You are on page 1of 12

Journal of Hazardous Materials 235236 (2012) 1728

Contents lists available at SciVerse ScienceDirect

Journal of Hazardous Materials


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

Review

A review on immobilization of phosphate containing high level nuclear wastes


within glass matrix Present status and future challenges
Pranesh Sengupta
Materials Science Division, Bhabha Atomic Research Centre, Mumbai 400 085, India

h i g h l i g h t s
 Technical review.
 High level nuclear waste immobilization within phosphate glasses.
 Integration of data from laboratory scale experiments, plant scale observations and natural rock information.

a r t i c l e

i n f o

Article history:
Received 10 April 2012
Received in revised form 12 July 2012
Accepted 18 July 2012
Available online 27 July 2012
Keywords:
High level nuclear waste
Immobilization
Phosphate glass
Review

a b s t r a c t
Immobilization of phosphate containing high level nuclear wastes within commonly used silicate glasses
is difcult due to restricted solubility of P2 O5 within such melts and its tendency to promote crystallization. The situation becomes more adverse when sulfate, chromate, etc. are also present within the waste.
To solve this problem waste developers have carried out signicant laboratory scale research works in
various phosphate based glass systems and successfully identied few formulations which apparently
look very promising as they are chemically durable, thermally stable and can be processed at moderate
temperatures. However, in the absence of required plant scale manufacturing experiences it is not possible to replace existing silicate based vitrication processes by the phosphate based ones. A review on
phosphate glass based wasteforms is presented here.
2012 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

4.
5.
6.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Generation of phosphate rich HLW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Bismuth process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
PUREX process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
P2 O5 -HLWs storage as neutralized solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
P2 O5 -HLWs storage as calcined powder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Phosphate in silicate melts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Lessons learnt from natural melts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Difculties associated with phase separations within HLW loaded silicate melts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Networking within phosphate melt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Immobilization of phosphate containing HLW within glass matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Future challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

E-mail address: praneshsengupta@gmail.com


0304-3894/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jhazmat.2012.07.039

18
20
20
20
20
21
21
21
21
22
23
25
26
26

18

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

1. Introduction
Eco-friendly and proliferation resistant management of spent
fuels is one of the most sensitive issues in the world today. Options
available at the back-end of nuclear fuel cycles are either
(i) direct disposal of spent fuel within suitable deep geological
repositories (better known as open fuel cycle), or
(ii) reprocessing it to extract valuables, followed by immobilizing
the process generated nuclear high level liquid wastes (HLWs)
within suitable inert matrice(s) for its interim storage and
disposal inside deep geological repositories (better known as
closed fuel cycle). Reaserch is also being pursued to segregate
and separately irradiate long lived minor actinides and ssion
products within critical and subcritical reactors to convert them
into shorter lived radionuclides (partitioning and transmutation; advanced fuel cycle; Fig. 1).
Recently, closed fuel cycle option has gained more importance
as it is environment friendly and economical [1]. In a simplied way, reprocessing begins with dismantling of water cooled
spent fuel bundles into pins and separating irradiated pellets from
Zircaloy (ZrSn alloys)/Al/stainless steel (FeCrNi alloys) clads
(protective metal jacket surrounding the fuel pellets) through
chopping (mostly by mechanical shearing) and dissolving the
small slices within concentrated acids (Fig. 2). The resultant solutions so obtained are highly hazardous as they contain various

radioisotopes (radioactivity more than few Ci/litre and posses


nearly 99% of the total radioactivity witnessed in a given nuclear
fuel cycle) of ssile and fertile materials, minor actinides, ssion
elements, activation products, etc. extracted from the spent fuels
(Table 1; [2]). Needless to say, the actual compositions (chemistry and radiochemistry) of HLWs depend on reprocessing route
adopted as well as on spent fuel compositions [37] and their irradiation history (type of reactors, pellet-cladding interactions [8],
neutron ux, burn up, cooling period, etc. [9]). Fig. 3 shows the four
different possible modes of occurrences of radionuclides within a
spent fuel. For example they can occur (i) within the irradiated
matrix fuel, (ii) along the grain boundaries, pores, cracks, dislocations, etc., (iii) diffused inside clad and (iv) within the gap in
between fuel and clad. With passage of time, relative contributions
of each of the radionuclides on the overall radioactivity of the spent
fuel changes due to differences in their respective half-lives (T1/2 ;
Table 2). Among all the radioisotopes present, those having T1/2 of
the order of years to decades and capable of getting incorporated
within tissues or organs (e.g. 90 Sr) are biologically most dangerous
ones. On the other hand, radionuclides having high radiotoxicity,
geochemical mobility, and long half-lives (e.g. 99 Tc, 129 I, 79 Se, 135 Cs,
239 Pu, 237 Np, 235 U, 36 Cl, and 14 C) are elements of deep concern from
environmental pollution point of view [10,11]. Therefore, for protection of biosphere, HLWs need to be concentrated and contained.
In general, HLWs are concentrated by evaporation and neutralized by addition of alkali (NaOH), and stored in large underground
steel tanks. Although such storage is acceptable for few years, but

Fig. 1. Schematic diagram showing basic differences between open and closed (advanced) nuclear fuel cycles.

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

Cooled
spent fuel

Shearing

19

U storage/
disposal

U oxide

U nitrate
solution

Chopping and
dissolution

Product
conversion

Reprocessing

Np, Am, Cm, remaining


fission products
Repository

Compaction/
Alloy formation

Vitrification

Pu nitrate
solution

MOX fuel
fabrication

Pu oxide

Product
conversion

Fig. 2. Schematic diagram showing major steps in the back-end part of closed nuclear fuel cycles.

on long time scale the wastes need to be immobilized within appropriate inert host matrices (wasteform), stored and disposed off
extremely carefully within suitable deep geological repositories so
as to isolate them from biosphere for as many as 104 106 years.
For conditioning of the HLWs, various amorphous (borosilicate
and aluminosilicate glasses), crystalline (synthetic rock SYNROC, titanate ceramics, phosphate ceramics (monazite, apatite and
related phases), calcines, alloys, etc.) and crypto-crystalline (glass
ceramics based on sphene, zirconolite, monazite, zircon, etc.) matrices [1224] have been proposed but the nal selection of wasteform
materials depends on several scientic and technological merits
and demerits associated with HLW compositions [25,26], product
durability factors, processing constraints [2733], service condition
demands [34,35], etc. Thus identication of a suitable wasteform
for any given HLW is a difcult task, and too much of expectations
in terms of its long term performance within geological repository
make the selection procedure even tougher. It is argued that the
environment within deep geological repository (constructed at a
depth of 5001000 m from the surface) is expected to be harsh due
to simultaneous interplay between (i) thermal eld, (ii) thermomechanical eld, (iii) biological eld, (iv) hydrological eld, (v)
chemical eld and (vi) radiation eld [36].

Of the various matrix-options available, sodium borosilicate


glasses are mostly favored as they offer wide compositional exibilities, high order of product durabilities and are easy to manufacture
through remote and robotic operations. Extensive scientic and
technological database already exist for the system and the glasses
are well accepted by general public. However in reality, borosilicate glasses cannot be considered as a universal wasteform matrix
as in many cases HLWs contain elements having poor solubility
within it [37,38]. A good example of this is phosphate rich HLWs
(P2 O5 -HLWs) which have only 23 wt% (P2 O5 ) solubility within
borosilicate melts [39]. The only possible way to condition such
HLWs within conventional borosilicate glass metrices is to dilute
them to a level so that the concentration(s) of the troubleshooting
element(s) become lower than their respective solubility limits. But
such an approach nally increases the volume of the vitried waste
products and makes the waste management procedure more costly.
An alternative approach to this is to nd out a novel amorphous
matrix having higher solubility for P2 O5 -HLWs. Toward this a number of efforts have been made across the world over the last fty
years, but so far limited progresses have been made to produce such
matrices in actual plant scale on regular basis. The probable reason
behind this is lack of complete comprehensions of the problem, its

Table 1
Radionuclides commonly found within spent nuclear fuel.
Major actinides

234

U, 235 U, 236 U, 237 U, 238 U


Th
236
Pu, 237 Pu, 238 Pu, 239 Pu, 240 Pu, 241 Pu, 242 Pu
232

Minor actinides

237

Np, 239 Np
Am, 242 Am, 243 Am
242
Cm, 243 Cm, 244 Cm, 245 Cm, 246 Cm
241

Fission materials

79

Activation products

Se, 85 Kr, 87 Rb, 89 Sr, 90 Sr, 93 Zr, 95 Zr, 95 Nb 99 Tc, 107 Pd, 115 In, 126 Sn, 129 I, 131 I 135 Cs, 137 Cs, 141 Ce, 142 Ce, 144 Ce, 144 Pr 144 Nd, 147 Sm, 147 Pm, 148 Sm,
149
Sm, 151 Sm, 133 Xe, 140 Ba, 134 Te, 93 Mo, 106 Ru, 106 Rh, 107 Pd, 140 La, 154 Eu

H, 10 Be, 14 C, 24 Na, 36 Cl, 39 Ar, 55 Fe, 59 Ni, 60 Co, 63 Ni, 93 Mo, 93m Nb, 94 Nb, 99 Tc, 108m Ag, 113m Cd, 121m Sn, 205 Pb, 210 Po

20

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

P2 O5 -HLWs were generated due to two different reprocessing


routes (a) bismuth process [41] and (b) PUREX process [40]. Subsequently for safe temporary storage, these wastes were either
stored as neutralized liquid or as calcined powder. Short descriptions on each of these items are given below. It may be added
here that besides the above mentioned two reprocessing routes,
currently various other innovative extraction processes involving
octyl(phenyl)-N,N-diisobutylcarbamoylmethyl phosphine oxide;
trialkyl phosphine oxides, etc. are being tried to reprocess spent
fuels more efciently. These may also yield P2 O5 -HLWs in future.
2.1. Bismuth process

Fig. 3. Schematic diagram (not to scale) showing elemental distribution within a


fuel-clad assembly after irradiation.

