0% found this document useful (0 votes)
12 views

Geothermal resources classification review Saeid

Uploaded by

camila.medele
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
12 views

Geothermal resources classification review Saeid

Uploaded by

camila.medele
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 45

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

net/publication/363586429

Geothermal Resources Classification – A Review

Conference Paper · September 2022

CITATION READS
1 1,251

1 author:

Saeid Jalilinasrabady
GNS Science
85 PUBLICATIONS 1,326 CITATIONS

SEE PROFILE

All content following this page was uploaded by Saeid Jalilinasrabady on 26 January 2023.

The user has requested enhancement of the downloaded file.


TRANSACTIONS
VOLUME 46

USING THE
EARTH TO SAVE
THE EARTH
2022 Geothermal Rising Conference

GEOTHERMAL RISING
1120 Route 73, Suite 200
Mount Laurel, NJ 08054 • USA
GRC TRANSACTIONS, VOLUME 46

Using the Earth to Save the Earth

2022 GEOTHERMAL RISING CONFERENCE

GEOTHERMAL RISING
1120 Route 73, Suite 200
Mount Laurel, NJ 08054 · USA
Copyright © 2022 Geothermal Rising
Printed in the United States of America

Notice

All views and conclusions in this publication are those of the individual authors of the
papers in which they are contained, and should not be interpreted as necessarily representing the
official policies and/or views/recommendations of the Geothermal Resources Council (GRC). Draft
papers were reviewed by the GRC 2021 Technical Program Committee. Final paper submissions
were not reviewed for scientific and/or editorial correctness prior to publication. The papers
presented in this publication are reproduced as faithfully as possible by Geothermal Rising. The
contents of papers in this volume, however, are solely the responsibility of the authors.

ISSN: 0193-5933
ISBN: 0-934412-28-6
Geothermal Resources Classification – A Review

Saeid Jalilinasrabady

[email protected]
Dept. of Earth Resources Eng., Graduate School of Eng., Kyushu University, Fukuoka, Japan

Keywords:
Geothermal resources, Classification, Exergy, Sustainability and Resource assessment

Highlights:
A comprehensive review has been done on geothermal resources classification.
Details of key existing classification methods and their pros and cons were discussed.
The stored heat method is the most well-established.
A combination method was discussed which is easy to understand for technical and non-
technical counterparts.

ABSTRACT

Renewable energy resources are gaining popularity and they have proven their reliability to
achieve a sustainable future, their efficiency increases due to advancement in innovative
technologies, and that leads to less uncertainty and increased contribution to the energy mix of
future society.
Geothermal energy carries its own uncertainty due to its nature, it requires high capital
investment owing to the very high cost of well drilling and the journey from exploration to
exploitation can be very stressful for investors and decision-makers. Since the nature of
geothermal resources dictates their method of utilization, it is important to categorize available
resources. There is no consensus on the classification of geothermal resources. Most scientists,
from geologists to engineers, agree on the term temperature. However, temperature or enthalpy
alone cannot describe the nature of fluids. Other classification methods discussed in this paper
are assessment confidence, Australian, Canadian, and GEA Codes, Geologic Setting,

1651
thermodynamic properties, resources potential, stored heat, and geothermal resources utilization
including direct uses. Even within the framework of each classification, there are conflicting
opinions between scientists. Well-organized and easily understandable frameworks for
classifying geothermal resources are essential for the assessment, exploration, development, and
reporting and this paper aims to illustrate a comprehensive analysis of existing methods to come
up with a recommendation of the most appropriate method for classification of the geothermal
resources.

1. Introduction
Geothermal heating has been used since Roman times for bathing, cooking, and as a way of
heating buildings and spas using sources of hot water and hot steam that exist near the earth’s
surface. Water from hot springs is now used worldwide in spas, for space heating, and
agricultural and industrial uses [1].
Geothermal energy is a clean and renewable source of energy characterized by a reservoir from
which heat can be extracted economically and utilized for generating electric power or any other
suitable industrial, agricultural, or domestic application. The geothermal system is composed of a
hydrological system, the recharge zone, subsurface parts, and the outflow of the system.
Geothermal reservoir indicates the hot and permeable part of a geothermal system pressurized,
either by artesian or through boiling that may be directly exploited. Common rocks are volcanic
rocks as well as limestone, shale, and granites.
The importance of geothermal energy originates from the availability and continuity of the
source for heat exploitation, and the fact that it is clean energy that produces low harmful
emissions and ensures cost stability for end-users. Furthermore, geothermal energy presents no
geopolitical risk and can be used with the current industrial technology. The main challenges and
limitations for geothermal energy are that most of the high-temperature resources are
concentrated in specific areas usually active tectonics, and their remote optimal energy site
locations have made it difficult for direct use utilization in cities and potential customers.
Technological barriers lie mainly in high exploration and high initial investment costs for
electricity production, long-term investment return, and the risk of failure during the exploration
and drilling phase as well as low public acceptance in some countries.
Exploration of geothermal resources has advanced within the last decades employing
applications from new multidisciplinary science and technologies. GIS-supported certainty factor
(CF) models were proposed for the assessment of the geothermal potential of Tengchong County
in the southwest of China [2]. GIS was used to assess the geothermal potential in western
Anatolia, Turkey [3]. Geothermal favorability mapping by advanced geospatial overlay analysis
was applied to a case study in Tuscany, Italy [4]. A quantitative method for the assessment and
mapping of the shallow geothermal potential was proposed by Casasso and Sethi [5]. Artificial
neural networks for the generation of geothermal maps of ground temperature at various depths
by considering land configuration were proposed for Cyprus [6]. Estimations of heat mining
potential using sCO2 (Supercritical CO2) were performed using the TOUGH2 computer
software [7].

