You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/236737240

Geothermal Energy

Chapter · May 2013


DOI: 10.1007/978-1-4020-8939-8_120

CITATIONS READS

2 17,254

4 authors:

Hakim Saibi Stefan Finsterle


United Arab Emirates University Finsterle GeoConsulting
183 PUBLICATIONS 844 CITATIONS 234 PUBLICATIONS 4,104 CITATIONS

SEE PROFILE SEE PROFILE

Ruggero Bertani Jun Nishijima


Enel green power SpA Kyushu University
23 PUBLICATIONS 1,229 CITATIONS 122 PUBLICATIONS 679 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Geothermal Reconnaissance and Potential Investigations (Algeria, Egypt, Rwanda, Ethiopia, Afghanistan, Japan, Indonesia, ...) View project

groundwater issue including exploration View project

All content following this page was uploaded by Hakim Saibi on 20 October 2015.

The user has requested enhancement of the downloaded file.


Geothermal Energy
54
Hakim Saibi, Stefan Finsterle, Ruggero Bertani, and Jun Nishijima

Abstract
This chapter presents general information about the recent methods applied for
geothermal systems. Geothermal engineering can be separated into two groups:
research about the underground geothermal reservoir using geophysical and
numerical methods and the use of a geothermal power plant as a technology to
produce electricity from the underground hot waters. In this chapter, both aspects
are presented.
Twenty-four countries are currently generating electricity from geothermal
resources and 78 countries are using geothermal energy for heating purposes.
The total installed geothermal capacity worldwide is 10.7 GWe.
This chapter is divided into four parts:
The first part, the introduction, discusses the current use of geothermal
electricity and the trend of installed geothermal capacity in the world. It also
explains the main concepts of geothermal engineering and presents the different
types of hydrothermal systems.

H. Saibi ()
Earth Resources Engineering, Laboratory of Exploration Geophysics, Faculty of Engineering,
Kyushu University, Nishi-ku, Fukuoka, Japan
e-mail: saibi-hakim@mine.kyushu-u.ac.jp, saibi.hakim@gmail.com
S. Finsterle
Hydrogeology Department, Lawrence Berkeley National Laboratory, Earth Sciences Division,
Berkeley, CA, USA
e-mail: safinsterle@lbl.gov
R. Bertani
Geothermal Center of Excellence, Enel Green Power S.p.A., Pisa, Italy
e-mail: ruggero.bertani@enel.com
J. Nishijima
Faculty of Engineering, Laboratory of Geothermics, Kyushu University, Nishi-ku, Fukuoka,
Japan
e-mail: nisijima@mine.kyushu-u.ac.jp

J. Kauffman, K.-M. Lee (eds.), Handbook of Sustainable Engineering, 1019


DOI 10.1007/978-1-4020-8939-8 120,
© Springer Science+Business Media Dordrecht 2013
1020 H. Saibi et al.

The second part describes geothermal engineering technology and its compo-
nents. This part presents direct utilization, geothermal heat pumps, electric power
generation and combined heat and power generation, the numerical modeling of
geothermal systems, the current state of practice, recent advances, and emerging
trends in geothermal reservoir simulation and hybrid-microgravity monitoring
applications at geothermal field.
The third part presents a case study of Húsavı́k Energy in Iceland.
In the fourth part, the economic analysis is presented.

1 Introduction

Geothermal energy, defined as heat from the Earth, is clean and sustainable. The
energy demand by the world population is increasing due to the expansion of
economies and population growth of the planet Earth and advancements in energy-
intense technologies. Between 2008 and 2035, the world’s energy use is expected
to grow by approximately 53 % with half of the increase attributed to India and
China (International Energy Outlook 2011). Much of the energy growth comes
from developing countries. Fossil fuels are expected to supply approximately 80 %
of world energy use in 2035. Renewables are the world’s fastest-growing energy
source, at 2.8 % per year, with the renewable’s share of the world energy supply
growing to approximately 15 % in 2035 (International Energy Outlook 2011).
Renewable energy in general and geothermal energy in particular could play a
significant role in supplying a clean and environmentally sustainable source of
energy to satisfy the world energy demand and its challenges.
The Intergovernmental Panel on Climate Change (IPCC) performed a compre-
hensive study on climate. Their models assumed an increase in temperature of
approximately 2–4ı C for the period 2000–2100 with high CO2 concentration values
ranging from 550 ppm (which is twice the preindustrial level) to 850 ppm. The
increases in temperature and CO2 concentrations are mainly attributed to CO2
emissions from fossil-fuel energy production systems. Geothermal energy is one
recommended approach to decrease CO2 emissions.
Geothermal energy is the most versatile renewable energy and has been used for
thousands of years for washing, bathing, cooking and health: the direct utilization
of hot water is long lasting and still growing with different application ranges.
The world’s first geothermal district heating system was started in the fourteenth
century at Chaudes-Aigues, France, and the first geothermal well was drilled
near Reykjavik, Iceland, in 1755. However, only in the twentieth century has
geothermal energy been harnessed on a large scale for space heating, electricity
production, and industrial use. The first large municipal district heating service
was initiated in Iceland in the 1930s and currently provides geothermal heat to
approximately 99 % of the 200,000 residents of Reykjavik. The first commercial
plant to produce electricity became available in 1913 in Larderello, Italy. The use of
geothermal energy has increased rapidly since the 1970s. During the period 2000–
2010, the globally installed direct-use capacity tripled from 15 to 50 GWth , whereas
the installed capacity for electricity production increased from 8.0 to 10.7 GWe .
54 Geothermal Energy 1021