associated difculties and bottle neck situations. Hence for better


understanding of the subject and to take right steps towards feasible solutions it is absolutely necessary to put together the lessons
we have learnt from past experiments. So far no such attempt has
been made and the present article tries to ll up this lacuna. It
has compiled all relevant past experimental results, from 1950s till
date, in a single source so as to enable the researchers, the waste
immobilizers and the policy makers to judge the current state-ofthe-art. The review starts by summarizing the various reprocessing
techniques leading to P2 O5 -HLW generations, followed by the difculties encountered while immobilizing the same in borosilicate
matrices (also discusses observations made from natural rocks).
Subsequently a detailed note on the various wasteform matrices
tried in the laboratory scale is described and nally certain future
challenges associated with the problem are mentioned.
2. Generation of phosphate rich HLW
Spent fuel reprocessing was rst done in 1940s, and since
then various aqueous (wet; hydrometallurgical) and non-aqueous
(dry; pyrometallurgical) routes have been established or being
developed, keeping in mind the future requirements [40]. In past,
Table 2
T1/2 (half-lives) of selected radionuclides present within spent nuclear fuel.
Isotope

T1/2

Relatively short T1/2


89
Sr
54 days
103
Ru
40 days
131
I
8 days
140
Ba
13 days

Isotope

T1/2

Isotope

T1/2

95

65 days
57 min
8 days
40 h

95

Nb
Rh
134
Te
141
Ce

39 days
30 s
42 min
32 days

90

Sr
Ce
241
Am
60
Co

29
1.3
440
5.27

239

24,000
5730
76,000
20,300

Zr
Rh
133
Xe
140
La
103

Intermediate scale (year to decades) T1/2 (in years)


3
85
H
12.3
Kr
10
106
137
Ru
1
Cs
30
147
238
Pm
2.3
Pu
85.3
224
55
Cm
17.4
Fe
2.73
Longer (century and above) T1/2 (in years)
129
99
Tc
2 106
I
240
243
Pu
6500
Am
36
39
Cl
301,000
Ar
63
93
Ni
100
Mo

1.7 107
7300
269
4000

106

144

Pu
C
59
Ni
94
Nb
14

In this process, decladded spent fuels were dissolved within concentrated HNO3 solutions containing Binitrate and H3 PO4 so as
to precipitate Pu4+ together with Biphosphate. The precipitates
were then centrifuged and dissolved in HNO3 , to which KMnO4
was added to convert Pu4+ to PuO2 2+ and kept it in solution while
allowing Biphosphate to precipitate out. In subsequent stages,
Pu was precipitated using ferrous sulfate (FeSO4 ), ferrous sulfamate [Fe(SO3 NH2 )2 ], sulfamic acid (HSO3 NH2 ), etc. while ensuring
formation of uranyl sulfate [UO2 (SO4 )2 2 ] such that U did not
coprecipitate with Pu. The precipitates were separated from ssion
products and sulfate containing solution through centrifugation,
and were subjected to decontamination in order to reduce overall
-activity; treated with Nabismuthate, H3 PO4 , HNO3 , HF, lanthanum salts, NaOH, oxalic acid, H2 SO4 , (NH4 )2 SO4 , etc. to nally
obtain Pu in the form of solid PuLa oxide. From this mixed oxide,
metallic Pu was derived via Punitrate formation. P2 O5 -HLWs generated at different stages were transferred to underground storage
tanks [42].
2.2. PUREX process
Although Bi-process was efcient enough to extract Pu in pure
form but it could not extract useful U, and also the volume of waste
generated was huge due to repetition of precipitation cycles. This
led to adaptation of continuous PUREX (Plutonium Uranium Redox
Extraction) solvent extraction process in plant scale where tributyl
phosphate (TBP; saturated in hydrocarbon (e.g. kerosene, etc.)) was
employed to coextract Pu and U from HLWs generated after dissolution of oxide-, carbide-, nitride-based spent fuels [40]. In this
process, chopped fuel pins were dissolved in concentrated HNO3
solution from which U (UO2 2+ ) and Pu (Pu4+ ) were coextracted
and subsequently separated by reducing Pu4+ to Pu3+ (either by
ferrous sulfamate or by hydroxylamine) while keeping UO2 2+ in
solvent phase. P2 O5 -HLWs obtained at the end of the process were
transferred to underground tanks.
2.3. P2 O5 -HLWs storage as neutralized solutions
Storage of P2 O5 -HLWs within steel tanks for decades led to
various physicochemical changes including (i) thermal and radiolytic breakdown of organic/inorganic compounds, (ii) generation
of hydrogen and other gases, (iii) release of large amounts of
heat due to interaction between ferrocyanides and nitrates, etc.
To avoid unwanted corrosion of storage tanks, the acidic liquids are now being neutralized with NaOH. However, ageing of
these HLWs together with addition of fresh waste batches, and
occasional evaporation to adjust HLW concentrations and volume, lead to segregation of the mass in the form of supernate
liquids, saltcakes and insoluble sludges ([42], Fig. 4). During the
course of evaporation, soluble Na salts present within supernate
liquids (e.g. nitrate, nitrite, aluminate, hydroxide, etc. [43]) get
crystallized out in the form of salt cake, which is dominantly
NaNO3 mixed with Na7 F(PO4 )2 19H2 O, Na3 FSO4 , Na2 C2 O4 , NaF,

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

High level waste


storage tank

Supernate liquid
Na salts (nitrate, nitrite, aluminate, hydroxide)

Salt cakes
Na -nitrate + chloride, fluoride, phosphate, sulphate

Sludge
ejkaite, clarkeite, nitratine, goethite, maghemite, amorphous
solids (Na, Al, P, O and C), spinels, quartz, Na-Al silicates
Fig. 4. Schematic diagram (not to scale) showing possible modes of compositional
variations within high-level waste storage tanks.

Na3 AlF6 , etc. [44]. Radioactivity present within supernate solution and salt cake is essentially due to the presence of Cs, Tc and
limited amount of Sr and transuranic elements. The sludge, on
the other hand is much more hazardous as it retains most of the
radionuclides present within HLW (excepting Cs) and essentially

constituted of Cejkaite
(Na4 (UO2 )(CO3 )3 ). Apart from this, presence of clarkeite (Na[(UO2 )O(OH)](H2 O)01 ), nitratine (NaNO3 ),
goethite [-FeO(OH)], magnetite (-Fe2 O3 ), amorphous solids (Na,
Al, P, O and C), (Fe, Cr, and Ni) oxides, SiO2 , Na Al silicates,
etc. have also been reported [45]. Thus, supernate liquid and salt
cakes together constitute low activity waste (LAW) and whereas
sludge constitute high level waste (HLW). The radioactive components present within LAW are recovered through isotopic dilution,
transuranic precipitation, ultraltration and ion-exchange process
and are added to HLW fraction [46].
2.4. P2 O5 -HLWs storage as calcined powder
Apart from storing as liquid, P2 O5 -HLWs are also stored as calcined powder. Reprocessing of Al/Zircaloy/stainless steel/graphite
cladded spent fuels following PUREX process results in generations of (a) F-rich and (b) Al-rich wastes [47]. To these (i) Aland Ca-nitrates are added for uoride complexations as well as
volatility suppression, and (ii) H3 BO3 is put to prevent insoluble Al2 O3 and -Al2 O3 formation. For volume reduction, P2 O5 -HLWs
are sprayed into hot (400600 C) uidized bed whereupon anions
such as nitrates, carbonates, etc. decompose leading to formation
and growth of multi-oxide calcined particles, which are then piped
pneumatically into underground storage tanks. The nal composition of calcined wastes therefore depends on initial P2 O5 -HLW
compositions and subsequent chemicals added during powder formations. For example, Al-calcines are usually rich in Al2 O3 and
B2 O3 , while Zr-calcines are enriched in CaF2 , ZrO2 , Al2 O3 , CaO
and B2 O3 [48]. Similar calcination route has also being considered for treating NaNO3 rich waste streams (commonly referred as
sodium bearing waste, SBW) which mostly arise from site clean-up
(decontamination and decommission) activities [47]. Compositionally SBW is a HNO3 solution with relatively high concentrations of
NO3 , Na+ , Al3+ , K+ , P2 O5 , and SO4 2 [49].
3. Phosphate in silicate melts
3.1. Lessons learnt from natural melts
As mentioned earlier, borosilicate glasses are generally preferred for HLW immobilization. However, natural analogue studies

21

backed up by experimental observations on equivalent synthetic melts show that P2 O5 even if present in small quantities
(12 wt%) can (i) promote liquid immiscibility (basic and acidic
melts [50,51]), (ii) control trace element partitioning within melts
[52,53], (iii) reduce melt viscosity [54,55], (iv) shift liquidus
boundaries towards silica decient domains within variably polymerized silicate melts [56] and (vi) depress solidus [57]. Raman
investigations in SiO2 P2 O5 system show the presence of isolated non-polymerizedP2 O5 (basic) rich domains connected to
polymerizedSiO2 (acidic) melts through formation of P O Si linkages [58]. In case of aluminosilicate system, the situation changes
as juxtaposition of Al3+ and P5+ promote quartz (SiO2 )berlinite
(AlPO4 ) substitution:
Al3+ + P5+ = 2Si4+

(1)

within melts. Bulk compositional analyses of silica rocks show


phosphate solubility within multicomponent aluminosilicate
melts depends on A/CNK (Al2 O3 /(CaO + Na2 O+ K2 O); [59]) or ASI
(Al/(Li + Na + K + Rb + 0.5Ca); [60]) molar ratios and are usually
higher in peraluminous (A/CNK > 1) melts in comparison to metaluminous (A/CNK = 1) ones. Wolf and London [61] estimated
that meta-aluminous haplogranite melts (ASI = 1.0) can dissolve
0.1 wt% P2 O5 at 750 C whereas in similar condition peraluminous melts (ASI = 1.35) can incorporate 0.7 wt% P2 O5 . Thus as far as
wasteform designing is concerned peraluminous melts have more
promise than others. The reason behind such variations in phosphate solubility is linked to its respective melt structures. Within
peralkaline melts (A/CNK < 1) phosphate addition increases its
polymerization through associating alkali metals with phosphate
tetrahedra and thereby decreasing the number of NBO [6264].
On the other hand, phosphate in peraluminous melts stabilizes
berlinite component and thereby enhances apatite solubility [59].
This scenario however becomes complicated when uxing elements (e.g. F and B), alkali (Li, Rb, and Cs) and alkaline earths
(Be, Sr, and Ba) are present within the system. It has been conrmed from both petrological as well as experimental studies
that these elements play contrasting roles in terms of promoting crystallization. For example, formation of F Al clusters within
silicate melts decreases berlinite (AlPO4 ) activity [65] whereas
reverse happens with development of BAl complexes (due to
competition for tetrahedral positions). Similarly, high Li content
together with virgilite (LiAlSi5 O12 ) substitution are known to stabilize berlinite solid solutions through enhancing the AlPO4 SiO2
mutual solubility. In addition to these, crystallization of apatite,
berlinite, amblygonitemontebrasite, lithiophilitetriphylite, spodumene, petalite, etc. instead of potential silicate minerals within
acidic melts containing phosphate is not uncommon.
3.2. Difculties associated with phase separations within HLW
loaded silicate melts
As observed from natural analogues and equivalent synthetic
melt experiments, phosphate is found to promote phase separations (liquid immiscibility as well as crystallization) within waste
glass melts even if present in minor amounts [66]. Such phase separations are not coveted from HLW immobilization point of view as
it deteriorates product durability factors severely and also hampers
plant scale manufacturing processes signicantly. The situation
becomes even more stringent when additional poorly soluble (in
silicate melt) anions like sulfates, chromate, halides, etc. are also
present within the HLW [31,67]. In case of P2 O5 -HLWs, presence of
sulfate anion within the waste is very common as reducing agents
such as ferrous sulfate, and ferrous sulfamates are used for Pu cation
reduction. It has been noted that due to poor solubility of sulfate
anion within silicate melts (<1 wt%), salt layers (popularly known
as yellow phase or gall) in the form of foam and froth separate