1652
Falcone et al., [8] critically assessed conventional and unconventional deep geothermal well
concepts, focusing on the basic Borehole Heat Exchanger (BHE) concept. A novel single-well
geothermal system for hot dry rock geothermal energy exploitation was developed by Huang et
al. [9].
Geothermal energy resources are characterized by geologic settings, intrinsic properties, and
viability for commercial utilization [10]. Multiple classification schemes have been devised to
explain and categorize geothermal energy systems. Example classifications include those
developed by the United States Geological Survey [11], [12], National Renewable Energy
Laboratory[13], and International Geothermal Associations [14]. These classification systems
have mixed geology, engineering, and resource terminology together in a manner that is difficult
to present to an investor or someone not familiar with the geothermal industry. The broadest
geothermal classification and the easiest to understand is discussed at the Geo-Heat Center at the
Oregon Institute of Technology [15]. This classification is based entirely on temperature ratings
of low (<90oC or 194oF), moderate (90 – 150oC or 194 – 302oF) and high (>150oC or >302oF)
and is loosely tied to the potential uses of the geothermal heat energy [16]. This classification
approach is important for the public who is familiar with temperature but has limitations among
professionals.
Yuafrinaldi et al., [17] provided an overview of three different established resource
classifications in line with the Indonesian National Standard (SNI-1998), they also used the
petroleum resource classifications system as a guideline for leveling geothermal resource
classifications. The power density method for estimating potential power generation capacity
was recommended for geothermal prospects at an exploration or early appraisal stage by
Cumming [18]. A study by Wilmarth and Stimac [19] suggested that fault-based geothermal
systems tend to be low to moderate-temperature and have low power density, volcanic arc
systems tend to be moderate to high-temperature and have moderate to high power density, and
rift systems tend to be high-temperature and have high power density.
Suyama [17] proposed that there are three categories of geothermal resources with huge resource
bases including the hydrothermal convection system, the hot igneous system, and the regional
conductive environment.
At least four other classification systems were suggested by different organizations over more
than the past three decades of geothermal investigations. These schemes were devised to describe
geothermal resource type using scientific and engineering terms that are more closely related to
the geology of the resource when compared to a solely temperature-based description.
There are fundamental differences in prescribing a classification system that depends on whether
it aims at being a reporting standard, a set of guidelines, a set of definitions, a code, or a protocol
[20]. Each of the above options carries a substantially different ability to reinforce the goal of
reducing the uncertainty in defining the value of a given geothermal resource and thus, the risk to
investors.
It is also a challenge to develop a system that can equally satisfy the requirements of different
potential end-users of the same geothermal resource including governments, field owners,
operators, investors, reserves auditors, insurance companies, international energy associations,
agencies, and councils.

1653
To achieve a feasible and sustainable geothermal resources development, a sound and reliable
classification could be a good starting point for the investors and developers. If one considers the
different possible uses of geothermal energy, and different investment targets are equally
possible, yet they carry substantially different risks and uncertainties in their identification and
estimation such as the heat source, the reservoir, the fluids stored within it, and/or their pressure,
the temperature of the resource, the heat stored within the reservoir, the recoverable fluid
volume, the recoverable heat, the recoverable power, and the net profit/revenue from a given
development project.
Resource classification is a key element in the characterization, assessment, and development of
energy resources, including geothermal energy. Geothermal resource issues such as location,
quality, the feasibility of development, and potential impacts are important to the geothermal
stakeholders [10]. Their understanding and use of globally recognizable and easily understood
terms in describing and addressing these issues are important. These terms must cover the main
aspects of geothermal parameters from the fundamental geological nature of geothermal
resources to the practical technological and economic aspects of resource exploitation while
remaining understandable to the broad community of non-specialists [21].
1.1 Geothermal classification methods
Well-organized and easily understandable frameworks for classifying geothermal resources are
necessary for the assessment, exploration, development, and reporting. Stakeholders at all levels
of government, within the geothermal industry, and among the general public need to be able to
use and understand consistent terminology when addressing geothermal resource issues such as
location, quality, the feasibility of development, and potential impacts [21]. Up to date, several
geothermal resource classifications are proposed by different authors and institutions. Several
classification methods have been proposed and have been changing over time with geothermal
research and development as a result of the expanded scope of exploitable geothermal resources.

Saemundsson [22] emphasized that geothermal systems and reservoirs are classified based on
various aspects, such as reservoir temperature, entropy, physical state, or their nature and
geological setting. The sub-classification divided low-temperature systems into shallow
resources, sedimentary low-temperature systems, geo-pressured systems, and convective low-
temperature systems. The high-temperature fields are found without exception in volcanically
active areas; their sub-classification includes rift zone regime geothermal systems, hotspot
volcanism, and compression regions.
Since the 1960s and 1970s, numerous researchers have focused on the classification of
geothermal resources (i.e.[23];[24];[25]). Some of those definitions are still in use by developers.
For example, the basic framework for geothermal resource characterization and assessment of
Muffler and Cataldi [26] is foundational to recent resource assessments by the United States
Geological Survey and other organizations [27],[28].
Armstead [29] classified geothermal fields as semithermal (hot water up to 100°C at the surface),
hyperthermal wet field (producing hot water and steam at the surface), or hyperthermal dry fields
(producing dry saturated or superheated steam).