Fig. 54.1 Geothermal electricity production in the world

Fig. 54.2 Historical trend of geothermal installed capacity

At present, geothermal energy is used by 78 countries for heating purposes (called


“direct use”) and by 24 countries for electricity production. Figure 54.1 shows the
use of geothermal electricity throughout the world (Bertani 2012).
The increase in the installed capacity over the last century is clearly visible in
Fig. 54.2, with an impressive increase in new plants started in the 1980s, following
the energy crisis.
Direct utilization is present almost worldwide, with an impressive growing rate,
as shown in Fig. 54.3 (Lund et al. 2010).
1022 H. Saibi et al.

Fig. 54.3 Direct use of geothermal energy

1.1 Generalities of Geothermal Energy

“Geothermal” comes from the Greek words geo (Earth) and therme (heat). Thus,
geothermal means Earth heat. Geothermics can be defined as the study of the tem-
perature distribution in the Earth and the phenomena that influence that distribution.
The Earth’s heat is continuously radiated from within, and each year rainfall
and snowmelt supply new water to geothermal reservoirs. Production from in-
dividual geothermal fields can be sustained for decades and perhaps centuries,
which explains the renewability of this resource (Geothermal Education Office
2001) (Fig. 54.4).

1.1.1 Geothermal Gradient


A geothermal gradient expresses the rate of increasing temperature with depth in the
Earth’s crust. The average geothermal gradient is approximately 2.5ı C per 100 m.
However, these gradients are much higher at hot and active geothermal regions.

1.1.2 Geothermal Systems


Geothermal systems are located in regions with a normal or slightly above normal
geothermal gradient, creating low-to-medium-enthalpy geothermal reservoirs (less
than 150ı C). The high-enthalpy geothermal reservoirs (up to or above 400ıC) are
especially located in regions around plate margins with high geothermal gradients.
The circulation of water in terrestrial geothermal systems can reach depths of
approximately 5 km (Pirajno 1992) (Fig. 54.5).
54 Geothermal Energy 1023

Fig. 54.4 Geothermal reservoir fed by rain water

Fig. 54.5 Geothermal regions in the world (in red)

The major characteristics of geothermal energy are its renewability and sustain-
ability and its environmental friendliness (with few CO2 emissions). Its various
utilizations include power generation (conventional steam turbine, combined cycle),
direct heat use for space heating, greenhouses, aquaculture and tourism (swim-
ming pool).
There are mainly two types of geothermal systems: convective and conductive.
The convective geothermal systems include vapor-dominated and water-dominated
systems. The conductive geothermal system is represented by the hot dry rock
system. To form a geothermal reservoir, three components are necessary:
– Heat represented by the thermal energy itself, conductive heat transfer, and
convective heat transfer (fluid transport)
1024 H. Saibi et al.

– Fluid that helps to transfer energy by convection and recharging the water
– Open, permeable paths, which allow fluid to flow through porous and fractured
systems

1.2 Hydrothermal Systems

1.2.1 Water-Dominated Systems


The main characteristics of this type of geothermal system are that the liquid water is
continuous and the pressure controls the phase changes in the reservoir. Geothermal
production wells produce a fluid mixture of steam and water. The temperature of
a liquid-dominated reservoir ranges from 210ı C to greater than 300ıC. There are
many water-dominated geothermal systems all over the world, such as in Wairakei
in New Zealand, Olkaria in Kenya, Tongonan in the Philippines, Momotombo
in Nicaragua, and Hatchobaru in Japan. The advantages of the water-dominated
systems from an engineering point of view are its renewability if the separated water
is properly reinjected and the ability to use the separated water for cascade use.

1.2.2 Vapor-Dominated Systems


Liquid water and vapor coexist in the geothermal reservoir, but vapor is continuous,
and the pressure controls the phase condition. Geothermal wells in vapor-dominated
reservoirs produce dry steam only, with temperatures ranging from 230 to 260ıC.
Vapor-dominated systems are uncommon; some examples are The Geysers in the
USA, Kamojang in Indonesia, Matsukawa in Japan, and Larderello in Italy. The
advantages from an engineering point of view are that reinjection is not required
and the surface facilities of the power plant are simple compared with the water-
dominated geothermal power plants.

1.2.3 Hot Dry Rock Systems


The main characteristic of this conductive type of geothermal systems is that no fluid
exists to transport the large amount of heat stored in the hot rock to a production
well for extraction. Moreover, the reservoir rocks are very tight. Fractures with
sufficient permeability and connectivity need to be generated artificially (through
reservoir stimulation) to create flow paths through which an injected working fluid
can circulate. The temperatures are higher than 250ıC. There are some experimental
geothermal plants, such as Los Alamos in the USA, Hiijori and Ogachi in Japan, and
Soultz in France.