22

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

Fig. 5. Formation of yellow phase within borosilicate glass matrix.

out (formed essentially due to release of oxygen upon reduction


of multivalent cations like Fe, Ni, Mn, etc.) during vitrication of
HLW [38,6872]. The yellow phase is essentially a cluster of sulfates, chromates and aluminates of which water soluble thenardite
(Na2 SO4 ) is the most prominent one and acts as a sink for hazardous radionuclides ( emitting, heat generating, radiotoxic)
such as 137 Cs and 90 Sr. Fig. 5 (hand specimen scale) shows the
development of yellow phase within borosilicate melt upon excess
amount of sulfate containing waste loading. Apart from this, yellow
phase formations also severely impede vitrication process as they
(i) do not allow the release of gaseous reaction products from melt,
(ii) corrode melter liners, (iii) diverts electrical eld and can even
short electrodes, (iv) go on accumulating and occupying signicant
portion of vitrication furnace unless physically separated out, etc.
[3032,73]. Adaptation of high temperature vitrication process is
also not a solution for such type of problems as the decomposition
temperatures for salt layers are usually very high (above 1300 C).
Hence the best solutions for such cases is to go for an alternative
matrix where it is possible to maintain the liquidus temperature of
the melt at least 100 C below than the plant operating temperatures (1100 C) and the melt should not undergo phase separation
during production and even under prolonged idling conditions.
Keeping all these factors in mind a group of wasteform designers
opted for P2 O5 glasses (instead of silicate glasses) as an alternative
matrix since previous structural studies revealed that the system
offer wider range of network structures suitable for radionuclide
entrapment.
4. Networking within phosphate melt
Long back Zachariasen [74] and van Wazer [75] identied
(PO4 )3 tetrahedra as the basic unit of P2 O5 glasses which upon
linking through covalent P O P bonds and others were found to
dene various types of network structures (Table 3 and Fig. 6). Each
of the tetrahedra, results through formation of sp3 hybrid orbitals
by the outer electron (3s2 3p3 ). Usually, the fth outer electron of
P gets promoted to 3d orbital where together with 2p electron of
O forms strong -bond. Within each (PO4 )3 tetrahedra, the P ion
is linked to three bridging oxygens (OB ) through P OB bonds and
to one terminal oxygen (OT ) by doubly bonded P OT bonds. For
any given glass, the P OT bonds are always shorter than P OB
bonds and their respective lengths depend on P2 O5 contents and
nature of the modifying cations [76]. Cross linking of three out of

Fig. 6. Phosphate structural units reported from glasses (Qi : i represents number of
bridging oxygen per tetrahedron).

the four vertices of (PO4 )3 tetrahedra leads to the formation of


Q3 ultraphosphate structural group, which upon additions of modifying oxides (e.g. alkali oxides (R2 O), alkaline earth oxides (R O),
etc.), get depolymerized through conversion of bridging oxygens
(P O P) to non-bridging oxygens (P O R) [77]. Thus in case of
binary xR2 O (or R O)(1 x)P2 O5 glasses (where x is in mole fractions), with increasing molar R2 O/P2 O5 ratios (from 0 to 1 to 2 to
nally 3) or [O]/[P] ratios (from 2.5 to 3 to nally >3.5), the structural groups (represented by Qi where i represents the number
of bridging oxygens per tetrahedron) passes from cross linked Q3
(ultraphosphate; O/P = 2.5) to Q2 (metaphosphate; O/P = 3.0) to Q1
(pyrophosphate; O/P = 3.5) to Q0 (orthophosphate; O/P = 4.0), keeping P always in four fold coordination [78,79]. Recently it has been
noted that depending on the chemical nature, the cations present
within a given phosphate glass can act either as network modier
(e.g. Na+ and Ca2+ ) or network former (e.g. Al3+ , Bi3+ and Nb5+ ; [80]
and their coordination numbers may change with compositional

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

23

Table 3
Compositional dependence of phosphate structure binary in binary xR2 O (or R O)(1 x)P2 O5 (x in mol%) system.
Structure

Ultraphosphate

Mesophosphate

Polyphosphate

Pyrophosphate

Orthophosphate

Compositional
range
Dominating
structural groups
Proportion of
structural groups

0 < x < 0.5

x = 0.50

x > 0.50

x = 0.67

x = 0.75

Q2 and Q3

Q2 chains terminated
Q2 form chains
and rings
by Q1 tetrahedra
f(Q1 ) = ((2x 1)/(1 x))
and f(Q2 ) = ((2 3x)/(1 x))

Phosphate dimmers, a pair of


linked Q1 tetrahedra
f(Q0 ) = ((3x 2)/(1 x)) and
f(Q1 ) = ((3 4x)/(1 x))

Isolated Q0

3.5

>3.5

Molar O/P ratio

f(Q2 ) = (x/(1 x))


f(Q3 ) = ((1 2x)/
(1 x))
2.5

>3

modications [81]. Thus, in principle a variety of glasses can be


prepared within phosphate based systems, some of which do have
various advantageous properties suitable for plant scale HLW vitrication such as (i) low melting temperatures; (ii) fast melting, (iii)
low viscosity, and (iv) ability to dissolve wide range of elements.
However from conventional melt processing point of view, glass
formations is typically limited from vitreous P2 O5 to compositions
between the metaphosphate and pyrophosphate compositions
(i.e. in case of binary xR2 O (or RO) (1 x)P2 O5 systems within
0 < x < 0.550.60). Keeping these aspects in mind various attempts
have been made to condition P2 O5 -HLWs within phosphate glasses
and a brief description on this is given below.
5. Immobilization of phosphate containing HLW within
glass matrices
Recognizing the various advantageous features of P2 O5 -based
glasses, attempts were made (in 1960s) to condition Na rich HLWs
within Naphosphate glasses. These wasteforms however suffered
from poor chemical durability, rapid devitrication and high corrosiveness in molten stage which made them less attractive within
wasteform designer community [8285]. The poor chemical durability of the waste glasses was attributed to dominance of easily
hydrolysable P O P bonds, which upon hydration get depolymerized through formation of terminal OH groups [86]. Bunker
et al. [87] studied the interactions between different hydrogen
species, (H2 O, OH and H+ or H3 O+ ) and phosphatic glasses and concluded that metaphosphate glasses dissolve congruently through a
two stage process involving (a) surface hydration (due to water
diffusion) and (b) subsequent release of metaphosphate chains
in solution (due to surface reactions). Subsequent experimental

studies showed that addition of Al2 O3 improves chemical durability


of phosphate glasses [88] and this re-established the credibility of the system. In fact, former Soviet Union came up with a
new Na Al phosphate glass formulation for immobilizing its sulfate containing HLWs [89]. 27 Al-(MQ) MAS NMR spectroscopic
analysis of aluminophosphate glasses showed that replacement
of P O P bonds by Al O P linkages is principally responsible
for the improved chemical durability of the matrices [90]. The
idea of employing phosphate glasses for waste immobilization
was once again tried by United States, and this time it yielded
good results in terms of conditioning Na and sulfate rich mixed
waste from Solar Evaporation Basin of Hanford ([91]; Table 4).
Parallel to this, Kushinikov et al. [92] and Matyunin and Jardine [93] tried to immobilize Pu rich waste, both in the form
of PuO2 powder and Pu nitrate solution, within Na Al phosphate glass (Na2 O: 24 wt%, Al2 O3 : 22 wt% and P2 O5 : 54 wt%)
and respectively could load 0.240.31 wt% (for PuO2 powder
waste) and 0.70.8 wt% (for Pu nitrate waste) of waste homogeneously. Lately, Donald et al. [94] considered Na Al phosphate
glasses (41.0Na2 O20.5Al2 O3 38.5P2 O5 (mol%)) for encapsulating
Ca phosphate ceramic (containing chlorapatite (Ca5 (PO4 )3 Cl) and
spodiosite (Ca2 (PO4 )Cl)) loaded with chloride enriched HLWs [95].
They studied the devitrication tendency of the encapsulating glass
(AlPO4 , Na3 PO4 , NaAlP2 O7 and Na7 (AlP2 O7 )4 PO4 ) and hardly found
any detrimental effect of the same on the overall properties of the
nal monolithic wasteform. Further, to improve thermal stability of the material and suppress devitrication, several trials were
taken with FePO4 (as a source of Fe2 O3 ), ZnO and B2 O3 , of which
the later (>2 mol% B2 O3 ) yielded the best result [94,96]. The reason
behind such improvement is obviously better linkages of (PO4 )3
tetrahedra with trigonal (B3 ) and tetrahedral (B4 ) borate units [97].

Table 4
Compositions (wt%) of some well studied simulated waste loaded phosphate glasses.
Oxide

Na2 O
Al2 O3
P2 O5
B2 O3
BaO
CaO
CeO2
Cr2 O3
CuO
Fe2 O3
K2 O
MnO
PbO
SO3
SiO2
ZrO2
Cl
F
a

Hanford mixed waste [91]

Hanford LAW [130]

SPP22 Pb

SPP23 Pb

IP30LAW

IP30LAW-A

IP30LAW-C

IP40WG

IP40WG-CCIM

21.17
18.39
35.55
4.49
0.01
5.25
0.21
0.04
4.17
.51
0.94
0.43
0.67
1.62
3.79
2.73
0.0
0.0

17.36
21.08
42.34
0.01
5.61
0.20
0.07
4.33
0.52
0.78
0.45
0.53

1.3
52.2
0.0

0.0

0.1

20.0
0.0
0.0

2.9
0.2

0.2
0.5

11.3
52.2
0.0

0.0

0.1

10.0
0.0
0.0

2.9
0.2

0.2
0.5

1.3
52.2
0.0

0.0

3.1

17.0
0.0
0.0

2.9
0.2

0.2
0.5

10.9
51.7
0.1

0.9

0.1

10.3
3.0
0.3

1.4
0.0

0.3
0.3

11.6
50.0
0.0

0.9

10.6
3.2
0.0

1.8
0.0

0.0

Spiked with SO3 : 217 wt%.