1654
In most classification methods, the temperature is the main parameter. The main reason for this
popularity is its simplicity and ease of measurement. Further, most scientists, from geologists to
engineers, agree on the term ‘temperature.’ However, in each classification method, there are
different boundaries for classifying resources, and there is no clear agreement on the temperature
range of each category. In addition, temperature or enthalpy alone cannot describe the nature of
fluids. Two geothermal fluids at the same temperature may have completely different abilities to
do work; they can have the same temperature with different phases, such as saturated water or
saturated steam [30].
Geothermal classification has no globally agreed standards, guidelines, and codes to assess,
quantify and classify the resources [31]. This may lead to increased resource uncertainty, higher
investment risk, less confidence in development, and misunderstanding among different
stakeholders.
The following are the geothermal stakeholders who are interested in geothermal classification
information to inform their decisions on different issues [32].
a. Governments
Governments need clear and consistent information on geothermal classification and resource
estimates to better understand the total resource base, to make well-informed decisions on energy
security, to formulate an energy strategy and policies that will guide the country on geothermal
development and regulation.
b. Global Organizations
These organizations and institutions need to assess and compare global energy systems and
different energy sources including renewable energy.
c. Renewable Energy Sector Participants
Enhanced overview of asset values, provide a measure of comparability with traditional energy
systems, offer a basis to estimate the scale of each renewable resource, provide reliable estimates
based on best practices and common standards.
d. Investment community
They need the information to guide in comparing and contrasting different investment
opportunities and enhance portfolio valuation.
e. Locals
Clear and easy to understand methods of geothermal classification will lead to better
communication with the local population to increase social acceptance of geothermal energy
development.
f. Academic and Research Institute
Research and development of educational institutes play an important role to achieve better
classification methods for geothermal resources, this has been a gradual and continuous process.

1655
Therefore, geothermal resource classification systems should reflect the current state of
knowledge regarding geothermal technology and serve as an effective means of characterizing,
quantifying, assessing, and providing consistent communication within and outside the
geothermal community.
The main objective of this study is to review the proposed geothermal resource classification
methods and to make a comparative analysis. Furthermore, the review aims to recommend the
best classification method. In this paper, a comprehensive review of the existing geothermal
classification methods has been done from the widely published papers in international journals
in addition to other research findings funded by renewable energy agencies on the classification
and assessment of geothermal resources. A comparative analysis of the methods dependent on
the parameters utilized in the classification has also been done to come up with a
recommendation of the most appropriate method for the classification of the geothermal
resources.
The following are the main classifications categories that will be discussed in this review:
a) Assessment Confidence
b) Geologic Setting
c) Temperature and other thermodynamic properties
d) Other classification methods

2. Assessment Confidence
Several authors have classified the geothermal resources under this category.
2.1 Classification by accessibility and discovery status
The subdivision of the geothermal resource base was done by Muffler and Cataldi [26] through a
McKelvey diagram as shown in Figure 1. The degree of geologic assurance regarding resources
is set along the horizontal axis whereas the economic/technological feasibility which is
controlled by the depth is set along the vertical axis. According to this classification system, the
geothermal resource base is all of the thermal energy in the Earth’s crust beneath a specific area,
measured from the local mean annual temperature.
The geothermal resource is that fraction of the resource base at depths shallow enough to be
tapped by drilling in the foreseeable future that can be recovered as useful heat economically and
legally at some reasonable future time. The geothermal reserve is the identified portion of the
resource that can be recovered economically and legally at present using existing technology
[26].

1656
Figure 1: McKelvey diagram [26].
2.2 Geothermal classification by potential
Rybach [33] proposed a method to address the challenges posed by previous potential
classifications methods which estimated geothermal potential for each country without stating
which kind of potential referred to the recovery factor applied. In this method, the resource is
categorized into theoretical, technical, economic, sustainable, and developable; decreasing
successively in size with the potentials becoming more and more realizable and rewarding
financially.

Figure 2: Classification by potential [33].

1657
As it can be seen from Figure 2, theoretical is the physically usable energy supply often referred
to as the heat in place while the technical is the percentage of theoretical potential that can be
used with current technology. Economic is the time and location dependent percentage of
technical potential that can be economically used whereas sustainable is the percentage of
economic potential that can be used by applying sustainable production levels subject to
regulations, environmental restrictions [34].
Enhanced geothermal systems in Europe were evaluated with estimation and comparison of the
technical and sustainable potentials [35].
Figure 3, is an example from Goldstein et al., [36] for practical classification of geothermal
resources based on the method proposed by Rybach [33]. These were done to define and describe
global geothermal potentials but it can also be applied to the regional or local level.

Figure 3: Global geothermal potential [36].

2.3 Geothermal assessment and estimation protocols


2.3.1 Australian and Canadian Protocol Terms
The Australian Geothermal Reporting Code Committee (AGRCC) [37] produced the world’s
first uniform guide on how to report geothermal data to the market in 2008 with a revision in
2010. The main aim of the Geothermal Reporting Code and Geothermal Lexicon for Resources
and Reserves “Definitions and Reporting” was to provide “a methodology for estimating,
assessing, quantifying and reporting geothermal resources and reserves”. The key aspects of the
code were later adopted by the Canadian Geothermal Code for Public Reporting, by the
Canadian Geothermal Code Committee [38].

1658
They classified a geothermal resource into three categories namely; inferred, indicated, and
measured, which represent three different levels of geological knowledge and probability of
occurrence as shown in figure 4. The reserves were defined as probable and proven based on the
likelihood and reliability of the modifying factors and the type of resource. The modifying
factors depend on economic, environmental, and political context, and assess the commercial
viability of the resources.

Figure 4: Classification Method (AGRCC) [39].

2.3.2 Protocol for Estimating and Mapping Global EGS Potential [37]
Estimates of EGS potential derived using the protocol will not be ‘final’. They will continue to
be refined as more relevant data become available. Theoretical Potential will be refined as new
geological and geophysical data improve our understanding of the thermal structure of the crust.
Refinements here are expected to be gradual. Technical Potential will be refined as technological
advancements in drilling, power conversion and legal regimes allow greater amounts of the
theoretical potential to be realized. Changes here are expected to be sudden and dramatic.
2.3.3 Resource Assessment Protocol for GEOELEC, [38].
In this project, an overview of the location of geothermal resources which can be developed in
the 2020 and 2050 timeline horizons for electricity production in the pan-European region, has
been investigated. This protocol has been based on the following work:
• A protocol for estimating and mapping the global EGS potential [37].

