You are on page 1of 70

Chemistry and sorption capacity of two

alternative bentonites for use in the Finnish


spent nuclear fuel repository

Master’s thesis

University of Helsinki

Department of Chemistry

Master’s Programme in Chemistry and Molecular Sciences

Karola Silenius
February 2024
Tiedekunta — Fakultet — Faculty Koulutusohjelma — Utbildningsprogram — Degree program

Faculty of Science Degree programme in Chemistry and Molecular Sciences


Tekijä — Författare — Author

Karola Silenius
Työn nimi — Arbetets titel — Title

Chemistry and sorption capacity of two alternative bentonites for use in the Finnish spent nuclear
fuel repository

Työn laji –– Arbets art –– Aika –– Datum –– Month and year Sivumäärä –– Sidoantal –– Number of
Level pages

Master’s thesis 2/2024 63

Tiivistelmä — Referat — Abstract

As part of the SAFER2028, ABCRad (Alternative Buffer/Backfill Characterization and


Radionuclide Interactions) research project, two alternative bentonite materials supplied by
Posiva Oy were investigated in this thesis. The aim of the thesis was to investigate and determine
the sorption behavior of these two buffer material candidates, with a deliberate reference to a well-
known Na-Wyoming type bentonite serving as a benchmark. In order to closely imitate conditions
relevant to repository settings, a synthetic reference water was prepared, and the experiments were
conducted within a glove box in N2 atmosphere excluding CO2 and O2. This thesis provides
valuable perspectives on the behavior and attributes of the alternative bentonite materials, which
is crucial for guiding decisions in the design of repositories for radioactive waste and strategies
for managing spent nuclear fuel. More specifically, this thesis provides thermodynamic sorption
models (TMS) for two risk-driving radionuclides, uranium (U) and cesium (Cs). Batch sorption
isotherms were made using a 1:20 solid-to-liquid ratio including 0.5 g of bentonite in 10 cm3 of
reference water. Gamma spectroscopy and Liquid Scintillation Counting (LSC) were employed
for the analysis of reaction supernatants. Complementary to these techniques, additional bentonite
properties, including Cation Exchange Capacity (CEC) and Exchangeable Cations (EC), were
determined. Pre-characterization was done for the bentonites using Fourier Transform Infrared
Spectroscopy (FTIR) and the Specific Surface Areas (SSA) were determined. These analyses
provide a comprehensive characterization of the alternative backfill materials under investigation.

ii
This thesis focuses on combining quantitative sorption data (e.g., distribution coefficient, Kd) with
mechanistic understanding (e.g., FTIR spectroscopy). This contributes to an improved
understanding of radionuclide sorption mechanisms, thereby bolstering safety considerations.
The CEC determined for the bentonites, Laviosa, LMS, and Na-Wyoming were 87 (±0,048)
meq/100 g, 95 (±0,34) meq/100 g, and 91 (± 1,23) meq/100 g, respectively. The distribution
coefficient (Kd) values of uranium ranged from 130–135 cm3/g with Laviosa and 78–110 cm3/g
with LMS, while those of cesium ranged from 130–280 cm3/g with Laviosa and from 150–425
cm3/g with LMS. Cesium demonstrated sorption of 95% within the 10-10 to 10-6 M range,
decreasing slightly to 85–95% at concentrations up to 10-2 M. Uranium showed sorption in the
range 80–100% across both clays, peaking at lower concentrations and declining at higher
concentrations. These data align with those of the reference buffer material, indicating that these
bentonites could potentially serve as feasible alternatives if they exhibit additional favorable
sorption capacity with other risk-driving radionuclides (e.g., Eu, Ni, Th).
Avainsanat — Nyckelord — Keywords

Final disposal of spent nuclear fuel, Bentonite, Sorption isotherms, Distribution coefficient,
Percent sorption, Cation exchange capacity, Zeta potential, BET, SSA, FTIR

Säilytyspaikka — Förvaringsställe — Where deposited


Kumpulan kampuskirjasto, Helsingin yliopiston kirjallinen arkisto, E-thesis
Muita tietoja — Övriga uppgifter — Additional information

iii
Table of contents
Acronyms ........................................................................................................................... vi
1. Introduction..................................................................................................................... 1
1.1 Nuclear energy in Europe ........................................................................................... 1
1.2 Nuclear energy in Finland ........................................................................................... 2
1.3 Generating nuclear energy.......................................................................................... 4
1.4 Nuclear waste in Finland ............................................................................................. 5
1.4.1 Classification of radioactive waste ........................................................................ 5
1.4.2 Spent nuclear fuel and its disposal in Finland ...................................................... 9
1.4.3 The KBS-3 design ............................................................................................... 12
1.5 Bentonite ................................................................................................................... 14
1.6 Radiochemistry ......................................................................................................... 17
1.6.1 Uranium .............................................................................................................. 17
1.6.2 Uranium – bentonite chemistry ........................................................................... 18
1.6.3 Cesium ................................................................................................................ 24
1.6.4 Cesium – bentonite chemistry ............................................................................ 24
2. Project aims and objectives ........................................................................................ 28
3. Materials and methodology ......................................................................................... 28
3.1 Bentonite pre-characterization .................................................................................. 28
3.1.1 Cation exchange capacity ................................................................................... 29
3.1.2 The exchangeable cations .................................................................................. 30
3.1.3 X-ray diffraction ................................................................................................... 31
3.1.4 The zeta potential ............................................................................................... 31
3.1.5 BET ..................................................................................................................... 32
3.1.6 FTIR .................................................................................................................... 33
3.1.7 Gamma HIDEX ................................................................................................... 33
3.1.8 Liquid scintillation counter ................................................................................... 34
3.1.9 MP-AES .............................................................................................................. 34
3.2 Sorption experiments ................................................................................................ 35
3.2.1 Reference water ................................................................................................. 35
3.3 Sorption isotherms .................................................................................................... 36
3.3.1 Sorption kinetics ................................................................................................. 36

iv
3.3.2 Varying concentration ......................................................................................... 37
3.3.3 The distribution coefficient .................................................................................. 37
3.3.4 The percent sorption ........................................................................................... 38
3.4 pH.............................................................................................................................. 38
4. Results and discussion ............................................................................................... 39
4.1 Results of sorbent pre-characterization .................................................................... 39
4.1.1 The CEC and exchangeable cations .................................................................. 39
4.1.2 XRD .................................................................................................................... 40
4.1.3 The zeta potential ............................................................................................... 41
4.1.4 SSA ..................................................................................................................... 43
4.1.5 FTIR .................................................................................................................... 43
4.2 The sorption isotherms.............................................................................................. 44
4.2.1 Reaction kinetics ................................................................................................. 44
4.2.2 The distribution coefficient .................................................................................. 46
4.2.3 The percent sorption ........................................................................................... 48
4.3 The effect of pH......................................................................................................... 50
5. Conclusions and future work ...................................................................................... 51
5.1 Conclusion remarks .................................................................................................. 51
5.2 Future work ............................................................................................................... 52
References ........................................................................................................................ 54

v
Acronyms

BET Brunauer-Emmett-Teller

BWR Boiling water reactor

CEC Cation exchange capacity

CREA Centre for Research on Energy and Clean Air

EBS Engineered barrier system

EC Exchangeable cations

EPR European pressurised reactor

FES Frayed edge site

FTIR Fourier Transform Infrared Spectroscopy

HLW High-level waste

IEA International Energy Agency

ILW Intermediate-level waste

LILW Low- and intermediate-level waste

LLW Low-level waste

MEAE Ministry of Economic Affairs and Employment

MP-AES Microwave plasma atomic emission spectrometer

PWR Pressurized water reactor

SNF Spent nuclear fuel

vi
SSA Specific surface area

STUK Säteilyturvakeskus

TRLFS Time resolved laser fluorescence spectroscopy

TOT Tetrahedral-octahedral-tetrahedral

TVO Teollisuuden voima OY

VLLW Very low-level waste

XAS X-ray absorption spectroscopy

XRD X-ray diffraction

vii
1. Introduction
This thesis begins with a brief introduction to nuclear energy, energy sources and disposal
of nuclear waste in Europe and Finland. The experimental methodology employed in this
thesis investigates the interactions between two environmentally significant radionuclides,
uranium (U) and cesium (Cs), and bentonite. The project provides valuable insights into
how the physical and chemical properties, including swelling capacity and cation exchange
capacity, influence the sorption behavior of two different types of bentonites in conditions
relevant to repository sites. Given the current focus on safety considerations revolving
around a singular type of bentonite, the exploration of various variables becomes essential.
This comprehensive investigation of two distinct bentonite types underscores the necessity
for a thorough understanding of their properties to enhance safety measures.

1.1 Nuclear energy in Europe

Nuclear power plays a vital role in the European Union's energy landscape, contributing to
approximately 25% of its electricity generation (World Nuclear Association, 2023). As of
2022, nuclear power plants accounted for 22% of the EU's electricity, with fossil fuels and
biomass contributing 42%, wind and solar sources providing 22%, and hydro contributing
10% (World Nuclear Association, 2023). This dynamic energy mix underlines the
significance of understanding and addressing challenges related to nuclear power, which
forms the focus of this discussion. Currently, 12 out of 27 EU member states host nuclear
power plants on their territory. The member states that produce nuclear energy are Belgium,
Bulgaria, Czech, Finland, France, Hungary, Netherlands, Romania, Slovakia, Slovenia,
Spain and Sweden (Dulian, 2023).

The COVID-19 pandemic and the Russia-Ukraine conflict has significantly influenced the
worldwide energy scenario, particularly in Europe. Supply chain problems and economic
slowdowns cause energy prices to increase. In this context, Finland encounters difficulties
in ensuring a reliable, cost-effective, and sustainable energy supply. This becomes more
important when we think about how Europe depends on nuclear energy and how different
factors can influence the stability of the energy supply.

1
1.2 Nuclear energy in Finland

Nuclear power has attracted considerable global attention, and its utilization as an energy
generation option has risen, driven in part by the National Energy and Climate Strategy
implemented in 2013 (Olkkonen et al., 2018). Approximately one third of Finnish electricity
comes from nuclear energy, generating 4390 MW in 2023 (MEAE, 2023)(Figure 1;Table 1).
Other sources of energy production in Finland are hydro, wind, solar, thermal and fossil fuel
power. All nuclear fuel used in Finnish facilities is imported and Säteilyturvakeskus (STUK,
Radiation and Nuclear Safety Authority) grants licenses for importing fuel and approves of
the fuel consignments (MEAE, 2023).

Finland and Sweden have the highest renewable energy manufacturing in the European
Union. The increased concern about global warming and climate change has led to shifting
towards greener options such as low carbon energy and biofuels. In June 2019, the
government announced a new policy of going fully carbon neutral by 2035 (World Nuclear
Association, 2023). Opting for nuclear energy, as opposed to fossil fuels, enhances electricity
supply security while maintaining carbon neutrality. However, the construction of nuclear
power plants is both expensive and time-consuming, spanning several decades. Balancing
these challenges with the benefits, such as a low carbon footprint and heightened energy
security, is crucial for pursuing sustainable and dynamic energy solutions.

Nuclear Hydro power


power 23 %
33 %

Wind and
solar power
12 %

Thermal Thermal
power, fossil power,
15 % renewables
Hydro power 17 %
Wind and solar power
Thermal power, renewables

Figure 1. Energy production in Finland. Adapted from: (Statistics Finland, 2022)

2
In Finland, nuclear power plants utilize either boiling water reactors (BWR) or pressurized
water reactors (PWR). In a BWR, the reactor’s core generates heat, which is used to boil
water. The resulting steam is then directed to drive a steam turbine, which in turn generates
electricity through a generator. Conversely, a PWR heats water to high temperatures, but
due to its elevated pressure, it prevents boiling. The heated and pressurized water flows
through tubes within a steam generator. The pressurized water reactor is the most widely
utilized reactor type globally. Finland has a total of five reactors (Table 1). Olkiluoto has two
BWR operating power plants: Olkiluoto 1 (OL1), Olkiluoto 2 (OL2) and a PWR, Olkiluoto 3
(OL3). These reactors are operated by Teollisuuden Voima Oyj (TVO). OL1 and OL2 each
have net capacity of 890 MW (Table 1). Finland’s newest and fifth pressurized water reactor,
OL3 has been operating by TVO since 2022. OL3 has a capacity of approximately 1,600 MW,
which is larger than any other reactor in Finland has (Table 1). Two PWR plants in Loviisa
are operated by Fortum Power and Heat Oy (Fortum). These units were opened for
commercial use in 1977 and 1980, respectively (MEAE, 2023)(Table 1). The net capacity of
Loviisa’s reactors for LO1 and LO2 is 507 MW each.

Table 1. Existing nuclear power plants in Finland. Adapted from: (World Nuclear Association, 2023)

Reactor Reactor Net Construction First Grid


type Capacity start Connection
(MW)
Loviisa 1 PWR 507 1971–05 1977–02
Olkiluoto 1 BWR 890 1974–02 1978–09
Olkiluoto 2 BWR 890 1975–11 1980–02
Olkiluoto 3 PWR 1,600 2005–08 2022–03
Loviisa 2 PWR 507 1972–08 1980–11

The potential construction of the sixth and seventh units Hanhikivi 1 and Olkiluoto 4 have
been cancelled. Hanhikivi 1 construction is cancelled due to the ongoing Russia-Ukraine
war. According to TVO, the cancellation of Olkiluoto 4 project is due to delays that occurred
during the construction of Olkiluoto 3.

3
Liberalization and the aim to become more sustainable as a society have changed the
European electricity markets in the last decades. Electricity consumption in Finland has
surged significantly, driven not only by the increasing population but also by economic and
technological advancements. The continuity of electricity supply is important for economic
growth and developing countries depend heavily on electricity. In order to reach this, we
need nuclear power. (Aydin, 2019)

1.3 Generating nuclear energy

Nuclear power plants use the fission of uranium to generate energy. Fission is a reaction
where nucleus of an unstable atom splits into two or more smaller nuclei, releasing energy.
During uranium fission, UO2 fuel, enriched in fissile U-235 undergoes fission releasing
excess energy, an average of 3.5 neutrons, and forming lighter atoms known as fission
products. Equation 1 gives an example of U-235 fission forming Cs-137. The neutrons
generated in the reaction are slowed down by a moderator. This increases their interaction
with the fuel, which contributes to a chain reaction. Each time the reaction happens, excess
energy is released in form of heat and radiation. The produced heat is controlled by a coolant
that circulates through the reactor. In a nuclear power plant, released energy is harnessed
by using it to heat water. Thermal energy is then converted into kinetic energy, which drives
turbines, which in turn produces electricity.

235U + 1n → 137Cs + 96Rb + 3 1n


Equation 1. Fission reaction from U-235 forming Cs-137, Rb-96, and 3 neutrons.

Generally, nuclear fuel can be in the reactor for 18-36 months until the U-235 content
becomes too low for the fuel to continue functioning (World Nuclear Association, 2021). U-
235 fission leads to the formation of fission products (e.g., Cs-137 and Sr-90) and
transuranics (e.g., Np-239 and Pu-239) in the spent fuel and excess neutrons can be
absorbed by other materials present in the reactor generating activation products. After the
fuel is depleted, it is still highly radioactive and must be transferred to a storage pond where
the radiation levels gradually decrease. Further details about this process are elaborated in

4
the subsequent section (1.4). In Finland, spent fuel is regarded as radioactive waste,
necessitating secure disposal measures to ensure the safety of both people and the
surrounding environment.

