You are on page 1of 62

Geo Energy and CCS

Global economic developments cause increasing energy demands


and require adequate response to the effects of growing energy CCS
CO2 Capture and Storage
consumption. Therefore, the main challenges for society in geo-energy
are security of supply and sustainability. Research by TNO strongly
contributes to realisation of these objectives.
Report | TNO-034-UT-2009-02240/A
TNO
P.O. Box 80015
3508 TA Utrecht
The Netherlands

T +31 30 256 46 00
F +31 30 256 46 05
E info-oilandgas@tno.nl
tno.nl
TNO-report
TNO Built Environment and Geosciences

TNO-034-UT-2009-02240/A

CCS
CO2 Capture and Storage

Date: December 2009


Authors: Henk Pagnier, Ton Wildenborg, Manuel Nepveu,
Bert van der Meer, Cor Hofstee, Rob Arts,
Filip Neele, Vincent Vandeweijer, Filip Neele,
Chris Hendriks (Ecofys), Ruut Brandsma (Ecofys),
Earl Goetheer, Erwin Giling, Sven van der Gijp,
Menso Molag
Number of pages: 60
Report
TNO-034-UT-2009-02240/A

TNO Built Environment and Geosciences


Princetonlaan 6
PO Box 80015
3508 TA Utrecht
The Netherlands

Colophon
Published by
TNO Built Environment and Geosciences
Business unit Geo-energy and Geo-information
PO Box 80015
3508 TA Utrecht
The Netherlands

All rights reserved.

No part of this publication may be reproduced and/or published by print, photoprint,


microfilm or any other means without the previous written consent of TNO.

In case this report was drafted on instructions, the rights and obligations of contracting
parties are subject to either the Standard Conditions for Research Instructions given to
TNO, or the relevant agreement concluded between the contracting parties. Submitting
the report for inspection to parties who have a direct interest is permitted.

ISBN/EAN: 978-90-5986-333-0

© Copyright 2009 TNO


Report|TNO-034-UT-2009-02240/A

Content

Introduction 5
Henk Pagnier

Regulating safety and effectiveness of CO2 storage 7


Ton Wildenborg, Manuel Nepveu

An aquifer CO2 injection reservoir engineering feasibility study 17


Bert van der Meer, Cor Hofstee

Monitoring underground CO2 storage  23


Rob Arts, Filip Neele and Vincent Vandeweijer

Geocapacity: feasibility of CO2 capture, transport and storage  33


Filip Neele, Chris Hendriks (Ecofys), Ruut Brandsma (Ecofys)

Development of Post Combustion Capture of CO2 at TNO 41


Earl Goetheer, Erwin Giling and Sven van der Gijp

Safe transport of CO2 53


Menso Molag

3
4
Report|TNO-034-UT-2009-02240/A

Introduction

Besides progress in energy efficiency and sustainable energy, like wind, solar and
geothermal energy, CCS is a necessary step to combat human-induced climatic change.
 
For more than fifteen years TNO has been actively researching a broad range of topics
related to Carbon Capture and Storage (CCS) and has been involved in many projects both
at home and in  various countries  from  Norway, Poland and Germany to China.

In this book report we present a study of CO2 injection in an aquifer, the storage of CO2
and the monitoring of underground CO2 storage. But since TNO tackles the whole CCS
chain, you will also find an article about the safe transport of CO2 and the development
of post-combustion capture of CO2. TNO is coordinator of CATO 2, the second Dutch
National Programme on CCS.

 
We hope that you will enjoy this book report.

If you have any questions about CCS, please contact TNO.


 
With kind regards,

Henk Pagnier
henk.pagnier@tno.nl
+31 30 256 46 06

5
6
Report|TNO-034-UT-2009-02240/A

Regulating the safety and effectiveness of CO2


storage

Ton Wildenborg, Manuel Nepveu

Abstract

The regulatory bodies for CO2 storage have been “inspired” by both societal needs
and by technological input. This contribution describes the regulatory background
and presents an outline of how these views on regulation are merged with a
practicable risk management scheme.

Introduction

The aim of CO2 storage is to reduce the emissions of greenhouse gases and thus
contribute to the abatement of global climate change, an obvious supra-national
problem. For that reason it is of prime importance that the stored CO2 remains
isolated from the biosphere for as long as global climate change is considered to
be a problem. At the same time, the storage of CO2 should not lead to unacceptable
effects on human safety and the local environment. In brief, then, the main focus of
regulating CO2 storage is to assure its effectiveness and safety in a practicable and
cost-effective way.

Regulatory framework

The primary focus of the EU Storage Directive [1] and the OSPAR Guidelines [2] is
on safety and the environment; the IPCC Guidelines [4] and ETS MRG [3] target the
effectiveness of emission reduction.

Storage Directive

The provisions of the Storage Directive cover the whole lifecycle of a CO2 storage
site, whereby the suitability of a reservoir for geological storage of CO2 is
determined initially with regard to technical, geological and economic issues. The
Storage Directive requires potential site operators to apply for an exploration permit
for this phase.
Application for a storage permit required to operate a storage site has to include a
thorough risk assessment analysis, including a model of the behaviour of injected
CO2 in the reservoir, a monitoring plan for both the operational phase and the
time frame after injection has ceased. Furthermore, a plan for corrective measures

7
Report|TNO-034-UT-2009-02240/A

in the event that identified risks materialise, including the monitoring of their
effectiveness, has to be provided.
The operator is required to have in place financial security enabling him to
meet obligations arising from the permit, e.g. corrective measures, post-closure
monitoring or acquiring certificates under the EU-ETS in the event of leakage. The
permit application is assessed at national level.

During operation of the site, the operator has to report annually on the results of
monitoring and the amounts and composition of the CO2 injected. Competent
authorities will carry out inspections of the storage site at least once annually.
The monitoring plan has to be updated at least every five years and submitted
to the competent authority for approval. In the event of leakage or significant
irregularities, the operator is obliged to take corrective measures.

After injection has been ceased, monitoring can be reduced. Responsibility for the
storage site can be transferred to the respective EU member state after a minimum
time frame of 20 years. Under certain circumstances, a shorter time frame is
applicable. As a prerequisite for the transfer, the storage site operator has to show
that the CO2 in the reservoir will be completely and permanently contained. Even
after transfer has taken place, a minimum level of monitoring has to be carried out.

The CCS directive prescribes mitigation of risks as the red-line along the lifecycle of
a storage site. A thorough risk assessment analysis has to be carried out during the
exploration phase. Monitoring must be based on the risk assessment and the same
applies to the design of corrective measures, their monitoring as well as the financial
mechanism structure.
Monitoring is required in order to check whether the stored CO2 behaves as expected,
migration or leakage occurs or an identified leakage threatens the environment or
human health.

The Storage directive refers to the OSPAR Convention 2007, which was amended to
allow for geological storage under the seabed. The directive stipulates the inclusion of
OSPAR Guidelines in the directive for storage under the seabed.

OSPAR Guidelines

These guidelines provide generic guidance for contracting parties, when considering
applications for permits for the storage of CO2 streams in geological formations deep
under the seabed. The guidelines specify that the monitoring programme should
encompass the monitoring of effectiveness of CO2 storage as a GHG mitigation
technology. This objective is not in line with the objectives of the Storage directive but
is covered by the IPCC guidelines and the ETS MRG.

8
Report|TNO-034-UT-2009-02240/A

MRG in ETS Directive

The MRG specifies how emissions of the CO2 storage activity have to be reported
and emphasises the verification, accounting and reporting of any emissions or
releases towards the water column. The MRG states that a monitoring plan should
be established. This includes a detailed, complete and transparent documentation
of the monitoring methodology of a specific installation, including documentation
of the data acquisition and data handling activities as well the corresponding
verification control system. It should include the following specific items:
Quantification approaches for emissions or CO2 release into the seawater from
potential leakages as well as the applied and possibly adapted approaches for actual
emissions or CO2 release into the spawater. Descriptor of the calculation or measure-
ment-based method for quantifying emissions.
Furthermore the operator must provide an assessment to show understanding of the
main sources of uncertainty when calculating emissions. If there is an indication that
CO2 is emitted or released into the seawater (additional) monitoring techniques must
be applied to enable the quantification of the leakage term(s).

IPCC Guidelines

The IPCC guidelines consist of a number of steps leading to the inventory and
quantification of emission terms during injection and storage of CO2. The most
important step is to determine whether each site has a suitable monitoring plan. Each
site’s monitoring plan should describe monitoring activities that are consistent with
the leakage assessment and modelling results.

Types of risk

Three basic groups of risk can be discerned [7]:


CO2 leakage
Ground movement
Brine displacement

Potential mechanisms of leakage

In order to assess the potential safety or environmental impact of CO2 leaking


out of the subsurface storage container, the possible leakage mechanisms from
the reservoir and subsequent flow pathways to the surface are considered. Four
individual potential leakage mechanisms can been distinguished (see also Figure 1):
Leakage through the sealing caprock: It is unlikely but still possible that CO2 will leak
at relatively small constant fluxes from the reservoir through the seal as a consequence
of fracturing, chemical reactions or exceeded entry pressure. These processes depend

9
Report|TNO-034-UT-2009-02240/A

on the pressure development in the reservoir and the potential of mineral and residual
trapping in the seal.

Leakage from the reservoir spill-point: In theory CO2 may reach the reservoir’s spill
point during injection, thereby spilling CO2 into an adjacent aquifer at the same
stratigraphic level. Spill leakage of CO2 could occur if the total injected volume of CO2
exceeds the volume of the trapping structure. This would result in uncontained flow
of CO2. The probability of this scenario to occur could be reduced by using a detailed
characterisation of the storage reservoir.

Leakage through or along geological faults: Faults are planar zones at which strata or
layers are discontinuous and displaced. This displacement, and subsequent alteration
of the displaced rock, can result in changes of the permeability in the fault zone. Due
to deformation of the reservoir during the production of hydrocarbons or injection of
CO2, faults can also be reactivated. It is unlikely, however, that these reactivations will
increase the permeability over a large area as maximum displacements are within the
centimetre range.