4.32
2.35
0.0
0.0

INEEL-SBW [131]

24

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

But once again, the encapsulation suffered from hydrolyzation of


the network as borate units are also found to be hygroscopic as
phosphate units are [98].
Parallel to the studies on Na Al phosphate glasses, Sales and
Boatner [99] came up with a new formulation from Pb Fe phosphate glass system to immobilize chloride rich HLWs. Pb and Fe
were purposefully added to the phosphate glass system to decrease
its melting temperature and viscosity, and to increase its chemical durability and resistance towards devitrication [100,101].
Experimental studies indicated that addition of PbO should be
restricted within 3566 wt% as beyond this range phosphate glasses
are either prone to devitrication or are highly viscous in molten
stage. Hence Sales and Boatner [99] selected an intermediate composition of 45 wt% PbO55 wt% P2 O5 and added 9 wt% Fe2 O3
(40PbO10Fe2 O3 50P2 O5 (mol%)) to further improve its resistance
towards corrosion and devitrication, and to increase its softening
temperature. Thus they could ultimately formulate a pyrophosphate glass (Fe2 Pb(P2 O7 )2 and Fe3 (P2 O7 )2 ; O/P molar ratio 3.7)
capable of incorporating 1015 wt% of simulated waste and still
having a leach resistance better than borosilicate waste glasses
[102104]. Mssbauer [105] and electron paramagnetic resonance
(EPR; [106,107]) studies indicated that Fe2+ cations are mostly
restricted within interstitial positions with octahedral coordination
whereas Fe3+ cations can occupy both substitutional and interstitial positions with octahedral coordination. The improved chemical
durability is due to replacement of hygroscopic P O P linkages
by P O Fe bonds [108]. With increasing waste limit, the overall pyrophosphate structure (P2 O7 )4 was gradually replaced by
orthophosphate (PO4 )3 and metaphosphate (PO3 ) species, and
NaFeP2 O7 and SiP2 O7 crystallized within the matrix adversely
affecting its chemical durability.
Although PbFe phosphate glass exhibited high potentiality
towards waste immobilization but still it suffered from (i) low
waste loading capacity (<20 wt%), (ii) not adequate solubility
for alumina, silica, zirconia, uranium sulfates, chloride, uoride,
chromium oxide, heavy metals, etc., (iii) incompatibility with zeolite, (iv) highly corrosive towards ceramic melter components
(especially for Inconel 690 electrodes and refractories), heterogeneity and (v) poor chemical durability when devitried (especially in
NaCl solutions) [108]. However, further research in the eld led to
the development of Fe phosphate glasses (40Fe2 O3 60P2 O5 mol%)
having (i) better chemical durability, (ii) higher solubility for heavy
metals, noble metals, and rare earths; (iii) low melting temperatures (9501100 C), rapid melting rates (few hours), high melt
uidity (viscosity typically below one poise), (iv) higher waste loading capacity (2550 wt% depending on the waste), and higher
density (3.03.4 g/cm3 ), which combined to minimize the volume
of vitried waste (Tables 4 and 5; [84,108113]. It was observed
that glass formation region within Fe phosphate multicomponent

system is a function of the size of monovalent/bivalent modier


cations and increases through the series of Mg Ca < Pb < Ba and
Li < Na < K [114].
Structurally, Fe phosphate glasses are constituted of (Fe3 O12 )16
clusters connected via pyrophosphate groups, i.e. (P2 O7 )4 [115].
When prepared in air, Fe-phosphate glasses always contain both
Fe2+ and Fe3+ irrespective of the Fe-oxide precursor (FeO or Fe2 O3 ).
The Fe2+ /Fe3+ ratio depends on many factor including melting temperature, process, time, furnace atmosphere, raw materials, etc.
[116,117]. For example the Fe2+ /Fetotal fraction in 40Fe2 O3 60P2 O5
(mol%) glass varies from 17% to 50% as the temperature increases
from 1150 to 1200 C [118]. A careful analysis of the Fe phosphate
glasses shows that only a minority, 1626% of the oxygens form
hygroscopic P O P bond, whereas others are bonded via hydration
resistant Fe2+ O P and Fe3+ O P [112]. Dependence of chemical durability on O/P molar ratio shows that maximum durable
Fe-phosphate glasses can be obtained when O/P ratio is 3.63.7
(pyrophosphate glasses); on either side of this range chemical durability tends to decrease. Further, Fe3 (P2 O7 )2 type structure shows
that it contains interstitial voids large enough to accommodate ions
as large as U4+ and Cs+ [119]. The ionic crystal radii (for coordination number 6; in ) of some important HLW constituents within
Fe-phosphate glasses are P5+ : 0.35, O2 : 1.40, Si4+ : 0.40, Fe2+ : 0.77,
Fe3+ : 0.65, U4+ : 0.890.97, U5+ : 0.79, Al3+ : 0.53 and Na+ : 1.02. This
is why the Fe-phosphate glass networks remain largely insensitive to the additions of P2 O5 -HLW cations and can solubilize them
in high quantity without sacricing much on its chemical durability aspects [119]. Previously, Day et al. [110] could load simulated
waste of 32 wt% (against 10 wt% in borosilicate glass) for the high
sulfate LAW (Hanford), to 40 wt% (against 20 wt% in borosilicate
glass) for the SBW (INEEL), to 75 wt% (against 15 wt% in borosilicate glass) for the high chrome HLW (Hanford). The better solubility
of Cr within Fe phosphate melts is due to its occurrence in +3 state
only (instead of Cr3+ , Cr4+ , Cr6+ as found in silicate melts; [37])
which makes it possible to incorporate the cation through break
c et al. [121] investigated the
ing P O double bonds [120]. Santi
effect of Cr2 O3 additions on xCr2 O3 (40 x)Fe2 O3 60P2 O5 (where
0 < x < 10, x is in mol%) pyrophosphate (Q1 , O/P molar ratio 3.5) and
noted that replacement of Fe2 O3 by Cr2 O3 up to 5 mol% does not
produce any signicant structural changes, however higher loading
(e.g. 10 mol%) promote crystallization of chromium orthophosphate (-CrPO4 ) and iron pyrophosphate (Fe3 (P2 O7 )2 ) within the
matrix. To simulate the partitioning of actinides, Zhang et al. [122]
and Lopez et al. [123] assessed the HfO2 and Gd2 O3 solubility limits
using X-ray diffraction and Mssbauer spectroscopic analyses and
noted that additions up to 5 mol% have hardly any effect in the
Fe-phosphate melt structures.
This is however not the case when alkali (R2 O) and alkaline
earth (RO) elements are concerned. Bingham and Hand [114] noted

Table 5
Some thermal durability parameters for simulated waste loaded Fe phosphate glasses.
Wasteform

Density
(g/cm3 )

Glass transition
( C)

Crystal
formation ( C)

IP30LAW
IP30LAW-A
IP30LAW-C

2.88
2.77
2.871

483
453
487

572
561
605

IP40WG
IP40WG-CCIM

2.76
2.76

430
420

CAB1
CAB2
DR1
DR2
PFeOx
PFeUOx

500
496
497
506
515
535

Liquidus ( C)

Thermal expansion coefcient


(107 / C1 )

Reference

168
165
163

[130]

590
560

740
718

[131]

67.1
78.2
65.3
52.4
90.4
84.7

[119]

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

that although additions of both (R2 O and RO) in multicomponent


P2 O5 Fe2 O3 FeORy O2 (R = Li, Na, K, Mg, Ca, Ba, or Pb and y = 1 or 2)
systems induces depolymerization, R2 O hardly modies network
structures, thermal properties and redox state whereas RO does.
It has been noted that up to 26 mol% of CsCl, 31 mol% SrF2 , and
34 mol% CsCl + SrF2 could be loaded within Fe-phosphate glasses
without much sacrice on product durability aspects [124,125],
whereas for higher loading (e.g. Cs2 O of 55 mol%, [126]) the
matrix becomes moisture sensitive and the glass-transition temperature reduces drastically. In case of chemical durability, R2 O
additions reduce it following the series K > Na > Li whereas presence of RO enhances it. Bulk crystallizations of Fe-phosphate
glasses are also observed to take place at much lower temperatures
(600 C) in presence of R2 O (3040 mol%) than when RO is present
(700 C). For example, in case of Cs2 O doped Fe-phosphate glasses
(36 mol% Cs2 O26 mol% Fe2 O3 P2 O5 ) CsFe(P2 O7 ), Cs7 Fe7 (PO4 )8 O2
and Cs3 PO4 are found to crystallize out at 583 C [127].
Phosphate glasses have long been known for better solubility
of sulfate [128] and sulfate containing HLW [68,129]. However
when tried for sulfate and phosphate containing mixed HLW
(17 wt% sulfate; Hanford mixed waste) it is observed that retention of sulfate within otherwise homogeneous Na2 OAl2 O3 P2 O5
glasses (loaded with 38 wt% of simulated HLW, 1050 C)
are rather poor [91]. To understand the reason, Bingham
and Hand [114] carried out in-depth experimental studies on
[(1 x)(0.6P2 O5 0.4Fe2 O3 )] + xRy SO4 (where R = Li, Na, K, Mg, Ca,
Ba, Pb; x = 00.5; y = 2 and 1) and P2 O5 Al2 O3 Na2 OFe2 O3 SO3
glasses and noted that sulfate solubility (log[SO3 ]) is linearly
related to normalized cation eld strength index (z/a2 ), theoretical optical basicity th , and [O]/[P] molar ratio. It has
been postulated that under general plant scale operating conditions (1150 C), Fe-phosphate glasses for which (z/a2 ) < 1.2;
th > 0.5; [O]/[P] > 3.8; or P2 O5 content <40 mol% should be
able to incorporate more SO4 2 than it usually retains (<0.5 mol%).
In continuation to these studies, Kim and Day [130] and Kim et al.
[131] could successfully load 30 wt% of simulated LAW and 40 wt%
of simulated SBW within Fe-phosphate glass respectively, and satised all necessary product consistency and vapor hydration test
requirements [132,133] even when deliberately crystallized (containing Fe3 (P2 O7 )2 , Fe4 (P2 O7 )3 , BiPO4 , Na3 Fe2 (PO4 )3 , Ca5 F(PO4 )3 ,
Al(PO)4 and Al2 O3 ) [110,119]. Surcial layers developed on the
waste glasses were found to consist of NaFe3 P2 O7 (OH)3 H2 O,
Na3 Fe(PO4 )2 H2 O, NaFeH2 (PO4 )2 and Na3 Al(PO4 )2 1.5H2 O, which
according to Ewing and Wang [21] depends on the metal (M)-tophosphorous (P) ratio of the glass. A high M/P ratio suppresses
corrosion, whereas when it is low the corrosion products do not
build up on the surface leading to release of H3 PO3 in adjacent
solution and thereby lowering the pH and accelerating the direct
hydrolysis of the P O P bonds on the glass surface [134].
Although good amount of studies have been carried out
on compositionstructure correlation of phosphate glasses but
hardly any investigation has been done on radiation damage
related aspects. Sun et al. [135] noted that Fe-phosphate (in
mol%, 45% Fe2 O3 55% P2 O5 , 20% Fe2 O3 80% P2 O5 , and 20%
Fe2 O3 20%Na2 O60% P2 O5 ) and aluminophosphate glasses (in
mol%, 44.5% P2 O5 31.5% Al2 O3 20.2% Na2 O3.8% K2 O) get decomposed under 200 kV electron irradiation (at doses higher than
1.0 1026 e m2 ). Alkali elements, if present, are found to migrate
from the irradiated centres to the peripheries and thereby making
the glasses compositionally heterogeneous. Formation of bubbles
within irradiated Fe-phosphate and aluminophosphate glasses has
also been reported. In a separate study Ezz-Eldin [136] showed that
-irradiation can signicantly change various physical properties
of phosphate glasses.
Once again, not much work has been done on plant scale processing related aspects also. The estimated parameters are mostly