1.4 Nuclear waste in Finland

Finland is an international pioneer in nuclear waste management. It is one of the first


countries to solve a secure and safe disposal method for spent nuclear fuel (Auffermann,
2015). After termination of the fission reaction, SNF generates residual heat due to the decay
of radioactive isotopes contained within. Consequently, prior to disposal, it must undergo a
cooling period outside the reactor. Cooling ponds serve this purpose, where the surrounding
water acts as a radiation shield and absorbs generated heat. During this cooling phase,
isotopes with shorter half-lives decay. The heat absorbed by the water is subsequently
dissipated through external heat exchangers. This process can take several months or even
many years before it can be placed in long-term storage and final disposal (World Nuclear
Association, 2021).

Most of the radioactive nuclear waste formed in Finland is from Loviisa’s and Olkiluoto’s
nuclear power plants. Apart from nuclear power generation, radioactive waste stems from
diverse sectors like mining, technology, healthcare, and research. Mining involves the
extraction of naturally occurring radioactive materials, while technology and industrial
processes produce waste with radioactive components. Research departments using
radioactive materials also contribute to this spectrum of waste generation. Radioactive waste
and its safe disposal in Finland are regulated by the radiation and nuclear safety authority,
STUK. Loviisa and Olkiluoto, operated by Fortum and TVO, respectively are responsible for
the safe disposal of the radioactive wastes they generate according to Finnish law. Together,
they founded Posiva Oy in 1995 to handle the final disposal of the spent nuclear fuel. In 2015
Posiva Oy got a license from the Finnish Government for the construction of a disposal
facility, Onkalo. This collaboration is discussed more in detail in section (1.4.2).

1.4.1 Classification of radioactive waste

Radioactive waste necessitates isolation from the environment until its radioactivity
diminishes to safe levels, safeguarding ecosystems and human health. Owing to intrinsic

5
chemotoxic and radiotoxic characteristics, radioactive waste is a considerable risk to the
environment. Contamination of the environment by nuclear waste may result in migration
of radionuclides to the biosphere and incorporation into the food chain, presenting a direct
threat to human health. Nuclear waste contains particles and elements, which can build up
in human tissues (Philipp et al., 2019). The management of nuclear waste involves
considerations of its activity, radionuclide composition, and their half-lives, influencing the
processes of disposal. The Finnish government grants licenses for nuclear waste
management facilities and composes regulations regarding safe disposal of nuclear waste.
STUK has determined the classification of radioactive wastes in Finland (Table 2) and here,
radioactive waste is categorized into very low-, low-, intermediate-, and high-level wastes.

6
Table 2. The classification of radioactive wastes. Adapted from: (STUK, 2020)

Waste Definition Examples

Very low-level waste (VLLW) • Specific activity: As < • Residual activity, naturally
100 kBq/kg occurring radionuclides,
protection equipment
• Total activity: At < 1 TBq

• or Alpha activity: Aa of
10 GBq

Low-level waste (LLW) • Concentration activity of • Contaminated items: work


LLW is not more than 1 clothing and tools
MBq/kg containing 1% of
radioactivity
• Waste can be
processed without
special sorting

Intermediate-level waste • 1 MBq/kg < A (ILW) < 1 • Steel equipment from


(ILW) GBq/kg reactors, filters and
effluents from
• No cooling process reprocessing containing
needed: Doesn’t 4 % of radioactivity
produce enough heat

• Processing ILW requires


special arrangements
for radioactivity
protection

High-level waste (HLW) • A > 10 GBq/kg • Spent nuclear fuel (SNF)


containing ̴ 95 % of
• Cooling solutions radioactivity in the nuclear
needed due to heat waste
production
• Transuranic (TRU)
radionuclides from
reprocessing of SNF

Very low-level waste (VLLW) represents a subclassification within the low-level waste
category, primarily comprising naturally occurring radionuclides along with radionuclides
characterized by half-lives shorter than 30 years. The wastes in this category have a
concentration activity of <100 kBq/kg (Table 2). The main radionuclide in this waste

7
category is Sr-90 (t1/2=28.8 y) which is a pure beta emitter. VLLW is generated as a result of
operation and decommissioning of nuclear facilities (Ojovan, 2011). Very low-level waste
includes residual radioactivity and waste from households and industry that include
radioactive particles. Soil and steel items are major components of VLLW (NRC, 2021;
UKRWI, 2023). Currently, Finnish VLLW is disposed of in underground repositories, that
were originally designed for low- and intermediate-level waste (LILW) disposal (Ho et al.,
2023). According to the Finnish disposal concept, VLLW is packed into canisters made from
carbon and steel and they are disposed in a repository with multi-layer barrier system
(Figure 3) to isolate the waste from the environment. The burial concept is only required to
isolate the radionuclides from surrounding biosphere for several hundred years (Ho et al.,
2023). The near-surface disposal of VLLW is a waste management option that will be
employed by many countries. The Finnish Nuclear Energy Act (990/1987) permits the
storage of VLLW in near-surface repositories and TVO is in the advanced stages of planning
a disposal facility at the Olkiluoto nuclear plant (Ho et al., 2023).

Low-level waste (LLW) has a concentration activity of not less than 1 MBq/kg (STUK , 2022;
World Nuclear Association., 2023). This type of waste includes insulation materials, packaging
and protective equipment accumulated in the operating of nuclear plants (MEAE, 2022).
The radionuclides including in LLW are e.g., I-129 (t1/2=16.1 x 106 y) and Co-60 (t1/2=5.3 y)
(García-Toraño et al., 2018; Ahmad, H. et al., 2022). LLW accounts for the majority of
produced nuclear waste by the volume of approximately 90 %. Low level solid waste is placed
in metal containers and buried in shallow pits for disposal at the site or at licensed disposal
facility (Shchipalkina and Smirnova, 2023). LLW is typically compressed in production site
and it is then packaged for disposal (STUK, 2022).

Low- and intermediate-level wastes (LILW) encompass waste materials falling within the
classification of either low-level or intermediate-level waste. In Finland, the disposal of such
waste involves placing it in geological repositories situated 60–100 meters below the
surface, within concrete chambers excavated into the bedrock near the Olkiluoto and Loviisa
nuclear power plant sites. The main radionuclides contributing to the activity in the Finnish
LILW inventory are reported to be Cs-137 (t½=30.07 y), Co-60 (t½=5.27 y) and Ni-63
(t½=100 y)(STUK, 2022; Kumpula L. et al, 2022). The concern with LILW disposal lies in
effectively isolating and containing these waste materials to prevent the release of harmful

8
substances such as organic waste into the environment. Organic waste, such as ion exchange
resins, may contain hazardous chemicals or radioactive isotopes that can pose risks to
human health and the environment if not properly managed (IAEA, 2002). Low and
intermediate level nuclear waste has underground final disposal facilities in the power plant
areas and LILW require licensing for disposal.

Finnish HLW consists of mostly spent nuclear fuel (SNF) and will be disposed of in Posiva
Oy’s final disposal facility in Olkiluoto. Here, uranium is the dominant radionuclide by mass,
making up 97% of the spent fuel. Other components of HLW are the fission products (e.g.,
Cs-137, t1/2=30 y) and transuranic elements (e.g., Np-237, t1/2=2.1 x 106 y and Pu-239,
t1/2=2.4 x 104 y) generated in the reactor core. SNF is not reprocessed in Finland and
accounts for the vast majority of HLW due to this fact. In other countries that do, or have
previously reprocessed SNF, (e.g., the UK) the handling and storage of HLW is significantly
more complex. HLW is not only radioactive, but constantly generates heat (Table 2). High-
level waste must be cooled in ponds for a long period of time before permanent and secure
long-term disposal. HLW must take heat management into consideration for storage and
disposal. The cooling of HLW can take up to decades (World Nuclear Association, 2023).
Following the cooling phase of HLW, significant considerations relate to the selection of the
host rock surrounding the waste and determining the depth of the repository (Pusch, Yong
and Nakano, 2018). The spent fuel is divided into two general categories of radionuclides:
fission and activation products plus actinides (Hedin, 1997). This thesis focuses on SNF
disposal, so it is discussed in further detail in the next section.

1.4.2 Spent nuclear fuel and its disposal in Finland

The disposal of spent nuclear fuel represents a critical aspect of nuclear energy management,
requiring careful consideration of safety, environmental impact, and long-term stability.
Spent nuclear fuel is a byproduct of nuclear energy production, containing a complex
mixture of fission products, transuranic elements, and other radioactive isotopes.

All Finnish nuclear reactors are fueled using uranium dioxide (UO2) (Galahom and Sharaf,
2021). When the fuel is spent its composition is still mostly UO2, however fission products
are formed in addition to minor actinides. One key, risk-driving, fission product is Cs-137

9
(6.3% yield fission product from fission of U-235, t1/2=30.2 y), which is one of the
radionuclides of interest in this thesis. Other products created from electricity production
process in a reactor are plutonium (Pu-239), neptunium (Np-239), americium (Am-241) and
curium (Cm-244). These actinides are heavier elements than uranium and they are formed
when uranium captures a neutron without splitting (De Jesus et al., 2021; Galahom et al.,
2021). The typical composition of SNF after 10 years of cooling is presented in Figure 2 and
it is determined by the enrichment level of U-235. Uranium takes up to 95.6% of SNF and
the rest consists of minor actinides, plutonium, long-lived iodine, technetium, cesium,
strontium, other fission products and stable fission products. The safe and secure disposal
of SNF is imperative to prevent environmental contamination and safeguard public health.

Figure 2. Approximate composition of spent fuel after 10 years of cooling. Adapted from: (Blue Ribbon Commission on
America’s Nuclear Future Report to the Secretary of Energy | Department of Energy, 2012)

Currently, the operating Finnish nuclear power plants at Olkiluoto and Loviisa generate
spent nuclear fuel. This is stored for cooling temporarily in the power plant areas. During
this time the initial radioactivity and heat output of SNF decreases to a level required for

10
safe and final disposal in Onkalo. The storage in Onkalo has been expanded to meet the
needs for the newest reactor, Olkiluoto 3 (MEAE, 2022).

In addition to understanding the conversion of radioactive element activity into


radiotoxicity and assessing the radiotoxicity of the radionuclides, it is also essential to
evaluate their potential migration pathways, interaction with geological formations, and
long-term environmental impact (Hedin, 1997). HLW generates heat as a result of
radioactive disintegrations. This heat primarily originates from the fission reaction of the
fuel and the radioactive decay of the components that underwent fission. Some of the
radioactive materials in the fuel have long half-lives, requiring their isolation from the
natural environment for long periods of time. For this reason, the final disposal canisters
(Figure 3) are designed to remain water-tight in their final deposition place long enough for
the radioactivity of spent fuel to decrease to a level not harmful to the environment.

According to Finnish legislation, the companies that operate nuclear power plants are
accountable for the disposal of their nuclear waste. In 1995, the two Finnish power
companies, TVO and Fortum, co-founded Posiva Oy, which in turn, submitted a license
application for SNF encapsulation to the Finnish Government. In 2021, Posiva Oy submitted
to the Finnish Government an operating license application for SNF encapsulation plant and
disposal facility. Posiva Oy was the first in the World to submit such a licence and aims to
start final disposal in approximately 2025 (Eronen et al., 2023). Finland has adopted a
geological disposal approach for its spent nuclear fuel, establishing the Onkalo repository as
a key component of this strategy. Onkalo, located in Olkiluoto, is designed to provide a stable
geological environment for the isolation of high-level radioactive waste. The selection of
suitable geological formations and engineered barriers is essential to ensure the
containment of radionuclides over extended periods.

Finland's geological conditions play a remarkable role in the success of the Onkalo
repository. The Olkiluoto site has crystalline bedrock, characterized by low groundwater
flow and minimal seismic activity. This geological stability enhances the long-term isolation
potential of the repository, which is crucial for the safety of the disposal strategy. Onkalo has
been built at a depth of ~420 meters in the bedrock. The final disposal concept consists of
the waste repository and related above- and underground facilities. The encapsulation
facility located above ground is connected via an elevator and canister shafts with the
underground facilities. Onkalo will eventually store spent fuel in a multilayer disposal

11
concept that includes natural and engineered barrier systems (EBS). The facility consists of
spiral-shaped tunnels and vertical shafts underground.

The construction of Onkalo started in June 2004. By 2020 approximately 10 km of tunnels


had been excavated in the bedrock (POSIVA, 2012). Posiva Oy’s research and underground
facility has attracted interest internationally and it is a hot topic now, while nuclear energy
is produced increasingly. Drying and packaging of the spent nuclear waste into disposal
canisters is done above ground. The underlying necessity of secure disposal of SNF is
isolating the radioactive material in a deep repository with physical and chemical barriers
and buffers (IAEA 2022). The final disposal is based on a concept called KBS-3. This is
described in the next chapter.

1.4.3 The KBS-3 design

The generally agreed plan for disposal of spent nuclear fuel in Finland follows the Swedish
methodology KBS-3, which is developed by Swedish Nuclear Fuel and Waste Management
Company, SKB (SKB, 2010; POSIVA, 2012). The KBS-3 system consists of a central facility
for intermediate storage and encapsulation of spent nuclear fuel, a system for transportation
of canisters with spent nuclear fuel, and a final repository facility (SKB, 2010). The purpose
of the design is to store SNF until its radioactivity and heat production has decreased to
levels suitable for deposition in a KBS-3 style repository. The KBS-3 system consists of both
natural and engineered barriers that both ensure a safe disposal of waste. It is a multi-
layered system, with each layer possessing unique protective capabilities. The design
ensures comprehensive isolation: if one layer in the engineered barrier system (EBS) were
to fail, the presence of additional layers serves to buffer the system (Figure 3).

12
Figure 3. The disposal concept KBS-3 for final storage of spent nuclear fuel. (SKB)

To ensure a prolonged period of containment of the SNF, the canister consists of cast iron
insert and copper canister. The outer copper canister provides a robust barrier against
corrosion, thereby safeguarding the longevity of the containment system. The copper
canister, acknowledged for its low corrosion rate prevents water ingress, thereby
safeguarding the contained radioactive materials and mitigating environmental risks.
(POSIVA, 2012)

The function of the secondary barrier, bentonite, involves providing mechanical shielding
and bolstering canister stability. It also adsorbs heat generated from the wastes inside the
canister. The material surrounding the canister should have mechanical, geochemical and
hydrogeological conditions that are predictable and are favorable to the canister (POSIVA,
2012). The material surrounding the canister, acting as a secondary barrier and filling within
the host rock, is expected to primarily consist of bentonite (Yoon, Jeon and Lee, 2023; Peng
et al., 2024). Bentonite has multiple properties that make it a suitable buffer material. It has
favorable mechanical properties such as swelling capacity and plasticity. In addition to these
properties, bentonite has low permeability, small pore size, good buffering capacities,
limited microbial activity and high sorption capacity for radionuclides. Because of its
swelling properties, bentonite has the ability to absorb nearby water and expand
accordingly. The small pore size is beneficial as it limits the flow of water. Bentonite as a
buffer material is extensively utilized as a barrier due to its high capacity to sorb
radionuclides, limiting their transport into the environment. These attributes play a critical
role in the secure disposal of radioactive waste. The subsequent chapter (1.5) provides a

13
more detailed discussion of bentonite, exploring its properties to ensuring the safe
containment of radioactive materials. The effectiveness of the multilayer barrier system
hinges on the careful selection of the host rock surrounding the bentonite material. This is a
complex process affected by technical and nontechnical factors (Cao et al., 2019).