Leakage through or along wells: Well flaws have been the major cause of leakage in
underground gas storage facilities. Injection well failures thereby result from the use
of construction materials incompatible with the injected waste, leading to excessive
corrosion of the well casing. Generally, repairing or reconditioning the wells fixes this
problem.
Large-scale industrial storage of CO2 leads to new angles regarding well integrity
issues, due to the dissimilar behaviour of CO2 relative to natural gas. The primary
difference between the two gases is the propensity of CO2 to form a slightly acidic
solution in combination with water that may chemically interact with the well
structure material. In order to prevent leakage of CO2 through or along the well,
the applied materials need to be highly resistant to both short-term and long-term
degradation processes.

Brine displacement

Injection of CO2 may lead to the displacement of fluids that were already present in
the reservoir pores. Displacement should not affect the quality of fresh groundwater
reserves. Generally this is not an issue as shallow fresh groundwater is disconnected
from the deep saline formation water.

Ground movement

Injection generally leads to expansion of the pore volume and subsequent ground
movement. Injection might induce a sudden stress release which could be detected as
an induced seismic event. The intensity of induced seismicity is generally low.
10
Report|TNO-034-UT-2009-02240/A

Figure 1: Main leakage paths for CO2 to move towards the surface;
1) seal leakage; 2) spill leakage; 3) fault leakage; 4) well leakage

Site characterization and assessment workflow

a. b. c. d. e. f. c.
Compile Define Risk Collect Building CO2 Risk
and assess- identifi- and static dynamic identifi-
evaluate ment cation process earth modelling cation
available basis and site data model and
datapile eva- eva-
luation luation

Main steps in Storage Directive:

1. Data collection

2. Building static earth model

3. Dynamic behaviour & sensitivity characterization and risk assessment

Operation re-iteration

Figure 2: Flow diagram illustrating the major phases of site characterisation and risk
assessment and the relation with the 3-step process in the Storage Directive

Risk management workflow

Site characterisation and risk assessment are strongly intertwined. The activities in
site characterisation and risk assessment are part of an iterative process, which means
that a number of loops exist between data acquisition and processing and modelling
11
Report|TNO-034-UT-2009-02240/A

and risk assessment.


The proposed risk management workflow is very flexible: the actual risk management
activities very much depend on the amount of data already available, the initial status
of the storage site and the complexity of the storage concept. A well documented
depleted gas field with a long production history needs less effort in terms of risk
management than an aquifer with far less data and evidence for safe performance.

Risk management of CO2 storage consists of the following main components:

Site characterisation and assessment;


Monitoring;
Preventive and corrective measures.

Site characterisation and assessment

Site characterisation and assessment are the crucial activities underpinning the
qualification of a site for the safe storage of CO2. The following steps are needed
(Figure 2; see also [6]):
1. Compile and evaluate available data: All accessible site data relevant for the
assessment of CO2 containment and possible irregularities in both the short and long
term need to be gathered. Much of the available data will come from hydrocarbon
exploitation, the focus of which is on the storage compartment.

2. Define assessment basis: Based on the evaluation of available data in Step 1 the
containment concept (= the trap mechanism and extent) is to be defined as well
as the objective of the assessment. Furthermore, the geological and geographical
setting of the storage site needs to be described. This step is crucial for successful risk
assessment.

3. Risk identification and evaluation: The potential risks for safety and environment
must be identified starting from the assessment basis in Step 2. This requires the
cooperation of many experts and specialists (Figure 3).

4. Acquire and process site data: The site characterisation programme is defined on the
basis of the outcome of steps 1 to 3 whereby the gaps in the available data (Step 1) are
filled to enable a full assessment of CO2 containment and potential risks. Additional
data is acquired and processed through seismic or other types of geophysical survey
and/or drilling, coring and logging campaigns.

5. Building a static earth model: The earth model is constructed on the basis of
the site data assembled in steps 1 and 4. The model should encompass the storage
compartment, the caprock and the overburden. The static earth model should allow

12
Report|TNO-034-UT-2009-02240/A

the dynamic behaviour of CO2 to be modelled and subsequent risk to be assessed


(Figure 4).

6. CO2 dynamic modelling: A physico-mathematical tool capable of simulating the


dynamic behaviour of CO2 in the storage compartment and its potential interactions
with the adjacent fluids and rocks needs to be set up. A series of modelling exercises
has to be performed to evaluate the containment of CO2 in the pre-defined storage
space. The required size and complexity of the models depend on the chosen
containment concept and the site characteristics. The selected tool or tools may range
from simple analytical prescriptions to complex numerical schemes with full coupling
of two or more physical processes.

7. Quantitative risk assessment: The most critical risk or risks identified in step 3
must be quantitatively evaluated with the help of the geological models and the
numerical tools. The complexity depends on the nature of the risks to be assessed;
they can be deployed in a conservative deterministic or in a probabilistic manner.
The outcome of the process modelling is assessed in terms of risk acceptance and risk
management activities. The main types of risk (see above) need to be evaluated in a
hazard assessment. If needed, the consequences for exposure and adversity on people,
animals, vegetation, infrastructure or buildings will be determined. The outcome of
the modelling is assessed in terms of risk acceptance and risk management activities.

Figure 3: Screen view of supporting FEP identification Figure 4: Static earth model of the
software K12-B CO2 injection pilot.

Monitoring

This important issue is described in depth in another contribution (Arts et al.; see
also [5]) elsewhere in this booklet. We refer the reader to that contribution for more
details.

13
Report|TNO-034-UT-2009-02240/A

Preventive and corrective measures

With all uncertainties present, it is sensible to focus on mitigation of leakage by


reservoir selection, operational practice and applied design of well and plugs, trying
to realise a ‘no regret’ design for which the likelihood of CO2 leakage is negligible.

Risks related to well leakage, for example, can be mitigated by specially designed well
casing and plugs for CO2 storage. Using pancake plugs, for which casing and part of
the cap rock is drilled away and substituted with a resistive cement plug, is an option.
Using such a plug might be combined with filling the entire casing and tubing with
cement.

Leakage through cap rock and faults and from spill points can be prevented by careful
reservoir selection, limited injection and ultimate reservoir pressure as well as limiting
the concentrations of impurities in the injected CO2.

Risks of leakage through wells can be mitigated by limiting the number of wells and
by adapting an abandonment strategy that includes well casing, tubing, cementing
and plugging that are thoroughly resistant to corrosion and chemical reactions with
CO2 and impurities.

References

[1] European Parliament (2008). Directive on the Geological storage of carbon dioxide; text adopted at the
sitting of Wednesday 17 December 2008, provisional edition, P6_TA-PROV(2008)0612.
[2] OSPAR (2007). OSPAR Guidelines for Risk Assessment and Management of Storage of CO2 Streams in
Geological Formations (Reference Number: 2007-12), Meeting of the OSPAR Commission, Ostend, 25-29
June 2007.
[3] Commission of the European Communities (2009).Draft Commission Decision amending Decision
2007/589/EC as regards the inclusion of monitoring and reporting guidelines for greenhouse gas
emissions from the capture, transport and geological storage of carbon dioxide, D004164/02.
[4] IPCC (2006) Guidelines for National Greenhouse Gas Inventories. Chapter 5: Authors: Sam Holloway
(UK), Anhar Karimjee (USA), Makoto Akai (Japan), Riitta Pipatti (Finland), and Kristin Rypdal (Norway)
carbon dioxide transport, injection and geological storage.
[5] North Sea Basin Task Force -NSBTF (2009). Monitoring Verification Accrediting and Reporting (MVAR)
Report for CO2 storage deep under the seabed of the North Sea, October 2009, 37 pages.
[6] Association ASPEN (2009). Support to the Introduction of the Enabling Legal Framework on Carbon
Dioxide Capture and Storage (CCS), Final draft report, 166 pages. A Service Request under the
atmospheric emissions Framework contract ENV.C.5/FRA/2006/0071.
[7] Haskoning et al. (2007). AMESCO - Generic Environmental Impact Study on CO2 Storage, Final report,
208 pages.

14
Ton Wildenborg
ton.wildenborg@tno.nl
+31 30 256 46 36

Manuel Nepveu
manuel.nepveu@tno.nl
+31 30 256 46 36

15
16
Report|TNO-034-UT-2009-02240/A

An Aquifer CO2 Injection Reservoir Engineering


Feasibility Study

Bert van der Meer, Cor Hofstee

Introduction

An aquifer CO2 storage reservoir engineering study has been conducted with the
aid of numerical simulation to investigate the possibilities of injecting CO2 into the
Slochteren sandstone. The simulation study had to answer the following questions:
l How will reservoir conditions change as a result of a limited injection pilot
activity?
l Is it possible to inject a total of 60 Mton of CO2 into the Slochteren sandstone in a
total period of seven years, with an injection rate of some eight and a half times
more than the Sleipner CO2 site.
l Are special well conditions required?
l Is it possible to predict the pressure and CO2 distribution for the total storage
period?
The following sections summarise all the assumptions made, procedures and tools
used and conclusions drawn.

Reservoir Simulation Model

From the outset it was decided to use the ECLIPSE (Schlumberger) reservoir simulator
as a logical step after a PETREL (Schlumberger) geological modelling activity. Almost
all standard reservoir simulation software is based on finite differentiation of the
relevant equations. They solve Darcy’s law, which relates volumetric velocity to the
pressure gradient and gravitational force between so-called gridblocks. At the same
time the simulator uses a generalised conservation equation to account for material
balance.

Figure 1: Original grid layout, with indication of Figure 2: Model grid layout used, with indication of
the well locations the well locations
17
Report|TNO-034-UT-2009-02240/A

One of the limitations of this approach is the fact that the simulator running times are
directly related to the number of gridblocks used. So, in nearly all cases some degree
of upscaling is required. In the present study, we opted to reduce the number of
layers in order to end up with a reasonable number of simulation gridblocks. The five
layers used represent the Slochteren sandstone member only. The resulting original
grid was 355 x 338 x 5, a total of almost 600,000 gridblocks (fig. 1). In order to reduce
the number of gridblocks, while considering the boundary of the targeted formation
(affected space, van der Meer, 2008), a major upscaling of grids in the areas where just
compression of the brine takes place, was conducted. This resulted in a reduction of
the number of gridblocks to some 400,000 active gridblocks (fig. 2).