25

obtained from laboratory based experiments rather than pilot scale


or plant scale studies. It has been noted that homogeneous and
easily pourable phosphate melts, loaded with waste, can be produced between 950 and 1050 C using conventional vitrication
furnaces with short melting times (usually <4 h) and less tendency
towards insulating foam formation [130,131]. This is largely due
to lower activation energy (Q) for viscous ow (35.367.0 kJ/mol;
[130,131]) of Fe-phosphate waste melt as compared to those for
SiO2 (514.9 kJ/mol 16002500 C; 237 kJ/mol according to Ojovan
et al. [137]) and P2 O5 (173.7 kJ/mol 545655 C) melts. In terms of
viscosity at melting point (1000 C), Fe-phosphate waste melts are
much more uid 200400 centipoises for LAW waste glasses [130]
and 500900 centipoises for SBW glasses [131] compared to the
borosilicate melts (45009000 centipoises). Similarly activation
energy (Q) of electrical conductivity is also found to be much less for
Fe-phosphate melts (20 kJ/mol, [130]) which assigns reseaonably
good electrical conductivity (for LAW glasses: 70 S/m at 800 C to
135 S/m at 1200 C, [130]; for SBW glasses: 40 S/m at 900 C to
160 S/m at 1300 C [131]) within the melt probably through hoping
of electrons from Fe2+ to Fe3+ ions [131].

6. Future challenges
As indicated earlier, restricted solubility of P2 O5 -HLWs (+sulfate) within silicate melts has encouraged wasteform designers
once again to judge the potentiality of P2 O5 based glass systems
as host matrices. A good amount of laboratory based experimental studies on compositionpropertystructure correlations in
various phosphate based glasses and waste glasses have already
been done and the results have identied few formulations within
alumino phosphate and Fe-phosphate systems. Some of these phosphate glasses look very promising as they are chemically durable,
thermally stable and can be processed at moderate temperatures. However, general tendency of phosphate glasses towards
devitrication often considered as a serious shortcoming. Experimental studies dealing with the effect of devitrication on chemical
durability show contrasting results which make it all the more
important to study the dependence of various product durability
parameters (waste loading, density, glass transition temperature,
melting temperature, Youngs modulus, Poisson ratio, Vickerss
hardness, fracture toughness, thermal expansion, thermal conductivity, specic heat, etc.) on crystallization. To reduce the
investigation period substantially it will be a good idea to construct
models capable of predicting crystal nucleation and growth, crystal
compositions and melting behavior for a given formulation, taking
clues from laboratory scale tests. Such models however need to be
calibrated and veried with experimental data obtained from pilot
scale melter trails. While doing this exercise, due considerations
should also be given to the various inuences of radiation damage
processes (with varying doses over different temperature regimes)
on modication of product durability parameters, an aspect so far
neglected in the context of phosphate glass based wasteform.
Another eld where signicant thrust needs to be given is
relevant data acquisition and experience gathering from plant
scale level synthesis of phosphate waste glass matrices. Gaining
such experiences prior to actual plant scale production is necessary so as to avoid severe consequences of faulty manufacturing
operations and/or unsatisfactory wasteform performance. During
laboratory scale experiments with simulated HLWs it is not possible to encounter or anticipate all possible issues that one may
has to face during hot operations. One good example for this is
rapid corrosion and pre-mature failures of furnace components,
especially Alloy 690 components (a variety of NiCrFe superalloy), refractory bricks, etc. during plant scale vitrication of HLWs
over long time scale. Such melt and/or process induced corrosions

26

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

is known to be much more aggressive in case of phosphatic melts


as compared to silicate ones. As a solution towards this, one may
suggest to modify phosphate melt compositions by adding Al2 O3 ,
B2 O3 , etc. [138,139] or by putting chemical diffusion barriers on the
alloy components [140,141], but the actual progress in this direction can only be done after getting feed back from plant operators.
Due to lack of such data or experiences it is also not possible to
decide various important processing parameters such as type of
furnace need to be used, nature of melter feed (slurry or calcined
mass), requirement of external agitation (using stirrer, bubbler,
etc.) for homogenization of the melt, etc. Although judging from
experimentally determined viscosity, electrical conductivity data
one can suggest to adapt induction furnace based technologies
but it will not be prudent to decide on this without having done
required plant scale testing. Such an approach will also help in
establishing exibility within process ow sheets as variations in
HLWs and/or melter feed chemistry and various processing parameters can impact melting rate and steady state melting conditions
signicantly.
All said and done, another big challenge in nal adaptation
of phosphate melt based vitrication technique in plant scale is
mental blockage. Over the decades huge investments of time and
resources have been made to establish borosilicate glasses as a
suitable waste form matrix. Thus waste immobilizers are reluctant to adopt any new vitrication technique in plant scale level
until and unless all routes of modifying borosilicate glass matrices are tried. On one hand the possibility of modifying existing
borosilicate glass compositions with better compositional exibility, durability and processing characteristics cannot be ruled
out but on the other hand it is also equally possible to identify
an alternative matrix having wider processing windows, greater
tolerance towards waste stream compositional uctuations and
yet having acceptable product durability properties. In such cases
of course one has to take decision after giving proper weightage
to the overall costs of developing a new wasteform along with
duly required acquisition of repository performance data, licensing
amendments, production change-over, revision of waste acceptance requirements, identication of new consistency/durability
test specications, rescheduling storage, disposal plans, etc. But the
data and information required to do such appraisals is still lacking
and until and unless these are made available the preferred choice
will be to stick to already established vitrication technique. Thus
to establish the credentials of phosphate glass based vitrication
process in plant scale it is important to assure all required information and experiences on a faster track for the already identied bulk
formulations and the newer ones [142,143]. Additionally, in case
of waste streams containing signicant amount of low-solubility
components, due attentions should be given to glass-composite
materials also [144,145]. Finally, it is hoped that with such an
approach, practicing professionals and interested students will be
able to appreciate the intricacies of the subject and come forward
to solve the problem in future.

Acknowledgements
Author gratefully acknowledges Alexander von Humboldt Foundation, Germany. Prof. Andrew J. Daugulis is thanked for carrying
out editorial responsibility very effectively.

References
[1] A. Dahhan, H. Muthanna, Radioisotopes applications in industry: an overview,
Atom Peace 2 (2009) 324337.
[2] Y. Ando, K. Nishihara, H. Takano, Estimation of spent fuel compositions from
light water reactors, J. Nucl. Sci. Technol. 37 (2000) 924933.