1659
• Australian code for reporting of exploration results, geothermal resources, and
geothermal reserves: the geothermal reporting code [39].
• The Canadian geothermal code for public reporting [40].
• Application of the United Nations Framework Classification for Resources (UNFC) to
Geothermal Energy Resources [31].
A set of 14 case studies from Australia, Germany, Hungary, Iceland, Italy, Netherlands, New
Zealand, Philippines, and the Russian Federation are presented in this report to facilitate a better
understanding of the specifications and the uniform application of UNFC to geothermal
resources. These application examples illustrate the classification of a range of different
geothermal resource scenarios in a manner consistent with other energy resources.

3. Geologic Setting
Geothermal systems appear in diverse geological environments all over the world. The
geological setting of a geothermal system has a fundamental effect on the potential temperature,
fluid composition, and reservoir characteristics. Geothermal systems are commonly associated
with areas of young tectonism and volcanism for example along active plate margins and inter-
plate hot spots than in stable cratonic regions.
Important factors for geothermal utilization are how much heat is stored at a drillable depth and
the ability to recover this heat in the surface at an economic rate for specific use (feasible
development in depth of interest). Geothermal systems can be classified based on the effects of
geological controls and structural plate tectonic positions on the thermal regime and heat flow,
hydrogeological regime, fluid dynamics, fluid chemistry, faults and fractures, stress regime, and
lithological sequence.
Several geothermal classification methods based on geological controls have been proposed by
several authors.
3.1 Geological classification method proposed by Inga [41]
These classification schemes proposed classifying geothermal resources based on their heat
transfer regimes and geological settings, into two broad categories of conductive systems and
convective Systems.
3.1.1 Conduction Dominated Geothermal Systems
Conduction dominated geothermal systems are characterized by low to medium enthalpy
resources. They occur in passive tectonic plate settings with no significant recent tectonics,
volcanism occurrence, or no asthenospheric anomalies.
Conduction dominated geothermal systems occur in low permeability domains such as tight
sandstone carbonates or crystalline rock and the working fluid can be present or should be
supplied through EGS technology to be utilized on an economic level.

1660
3.1.1.1 Igneous geothermal plays-basement type
Crystalline rocks such as granitic rocks host vast resources of heat energy in igneous provinces,
which often underlie large areas of continents. These low porosity low permeability rocks require
reservoir development by stimulation techniques to allow circulation between injector and
producer wells, with the rock mass acting as the heat exchanger, referred to as Hot Dry Rock
(HDR). Reasonably efficient electricity production requires water temperatures exceeding 180°C
[70]. The generally accepted performance target for a well doublet is a production rate of 50 l/s
and minimum rock temperature of 200°C (e.g. [71,3]).
3.1.1.2 Conduction geothermal systems in orogenic belts
Conduction dominated geothermal play systems without active igneous activity cover different
types of geologic settings located within orogenic belts sand associated foreland basins.
Advective heat transport may play a role in mountainous areas of the orogenic belt type, where
high permeability domains and deeply rooted faults allow deep circulation of meteoric water and
form hot springs.
In the adjacent mountain belt, groundwater flow and thermal gradient are strongly influenced by
large hydraulic heads resulting from the pronounced topographic relief [42] & [43]. The great
depth and small width of mountain belt valleys result in relatively shallow penetration of
recharged water, which then discharges on valley floors or shallow valley slopes [44]. Thermal
highs occur underneath high mountains and thermal lows beneath the valleys, resulting in
varying local thermal gradients due to meteoric water circulation. Beneath high mountains at
about 15–20°C and beneath deep valleys 30–50°C [43] & [45].
3.1.1.3 Conduction geothermal systems in intracratonic basins
In sedimentary basin settings, conduction-dominated hydrothermal plays are located in deep
aquifers heated by a near normal heat flow. Effectively, sedimentary basins host prime aquifer
systems from where the thermal water can be produced and utilized. The exploration target is to
identify high porosity high permeability or high porosity/low permeability domains at different
temperature levels and normally at great depths.
The success of EGS in tight, hot sedimentary aquifers may be strongly affected by the storage
capacity expressed by the porosity of the host rock [4]. The heat content of the fluid in the
porous layer setting is strongly affected by the basin geometry, an artefact of the basin type, and
evolution.
3.1.2 Convection dominated geothermal systems
Convection dominated geothermal systems are associated with high enthalpy resources and
occur at plate tectonic margins, or settings of active tectonism or volcanism [46] & [47].
Different tectonic settings where convection dominated geothermal systems occur include:
a. Magmatic arcs above subduction zones in convergent plate margins e.g. the Sunda arc or
the Philippine-Japan arc;

1661
b. Divergent margins located within oceanic settings e.g. the mid-Atlantic ridge, or
intracontinental settings e.g. East African rift;
c. Transform plate margins with strike-slip faults e.g. the San Andreas or Alpine faults;
d. Intra-plate ocean islands formed by hotspot magmatism e.g. Hawaii.

These convective systems have two basic classes depending on the source of the thermal energy;
volcanic and non-volcanic. A volcanic convective system derives its thermal energy from a
magma body dominated by heat convection. The heat sources for such systems are hot intrusions
or magma in volcanic complexes such as calderas, at plate hot spot areas.
Non-volcanic convective systems derive their thermal energy from meteoric water that has
heated up by deep circulation in high heat flow areas of the earth hence there is no magma body
associated with this system [41].
Igneous systems can induce both conduction and convection-dominated geothermal systems.
The difference is that conduction-dominated systems in or close to igneous rocks are related to
high radiogenic heat production from high heat-producing element-rich granites, but no active
volcanism and minor or no active tectonism occurs [48] & [49].
3.1.2.1 Magmatic geothermal systems with a magma heat source
Magmatic play systems can be found in regions with active basaltic volcanism at divergent plate
margins [47] and along continent-continent convergent margins with recent plutonism [50].
Magma chambers in volcanic fields, with their parental melts, recharge of basalt and crystallized
melts, control fluid chemistry, fluid flow, and the overall geothermal play system. These systems
can be separated into an up-flow zone and an outflow zone [9].
3.1.2.2 Magmatic geothermal systems with Plutonic heat source
Plutonism controlled geothermal systems are typically located along continent-continent
convergent or transform margins with recent magmatism and with or without recent recharge of
meteoric water [44]. The heat source pluton crystallized from magma and slowly cooling below
the surface, ranging from hundreds of meters to some kilometers in dimension. The presence and
scale of a heat source in these play types are controlled by the age of magmatism. Active and
recent magmatism commonly indicates a viable underlying heat source [39], while inactive or
extinct magmatism may be reflected by large scale igneous intrusions at greater depth, with
remnant heat and heating by radioactive decaying granitic rock [41].
3.1.2.3 Non-Magmatic convection controlled geothermal systems
Non-magmatic convection dominated geothermal play systems are characterized by fluids
circulation deep within the crust and transport heat to earth surface along permeable faults by
either fault-controlled or fault leakage controlled [11]. The heat source is from zones of thinner
crust corresponding to regions of upwelling asthenosphere and higher heat flow where the Moho
has moved closer to the surface [51].
In purely fault-controlled play systems, convection occurs along the fault and is commonly
combined with infiltration of meteoric water along the fault [11]. In fault leakage controlled play

