You are on page 1of 8

Journal of Nuclear Materials 452 (2014) 457–464

Contents lists available at ScienceDirect

Journal of Nuclear Materials


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

The suitability of a supersulfated cement for nuclear waste


immobilisation
N.C. Collier a,⇑, N.B. Milestone a,b, L.E. Gordon a,c, S.-C. Ko d
a
Immobilisation Science Laboratory, Department of Materials Science and Engineering, The University of Sheffield, Mappin Street, Sheffield S1 3JD, UK
b
Callaghan Innovation, 69 Gracefield Road, PO Box 31310, Lower Hutt 5040, New Zealand
c
Geopolymer and Minerals Processing Group, Department of Chemical and Biomolecular Engineering, University of Melbourne, Parkville, Victoria 3010, Australia
d
Holcim Technology Ltd, Hagenholzstrasse 85, CH-8050 Zurich, Switzerland

h i g h l i g h t s

 We investigate a supersulfated cement for use as a nuclear waste encapsulant.


 High powder fineness requires a high water content to satisfy flow requirements.
 Heat generation during hydration is similar to a control cement paste.
 Typical hydration products are formed resulting in a high potential for waste ion immobilisation.
 Paste pH and aluminium corrosion is less than in a control cement paste.

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

Article history: Composite cements based on ordinary Portland cement are used in the UK as immobilisation matrices for
Received 10 April 2014 low and intermediate level nuclear wastes. However, the high pore solution pH causes corrosion of some
Accepted 31 May 2014 metallic wastes and undesirable expansive reactions, which has led to alternative cementing systems
Available online 11 June 2014
being examined. We have investigated the physical, chemical and microstructural properties of a
supersulfated cement in order to determine its applicability for use in nuclear waste encapsulation.
The hardened supersulfated cement paste appeared to have properties desirable for use in producing
encapsulation matrices, but the high powder specific surface resulted in a matrix with high porosity.
Ettringite and calcium silicate hydrate were the main phases formed in the hardened cement paste
and anhydrite was present in excess. The maximum rate of heat output during hydration of the supersulf-
ated cement paste was slightly higher than that of a 9:1 blastfurnace slag:ordinary Portland cement paste
commonly used by the UK nuclear waste processing industry, although the total heat output of the
supersulfated cement paste was lower. The pH was also significantly lower in the supersulfated cement
paste. Aluminium hydroxide was formed on the surface of aluminium metal encapsulated in the cement
paste and ettringite was detected between the aluminium hydroxide and the hardened cement paste.
Ó 2014 Elsevier B.V. All rights reserved.

1. Introduction 300 m2/kg specific surface [7]) and is much coarser than the
ground granulated blastfurnace slag (GGBS) commonly used in
Composite cements based on ordinary Portland cement (OPC) the construction industry (typically 450–550 m2/kg [8]). This
have been used in the UK for approximately 30 years to encapsu- coarse BFS is used to produce a composite cement with low heat
late intermediate level and low level nuclear wastes (ILW and of hydration. Little literature or data is publicly available detailing
LLW respectively) [1–6]. These composites contain very high levels the assessment criteria and performance of nuclear encapsulation
of supplementary cementing materials (SCMs) such as blast grouts used in the UK, but the formulations currently used are
furnace slag (BFS) or pulverised fuel ash (PFA) with typical ratios primarily based around several criteria [6,9]:
of 9:1 and 3:1 BFS:OPC and 3:1 and 5:4 PFA:OPC being used. The
slag used in these formulations is very coarse (typically 260– 1. Low heat of hydration (this is generally required to be less than
that of a pure OPC paste in order to prevent detriment to
wasteform integrity caused by the loss of water during the exo-
⇑ Corresponding author. Tel.: +44 114 222 6021; fax: +44 114 222 5943. thermic reaction between the cement and the water, and to
E-mail address: nick.collier@sheffield.ac.uk (N.C. Collier). avoid radionuclide release).

http://dx.doi.org/10.1016/j.jnucmat.2014.05.078
0022-3115/Ó 2014 Elsevier B.V. All rights reserved.
458 N.C. Collier et al. / Journal of Nuclear Materials 452 (2014) 457–464