2 Technology and Components

Geothermal power generation is to produce electricity by rotating a turbine directly


with the steam taken out of the deep underground. While in the thermal power
generation the steam is generated by the combustion heat of coal, oil, or LNG, it
54 Geothermal Energy 1025

can be said that in the geothermal power generation, the Earth itself plays a role
of a boiler. Generally, the deeper in the Earth, the higher the temperature gets. It is
assumed to be about 1,000ıC at the depth of 30–50 km, where it can be regarded
as a large thermal storage. To use geothermal energy taken to the ground surface
from production wells for power generation, there are some generation systems
depending on characteristics of geothermal fluid (temperature, pressure, flow rate,
etc.) such as single-flash system, double-flash system, dry steam system, and binary
cycle system.

2.1 Direct Utilization

Geothermal energy for direct-use projects is in the low-to-intermediate-temperature


range, and these resources are more widespread and exist in about 80 countries at
economic drilling depths (from GEA 2012). These projects can use conventional
water-well drilling and commercial heating and cooling equipment, and there are
no conversion efficiency losses. Most projects can be online in less than a year. The
projects can be on a small scale, such as for an individual home, single greenhouse,
or an aquaculture pond, but can also be a large-scale commercial operation, such as
for district heating/cooling, food, lumber drying, and mineral ore extraction.
In addition, carbon dioxide, which often occurs in the geothermal water, can be
extracted and used for carbonated beverages or to enhance growth in greenhouses.
The typical equipment for a direct-use system is illustrated in the Fig. 54.6 and
includes a downhole and circulation pumps, heat exchangers (normally the plate
type), transmission and distribution lines (normally insulated pipes), heat extraction
equipment, peaking or backup plants (usually fossil fuel-fired), and fluid disposal
systems (injection wells). The geothermal energy can usually meet 80–90 % of the
annual heating or cooling demand despite only being sized for 50 % of the peak load.

2.2 Geothermal Heat Pumps

Geothermal heat pumps (GHPs) use the relatively constant temperature of the Earth
to provide heating, cooling, and domestic hot water for buildings. A small amount
of electricity input is required to run a compressor.
A closed loop of pipe is placed vertically (50–70 m deep) in the ground, and
a water-antifreeze solution is circulated through the plastic pipes to either collect
heat from the ground in the winter or reject heat to the ground in the summer. The
efficiency of GHP units is described by the coefficient of performance (COP), which
is the ratio of the output energy divided by the input energy (electricity for the
compressor). The ratio varies from 3 to 6 with the present equipment (the higher
the number, the better the efficiency). In comparison, an air-source heat pump has
a COP of approximately 2 and is dependent upon backup electrical energy to meet
peak heating and cooling requirements (see Fig. 54.7).
1026 H. Saibi et al.

Fig. 54.6 Typical direct-use geothermal heating system configuration

2.3 Electric Power Generation

Geothermal power is generated using steam or a hydrocarbon vapor to turn a


turbine-generator set to produce electricity. A vapor-dominated (dry steam) resource
can be used directly, whereas a hot water resource needs to be flashed by reducing
the pressure to produce steam, normally in the 15–20 % range. Some plants use
double and triple flashes to improve the efficiency. In some cases, using a bottoming
cycle (a small binary plant using the wastewater from the main plant) may be more
efficient. In the case of low-temperature resources, those that are generally below
180ıC, the use of a secondary low-boiling-point fluid (hydrocarbon) to generate the
vapor is needed, resulting in a binary or organic Rankine cycle (ORC) plant. The
different configurations are shown in the Figs. 54.8–54.10.
Usually, a wet or dry cooling tower is used to condense the vapor after it leaves
the turbine to maximize the temperature and pressure drop between the incoming
and outgoing vapors and thus increase the efficiency of the operation. Dry cooling
is often used in arid areas where water resources are limited. Air cooling normally
has lower efficiencies during the summer months when the air temperatures are high
and humidity is low.
54 Geothermal Energy 1027

HEAT EXCHANGER
REFRIGERANT/AIR
(EVAPORATOR)

COOL SUPPLY AIR TO WARM RETURN AIR


CONDITIONED SPACE FROM CONDITIONED
SPACE

EXPANSION VALVE

REFRIGERANT
DOMESTIC HOT WATER
REVERSING VALVE
EXCHANGER
(DESUPERHEATER)
HEAT EXCHANGER
REFRIGERANT/WATER
(CONDENSER)

IN
OUT
DOMESTIC WATER
REFRIGERANT TO / FROM GROUND
COMPRESSOR HEAT EXCHANGER
(GEOTHERMAL)

Fig. 54.7 Geothermal heat pumps in the cooling cycle

The share of each category in the total installed capacity, the annually produced
electricity, and the total number of units is are presented in Table 54.1. In addition,
the average values per unit for the installed capacity and the annually produced
electricity are given. From this overview, the hybrid plants (those using more than
one form of energy) are excluded, but their share is currently approximately zero.

2.4 Combined Heat and Power Generation

More recently, the use of combined heat and power plants (CHP) has made low-
temperature resources and deep drilling more economical. District heating using the
spent water from a binary power plant can make a marginal project economical,
as demonstrated at Neustadt-Glewe, Landau, and Bad Urach in Germany and Bad
Blumau in Austria. A similar outcome was found for high-temperature combined
heat and power plants in Iceland. Options for cascading are shown in the Fig. 54.11,
1028 H. Saibi et al.