For initial reservoir model testing a much smaller subset was used. The reduced test
model was basically an excerpt from the large model with the following dimensions:
197 x 162 x 5. The block sizes measured 200 by 200 metres and the total of almost
400,000 gridblocks is exceptionally high for a compositional reservoir simulation
activity. Such a problem size requires a large amount of effort to check the model and
resolve possible model discrepancies.
The final model properties are directly copied from the geological model with tops of
the Slochteren formation covering a depth range of approx 1390 m in the northern
section. The gridblock porosities run from 27.3 % to 0.5% and the permeability ranges
-5
from 2.2 to 1231 mD. The rock compressibility was assumed to be 5*10 /Bar.

Model comparison

The aquifer CO2 injection test case is basically a straightforward gas /water problem.
We can model the injection of CO2 simply if we are able to model the single phase
behaviour for water and CO2 within the pressure and temperature range of the
proposed operation. For this purpose TNO has two types of simulator available: the
compositional reservoir simulator (Eclipse300) that handles individual components
and their interaction by means of an Equation of State (EOS) and the simulator with
black oil (Eclipse100) type fluid definitions that can handle two liquids, i.e. water and
oil, and one gas phase. Any dissolved gas in the oil phase is simply handled by a gas
volume per oil volume versus pressure function. We have tested both approaches.
The potential problem here relates to the temperature sensitivity of CO2 behaviour
under reservoir conditions. Both simulators work iso-thermally, so heat exchange
is igored, which is an acceptable approach in the case of a nearly flat hydrocarbon
reservoir, such as the current. In this study TNO has attempted to inject CO2 into a
deeper part of the formation with the assumption that CO2 migrates up dip (southerly
direction
We have used ECLIPSE 100 (black oil) and ECLIPSE 300 (compositional) both with the
assumption of iso-thermal behaviour. Although the predictions were comparable,
some differences were observed. The ECLIPSE 100 simulator predicts somewhat lower
flowing bottomhole pressures (FBHP) and a larger CO2 gas bubble.
18
Report|TNO-034-UT-2009-02240/A

Operational parameters

Two long horizontal wells (Figure 5) were planned. To minimise possible injectivity
problems the wells are drilled from the same locations in opposite directions. We
assumed that for both wells nearly 2000 meters of the horizontal section can be used
for injection, probably by means of intermittent perforated intervals, for instance 25
metres of perforation per 100 metres of reservoir penetrated horizontal well section.
A lack of data implies that any feasibility study for aquifer storage must be followed
by the evaluation (including simulations) of CO2 injection pilot tests before up-scaling
to full-scale injection.

From the beginning of this project it was decided that an initial pilot CO2 injection
would have to should prove the reliability of the storage, so the development team
proposed making 0.2 Megaton/year of CO2 available. Furthermore, it was decided that
a successful two-year pilot could be followed by a full-scale injection schedule of 8.5
Megatons/year up to a total forecasted formation injection capacity of 60 Megatons,
3
i.e., a total of seven years at a full injection rate. In the first two years 292,000 Nm /
day is being injected into one well to represent the pilot period to be followed by the
full-scale injection period with a two-well injection plan with for each well at a CO2
injection rate of 6,210,000 Nm3/day.

Performance prediction

The ECLIPSE 300 simulation results are presented here. Figure 3 shows the CO2
injection rate (blue line) and the total volume of CO2 for the total (green line) of nine
years of injection.

Figure 4 shows the pressure performance prediction. The red and green curves are the
predicted FBHP for the two wells, while the blue line represents the average pressure
of the gas bubble. Figure 5 shows an impression of the free CO2 bubble size at the top
of the storage formation totalling 8.5 by 4.2 km. In order to get a clear impression
of the pressure development in time we have made two pressure profiles. The cross-
sectional profiles (figures 6 and 7) show the incremental pressure as a result of CO2
injection. The influence of faults on pressure jumps is clearly visible. It is expected that
after injection is stopped the pressure will stabilise at a constant average pressure of
around 17 bar for the whole affected area. A further pressure reduction is possible if
CO2 solubility is active over a long storage period. In this simulation solubility is not
accounted for due to the assumption that solubility as a result of molecular diffusion
is a very slow process and probably not yet noticeable, especially in a highly saline
storage location, over the nine years of the simulated period.

19
Report|TNO-034-UT-2009-02240/A

FGIT vs. TIME (FRIES300)


FGIR vs. TIME (FRIES300)

4E+10 14000000

12000000

3E+10
10000000

FGIR SM3/DAY
8000000
FGIT SM3

2E+10
6000000

4000000
1E+10

2000000

0E+0 0
0 1000 2000 3000 4000
TIME DAYS

Figure 3: Total CO2 injection performance

WBHP:CO2ANN1 vs. TIME (FRIES300)


WBHP:CO2ANN2 vs. TIME (FRIES300)
FPR vs. TIME (FRIES300)

280
WBHP:CO2ANN1, WBHP:CO2ANN2,FPR BARSA

260

240

220

200

180

160
0 1000 2000 3000 4000
TIME DAYS

Figure 4: Pressure performance plot


20
Report|TNO-034-UT-2009-02240/A

Figure 5: CO2 saturation plot after 60 Megaton of CO2 injection. The size of every block is nearly 200 x 200 m

70
1 YEAR
60 2 YEAR
3 YEAR
4 YEAR
50 5 YEAR
Presure increase [bar]

6 YEAR
7 YEAR
40 8 YEAR
9 YEAR

30

20

10

0
3900 6900 13900 18900 23900 28900 33900 38900 +3900 +8900
Cross-sectional distance in the Y direction [along wells] [m]

Figure 6: Incremental pressure profile along the wells

70
1 YEAR
60 2 YEAR
3 YEAR
4 YEAR
50 5 YEAR
Presure increase [bar]

6 YEAR
7 YEAR
40 8 YEAR
9 YEAR

30

20

10

0
6300 16300 26300 36300 46300 56300 66300
Cross-sectional distance in the X direction [across wells] [m]

Figure 7: Incremental pressure profile across the wells

21
Bert van der meer
bert.vandermeer@tno.nl
+31 30 256 46 35

Cor Hofstee
cor.hofstee@tno.nl
+31 30 256 47 04

22
Report|TNO-034-UT-2009-02240/A

Monitoring underground CO2 storage

By Rob Arts, Filip Neele and Vincent Vandeweijer

Introduction

Monitoring and verification are key issues for the implementation of large-scale
underground CO2 storage. According to the EU directive on CCS, CO2 storage
will be admitted to the ETS only if monitoring and verification can be carried out
satisfactorily.

Verification will essentially take place by comparing monitoring data with simulated
predictions. Where deviations from the predicted behaviour are significant, mitigating
measures must be taken. Quantification of leakage, for example, is important first of
all for HSE reasons but also for economic reasons. Emitted CO2 must be deducted from
the Carbon Credits provided under the ETS scheme.
Current demonstration projects like Sleipner, K12-B and InSalah have clearly
demonstrated that CO2 can be monitored in the subsurface. A large suite of techniques
has been tested at these and other sites. However, current experiences have also shown
that it is very difficult to pre-define a monitoring programme for CO2 storage because
of local site-specific (geological) conditions and circumstances. Regulators clearly have
a need for scientific input on the definition of a suitable monitoring programme and
on the techniques required for different types of underground storage.
This contribution aims to give an overview of monitoring requirements combined
with examples of monitoring. Part of this work has also been reported in a North Sea
Basin Taskforce report (NSBTF report: Criteria for North Sea CGS Risk Acceptance and
Site Qualification).

Monitoring categories

Monitoring serves several important purposes: confirmation of CO2 containment,


alerts for corrective measures in the event of increased leakage risk and gathering
evidence for the long-term containment of CO2. This can be achieved either by
measuring the absence of any leakage through direct detection methods or by
verifying indirectly that the CO2 is behaving as expected in the reservoir based on
static and dynamic modelling and updates corroborated by monitoring data. The
main challenge for measuring the absence of leakage consists of spatial and temporal
coverage of the monitoring method. In other words, where and when do we need to
monitor in order to be sure that no leakage occurs? The strategy should therefore be
based on identified risks.

23
Report|TNO-034-UT-2009-02240/A

For the indirect monitoring model, the emphasis is more on scenario confirmation.
As long as predictive models are behaving in accordance with monitoring data,
understanding of both the processes occurring and the behaviour of the storage
complex can be considered sufficient. In the case of deviations, one should find the
causes of the deviations and where necessary recalibrate the models. If, however,
the deviations fall well beyond the uncertainty ranges of the predictive models, then
additional monitoring and possibly contingency measures need to be taken.
In practice, a combination of approaches will often be required and the optimal
monitoring plan will be guided by the risk assessment and the site characterisation.
As an example, for a depleted Rotliegend gas reservoir with a thick cap rock of
Zechstein salt, the risk of leakage through the cap rock can be considered minimal.
In such an example no particular monitoring for direct leakage detection is deemed
necessary. If, however, the predicted behaviour of the CO2 in the reservoir gives rise to
irregularities like unexpected pressure drops, then additional monitoring for leakage
detection might be required.

The following monitoring categories can be identified:


1. Mandatory monitoring (in any case for all sites): a number of parameters to be
monitored is mandatory based on the storage directive.
2. Required monitoring (site specific): this monitoring group is directed to gathering
evidence for containment in the reservoir and to demonstrate the integrity of seal,
fault and wells in the case of regular development.
3. Optional contingency monitoring: the third group refers to a contingency
monitoring system which will only be installed if significant irregularities show up.
In the Storage directive a “significant irregularity” is defined as '…any irregularity
in the injection or storage operations or in the condition of the storage complex
itself, which implies the risk of a leakage or risk to the environment or human
health’.