2. Long period of workability (this is to allow the infilling of metal. In performing this assessment, it should be noted that very
various waste streams by the encapsulation grout which is few specific data exist publicly which detail the performance
required to be complete by the initial grout setting time). criteria of cement powders for use in the UK nuclear waste pro-
3. Sufficient time of fluidity to allow placement (P200 mm for cessing industry. A comparison has been made where these data
encapsulation grouts used in the Magnox Encapsulation Plant do exist, but to aid further assessment of applicability, other char-
(MEP) when measured using a Colcrete Flow Channel). acteristics more general to the chemistry of cement powders have
4. Little or no bleed water. been investigated.
5. Sufficient strength after 24 h with a subsequent strength
increase to meet moving and transport requirements. This
2. Experimental
strength is low and is usually of the order of approximately
0.7 MPa after 24 h and increasing over a time period of at least
2.1. Materials
90 days.
The SSC was supplied by Aggregate Industries Ltd., a member of
The historical composite cement formulations developed have
the Holcim group of companies. The composition of the powder is
proved suitable for immobilising ILW arising from fuel reprocess-
commercially sensitive and cannot be reported in this paper.
ing and large numbers of drums are currently curing in storage
Analysis by X-ray diffraction (XRD) (Fig. 1) showed that the main
awaiting ultimate disposal [10–14]. These systems have advanta-
crystalline phase present was anhydrite with a small amount of
geous properties such as a high pH pore water which is reported
gehlenite present. What may have been very small reflections for
to provide an environment in which the solubility of many heavy
alite and belite were identified in the XRD trace (presumably from
metal oxides, carbonates and hydroxides is reduced, limiting their
the OPC), but the intensities of these reflections were very low so
migration through the porous microstructure of the hardened
the presence of these phases could not be confirmed. Very small
cement paste [15]. Additionally, the calcium silicate hydrate bind-
reflections that may have been due to the presence of hemihydrate
ers formed (known as CASAH1) have a high surface area onto which
(possibly formed from the absorbance of water by the powder
many waste species can be sorbed [16,17].
during storage and the subsequent hydration of the anhydrite)
Difficulties in using the current formulations for immobilising
were also detected. The specific surface of the powder was
some legacy or historic wastes have been highlighted and are
684 m2/kg when measured using the Blaine test.
due to adverse reactions between the waste and the cement
matrix. Specific problems occur when reactive metals such as
aluminium and magnesium arising from fuel rod claddings and 2.2. Mixing and curing
assemblies are encapsulated, as well as any uranium metal, all of
which corrode in the highly alkaline environment of the pore Samples of cement paste were mixed at water:cement (w:c)
solution [13,18]. Difficulties also arise when considering the encap- ratios varying between 0.35 and 0.7 based on weight. All pastes
sulation of zeolites, desiccants and possibly ion exchange resins were mixed in a planetary mixer adding the powder to distilled
which can be destroyed by the alkaline matrix, releasing any water over a 1 min period and, after scraping any unmixed solids
adsorbed species into the pore solution which are then available from the sides of the mixing bowl and paddle into the bowl, mixing
for leaching [19]. In the high replacement composites, reaction of was continued to give a total powder addition plus mixing time of
the SCMs is limited because the small amount of OPC present is 10 min. All samples were cured in an environmental chamber at
not sufficient to activate significant amounts of the SCMs used. A 20 °C and 95% relative humidity for hydration times up to 90 days.
large amount of alkaline pore water remains and is available for To measure compressive strength, the freshly mixed paste was
on-going reactions such as metal corrosion. This corrosion gives placed into three pre-oiled 50  50  50 mm steel moulds which
rise to hydrogen gas evolution as well as expansive corrosion were vibrated for 1 min to remove entrapped air from the paste.
products which leads to concerns about long-term durability. The moulds were then sealed in zip lock plastic bags, stored in
Several alternative cementing systems have been identified the environmental chamber, demoulded after 72 h and then
where the hydration chemistry is significantly different from that replaced in the plastic bags inside the environmental chamber.
of OPC composites and which may provide viable alternative Samples for the measurement of permeability were cast into
immobilising media [20]. Among these are sulfate activated slags, 25 mm diameter  70 mm long polypots. To investigate the
where, apart from CASAH, a significant amount of ettringite encapsulation of aluminium in the cement paste, the freshly
(a calcium aluminosulfate hydrate (Ca6Al2(OH)12(SO4)326H2O)) prepared paste was placed in polypots and a strip of grade 1050
which contains a high quantity of water within its structure is
formed and the pore solution pH is approximately 10.5–11
[21,22]. It has been reported that aluminium corrosion is limited 800
A/H
in such a matrix, although further work is needed to confirm the 700
reason for this [23]. A similar effect of limited aluminium corrosion
600
has also been noted with calcium sulfoaluminate cement (CSA)
[24] where ettringite is also the principal binder [25]. Supersulfat- 500
Counts

A/ C2S/
ed cement (SSC) has a composition similar to that of sulfate 400 C2AS C3S
C3 S
activated slag, and consists of a blend of GGBS, anhydrite (CaSO4) 300 H H
and OPC clinker ground together to form a single powder [26– A A
200 A/H
29]. Upon hydration the main binding phases are ettringite and AA A A
CASAH and the properties correspond to BS4248 [30]. 100 H