Fig. 54.8 Steam plant using a vapor- or dry steam-dominated geothermal resource

Fig. 54.9 Single-stage flash steam plant using a water-dominated geothermal resource with a
separator to produce steam
54 Geothermal Energy 1029

Fig. 54.10 Binary power or ORC plant using a low-temperature geothermal resource and a
secondary fluid of a low-boiling-point hydrocarbon

Table 54.1 Average capacity and electricity produced per plant category and the share of each
category
Average
electricity
Average production Share of the Share of the Share of the
capacity per per unit number of total capacity electricity
Type of plant unit (MW) (GWh/year) plants (%) (%) produced (%)
Binary plant 5 27 44 4 4
Back pressure plant 6 50 5 5 6
Single-flash plant 31 199 27 25 26
Double-flash plant 34 236 12 28 30
Dry steam plant 46 260 12 38 34

where the geothermal fluid is used for a number of applications at progressively


lower temperatures to maximize energy use.

2.5 Numerical Modeling of Geothermal Systems

2.5.1 The Role of Numerical Modeling


Numerical modeling plays a crucial role to support the exploration, characterization,
operation, and optimization of geothermal reservoirs. Evaluating the sustainability
of geothermal energy production requires predictive simulation capabilities that
1030 H. Saibi et al.

Fig. 54.11 Cascade utilization of geothermal energy

capture the key features and processes affecting coupled fluid and heat flow in
the reservoir and in the injection and production wells. Numerical models can
be employed to study generic issues of geothermal reservoir dynamics, including
two-phase water-steam flow through fracture networks; heat transfer mechanisms
between the hot rock matrix and fluids in the fractures; reactive geochemical
transport including scaling due to mineral dissolution and precipitation; coupled
thermal-hydrological-mechanical processes, specifically microseismicity associated
with reservoir stimulation; and interactions between the reservoir, wells, and surface
facilities. In addition to the use of numerical models to increase the fundamental
understanding of these complex systems, simulators are extensively used to model-
specific geothermal fields (O’Sullivan et al. 2001). During exploration, preliminary
geological information and data from geophysical surveys, surface manifestations,
and exploration boreholes can be used for resource estimation and the initial
well field design. Once additional data are available from drilling, well testing,
and well completion, the reservoir properties can be determined, and a natural-
state simulation can be performed to examine the convection patterns, upflow
and recharge regions, the presence of flow boundaries, and the initial condi-
tions prior to exploitation. Predictive simulations can be performed for research
estimation and planning of the field development. Once production is initiated, the
54 Geothermal Energy 1031

pressure, temperature, enthalpy, and chemical signals measured in the observation


or production wells can be used in conjunction with the reservoir simulator to
optimize the day-to-day field operations and to further refine the long-term resource
estimation. Each of these application modes has its specific demands regarding
model complexity and data needs, but all require the modeler to have an excellent
understanding of the tectonic and geologic situation, quantitative information about
the hydrothermal properties of the reservoir rock and bounding units, and knowledge
of the natural-state conditions prior to exploitation.

2.5.2 Modeling Challenges


Simulating hydrothermal reservoirs or enhanced geothermal systems is both con-
ceptually and numerically challenging. The development of a conceptual model may
be the most difficult and most important step. It requires inferring the large-scale
structure of the reservoir from the available geologic information and geophysical
data. At the same time, smaller-scale discrete features need to be detected and
characterized because they may dominate the reservoir behavior under production
conditions. The initial and boundary conditions usually have a significant impact on
the long-term performance of the reservoir, but they are often highly uncertain and
difficult to determine.
A fundamental challenge of geothermal reservoir simulation lies in the strong
coupling of fluid flow and heat transfer under two-phase conditions, involving geo-
logic formations that are fractured with property values that vary over many orders
of magnitude. Phase transitions due to boiling and condensation are associated with
strong latent heat effects and lead to significant changes in the fluid properties, such
as density, viscosity, and internal energy. Moreover, phase interference needs to be
captured by the relative permeability and capillary pressure curves that are highly
nonlinear.
Additional simulation challenges arise as more coupled processes are considered,
specifically reactive geochemical transport and thermally or hydrologically induced
stress changes. Finally, accounting for the flow dynamics in the injection and
production wells, which are strongly linked to the conditions encountered at the
feed points, is numerically demanding because these processes occur on temporal
and spatial scales that are significantly smaller than those characteristic for the flow
and transport in the reservoir.

2.5.3 Modeling Workflow


The typical workflow for developing, calibrating, and testing a geothermal reservoir
model using the available geological, geophysical, hydrochemical, and thermal data
is shown in Fig. 54.12. Realizing that each model is developed for a specific purpose
is essential, i.e., the complexity of the model, the included features, the simplifying
assumptions considered acceptable, and the calibration effort all depend on the
ultimate objectives of the simulations. As mentioned above, the development of a
conceptual model based on the available information is potentially the most crucial
1032 H. Saibi et al.