For contingency monitoring especially, it is worthwhile noting that this begins at


a later stage of the project after injection has started. This implies that a baseline
dataset is not necessarily available for comparison with a dataset acquired during
contingency monitoring activities. This is not necessarily a problem in cases where
large anomalies are expected, but careful planning of the entire monitoring plan
prior to the injection phase is important. Contingency monitoring based on “what
can go wrong” risk analysis needs to be taken into account already at the very early
pre-injection stage.

Quantification of leakage for ETS

The quantification of leakage at seabed level or ground surface level for ETS purposes
is considered part of the contingency monitoring. Quantitative monitoring for

24
Report|TNO-034-UT-2009-02240/A

ETS will only be required if there is an indication of leakage. The key question for
quantitative monitoring is, of course, to what extent state-of-the-art technology
allows for an accurate quantification.
In that respect, several issues need to be taken into account for individual monitoring
techniques:
Resolution of individual monitoring methods. Resolution can be translated
into the smallest amount of CO2 that can be detected by the individual method.
Resolution generally depends on the instrument specifications but also on the local
environmental circumstances.
Accuracy of the individual monitoring methods. Accuracy can be translated into the
uncertainty margin on the measurements. As for resolution, accuracy can be divided
into the accuracy of the instrument and accuracy determined by the local environment
(i.e. ambient noise, measurement circumstances). As an example, specifications of
geophones for acquiring seismic data provide detailed information on the accuracy
of the measured signal. However, it makes a large difference if a geophone is placed
in a soil with good coupling compared to an unconsolidated environment, where the
transfer of the seismic signal to the geophone may be very bad. And even when the
coupling is perfect, the geophones will also pick up local noise such as that caused by
traffic or industry in the neighbourhood.

Parameters measured by the monitoring method. Most of the currently available


monitoring techniques do not measure CO2 concentrations or fluxes directly. They
measure indirect parameters that can be related to the presence of CO2 through a
model. Such a model will add additional uncertainty to the quantification. Again,
seismic data is a good example. The seismic signal picks up differences in density and
wave velocity. A model is required to link the seismic signal to CO2 concentrations.
Acquisition pattern deployed by the method (spatial sampling). To quantify leakage
measurements over an area of variable size must be integrated. Suppose that the
aspects discussed in the first three bullets are perfectly known and that we have a
highly accurate method to measure CO2 with a high resolution directly at a location
on the surface (i.e. a CO2 sniffer), this would still not guarantee a proper quantification
of leakage over a large area. The sampling density will add another uncertainty to the
quantification of a leakage.
Continuity of the measurements in time (temporal sampling). As for the spatial
sampling, time sampling will add an uncertainty to the quantification of leakage.
Separation from background noise. Even if CO2 could be detected and perfectly
quantified, there is still a need to distinguish between CO2 having migrated from the
actual storage reservoir and CO2 coming from other sources, naturally or man-made.

Taking into account these different aspects, we suggest combining a model-driven


approach and a scenario/risk-based monitoring strategy to best estimate the leakage
for ETS purposes.

25
Report|TNO-034-UT-2009-02240/A

(Near) surface methods


repeated 2D/3D Gravity EM Seabed Concentration Flux Isotope content Groundwater Hyperspectral InSar
seismics time-lapse time-lapse echosounding measurements measurements of CO2 samples (Vegetation stress)
(Sniffers)
Parameter to be monitored Offshore Onshore Onshore

Seal integrity (or leakage) Gas pocket Anomalies Anomalies Pockmarks at Only when Only when Only when Only when Only onshore highly -
detection in in shallow the seabottom seafloor is seafloor is seafloor is aquifer below indirect measure)
overburden overburden reached at reached at reached at seafloor is
(low (low measured measured measured reached
resolution) resolution) location location location

Fault integrity (or leakage) Gas chimney Possibly Not likely Pockmarks at Only when Only when Only when Only when Only onshore highly -
gas the seabottom seafloor is seafloor is seafloor is aquifer below indirect measure)
chimneys reached at reached at reached at seafloor is
measured measured measured reached
location location location

Well integrity (or leakage) Only Only - Pockmarks at Only when Only when Only when Only when Only onshore highly -
accumulations at accumulatio the seabottom seafloor is seafloor is seafloor is aquifer below indirect measure)
intermediate ns at reached at reached at reached at seafloor is
levels intermediate measured measured measured reached
levels location location location

Ground movement (induced seismicity) Not likely Possible - - - - - - - Only


onshore

Leakage of saline fluids Not likely Not likely Possibly - - - - Only when Only onshore -
very shallow aquifer below (possibly)
seafloor is
reached

Buried above top-seal Injection wells


Pressure Chemical Inj. P, T & Flow rate DTS Integrity logs such as
(sniffer) CBL, caliper
ultrasonic casing imager
Parameter to be monitored

Seal integrity (or leakage) Pressure If - - -


increase in concentration
case of above
accumulatio threshold
n

Fault integrity (or leakage) Pressure If - - -


increase in concentration
case of above
accumulatio threshold
n

Well integrity (or leakage) Pressure If Irregularities will be Detection Yes


increase in concentration detected of
case of above anomalies
accumulatio threshold
n

Ground movement (induced seismicity) - - - - -

Leakage of saline fluids Not likely - - - -

Monitoring wells
(offset) VSP Cross-well Cross-well Micro-seismic Pressure DTS or repeated Fluid Ph CO2 detection logs Fluid samples
seismics EM Monitoring (BHP) Temp-logging (BH) (neutron, resisitivity, gravity,
acoustic, …)
Parameter to be monitored

Seal integrity (or leakage) Gas pocket Gas pocket Gas pocket In case the Anomaly in Anomaly in Only when In the near-well region Only when
detection detection detection seal is behavior behavior indicates measured measured above the seal measured above
(low fractured by indicates leakage above the the seal
resolution) the CO2 leakage seal
pressure

Fault integrity (or leakage) Gas chimney Gas chimney Possibly gas In case of fault Anomaly in Anomaly in Only when - Only when
chimney (re-)activation behavior behavior indicates measured measured above
detected as indicates leakage above the the seal
anomaly leakage seal

Well integrity (or leakage) Only Anomalies in - - Anomaly in Anomaly in - In the near-well region -
accumulations first arrivals behavior behavior indicates measured above the seal
at intermediate indicates leakage
levels leakage

Ground movement (induced seismicity) - - - Yes (if - - - -


detectable)

Leakage of saline fluids - - Possibly - Anomaly in - - - If taken from


changes is behavior shallower layers
signal (not indicates and detectable
likely) leakage difference

Table 31, 2, 3: Monitoring tools available and their suitability to monitor seal integrity, fault integrity,
well integrity, ground movement and/or leakage of saline fluids.

26
Report|TNO-034-UT-2009-02240/A

Examples of monitoring

Tables 1, 2 and 3 show a suite of monitoring tools available and their suitability
to monitor seal integrity, fault integrity, well integrity, ground movement and/or
leakage of saline fluids. A qualitative description is provided on what the different
monitoring tools actually show. Figures 1 to 5 are specific examples of monitoring
as circled in the tables 1-3 taken from current experiences at demo-sites and natural
analogue sites.

Figure 1: Tracking the plume at Sleipner. Migration of the CO2 is interpreted on 4D seismic data. The flow simulation
is constrained by the 4D seismic model. From the flow simulation synthetic seismic data are generated to
corroborate the match with the real observations.

Figure 1 shows an example of Sleipner (Arts, R.J., et al., 2007; Arts, R.J., et al.,
2008) where CO2 has been injected in the Utsira formation since 1996 at a rate of
approximately 1 Mt/year. So far migration of the CO2 has been monitored through 7
time-lapse seismic datasets. The interpreted seismic data have been used for a full 3D
simulation model. The construction of such a model was problematic for two main
reasons: The intra-reservoir shale layers could only be interpreted on the time-lapse
surveys when CO2 was present below them.
The conversion of these intra-reservoir shale layers from the time domain to the
depth domain is not sufficiently accurate due to the large velocity variations caused
by the presence of CO2.

As a consequence, an assumption has been made for the depth structure of the intra-
reservoir shale layers, assuming them to be parallel to the upper shale layer, the only
one that has been interpreted on the baseline seismic data.
27
Report|TNO-034-UT-2009-02240/A

Figure 2: Tracking the plume at K12-B. Tracers have been used to determine the breakthrough of injected CO2
at the production well K12-B1. The graph shows the tracer concentration, CO2 concentration and the simulated
incremental CO2 through time. The model matches not only the breakthrough time but also the increase of CO2
very well.

Figure 3: Detecting leakage. An example of a natural gas chimney (methane) in the North Sea above a shallow
gas accumulation seen with seismic data. The time slice at 300 ms shows the extent of the chimney.

Figure 4: Detecting leakage. An example of a natural CO2 seepage at Latera (Italy). The peaks show the
CO2 fluxes measured at the surface. Note the good correlation with the EM data, measuring the change in
conductivity in the soil due to the CO2.

28
Report|TNO-034-UT-2009-02240/A

The saturation model resulting form the reservoir simulation has been transformed
into an acoustic impedance model using the Gassmann model, which in turn has
been used to create synthetic seismics. Taking into account the assumptions made to
construct the model, the synthetic seismic data match the field data fairly well.

Another example of tracking the plume is shown in Figure 2 based on the K12-B
project (Van der Meer, L. G. H., et al., 2005; Vandeweijer, V.P., et al, 2009). Two
seperate CO2 injection field tests are conducted in different compartments in the
nearly depleted K12-B gas field offshore the Netherlands. The CO2 originates from
the produced gas (the methane contains 13% CO2), is separated on the production
platform and re-injected into the reservoir. The first test, finished in 2004, consisted
of CO2 injection through a single well in a depleted reservoir compartment to test the
injectivity. The second test in a nearly depleted reservoir compartment comprising
two gas production wells and one CO2 injection well is still ongoing. In March 2005,
two chemical tracers added to the injection stream such that injected CO2 could be
discriminated from resident CO2. A breakthrough of the tracers was observed in the
nearest production well after 130 days (in 2005). The simulation results for the CO2
breakthrough have been compared to the first arrival of the chemical tracers as well
as the simulated CO2 concentration and the CO2 concentrations measured at the
production well. The simulation results were very accurate.

