energy 19 (1), 7–68. Kissane, M.P., Bowsher, B.R., Drossinos, Y., Powers, D.A., 1994. Final report of the
Burson, S.B., Bradley, D., Brockmann, J., Copus, E., Powers, D., Greene, G., Alexander, Phebus-FP boric acid task force. IRSN Report (IPSN Note Technique SEMAR
C., 1989. United States Nuclear Regulatory Commission Research Program on 94/1007).
molten core debris interactions in the reactor cavity. Nucl. Eng. Des. 115 (2–3), Kissane, M.P., Drosik, I., 2006. Interpretation of fission-product behaviour in the
305–313. Phébus FPT0 and FPT1 tests. Nucl. Eng. Des. 236 (11), 1210–1223.
Clément, B., Hanniet-Girault, N., Repetto, G., Jacquemain, D., Jones, A.V., Kissane, Leveque, J.P., Boulaud, D., 1994. Fission product aerosols in the programmes HEVA
M.P., Von der Hardt, P., 2003. LWR severe accident simulation: synthesis of the and VERCORS. J. Aerosol Sci. 25 (Suppl.1), 87–88.
results and interpretation of the first Phebus FP experiment FPT0. Nucl. Eng. Des. Mulpuru, S.R., Pellow, M.D., Cox, D.S., et al., 1992. Characteristics of radioactive
226, 5–82. aerosols generated from a hot nuclear fuel sample. J. Aerosol Sci. 23 (suppl.1),
CSNI (report by a group of experts), 1993. Physical and chemical characteristics of S827–S830.
aerosols in the containment. OECD/NEA/CSNI/R(93)7. Petti, D.A., Hobbins, R.R., Hagrman, D.L., 1994. The composition of aerosols
Csordás, A.P., Matus, L., Czitrovszky, A., Jani, P., Maróti, L., Hózer, Z., Windberg, P., generated during a severe reactor accident: experimental results from the
Hummel, R., 2000. Investigation of aerosols released at high temperature from Power Burst Facility severe fuel damage test 1-4. Nucl. Technol. 105, 334–
nuclear reactor core models. J. Nucl. Mater. 282, 205–215. 345.
Dick, J.E., Nath, V.I., Kohn, E., Min, T.K., Prawirosoehardjo, S., 1990. Event sequence of Petti, D.A., Martinson, Z.R., Hobbins, R.R., Allison, C.M., Carlson, E.R., Hagrman, D.L.
a severe accident in a single-unit CANDU reactor. Nucl. Technol. 90 (2), 155–167. et al., 1989. Power burst facility (PBF) severe fuel damage test 1-4 test results
Dickinson, S., Baston, G.M., Sims, H.E., Funke, F., Cripps, R., Bruchertseifer, H., report. US NRC Report NUREG/CR-5163 (EGG-2542), April 1989.
Jäckel, B., Güntay, S., Glänneskog, H., Liljenzin, J-O., Cantrel, L., Kissane, M.P., Powers, D.A., Brockmann, J.E., Shiver, A.W., 1985. VANESA: a mechanistic model of
Krausmann, E., Rydl, A., 2003. Iodine chemistry and mitigation mechanisms: radionuclide release and aerosol generation during core debris interactions with
ICHEMM final synthesis report. European Commission report SAM-ICHEMM- concrete. US NRC Report NUREG/CR-4308.
D021, ftp://ftp.cordis.europa.eu/pub/documents r5/natdir0000048/s 2119005 Spencer, B.W., Thompson, D.H., Fink, J.K., Gunther, W.H., Sehgal, B.R., 1988. Results
20030910 104224 2119en.pdf. of fission product release from intermediate-scale MCCI tests. In: Proceedings
Fink, J.K., Thompson, D.H., Spencer, B.W., Sehgal, B.R., 1995. Aerosol and melt chem- of the International Conference on Thermal Reactor Safety, Avignon, France,
istry in the ACE molten core–concrete interaction experiments. High Temp. October 2–7, 1988.
Mater. Sci. 33 (1), 51–76. Strecker, J.J.F., Roth, P., 1994. Particle breakup in shock waves studied by single par-
Froeschke, S., Kohler, S., Weber, A.P., Kasper, G., 2003. Impact fragmentation of ticle light scattering. Part. Part. Syst. Char. 11 (3), 222–226.
nanoparticle agglomerates. J. Aerosol Sci. 34 (3), 275–287. Taylor, P., 1998. Calculations on the volatility of control-rod elements in various
Güntay, S., Suckow, D., Dehbi, A., Kapulla, R., 2004. ARTIST: introduction and first atmospheres at 1200 to 2500 K. IRSN Report (IPSN Note Technique SEMAR 98/
results. Nucl. Eng. Des. 231 (1), 109–120. 160).

Вам также может понравиться