The final part of the KBS-3 design is the surrounding bedrock and its function is to help
isolate the SNF repository from the surface environment. This buffer layer provides
mechanical, geochemical and hydrogeological conditions for the engineered barriers below
it (SKB, 2021). The bedrock structure has pores and fractures where radionuclides
eventually released from the SNF would hopefully be caught (SKB, 2021). The Finnish
bedrock comprises a variety of mineral species, each possessing specific cation exchange
capacity (CEC). The characteristics of the bedrock, coupled with its deep location, make it
suitable for the safe disposal of SNF. The unique mineral composition and exchange
capabilities of the northern bedrock contribute to its effectiveness as a secure repository for
radioactive waste.

1.5 Bentonite

Bentonite is a clay mineral of the smectite group that includes approximately 50-80% of
montmorillonite mineral. Clay minerals generally consist of minerals such as kaolinite, illite,
vermiculite and montmorillonite (Liu et al., 2019). These minerals belong to the
phyllosilicate group and are composed of silica tetrahedral sheets combined with alumina
octahedral sheets. Certain minerals, such as vermiculite and montmorillonite, are
recognized as expansive clay components due to their expansion properties. This unique
characteristic contributes to the advantageous sorption behavior of bentonite material. The
capacity of bentonite to swell upon contact with water enhances its ability to absorb and
retain large quantities of water, making it an ideal choice particularly in the context of
radioactive waste disposal, where effective sorption is necessary for maintaining the safe
disposal of hazardous waste and minimizing potential environmental impact.

Bentonites are composed mostly of montmorillonite [(Na, Ca)0.33(Al, Mg)2(Si4O10)(OH)2·n


H2O] (Shetti et al., 2018). Montmorillonite is an alumina phyllosilicate mineral that has a
tetrahedral-octahedral-tetrahedral (T-O-T) layer structure. In a T-O-T structure, the two
tetrahedral layers surround an octahedral layer. The basal layer thickness of each platelet in
the montmorillonite layer structure is in the range of 1 nm and the lateral dimension is

14
approximately 200–600 nm (García-Guzmán et al., 2023). The net negative charge is
produced via substitutions of Si4+ in the Si-tetrahedral sheet by Al3+ and substitutions of Al3+
in the Al-octahedral sheet by Mg2+(Shetti et al., 2018; García-Guzmán et al., 2023). The
overall negative charge can is compensated by the cation adsorption of other cations (e.g.,
Li+, Na+ and Ca2+) located between the TOT sheets (García-Guzmán et al., 2023). These
cations may be easily replaced through cation exchange. Because of these isomorphic
substitutions, positively charged cations can retain into it (Zhou, C. et al., 2019). The cations
attract to the negatively charged surface surrounding the interlayer (Muurinen, 1990;
Pourhakkak et al., 2021). Having layer charges with varying densities explain the anisotropic
elastic properties of dry montmorillonites (Muurinen, 1990; Shetti et al., 2018).

Presenting a significant group of natural fine-grained particles and being a porous mineral,
montmorillonite is an important part of the EBS. In addition to good swelling properties, it
has other beneficial properties such as abundance in nature, environmental friendliness,
large internal surface area, high cation exchange capacity (CEC), high adsorption capacity,
low to no toxicity, and good biocompatibility (García-Guzmán et al., 2023). Low
permeability of the montmorillonite mineral is a preferable property since heat and
mechanical forces can increase the permeability of the material that is used (Galamboš et
al., 2010).

Montmorillonite is amphoteric and the edge sites show variable charges because they can
exchange protons (Liu et al., 2013). The functional groups in the edge sites may include
entities of –OH or –OH2+ (Liu et al., 2013). Bentonite sorption capacities are therefore
highly sensitive to pH changes. The effect of pH on sorption process has been studied well
in previous years (e.g., Bayülken et al., 2011; Philipp et al., 2019). pH as a factor affects the
charges on the bentonite surface: the increase in pH creates greater net negative charge on
the surface of the bentonites and there are less H+ ions present in the reference solution
which leads to less competition for the positively charged radionuclides e.g., Cs and U to
sorb onto the material. At pH above 4, the montmorillonite generates more negative sites
for cation sorption (Abollino et al., 2008). The content of montmorillonite in bentonite
varies between bentonite types and this affects their swelling abilities.

15
Figure 4. An illustration of montmorillonite interlayers at different hydration stages. Adapted from (Birgersson et al.,
2017)

The particles of the montmorillonite mineral form layers which can expand from hydration.
Dry montmorillonite layers have a spacing of ~9.7 Å (Birgersson et al., 2017)(Figure 4). The
spacing is increased due to hydration. Basal spacing of values 12.5 Å and 15.0 Å signify a
monolayer and a double-layer of water, respectively (Birgersson et al., 2017). Generally, the
predominant counter cation in bentonite is either Ca2+ or Na+ (Muurinen, 1990). As
discussed above, cations like Na+ or Ca2+ balance the net negative charge of the Al and Si
sheets. The cations are hydrated and the spacing of montmorillonite is a function of water
content. Compacted sodium bentonite is recognized as an advantageous backfill material for
the disposal of spent fuel in various countries. The most utilized bentonite material for this
purpose is a sodium bentonite with commercial name MX-80, from Wyoming, USA
(Muurinen, 1990; Youssef, 2017). However, some research has suggested that calcium
bentonites may be more stable towards chemical erosion (Schatz, T. et al., 2013). The
distribution, hydration, and diffusion properties within the interlayer of montmorillonite
vary depending on the types of interlayer cations present (Wu, Cao and Lv, 2018). This
conclusion highlights that different cations influence the properties of bentonite. The
examination of the interlayer cations is specifically investigated through X-ray diffraction
analysis (XRD), a detailed process that helps unravel the interactions within the bentonite
structure. This is discussed in more detail in the results section (4.1.2).

Given the widespread utilization of bentonite on a global scale, it becomes imperative to


prioritize and enhance safety measures as part of the safety case. Bentonite emerges as a
cost-effective and highly promising option, possessing remarkable attributes, as discussed
earlier. Its predominance as a backfill and buffer material in geological disposal scenarios
worldwide underscores its significance. Within a barrier system, bentonite has versatile
roles, primarily focused on the crucial task of retaining radionuclides. Recognizing the

16
importance of bentonite in these applications, continual efforts are warranted to refine
safety protocols. Notably, existing safety considerations predominantly revolve around Na-
Wyoming type bentonite. Thus, this thesis delves into investigating the chemistry of two
alternative options, Laviosa and LMS.

1.6 Radiochemistry

Given the presence of radionuclides in spent nuclear fuel (SNF), a comprehensive


understanding of their interactions with bentonite is essential to ensure the secure disposal
of SNF. Two risk-driving radionuclides present in SNF, uranium and cesium, will be
discussed in subsequent chapters (1.6.1-1.6.4) with a focus on their nuclear chemistry and
interactions with bentonite. The radionuclides (U and Cs) of study in this thesis were chosen
following discussions with by Posiva Oy.

The transportation of radionuclides to bentonite are mostly controlled by diffusion and


sorption processes (Muurinen, 1990). Diffusion is a spontaneous phenomenon
characterized by the directional flow of matter from regions of higher concentration to those
of lower concentration. Sorption, on the other hand, is a physico-chemical process involving
the adherence of a substance (sorbate) onto another material (sorbent). Sorption can be as
either adsorption or absorption. In absorption, there is an exchange of particles, while in
adsorption, the atoms or ions of the substance adhere to the surface of the adsorbent. Other
adsorption mechanisms include ion exchange, surface complexation and hydrogen bonding
(Liu et al., 2023).

1.6.1 Uranium

Uranium is the most abundant radionuclide, by mass, in SNF (approximately 97%). Having
high chemotoxicity and a long half-life makes it a serious hazard for the environment.
Uranium is a kidney toxin, neurotoxin, immunotoxin, mutagen, carcinogen and teratogen
(Majdan et al., 2010).

Uranium is a naturally occurring actinide with an average concentration of 2.8 parts per
million in the Earth's crust (World Nuclear, 2023). Uranium is characterized by a silvery-
grey appearance with an atomic number of 92. Its melting point is 1132 °C and it has a high
density of 19.1 g/cm3, making it a viable metal for manufacturing products and in industry.

17
There are three dominant uranium isotopes present in SNF. U-238 is the dominant isotope,
with a long half-life (t1/2=4.5 x 109 y). The other isotopes of uranium in SNF are U-235
(t1/2=700 x 106 y) and U-234 (t1/2=2,47 x 105 y). The isotopic composition of uranium in SNF
is as follows: 95% U-238, 1.0% U-235, and 0.001% U-234, respectively (World Nuclear,
2021). The isotope U-238 is fertile and U-235 is fissile, meaning it is capable of undergoing
fission.

Uranium has 6 valence electrons and is generally found in the environment as U(IV) and
U(VI). U(IV) is poorly soluble and relatively immobile. U(VI) dominates under oxidizing
conditions and is present as the uranyl cation, [UO2]2+. The uranyl cation is readily soluble
and therefore mobile and a concern in safety cases. U(IV) dominates reducing conditions
generally present as relatively insoluble U(IV) minerals (e.g., uraninite, UO2). This thesis
focuses on the chemistry of aqueous uranium present as the uranyl cation. Although U is
present in fuel as U(IV) (as UO2 fuel), upon leaching from the spent fuel the U is expected to
oxidize to form more soluble U(VI). In the uranyl cation, uranium (+VI) is bonded with two
axial oxygen atoms, [UO2]2+, and ligands bind around the equatorial ring.

1.6.2 Uranium – bentonite chemistry

In the context of nuclear waste management and environmental protection, understanding


the interaction between uranium and bentonite is crucial. When uranium enters the
geological environment, uranyl ions can be adsorbed by minerals, which affect the migration
and transformation of uranium in the geological environment (Yu et al., 2020). The
adsorption of uranium onto bentonite is a complex process affected by several mechanisms.
Factors influencing the mechanisms are pH, Eh, mineral composition, microbial activity,
and ionic strength.

Cations like uranyl adsorbs onto bentonite at three positions: the planar, interlayer and
frayed edge site (FES). Experiments examining uranyl adsorption onto smectites have
revealed that the dominant mechanisms of U(VI) uptake are both ion exchange and surface
complexation and, the pivotal factors influencing the adsorption of U(VI) by bentonite
include pH and the ionic strength of the environment (Pabalan and Turner, 1996; Chisholm-
Brause et al., 2001; Korichi and Bensmaili, 2009; Schlegel and Descostes, 2009; Wang et
al., 2017; Tran et al., 2018a; Philipp et al., 2019, 2022; Yu et al., 2020).

18
Ion exchange is a mechanism where uranium ions U(VI) replace exchangeable cations, such
as Ca2+, K+ or Na+ on the surfaces of bentonite. The U(VI) ions in the surrounding solution
interact with the charged sites on the bentonite surfaces, which results in the adsorption of
U(VI) onto the bentonite. Uranyl ions bind via cation exchange on negatively charged
surfaces (Tran et al., 2018b; Philipp et al., 2019, 2022) in competition with other cations or
radionuclides which sorb through a similar mechanism (Tran et al., 2018).

Once the ions are bound through cation exchange, a common mechanism of sorption is
surface complexation (Yu et al., 2020; Liu et al., 2023). This mechanism involves the
formation of surface complexes between U(VI) ions and hydroxyl groups on the bentonite
surface at specific sites. The mechanism of the adsorption process by surface complex
formation between the mineral (e.g., montmorillonite) and the radionuclide is based on
inner-sphere and outer-sphere surface complexes. The inner-sphere complexes form a
direct bond between the surface and the absorbed radionuclide (e.g., U(VI)) (Payne et al.,
2013). These complexes involve a strong chemical bond. The outer-sphere complexes form
a bond with the water molecules in the hydration sphere of the absorbed radionuclide
involving weaker physical bonds via electrostatic forces of attraction (Payne et al., 2013).
Schlegel et al. (2009) discovered that at near-neutral pH, U(VI) formed inner-sphere surface
complexes at montmorillonite’s layer edges. The exact nature of the sorption complex that
forms remained unclear since U(VI) was proposed to bind either Al or Fe octahedra (Schlegel
and Descostes, 2009). Aqueous uranyl species may occupy more than one surface site
because they may be larger than a single surface site and form bidentate surface complexes
under certain conditions, thereby occupying two surface sites (Chisholm-Brause et al.,
2001). The sorption sites of the bentonite are pH dependent, and this affects the adsorption
of the radionuclide.

19
Figure 5. A structural model for U(VI) sorption on montmorillonite (a). Side view of the edge complex formed on the
border sites of montmorillonite (b). Structure model for U(VI) bonding to Si sites at hectorite edges (c). (Schlegel and
Descostes, 2009)

A polarized Extended X-ray absorption fine structure (EXAFS) study was completed by
Schlegel et al. (2009) to understand the nature of the adsorbed species of U(VI) with and
without complexing ligands. The EXAFS results showed that U(VI) binds to the edge of Al
octahedra at the montmorillonite surface (Schlegel and Descostes, 2009)(Figure 5).
Montmorillonite, a major constituent of bentonite, facilitates diverse U(VI) binding to
amphoteric sites like silanol and aluminol, potentially via edge or corner sharing
mechanisms, involving mono/bi-dentate coordination (Verma et al., 2015). This means that
ligands with a single or double donor site are arranged in a way that they share a common
corner around the metal atom in the middle. The dependence of uranium sorption modeling
results on specific surface area and edge site surface is useful to describe and predict U(VI)
retardation as a function of chemical conditions in the field-scale reactive transport
simulations. Therefore this approach can be used in the environmental quality assessment
(Korichi and Bensmaili, 2009).

In their study, Yu et al. (2020) investigated the adsorption behavior of U(VI) on


montmorillonite colloids using a test tube setup consisting of 9 mL of montmorillonite
colloids and 1 mL of U(VI) solution (30 mg/L) in a total volume of 10 mL (Yu et al., 2020).

20
They examined the impact of contact time on adsorption and observed that initially, U(VI)
adsorption occurred rapidly onto the colloids. Once it reached around 72%, the adsorption
rate stabilized, indicating a saturation point in the adsorption process (Yu et al., 2020).
Youssef, (2017) discovered similar behavior on U(VI) removal from uranium solution (200
mg/L) by compacted sodium bentonite at room temperature. Results showed that increasing
the contact time from 1 to 120 min, resulted in U(VI) adsorption efficiencies increasing
significantly from 22.5% to 91% from 1 to 120 min, respectively (Youssef, 2017). This
suggests that sorption of uranyl (VI) ions onto bentonite reaches equilibrium in
approximately 1 hour.

Uranyl's sorption to surfaces is highly influenced by its complexation with carbonate (Tran
et al., 2018). Reduced sorption in the presence of carbonate can be explained due to the
formation of ternary calcium-uranyl-carbonate compounds, which remain stable in solution
and do not interact with mineral or colloid surfaces. The formation of these complexes is
likely to influence U(VI) mobility through a carbonate rock aquifer (Tran et al., 2018). This
behavior is also depicted in Figure 6, where the adsorption of U(VI) is higher in the absence
of CO2 (Philipp et al., 2022).

Figure 6. Percentage of U(VI) adsorbed on a compacted calcium bentonite (10 g/L). (Philipp et al., 2022)

Sorption experiments have demonstrated strong U(VI) retention on calcium-rich bentonites


at the pH range of 8 or above (Philipp et al., 2022). In Figure 7, calcium is represented by
green dots, uranium by blue dots, silicon by yellow dots, and oxygen by red dots. Calcium,

21
functioning as an exchangeable cation, can be replaced by the uranyl cation with an
equivalent charge at the montmorillonite edge sites. This phenomenon enhances the
sorption behavior, signifying that the presence of calcium plays a role in intensifying the
overall sorption characteristics. In the presence of calcium in a bentonite, it is known that
stable Ca-UO2-CO3-complexes are formed (Tran et al., 2018).