Figure 3 shows an expression of leaking gas (chimney) on seismic data (Schroot, B.M.,
et al., 2005). It clearly reveals the ability of seismic data to show shallow gas. In this
particular case no CO2 is involved, but the natural leakage of methane from a shallow
accumulation. However, based on the compressibility and density of CO2 at these
depths, a similar expression can be expected.

Since the current demo sites fortunately do not appear to “leak” CO2, it is difficult to
test the most suitable monitoring techniques for their ability to detect CO2 migration
pathways. Figure 4 shows one of the results of a study where different monitoring
methods are being evaluated at a site in the Latera caldera (central Italy) where
natural, thermo-metamorphically produced CO2 finds its way to the surface (Arts, R.J.,
et al., 2008). The aim of the study is to identify which monitoring methods can detect
the migrating CO2 and to gain understanding of the preferential migration pathways
of the CO2. Different geophysical monitoring techniques have been deployed at a
small, 200 x 500 m study area located in the centre of the caldera. Figure 4 shows the
good correlation between measured CO2 fluxes and EM measurements.

29
Report|TNO-034-UT-2009-02240/A

Conclusions

Current experiences at demo-sites demonstrate our ability to monitor underground


CO2 storage. However, as is also admitted in the EU directive, since the success of
most monitoring tools remains site-specific, more experience with a wide variety of
sites is required to draw more generic conclusions.
TNO has numerous activities focused on monitoring, both from a technical point
of view and from a regulatory point of view. This is clearly demonstrated through
considerable involvement in projects(past and present) like SACS, CO2STORE, CASTOR,
RECOPOL, MOVECBM, CO2REMOVE, ORC, MONK, CATO and CATO-2.

References

North Sea Basin Taskforce on CO2 Geological Storage Legal and Regulatory Frameworks: Criteria for North Sea CGS
Risk Acceptance and Site Qualification (March 2009 Draft)

EU directive, http://eur-lex.europa.eu/LexUriServ/LexUriServ.do?uri=OJ:L:2009:140:0114:0135:EN:PDF.

Schroot, B.M., G.T. Klaver and R.T.E. Schüttenhelm, "http://www.sciencedirect.com/science?_ob=ArticleURL&_


udi=B6V9Y-4FY3P5Y-1&_user=603085&_coverDate=04%2F30%2F2005&_alid=1077559208&_rdoc=1&_
fmt=high&_orig=search&_cdi=5911&_sort=r&_docanchor=&view=c&_ct=6&_acct=C000031079&_
version=1&_urlVersion=0&_userid=603085&md5=22abb829e75f60042b52e823edf86397" Surface and subsurface
expressions of gas seepage to the seabed—examples from the Southern North Sea, Marine and Petroleum Geology,
Volume 22, 499-515, 2005

Van der Meer, L. G. H., Kreft E., Geel C., Hartman J., 2005, K12-B: A Test Site for CO2 Storage and Enhanced Gas
Recovery. SPE Europec/EAGE Annual Conference, Madrid, Spain. Paper number 94128-MS

Vandeweijer, V.P., Van der Meer, L.G.H., Hofstee, C., D’Hoore D., Mulders, F., 2009, CO2 Storage and Enhanced Gas
Recovery at K12-B. 71st EAGE Conference and Exhibition, Amsterdam, the Netherlands. Extended abstract.

Arts, R.J., Chadwick, A., Eiken, O., Trani, M.,Dortland, S., 2007. Synthetic modelling of time-lapse seismic data
versus real data at the Sleipner CO2 injection site. 77th SEG Annual Conference & Exhi-bition, San Antonio, Texas
(US). Expanded abstracts paper TL P1.3.

Arts, R.J., Chadwick, A., Eiken, O., Thibeau, S. and Nooner, S. 2008. Ten years of experience with CO2 injection in the
Utsira Sand at Sleipner (offshore Norway). In Goulty, N. and Arts, R.J., Special topic on underground CO2 storage,
First Break, January 2008.

Arts, R.J., Baradello, L., Girard, J.F., Kirby, G., Lombardi, S., Williamson, P. and Zaja, A., 2008. Results of geophysical
monitoring over a “leaking” natural analogue site in Italy. 9th International Conference on Greenhouse Gas Control
Technologies, Washington, 2008.

30
Rob Arts
rob.arts@tno.nl
+31 30 256 46 38

Filip Neele
filip.neele@tno.nl
+31 30 256 48 59

Vincent Vandeweijer
vincent.vandeweijer@tno.nl
+31 30 256 47 87

31
32
Report|TNO-034-UT-2009-02240/A

Geocapacity: feasibility of CCS projects

Filip Neele, Chris Hendriks (Ecofys), Ruut Brandsma (Ecofys)

The Geocapacity CCS tool evaluates the economic feasibility of CO2 capture, transport
and subsurface storage (CCS). The tool can be used in the first stages of planning CO2
projects and was developed to help find options for CCS throughout Europe. Each
element in the CCS chain is represented in the tool. The geological properties of the
subsurface storage reservoirs, key cost drivers in a CCS project, are essential input
data. A CCS project may consist of any number of CO2 sources and sinks, connected
by a pipeline network. The tool uses the database of CO2 emission points and storage
locations in Europe that has been compiled in the EU Geocapacity project.

Introduction

The aim of the EU Geocapacity project (2006 – 2008) was to assess the European
capacity for geological storage of CO2 in deep saline aquifers, oil and gas structures and
coal beds. Other priorities were the development of methods for capacity assessment,
economic modelling and site selection as well as international cooperation, especially
with China. This paper presents the economic modelling tool that was developed to
model and assess the economic feasibility of CCS projects.
The CCS analysis tool can handle scenarios with multiple sources and multiple storage
locations. In addition, the tool clearly shows the uncertainties associated with the
analysis of the economic feasibility of CCS projects. As CCS initiatives are being
developed in Europe, the uncertainties at various levels in the CCS chain become
apparent; it is essential that a feasibility assessment takes these uncertainties into
account. The updated and extended database of European CO2 emission and storage
capacity provides the basis for the Geocapacity CCS tool.

Geocapacity CCS tool highlights

The economic tool has the following characteristics.


Many sources + many sinks in a pipeline network. The tool can handle multiple
sources and multiple sinks (storage locations). CCS projects that are currently being
planned are more complex than one-to-one configurations. Storing the CO2 of even a
single coal-fired power plant typically requires in the order of 2-5 MtCO2 to be stored
each year over the expected lifetime of the power plant, which is typically 30 or 40
years. This requires a total storage of up to about 200 Mt for a single power plant.
Although this could be stored in a single (large) aquifer or depleted gas field, it is
reasonable to assume that multiple fields will be used. In the Netherlands, storing this
amount of CO2 will require a cluster of depleted gas fields. On the source side, current

33
Report|TNO-034-UT-2009-02240/A

plans assume a network, linking several sources to a cluster of storage locations.


Stochastic analysis of the cost of all elements in the CCS chain. The uncertainty
inherent in economic feasibility studies must be represented in the results of the tool.
In the case of CCS projects, uncertainty exists not only in the economic parameters
but also in the calculation of capture, transport and storage costs. Additionally, the
geological properties of the storage locations can be highly uncertain, especially in
the case of storage in saline aquifers. A stochastic approach is required to propagate
uncertainty in input parameters to the results.
The timing of costs is driven by properties of subsurface storage reservoirs. Also
in this sense, the geological properties of the subsurface storage reservoirs are the
ultimate drivers for storage and transport costs. These properties are key elements in
the tool’s input.
Modules for elements of the CCS chain are based either on first principles or on
experience-based relationships. The user can define the level of detail in the
computations, depending on the amount and quality of the data available.
The tool uses the EU Geocapacity database of CO2 sources and sinks (oil / gas fields,
aquifers).

Decision Support Software (DSS) set-up

The tool consists of two parts: a web-based application and a standalone program. The
first part of the tool (web-based) serves as a scenario builder. The user selects sources
and sinks from a database; the tool computes the connecting pipeline network. The
second part of the tool is a standalone application that uses data downloaded from
the web-based part as input. It performs a Monte Carlo analysis of the CCS project,
at both technical and economic level, and provides the user with an estimate of the
economic costs of the CCS project.

Web-based tool

The web-based tool defines the CCS scenario: sources, sinks and network. The
Geocapacity database provides European CO2 emission points and storage locations
– oil and gas fields, aquifers (Figure 1). In a Google Maps window, sources and sinks
are shown for all participating countries. A network connecting all sources and sinks
can be computed. The network can be edited by the user, for example, to align the
network with existing pipeline corridors (Figure 2).

Tool on local computer

The CCS scenario constructed with the web tool is downloaded to a local computer,
where a standalone application performs a stochastic analysis of the economic costs
of a CCS project. This tool is an application of decision support software
developed at TNO. A series of modules represents the entire CCS chain, to compute
34
Report|TNO-034-UT-2009-02240/A

the costs associated with capture, compression, transport and storage. The modules
representing the elements of the CCS chain are briefly discussed here, in the order in
which they are executed in the tool.

Capture

The capture module estimates capture parameters and cost for a variety of sources
types and capture technologies. The latter include post-combustion (capture of CO2
from flue gases), pre-combustion (capture by fuel conversion), oxyfuel combustion
(capture by fuel combustion using pure oxygen) and high-purity CO2 (sources that

Figure 1: Screen shot of the Google Maps interface showing Netherlands sources (blue),
aquifers (red) and hydrocarbon fields (purple) from the Geocapacity database.

Figure 2: Hypothetical example of a CO2 network, before (inset) and after aligning the
network produced by the economic tool with existing pipelines (purple).

35
Report|TNO-034-UT-2009-02240/A

emit almost pure CO2). Important parameters are plant type (both power and non-
power plants), new versus existing (retrofit) installations, plant technology (gas
turbine, combined cycle, etc.), the type of fuel (gaseous or liquid).
The capture module calculates the energy required for the capture process, the
amount of CO2 captured per year, any additional CO2 generated by the process itself,
remaining CO2 emissions and the economics of the capture process.