[3] V. Grover, P. Sengupta, K. Bhanumurthy, A.K. Tyagi, Electron probe microanalysis (EPMA) investigations in the CeO2 ThO2 ZrO2 system, J. Nucl. Mater. 350
(2006) 169172.
[4] V. Grover, P. Sengupta, A.K. Tyagi, Sub-solidus phase relations in CeO2 YSZ
and ThO2 YSZ systems: XRD, high-temperature XRD and EPMA studies,
Mater. Sci. Eng. B 138 (2007) 246250.
[5] V. Grover, A. Banerji, P. Sengupta, A.K. Tyagi, Raman, XRD and microscopic
investigations on CeO2 Lu2 O3 and CeO2 Sc2 O3 systems: a sub-solidus phase
evolution study, J. Solid State Chem. 181 (2008) 19301935.
[6] T.R.G. Kutty, R.V. Kulkarni, P. Sengupta, K.B. Khan, K. Bhanumurthy, A.K. Sengupta, J.P. Panakkal, A. Kumar, H.S. Kamath, Development of CAP process for
fabrication of ThO2 UO2 fuels. Part II: Characterization and property evaluation, J. Nucl. Mater. 373 (2008) 309318.
[7] T.R.G. Kutty, M.R. Nair, P. Sengupta, U. Basak, A. Kumar, H.S. Kamath, Characterization of (ThU)O2 fuel pellets made by impregnation technique, J. Nucl.
Mater. 374 (2008) 919.
[8] P. Sengupta, P.S. Gawde, K. Bhanumurthy, G.B. Kale, Diffusion reaction
between Zircaloy 2 and Thoria, J. Nucl. Mater. 325 (2004) 180187.
[9] P. Hejzlar, M.J. Driscoll, N.E. Todreas, Impact of fuel choices on spent fuel
characteristics for once through heavy metal cooled reactors, in: Proceedings
of a Committee Meeting (TCM) on Core Physics and Engineering Aspects
of Emerging Nuclear Energy Systems for Energy Generation and Transmutation, Argonne, IL, USA, 28 November 1 December, IAEA-TECDOC-1356,
IAEA, Vienna, 2000, pp. 168185.
[10] E.C. Buck, B.D. Hanson, B.K. McNamara, The geochemical behavior of Tc, Np
and Pu in spent nuclear fuel in an oxidizing environment, in: R. Gier, P. Stille
(Eds.), Energy, Waste, and the Environment: a Geochemical Perspective, vol.
236, The Geological Society of London Special Publication, 2004, pp. 6588.
[11] M. Benedict, T.H. Pigford, H.W. Levi, Nuclear Chemical Engineering, 2nd ed.,
McGraw Hill, New York, 1981, 369 pp.
[12] R. Roy, Ceramic science of nuclear waste xation, Am. Ceram. Soc. Bull. 54
(1975) 459.
[13] N. Das, P. Sengupta, S. Roychowdhury, G. Sharma, P.S. Gawde, A. Arya, V. Kain,
U.D. Kulkarni, J.K. Chakravartty, G.K. Dey, Metallurgical characterizations of
FeCrNiZr base alloys developed for geological disposal of radioactive hulls,
J. Nucl. Mater. 420 (2012) 559574.
[14] P. Sengupta, S. Fanara, S. Chakraborty, Preliminary study on calcium aluminosilicate glass as a potential host matrix for radioactive 90 Sr an approach
based on natural analogue study, J. Hazard. Mater. 190 (2011) 229239.
[15] Parveena, P. Sengupta, R.K. Mishra, C.P. Kaushik, S. Sai, G.K. Dey, Feasibility
study on immobilization of nuclear power plant waste in borosilicate glass
matrix doped with calcium aluminosilicate and titania, J. Nucl. Mater. (2012)
(under revision).
[16] A.E. Ringwood, S.E. Kesson, N.G. Ware, W. Hibberson, A. Major, Immobilization of high level nuclear reactor wastes in SYNROC, Nature 278 (1979) 219
223.
[17] I.W. Donald, B.L. Metcalfe, R.N.J. Taylor, Review: the immobilization of high
level radioactive wastes using ceramics and glasses, J. Mater. Sci. 32 (1997)
58515887.
[18] R.C. Ewing, Radioactivity and the 20th century, in: P.C. Burns, R.J. Finch (Eds.),
Mineralogical Society of America, Reviews in Mineralogy, vol. 38, 1999, pp.
122.
[19] R.C. Ewing, Natural glasses: analogues for radioactive waste forms, in: G.J.
McCarthy (Ed.), Scientic Basis for Nuclear Waste Management, vol. 1, Plenum
Press, New York, NY, 1979, pp. 5768.
[20] R.C. Ewing, L.M. Wang, Phosphates as nuclear wasteforms, Rev. Miner.
Geochem. 48 (2002) 673699.
[21] L.L. Hench, D.E. Clark, J. Campbell, High level waste immobilization forms,
Nucl. Chem. Waste Manag. 5 (1984) 149173.
[22] W.E. Lee, M.I. Ojovan, M.C. Stennett, Immobilization of radioactive waste in
glasses, glass composite materials and ceramics, Adv. Appl. Ceram. 105 (2006)
312.
[23] M.I. Ojovan, W.E. Lee, An Introduction to Nuclear Waste Immobilization,
Elseiver, 2005, p. 315.
[24] M.I. Ojovan, W.E. Lee, New Developments in Glassy Nuclear Wasteforms, Nova
Science Pub Inc., 2007, p. 131.
[25] International Atomic Energy Agency, Spent Fuel and High Level Waste: Chemical Durability and Performance Under Simulated Repository Conditions,
IAEA-TECDOC-1563, IAEA, Vienna, 2007, 29 pp.
[26] International Atomic Energy Agency, Status and Trends in Spent Fuel Reprocessing, IAEA-TECDOC-1467, IAEA, Vienna, 2005, 101 pp.
[27] P. Sengupta, D. Rogalla, H.W. Becker, G.K. Dey, S. Chakraborty, Development
of graded NiYSZ composite coating on Alloy 690 by pulsed laser deposition
technique to reduce hazardous metallic nuclear waste inventory, J. Hazard.
Mater. 192 (2011) 208221.
[28] P. Sengupta, Interaction study between nuclear waste glass melt and ceramic
melter bellow liner materials, J. Nucl. Mater. 411 (2011) 181184.
[29] International Atomic Energy Agency, Design and Operation of High Level
Waste Vitrication and Storage Facilities, Technical reports series no. 339,
1992.
[30] P. Sengupta, J. Mittra, G.B. Kale, Interaction between borosilicate melt and
Inconel, J. Nucl. Mater. 350 (2006) 6673.
[31] P. Sengupta, C.P. Kaushik, R.K. Mishra, G.B. Kale, Microstructural characterization and role of glassy layer developed on process pot wall during
nuclear high-level waste vitrication process, J. Am. Ceram. Soc. 90 (2007)
30573062.

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728


[32] P. Sengupta, C.P. Kaushik, G.B. Kale, D. Das, K. Raj, B.P. Sharma, Evaluation
of Alloy 690 process pot at the contact with borosilicate melt pool during
vitrication of high level nuclear waste, J. Nucl. Mater. 392 (2009) 379385.
[33] P. Sengupta, N. Soudamini, C.P. Kaushik, R.K. Mishra, D. Das, Jagannath, G.B.
Kale, K. Raj, B.P. Sharma, Corrosion of alloy 690 process pot by sulfate containing high level radioactive waste at feed stage, J. Nucl. Mater. 374 (2008)
185191.
[34] International Atomic Energy Agency, Monitoring of Geological Repositories
For High Level Waste, IAEA-TECDOC-1208, IAEA, Vienna, 2001, 25 pp.
[35] N.T. Rempe, Deep Geologic Repositories, The Geological Society of America,
2008, 123 pp.
[36] P. Sengupta, C.P. Kaushik, G.K. Dey, Immobilization of high-level nuclear
wastes: the Indian scenario, in: Mu. Ramkumar (Ed.), On the Sustenance of
Earths Resources, Springer-Verlag, 2012.
[37] W. Huang, D.E. Day, C.S. Ray, C.W. Kim, A.M. Milankovic, Vitrication of high
chrome oxide nuclear waste in iron phosphate glasses, J. Nucl. Mater. 327
(2004) 4657.
[38] C.P. Kaushik, R.K. Mishra, P. Sengupta, D. Das, G.B. Kale, K. Raj, Barium borosilicate glass: a potential matrix for immobilization of sulfate bearing high level
radioactive waste, J. Nucl. Mater. 358 (2006) 129138.
[39] X. Fang, P.R. Hrma, J.H. Westsik Jr., N.R. Brown, M.T. Schweiger, H. Li, Glass
Optimization for Vitrication of Hanford Site Low Level Tank Waste, PNNL10918, Pacic Northwest Laboratory, Richland, WA, 1996.
[40] International Atomic Energy Agency, Spent Fuel Reprocessing Options, IAEATECDOC-1587, IAEA, Vienna, 2008, 144 pp.
[41] J.M. Cleveland, The Chemistry of Plutonium, Gordon and Breach Science Publishers, New York, 1970, pp. 500503.
[42] M.E. Johnson, Origin of Wastes in C-200 Series Single Shell Tanks. RPP-15408,
CH2MHILL Hanford, Washington, 2003, p. 28.
[43] J.M. Tingey, G.H. Bryan, J.R. Deschane, Dangerous Waste Characteristics of
Waste from Hanford Tank 241-S-109, Pacic Northwest National Laboratory
(PNNL), 2004, p. 29.
[44] D.L. Herting, G.A. Cooke, R.W. Warrant, Identication of Solid Phases in Salt
Cake from Hanford Site Waste Tanks, HNF-11585, Flour Hanford, USA, 2002,
p. 140.
[45] K.M. Krupka, H.T. Schaef, B.W. Arey, S.M. Heald, W.J. Deutsch, M.J. Lindberg,
K.J. Cantrell, Residual waste from Hanford tanks 241-C-203 and 241-C-204.
1. Solids characterization, Environ. Sci. Technol. (2006) 37493754.
[46] G.J. Lumetta, B.K. McNamara, E.C. Buck, S.K. Fiskum, L.A. Snow, Characterization of high phosphate radioactive tank waste and stimulant development,
Environ. Sci. Technol. 43 (2009) 73487843.
[47] National Research Council (NRC), Alternative High Level Waste Treatments
at the Idaho National Engineering and Environmental Laboratory, Committee
on Idaho National Engineering and Environmental Laboratory (INEEL) High
Level Waste Alternative Treatments, National Academy of Press, USA, 1999,
182 pp.
[48] National Research Council (NRC), Tank Waste Retrieval, Processing and Onsite Disposal at Three Department of Energy sites: Final report. Committee
on the Management of Certain Radioactive Waste Streams Stored in Tanks at
Three Department of Energy sites, National Academy of Press, US, 2006, 201
pp.
[49] J. Ahearne, J.A. Gentilucci, D. Pye, E.T. Weber, F.E. Woolley, High Level Waste
Melter Review Report, TFA 0108, Pacic Northwest National Laboratory, Battelle, 2001, p. 61.
[50] E.B. Watson, Two liquid partition coefcients: experimental data and geochemical implications, Contrib. Mineral. Petrol. 56 (1976) 119134.
[51] I.C. Freestone, Liquid immiscibility in alkali rich magmas, Chem. Geol. 23
(1978) 115124.
[52] G. Mahood, W. Hilderth, Large partition coefcients for trace elements in high
silica rhyolites, Geochim. Cosmochim. Acta 47 (1983) 1130.
[53] J. Sommerauer, K.K. Lehnert, Trapped phosphate melt inclusions in
silicatecarbonatehydroxyapatite from comb-layer alvikites from Kaiserstuhl carbonatite complex (SW-Germany), Contrib. Mineral. Petrol. 91 (1985)
354359.
[54] B.O. Mysen, D. Virgo, C.M. Scarfe, Relations between anionic structure and
viscosity of silicate melts a Raman spectroscopic study at 1 atm and at high
pressure, Am. Mineral. 65 (1980) 690711.
[55] D.B. Dingwell, K.U. Hess, R. Knoche, Granite and granitic pegmatite melts:
volumes and viscosities, Trans. R. Soc. Edinb.: Earth Sci. 87 (1996) 6572.
[56] I. Kushiro, On the nature of silicate melt and its signicance in magma genesis.
Regularities in the shift of liquidus boundaries involving olivine, pyroxene,
and silica minerals, Am. J. Sci. 275 (1975) 411431.
[57] H.S.C. Liu, ONeill, Effects of P2 O5 and TiO2 on the partial melting of spinel
lherzolite in the system CaOMgOAl2 O3 SiO2 at 1.1 GPa, Can. Mineral. 45
(2007) 649655.
[58] B.O. Mysen, F.J. Ryerson, D. Virgo, The structural role of phosphorous in silicate
melts, Am. Mineral. 66 (1981) 106117.
[59] B.O. Mysen, F. Holtz, M. Pichavant, J.M. Beng, J.M. Montel, Solution mechanisms of phosphorous in quenched and anhydrous granitic glass as a function
of peraluminosity, Geochim. Cosmochim. Acta 61 (1997) 39133926.
[60] M. Pichavant, J.M. Montel, L. Richard, Apatite solubility in peraluminous liquids: experimental data and an extension of the HarrionWatson model,
Geochim. Cosmochim. Acta 56 (1992) 38553861.
[61] M.B. Wolf, D. London, Apatite dissolution into peraluminous haplogranitic
melts: an experimental study of solubilities and mechanisms, Geochim. Cosmochim. Acta 58 (1994) 41274145.