Figure 7. Demonstration of the effect of calcium in bentonite in U(VI) retention experiments. Color code: uranium (blue),
calcium (green), silicon (yellow) and oxygen (red). (Philipp et al., 2022)

A study conducted by Bachmaf et al., (2008) demonstrated that the efficient sorption of
U(VI) occurs at pH 8, with nearly complete sorption, approaching 100%, observed at this
particular pH level (Bachmaf, Planer-Friedrich and Merkel, 2008). At pH 8–9, uranium is
present as the U(VI) ion and at this pH ion exchange processes are relevant in addition to
surface complexation mechanism (Amphlett et al., 2020; Ye et al., 2022). Consistent with
expectations, the experimental setup in this thesis maintained in a similar pH level of
approximately 8. However, Philipp et al., (2022) reported that in the absence of carbonate,
the adsorption of U(VI) remains at approximately 90% at the same pH level. These findings
suggest that the presence or absence of carbonate can influence the sorption behavior of
U(VI) at pH 8 (Philipp et al., 2022). Further research is warranted to explore the
mechanisms underlying this difference and its implications for environmental remediation
strategies involving uranium-contaminated sites. Majdan et al., (2010) observed that the
sorption of U(VI) on modified bentonite, specifically the distribution coefficient Kd,
decreases as the percentage of mineral modification increases until it reaches a minimum at
76% of CEC, after which it increases again. The effective sorption of U(VI) occurred within

22
the pH range of 6–10 for the modified bentonite, and this behavior was attributed to the
sorption of U(VI) anionic hydroxy complexes. (Majdan et al., 2010)

At low concentrations, uranium is adsorbed to specific adsorption sites. With increasing


concentration, these sites are saturated, and the exchange sites are filled (Bachmaf, Planer-
Friedrich and Merkel, 2008; Yu et al., 2020). At high ionic strength (1 M), the equatorial
oxygen shell of uranyl is split, indicating inner-sphere binding to edge sites of
montmorillonite, primarily on exposed [Al(O,OH)6] sites from the octahedral sheet.
(Catalano and Brown, 2005). Figure 8 illustrates the relationship between uranium
concentration and the quantity of U(VI) adsorbed onto the compacted sodium smectite. The
results present relatively high adsorption percent at lower concentrations of uranium and
adsorption percent decreases with increasing uranium concentration in the solution. At low
initial uranium concentrations, the mobility of uranyl ions is high, which affects the U(VI)
adsorbed on Na-smectite (Korichi and Bensmaili, 2009).

Figure 8. The effect of U(VI) concentration on the U(VI) adsorption by Na-smectite. (Korichi and Bensmaili, 2009)

23
1.6.3 Cesium

Cesium (Cs) is an alkali metal with the atomic number 55 and a melting point of 28.5 °C. It
is therefore in liquid state around room temperature. Cesium ions, Cs+ have high solubility
and mobility in groundwater systems (Liu et al., 2019). Cs has only one common oxidation
state of +I and therefore most Cs compounds include Cs at this state. Cesium is highly
reactive and therefore an interesting metal occurring in radioactive waste.

Cesium isotopes, namely Cs-134 (t1/2=2.3 x 106 y) and Cs-137 (t½=30.07 y) are present in
high abundance. Cs-137 poses potential risks, particularly in the event of containment failure
due to its persistence. Cs-137 isotope decays to Ba-137m while emitting a strong beta ray.
The meta-stable Ba-137m is short lived and decays to stable non-radioactive Ba-137. If Cs-
137 enters the environment, it may enter the food chain, where it poses as a direct risk to
human health (Galamboš et al., 2010). However, Cs-135 (t1/2=2.3 x 106 y), despite its lower
abundance, presents a distinct hazard owing to its extended half-life, rendering it a
significant long-term risk. Therefore, it is essential to carefully assess and implement
measures for Cs-135 in radioactive waste handling procedures.

1.6.4 Cesium – bentonite chemistry

The chemistry between cesium and bentonite has been a subject of significant investigation,
with researchers delving into the complex interactions in between them. Understanding the
interactions between cesium and bentonite is crucial for elucidating its sorption behavior,
which in turn, holds implications for hazardous waste disposal strategies. The evaluation of
cesium adsorption on bentonite is crucial for assessing the long-term safety of radioactive
waste disposal, primarily due to the half-lives and mobility of Cs isotopes in the spent
nuclear fuel (SNF)(Liu et al., 2019). This thesis aims to investigate the fundamental
principles governing the chemical interactions between cesium and bentonite, while also
evaluating the potential application of bentonite in the disposal of spent nuclear fuel (SNF).

Similar to uranium, the adsorption of cesium onto bentonite is predominantly influenced by


ion-exchange mechanisms (Siroux et al., 2017). Bayülken et al., (2011) investigated and
modelled the adsorption of Cs on different type of clays (e.g., bentonite). In alignment with
investigations made by Siroux et al., (2017), the outcomes of this investigation similarly
underscore that the adsorption of Cs involves ion exchange. The results showed that

24
especially bentonite and zeolite can be used as effective barrier materials (Bayülken et al.,
2011). The capacity of clays for sorption through ionic exchange can be characterized by
cation exchange capacity (CEC) (Volkov et al., 2021).

Cesium (Cs) has the capacity to undergo sorption within the interlayer regions of bentonite
materials, similar to the behavior observed for the uranyl cation. Specifically, Cs can sorb at
three distinct sites: the planar, interlayer, and frayed edge sites (Lee et al., 2017). In fact, the
amount of FES on the mineral has been recognized as a key factor of the Cs environmental
dynamics in soils (Delvaux, Kruyts and Cremers, 2000). The adsorption of Cs is influenced
by charge density, as demonstrated by Lee et al. (2017). Cesium ions enter the interlayer
space of the mineral via ionic exchange, facilitated by the formation of Cs-O bonds within
montmorillonite's tetrahedral lattice (Volkov et al., 2021). This exchange occurs readily with
hydrated cations like K+, Na+, Ca2+ and Mg2+. Moreover, the replacement of silicon by
aluminum in the silica tetrahedron directly impacts cation exchange sites and ion exchange
potential. Additionally, the internal pore size of bentonite surfaces plays a crucial role in its
sorption behavior, as noted by Chen et al., (2023). The adsorption process on bentonite is
pH-dependent, as variations in pH can affect the speciation of cesium (Cs) ions in solution.
Elevated pH levels (pH > 6) have been found to enhance the presence of pH-dependent sites
on the bentonite surface, thereby facilitating a stronger attraction of Cs ions. This occurs
through the formation of inner-sphere complexes, along with other cations (Izosimova et
al., 2022).

Abou-Lilah et al., (2022) investigated the kinetics of Cs adsorption on bentonite across


different contact time intervals. Their findings revealed a rapid adsorption of Cs on
bentonite, particularly within the initial 5 to 60-minute period. Notably, approximately 81%
of Cs was adsorbed within the first 15 minutes, reaching a plateau of over 95% of the
equilibrium sorption capacity by the end of the 60-minute duration. (Abou-Lilah et al.,
2022). Experiments conducted by Wu et al., (2009) also showed efficient adsorption of Cs
on montmorillonite: the equilibrium which corresponded to the adsorption rate of 78.2%
could be attained after only 5 min of reaction time (Wu et al., 2009).

The impact of salinity on Cs sorption has been demonstrated with experiments where high
ionic strengths from monovalent Na+ or K+ -ions reduced Cs sorption onto micaceous
aquifer sediments under neutral pH (Tran et al., 2018). Also other cations present in the
reference solution, such as Ba2+, despite of being a divalent cation where Cs+ is monovalent,

25
decrease the adsorption of Cs (Missana et al., 2014; Seki et al., 2024)(Figure 9). Calcium
however has an opposite effect on the adsorption of Cs (Seki et al., 2024).

Figure 9. The effect of the presence of Ba on the sorption ratio and distribution coefficient of Cs on Mg-containing
calcium silicate hydrate. (Seki et al., 2024)

There has been limited investigation into the adsorption of Cs onto pure Na-
montmorillonite. However, Chen et al., (2023) investigated the sorption behavior of three
inorganic modified bentonites (Na-bentonite, K-bentonite and Mg-bentonite). In terms of
adsorption rate, Na-Bentonite showed relatively stable adsorption rate under the condition
of high initial concentration of Cs. The results indicated a maximum adsorption rate of
95.30% in an alkaline environment with Na-bentonite. (Chen et al., 2023)

Figure 10 (a) demonstrates the effect of the varying Cs concentration sorption onto two
different sorption sites of a bentonite. The black dots represent K-smectite, red dots
represent Ca-smectite and blue dots indicate Na-smectite. From the plottings, it can be
observed that the sorption behavior is mainly consistent with the existence of two different
sorption sites (Missana et al., 2014). The T1 sites stand for frayed edge sites (FES) and the
figure presents very low capacity of T1 and up to the saturation of these sites, sorption is
linear (Missana et al., 2014). Figure 10 (b) shows the same data represented as the logarithm
of the distribution coefficient (Log (Kd)). The Kd values are higher in T1 sites than in T2 sites,
which are the planar sites of the smectite. This observation is due to having lower capacity
but higher affinity for cesium (Missana et al., 2014). Type T2 sites on the contrary have lower
affinity but higher capacity for cesium, which could be explained with the existence of
interstratified mixed layers which means that the layers of the clay are stacked in various
ways forming new structures (Missana et al., 2014).

26
Figure 10. The effect of varying concentration in cesium retention onto Ca-, K- and Na-smectites. (Missana et al., 2014)

For bentonites with smectite composition, the sorption is expected to occur at planar sites
(T2 sites), where the exchangeable cations exist (Missana et al., 2014). The presence of large,
hydrated cations induces the expansion of clay layers which creates additional sorption sites.
The bentonites employed in this thesis predominantly comprise montmorillonite-smectite,
leading to anticipated similarities in sorption behavior. The interlayer cation (Ca2+, K+, or
Na+) is determined through XRD analysis, and the findings are elaborated upon in the
results section within the context of pre-characterization of the bentonites.

As the concentration increases, there is a concurrent reduction in the occupancy of sorption


sites. This is visually presented in Figure 10 as a descending trend. When the concentration
is low, the affinity of the mineral is higher and it’s capable of retarding more cations. This
observed relationship between concentration and sorption site occupancy suggests a
dynamic interaction between bentonite and radionuclide, Cs in this case. This observation
enhances the comprehension of the concentration-dependent dynamics of sorption sites.

27
2. Project aims and objectives
The overall objective of ABCRad is to provide physico-chemical data on the thermal effects
and chemical interactions between key, risk-driving radionuclides and two alternative
bentonite materials for use in the Onkalo repository. This thesis forms a collaborative part
of this research project and, specifically it aims to attain an understanding of the interactions
between two radionuclides and the two chosen bentonites suitable as alternative buffer and
backfill materials to bentonite form Wyoming for the safe storage of spent nuclear fuel in
Finland.

The two alternative bentonites were supplied by a Posiva Oy, and a detailed characterization
of these materials along with physico-chemical data elucidating their chemical interactions
with two risk-driving radionuclides (U and Cs) is presented. The aim is to first characterize
the bentonites using Powder X-ray Diffraction (XRD), Fourier Transform infrared
Spectroscopy (FTIR) and determine their Specific Surface Areas (SSA) and Cation Exchange
Capacities (CEC). The second aim is to determine the distribution coefficients (Kd), a ratio
that characterizes the partitioning of an element between the solution and the solid phase,
for the two radionuclides and two bentonites under a broad range of concentrations. This,
alongside the calculation of percent sorption, contributes to the formulation of sorption
isotherms for two bentonite candidates. Combining quantitative sorption data with
mechanistic understanding from pre-characterization results in a broader understanding of
retention capacities for U and Cs. The data helps to define possible issues on long-term
material stability and radionuclide transportation. The calculated parameters give
important information for the radiation safety authority, STUK and implementing
organization, Posiva Oy. The goal is to determine if the examined bentonites fulfill the
criteria for serving as safe materials capable of withstanding external conditions in a disposal
facility, potentially replacing the Na-Wyoming bentonite reference material.

3. Materials and methodology


3.1 Bentonite pre-characterization

Prior to the sorption isotherm experiments, comprehensive pre-characterization of the


bentonites was carried out. This pre-characterization involved Powder X-ray Diffraction
(XRD), Specific Surface Analysis (SSA), and Fourier Transform Infrared Spectroscopy

28
(FTIR). In addition to these analyses, determination of Cation Exchange Capacity (CEC) and
the examination of Exchangeable Cations (EC) were conducted.

3.1.1 Cation exchange capacity

Cation exchange capacity (CEC) is a parameter defined as the capability of a soil or clay to
absorb cations. It serves as a pivotal metric for determining mineral properties, including
their swelling capacity. The desorption of adsorbed cations can be influenced by competing
ions. This parameter is linked with the interlayer cations (Figure 11), the characteristics of
which are vital, as the composition of cations at exchange sites influences exchange
properties, plasticity, swelling capacity, and rheological behavior (Fernández et al., 2022).
For smectite, cation exchange is expected to occur mainly at the planar sites, where
exchangeable cations reside and these sites likely correspond to the total CEC of the clay
(Missana et al., 2014). It is expected that the mineralogical composition will have an effect
on the CEC value obtained (Choo and Bai, 2016). The CEC of a bentonite was expected to be
in the range of 87–97 meq/100 g at the highest, depending on the methodology (Dohrmann
et al., 2012; Fernández et al., 2022).

Figure 11. CEC between interlayers of smectite minerals. Tetrahedral-octahedral-tetrahedral (TOT) is representing the
configuration of the mineral. (Zeyen et al., 2022)

29
The CEC of clays was assessed through the preparation of triplicate samples. The
methodology, as described by Fernandez et al. (2022), employs cesium as the displacing
cation. A solution of CsCl was prepared by dissolving 21.045 mg of solid CsCl in 250 mL of
18 MQ deionized water. Subsequently, 1 g of solid air-dried sample was mixed with 4 mL of
0.5 M CsCl in a 1:4 solid to liquid ratio. After phase separation by centrifugation (2,522 g,
30 min), the supernatants of the solutions were filtered using a 0.2 µm pore size syringe
filter to prevent larger bentonite particles being present in the final samples. This step is
crucial for maintaining sample purity and avoiding potential interference from oversized
particles during analyses.

The concentration of the major and trace cations: Ba2+, Ca2+, Cs+, K+, Mg2+, Na+, Sr2+ and
were analyzed with Microwave Plasma Atomic Emission Spectroscopy (MP-AES). The
samples underwent a tenfold (x10) dilution for the MP-AES measurement. The CEC results
for the bentonites are detailed in the subsequent section (4.1.1). The results provide valuable
information on bentonite’s cation retention capabilities and overall suitability for specific
applications. The CEC (meq/100 g) is calculated with the following equation:

𝑐𝑉𝑧
𝐶𝐸𝐶 =
𝑚𝑀
Equation 2. Equation representing cation exchange capacity (CEC) calculation.

In Equation 2, c is the concentration of the exchangeable cation (g/mL), V is the volume of


the water (mL), Z is the charge of the ion, m is the mass of bentonite (g), and M is the
molecular weight of the cation (g/mol).