1662
systems, the fluid leaks from the fault into a permeable concealed layer where the leak may
move to the fault zone and surface.
3.1.2.4 Case Study: Application of the method proposed by Inga, 2014 [41].
Moeck, et.al, [52], applied the geothermal play method in the German case, the study was carried
out on molasses basin as a modern concept to categories geothermal resources related to crustal
permeability. This study presents the first geothermal play cluster of Germany and new terms for
play-based geothermal exploration such as play level and play element, required for an ongoing
categorization of geothermal play types. The focus is on the depth relation of the porosity –
permeability/transmissivity variation and compiles data from the geothermal prospects in the
Molasse Basin. New data interpretation of the deep well Gerestried, drilled in 2017 south of
Munich, is integrated into this study and presented for the first time in a geothermal play context.
Table 1, summarizes the application of the geothermal play method in convection and
conduction dominated geothermal resources.

Table 1: Conduction and convection dominated geothermal resources

Conduction Dominated Systems

Basement Type Intracratonic Basin Orogenic Belt

Cooper basin Australia North German basin Molasse basin, Germany


[49] [53]

Fenton hill new Mexico/USA, Williston Basin of The Western Canada


Rosemanowes/UK Saskatchewan, Canada Sedimentary
Basin of the North Basin associated with the
Dakota, U.S.A Rocky Mountains

Convection Dominated Systems

Volcanic field type Plutonic type Extensional domain type

Mid ocean Ridges, Reykjanes Recent magmatism, Metamorphic core


Iceland [54] Geysers, no recharge complexes, Great Basin in
U.S.

Subduction zone Taupo New Recent plutonism Extensional terrain,


Zealand [55], Java Indonesia Laderello [56] ,Italy, Western Turkey

1663
[99] recharge

Mantle plumes, Hawaii USA. Recent plutonism, Intra continental rift


southern periphery grabens, E.A rift, Upper
[57] Alpine orogeny Rhine graben in Central
Europe, Sous-Foret,
France

3.2 Geological classification method proposed by Richard J. [58]


These classification methods consider five parameters as key to geothermal classification
namely;
a. Geologic environment and features;
b. Geological features;
c. The crustal heat source;
d. Resource category;
e. Rock type.

Table 2, explains details of this geological classification method.

Table 2: Geological classification method by [58]

1664
a. Geologic environment

The geothermal resources are categorized into two basins on plate tectonics and geology as plate
margin and intraplate-related environments. The plate margin environment is further subdivided
into convergent (compressional), divergent (extensional), and transform (strike-slip)
environments. Large-scale geologic features exist within each of these environments that
represent the target regions for geothermal power exploration.
The intraplate environment are areas that are not presently along or very near to active plate
margin environments. These are characterized by mantle plumes or hot spots, cratonic basins,
passive margin basins, extensional terrain, and the basement complex.
b. Geological features
This category is the large-scale geologic feature that can be found in either the plate margin or intraplate
environment. Thus a volcanic arc complex can be expected to form in a convergent plate margin whereas
a pull-apart basin will be within a strike-slip environment.

c. The crustal heat source

The magmatic activity would tend to dominate along with plate margin environments where
shallow depths into the Earth to where tectonic activity occurs provides conduits for magma to
rise into the crust heating up a broad region or provides more local sources of heat and intraplate
environment where mantle plumes and extensional terrain is found.
Another source of heat is a conductive gradient from deeper layers for example in sedimentary
basins. If the rate of heat removal from some regions is constant, then a continuous flow of heat
will be established into this disturbed region as a thermal gradient established between this
cooler location and the hotter rock forming which is the source of heat that is extractable for
power generation.
Basement rocks such as granite contain higher concentrations of radiogenic minerals when
compared to other types of crystalline rocks. The heat generated within these rocks becomes a
target for extraction if overlying rock strata have lower conductivity and act as a thermal blanket,
trapping the heat within the granite, not allowing it to readily escape.
d. Resource Category

Resource category refers to the medium within which the heat is to be found and produced.
Tapping heat within the plate margin environment and within mantle plumes and extensional
features will be dominated by steam or hydromagmatic production.
Within sedimentary basins, two distinct geothermal resource categories are identified i.e stranded
and co-produced. The stranded geothermal resource represents warm to a hot fluid that is in a
geopressured or hydrostatic condition that was not produced when the basin was drilled
originally for oil and gas. A stranded geopressured resource consists of hot brine often saturated
with methane and found in large, deep aquifers that are under higher pressure due to water
trapped in the burial process. These resources are often found in sedimentary strata at depths of 3
km to 6 km with water temperature could be ranging from 900°C to 2000°C.