This paper reports initial results examining the potential of 0


using a UK supplied SSC as a viable immobilising media for ILW, 5 15 25 35 45 55 65

and contains some results from the encapsulation of aluminium Degrees 2 Theta

Fig. 1. XRD trace of SSC powder. Notes: H – Hemihydrate (CaSO40.5H2O), A –


Anhydrite (CaSO4), C2AS – Gehlenite (2CaOAl2O3SiO2), C2S – Belite (2CaOSiO2), C3S
1 
Cement chemist’s notation, where C@CaO, S@SiO2, A@Al2O3, S@SO3, H@H2O. – Alite (3CaOSiO2).
N.C. Collier et al. / Journal of Nuclear Materials 452 (2014) 457–464 459

aluminium (>99.0 wt% purity), approximately 2  15  50 mm, 3. Results and discussion


was pushed into each polypot ensuring complete immersion. All
other samples were hydrated as pastes in polypots. 3.1. Composition and chemical properties

2.3. Sample analysis The results for 90 day samples made at different w:c ratios are
shown in Fig. 2 (XRD) and Fig. 3 (TGA/derivative thermogravimetry
The pH of the freshly prepared cement paste was measured (DTG)) and all results at all ages are summarised in Table 1.
immediately after mixing using a hand held pH meter. Pore expres- Although not shown numerically here, the results demonstrated
sion equipment at Queen’s University Belfast was used to measure that the quantities of anhydrite and ettringite detected decreased
pH of the pore solution of the hardened cement paste immediately and increased with age respectively as the anhydrite reacted with
after expression. The flow of the fresh cement paste was measured the GGBS to form ettringite. The quantity of ettringite detected
immediately after mixing using Colflow testing equipment and ini- increased with water content at all ages indicating that there
tial and final setting time of the cement paste was measured using was always unreacted powder available to react with the increas-
a Vicat Automatic Setting Time Recorder. Heat generation during ing quantity of water used. This was due to the high specific
hydration was measured in a JAF isothermal conduction calorime- surface of the powder causing a high water demand. C-S-H, formed
ter (ICC) maintained at a temperature of 40 °C using 15 g of cement from the hydration of the OPC and from the activation of the GGBS
powder over 72 h which measured heat flow and total heat output. by the hydration of the OPC, and which loses water gradually
To analyse the hardened cement paste, the samples were between approximately 50 and 500 °C [33], was detected in all
removed from their polypots and lightly crushed by hand with a per- samples. The DTG peak for ettringite was superimposed onto that
cussion mortar to a particle size of approximately 8  8  8 mm. of CASAH making it difficult to quantify the relative amounts of
Hydration was then arrested using a solvent exchange technique these two phases. A small quantity of C2AS (from the GGBS),
[31] where the pieces of hardened cement paste were immersed in gypsum (formed from the hydration of the anhydrite) and calcite
acetone for 7 days before drying in a vacuum desiccator (1  102 - (probably formed from the carbonation of Ca(OH)2) were also
mBar) for 3 days to remove the acetone. The samples were subse- detected. The small DTG peak at approximately 370 °C could have
quently stored in sealed glass jars. The generation of bleed water been caused by the release of water from a katoite-type phase, pos-
was monitored visually during the initial (7 days) curing process sibly formed during hydration by a reaction between the alumin-
and any bleed water was removed and weighed. ium from the slag and the calcium and possibly silicon from the
For analysis by XRD and thermogravimetry (TGA), after arresting OPC, but such a phase was not detected by XRD. TGA weight losses
hydration, the particles of dried hardened cement paste were at approximately 900 °C were detected in all samples and were
crushed and ground in an agate mortar to pass a 63 lm sieve. XRD probably due to the loss of SO2 from oxidised sulphides present
spectra were obtained using a Siemens D500 diffractometer operat- in the GGBS. These results demonstrate that the hydration
ing at 2°/min with a step size of 0.02° between 5–65° 2h. For TGA, a products of the cement powder were typical of a SSC.
Perkin Elmer Pyris 1 Thermogravimetric Analyzer programmed with Calorimetry results (Fig. 4) show that the maximum rate of heat
a heating profile of 30–1000 °C at 10 °C/minute in flowing nitrogen output for the sample made with w:c 0.5 was slightly higher than
was used. The corrosion product formed during encapsulation of that made with w:c 0.4 (2.706 W/kg at 3.4 h and 1.915 W/kg at
aluminium in the cement paste was analysed by using a section of 4.0 h respectively) with the former having a higher total heat
the corroded aluminium strip removed from the hardened cement output (101.76 kJ/kg compared to 89.55 kJ/kg for the sample made
paste and analysed as above. For TGA analysis of the aluminium cor- with w:c 0.4). Compared to other results [14], the heat generated
rosion product, material was scraped from the surface of the cor- from these systems is considerably lower than from an OPC paste
roded aluminium after removal from the hardened cement paste. mixed at w:c 0.37 from which the maximum rate of heat output
For compressive strength measurement, the hardened cement when hydrated at 40 °C approached 10 W/kg and generated nearly
samples were removed from the environmental chamber after 300 kJ/kg total heat output. The maximum rate of heat output and
the desired period of hydration and a Hounsfield Model H100KS total heat output of a 9:1 BFS:OPC mixed at w:c 0.37 was approx-
Compressive Strength Testing Machine was used to test the imately 1.6 W/kg and 120 kJ/kg respectively. This suggests that the
samples in triplicate. The porosity of the hardened cement paste SSC paste is more reactive and has a higher maximum rate of heat
was measured using the following water absorption technique. output than a 9:1 BFS:OPC paste, probably due to the higher
After hydration of the cement paste for the desired period of time,
the samples were broken by hand to obtain a piece of hardened
cement paste approximately 20 mm diameter  20 mm long.
These were then immediately immersed in distilled water to G/C2AS/C
prevent drying and placed in an oven overnight at 60 °C to ensure A/G
A/E
full surface saturation. After cooling, the samples were removed E E E E
E G A/E
from the water, dried in a damp towel, weighed and then placed E G
G E w:c 0.6
in an oven overnight at 105 °C to remove free water. Finally, after
100 Counts