27

[62] C. Nelson, D.R. Tallant, Raman studies of sodium silicate glasses with low
phosphate contents, Phys. Chem. Glasses 25 (1984) 3139.
[63] R. Dupree, D. Holland, M.G. Mortuza, The role of small amounts of P2 O5 in the
structure of alkali disilicate glasses, Phys. Chem. Glasses 29 (1988) 1821.
[64] R. Dupree, D. Holland, J.A. Mortuza, J.A. Collins, M.W.G. Lockyer, Magic
angle spinning NMR of alkali phosphoralumino-silicate glasses, J. Non-Cryst.
Solids 112 (1989) 111119.
[65] D. London, M.B. Wolf, G.B. Morgan VI, M.G. Garrido, Experimental silicatephosphate equilibria in peraluminous granitic magmas, with a case study of
the Alburquerque batholith at Tres Arroyos, Badajoz, Spain, J. Petrol 40 (1999)
215240.
[66] C.M. Jantzen, Investigation of lead-iron phosphate glass for SRP waste, in: D.E.
Clark, W.B. White, A.J. Machiels (Eds.), Nuclear Waste Management, vol. 20,
Adv. Ceram., 1986, pp. 157165.
[67] R.K. Mishra, P. Sengupta, C.P. Kaushik, A.K. Tyagi, G.B. Kale, K. Raj, Thorium in
barium borosilicate glass, J. Nucl. Mater. 360 (2007) 143150.
[68] S.V. Stefanovsky, F.A. Lifanov, Glasses for immobilization of sulfate containing
radioactive wastes, Radiokhimiya 31 (1989) 129134.
[69] W.K. Kot, H. Gan, I.L. Pegg, Ceram. Trans. 107 (2000) 441449.
[70] C.M. Jantzen, M.E. Smith, D.K. Peeler, Dependency of sulfate solubility on melt
composition and melt polymerization, Ceram. Trans. 168 (2005) 141152.
[71] D. Manara, A. Grandjean, O. Pinet, J.L. Dussossoy, D.R. Neuville, Sulfur behavior
in silicate glasses and melts: implications for sulfate incorporation in nuclear
waste glasses as a function of alkali cation and V2 O5 content, J. Non-Cryst.
Solids 353 (2007) 1223.
[72] R.K. Mishra, V. Sudarsan, P. Sengupta, R.K. Vatsa, A.K. Tyagi, C.P. Kaushik,
D. Das, K. Raj, Role of sulphate in structural modications of sodium barium borosilicate glasses developed for nuclear waste immobilization, J. Am.
Ceram. Soc. 91 (2008) 39033907.
[73] V. Kain, P. Sengupta, P.K. De, S. Banerjee, Case reviews on the effect of
microstructure on the corrosion behavior of austenitic alloys for processing and storage of nuclear waste, Metall. Mater. Trans. 36A (2005) 1075
1084.
[74] W.H. Zachariasen, The atomic arrangement in glass, J. Am. Chem. Soc. 54
(1932) 38413851.
[75] J.R. van Wazer, Structure and properties of the condensed phosphates. II. A
theory of the molecular structure of sodium phosphate glasses, J. Am. Chem.
Soc. 72 (1950) 644647.
[76] U. Hoppe, G. Walter, R. Kranold, D. Stachel, Structural specics of phosphate
glasses probed by diffraction methods: a review, J. Non-Cryst. Solids 263 and
264 (2000) 2947.
[77] Y.M. Moustafa, K. El-Egili, Infrared spectra of sodium phosphate glasses, J.
Non-Cryst. Solids 240 (1998) 144153.
[78] R.K. Brow, Review: the structure of simple phosphate glasses, J. Non-Cryst.
Solids 263 and 264 (2000) 128.
[79] G. Walter, G. Goerigk, C. Rssel, The structure of phosphate glass evidenced
by small angle X-ray scattering, J. Non-Cryst. Solids 352 (2006) 40514061.
[80] F.F. Sena, J.R. Martinelli, L. Gomes, Synthesis and characterization of niobium
phosphate glasses containing barium and potassium, J. Non-Cryst. Solids 348
(2004) 3037.
[81] B. Tiwari, M. Pandey, V. Sudarsan, S.K. Deb, G.P. Kothiyal, Study of structural modication of sodium aluminophosphate glasses with TiO2 addition
through Raman and NMR spectroscopy, Physica B: Phys.: Condens. Matter
404 (2009) 4751.
[82] A.G. Blasewitz, G.L. Richardson, J.L. McElroy, J.E. Mendel, K.J. Schneider, Management of Radioactive Wastes from Fuel Reprocessing, Organization for
Economic Cooperation and Development, Paris, 1973, pp. 615654.
M. Gotic,
S. Popovic,
Corrosion of stainless steel in contact with a
[83] S. Music,
melt of phosphate glasses, J. Mater. Sci. Lett. 8 (1989) 13891390.
[84] D.E. Day, C.S. Ray, K. Marasinghe, M. Karabulut, X. Fang, An alternative host
matrix based on iron phosphate glasses for the vitrication of specialized
waste forms, Final report for DE-FG07-96ER45618, Project No: 55110; submitted to The Environmental Management Science Program, US-DOE, 2000,
pp. 39.
[85] D. Read, C.T. Williams, Degradation of phosphatic waste forms incorporating
ling-lived radioactive isotopes, Miner. Mag. 65 (2001) 589601.
[86] A.E. Hadrami, M. Mesnaoui, M. Maazaz, J.J. Videau, Kinetic dissolution of phosphate glasses containing toxic heavy metals, J. Non-Cryst. Solids 331 (2003)
228239.
[87] B.C. Bunker, G.W. Arnold, J.A. Wilder, Phosphate glass dissolution in aqueous
solutions, J. Non-Cryst. Solids 64 (1984) 291316.
[88] R.K. Brow, Nature of alumina in phosphate glass: I. Properties of sodium
aluminophosphate glass, J. Am. Ceram. Soc. 76 (1993) 913918.
[89] S.V. Stefanovsky, I.A. Ivanov, A.N. Gulin, in: T. Murakami, R.C. Ewing (Eds.),
Scientic Basis for Nuclear Waste Management, vol. XVIII, Materials Research
Society, Pittsburgh, PA, 1995, pp. 101106.
[90] S. Wegner, L. van Wllen, G. Tricott, The structure of aluminophosphate
glasses revisited: application of modern solid state NMR strategies to determine structural motifs on intermediate length scales, J. Non-Cryst. Solids 354
(2008) 17031714.
[91] R.A. Merrill, K.F. Whittington, R.D. Peters, Vitrication of High Sulfate Wastes,
Report no. PNL-SA-24672, Pacic Northwest Laboratory, Richland, Washington, 1994, pp. 16.
[92] V.V. Kushinikov, Y.I. Mayunin, N.V. Krylova, The behavior of alpha-emitting
radionuclides in the solidication of high activity waste, Atom. Energy 70
(1991) 239243.

28

P. Sengupta / Journal of Hazardous Materials 235236 (2012) 1728

[93] Y.I. Matyunin, L.J. Jardine, Investigations of Plutonium Immobilization into


the Vitreous Compositions, Report no.: UCRL-JC-130180, Lawrence Livermore
National Laboratory, USA, 1998, pp. 111.
[94] I.W. Donald, B.L. Metcalfe, S.K. Fong, L.A. Gerrard, D.M. Strachan, R.D. Scheele,
A glass-encapsulated calcium phosphate wasteform for the immobilization
of actinide-, uoride-, and chloride containing radioactive waste from the
pyrochemical reprocessing of plutonium metal, J. Nucl. Mater. 361 (2007)
7893.
[95] I.W. Donald, B.L. Metcalfe, Thermal properties and crystallization kinetics
of a sodium aluminophosphate based glass, J. Non-Cryst. Solids 348 (2004)
118122.
[96] I.W. Donald, B.L. Metcalfe, S.K. Fong, L.A. Gerrard, The inuence of Fe2 O3 and
B2 O3 additions on the thermal properties, crystallization kinetics and durability of a sodium aluminum phosphate glass, J. Non-Cryst. Solids 352 (2006)
29933001.
[97] M.Z. Dsterer, L. Montagne, G. Palavit, C. Jger, Combined 17 O NMR and
11
B31 P double resonance NMR studies of sodium borophosphate glasses,
Solid State Nucl. Magn. Reson. 27 (2005) 5064.
[98] D. Qiu, P. Guerry, I. Ahmed, D.M. Pickup, D. Carta, J.C. Knowles, M.E. Smith, R.J.
Newport, A high energy X-ray diffraction, 31 P and 11 B solid state NMR study
of the structure of aged sodium borophosphate glasses, Mater. Chem. Phys.
111 (2008) 455462.
[99] B.C. Sales, L.A. Boatner, Lead iron phosphate glass: a stable storage medium
for high level nuclear waste, Science 226 (1984) 4549.
[100] A.E. Hadrami, H. Aouad, M. Mesnaoui, M. Maazaz, J.J. Videau, Storage of toxic
heavy metals in phosphate glasses: physical and water durability properties,
Mater. Lett. 57 (2002) 894898.