3.1.2 The exchangeable cations

The determination of exchangeable cations (EC) in addition to the Cation Exchange Capacity
(CEC) of bentonites, despite employing a similar methodology described by Fernandez et al.
(2022), offers a more comprehensive understanding of the bentonite’s ion exchange
dynamics. While CEC focuses on the total capacity for cation exchange, assessing EC
provides information of the types and quantities of cations that are available for exchange
within the clay structure. This dual analysis enhances the knowledge of the clay’s sorption
properties.

30
The exchangeable cation population was determined through a CsCl method (Fernández et
al., 2022). In the experimental procedure, 0.5 g of solid, air-dried samples of Laviosa and
LMS were manually mixed with 2 mL of 0.5 M CsCl for a duration of 3 minutes. After phase
separation by centrifugation (2,522 g, 30 min), the supernatants of the solutions were
filtered through 0.2 µm pore size syringe filter, and the concentration of the major cations
were analyzed using MP-AES. The results are presented in section (4.1.1).

3.1.3 X-ray diffraction

X-ray diffraction (XRD) analyses were conducted to determine the bulk mineral
composition of the bentonites, utilizing a Panalytical X’Pert3 Powder X-ray diffractometer
with CoKα-radiation (𝜆 = 1.5406 Å, 40 kV, 40 mA). These analyses facilitated not only the
identification of bulk minerals within the bentonites but also the prediction of interlayer
cations. The measurement range spanned from 5° to 75° 2θ, using a step size of 0.02° 2θ
during the analysis. The samples were prepared as random powders without any specific
particle orientations by being placed on glass slides. Bragg’s law was employed to calculate
the interlayer spacing (d) using Equation 3, where λ represents the wavelength of the
incident X-ray and θ is the angle of diffraction. Mineral identification was conducted using
HighScore Plus software.

2𝑑𝑠𝑖𝑛𝛳 = 𝑛𝜆
Equation 3. Equation depicting Bragg’s law.

3.1.4 The zeta potential

The electrokinetic behavior of a particle, i.e., the zeta potential (ZP), can give a valuable
indication of surface charge state, which explains why the zeta potential is a decisive
parameter regarding understanding and modelling of particle-particle interactions (Lowke
and Gehlen, 2017). Measuring ZP can reveal possible rheological differences between
bentonite types. The ZP is the charge developing at the space between the solid surface and
the liquid. The potential developed in the interface is highly dependent on pH. High ZP
correlates to highly charged particles, which prevents aggregation of the particles due to
electric repulsion and if the ZP is low, attraction overcomes repulsion, and it is likely that

31
the mixture forms coagulates (Samimi et al., 2019). The liquid layer surrounding the particle
exists as inner region, where the ions are strongly bound and an outer, diffuse region where
the ions are less bounded. When a particle moves, ions within the boundary move it and
those ions beyond the boundary stay with the bulk dispersant. The potential at this boundary
is called the zeta potential.

Şans et al., (2017) observed variations in the zeta potentials of Na+ and Ca2+ saturated
bentonites. The zeta potentials of compacted sodium bentonites exhibited more negativity
compared to Ca-bentonites. Yu et al., (2020) through batch experiments, determined that a
ZP value within the range of -30 mV to -50 mV indicated robust stability of the
montmorillonite colloid in a pH range of 6.0–6.5. These findings highlight the importance
of zeta potential in understanding the stability of bentonite colloids under different ionic
compositions and pH conditions.

A pH series was prepared for Laviosa, LMS, and Na-Wyoming bentonites. The solid
specimens (0.5 g) were mixed with 10 cm3 of reference water (section 3.2.1) in polyethylene
centrifuge tubes taken into glove box for degassing overnight. The pH series was conducted
within the controlled environment of the glove box, imitating repository conditions. The pH
levels were precisely adjusted within the range of 4 to 13 for the bentonite series, through
the addition of either 0.1 M HCl, 0.1 M NaOH, or 6 M NaOH. The adjustment of pH for each
sample within the three series spanned a duration of three days to allow for equilibration to
the desired values. By systematically adjusting and equilibrating the pH of the clay samples,
these steps serve to create a well-defined experimental environment for subsequent zeta
potential measurements. This allows for precise determination of surface charge properties.
The zeta potential was measured for Laviosa, LMS and Na-Wyoming bentonite samples
using a Zetasizer Nano ZS (Malvern Instruments Ltd.).

3.1.5 BET

Surface area is an important property of porous materials such as bentonites. The Brunauer-
Emmett-Teller (BET) analysis is an analytical technique, where the specific surface area
(SSA) and porosity distribution of a solid material is determined. The BET model is a
generalization of the Langmuir's theory of multilayer adsorption of pure gases (Saberi and
Rouhi, 2021). The method is based on adsorption isotherms of inert gases such as argon or
nitrogen (Saberi and Rouhi, 2021; Zou et al., 2021). The unreactive gas flows over the solid

32
sample. The molecules of the gas adsorb onto the surface of the solid sample through weak
van der Waals forces. The amount of gas adsorbed onto the solid is directly proportional to
the SSA of the sample.

Measurements were performed using a Quantachrome Autosorb iQ Station 3 with ~0.5 g of


bentonite material. Standard pre-treatment conditions in the measurement were 110 °C and
9 h. The surface areas of the bentonites determined using BET analysis are presented in
(Table 4).

3.1.6 FTIR

Fourier Transform infrared Spectroscopy (FTIR) is an analytical technique which uses


infrared (IR) light to scan a sample and to detect chemical properties such as functional
groups. It is used to identify unknown substances and is well suited to analysis of organic
materials. In FTIR analysis, the sample is exposed to IR-radiation after which the bonds in
the molecules of the sample absorb at characteristic wavelengths. The amount of light
absorbed and transmitted through the sample is detected. The raw data is transformed to an
IR-spectrum with a mathematical formula. The FTIR application has many advantages over
IR-spectrometer such as high speed which makes it a more efficient method in analyzing the
chemical components of the sample.

3.1.7 Gamma HIDEX

The determination of cesium (Cs-134) activity in supernatants after interaction with


bentonite and centrifugation was conducted using a gamma (HIDEX) counter. Following
centrifugation, 5 mL of the supernatant was transferred into a vial, and samples were
measured for 30 minutes each using the gamma (HIDEX) counter. The resulting Cs-134
activity value obtained from these measurements was employed in the calculation of the
distribution coefficient (Kd) and percent sorption.

This instrument was chosen for Cs activity assessment due to cesium’s gamma radiation
emission, making its activity detectable. The gamma counter consists of NaI detector,
scintillation crystal, and a photomultiplier. In the presence of radiation within the sample,
emitted photons traverse a crystal, leading to their absorption and subsequent light
production, which is quantified by a detector. For samples containing Cs, a measurement

33
duration of 30 minutes was used based on the sorption kinetics that were determined via
time interval experiments. Given that uranium emits only weak gamma rays, the activity of
samples containing U was determined using a liquid scintillation counter (LSC), a method
which will be introduced next.

3.1.8 Liquid scintillation counter

In liquid scintillation counting (LSC), energy is absorbed in the counter system, creating
excited states in electrons. These excited states eventually return to their normal state,
emitting light pulses unique to the scintillator. The LSC cocktail, an organic compound,
soaks up the emitted energy, passing it on to a scintillator molecule. This molecule then
releases the energy as light, which is transformed into an electric pulse for detection.

Following the interaction with bentonite and subsequent centrifugation, 5 mL of


supernatants were combined with 15 mL of LSC cocktail, Ultima Gold AB (Perkin Elmer), in
a vial. Subsequently, the samples were equilibrated in the dark overnight after the addition
of 0.1 mL of 1 M hydrochloric acid, following thorough mixing. The radioactivity of U-233 in
the samples was measured using a Tri-Carb2910 TR liquid scintillation analyzer (Perkin
Elmer). Each sample containing U was measured for 3 hours. The counter consists of an
autosampler, where the samples are measured in a light tight chamber. The chamber’s walls
are made of aluminum mirrors painted with titanium oxide. The scintillations of light that
the sample emits are collected to a photon multiplier detector’s photocathode. Liquid
scintillation counting is used to measure low beta energy.

3.1.9 MP-AES

Microwave Plasma Atomic Emission Spectroscopy (MP-AES) was used to determine the
quantity of elements in the samples. In principle, the instrument uses microwave energy to
produce a plasma discharge. An aerosol is created from the liquid sample at high
temperature. The aerosol is then introduced into the center of the hot plasma after which, it
dries, atomizes and excites. The atoms emit light at wavelengths characteristic for each
element when they return to lower energy levels.

An agilent MP-AES 4200 instrument was used in this project to measure the concentrations
of bulk cations. The cations measured with MP-AES were Na+, K+, Mg2+, Ca2+, Sr2+ and Ba2+.

34
The samples were prepared by acidifying the sample and adding an internal standard,
scandium (Sc2+) was chosen as the internal standard to compensate for variations in sample
introduction, atomization efficiency, and instrumental drift, thereby enhancing the accuracy
and precision of the elemental analysis. The aqueous samples were diluted in analytical
grade HNO3 (5% v/v, 100x dilutions for most cations and a 1000x dilution for barium cation)
for MP-AES.

3.2 Sorption experiments

The experiments were conducted under site-specific conditions, simulating the repository
environment of the Onkalo repository, through the utilization of the batch technique in
polyethylene centrifuge tubes. The sorption experiments were done in a glove box including
inert gas N2. All bentonite samples and ONKALO water were prepared the day prior to the
experiment and taken into the glove box to degas overnight. All solutions needed in the
experiment were prepared using 18 MQ water and were degassed before importing them
into the glovebox. The pipettes used for spiking with radiotracers underwent calibration
prior to the experiments. The calibration ensured the accuracy and precision of volume
measurements, which is essential for maintaining the reliability of experimental results.

3.2.1 Reference water

To get realistic results, preparations were done before sorption isotherm experiments.
ONKALO water was used to mimic the flowing groundwater in the repository area in the
Onkalo geological repository in Olkiluoto. The recipe for 1 L of saline ground water used in
the experiments is described next.
Table 3. Composition of 1 L of 0.365 M reference water (ONKALO water) utilized in experimental procedures.

Salt Mass (g) Molarity (M)

NaCl 6.4739 0.111

CaCl2 · 5 H2O 3.5261 0.320

35
The ONKALO water was degassed to eliminate atmospheric oxygen from the solution. This
was achieved by introducing an inert gas flow of either nitrogen or argon into the solution
with the duration adjusted based on the volume of the solution. By eliminating atmospheric
oxygen and carbon dioxide, potential unwanted reactions were minimized.

3.3 Sorption isotherms

The chemical interactions between alternative buffer materials and radionuclides lack
comprehensive investigation when compared to Na-Wyoming bentonite (e.g., MX-80).
Sorption isotherms serve as predictive models for the interactions between radionuclides
and the surfaces of bentonite materials but are specific to the conditions under which they
were applied. As the groundwater chemistry and bentonite composition can vary widely,
sorption isotherms must be carried out under representative conditions. These data provide
safety authorities with direct and valuable insights, aiding in the assessment of whether a
given buffer material aligns with established safety limits.

3.3.1 Sorption kinetics

Kinetic experiments were conducted to evaluate how long it takes for the radionuclides,
U(VI) and Cs to reach the equilibrium in the solution including ONKALO water and either
Laviosa or LMS type bentonite. Triplicates of each 0.5 g of bentonite in 10 cm3 of ONKALO
water were put into a mechanical shaker in a glove box with a given concentration (10-3 M)
of uranium and cesium in time intervals: 0.5 h, 2 h, 16 h, 24 h and 72 h at a fixed temperature
(293 ± 3 K). After contact time, the samples were centrifuged at 2,522 g for 30 min.
Following each experimental run, the samples were weighed in order to precisely
determinate the added reference water's exact volume. This information is crucial for
subsequent calculations of parameters, including the distribution coefficient (Kd) and
percent sorption. The activity of Cs was assessed using a HIDEX gamma spectrometer and
activity of U was determined using LSC. By utilizing the determined activity counts, time
intervals, and calculated percent sorption values, the optimal reaction time for each
radionuclide could be selected for subsequent experiments where the radionuclide
concentration was changed.

36
3.3.2 Varying concentration

The experiments involving varying radionuclide concentrations encompassed concentration


ranges of 10-3 to 10-10 M for Cs and 10-2 to 10-10 M for U. Cesium chloride (CsCl) and uranyl
nitrate (UO2(NO3)2) served as the selected tracers, and solutions were prepared by diluting
a stock solution of higher activity with ONKALO water to achieve lower concentrations. The
samples were then spiked with a known concentration of a mixture containing stable and
radioactive cesium or uranium. Conducted in triplicate, these experiments followed a similar
protocol to the varying time experiments, with subsequent analysis of the supernatants
utilizing the same instruments.

3.3.3 The distribution coefficient

The distribution coefficient, Kd is a ratio value which describes the distribution of the
dissolved element between solution and solid phase. It can be determined when the mass of
the bentonite, the volume of added solution and the activity counts of the radionuclide are
known. This value not only describes the partitioning between the phases but also surface
complexation and ion-exchange (Payne et al., 2013). The distribution coefficient Kd of both
uranium and cesium have been found to decrease with increasing bentonite colloid
concentration (Tran et al., 2018).

The distribution coefficient was calculated from the following equation:

𝐴" − 𝐴# 𝑊
𝐾! = ∗
𝐴# 𝑉

Equation 4. The distribution coefficient equation. (Khan et al., 1994)

In Equation 4, A0 is the initial activity of the solution, A1 is the activity of the solution after
the reaction has reached equilibrium, W is the mass of bentonite, and V is the volume of the
solution.

The uncertainties in the Kd value during batch experiments can arise from factors such as
variations in phase separation efficiency and mineralogical heterogeneity in the samples
(Vejsada, 2006). These uncertainties were mitigated by subjecting the samples to a 30-

37
minute centrifugation process and carefully pipetting the supernatant. Additionally, the
solid samples underwent sieving through a 200 µm sieve to exclude particles larger than the
specified size, thus preventing their interference with the experimental outcomes.

3.3.4 The percent sorption

The percent sorption serves as a critical parameter reflecting the clay material's capability to
adsorb radioactive elements. This key metric describes how negatively charged sorption sites
on the bentonite surfaces accommodate and immobilize radionuclides present in the
solution. As the concentration of the solution increases, more sorption sites become
occupied, influencing the percent sorption observed. Understanding this sorption behavior
is pivotal for assessing the effectiveness of bentonite as a barrier material.

The percent sorption is expected to be approximately 90% under room temperature and 10-
2 M Cs concentration (Ali Khan, Riaz-ur-Rehman and Ali Khan, 1994). This factor can be
calculated from the following equation when Kd value is known:

100 ∗ 𝐾!
𝑃=
𝑉
𝐾! + 𝑊

Equation 5. Percent sorption equation. (Khan et al., 1994)

In Equation 5, Kd is the distribution coefficient (Equation 4), V is the volume of the solution,
and W is the mass of bentonite.

3.4 pH

In the context of the experimental protocol of this thesis, no pH adjustments were done
during the formulation of sorption isotherms. The pH measurements were conducted on
individual samples of 0.5 g of bentonite in reference water after reactions and subsequent
centrifugation, utilizing the Orion 5 Star (Thermo Scientific) multifunction meter.

38
4. Results and discussion
4.1 Results of sorbent pre-characterization

The bentonite materials were provided by Posiva Oy for characterization and further
experiments. Laviosa bentonite originates from Italy, LMS from Georgia and the Na-
Wyoming reference material (Barakade) comes from the USA. Their main minerals as
identified by XRD, and surface areas measured by BET are presented in Table 4. Na-
Wyoming type bentonite is the current reference material for most geological disposal
facilities. It is used as a reference for the two alternatives investigated in this project.