cooling, the samples were reweighed to calculate the mass of


water removed during heating. This loss in mass was then
converted to volume % porosity (assuming the density of water
w:c 0.5
to be 1 g/cm3). To measure saturated permeability, the 25 mm
diameter cast cores of hardened cement paste were hydrated for
the desired length of time and, after removal from the polypots, w:c 0.4
were cut to a length of 20 mm using a diamond saw and isopropa-
5 15 25 35 45 55 65
nol (IPA) as lubricant. The samples were immediately placed into
beakers of distilled water which were then placed in an oven over- Degrees 2 Theta
night at 60 °C to ensure full saturation. The saturated permeability Fig. 2. XRD traces for 90 day samples with varying w:c ratios. Notes: E – Ettringite
of the cooled samples was then measured using Hassler Cell (C3A3CaSO432H2O), G – Gypsum (CaSO42H2O), A – Anhydrite (CaSO4), C – Calcite
equipment [32]. (CaCO3), C2AS – Gehlenite (2CaOAl2O3SiO2).
460 N.C. Collier et al. / Journal of Nuclear Materials 452 (2014) 457–464

K
C
From GGBS

Legend
- w:c 0.4
- w:c 0.5
- w:c 0.6

C-S-H
Fig. 3. TGA/DTG traces for 90 day samples with varying w:c ratios. Notes: CASAH – calcium silicate hydrate, E – Ettringite (C3A3CaSO432H2O), G – Gypsum (CaSO42H2O), K
– katoite-type phase, C – Calcite (CaCO3).

Table 1
Summary of main phases detected in all samples.

Sample CASAH E C2AS A G K C


p pppppp
Powder
Hardened cement paste
1 day hydration
p pp p ppp p p p
w:c 0.5
p pp p ppp p p p
w:c 0.55
3 days hydration
pp ppp p ppp p p p
w:c 0.4
pp ppp p ppp p p p
w:c 0.45
pp ppp p ppp p p p
w:c 0.5
pp pppp p ppp p p p
w:c 0.6
28 days hydration
pp ppp p pp p p p
w:c 0.4
pp ppp p pp p p p
w:c 0.5
pp pppp p pp p p p
w:c 0.65
90 days hydration
pp pppp p pp p p p
w:c 0.4
pp pppp p pp p p p
w:c 0.5
pp ppppp p pp p p pp
w:c 0.6

Notes:
1. CA SAH – calcium silicate hydrate, E – ettringite (C3A3CaSO432H2O), C2AS – Gehlenite (2CaOAl2O3SiO2), A – Anhydrite (CaSO4), G – Gypsum (CaSO42H2O), K – Katoite-
type phase, C – calcite.
p
2. The number of ’s represents relative quantities of phases detected.
3. The phase responsible for weight loss at approximately 900 °C is not shown.

3.0 specific surface area of the powder and the small amount of OPC
100.0 acting as a slag activating agent, but is significantly less than that
2.5 of an OPC paste.
Rate of Heat Output (W/kg)

Total Heat Output (kJ/kg)

80.0 pH results (Table 2) show that after 28 days hydration, the pHs
2.0
of samples made with w:c 0.5 and 0.6 were lower than that of an
1.5
60.0 OPC paste cured for the same length of time where the pH was
Legend approximately 13.9 (after [1]) and were lower than a 9:1 BFS:OPC
- w:c 0.4
1.0 - w:c 0.5
40.0 made with a w:s ratio of 0.37 (pH 12.61). Cement paste with pH
approximately neutral will reduce the amount of corrosion of
0.5 20.0 encapsulated reactive metals such as aluminium (which is passive
in the pH region 4 to 8.5 [13]), so the lower pHs of the SSC pastes
0.0 0.0 investigated here are advantageous compared to OPC based
0 10 20 30 40 50 60
Time (hours) cements. In the SSC samples cured for 28 days, the pH increased
by approximately 1 pH unit as the w:c ratio increased from 0.5
Fig. 4. Heat of hydration curves of the SSC paste. to 0.6. This suggests that at the lower w:c ratio, there is insufficient
N.C. Collier et al. / Journal of Nuclear Materials 452 (2014) 457–464 461