Fig. 54.12 Geothermal reservoir modeling workflow

step in the process because it ultimately determines the mathematical model, the
selection of the computer code, and the data needed for the model calibration and the
confidence-building exercise. Simulations usually proceed in multiple steps, starting
with reproducing the pre-exploitation, natural state of the reservoir, which yields the
initial conditions for the subsequent simulation of the exploitation phase, provides
the main calibration data, and is used for reservoir management purposes. The time
horizon of these simulations can be extended to look at additional scenarios, the
sitting of additional wells, and long-term predictions of reservoir performance for
resource assessment. These predictive simulations may be accompanied by a formal
uncertainty analysis, which in turn may be used to improve the monitoring systems
and the collection of additional calibration data to improve parameter estimations.
This workflow is approximate as certain stages may be skipped or executed in a
different order, depending on the modeling purpose. Moreover, the process of model
development, calibration, and prediction is iterative, accompanying the exploration,
development, exploitation, and assessment of the actual geothermal reservoir.
54 Geothermal Energy 1033

2.5.4 Modeling Tools


To address the modeling challenges outlined above, a number of numerical sim-
ulators have been developed in academia, government organizations, and private
industry. According to the review by O’Sullivan et al. (2001), the TOUGH suite
of codes appears to be the most widely used simulator for geothermal reservoir
engineering. Most of these simulators handle coupled multiphase fluid and heat
flows using equation-of-state modules to describe the thermodynamic properties of
water and steam. They provide the means to represent fractures, either discretely
or using a dual-continua approach, and wellbore flow is either fully integrated
or enabled through a link to an external wellbore simulator. Tracer and reactive
geochemical transport and coupled mechanical processes are accounted for by
some of the simulators. The inverse modeling capabilities for historical matching
are available, either fully integrated or through a link to a general parameter
estimation package. The governing equations consist of mass balance equations for
each considered component and the use of a multiphase extension of Darcy’s law.
In addition, an energy balance equation is formulated, which includes conductive
and convective terms and accounts for latent heat effects. These balance equations
are discretized in space using finite volume or finite element methods and are
usually solved and fully coupled using an implicit scheme. Efficient and robust
sparse linear equation solvers are used to solve the set of algebraic equations arising
in each Newton-Raphson iteration, which are needed to handle the nonlinearities
inherent in the governing equations. Various pre- and postprocessors are available
to assist with model development and the visualization of results and also to
calculate the properties, such as geophysical attributes, that are needed for further
analysis.

2.5.5 Model Calibration


The calibration of the model against the measurements of the temperatures,
geochemical signals, and directly observable liquid and steam outflows at the
surface is often used to adjust the model structure to better reflect the natural state.
Geophysical observations, specifically resistivity, microgravity, and self-potential
data, contain indispensable information about the structure of the reservoir. These
data can be used for the development of the conceptual model or (in a more
quantitative manner) be included in the numerical model as part of an iterative
or joint inversion framework. In the latter approach, the calculated heat and fluid
flow is translated using a petrophysical relationship into geophysical attributes that
are then either directly compared with corresponding maps created by geophysical
inversions or used to predict the geophysical raw data themselves, which are then
matched in a formal joint inversion along with the thermal, hydrological, and
geochemical data.
The historical matching of production data has the advantage that the data
used for model calibration are on the appropriate scale and reflect the processes
that are relevant to and of interest for the subsequent model predictions. This
1034 H. Saibi et al.

consistency of scale and process is important as the parameters estimated by the


inverse modeling always refer to the specific structure of the calibration model.
Any conceptual error in the model structure inevitably leads to errors in the
estimated parameters, and these errors are propagated through the prediction model
(i.e., the model used to evaluate the long-term performance and sustainability
of the geothermal reservoir). Minimizing the conceptual differences between the
calibration and prediction models is thus essential. Moreover, using complementary
data (e.g., thermal, hydrological, geochemical, and geophysical data) that contain
information about both the processes and the geologic structure of the reservoir
is advantageous. If all these data are included in an iterative model development
framework or a formal joint inversion approach, the resulting model is likely to be
more accurate and more robust against residual uncertainties and will thus be able
to make more reliable predictions of reservoir behavior.

2.5.6 Example
An example of a numerical model of a synthetically enhanced geothermal system
is shown in Fig. 54.13. It consists of a central injection well and two peripheral
production wells. The reservoir is hydraulically and thermally stimulated, leading
to an ellipsoidal region of fractured rock that is modeled using a dual-permeability
approach. In addition, a discrete wide-aperture zone with increased permeability
is formed, providing potential pathways for fast fluid flow that lead to an early
thermal breakthrough. The heat exchange between the wells and the formation
above the model domain is accounted for using an efficient semi-analytical solution
for radial heat transfer. The figure shows the computational mesh on the faces of the
model domain, along with the temperature isosurfaces that reveal reservoir volume
undergoing heat mining. The simulations were performed using the TOUGH code
(Pruess et al. 1999).

2.5.7 Outlook
Although numerical modeling in support of geothermal reservoir engineering can
be considered a mature technology, the need to manage reservoirs in a sustainable
manner or to engineer them poses significant challenges that can be partly addressed
by advanced simulation capabilities. In particular, the ability to numerically examine
coupled thermal-hydrological-geochemical-mechanical effects allows scientists and
engineers to better understand the reservoir behavior to analyze the monitoring data
for an improved characterization of the formation and the fluid and heat flows
under natural state and production conditions and to manage and optimize the
reservoir operations with a reduced risk to the resource and the nearby communities.
Integrating the geophysical information in a quantitative manner into the reservoir
models is another recent advance that has the potential to further improve the
reliability of the predictions made in support of the exploration, characterization,
and prediction of the geothermal reservoir behavior.
54 Geothermal Energy 1035

T [°C]
190
−4000 180
170
160
Depth [m]