Table 4. Properties of the three bentonites. XRD analysis results are presented in Figure 13.

Bentonite Country of Surface area Main minerals


2
origin (m /g)

Laviosa Italy 33.165 Montmorillonite, Feldspar, Quartz,


Cristobalite
LMS Georgia 76.157 Montmorillonite, Feldspar, Mica,
Zeolinite
Na-Wyoming USA (Wyoming) 29.862 Montmorillonite, Calcite, Quartz,
Feldspar

4.1.1 The CEC and exchangeable cations

The sorption of U(VI) and Cs on bentonite can be affected by a variation in cations, and the
number of exchangeable cations determines the CEC of the bentonite. Laviosa favors cations
such as Na+, Mg2+, and Ba2+ whereas LMS, respectively shows a higher content of Ca2+ and
Sr2+ cations as well as Na+ cations.

The CEC for Laviosa, LMS and Na-Wyoming was 87 (±0,048) meq/100 g, 95 (±0,34)
meq/100 g, and 91 (± 1,23) meq/100 g, respectively. The results are consistent with the
literature where the CEC of bentonites is generally in the range of of 87–97 meq/100 g
(Dohrmann et al., 2012; Fernández et al., 2022)(Figure 12). Na-Wyoming (Barakade)
bentonite serves as a reference material for the alternative bentonites. Figure 12 illustrates

39
that the cation exchange capacities (CEC) of Laviosa and LMS are greater than those of the
Na-Wyoming type bentonite commonly utilized in the industry and disposal facilities.

Figure 12. Cation exchange population (meq/100 g) of Barakade (Na-Wyoming), Laviosa and LMS bentonites. The error
bars signify one standard deviation from triplicate samples.

4.1.2 XRD

The XRD patterns of Laviosa, LMS and Barakade bentonites are shown in Figure 13. The
strongest peaks of montmorillonite, feldspar and quartz are indicated. Other minerals such
as cristobalite, mica, zeolinite and calcite were also detected (Table 4). The interlayer spacing
on the bentonite can be seen in Figure 13 as the first peak for each bentonite. The position
of the first peak was determined by reading 2q value with Origin program, which was then
converted to Ångstroms using Equation 3, with a CuKa wavelength (l) value of 1.5406 Å.

40
Figure 13. XRD analysis results for Na-Wyoming and alternative bentonites LMS and Laviosa.

For LMS, Laviosa and Na-Wyoming the distances are 12.5 Å, 12.6 Å, and 12.1 Å, respectively.
The interlayer cation could be predicted from the results. Since 12.4 Å is a typical
characteristic peak for sodium bentonite, the results reveal that the predominant
exchangeable cation is sodium (Moma, Baloyi and Ntho, 2018; Salah, Gaber and Kandil,
2019). The interlayer cation is therefore mostly Na+ with Ca2+ content, and Figure 12 shows
the molar equivalence. The actual proportion moles of Ca2+ can be obtained by dividing the
number by its charge (two). The CEC value is presented in molar equivalents so by doing
this, the values of Na+ and Ca2+ are comparable.

4.1.3 The zeta potential

In order to assess the stability of the suspensions and surface charge of particles, zeta
potential measurement was completed. The zeta potential measurement results for Laviosa,
LMS, and Na-Wyoming bentonites are shown in Figure 14. The values reflect the stability of
the bentonite solution. The determination of the zero point charge (isoelectric point),
representing the pH at which the surface achieves net electrical neutrality, was undertaken.

41
The surface charge significantly influences the adsorption of ionizable pollutants onto the
catalyst from aqueous solutions (Moma, Baloyi and Ntho, 2018). The isoelectric point was
observed at pH 12, 11.5, and 12 for Laviosa, LMS, and Na-Wyoming, respectively.

Figure 14. The zeta potential of Laviosa, LMS and Na-Wyoming bentonites. The error bars signify one standard
deviation from triplicate samples.

Figure 14 shows how increasing the pH creates more negative charge. The ZP for LMS shows
a decreasing trend when pH is increased, whereas Na-Wyoming bentonite has relatively low
ZP values in between pH 7.5-8.5. The decrease in negative ZP reflects to pH dependent
charge (Moma, Baloyi and Ntho, 2018). The pH of the sorption isotherm experiments done
in this project were between 8-9 for Laviosa and LMS. The ZP values for these bentonites
are between -11 mV and +6 mV, respectively.

According to investigations done by Moma et al., (2018), and the results observed in this
study, all bentonites displayed mainly negative zeta potential values over the entire pH
range. The negative zeta potential values of natural bentonite clay could be due to permanent
negative charge of the isomorphous substitution of structural Si4+ by Al3+ and to the OH
group charge, that are pH dependent at the edges of the bentonite clay particles (Moma,
Baloyi and Ntho, 2018).

The bentonite’s zeta potential plays an important role in its versatile applications. The
repulsive force between particles ensures stability between particles by preventing

42
aggregation. The negative ZP is also linked to bentonite’s exceptional swelling capacity when
exposed to water. In addition to these properties, the negative charge attracts cations and by
this, enhances bentonite’s effectiveness as a buffer in a safe disposal system.

4.1.4 SSA

The specific surface areas (SSA) of the bentonites were determined using BET analysis. The
results are presented in Table 4. Laviosa, LMS and Na-Wyoming type bentonites the SSA
determined were 33.165 m2/g, 76.157 m2/g, and 29.862 m2/g, respectively. A larger surface
area provides more available sites for sorption, and it leads to increased potential for ion
exchange process. LMS is expected to have good sorption capability due to its larger SSA.

4.1.5 FTIR

In order to assess the functional groups present on the bentonite with the capacities to bind
radionuclides, the bentonites were characterized using FTIR spectroscopy. The results are
shown in Figure 15. Yu et al., (2020) conducted FTIR spectroscopy on the montmorillonite
colloid both pre- and post-adsorption of U(VI). Despite the adsorption process, minimal
alterations were observed in the spectral profile. The spectrum exhibited persistent broad
bands at 3625 and 3436 cm-1, denoting vibrations linked to free water molecules adsorbed
onto the solid surface, along with distinct bands at 1640 cm-1 and 1044 cm-1 (Yu et al., 2020).
These are due to bending of H-OH bond which are retained in the matrix and Si-O-Si group
in the tetrahedral sheet, respectively (Yu et al., 2020; Abou-Lilah et al., 2022; Fernández et
al., 2022). Based on these observations, the broad bands observed in ~3618 cm-1 (3600-3750
cm-1) in the FTIR spectrum of Na-Wyoming, Laviosa, and LMS, are most likely due to H-OH
vibrations (Yu et al., 2020; Fernández et al., 2022). The strong band at approximately 1012
cm-1 corresponds to Si-O-Si groups in the tetrahedral sheet of montmorillonite (Yu et al.,
2020). In this band, the occupancy of the sheet can be distinguished due to each cation
strongly influencing the position of the OH bending bands, which arise from vibrations of
the inner and surface OH groups (Fernández et al., 2022). According to the FTIR analysis
done by Yu et el. 2020, the peak at ~1639 cm-1 indicates the retention of water molecules H-
OH bond after the adsorption (Yu et al., 2020).

43
Figure 15. The FTIR spectrum of Na-Wyoming, Laviosa and LMS.

The FTIR spectra of Na-Wyoming, Laviosa, and LMS type bentonites exhibit intensity
variations influenced by diverse factors. These include the mineral composition, chemical
impurities, particle size distribution, and cation exchange capacity (CEC). Such factors
modulate the detected peak intensities by influencing the vibrational modes, resulting in
either amplification or attenuation of the peaks. Higher transmittance values correlate
with reduced absorption, indicative of potentially higher bentonite purity. Across the
analyzed bentonite samples, the transmittance diminishes in the sequence of Laviosa,
LMS, and Na-Wyoming.

4.2 The sorption isotherms

4.2.1 Reaction kinetics

The primary objective was to determine the optimal duration for the subsequent
experiment. Selecting an appropriate shaking duration is essential as it ensures the stability
of equilibrium in the reaction. This was done in following time intervals: 30 minutes, 2
hours, 16 hours, 24 hours and 72 hours. Figure 16 shows how the reaction time effects the
percent sorption. The percent sorption as a function of contact time for uranium is on the

44
left and for cesium on the right. It was expected that the adsorption of uranium is rapid
during the first few hours after which the equilibrium is reached (Yu et al., 2020). The
percent sorption increases until 72 hours of contact time and this time was chosen as a
suitable contact time for the varying uranium concentration experiments. This observation
deviated from prior research, where the attainment of equilibrium with uranium adsorption
onto bentonite or colloid typically occurred within approximately one hour (Youssef, 2017;
Yu et al., 2020). The prolonged uranium adsorption duration in this study compared to prior
research may be due to various factors, including differences in experimental conditions and
adsorbent characteristics. Variations in U(VI) concentration, pH and surface area can
influence adsorption kinetics.

The adsorption was fast for Cs in the first time periods and reaction reached equilibrium at
approximately 24 hours. For cesium, this was the suitable contact time. This contact time
goes well together with a study (Liu et al., 2019) where the equilibrium was reached after 24
hours of reaction time. Liu et al. (2019) study’s results showed that the adsorption ratio of
cesium on bentonite reached above 90%. The maximal percent sorption recorded for
Laviosa reached 89.8%, while LMS exhibited a peak sorption of 86.6% at a cesium
concentration of 10-3 M. In the case of uranium at a concentration of 10-3 M, Laviosa
demonstrated nearly complete sorption, approaching 100%, whereas LMS exhibited a
maximal sorption percentage of approximately 85%. The observed kinetics of the reaction
align with prior research, indicating that the adsorption of Cs reaches equilibrium more
rapidly than that of U with both bentonites.

45
Figure 16. The effect of shaking time for U (left) and Cs (right).

4.2.2 The distribution coefficient

The distribution of a radionuclide between bentonite and the reference water was
investigated in varying U and Cs concentration experiments. The variation of the
concentration helped to elucidate how the adsorption behavior of two bentonites changes
under different U and Cs concentrations. By altering the concentration, trends in sorption
efficiency can be observed. Overall this leads to providing valuable information for
understanding the interaction between radionuclides and clay materials. Notably, the
observed dependency of the distribution coefficient (Kd) on concentration was a distinctive
feature for both studied radionuclides and the results are presented in Figure 17.

The experimental setup involved a range of concentrations spanning from 10-10 M to 10-2 M
for uranium and 10-10 M to 10-3 M for cesium. The experiment involving uranium revealed
Kd values in the range of 130–135 cm3/g with Laviosa and 78–110 cm3/g with LMS. In the
concentration range of 10-10 M to 10-6 M, the activity of the U on the solids, after sorption,
was below the detection limit of the LSC used. This meant that the activity counts could not
be determined, and therefore Kd could not be calculated in this concentration range for U.

46
In Figure 17, we assume the percent sorption of uranium, assessed through liquid
scintillation counting (LSC), to be 100% within the concentration bracket of 10-10 M to 10-6
M. N. Megouda et al., (2007) demonstrated that the Kd values at equilibrium of natural
uranium ranged from 30 to 600 cm3/g and 50 to 1100 cm3/g when utilizing natural
bentonite and drilling bentonite, respectively (Megouda et al., 2007). These findings align
with the results reported in the referenced study.

Figure 17. The distribution coefficient for U (left) and Cs (right) as a function of concentration.

In Figure 17 on the right graph, the distribution coefficient is presented as a function of Cs


concentration, showcasing the relationship between the two. The distribution coefficient
values of Cs ranged from 130 to 280 cm3/g with Laviosa, while with LMS, the range was 150
to 425 cm3/g. Correspondingly, Khan et al., (1994) reported Kd values of approximately 150
to 1100 cm3/g for Cs adsorption on bentonite, with concentrations ranging from decreasing
concentration 10-2 M to 10-7 M, respectively and a solution/bentonite ratio of 40:1 (Ali Khan,
Riaz-ur-Rehman and Ali Khan, 1994). The findings are consistent with the results obtained
in this thesis. The data presented in Figure 17 implies that the higher specific surface area
(SSA) of LMS correlates with an increased presence of frayed edge sites (FES), which are
sites where Cs ions tend to bind. The frayed edge sites (FES) of the bentonite, characterized
by their heightened reactivity exhibit saturation before the non-specific and planar sites
(Delvaux, Kruyts and Cremers, 2000; Lee et al., 2017). This phenomenon is indicative of the
preferential adsorption of Cs onto the more reactive FES. As the FES reach saturation,
subsequent Cs sorption transitions to the non-specific and planar sites. Here, the other
cations can compete and this results to the reduce of the total Cs adsorption process. This

47
behavior can be seen in Figure 17, where the Kd value decreases similarly for both bentonites
at Cs concentrations above 10-6 M. This behavior was expected based on the study made by
Tran et al., (2018) and Fuller et al., (2014).

The LMS type bentonite exhibits a relatively high specific surface area of 76.157 m2/g in
contrast to Laviosa and Na-Wyoming variants. This characteristic may facilitate enhanced
uranium sorption, as supported by prior studies (Pabalan and Turner, 1996; Barnett et al.,
2000; Korichi and Bensmaili, 2009; Majdan et al., 2010). The findings align with this
notion, indicating a noticeable disparity between Laviosa and LMS in cesium sorption, with
the latter one demonstrating a larger distribution coefficient.

4.2.3 The percent sorption

The correlation between radionuclide concentration and the percent sorption was examined
to assess the efficacy of the buffer material in retaining uranium and cesium. The results are
presented in Figure 18. In the case of uranium sorption, the bentonites exhibit a percent
sorption range of 87–100%, with 100% sorption observed at lower uranium concentrations
and a gradual decrease to 80–90% at higher concentrations. Laviosa and LMS demonstrate
percent sorption of around 95% for cesium within the concentration range of 10-10 to 10-6 M.
At higher concentrations up to 10-2 M, the percent sorption decreases to a range of 85–95%
for both bentonites. Overall, a notable uptake of uranium was observed on both bentonites,
Laviosa and LMS. In the lower uranium concentration ([U] < 10-6 M) supernatants, no
counts were observed that could be distinguished from background counts using liquid
scintillation counting (LSC). The measurement results show efficient percent sorption for
uranium in concentrations of 10-6 and lower, reaching 100%. However, as the concentration
of both uranium and cesium increases in the solution, there is a corresponding decrease in
the percent sorption. These data compare well with relevant literature for U(VI) adsorption
to bentonite done by Yu et al., (2020) and Tran et al., (2018).

48
Figure 18. The percent sorption for U (left) and Cs (right) as a function of concentration.

The percent sorption remains stable with cesium concentrations from 10-10 M until 10-6 M
(Figure 18). At lower concentrations, the bentonites adsorb Cs more efficiently than in
higher concentrations. This is due to the saturation of the sorption sites. Previous studies
have shown the connection between the adsorbents CEC and radionuclide uptake. Missana
et al., (2014) presented that cesium uptake has been largely dependent on the CEC of the
clay in a study where the cesium retention onto Na-, K- and Ca-smectite minerals were
studied (Missana et al., 2014). The experimental findings in this thesis demonstrate similar
patterns. Laviosa, demonstrating a greater affinity for Ba2+ cations in the cation exchange
capacity (CEC) experiment, displayed a reduced percent sorption in the adsorption of
cesium when compared to LMS, which exhibited a lower preference for Ba2+ cations. This
observation aligns with the findings discussed earlier by Seki et al., (2024). As seen in Figure
18, the percent sorption of cesium (10-3 M) onto bentonite is between 80–90%. The results
go well with the observations Khan et al., (1994) discovered. The main composition of both
bentonites in this study is montmorillonite, which is a subcategory of smectite so similar
results were expected. The expectation was that the material would exhibit nearly 100%
sorption due to its high CEC, facilitating effective Cs adsorption. The CEC values determined
for Laviosa and LMS were relatively high: 95 meq/100 g and 91 meq/100 g, respectively
(Figure 12) so the observed results support this expectation.