Table 2 The hydrogen generated during the corrosion of the aluminium


Summary of pore solution pH and total porosity. metal in the cement paste was not determined quantitatively, but
Sample pH Total Porosity (vol.%) the volume produced was insufficient to burst the lids of the
28 days hydration polypots. Other experiments performed by the authors investigat-
w:c 0.5 10.81 28.35 ing the generation of hydrogen in a 9:1 BFS:OPC system containing
w:c 0.6 11.83 32.38 the same grade aluminium caused the lids of the polypots in which
90 days hydration the samples were hydrating to burst open with the pressure of
w:c 0.5 11.07 27.97 hydrogen generated within a few hours of hydration [18,23]. Pho-
w:c 0.6 11.52 31.91 tographs of the samples examined after hydrating for 28 days are
28 days hydration shown in Fig. 7 and show considerable corrosion product on the
9:1 BFS:OPC 12.61a 40.60b surface of the aluminium. The aluminium strip was loose within
a
Made with a water:cement solids ratio (w:s) of 0.37. the hardened cement paste indicating that the corrosion product
b
Made with a w:s ratio of 0.33 [34]. was not strongly adhered to the surrounding hardened cement
paste.

water to allow full hydration of the paste, whereas at the higher


w:c ratio there is water available to allow hydration of more 3.3. Physical properties
cement powder which releases more hydroxide anions into
solution. The pHs of the pastes cured for 90 days were very similar The 3 day compressive strengths were reduced with increasing
to those of the 28 day pastes indicating that most of the alkalis had w:c ratio (Fig. 8). A similar trend was seen for the samples
been released by 28 days. hydrated for 28 days as were those for the 90 day samples apart
from that made with a w:c ratio of 0.4 where the compressive
strength was lower than that of the 28 day sample. This was
3.2. Aluminium corrosion product
probably due to difficulties arising with mixing the cement paste
at this lowest w:c ratio where it was difficult to produce a
The XRD and TGA results of the aluminium corrosion product
well-mixed homogenous paste and the subsequent problems with
after encapsulation in the cement paste for 90 days (Figs. 5 and 6
loading the paste into the compressive strength testing moulds to
respectively) show that the main cement hydration product
ensure the moulds were filled to the same extent as with the
detected was ettringite. (The temperature at which water was
samples made using higher w:c ratios. Apart from the problems
removed from ettringite in these samples when analysed by TGA
described above for the samples mixed with a w:c ratio of 0.4,
was slightly lower than reported in the literature, but the presence
the compressive strengths at all w:c ratios generally increased
of so many XRD reflections confirmed the presence of large
with age, although this trend was difficult to confirm for the
amounts of ettringite.). This was by far the most predominant
samples made with w:c ratios of 0.5 and 0.55. It is difficult to
phase, but a small amount of Al(OH)3 (bayerite) was also identified.
compare these results with those of a BFS:OPC formulation cur-
These results suggest that the aluminium strip corrodes to produce
rently used for nuclear waste encapsulation because of the differ-
a layer of Al(OH)3 and that ettringite is formed at the interface
ences in w:c ratios used for each type of cement paste, but a
between the Al(OH)3 and the hydrating cement. The sharp XRD
comparison can be made with compressive strength data
reflection at approximately 45° 2h is from the aluminium strip.
obtained by the authors for a 9:1 BFS:OPC paste made with a
The intensities of the reflections for ettringite were greater than
w:c ratio of 0.37 (the compressive strengths of these samples
those for bayerite. When these data are compared with similar
cured for 3, 7 and 28 days were 7.5, 17.4 and 34.4 MPa respec-
results obtained from encapsulating aluminium metal in a 9:1
tively). From this it can generally be observed that the early age
BFS:OPC [18,23], it is seen that the relative intensities of the
compressive strengths of all SSC samples were higher than those
bayerite/ettringite reflections in the SSC samples were significantly
of the BFS:OPC formulation referred to here, which is likely to be
lower than those in the BFS:OPC samples. This suggests that the
due to the high fineness of the SSC powder causing accelerated
quantity of bayerite formed during corrosion of aluminium
hydration.
encapsulated in the SSC paste was less than in the BFS:OPC paste.
The flows of the freshly mixed cement pastes (Fig. 9) increased
Small amounts of calcite were detected which were likely to have
relatively linearly with w:c ratio up to a w:c ratio of approximately
been formed from the carbonation of portlandite (arising from OPC
0.55, after which the rate of paste flow increased sharply with
hydration).
increasing w:c ratio. This suggests that an excess of water may
be present in the grout at w:c ratios above 0.55 causing the higher
flow properties of the grout. For a grout made with the SSC studied
Al here to achieve the flow requirement for MEP of P200 mm, a w:c
ratio of greater than approximately 0.55 must be used.
E/A E E
E B Initial and final setting time results (Fig. 10) show an increase in
Counts