Compression

The compression module calculates the energy required for the compression process,
the amount of CO2 produced additionally per year due to compressing the captured
CO2 and the costs (investment and operational costs).

Storage

A significant part of the uncertainty in CCS projects is associated with the properties
of the subsurface reservoirs. Whereas depleted gas fields are generally well studied
and have a relatively well-defined storage capacity (although current increasing
oil prices result in highly uncertain production end times), saline aquifer storage
is associated with uncertain reservoir properties. As the timing of costs strongly
affects the net present value (NPV) of a project, all costs must be placed at the correct
position in time in the lifetime of the project. The key driver here is the behaviour of
the sinks, which determine if and when each sink is developed and activated, and
which parts of the network are to be constructed. The storage module was developed
for injection of supercritical CO2 into aquifers and abandoned oil and gas reservoirs
and computes storage capacity and injection rates. Storage potential (MtCO2) and
storage rates (MtCO2/yr) can be taken directly from the database, or computed from
geological properties (such as permeability and reservoir thickness).
EOR is not included in the current version but can easily be added.

Source – sink matching

When the volume of captured CO2 (throughout the lifetime of the scenario) is known
and the reservoir properties (storage capacity and injection rates) are available, the
flow of CO2 through the network can be computed. Whereas the network structure is
computed by the web-based tool, the size of the pipeline segments is not known, nor
is the year in which each of the sinks in the scenario is to be developed for injection.
Both these data are required for the computation and timing of transport and storage
costs; they are computed by the source-sink matching algorithm.
The source-sink match algorithm determines where the captured CO2 is stored. A
choice can be made to use the closest available sink, or the largest available sink. The
former option minimises the total length of pipeline used at any given time during

36
Report|TNO-034-UT-2009-02240/A

70

60

50
e/tCO2 avoided

40

30

20

10

0
NPV NPV capture NPV NPV transport NPV storage
compression
Figure 3: Net present value for a hypothetical CCS project, in e/MtCO2. Shown is the NPV for the entire
project, as well as for capture, compression, transport and storage separately.

Figure 4: Investment costs (CAPEX) during the lifetime of the CCS project, and
operation and maintenance cost (OPEX), in Me/yr.

the lifetime of the scenario while the latter option delays the development of any
additional sinks. In both cases, investment costs are delayed, which improves the net
present value of the CCS project.
During the scenario lifetime, the injection rate of an active sink decreases as the
volume of injected CO2 increases. When the maximum possible injection rate becomes
insufficient to store the annual volume of CO2, additional wells can be developed
when appropriate. When additional wells are insufficient, new sinks are developed.

37
Report|TNO-034-UT-2009-02240/A

Transport

The result of the source-sink match algorithm is the time interval in the scenario that
each segment of the network is active. This is used to compute the investment and
operation and maintenance cost for the entire network.

Economy

The economic module adds the cost of all elements in the CCS chain to compute
key performance indicators (KPI) of the project, such as the net present value (NPV),
unit cost and internal rate of return. Income from emission credits can be included
and the tax regime can be defined. Additional results include the NPV broken down
into the capture, compression, transport and storage elements as well as the total
costs for each source and each sink in the project. This helps identify dominant cost
factors. An important aspect of the results produced by the tool is the estimate of
the uncertainty associated with each cost result. All results are given in terms of a
probability distribution function (pdf) or, more concisely, as a p10, p50 and p90 values.

Results

Figure 3 shows an example of the output: CCS cost in units of EUR/ton of CO2 avoided,
for the entire chain as well as for the elements of the chain. The stochastic approach
results in an estimate of the uncertainty in the costs.
Figure 4 shows another example of output: investment and maintenance costs during
the CCS lifetime of 30  years. Each curve shown in the figure represents the result
of one Monte Carlo run. Such output shows the uncertainty in the timing of future
costs as well as in the magnitude of costs. The high investment costs at the start of
the project are due to capture installations, construction of the first elements of the
network and development of the first sinks. Investments required between about 15
and 30 years into the projects are due to additional network construction and new
storage site developments.

Conclusions

A tool has been developed that can be used to estimate the economic feasibility of a
CCS project, which includes any number of CO2 sources and sinks and a connecting
network of pipelines. Given the location of the sources and sinks, the tool computes
the network of pipelines and the total cost (net present value) of the CCS project,
using a Monte Carlo approach to propagate the uncertainty in any of the input
variables into the results. The tool covers the entire CCS chain, producing results that
can be used to test in detail the capture, compression, transport and storage aspects
of a CCS project.

38
Filip Neele
filip.neele@tno.nl
+31 30 256 48 59

39
40
Report|TNO-034-UT-2009-02240/A

Development of Post-Combustion Capture of CO2


at TNO

Earl Goetheer, Erwin Giling and Sven van der Gijp

Introduction

The global climate appears to be changing as a result of the increased atmospheric


concentration of greenhouse gases such as CO2. There is a consensus among most
scientists that CO2 emissions need to be reduced worldwide by 30 to 60 % in 2050
compared to 2000 in order to keep CO2 concentration in the atmosphere below 450
ppmv. This would keep the temperature rise between 2.4 °C and 2.8°C compared to pre-
industrialised levels [1]. In order to achieve the Kyoto protocol targets, the European
Union (EU) emphasis on the necessity to reduce CO2 emissions by developed countries
by 30% in 2020 compared to 1990 levels [3,4]. In addition, the EU is committed to
achieving a 20% reduction of its greenhouse gas emissions by 2020 compared to 1990
[5]. The emissions are closely linked to the use of fossil energy, e.g., for the generation
of electricity (see Figure 1). The removal of CO2 at some point in the energy conversion
process and subsequent storage is attracting worldwide interest. It is now regarded as
an important addition to the existing portfolio of CO2 abatement technologies needed
to fight climate change.

Other, 8% Electricity
Residential and Generation, 41%
Commercial, 13%

Industry, 18%

Transportation, 20%

Figure 1: The global CO2 emissions shares per sector 2004 [6]

It is clear that reduction of CO2 emission to the atmosphere from the power generation
sector (main CO2 emitters) would require a combination of several solutions. Carbon
dioxide capture and storage can be part of the total solution. In this chapter the focus
is on the different technology options for CO2 capture. There are three basic systems
for capturing the emitted CO2 from the power sector, which are shown in simplified
form in Figure 2.

41
Report|TNO-034-UT-2009-02240/A

N2, O2, H2O

Fuel
Air Power and Heat Flue gas CO2 seperation CO2

Post-combustion capture
CO2

Gasification or
Fuel
partial oxidation H22 Power and Heat
shift + CO2 Air
CO2 dehydration,
seperation N2, O2, H2O compression
transport and
storage
O2

Air Air seperation N2

Pre-combustion capture

Fuel Power and Heat CO2 H2O

O2 Recycle

Air Air seperation N2


Oxy-fuel-combustion capture

Figure 2: the different options for CO2 capture

Post-combustion capture

Capture of CO2 from flue gases produced by combustion of fossil fuels is referred to as
post-combustion capture.. Post-combustion capture typically uses a solvent to capture
the CO2 from the flue gases after the combustion process. The main line of research is
on the development of reactive solvents for the capture of CO2. Reactive solvents are
the preferred choice due to the low partial pressure of CO2 (between 3-15 kPa) in the
flue gas [10] .

Pre-combustion capture

In this concept, a fuel is partially oxidised to give a syngas (mixture of carbon monoxide
and hydrogen). This process, which is well known in the chemical industry, is known
as the gasification step. In the second step, the carbon monoxide is reacted with steam
in a catalytic reactor (water-gas shift reactor) to produce CO2 and more hydrogen. A
physical or chemical absorption process then separates the highly concentrated CO2
from the hydrogen. The hydrogen-rich stream can be used as a fuel in a gas turbine
combined cycle plant. Another option is to distribute the hydrogen for use in fuel cells
or in the future to provide vehicle fuel [1,10].

Oxy-fuel combustion

In oxy-fuel combustion, instead of air nearly pure oxygen mixed with recycled flue
gas is used for combustion. This will result in a flue gas that is mainly CO2 saturated

42
Report|TNO-034-UT-2009-02240/A

with water vapour [12,13]. Having water vapour as the main component in the CO2
stream makes it possible to purify and store CO2 with less downstream processing. If
fuel is burnt in pure oxygen, the flame temperature is excessively high. However, by
recycling the CO2 rich flue gas the flame temperature will be similar to that of a normal
air-blown combustor. Oxygen is usually produced by cryogenic air separation. Novel
techniques with lower energy consumption and cost are being developed [1,10].

Capture routes and technologies evaluation

Significant effort is directed towards the comparison of the three different


technological routes for the CO2 capture from fossil fuel power plants [14-17]. The
majority of these studies concluded that the net efficiency of the power plant, the
overall cost of electricity and the cost of CO2 avoided are fairly comparable for the
different technologies. The efficiencies of the power plant reference cases without
capture are compared with the overall power plant net efficiency after implementing
the capture process. It is clear that the overall efficiency of the three CO2 capture routes
(post, pre and oxy-fuel combustions) is very similar. This supports the conclusion that
it is not sufficient to base the selection of one of the capture routes solely on the
criterium of overall efficiency.

90 No capture
80 Pre-combustion
Post-combustion
Cost of electricity [eMIWh]

70
Oxy-fuel
60
50
40
30
20
10
0
Hard coal Lignite Natural gas
Figure 3: Estimated cost of electricity in 2020 after applying different CO2 capture routes (excluding CO2
transport and storage costs) [18]

By evaluating the effect of the different CO2 capture routes on the overall estimated
cost of electricity, it was found that the difference between these different routes
using the same type of fuel is very limited and could not be used to select a preferred
capture route, see Figure 3. However, the advantage of one technology route over
the others could be related to the technology availability, the level of maturity of the
technology, the possibility of capture process retrofitting to the existing power plants,

43
Report|TNO-034-UT-2009-02240/A

the experience, and repetition in commercial and large-scale applications and, the
most important factor, the period needed for the technology implementation.
At TNO the focus is on post-combustion capture for the following reasons:
l The possibility of add-on to existing power plants (retrofit possibility).
l Capture technologies are considered available and the solvent technologies are
proven on a smaller scale.
l Capture readiness makes post-combustion capture relatively easy to incorporate
into power plants.
l More operational flexibility in switching between capture – no capture operation
mode.
l Learning by doing will lead to cost reductions similar to SO2 capture process
development.