1665
The hot dry rock (HDR) category represents dry rocks with high heat flow that can be produced
if the fluid is injected into the rock as a heat mining medium.
e. Rock Type

Geothermal resources are found mostly in three types of rocks namely igneous, metamorphic,
and sedimentary. Convergent and divergent environments, along with mantle plume and
extensional features, are probably dominated by igneous rocks. Transform environments have
both igneous and sedimentary strata. Basins are dominated by sedimentary rocks, while igneous
rocks are the target for the basement complex (Table 2).
Rajaobelison, et.al, [59] used methods described in this section to characterize the geological
environment by highlighting the tectonic and structural settings at the regional scale in each area.
The analysis of tectonic data was based on up-to-date geological maps and the available
literature for the tectono-metamorphic domains of Madagascar. This information was then used
to infer the dominant geologic control on fluid flow or the heat migration pathway, geothermal
play type, and heat source. These elements were used to develop the catalogue of geothermal
systems in Madagascar, similar to the manner proposed by Erdlac [58], for classifying
geothermal energy sources in a geothermal power classification system and by Moeck [41] &
[60] & [61] for cataloguing geothermal plays.

4. Temperature and other thermodynamic properties


4.1 Classifications of geothermal resources by temperature
Temperature is a fundamental measure of the quality of geothermal resources and consequently
is the primary element of most classification systems. The temperature of the geothermal
resource can be easily measured by drilling an exploration well.
Figure 5, shows the main temperature classification proposals by different researchers and it is
apparent that there is no standard classification method used internationally. The temperature
boundaries are set at different temperatures thought to be significant in either a thermodynamic
or an economic utilization context.
Table 3 shows the classification of geothermal resources by temperature for Italy, New Zealand,
Iceland, and the USA proposed with different researchers, this classification method offers 3
classes of low, intermediate, and high for Italy, New Zealand, and Iceland, and meanwhile, it
offers 7 classes for the USA. The lack of consistency and conflicting boundaries for temperature
values is obvious from this classification method. A geothermal fluid of 130°C is considered to
be intermediate in Italy, New Zealand, and Iceland, while it is considered to be a very low
temperature resource in the USA. Once again, this fact proves the demand for a unified method
for the classification and confirms that temperature classification cannot be a solution alone.

1666
Figure 5: Classifications of geothermal resources by temperature [10].

Table 3: Classifications of geothermal resources by temperature done by various researchers

Researcher Country Temperature boundaries


3 classes: low (<90℃), intermediate (90-150℃), high
Muffler and Cataldi [62] Italy
(>150℃)
3 classes: low (<125℃), intermediate (125-225℃), high
Hochstein [63] New Zealand
(>225℃)
3 classes: low (<100℃), intermediate (100-200℃), high
Bendritter and Cormy [64] Iceland
(>200℃)
7 classes: non-electrical grade (<100°C), very low (100-
Sanyal [65] USA 150°C), low (150-190°C), moderate (190-230°C), high
(230-300°C), ultrahigh (>300°C), steam fields

4.1.1 Classification method by Sanyal [65]


Temperature classification method proposed by Sanyal [65] on request by the US Department of
Energy is further discussed in this section. In this classification method, geothermal resources are
classified into seven classes based on temperature: non-electrical grade (<100°C), very low
temperature (100°C to <150°C), low temperature (150°C to 190°C), moderate temperature
(190°C to<230°C), high temperature (230°C to<300°C), ultra-high temperature (>300°C), and
steam fields (approximately 240°C with steam as the only mobile phase).

1667
Table 4: A possible classification scheme for geothermal resources [65].

Depending on each temperature class, the phase of geothermal fluid, power generation method,
and development problems will differ. Table 4, shows the temperature and additional important
attributes of each class. The Classes of reservoir temperature below 230°C contain liquid water,
and over 230°C contain liquid water or two-phase fluid. The geothermal resources over 100℃
can be used for power generation (Figure 6). Low temperature geothermal resources are used for
direct uses such as heating and bathing as well as binary power generation. Flash and hybrid
power plants utilize steam from high temperature geothermal resources whereas steam fields are
utilized by direct steam power plants. Scaling problems, high NCG content, and silica deposition
tend to occur with high-temperature geothermal resources. Class 7 is located to the right of the
saturated vapor line. The other six classes are simply classified by temperature, so it doesn't
consider enthalpy and pressure.

Figure 6: Classification scheme on pressure-enthalpy-temperature diagram for water [65]

1668
4.2 Classification of geothermal resources by exergy
Temperature and enthalpy alone cannot represent fluid properties (For fluids with different
phases such as saturated water or saturated steam).
Using exergy for resource classification has the advantage of comparing according to the ability
to do work. Exergy is expressed as equal to the maximum work when the stream of substance is
brought from its initial state to the environmental state defined by P0 and T0, by physical
processes involving only thermal interaction with the environment [66].

𝐸 = 𝑚𝑖 [(ℎ𝑖 − ℎ0 ) − 𝑇0 (𝑆𝑖 − 𝑆0 )] (1)

m: mass flow late (𝑘𝑔�𝑠)

h: enthalpy (𝑘𝐽�𝑘𝑔)

s: entropy (𝑘𝐽�𝑘𝑔 ∙ 𝐾)

T: temperature (K)
The subscript i and 0 denote initial and environmental states, respectively.
The specific exergies of saturated water and steam are listed in table 5 for sink conditions of the
triple point (0.01℃), 10℃, and 20℃.

Bejan [67] pointed out that the minimization of lost work in the system would provide the most
efficient system. By using exergy analysis method, magnitudes and locations of exergy
destructions (irreversibilities) in the whole system are identified, while the potential for energy
efficiency improvements is introduced [68].
Energy, environmental and economic analysis of the four defined geothermal power plants has
been performed by Chamorro et al. [71] and it was concluded that the geological structure, the
depth of the reservoir, and other factors vary in a significant way from one location to another.
Exergy values are sensitive to sink conditions according to Eq. (1), and they depend on location,
season, and altitude. This variation proves that specific exergy cannot be a reliable parameter in
classifying geothermal resources. To achieve a reliable and stable parameter, independent of the
sink condition, Lee, [72] suggested normalizing exergy values by the maximum exergy of the
corresponding sink condition. The normalized exergy values, known as the specific exergy index
(SExI), are from 0 to 1.0 for saturated steam and water, but can theoretically exceed 1.0 for
abnormally high superheated steam [30].