E A/C
A B both initial and final setting time with increasing w:c ratio with
E E/B P A
A the relationship between final setting time and w:c ratio being
P
particularly linear. Increasing the water content in the cement
20 paste will allow hydration of a larger quantity of cement grains
because of the provision of water to the powder facilitating hydra-
tion reactions, but higher water content will also generate more
5 15 25 35 45 55 65 pore water (water not involved in hydration reactions), which will
Degrees 2 Theta make the paste softer during the early setting process resulting in
longer setting times. A slight deviation from linearity in the initial
Fig. 5. XRD traces of aluminium corrosion product in samples made with w:c 0.5
(lower) and 0.6 (upper) after hydration for 90 days. Notes: E – Ettringite
setting time results between w:c ratios of 0.45 and 0.55 suggests
(C3A3CaSO432H2O), P – portlandite (Ca(OH)2), B – bayerite (Al(OH)3), A – that within this range, the quantity of water present has little
Anhydrite (CaSO4), C – calcite, Al – aluminium. influence on setting time.
462 N.C. Collier et al. / Journal of Nuclear Materials 452 (2014) 457–464

C
B/Gb

Legend
- w:c 0.5
- w:c 0.6

C-S-H

Fig. 6. TGA/DTG traces for aluminium corrosion product in samples made with w:c ratios of 0.5 and 0.6 after hydration for 90 days. Notes: CASAH – calcium silicate hydrate,
E – ettringite (C3A3CaSO432H2O), B – bayerite (Al(OH)3), Gb – gibbsite (Al(OH)3), C – calcite.

100
90
80
70
Flow (cm)

60
50
40
30
20
10
0
0.45 0.5 0.55 0.6 0.65 0.7
w:c ratio

Fig. 9. Flow of cement paste at various w:c ratios.

12
Final setting time
10
Setting Time (hours)

8
Fig. 7. Photographs of sample pots and corroded aluminium strip when encapsu-
lated in cement paste made at w:c 0.5 (top) and 0.6 (bottom) and hydrated for 6
28 days. Initial setting time
4

2
35 3 Days
0
28 Days 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Compressive Strength (MPa)

30
90 Days w:c Ratio
25
Fig. 10. Variation of setting time with w:c ratio.
20

15
The total porosity results (Table 2) show that samples mixed
10 using a w:c ratio of 0.6 were more porous than those mixed at
5
w:c 0.5 after curing for both 28 and 90 days because the amount
of pore water compared to solid material was higher. Porosity
0 decreased with increasing hydration time for both samples studied
0. 4 0.45 0.5 0.55 0.6 0.65
as the amount of cement hydration products formed increased
w/c Ratio
with time, filling the pore space previously occupied by water.
Fig. 8. Compressive strength of samples made with varying w:c ratio and hydrated However, this decrease was very small, suggesting that either most
for 3, 28 and 90 days. hydration had occurred by 28 days or that hydration was so slow
N.C. Collier et al. / Journal of Nuclear Materials 452 (2014) 457–464 463