The sections below focus on the technology aspects of post-combustion capture and
on the developments at TNO.

CO2 post-combustion capture

Carbon dioxide post-combustion capture is considered to be one of the most mature


capture technologies. Within many industrial applications there is much experience
with this type of capture technology and it has a good reputation [11].

State-of-art technology

Rao and Rubin 2002 [11] shows that amine-based CO2 absorption systems are the most
suitable for combustion based power plants for many reasons: for example, they can
be used for dilute systems and low CO2 concentrations, the technology is commercially
available, it is easy to use and can be retrofitted to existing power plants. Typically,
absorption processes are based on a thermally regenerable solvent, which has a strong
affinity for CO2. The process thus requires thermal energy for the solvent regeneration.
The benchmark absorption process is based on an aqueous solution with 30% by
weight Monoethanolamine (MEA) as the active ingredient. This amine-based process
is considered the state-of-the-art technology [20]. At this moment there are several
vendors available for large scale CO2 capture installations.

Process: short description

The general process flow diagram for amine absorption is shown in Figure 4. This is
the typical conventional MEA-CO2 capture flow sheet, which has been described and
discussed in a large number of commercial, industrial and research activities [9,11].
The underlying fundamental principle is the exothermic, reversible reaction between
a weak acid (e.g. CO2) and a weak base (e.g. MEA) to form a soluble salt.

44
Report|TNO-034-UT-2009-02240/A

The inlet gas is contacted counter-currently with lean solvent in the absorber. The acid
gases are preferentially absorbed by the solution. The solution, enriched with CO2,
is pre-heated before entering the stripper where, through the addition of heat, the
reaction is reversed. The lean solvent leaves the stripper and is lowered in temperature
through heat exchange with the rich solvent. The lean solvent is recycled back to the
absorber. From the top of the stripper, a high-purity (dry-basis) CO2 is produced. For
the regeneration of the rich solvent large quantities of heat are required. An important
aspect is the source of the heat and electricity. One approach is to produce the required
heat and electrical power using auxiliary equipment. The alternative is to extract the
required heat from the existing power plant.

Figure 4: General chemical absorption CO2 capture process flow sheet


.
Post-combustion capture process challenges

Several researchers have studied and evaluated the MEA absorption process [11,14,20].
An important aspect is the energy needed for the regeneration of the absorption liquid.
The main conclusion is that the thermal energy requirement for the regeneration of the
solvent is the Achilles heel of the process. To reduce cost it is imperative to lower the
heat demand for the regeneration. Typically, the current state-of-art capture process
reduces the overall power plant net efficiency by 8 to 12%[1]. This means a reduction
by almost 20% of the overall electricity output for the conventional hard coal power
plant. Furthermore, the overall economic evaluation of CO2 capture shows that the
cost of CO2 capture is currently very high. This implies a strong need for breakthrough
technology to realise the potential of CCS.

As a result of the different CCS projects, evaluations and studies, current CO2 capture
technologies can generally be concluded to have the following disadvantages:
l Significant efficiency reduction.
l Increased power generation costs.
l Large capture plant equipment dimensions and footprint, which requires further
development and modification.
l Lack of experience in CO2 capture process at a full power plant scale.
45
Report|TNO-034-UT-2009-02240/A

Development of post-combustion capture at TNO

The economic framework of post-combustion capture is highly dependent on the


solvent system, the equipment and how this equipment/capture process is integrated
with the power plant. In the TNO development strategy all three aspects are addressed
for the reduction of the capital and operational expenditure. An important approach
for the development of improved absorption systems, which can lead to a substantial
reduction in energy costs, is the solvent development workflow developed by TNO.
This solvent test street begins with detailed organic chemistry (including quantum
chemical modelling), laboratory tests to determine the main operating parameters
and pilot plant tests with real flue gas from a coal fired power plant. The workflow
consists of four main steps:

1. Design
2. Lab experiments
3. Mini plant process set-up
4. Pilot testing

Step 1 concerns the smart selection and design of absorption molecules on the basis
of a QSAR (Quantitative Structure-Activity Relationship) methodology developed by
TNO. The basic idea is to incorporate methodology developed for the pharmaceutical
sector in post-combustion research.

Step 2 is the characterisation of solvent-based thermodynamic and kinetic


considerations using lab-scale equipment. Based on these results and using in-house
models, it is possible to assess the potential of the newly developed solvent system.

Step 3 is focused on determining solvent stability and process parameters using an


autonomous continuous absorber/desorber installation.

Step 4 involves pilot-scale test runs using the CATO CO2 catcher to generate key
parameters, assess long-term stability and validate process models.

The CATO CO2 Catcher

The CATO CO2 Catcher pilot plant was opened in April 2008 and is part of CATO, a
national publicly and privately funded programme that unites efforts to research,
assess and prepare for the implementation of CO2 Capture & Storage (CCS)
technologies.
The objective is to test novel CO2 gas scrubbing  processes under real industrial
conditions. This includes solvents,  absorbers and  novel process concepts. The
construction of the pilot plant at the coal fired power plant at the Maasvlakte

46
Report|TNO-034-UT-2009-02240/A

(Rotterdam) has been a joint effort between TNO and E.ON Benelux.
The pilot facility enables the performance evaluation and benchmarking of different
CO2 capture methods under real industrial flue gas conditions. The pilot plant is
connected to the stack of the second unit of the power plant after desulphurisation.
Part of the flue gases is directed to the CO2 capture pilot plant for carbon dioxide
removal. The CO2 is then removed at a maximum capacity of 250 kg CO2 per hour.
Complete monitoring of the process conditions (temperature, pressure, flows, content
of CO2, SO2, soot, etc.) is possible. Other parameters such as solvent stability can also
be measured separately.

Figure 5: the CATO CO2 Catcher

Amino acid salts

Normal CO2 scrubbing solvents based on amines have the disadvantage of being
volatile and susceptible to degradation. To overcome these drawbacks TNO is
developing the CORAL family of absorption solvents based on amino acid salts. This
family of absorption solvents is dedicated to the removal of CO2 with a low amount of
energy consumption in the regeneration step and minimal environmental impact.

The amino acids are commercially available, biodegradable and, moreover, chemically
and thermally highly stable. Due to the fact that the CORAL family is based on the
salt of these amino acids, they have no vapour pressure. Therefore, no evaporation of
these components is possible and no atmospheric emissions will occur.

47
Report|TNO-034-UT-2009-02240/A

Figure 6: Amino acid salts versus traditional alkanol amines.

First pilot plant results

TNO has demonstrated the potential of some initial solvent compositions of the
CORALTM family in the pilot plant at the coal fired power station of E.ON at the
Maasvlakte. In total more than 3000 hours of operation suggest that CORAL XPT (one
of the first CORAL compositions) is remarkably stable under industrial conditions
(degradation losses below 0.15 kg per tonne of CO2). Minute amounts of degradation
products can be found in the vent gasses. Also great potential has been found for the
further reduction of energy consumption in the regeneration step.

60 relative cost of CO2 avoided


90 penalty point reduction
relative cost or penalty point reduction (%)

80
70
60
50
40
30
20
10
0
Benchmark CORAL XPT CORAL XPTu Under Under
(MEA) development development

Figure 7: relative cost for CO2 capture and the penalty point reduction for CORAL XPT systems
compared to the benchmark MEA
48
Report|TNO-034-UT-2009-02240/A

Figure 7 depicts the progress that has been made. It is important to note that the results
presented are based only on the bare performance of the solvent in a conventional CO2
scrubbing facility. Obviously an additional gain can be made by improving the process
design.

The indicated performances are relative to a benchmark with a 30wt% MEA solution
in a conventional capture process with only basic heat integration.

TNO has planned to employ the pilot to test other solvents. In this way industry can
be provided with a sound benchmark and support can be given to the development
and upscaling of different solvent systems. In the end, only a set of successful
solvent systems can provide a total solution for energy efficient and environmentally
responsible CO2 scrubbing under variable conditions.

It is TNO’s mission to support industry in finding the best scrubbing solvents for their
specific process and to provide data that will assist in the modelling and scaling up of
solvent systems

Future outlook
There is currently momentum for shifting from a research/design phase to an
engineering phase. In Europe several companies are involved in developing detailed
plans for large-scale capture installations. The European Commission has enabled an
additional push to realise, in the short term, large capture installations (demonstration
plants) as part of the Flagship approach. This means that by 2015 several installations
will be capturing CO2. However, because the technology will be continuously
improving it is important that for the large-scale demo plant possibilities exist to
adopt new technology such as improved solvent systems. Based on the observed
learning curve of the past decade, it can be concluded that significant cost reductions
are still possible.

Acknowledgements

The authors would like to acknowledge Mohammad Abu Zahra, Sanaz Allaie, Arjen
Huizinga, Yvonne Mergler, Eva Sanchez for their work on the continuous improvement
of CO2 capture processes.