Laviosa and LMS represent smectite clay minerals, containing 85 wt.-% and 84 wt.-% of
montmorillonite, respectively. This composition contributes to the observed high percent
sorption outcomes. Notably, Laviosa, characterized by a greater smectite content, exhibits

49
elevated percent sorption levels, as evidenced in experiments involving varied uranium
concentrations (Figure 18). Although LMS possesses a larger specific surface area (SSA), it
did not exhibit enhanced adsorption of U(VI) compared to Laviosa. The high percent
sorption observed for both bentonites, along with their effective adsorption of U and Cs,
reinforces their suitability for retaining these radionuclides from the surrounding
environment.

4.3 The effect of pH

The pH of the solutions in experiments involving varying radionuclide concentrations were


analyzed and the results are presented in Figure 19. The assessment of pH is crucial for
validating assumptions related to the adsorption mechanisms. Previous studies have
identified the effective sorption of uranium within the pH range of 6–10 (Philipp et al.,
2019). In the current thesis experiments, the pH of the solutions was maintained within the
range of 8-9, aligning with conditions conducive to similar results. The pH was set as the
natural equilibrium between the bentonite and the Onkalo reference water.

Figure 19. The pH of varying U and Cs concentration experiment was measured.

The pH of the supernatant from both bentonite samples remained stable within the range of
8.23–9.02 across uranium concentrations spanning from 10-10 to 10-3 M. Notably, pH
stabilization was observed in solutions containing 10-5 M uranium for both bentonites. In
the context of U(VI) adsorption onto montmorillonite colloids at low pH, cation exchange at

50
the negatively charged surface has been identified as a significant mechanism and at higher
pH levels, complexation emerges as the predominant process (Yu et al., 2020). The
effectivity of the adsorption of uranium can be explained with the consequence of anionic
hydroxyl complexes sorption of modified bentonites such as attaching large organic cations
to their surface (Majdan et al., 2010). However, when increasing the pH, U(VI) complexes
can dominate solution speciation. (Majdan et al., 2010; Saleh et al., 2018)

In the experiments conducted with cesium (Cs) concentrations ranging from 10-10 to 10-3 M,
the pH remained within the range of 8.20–8.92 (Figure 19). This observation is significant
when considering Cs solution concentrations exceeding 10-6 M, as the pH of the system plays
a crucial role due to the multi-site nature of Cs sorption (Fuller et al., 2014; Lee et al., 2017).
Both the distribution coefficient and percent sorption align with the trend observed in the
dataset. This indicates that change in pH in the measurement did not impact the sorption of
Cs at detectable levels. Notably, frayed edge sites exhibit a selective sorption of Cs,
demonstrating a preference for Cs over other solution cations (Lee et al., 2017). Additionally,
the elevation of pH levels (pH > 6) has been identified as a factor that enhances the pH-
dependent sites on the surface of bentonite. This enhancement results in a more secure
attraction of Cs ions, forming inner-sphere complexes (Izosimova et al., 2022). Despite not
adjusting the pH, the pH levels observed (8-9) align with the findings of studies mentioned
in section (1.6.4).

5. Conclusions and future work


5.1 Conclusion remarks
The aim of this thesis was to assess the interactions between the bentonites and risk-driving
radionuclides (Cs and U) under conditions relevant to the Onkalo repository. This thesis
provides batch technique results, theoretical calculations and spectroscopic analytical tools.
The sorption isotherms quantify the sorption capacity of both radionuclides and conclusions
for the mechanisms for radionuclide retention at molecular scale. Bentonites as a buffer in
an EBS should meet criteria involving long-term material stability and this project helps
defining issues in radionuclide transport. The parameters determined inform Posiva Oy
directly in the context of assessing the performance of the alternative bentonites in site-
specific conditions relevant to Finnish SNF disposal.

51
The study contributes valuable insights into the interactions between radionuclides and
bentonites, focusing on Laviosa and LMS. Both bentonites exhibit high CEC of 87 (±0.048)
meq/100 g, and 95 (±0.34) meq/100 g, respectively, showcasing their efficiency in retaining
cations on their intrinsic surfaces. XRD analysis highlights the fundamental 84 wt.-% and
85 wt.-% montmorillonite content in Laviosa and LMS, respectively, influencing their
swelling behavior and enhancing sorption performance. Interlayer spacing calculations
reveal Na+ as the predominant interlayer cation, with some Ca2+ in both Laviosa and LMS,
also impacting the swelling behavior. The experiment involving uranium revealed Kd values
in the range of 130–135 cm3/g with Laviosa and 78–110 cm3/g with LMS. Similarly, in the
experiment with cesium, Kd values were in the range of 130–280 cm3/g with Laviosa and
150–425 cm3/g with LMS. The bentonites demonstrate strong sorption of cesium (95%) in
the concentration range 10-10 to 10-6 M, decreasing to 85–95% at higher concentrations up
to 10-2 M. For uranium, Laviosa and LMS exhibit efficient sorption (87–100%), reaching
100% at lower concentrations and gradually decreasing to 80–90% at higher
concentrations.

The results of the experimental work demonstrate that the alternative bentonites proposed
for the Finnish repository have potential for key risk-driving radionuclide removal. The
sorption isotherms present valuable data, illustrating relatively high percent sorption for
both bentonites. The data inform the safety case for the storage of spent nuclear fuel and
results show that Laviosa and LMS could be suitable alternative buffer and backfill
materials. Of course, it is imperative to evaluate other key-risk driving radionuclides such as
Eu, Ni, and Th, along with the hydromechanical characteristics of the bentonites first.

5.2 Future work

Utilizing low-level liquid scintillation counters (LSC) for the quantification of uranium (U)
activity holds potential advantages. This approach has the capacity to address measurement
challenges and ensure the complete sorption of U onto analogous bentonite materials.
Moreover, exploring the optimal ratios between solid material and aqueous solutions
containing radionuclides presents promising prospects. This proposition is supported by the
research of Plecas et al., (2004), which examined the leaching behavior of Cs and revealed
that the addition of 1–5% bentonite led to the release of only 1–2% of the initial radionuclide
inventory into the environment (Plecas, P. et al., 2004). Exploring the mechanisms of

52
radionuclide removal through spectroscopic methods like X-Ray Absorption Spectroscopy
(XAS) and Time-Resolved Laser Fluorescence Spectroscopy (TRLFS) offers valuable
insights. While TRLFS is proficient at analyzing radionuclides exhibiting fluorescence, such
as U, it is not applicable for Cs. Extended X-ray Absorption Fine Structure (EXAFS) is a
synchrotron-based spectroscopic technique suitable for analyzing Cs, providing detailed
information on the element's oxidation state, local geometry, and speciation.

53
References
Abollino, O. et al. (2008) ‘Interaction of metal ions with montmorillonite and vermiculite’,
Applied Clay Science, 38(3–4), pp. 227–236. Available at:
https://doi.org/10.1016/j.clay.2007.04.002.

Abou-Lilah, R.A. et al. (2022) ‘Efficiency of bentonite in removing cesium, strontium,


cobalt and uranium ions from aqueous solution: encapsulation with alginate for column
application’, International Journal of Environmental Analytical Chemistry, 102(12), pp.
2913–2936. Available at: https://doi.org/10.1080/03067319.2020.1761348.

Ahmad, S., Hammad, R. and Rubab, S. (2022) ‘Gamma Radiation-Induced Synthesis of


Polyaniline-Based Nanoparticles/Nanocomposites’, Journal of Electronic Materials,
51(10), pp. 5550–5567. Available at: https://doi.org/10.1007/s11664-022-09823-0.

Ali Khan, S., Riaz-ur-Rehman and Ali Khan, M. (1994) ‘Sorption of cesium on bentonite’,
Waste Management, 14(7), pp. 629–642. Available at: https://doi.org/10.1016/0956-
053X(94)90035-3.

Amphlett, J.T.M. et al. (2020) ‘Insights on uranium uptake mechanisms by ion exchange
resins with chelating functionalities: Chelation vs. anion exchange’, Chemical Engineering
Journal, 392, p. 123712. Available at: https://doi.org/10.1016/j.cej.2019.123712.

Auffermann, B.& S.P.& K.J.& V.J.& L.J. (2015) A final solution for a big challenge? The
governance of nuclear waste disposal in Finland. 1st edn.

Aydin, M. (2019) ‘Renewable and non-renewable electricity consumption–economic


growth nexus: Evidence from OECD countries’, Renewable Energy, 136, pp. 599–606.
Available at: https://doi.org/10.1016/j.renene.2019.01.008.

Bachmaf, S., Planer-Friedrich, B. and Merkel, B.J. (2008) ‘Uranium sorption and
desorption behavior on bentonite’, in Uranium, Mining and Hydrogeology. Berlin,
Heidelberg: Springer Berlin Heidelberg, pp. 515–524. Available at:
https://doi.org/10.1007/978-3-540-87746-2_63.

Barnett, M.O. et al. (2000) ‘Adsorption and Transport of Uranium(VI) in Subsurface


Media’, Soil Science Society of America Journal, 64(3), pp. 908–917. Available at:
https://doi.org/10.2136/sssaj2000.643908x.

Bayülken, S. et al. (2011) ‘Investigation and modeling of cesium(I) adsorption by Turkish

54
clays: Bentonite, zeolite, sepiolite, and kaolinite’, Environmental Progress & Sustainable
Energy, 30(1), pp. 70–80. Available at: https://doi.org/10.1002/ep.10452.

Birgersson, M. et al. (2017) ‘Bentonite buffer’, in Geological Repository Systems for Safe
Disposal of Spent Nuclear Fuels and Radioactive Waste. Elsevier, pp. 319–364. Available
at: https://doi.org/10.1016/B978-0-08-100642-9.00012-8.

Blue Ribbon Commission on America’s Nuclear Future Report to the Secretary of Energy
| Department of Energy (2012). Available at:
https://www.energy.gov/ne/downloads/blue- ribbon-commission-americas-nuclear-
future-report-secretary-energy (Accessed: 9 January 2023).

Cao, X. et al. (2019) ‘On the long-term migration of uranyl in bentonite barrier for high-
level radioactive waste repositories: The effect of different host rocks’, Chemical Geology,
525, pp. 46–57. Available at: https://doi.org/10.1016/j.chemgeo.2019.07.006.

Catalano, J.G. and Brown, G.E. (2005) ‘Uranyl adsorption onto montmorillonite:
Evaluation of binding sites and carbonate complexation’, Geochimica et Cosmochimica
Acta, 69(12), pp. 2995–3005. Available at: https://doi.org/10.1016/j.gca.2005.01.025.

Chen, J. et al. (2023) ‘Insight into Adsorption of Cesium Ion in Aqueous Solution Based on
Inorganic Modified Bentonite’, Polish Journal of Environmental Studies, 32(2), pp. 1565–
1580. Available at: https://doi.org/10.15244/pjoes/158763.

Chisholm-Brause, C.J. et al. (2001) ‘Uranium(VI) Sorption Complexes on Montmorillonite


as a Function of Solution Chemistry’, Journal of Colloid and Interface Science, 233(1), pp.
38–49. Available at: https://doi.org/10.1006/jcis.2000.7227.

Choo, K.Y. and Bai, K. (2016) ‘The effect of the mineralogical composition of various
bentonites on CEC values determined by three different analytical methods’, Applied Clay
Science, 126, pp. 153–159. Available at: https://doi.org/10.1016/j.clay.2016.03.010.

Delvaux, B., Kruyts, N. and Cremers, A. (2000) ‘Rhizospheric Mobilization of Radiocesium


in Soils’, Environmental Science & Technology, 34(8), pp. 1489–1493. Available at:
https://doi.org/10.1021/es990658g.

Dohrmann, R. et al. (2012) ‘Interlaboratory CEC and Exchangeable Cation Study of


Bentonite Buffer Materials: I. Cu(II)-triethylenetetramine Method’, Clays and Clay
Minerals, 60(2), pp. 162–175. Available at:
https://doi.org/10.1346/CCMN.2012.0600206.

55
Dulian, M. (2023) EPRS. Available at:
https://www.europarl.europa.eu/RegData/etudes/BRIE/2023/751456/EPRS_BRI(2023)
751456_EN.pdf (Accessed: 23 January 2023).

Eronen, V.-P. et al. (2023) ‘Loading optimization algorithm for solving assembly
assignment problem in the final disposal of spent nuclear fuel in Finland’, Nuclear
Engineering and Design, 402, p. 112105. Available at:
https://doi.org/10.1016/j.nucengdes.2022.112105.

Fernández, A.M. et al. (2022) ‘Characterization of Bentonites from the In Situ ABM5
Heater Experiment at Äspö Hard Rock Laboratory, Sweden’, Minerals, 12(4), p. 471.
Available at: https://doi.org/10.3390/min12040471.

Fuller, A.J. et al. (2014) ‘Ionic strength and pH dependent multi-site sorption of Cs onto a
micaceous aquifer sediment’, Applied Geochemistry, 40, pp. 32–42. Available at:
https://doi.org/10.1016/j.apgeochem.2013.10.017.

Galahom, A.A. and Sharaf, I.M. (2021) ‘Finding a suitable fuel type for the disposal of the
accumulated minor actinides in the spent nuclear fuel in PWR’, Progress in Nuclear
Energy, 136, p. 103749. Available at: https://doi.org/10.1016/j.pnucene.2021.103749.

Galamboš, M. et al. (2010) ‘Cesium sorption on bentonites and montmorillonite K10’,


Journal of Radioanalytical and Nuclear Chemistry, 284(1), pp. 55–64. Available at:
https://doi.org/10.1007/s10967-010-0480-1.

García-Guzmán, P. et al. (2023) ‘Study of the cholesterol adsorption and characterization


of montmorillonite and bentonite clay’, Materials Today Communications, 35, p. 105604.
Available at: https://doi.org/10.1016/j.mtcomm.2023.105604.

García-Toraño, E. et al. (2018) ‘The half-life of 129I’, Applied Radiation and Isotopes, 140,
pp. 157–162. Available at: https://doi.org/10.1016/j.apradiso.2018.06.007.

Hedin, A. (1997) Spent nuclear fuel - how dangerous is it? A report form the project
‘“Description of risk”’. Stockholm.

Ho, M.S. et al. (2023) ‘Mechanisms Governing 90Sr Removal and Remobilisation in a
VLLW Surface Disposal Concept’, Minerals, 13(3), p. 436. Available at:
https://doi.org/10.3390/min13030436.

IAEA (2002). Available at: https://www.iaea.org/sites/default/files/gc/gc48inf-4-

56
att3_en.pdf.

IAEA 2022 (no date). Available at:


https://cnpp.iaea.org/countryprofiles/Finland/Finland.htm (Accessed: 8 November
2023).

Izosimova, Y. et al. (2022) ‘Adsorption of Cs(I) and Sr(II) on Bentonites with Different
Compositions at Different pH’, Minerals, 12(7), p. 862. Available at:
https://doi.org/10.3390/min12070862.