that the difference between 28 and 90 days was very slight. The requirements (P200 mm for encapsulation grouts used in MEP).
total porosities of both SSC samples made with w:c ratios of 0.5 If the fineness of the SSC powder is reduced during production it
and 0.6 were less than that of a 9:1 BFS:OPC made with a w:s ratio will reduce the amount of water required to satisfy paste mixing
of 0.33 (a formulation typically used in the UK nuclear waste and fluidity criteria, which in turn will reduce porosity, heat of
processing industry) despite the SSC samples having a higher hydration and permeability, and may also reduce pore solution
water content. This is likely to be due to the higher water demand pH. These changes should result in an increase in the long-term
of the high specific surface SSC powder which results in more solid durability of any nuclear wasteform made using the SSC powder.
hydrate material being formed at any age. In analysing these data, However, a significant reduction in SSC powder fineness is also
it should be noted that drying the samples of hardened cement likely to reduce the amount of slag reaction during hydration with
paste at 105 °C may lead to dehydration of the CASAH and ettring- a reduction in the quantity of the binding phases produced
ite [33] which may return higher readings of total porosity. (ettringite and CASAH), which in turn will reduce both the
The permeability of the hardened cement paste made with a quantity of waste ions immobilised and compressive strength.
w:c ratio of 0.6 after hydration for 28 days was 1.75  105 mm/ The low pH of the SSC paste pore solution when compared to
min which was low, considering that the total porosity of this that of appropriate BFS:OPC composites is advantageous when
sample was relatively high (32.38 vol.%). This suggests that the encapsulating metals known to corrode at high pH, but corrosion
pores in the matrix of this hardened cement paste have low only appears to be reduced and not prevented. Corrosion may be
connectivity and as such the flow of fluid through the hardened controlled by adjusting the water content of the SSC paste but
cement matrix is low. further work is required to establish a paste with optimum work-
No bleed water was generated in any sample investigated in ability and desirable pH environment.
this work. The lack of any bleed water in any sample produced in this
assessment is advantageous when assessing the powder for use
in producing cemented nuclear wasteforms. This lack of bleed
4. Further discussion water is likely to be due to the high fineness of the powder which
results in high powder reactivity.
Because of the varying chemical and physical properties of the Overall, the potential to use the SSC cement power studied here
different ILW waste streams that require encapsulation/immobili- in producing immobilisation matrices for the types of wet nuclear
sation in the UK (and the changes to these properties that are likely intermediate level wastes presently being generated by the nuclear
to occur in the future) and because of the availability of publicly- fuel re-processing industry is good and may produce nuclear
available performance data for nuclear wasteforms in the UK, it wasteforms with high long-term durability.
is difficult to infer the applicability of using the SSC studied here
for processing specific waste streams. However, it is appropriate
to comment generally on the applicability of using this cement 5. Conclusions
powder in this application.
The main hydration products of this cement are ettringite and  The supersulfated cement powder studied here formed a hard-
CASAH, both of which are reported to immobilise many waste ions ened cement paste that may have acceptable physical, chemical
[17] indicating good applicability for the use of this cement as a and compositional properties for use as an encapsulant for a
material for immobilising radionuclides present in ILW. The detec- number of intermediate level nuclear wastes. However, the high
tion of anhydrite at all periods of hydration studied in this work specific surface of the powder used required a high water addi-
indicates that the amount added is more than sufficient to activate tion to ensure adequate mixing and acceptable paste fluidity
the slag over a curing period of 90 days. Any anhydrite not involved which resulted in a relatively porous hardened cement paste.
in the reaction with the GGBS will itself react with water to Despite this, the permeability of the hardened cement paste
produce hemi-hydrate (CaSO4½H2O) or gypsum (CaSO42H2O) was relatively low. Because of the high water demand, use of
which will reduce free water from the hardening paste and may this cement powder may be appropriate in producing encapsu-
reduce the amount of corrosion of reactive metals such as alumin- lation matrices for aqueous nuclear waste streams such as
ium. However, this could give rise to dimensional stability sludges or flocculated wastes, giving less permeable wasteforms
problems in hardened cement wasteforms and in construction than those produced using BFS:OPC, which in turn will give
mortars and concretes. Additionally, excess sulfate present in hard- advantages in long-term durability.
ened wasteforms or construction monoliths may cause durability  The composition of the hardened cement paste was typical of a
problems under leaching conditions because of reactions between supersulfated cement with the main crystalline phase formed
leached sulfate ions and other components of the near-field being ettringite; CASAH, anhydrite and gypsum were also
environment. detected. The quantity of crystalline material present increased
The high specific surface of the cement powder allows hydra- with increasing w:c ratio. On hydration, the maximum rate of
tion reactions to occur more quickly than those generated during heat generated by the cement paste was slightly higher than
hydration of a coarser cement powder. This higher reactivity of that of a 9:1 BFS:OPC paste made at a w:c ratio of 0.37, but
the SSC powder produces high early age strength characteristics was substantially lower than that of an OPC paste made at the
and causes the maximum rate of heat output from the SSC paste same w:c ratio; the total heat generated by the hydration of
to be slightly higher than that of a 9:1 BFS:OPC formulation the SSC paste was lower than that of both the BFS:OPC paste
presently used to encapsulate some nuclear wastes. However, this and the OPC paste. The pH of the paste pore solution mixed at
maximum rate of heat output for the SSC paste is still significantly the ratios used was up to 2 units lower than that of a 9:1
lower than that of cement pastes produced using only OPC powder. BFS:OPC paste. The pH appears to be dependent upon water
It should also be noted that despite the SSC paste having a higher content, which may be primarily due to the fineness of the
maximum rate of heat output than that of a 9:1 BFS:OPC formula- powder and the quantity of hydration product formed.
tion, the total heat output of the SSC paste (approximately 100 kJ/  The main corrosion product of aluminium encapsulated in the
kg) is less than the BFS:OPC formulation (approximately 120 kJ/kg). hardened cement paste was aluminium hydroxide (bayerite
The high specific surface of the SSC powder also means that a high and gibbsite) but large amounts of ettringite were detected
w:c ratio of approximately 0.55 is required to satisfy the flow between the aluminium hydroxide and the hardened cement
464 N.C. Collier et al. / Journal of Nuclear Materials 452 (2014) 457–464