1669
Table 5: Specific exergy and specific exergy index values under different sink conditions (water triple point,
10℃ and 20℃) [69] & [70].

The sink condition is no longer an important issue if SExI values are used. It is preferable to use
the triple point as a sink condition because saturated liquid enthalpy and entropy are defined as
zero at the triple point. Assuming a maximum value of 1194 kJ/kg for exergy and using eq. (1),
SExI can be formulated as:
(h−273.16s)
SExI = (2)
1194

Practically, it can be assumed that 100° is the minimum temperature of saturated steam that can
produce electricity, with 1 bar, enthalpy (h) of 2676 kJ/kg, and entropy of 7.361 kJ/kgK. SExI
can be calculated by Eq. (2) as 0.557. Therefore, 0.5 is the lower limit of SExI for high-quality
resources. In the same manner, assuming 100° saturated water with 1 bar, enthalpy (h) of 419.1
kJ/kg and entropy (s) of 1.307 kJ/kgK as a boundary for direct use application, SExI may be
calculated by Eq. (2) as 0.052. Hence, the upper limit of SExI for low-quality resources is 0.05.
Resources with SExI between 0.05 and 0.5 are evaluated as medium quality.
Bodvarsson and Eggers [73] applied the exergy concept to thermal water and concluded that
improvement in efficiency for each additional flash stage decreases rapidly with the increasing
number of stages in geothermal flash cycles. Brook et al. [74] also applied the concept of exergy
to geothermal systems greater than 150°C.

1670
4.2.1 Case study: Worldwide application
Figure 7, explains Lee’s SExI plot with applications on some important worldwide geothermal
fields [72]. As seen from the figure most geothermal areas are located in high and medium
exergy areas.

Figure 7: Lee’s SExI plot with some important worldwide geothermal fields [75] modified from Lee [72] &
[76].

4.2.2 Case study: Application in Turkey


A significant development was achieved in Turkey in geothermal electricity production and
direct uses (district heating, greenhouse heating, and thermal tourism) for the last decade. These
are mainly attributed to new Geothermal laws on regulations and the feed-in tariff. About 450
geothermal fields have been discovered in Turkey, geothermal electricity installed capacity
reached up 1282 MWe as of the end of 2018 [77]. A liquid carbon dioxide and dry ice
production factory is integrated into the Kizildere geothermal power plant. The existing plants
are in the following areas: Çanakkale-Tuzla 7 MWe, Aydin-Hidirbeyli 68 MWe, Aydin-Salavath
35 MWe, Aydin-Germencik 98 MWe, Aydin-Gümüsköy 7 MWe, Denizli-Kizildere 107 MWe,
Aydin-Pamukören 48 MWe, Manisa-Alasheir 24 MWe, Denizli-Gerali 3 MWe) [78] [79]. Figure
8 is an exergic classification map for turkey and most geothermal areas are located in high and
medium exergy area.
Hepbasli and Ozgener [80] presented an overview of geothermal energy utilization, historical
development, and opportunities in Turkey. Finally, a thermodynamic optimization was
performed for the Denizli-Kızıldere power plant using real data, and the most suitable working
fluid for the binary cycle were also investigated by Daǧdaş et al., [81].

1671
The results suggest that exergy should be utilized by engineers and scientists, as well as decision
and policymakers, involved in green energy and technologies in tandem with other objectives
and constraints. Exergy clearly identifies efficiency improvements and reductions in
thermodynamic losses attributable to green technologies. Thus, exergy has an important role to
play in increasing the utilization of green energy and technologies [82] & [83] & [84].

Figure 8: Examples of geothermal fields plotted on classification map of geothermal resources [85] & [86].

4.2.3 Case study: Application in Poland


The geothermal potential of the country is well known, with important development for direct
heat supply. As yet, there is no operational geothermal power plant in Poland, but a small binary
pilot plant at Lodz is under evaluation [87]. Figure 9 shows most geothermal areas in Poland are
located in medium exergy area.
4.2.4 Case study: Application in the Slovak Republic
Geothermal waters are widely used for recreational purposes, mostly in very popular aqua parks
in many places of Slovakia, with several others direct utilization. A small pilot plant is under
evaluation at Kosice, for a combined heat and electricity project [92] and [93]. The resources fall
under the medium exergy area as shown in figure 10.

1672
Figure 9: Classification of geothermal resources by SEI in Poland and selected global fields plotted on Mollier
diagram, [88],[89], [90], [91].

Figure 10: Prospective geothermal localities map of Slovakia on a Mollier's diagram with the perspective of
global geothermal fields [92].

1673
4.2.5 Case study: Application in Japan
Despite the large geothermal potential of the country, estimated at around 20 GWe, the present
total capacity of the geothermal power plant is still around 500 MWe, almost unchanged for
more than a decade. After the nuclear accident in March 2011, the government restarted an
incentive scheme for geothermal development and mitigation of constraints in national parks,
encouraging new geothermal exploration activities by private sectors as well as quick installation
of small binary systems. About 40 projects are either in the exploration or development stage.
The following fields are active: Akita (88 MWe), Fukushima (65 MWe), Hachijojima (3 MWe),
Hokkaido (25 MWe), Iwate (103 MWe), Kagoshima (60 MWe), Kumamoto (2 MWe), Miyagi
(15 MWe), Oita (155 MWe), and Tokamachi (2 MWe) [94] & [95] &[96].
Japan has a large area with a viable geothermal resource. The geothermal resource potential is
the third in the world [97]. From figure 11, it can be seen that most of the geothermal resources
are located in high and medium exergy area. From this, it is understood that the use of low
exergy area is not advanced. Plans are in place for the development of hot springs (Onsen) binary
units hence the development of resources under low exergy area [30] and [70].