150
−4500 140
130
120
110
−5000

−1000
−2000 −500
−1500
−1000 0 ]
−500 [m
Y [m 0 500 X
] 500
1000 1000

Fig. 54.13 Numerical model of a synthetically enhanced geothermal system with the temperature
isosurfaces after 30 years of heat mining by the injection of cold water from the central well and
production from two peripheral wells that intersect the stimulated fracture zone

2.6 Hybrid-Microgravity Monitoring at Geothermal Reservoirs

The gravity method is a potential-field geophysical method. Over the last four
decades, many applications in geothermal, volcanological, and engineering prob-
lems have been applied. Time-lapse microgravity surveys in the geothermal field
showed good results by monitoring the gravity changes with time and estimating
the underground mass changes (Saibi et al. 2005). The gravity method is a nonde-
structive geophysical technique that measures differences in the Earth’s gravitational
field at specific locations.
Microgravity measurement is one of the procedures for geothermal reservoir
monitoring. The production and reinjection of geothermal fluid causes mass move-
ment and redistributions, which can cause measurable gravity changes on the
surface. The mass balance can be monitored, especially the relationship between
production and recharge, in the geothermal reservoir (Fig. 54.14). Microgravity
monitoring has been performed in some geothermal fields. Gravity decreased
approximately 1,000 gal after 30 years in the Wairakei geothermal field in New
Zealand (Allis and Hunt 1986).
1036 H. Saibi et al.

Fig. 54.14 Concept of microgravity monitoring

3 Case Study: Design, Efficiency, Emissions

3.1 Combined Heat and Power Plants: The Icelandic Showcase

Combined utilization of heat and power for geothermal production (CHP) is not a
new application, but recently it has been widely extended in the low-to-medium-
temperature range of the resource, covering different project sizes.
The main reason for a CHP plant is the more efficient use of the entire geothermal
energy extracted from the reservoir fluid, from its initial temperature down to that
of the water discarded to the reinjection stream. In this way, it is possible to strongly
improve the economics of the entire system, making even small temperature
resources exploitable, even if the electricity revenues alone are not enough for a
reasonable payback of the investment.
In the selected case study, a quite complex system is highlighted, the Húsavı́k
Energy CHP system in Iceland (Hjartason et al. 2005).
54 Geothermal Energy 1037

Húsavı́k is a 2,500-inhabitant town in Northeast Iceland, with an economy based


on fishery and services; as typical for that country, geothermal district heating is
the most economic source of domestic energy for the long Nordic winter. The
geothermal field is located about 20 km away from the town, and through three
wells, a total flow rate of 95 L/s of 124ı C hot water is channeled at the first stage
of utilization, a binary power plant (with Kalina technology, i.e., using a mixture of
water and ammonia instead of an organic working fluid), with 1.7 MW of capacity.
The insulation of the 16 km pipe is very good, resulting in only 3ı C of cooling at
the plant inlet. The cooling leg of the power plant is realized using a water supply
at 4ı C, which is heated up to 80ı C and stored in a tank. The power plant is used for
7,000 h/year.
At the second stage, the hot water is discarded from the plant, and the storage
tank is exploited at different temperature levels, obtained by appropriate mixing, for
the following utilizations:
– High-temperature glucosamine factory (121ıC), cooled down to 80ı C.
– Low-temperature dry fish industry (80ıC) (fed also with hot water from the
power plant and the tank).
– District heating of the town (65 L/s at 80ı C) for 4,400 h/year.
– Snow melting system.
– 190 L/s to the bathing lagoon at 30ı C.
– Fish farm with 20 L/s at 30ı C.
Moreover, high-temperature springs of 34 L/s at 100ı C, related to the geothermal
field, are directly used for the following applications:
– Farms (9 L/s at 65–75ıC)
– Greenhouses (9 L/s at 100ı C cooled down to 35ı C)
– Additional 16 L/s at 60ı C to a fish farm
The sketch of the system is given in Fig. 54.15.
The overall energy balance of the system is as follows: the wells produce
339 GWh yearly, and 120 GWh is available from the hot springs; 48 GWy is used
for space heating, 9 GWy for tap water, 12 GWh for electricity production, 10 GWh
for the two levels of industry (for 6,000 h/year), 45 GWh for fish farming, 1 GWh
for snow melting, and finally 177 GWh for the bathing lagoon. Thirty percent of the
energy is reinjected or lost during transportation.
The total capital investment for the entire project was 12 Meuro, with about 66 %
of the cost for the district heating network, and only 4 Meuro for the CHP plant. The
Húsavı́k municipality covered 92 % of the investment.

4 Economic Analysis

As with other renewable energy technologies, geothermal projects have high up-
front costs (mainly due to the cost of drilling the wells) and low operational costs.
These operational costs vary from one project to another due to the size, the quality
1038 H. Saibi et al.