paste. When assessed qualitatively, the corrosion of aluminium [12] N.C. Collier, N.B. Milestone, J. Hill, I.H. Godfrey, J. Nucl. Mater. 393 (2009)
92–101.
encapsulated in the hardened cement paste appeared to be little
[13] A. Setiadi, N.B. Milestone, J. Hill, M. Hayes, Adv. Appl. Ceram. 105 (2006)
and may have been less than that in a BFS:OPC paste, but 191–196.
further quantitative work is required to confirm this. [14] C.A. Utton, PhD thesis, The University of Sheffield, Sheffield, UK, 2006.
[15] F.P. Glasser, Chemistry of cement-solidified waste forms, in: R.D. Spence (Ed.),
Chemistry and Microstructure of Solidified Waste Forms, CRC Press, Florida,
1993, pp. 1–40.
[16] I.G. Richardson, G.W. Groves, Cem. Concr. Res. 23 (1993) (1993) 131–138.
Acknowledgements [17] M.L.D. Gougar, B.E. Scheetz, D.M. Roy, Waste Manage. (Oxford) 16 (1996)
295–303.
The authors thank Dr. Margaret Carter of the Construction [18] N.C. Collier, N.B. Milestone, P.D. Swift, Advances Appl. Ceramics 109 (2010)
269–274.
Science Research Group in the School of Mechanical, Aerospace
[19] L.E. Gordon, N.B. Milestone, M.J. Angus, Mater. Res. Soc. Symp. Proc. 1107
and Civil Engineering at The University of Manchester for the use (2008) 135–142.
of the Hassler Cell to measure saturated permeability. [20] N.B. Milestone, Y. Bai, P.R. Borges, N.C. Collier, J.P. Gorce, L.E. Gordon, A. Setiadi,
C.A. Utton, Q.Z. Zhou, Scientific Basis for Nuclear Waste Management XXIX,
Mater. Res. Soc. Symp. Proc. 932 (2006) 673–680.
References [21] Y. Bai, N.B. Milestone, Mater. Res. Soc. Symp. Proc. 932 (2005) 759–766.
[22] Y. Bai, N.C. Collier, N.B. Milestone, C.H. Yang, J. Nucl. Mater. 413 (2011) 183–
192.
[1] F.P. Glasser, Cem. Concr. Res. 22 (1992) 201–216.
[23] N.C. Collier, N.B. Milestone, P.D. Swift, Mater. Res. Soc. Symp. Proc. 1124 (2009)
[2] F.P. Glasser, J. Hazard. Mater. 52 (1997) 151–170.
345–350.
[3] F.P. Glasser, Mineral. Mag. 5 (2001) 621–633.
[24] Q. Zhou, N.B. Milestone, M. Hayes, J. Hazard. Mater. 136 (2006) 120–129.
[4] G.V. Hutson, Waste treatment, in: P.D. Wilson (Ed.), The Nuclear Fuel Cycle;
[25] F. Winnefeld, B. Lothenbach, Cem. Concr. Res. 40 (2010) 1239–1247.
from Ore to Waste, Oxford University Press, New York, 1996, pp. 161–183.
[26] M.C.G. Juenger, F. Winnefeld, J.L. Provis, J.H. Ideker, Cem. Concr. Res. 41 (2011)
[5] J.H. Sharp, J. Hill, N.B. Milestone, E.W. Miller, Cementitious Systems for
1232–1243.
Encapsulation of Intermediate Level Waste, in: Proceedings of the 9th
[27] A. Gruskovnjak, B. Lothenbach, F. Winnefeld, R. Figi, S.C. Ko, M. Adler, U. Mäder,
International Conference of Radioactive Waste Management and
Cem. Concr. Res. 38 (2008) 983–992.
Environmental Remediation, September 2003, Oxford, UK.
[28] A. Gruskovnjak, B. Lothenbach, F. Winnefeld, B. Münch, R. Figi, S.C. Ko, M.
[6] C. Wilding, Cem. Concr. Res. 22 (1992) 299–310.
Adler, U. Mäder, Advances Cem. Res 23 (2011) 265–275.
[7] M. Hayes, National Nuclear Laboratory, Personal communication, 2014.
[29] T. Matschei, F. Bellmann, J. Stark, Adv. Cem. Res. 17 (2005) 167–178.
[8] Civil and Marine, Production and use of GGBS, JR1/JUNE2007, http://www.civil
[30] British Standards Institution, Specification for Supersulfated Cement, BS 4248:
and marine.co.uk, accessed March 2014.
BSI, London, 2004.
[9] M.J. Angus, I.H. Godfrey, M. Hayes, S. Foster, Managing change in the supply of
[31] N.C. Collier, J.H. Sharp, N.B. Milestone, J. Hill, I.H. Godfrey, Cem. Concr. Res. 38
cement powders for radioactive waste encapsulation – twenty years of
(2008) 737–744.
operational experience, WM2010 Conference, Phoenix, USA, March 2010.
[32] K.M. Green, W.D. Hoff, M.A. Carter, M.A. Wilson, J.P. Hyatt, Rev. Sci. Instrum. 70
[10] N.C. Collier, N.B. Milestone, J. Hill, I.H. Godfrey, Waste Manage. (Oxford) 26
(1999).
(2006) 769–775.
[33] H.F.W. Taylor, Cement Chemistry, second ed., Thomas Telford, London, 1997.
[11] N.C. Collier, N.B. Milestone, J. Hill, I.H. Godfrey, J. Nucl. Mater. 393 (2009)
[34] P.R. Borges, PhD thesis, The University of Sheffield, Sheffield, UK, 2007.
77–86.

You might also like