De Jesus, K. et al. (2021) ‘Extraction of lanthanides and actinides present in spent nuclear
fuel and in electronic waste’, Journal of Molecular Liquids, 336, p. 116006. Available at:
https://doi.org/10.1016/j.molliq.2021.116006.

Korichi, S. and Bensmaili, A. (2009) ‘Sorption of uranium (VI) on homoionic sodium


smectite experimental study and surface complexation modeling’, Journal of Hazardous
Materials, 169(1–3), pp. 780–793. Available at:
https://doi.org/10.1016/j.jhazmat.2009.04.014.

Kumpula Linda, Huhtanen Iida, Palander Salla, Ylä-Mella Mia, K.V. (2022) MEAE.
Available at:
https://tem.fi/documents/1410877/86271436/Management+of+spent+nuclear+fuel+and
+radioactive+waste+in+Finland.pdf.

Lee, J. et al. (2017) ‘Selective and irreversible adsorption mechanism of cesium on illite’,
Applied Geochemistry, 85, pp. 188–193. Available at:
https://doi.org/10.1016/j.apgeochem.2017.05.019.

Liu, H. et al. (2023) ‘Recent progress in radionuclides adsorption by bentonite-based


materials as ideal adsorbents and buffer/backfill materials’, Applied Clay Science, 232, p.
106796. Available at: https://doi.org/10.1016/j.clay.2022.106796.

Liu, W.-T. et al. (2019) ‘An EXAFS study for characterizing the time-dependent adsorption
of cesium on bentonite’, Environmental Science: Processes & Impacts, 21(6), pp. 930–
937. Available at: https://doi.org/10.1039/C9EM00124G.

Liu, X. et al. (2013) ‘Acidity of edge surface sites of montmorillonite and kaolinite’,
Geochimica et Cosmochimica Acta, 117, pp. 180–190. Available at:
https://doi.org/10.1016/j.gca.2013.04.008.

57
Lowke, D. and Gehlen, C. (2017) ‘The zeta potential of cement and additions in
cementitious suspensions with high solid fraction’, Cement and Concrete Research, 95, pp.
195–204. Available at: https://doi.org/10.1016/j.cemconres.2017.02.016.

Majdan, M. et al. (2010) ‘Uranium sorption on bentonite modified by


octadecyltrimethylammonium bromide’, Journal of Hazardous Materials, 184(1–3), pp.
662–670. Available at: https://doi.org/10.1016/j.jhazmat.2010.08.089.

‘MEAE’ (2022). Available at: https://urn.fi/URN:ISBN:978-952-327-855-4 (Accessed: 9


January 2023).

MEAE (no date a). Available at: https://tem.fi/en/nuclear-energy/nuclear-waste-


management (Accessed: 8 November 2023).

MEAE (no date b). Available at: https://tem.fi/en/nuclear-energy (Accessed: 8 November


2023).

Megouda, N. et al. (2007) ‘Removal of natural uranium from water produced in the oil
industry using Algerian bentonite’, Journal of Radioanalytical and Nuclear Chemistry,
272(1), pp. 75–79. Available at: https://doi.org/10.1007/s10967-006-6560-6.

Missana, T. et al. (2014) ‘Modeling cesium retention onto Na-, K- and Ca-smectite: Effects
of ionic strength, exchange and competing cations on the determination of selectivity
coefficients’, Geochimica et Cosmochimica Acta, 128, pp. 266–277. Available at:
https://doi.org/10.1016/j.gca.2013.10.007.

Moma, J., Baloyi, J. and Ntho, T. (2018) ‘Synthesis and characterization of an efficient and
stable Al/Fe pillared clay catalyst for the catalytic wet air oxidation of phenol’, RSC
Advances, 8(53), pp. 30115–30124. Available at: https://doi.org/10.1039/C8RA05825C.

Muurinen, A. (1990) ‘Diffusion of uranium in compacted sodium bentonite’, Engineering


Geology, 28(3–4), pp. 359–367. Available at: https://doi.org/10.1016/0013-
7952(90)90020-2.

Ojovan, M.I. (2011) ‘Radioactive waste characterization and selection of processing


technologies’, in Handbook of Advanced Radioactive Waste Conditioning Technologies.
Elsevier, pp. 1–16. Available at: https://doi.org/10.1533/9780857090959.1.

Olkkonen, V. et al. (2018) ‘Utilising demand response in the future Finnish energy system
with increased shares of baseload nuclear power and variable renewable energy’, Energy,

58
164, pp. 204–217. Available at: https://doi.org/10.1016/j.energy.2018.08.210.

Our methodology - SKB.com (no date). Available at: https://skb.com/future-projects/the-


spent-fuel-repository/our-methodology/ (Accessed: 8 November 2023).

Pabalan, R.T. and Turner, D.R. (1996) ‘Uranium(6+) sorption on montmorillonite:


Experimental and surface complexation modeling study’, Aquatic Geochemistry, 2(3), pp.
203–226. Available at: https://doi.org/10.1007/BF01160043.

Payne, T.E. et al. (2013) ‘Guidelines for thermodynamic sorption modelling in the context
of radioactive waste disposal’, Environmental Modelling & Software, 42, pp. 143–156.
Available at: https://doi.org/10.1016/j.envsoft.2013.01.002.

Peng, F. et al. (2024) ‘Effect of granular structure and initial suction on shear strength of
GMZ bentonite for deep geological disposal’, Applied Clay Science, 249, p. 107249.
Available at: https://doi.org/10.1016/j.clay.2023.107249.

Philipp, T. et al. (2019) ‘U(VI) sorption on Ca-bentonite at (hyper)alkaline conditions –


Spectroscopic investigations of retention mechanisms’, Science of The Total Environment,
676, pp. 469–481. Available at: https://doi.org/10.1016/j.scitotenv.2019.04.274.

Philipp, T. et al. (2022) ‘Effect of Ca(II) on U(VI) and Np(VI) retention on Ca-bentonite
and clay minerals at hyperalkaline conditions - New insights from batch sorption
experiments and luminescence spectroscopy’, Science of The Total Environment, 842, p.
156837. Available at: https://doi.org/10.1016/j.scitotenv.2022.156837.

Plecas, I., Pavlovic, R. and Pavlovic, S. (2004) ‘Leaching behavior of 60Co and 137Cs from
spent ion exchange resins in cement–bentonite clay matrix’, Journal of Nuclear Materials,
327(2–3), pp. 171–174. Available at: https://doi.org/10.1016/j.jnucmat.2004.02.001.

POSIVA (2012). Eurajoki. Available at:


https://www.osti.gov/etdeweb/servlets/purl/22134704 (Accessed: 11 January 2023).

Pourhakkak, P. et al. (2021) ‘Fundamentals of adsorption technology’, in, pp. 1–70.


Available at: https://doi.org/10.1016/B978-0-12-818805-7.00001-1.

Pusch, R., Yong, R.N. and Nakano, M. (2018) Geologic Disposal of High-Level Radioactive
Waste. CRC Press. Available at: https://doi.org/10.1201/9781351256803.

Saberi, M. and Rouhi, P. (2021) ‘Extension of the Brunauer-Emmett-Teller (BET) model


for sorption of gas mixtures on the solid substances’, Fluid Phase Equilibria, 534, p.

59
112968. Available at: https://doi.org/10.1016/j.fluid.2021.112968.

Salah, Gaber and Kandil (2019) ‘The Removal of Uranium and Thorium from Their
Aqueous Solutions by 8-Hydroxyquinoline Immobilized Bentonite’, Minerals, 9(10), p.
626. Available at: https://doi.org/10.3390/min9100626.

Saleh, A.S. et al. (2018) ‘Uranium(VI) sorption complexes on silica in the presence of
calcium and carbonate’, Journal of Environmental Radioactivity, 182, pp. 63–69.
Available at: https://doi.org/10.1016/j.jenvrad.2017.11.006.

Samimi, S. et al. (2019) ‘Lipid-Based Nanoparticles for Drug Delivery Systems’, in


Characterization and Biology of Nanomaterials for Drug Delivery. Elsevier, pp. 47–76.
Available at: https://doi.org/10.1016/B978-0-12-814031-4.00003-9.

Schlegel, M.L. and Descostes, M. (2009) ‘Uranium Uptake by Hectorite and


Montmorillonite: A Solution Chemistry and Polarized EXAFS Study’, Environmental
Science & Technology, 43(22), pp. 8593–8598. Available at:
https://doi.org/10.1021/es902001k.

Seki, T. et al. (2024) ‘Sorption behavior of cesium ions to Mg-containing calcium silicate
hydrate in a co-precipitation process’, MRS Advances [Preprint]. Available at:
https://doi.org/10.1557/s43580-023-00757-1.

Shchipalkina, A. and Smirnova, E. (2023) ‘Disposal methods for radioactive waste from
nuclear power plants and environmental radiation monitoring methods’, E3S Web of
Conferences, 431, p. 04002. Available at: https://doi.org/10.1051/e3sconf/202343104002.

Shetti, N.P. et al. (2018) ‘Graphene-clay-based hybrid nanostructures for electrochemical


sensors and biosensors’, Graphene-Based Electrochemical Sensors for Biomolecules: A
Volume in Micro and Nano Technologies, pp. 235–274. Available at:
https://doi.org/10.1016/B978-0-12-815394-9.00010-8.

Siroux, B. et al. (2017) ‘Adsorption of strontium and caesium onto an Na-MX80 bentonite:
Experiments and building of a coherent thermodynamic modelling’, Applied
Geochemistry, 87, pp. 167–175. Available at:
https://doi.org/10.1016/j.apgeochem.2017.10.022.

SKB (2010). Stockholm. Available at: https://www.skb.se/publikation/2167363/TR-10-


12.pdf (Accessed: 8 January 2023).

60
Statistics Finland, 2022 (2022). Available at: https://www.stat.fi/uutinen/energy-in-
finland-2022-information-package-on-energy (Accessed: 13 December 2023).

STUK (2020). Available at: https://urn.fi/URN:ISBN:978-952-309-488-8 (Accessed: 22


January 2023).

STUK (no date). Available at: https://stuk.fi/en/information-on-nuclear-waste (Accessed:


20 December 2023).

STUK (2022).

Timothy Schatz, N.K.J.M.P.S.B.O.M.O.A.S.V.K.K.P.O. (2013) Buffer erosion in dilute


groundwater. Available at:
https://inis.iaea.org/collection/NCLCollectionStore/_Public/45/087/45087745.pdf
(Accessed: 22 November 2023).

Tran, E.L. et al. (2018a) ‘Uranium and Cesium sorption to bentonite colloids under
carbonate-rich environments: Implications for radionuclide transport’, Science of The
Total Environment, 643, pp. 260–269. Available at:
https://doi.org/10.1016/j.scitotenv.2018.06.162.

Tran, E.L. et al. (2018b) ‘Uranium and Cesium sorption to bentonite colloids under
carbonate-rich environments: Implications for radionuclide transport’, Science of The
Total Environment, 643, pp. 260–269. Available at:
https://doi.org/10.1016/J.SCITOTENV.2018.06.162.

Vejsada, J. (2006) ‘The uncertainties associated with the application of batch technique for
distribution coefficients determination—A case study of cesium adsorption on four
different bentonites’, Applied Radiation and Isotopes, 64(12), pp. 1538–1548. Available at:
https://doi.org/10.1016/j.apradiso.2006.05.016.

Verma, P.K. et al. (2015) ‘Spectroscopic investigations on sorption of uranium onto


suspended bentonite: effects of pH, ionic strength and complexing anions’, Radiochimica
Acta, 103(4), pp. 293–303. Available at: https://doi.org/10.1515/ract-2014-2309.

Very Low Level Waste | NRC.gov (no date). Available at: https://www.nrc.gov/waste/llw-
disposal/very-llw.html (Accessed: 9 November 2023).

Volkov, I.N. et al. (2021) ‘Sorption of 90Sr and 137Cs on clays used to build safety barriers
in radioactive waste storage facilities’, Nuclear Energy and Technology, 7(2), pp. 151–156.

61
Available at: https://doi.org/10.3897/nucet.7.69930.

Wang, J. et al. (2017) ‘Adsorption of U(VI) on bentonite in simulation environmental


conditions’, Journal of Molecular Liquids, 242, pp. 678–684. Available at:
https://doi.org/10.1016/j.molliq.2017.07.048.

World Nuclear (2021). Available at: https://world-nuclear.org/information-


library/nuclear-fuel-cycle/introduction/nuclear-fuel-cycle-overview.aspx (Accessed: 30
January 2024).

World Nuclear, 2023 (2023) Uranium Mining Overview - World Nuclear Association.
Available at: https://world-nuclear.org/information-library/nuclear-fuel-cycle/mining-of-
uranium/uranium-mining-overview.aspx (Accessed: 20 December 2023).

World Nuclear Association. (2023) Nuclear Power in the European Union, World Nuclear
Association. Available at: https://world-nuclear.org/information-library/country-
profiles/others/european-union.aspx (Accessed: 26 January 2023).

World Nuclear Association (2021). Available at: https://world-nuclear.org/information-


library/nuclear-fuel-cycle/introduction/nuclear-fuel-cycle-overview.aspx (Accessed: 9
January 2023).

Wu, J. et al. (2009) ‘Behavior and analysis of Cesium adsorption on montmorillonite


mineral’, Journal of Environmental Radioactivity, 100(10), pp. 914–920. Available at:
https://doi.org/10.1016/j.jenvrad.2009.06.024.

Wu, L., Cao, S. and Lv, G. (2018) ‘Influence of Energy State of Montmorillonite Interlayer
Cations on Organic Intercalation’, Advances in Materials Science and Engineering, 2018,
pp. 1–8. Available at: https://doi.org/10.1155/2018/3489720.

Ye, Y. et al. (2022) ‘Removal and recovery of aqueous U(VI) by heterogeneous


photocatalysis: Progress and challenges’, Chemical Engineering Journal, 450, p. 138317.
Available at: https://doi.org/10.1016/j.cej.2022.138317.

Yoon, S., Jeon, J.-S. and Lee, G.-J. (2023) ‘Hydraulic properties of bentonite buffer
material beyond 100 °C’, Heliyon, 9(8), p. e18447. Available at:
https://doi.org/10.1016/j.heliyon.2023.e18447.

Youssef, WM. (2017) ‘Uranium Adsorption from Aqueous Solution Using Sodium
Bentonite Activated Clay’, Journal of Chemical Engineering & Process Technology,

62
08(04). Available at: https://doi.org/10.4172/2157-7048.1000349.

Yu, S. et al. (2020) ‘Uranium(VI) adsorption on montmorillonite colloid’, Journal of


Radioanalytical and Nuclear Chemistry, 324(2), pp. 541–549. Available at:
https://doi.org/10.1007/s10967-020-07083-y.

Zeyen, N. et al. (2022) ‘Cation Exchange in Smectites as a New Approach to Mineral


Carbonation’, Frontiers in Climate, 4. Available at:
https://doi.org/10.3389/fclim.2022.913632.

Zhou, C., Tong, D. and Yu, W. (2019) ‘Smectite nanomaterials: Preparation, properties,
and functional applications’, Nanomaterials from Clay Minerals: A New Approach to
Green Functional Materials, pp. 335–364. Available at: https://doi.org/10.1016/B978-0-
12-814533-3.00007-7.

Zou, J. et al. (2021) ‘A preliminary study on assessing the Brunauer-Emmett-Teller


analysis for disordered carbonaceous materials’, Microporous and Mesoporous Materials,
327, p. 111411. Available at: https://doi.org/10.1016/j.micromeso.2021.111411.

63

You might also like