Orkuveita Húsavíkur–Multible use of geothermal energy Hot water storage tank


Flow and temerature at maximum demand 2001
20 l/s Water supply
95 l/s 80°C 1.7 MW
>110°C
60°C–100°C Energy center
4°C–40°C
El. Power Plant 190 l/s
4°C
Hveravellir – Geothermal site
District heating
in Húsavík town

H1
75°C
128°C
Gas

75 l/s 65 l/s 35°C


Snow
80°C 80°C melting
23°C
15°C

10 l/s
H10

80°C
124°C 95 l/s 121°C 80°C
Industry
Gas separator
35°C
H16
Hot springs

210 l/s
115°C 80°C

30°C
19 l/s 16 l/s
100°C 60°C Control
house
100°C

190 l/s
9 l/s

75°C
3 l/s

30°C
75°C

Bathing lagoon
6 l/s

Green houses
35°C Fish farm Industry
Farms Farms 20 l/s

35°C 35°C 30°C 35°C

Fig. 54.15 Húsavı́k cascade utilization system

of the geothermal fluids, and so on but are predictable compared with power plants
that use traditional energy sources, which are usually subject to market fluctuations
in the fuel price.

4.1 Electricity

The cost structure of a geothermal-electric project comprises the following compo-


nents (Goldstein et al. 2012):
– Exploration and resource confirmation: This component includes lease acquisi-
tion and the permitting, prospecting, and drilling of exploration and test wells.
Drilling of these types of wells has a success rate that is typically 50–60 %. The
confirmation costs are affected by the well parameters (depth and diameter), rock
properties, well productivity, rig availability, time delays in obtaining permits or
leasing land, and interest rates.
– Drilling of production and injection wells: Field expansion projects may cost
10–15 % less than a new (greenfield) project because the investments have
already been made in the infrastructure and exploration and valuable resource
information is available. Drilling the production and injection wells has a
54 Geothermal Energy 1039

success rate of 70–90 %. Factors influencing the cost include well productivity
(permeability and temperature), well depth, rig availability, vertical or directional
design, the use of air or special circulation fluids, the use of special drilling bits,
the number of wells, and the financial conditions in a drilling contract.
– Surface facilities and infrastructure: This component includes gathering steam
and processing brine, separators, pumps, pipelines, and roads. Vapor-dominated
fields have lower facility costs because brine handling is not required. Factors
affecting this component are reservoir fluid chemistry, commodity prices (steel,
cement), topography, accessibility, slope stability, average well productivity
and distribution (pipeline diameter and length), and fluid parameters (pressure,
temperature, chemistry).
– Power plant: This component includes the turbines, generator, condenser, electric
substation, grid hookup, steam scrubbers, and pollution abatement systems.
The power plant design and construction costs depend upon the type (flash,
back-pressure, binary, dry steam, or hybrid) and the cooling cycle used (water

Table 54.2 Breakdown of current capital costs for typical turnkey geothermal-electric projects
(2005 US$)
Component
Exploration Drilling Surface
and (1.5–3 km facilities and
Typea Concept confirmation depth) infrastructure Power plant Total
1 US$/kWe 475 1,275 350 1,225 3,325
% capex 14 % 38 % 11 % 37 % 100 %
2 US$/kWe 30 1,275 350 1,225 2,880
% capex 1% 44 % 12 % 43 % 100 %
3 US$/kWe 25 1,008 300 1,175 2,508
% capex 1% 40 % 12 % 47 % 100 %
4 US$/kWe 24 800 274 1,782 2,880
% capex 1% 28 % 10 % 61 % 100 %
5 US$/kWe 205–560 750–1,500 205–750 1,215–2,240 2,025–3,750
% capex 10–15 % 20–40 % 10–20 % 40–60 % 100 %
6 US$/kWe 275–425 750–1,700 425–850 1,500–2,600 3,400–4,300
% capex 8–12 % 20–40 % 10–20 % 40–60 % 100 %
7 US$/kWe 530 3,350 1,350 4,720 9,950
% capex 5% 34 % 14 % 47 % 100 %
a
Type:
(1) Greenfield project, 40-MWe single-flash power plant, 200ı C, wells to 2 km depth
(2) Expansion project, 40-MWe single-flash power plant, 200ı C, wells to 2 km depth
(3) Expansion project, 4  25 MWe single-flash power plant (100 MWe), wells to 2.2 km depth
(4) Expansion project, 25-MWe single-flash power plant, wells at 1.8 km depth on average
(5) Greenfield project, 10–50 MWe condensing power plants
(6) Greenfield project, 10–20 MWe binary cycle power plants
(7) Greenfield project, 4 MWe binary cycle power plant, low temperature, wells to 2,750 m depth
1040 H. Saibi et al.

or air cooling). Other factors affecting power plant costs are fluid enthalpy
(resource temperature) and chemistry, location, cooling water availability, and
the economies of scale (a larger size is cheaper).
Table 54.2 presents the breakdown of current capital costs (capex) for typical
geothermal-electric projects in 2005 US$.
Labor and material costs are estimated to account for 40 % each of the total
project construction costs. Labor costs can increase by 10 % when a resource is
remotely located. In addition to raw materials and labor, the choice of power plant
size is a key factor in determining the ultimate cost of a plant. For example, using
a single 50-MWe plant instead of multiple 10-MWe plants can decrease the power
plant costs per kilowatt by approximately 30–35 % for binary systems. The installed
cost per kilowatt for a 100-MWe flash steam plant can be 15–20 % less than that of
a 50-MWe plant.