49
Report|TNO-034-UT-2009-02240/A

References

1. Intergovernmental Panel on Climate Change (IPCC), 2005. Carbon dioxide capture and storage. Cambridge
university press.
2. Mauna Loa by Scripps Institute of Oceanography, CDIAC, and NOAA/ESRL at www.esrl.noaa.gov/gmd/
ccgg/trends, updated January 2008.
3. Intergovernmental Panel on Climate Change (IPCC), 2007: Climate Change 2007: Synthesis Report.
Contribution of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental
Panel on Climate Change [Core Writing Team, Pachauri, R.K and Reisinger, A. (eds.)]. IPCC, Geneva,
Switzerland, 104 pp.
4. European Union, 2007a. In: Press release 2785th council meeting environment Brussels.
5. European Environmental Agency (EEA) Report, 2008. Greenhouse gas emission trends and projections in
Europe, tracking progress towards Kyoto targets. Report number 5/2008.
6. M. AbuZahra, Carbon dioxide capture from flue gas: development and evaluation of existing and novel
process concepts, PhD Thesis technical University Delft 2009.
7. Birol, F., Argiri, M., 1999. World energy prospects to 2020. Energy, 24, 905-918.
8. International Energy Agency, 2005. World energy outlook 2005. OECD/IEA.
9. Kohl, A., Nielsen, R., 1997. Gas purification. 5th edition, Gulf publishing company.
10. International Energy Agency (IEA) greenhouse gas R&D programme, 2007. Capturing CO2. By Christopher
Lewis,© Separations research program 2007.
11. Rao, A., Rubin, E., 2002. A technical, economic, and environmental assessment of amine-based CO2
capture technology for power plant greenhouse gas control. Environ. Sci. Technol. 36, 4467-4475.
12. Zanganeh, K., Shafeen, A., 2007. A novel process integration, optimization, and design approach for
large-scale implementation of oxy-fired coal power plant with CO2 capture. J. GHG control, 1, 47-54.
13. Varagani, R., Chatel-Pelage, F., Gautier, F., Pranda, P., McDonald, D., Devault, D., Farzan, H., Schoff,
R., Ciferno, J., Bose, A., 2006. Oxy-combustion process for CO2 capture from coal fired power plants: An
overview of techno-economic study and engineering feasibility analysis. Proceedings of Inter. Con. GHG
Con. Tech., Trondheim, Norway.
14. Wall, T., 2007. Combustion processes for carbon capture. Proceedings of the combustion institute, 31,
31-47.
15. International Energy Agency (IEA) greenhouse gas R&D programme, 2004. Impact of Impurities on CO2
capture, transport and storage, Report No. PH4/32.
16. International Energy Agency (IEA) greenhouse gas R&D programme, 2005. Towards zero emission coal-
fired power plants. Clean coal centre, report No. CCC/101.
17. Stromberg, L., 2005. Technology benchmarking studies in Oxy-fuel Combustion. Presentation to IEA Oxy-
Fuel Combustion Research Network, Cottbus, Germany.
18. The EU technology platform for zero emission fossil fuel power plant (ZEP), strategic research agenda,
2006. Published online at www.zer-emissionplatform.eu/website/lbrary/index.html.
19. Feron, P., 2005. Progress in post-combustion CO2 capture. Presentation for the international symposium,
reduction of emissions and geological storage of CO2.
20.Chapel, D., Ernst, J., Mariz, C., 1999. Recovery of CO2 from flue gases: commercial trends. Can. Society of
Chem. Eng. (Oct 4-6).

50
Earl Goetheer
earl.goetheer@tno.nl
+31 15 269 28 46

Erwin Giling
erwin.giling@tno.nl
+31 15 269 23 01

Sven van der Gijp


sven.vandergijp@tno.nl
+31 15 269 23 60

51
52
Report|TNO-034-UT-2009-02240/A

Safe transport of CO2

Menso Molag

CO2 transport from capture to underground storage

The major users of fossil energy like refineries, petrochemical companies and power
plants that capture CO2, are mosttimes not located in the vicinity of underground CO2
storage sites, such as on shore and off shore exhausted gas fields. This means that
transport of CO2 will often involve substantial distances and in the populated areas of
Europe passing of the transport routes of residential areas cannot be avoided. For very
large quantities this transport will run via pipelines from the capture installation to
the underground storage site. For smaller capture installations transport by tankship
will have to be reviewed in terms of technical and economic feasibility.

Since the investments costs of pipeline are high, CO2 will be transported under the
highest possible pressure. High pressure causes the CO2 to liquefy, which means that
more CO2 can be transported through a narrower pipe than if CO2 is transported as
gas. In North America around 5,000 km of high pressure pipeline is available for the
transport of CO2 from natural sources or capture plants to oil fields for enhanced oil
recovery. In Europe there is still little experience with high-pressure transport of CO2
by pipeline. In the Netherlands there is an 85 km long underground pipeline with a
diameter of 0.65 m for the transport of 300,000 tonnes of CO2 per year from a refinery
to horticultural customers. In this case the transport pressure lies between 10 and
22 bar. A 20 km 44 bar underground pipeline is being planned for the transport of
400,000 tonnes of CO2 per year from the Shell refinery to an underground storage site
in Barendrecht.

There have not been reported accidents with the existing CO2 pipelines that caused
serious CO2 leakage. But if such an event were to happen this CO2 could spread and
be inhaled by people in the vicinity and be harmful to health if the concentration
exceed 10% of CO2. It is highly unlikely that such high concentrations of CO2 in the
atmosphere will occur in the event of an accident with the pipeline. But given the fact
that the pipelines may run through residential areas, safety will be a key issue in the
construction and the operation of CO2 pipelines.

53
Report|TNO-034-UT-2009-02240/A

Figure 1: Trajectory of NPM carbon dioxide storage and transport pipeline with detail showing Zoetermeer

Figure 2: Vicinity of OCAP underground CO2 pipeline

CO2 pipelines

Table 1 provides a summary of several existing and planned pipelines for the transport
of CO2 from capture installations to underground storage sites.

As is evident from table 1 a number of pipelines currently exists for the transport of
CO2. These are generally 1.5 – 2 m under the ground, have a diameter of 0.3 to 0.7 m
and a pressure varying from 10 to 200 bar. The number of pipelines for the transport

54
Report|TNO-034-UT-2009-02240/A

Pipeline Country Capacity Length Diameter Pressure


[1,000,000 [km] [m] [bar]
tonnes/year]

NEJD, Denbury USA 11.5 293 0.51 80-150

Free State, Denbury USA 6.7 138 0.51 80-150

Weyburn USA + Canada 5 328 0.35 150

OCAP Nl 0.3 85 0.65 10-22

Barendrecht Nl 0.4 20 0.36 44

Sleipner (off-shore) N 1.0 245 0.32 200

Table 1: Summary of several existing and planned pipelines for the transport of CO2.

of CO2 from capture installations to storage sites is expected to rise significantly. For
the Netherlands the current annual transport of 300,000 tonnes of CO2 is expected
to rise to 10-15 million tonnes in 2020 and 60 million tonnes by 2050. With such a
substantial rise in transport, it is inevitable that CO2 pipelines will also run along and
through residential areas. However, this is already the case for pipelines that carry
hazardous substances like petroleum and natural gas, so there is plenty of experience
of laying and using such transport pipelines safely. The issues below are focal points
for attention:

Preventing damage to the pipeline by third parties

Studies of incidents involving high-pressure natural gas pipelines reveal that 90% of
all incidents were caused by construction activities (soil excavation, pile-driving etc.)
by third parties that were unaware of a natural gas pipeline in the vicinity of their
contruction site. By marking pipelines clearly, laying them deeper, protecting them
with a fence or making the pipeline stronger, this kind of third party interference can
be prevented.

Preventing corrosion

If there is contamination in the gas (like water or hydrogen sulphide in the case of
CO2) the pipeline may become corroded. This can be prevented by the right choice of
material for the pipeline and by pre-empting this kind of contamination, removing
it from the CO2 in advance and controlling the entry concentrations. Corrosion from
outside may occur for underground pipelines but this can be prevented by selecting
the right material and cathodic protection.

55
Report|TNO-034-UT-2009-02240/A

Construction errors
Incorrect laying in an unstable subsoil or welding errors can cause leakages. This can
be prevented by adhering to construction standards and checking compliance with
the standards.

Restricting the impact of incidents

Despite all of the safety measures taken, should an incident result in a CO2 leak, then
the impact must be restricted. This can be done, for instance, by sectioning the pipeline
and equipping these sections with rapid shut-off valves so that only the inventory of
one pipeline section can be released and not the inventory of the whole pipeline. By
promptly warning residents in the vicinity of a leak, people can be advised to stay
indoor and close windows and doors. Exposure of people to the released CO2 can be
minimalised by these measures.

Harmonising the safety analysis and evaluation of CO2 pipelines

Many risk assessments of CO2 pipelines have been performed in North America and
Europe. The overall picture that emerges is that a high pressure CO2 pipeline is a
limited and manageable risk. Some risk assessments reveal no need for a safety zone
while others reveal the need for a safety zone for the built environment ranging from
0 to several hundred metres from the pipeline. Such differences can be explained by
the differences in the pipeline characteristics (e.g. pressure, temperature and pipeline
diameter) and the risk analysis method employed. Some examples of the differences
are the dispersion of released CO2 (will the released CO2 spread as dense gas in
the local environment or is it a vertical jet release with zero CO2 concentration at
ground level) and the the applied CO2 human toxicity threshold. Given the different
risk analysis methodologies it is difficult for the administrators to grant permits for
the construction and operation of CO2 pipelines. Within the CATO2 project in the
Netherlands the initiative has therefore been taken to investigate a number of aspects
of the safety of CO2 pipelines.

This investigation concerns experiments whereby liquid and supercritical CO2 leaks
from a transport pipeline. The aim of these experiments and supplementary theoretical
studies is to determine how quickly the CO2 leaks, the extent to which gaseous, liquid
and solid CO2 leaks, the shape of any crater in the event of a pipeline rupture and the
impact of this shape on the dispersion of CO2 in the vicinity of the pipeline. Separate
studies are examining the human toxicity of the CO2. This research will provide the
basis for a method to harmonise safety analysis and evaluation, something that will
enable the authorities to make a better decision in respect of the construction and
operation of a CO2 transport pipeline.

56
Menso Molag
menso.molag@tno.nl
+31 88 86 6261

57
58
Geo Energy and CCS

Global economic developments cause increasing energy demands


and require adequate response to the effects of growing energy CCS
CO2 Capture and Storage
consumption. Therefore, the main challenges for society in geo-energy
are security of supply and sustainability. Research by TNO strongly
contributes to realisation of these objectives.
Report | TNO-034-UT-2009-02240/A
TNO
P.O. Box 80015
3508 TA Utrecht
The Netherlands

T +31 30 256 46 00
F +31 30 256 46 05
E info-oilandgas@tno.nl
tno.nl

You might also like