A. Santi
S.T. Reis, K. Furic,
D.E. Day, Studies of leadiron
[101] A.M. Milankovic,
c,
phosphate glasses by Raman, Mssbauer and impedance spectroscopy, J. NonCryst. Solids 351 (2005) 32463258.
[102] B.C. Sales, L.A. Boatner, Lead phosphate glass as a stable medium for the
immobilization and disposal of high-level nuclear waste, Mater. Lett. 2 (1984)
301304.
[103] S.T. Reis, D.L.A. Faria, J.R. Martinelli, W.M. Pontuschka, D.E. Day, C.S.M. Partiti,
Structural features of lead iron phosphate glasses, J. Non-Cryst. Solids 304
(2002) 188194.
[104] S.T. Reis, M. Karabulut, D.E. Day, Structural features and properties of
leadironphosphate nuclear wasteforms, J. Nucl. Mater. 304 (2002) 8795.
[105] J. Kliava, R. Berger, J. Trokss, Electron paramagnetic resonance and Mssbauer
effect studies in iron-doped 57 Fe isotope enriched phosphate glasses, J. NonCryst. Solids 202 (1996) 205214.
[106] R. Berger, J. Kliava, E.M. Yahiaoui, J.C. Bissey, P.K. Zinsou, P. Beziade, Diluted
and non-diluted ferric ions in borate glasses studied by electron paramagnetic
resonance, J. Non-Cryst. Solids 180 (1995) 151163.
[107] G.K. Marasinghe, M. Karabulut, C.S. Ray, D.E. Day, D.K. Shuh, P.G. Allen, M.L.
Saboungi, M. Grimsditch, D. Haeffner, Properties and structure of vitried
iron phosphate nuclear wasteforms, J. Non-Cryst. Solids 263264 (2000) 146
154.
[108] D.E. Day, Z. Wu, C.S. Ray, P. Hrma, Chemically durable iron phosphate glass
waste forms, J. Non-Cryst. Solids 241 (1998) 112.
M. Rajic,
A. Drasner, R. Trojko, D.E. Day, Crystallization of
[109] A.M. Milankovic,
iron phosphate glasses, Phys. Chem. Glasses 39 (1998) 7075.
[110] D.E. Day, C.S. Ray, C.W. Kim, Iron phosphate glasses: an alternative for vitrifying certain nuclear wastes, Final report for DE-FG07-96ER45618, Project
No. 73976, submitted to The Environmental Management Science Program,
US-DOE, 2004, p. 20.
[111] M.G. Mesko, D.E. Day, Immobilization of spent nuclear fuel in iron phosphate
glass, J. Nucl. Mater. 273 (1999) 2736.
[112] M. Karabulut, G.K. Marasinghe, C.S. Ray, D.E. Day, O. Ozturk, G.D. Waddill,
X-ray photoelectron and Mssbauer spectroscopic studies of iron phosphate glasses containing U, Cs and Bi, J. Non-Cryst. Solids 249 (1999) 106
116.
[113] M. Karabulut, G.K. Marasinghe, C.S. Ray, G.D. Waddill, D.E. Day, Y. Badyal, M.L.
Saboungi, D. Haeffner, S. Shastri, A high energy X-ray and neutron scattering
study of Iron phosphate glasses containing uranium, J. Appl. Phys. 87 (2000)
21852193.
[114] P.A. Bingham, R.J. Hand, Sulphate incorporation and glass formation in phosphate systems for nuclear and toxic waste immobilization, Mater. Res. Bull.
43 (2008) 16791693.
[115] G.K. Marasinghe, M. Karabulut, C.S. Ray, D.E. Day, M.G. Shumsky, W.B. Yelon,
C.H. Booth, P.G. Allen, D.K. Shuh, Structural features of iron phosphate glasses,
J. Non-Cryst. Solids 222 (1997) 144152.
[116] C.S. Ray, X. Fang, M. Karabulut, G.K. Marasinghe, D.E. Day, Effect of melting
temperature and time on iron valence and crystallization of iron phosphate
glasses, J. Non-Cryst. Solids 249 (1999) 116.
[117] F.J.M. Almeida, J.R. Martinelli, C.S.M. Partiti, Characterization of iron phosphate glasses prepared by microwave heating, J. Non-Cryst. Solids 353 (2007)
47834791.

D.E. Day, Iron redox equilibrium, struc[118] X. Fang, C.S. Ray, A.M. Milankovic,
ture and properties of iron phosphate glasses, J. Non-Cryst. Solids 283 (2001)
162172.
[119] D.O. Russo, D.S. Rodrguez, J.M. Rincn, M. Romero, C.J.R.G. Oliver, Thermal properties and crystallization of iron phosphate glasses containing unto
25 wt% additions of Si-, Al-, Na- and U-oxides, J. Non-Cryst. Solids 354 (2008)
15411548.
[120] T. Murata, M. Torisaka, H. Takebe, K. Morinaga, Compositional dependence of
the valency state of Cr ions in oxide glasses, J. Non-Cryst. Solids 220 (1997)
139146.

A.M. Milankovic,
K. Furic,
V. Bermanec, C.W. Kim, D.E. Day, Structural
[121] A. Santi
c,
properties of Cr2 O3 Fe2 O3 P2 O5 glasses. Part I, J. Non-Cryst. Solids 353 (2007)
10701077.
[122] Y. Zhang, A. Navrotsky, H. Li, L. Li, L.L. Davis, D.S. Strachan, Energetics of dissolution of Gd2 O3 and HfO2 in sodium alumino-borosilicate glasses, J. Non-Cryst.
Solids 296 (2001) 93101.
[123] C. Lopez, X. Deschanels, J.M. Bart, J.M. Boubals, C.D. Auwer, E. Simoni, Solubility of actinide surrogates in nuclear glasses, J. Nucl. Mater. 312 (2003)
7680.
[124] M.G. Mesko, D.E. Day, B.C. Bunker, Immobilization of CsCl and SrF2 in iron
phosphate glass, Waste Manage. 20 (2000) 271278.
[125] X. Fang, C.S. Ray, D.E. Day, Glass transition and fragility of iron phosphate
glasses. II. Effect of mixed alkali, J. Non-Cryst. Solids 319 (2003) 314321.
[126] C.A. Click, R.K. Brow, T.M. Alam, Properties and structure of cesium phosphate
glasses, J. Non-Cryst. Solids 311 (2002) 294303.
[127] K. Joseph, R. Venkata Krishnan, K.V. Govindan Kutty, P.R. Vasudeva Rao, Crystallization kinetic of a cesium iron phosphate glass, Thermochim. Acta 494
(2009) 110114.
[128] K. Ghosh, G.K. DasMohapatra, N. Soodbiswas, Glass formation in
K2 SO4 CaOP2 O5 system, Phys. Chem. Glasses 44 (2003) 313318.
[129] I.A. Ivanov, Diffusion of sodium cations and water resistance of glasses for
immobilizing medium level wastes, Radiokhi 33 (5) (1991) 122127.
[130] C.W. Kim, D.E. Day, Immobilization of Hanford LAW in iron phosphate glasses,
J. Non-Cryst. Solids 331 (2003) 2031.
M.
[131] C.W. Kim, C.S. Ray, D. Zhu, D.E. Day, D. Gombert, A. Aloy, A.M. Milanokovic,
Karabulut, Chemically durable iron phosphate glasses for vitrifying sodium
bearing waste (SBW) using conventional and cold crucible induction melting
(CCIM) techniques, J. Nucl. Mater. 322 (2003) 152164.
[132] C.M. Jantzen, N.E. Bibler, D.C. Beam, C.L. Crawford, M.A. Pickett, Westinghouse
Savannah River Company Report, WSRC-TR-92-346, Rev. 1, 1993.
[133] U.S. DOE, Design, Construction, Commissioning of the Hanford Tank Waste
Treatment and Immobilization Plant, DOE Ofce of River Protection, Richland,
WA, Bechtel National, Inc., San Francisco, CA, 2001, Contract No.: DE-AC2701RV14136.
[134] E. Schiewer, W. Lutze, L.A. Boatner, B.C. Sales, Characterization of leadiron
phosphate nuclear waste glasses, in: L.O. Werme (Ed.), Scientic Basis for
Nuclear Waste Management IX, vol. 50, Mater. Res. Soc. Proc., 1986, pp.
231238.
[135] K. Sun, L.M. Wang, R.C. Ewing, W.J. Weber, Effects of electron irradiation in
nuclear waste glasses, Philos. Mag. 85 (4) (2005) 597608.
[136] R.M. Ezz-Eldin, Radiation effects on some physical and thermal properties of
V2 O5 P2 O5 glasses, Nucl. Instrum. Meth. Phys. Res. Sec. B159 (1999) 166175.
[137] M.I. Ojovan, K.P. Travis, R.J. Hand, Thermodynamic parameters of bonds in
glassy materials from viscositytemperature relationships, J. Phys.: Condens.
Matter 19 (2007) 112.
[138] P.A. Bingham, R.J. Hand, S.D. Forder, A. Lavaysierre, F. Deloffre, S. Kilcoyne, I.
Yasin, Structure and properties of iron borophosphate glasses, Phys. Chem.
Glass. B 47 (2006) 313317.
[139] P.A. Bingham, R.J. Hand, S.D. Forder, Doping of iron phosphate glasses with
Al2 O3 , SiO2 or B2 O3 for improved thermal stability, Mater. Res. Bull. 41 (2006)
16221630.
[140] S. Majumdar, P. Sengupta, G.B. Kale, I.G. Sharma, Development of multilayer
oxidation resistance coating on Nb and Ta, Surf. Coat. Technol. 200 (2006)
37133718.
[141] P. Mishra, P. Sengupta, S.N. Athavale, A.L. Pappachan, A.K. Grover, A.K. Suri,
G.B. Kale, P.K. De, K. Bhanumurthy, Brazing of hot isostatically pressed-Al2 O3
to stainless steel (AISI 304L) by MoMn route using 72Ag28Cu braze, Metall.
Mater. Trans. 36A (2005) 14871494.
[142] T. Okura, T. Miyachi, H. Monma, Properties and vibrational spectra of magnesium phosphate glasses for nuclear waste immobilization, J. Eur. Ceram. Soc.
26 (2005) 831836.
[143] L. Ghussn, M.O. Prado, D.O. Russo, J.R. Martinelli, Crystallization of a niobium
phosphate glass, J. Non-Cryst. Solids 352 (2006) 33913397.
[144] M.I. Ojovan, J.M. Juoi, A.R. Boccaccini, W.E. Lee, Glass composite materials for
nuclear and hazardous waste immobilization, Mater. Res. Soc. Symp. Proc.
1107 (2008) 245252.
[145] M.I. Ojovan, W.E. Lee, Glassy wasteforms for nuclear waste immobilization,
Metall. Mater. Trans. A 42 (2011) 837851.

You might also like