4.2 Direct Uses

Direct-use project costs have a wide range, depending upon the specific use, the
temperature and flow rate required, the associated O & M and labor costs, and the
income from the product. In addition, the costs for new construction are usually less
than those for retrofitting older structures. The cost figures given below in Fig. 54.16
are based on a temperature climate typical of the northern half of the United States or
Europe, and the heating loads would obviously be higher for more northern climates,

Fig. 54.16 Typical cost for direct utilization


*The costs for residential geothermal heat pumps do not include the drilling cost.
54 Geothermal Energy 1041

such as Iceland, Scandinavia, and Russia. Most figures are based on the cost in the
United States (expressed in 2005 US$) but would be similar in developed countries
and lower in developing countries.

5 Summary

Geothermal heat is an energy source that – if properly managed – has the potential to
be abundant, versatile, environmentally acceptable, cost-effective, and sustainable
Its sustainability is mainly a result of the vast amount of heat stored in the
subsurface. Extracting this heat, however, is challenging and requires advances in
exploration and drilling technology, novel approaches to stimulate reservoirs, and
potentially new working fluids. Moreover, exploitation of thermal energy must be
optimized and carefully managed, which in turn requires good characterization
methods of the geologic formation and fluid movements within the reservoir.
In addition to high-temperature hydrothermal or engineered geothermal systems
(EGS) for electricity production, residual heat from deep geothermal fluids or low-
temperature resources at shallow depths can be used for a vast variety of applications
and joint uses on different temperature levels for a high overall degree of efficiency.
Finally, reinjection of geothermal fluids closes the loop, providing pressure support
and the working fluid for a sustainable energy production system.
The use of computer modeling in the planning and management of the develop-
ment of geothermal fields has become standard practice during the last 20–35 years.
Geothermal reservoir simulation was recently used in reservoir engineering practice
with more complex three-dimensional models with a graphical interface. Reservoir
evaluation is an important phase prior to installing a geothermal power plant.
Geophysics is an important study phase of the perspective geothermal field. Main
methods include the following:
– Microgravity method: It is an effective geophysical technique that helps monitor
the geothermal reservoirs by measuring the underground mass changes using the
Gauss theorem in the geothermal reservoirs due to the production and injection
phases and natural recharge or discharge. Repeated microgravity measurements
at geothermal fields are recommended to maintain and advance the sustainable
utilization of geothermal resources.
– Electromagnetic method: Monitoring surveys can give information about fluid re-
distributions, flow rates, and flow directions caused by production and reinjection
in the geothermal reservoir from the changes of underground electrical resistivity
and spontaneous potential at the surface. The shallow survey is preferably
performed using transient electro magnetic soundings. Deep survey is made by
Magneto-Tellurics and can also be applied to delineate faults.
– Microseismic method for the evaluation of the heat source, fluid-flow channels’
permeability, and reservoir properties. Combining geothermal reservoir simula-
tors with geophysical postprocessors will enhance the numerical modeling of the
geothermal reservoirs and help managers to make good decisions about the future
plan of the geothermal power plant.
1042 H. Saibi et al.

Geothermal projects have high up-front costs (mainly due to the cost of drilling
the wells) and low operational costs. The geothermal plants have low recurring fuel
costs. The costs of heat from direct uses of geothermal heat are competitive with
market energy prices.
More international courses on geothermal energy need to be developed and also
need to increase the number of engineers all over the world by developing university
undergraduate programs in the geothermal field to teach younger generations to take
responsibility when using this natural and sustainable energy for the safety of the
human race.

6 Cross-References

Geothermal Energy

References
R.G. Allis, T.M. Hunt, Analysis of exploitation-induced gravity changes at Wairakei geothermal
field. Geophysics 51, 1647–1660 (1986)
R. Bertani, Geothermal power generation in the world 2005–2010 update report. Geothermics 41,
1–29 (2012)
GEO-Geothermal Education Office, Geothermal energy introduction (2001), http://geothermal.
marin.org/index.html
GEA, Global Energy Assessment – Toward a Sustainable Future. (International Institute for
Applied Systems Analysis, Vienna, Austria and Cambridge University Press, Cambridge and
New York, 2012), p. 1888
B. Goldstein, G. Hiriart, R. Bertani, C. Bromley, L. Gutierrez-Negrin, E. Huenges, H. Muraoka,
A. Ragnarsson, J. Tester, V. Zui, Contribution to special report renewable energy sources
(SRREN), International panel on climate change (IPPC) (2012)
H. Hjartason, R. Maack, S. Jóhannesson, GHC Bull. 26, 7–13 (2005)
International Energy Outlook, U.S. Energy Information Administration, 292 (2011)
J.W. Lund, D.H. Freeston, T.L. Boyd, Direct utilization of geothermal energy 2010 worldwide
review. Geothermics 40(3), 159–240 (2010)
M.J. O’Sullivan, K. Pruess, M.J. Lippmann, State of the art of geothermal reservoir simulation.
Geothermics 30, 395–429 (2001)
F. Pirajno, Hydrothermal Mineral Deposits (Springer, Berlin, 1992), p. 709
K. Pruess, C. Oldenburg, G. Moridis, TOUGH2 user’s guide, version 2.0, report LBNL-43134,
Lawrence Berkeley National Laboratory, Berkeley, CA (1999)
H. Saibi, J. Nishijima, S. Ehara, Reservoir monitoring by repeat microgravity measurement at
Obama geothermal field, southwestern Japan. Geothermal and volcanological research report
of Kyushu University, 1, No. 14, 27–31 (2005)

View publication stats

You might also like