Figure 11: Distribution of Japanese geothermal resources on SExI map, according to their specific entropy
and enthalpy [69] &[98].

Geothermal resources in Nigeria were classified using thermodynamic-based resource


classification [99] and the classification of geothermal resources in Indonesia by applying exergy
concept was done by [100].

1674
5. Other classification methods
5.1 Combinational Terminology for Geothermal Systems
Combinational terminology of geological and temperature-based classifications was proposed by
[101]. This method takes into consideration three parameters i.e geological setting, temperature,
and physical state of the geothermal reservoir. According to geological settings, in combinational
terminology, the class of geothermal resource’s name would be decided first based on the
geological environment, then the physical state of the reservoir which is liquid-dominated, two-
phase, or vapor-dominated, and finally, the class of the geothermal reservoir which is related to
its temperature. These encompass both the fundamentally geological nature of geothermal
resources and almost the practical technological and economic aspects of resource exploitation
while remaining understandable to the broad community of non-experts and experts [101].
5.1.1 Case study: Berlín Geothermal Field, El Salvador
The Berlín geothermal field is one of the geothermal fields in El Salvador. It is located 110 km
east of San Salvador, the capital city, in the District of Usulutan, near Berlín City. This
geothermal field is located on the northern flank of the Berlín-Tecapa volcanic complex, inside a
system of faults in the southern part of the east-west oriented Central American graben. Table 6
shows the classification of the Berlín Geothermal resource based on the combination
methodology and it can be classified as a volcanic liquid-dominated high temperature geothermal
system [101].
Table 6: Application of Combinational Method on Berlin Geothermal Resource.
Classification Class
Geology Volcanic geothermal system
The physical state of water Liquid-dominated geothermal system
Sanyal [65] High temperature geothermal system
USGS [102] High temperature geothermal system
Enthalpy [103] High enthalpy geothermal system
Combinational terminology Volcanic liquid dominated high temperature geothermal system

5.2 By Stored Heat


The heat in place approach was developed by Nathenson [104], White and Williams [105],
Muffler and Cataldi [106] & [107], and quickly became a well-established method for the
assessment of geothermal resources. The data is collected or calculated based on proven
correlations and then the thermal energy available in a volume of porous and permeable rock,
given the thickness, areal extent, porosity, average temperature, rock density and specific heat of
the rock in the reservoir, and physical properties of fluids are estimated. This method is the most
commonly used method for the assessment of green fields. Its accuracy depends on available
data and more data can lead to more reliable simulation. Stored heat calculations were modified
to account for the difference between hydrothermal and enhanced geothermal systems behavior
by incorporating natural heat input into the reservoir [108].
A case study was done by Agemar et al. [60] to assess geothermal resources in Germany, to
emphasize the renewability of geothermal energy, they proposed the reporting of geothermal

1675
capacities (per km2) instead of recoverable heat energy which depends very much on project
lifetime and other factors.
5.3 By Electric Power Generation Potential
The results from the stored heat estimation can be used to assess the electric power generation
potential of the identified geothermal occurrence. For electric power generation projects, the
potential is a function of the thermal energy stored in the reservoir, the thermal energy that can
be recovered at the wellhead, and the efficiency with which the latter can be converted into
electric power. The potential can be estimated from the stored heat through the application of a
recovery factor, an energy conversion factor, a power plant capacity factor, and power plant life.
Pezzuolo et al. [109] proposed a computer tool, called “ORC Plant Designer”, able to perform
the fluid screening and the design point analysis of different ORC plant configurations with
different heat sources. Kang [110] investigated the performance of an ORC with a two-stage
radial turbine.
Real-time energy convex optimization, via electrical storage, in buildings was performed by
Georgiou et al. [111]. A multidisciplinary approach for sustainable exploitation of medium to
low enthalpy sources was proposed by Franco and Vaccaro [112]. In the study by Astolfi et al.
[113] a techno-economic optimization of different ORC configurations operating with a number
of working fluids was performed, considering equipment cost correlations and a model to
estimate the turbine design and efficiency.
As a case study in the Republic of Croatia, the possibility of geothermal electricity was
investigated and ORC cycle was suggested to be suitable to utilize their geothermal potential
[114] & [115]. The traditional classification of working fluids, with merely three categories (wet,
dry, isentropic), is not sufficient to reliably predict or exclude the formation of liquid droplets in
the low-pressure stage of the turbine of an ORC power plant, therefore a novel classification of
pure working fluid selection was proposed by Gyorke et al. [116]. The study by Usman et.al
[117] compares the part-load operation of air-cooled and cooling tower based low-medium
temperature geothermal Organic Rankine Cycle (ORC) systems installed at different
geographical locations.

Figure 12: Geothermal power generation capacity from 2005 to 2015 [118].

1676
An investigation was done to determine the temperature classes application on geothermal
resources. These were achieved by investigating production wells used for geothermal power
generation or direct uses which were drilled in recent years by different countries. Figure 12,
shows trends in geothermal power generation capacity by major countries in 2005, 2010, and
2015.
The countries with the biggest generation capacity increase in this period i.e United States, New
Zealand, and Kenya were further investigated in detail and the results are shown in tables 7, 8,
and 9 respectively. In addition, the case of Japan was also further investigated within the same
period.
Table 7: Wells drilled for different uses in the USA from 2010 to 2014 [119] [120].

From table 7, drilled wells for geothermal uses between 2010 and 2014 in the USA are divided
as over 150°C, 100°C to 150°C, and below 100°C based on temperature. Production wells
ranging from 100°C to 150°C were the most numerous, and 67 wells were drilled for power
generation.
Table 8: Wells drilled for different uses in New Zealand from 2010 to 2014 [121].

It can be seen from table 8 that production wells of over 150°C are used for power generation
and combined utilization. There is no well for power generation in the production well below
150°C and twenty-six production wells were drilled in a temperature classification of over
150°C.
In Kenya, a large number of production wells were drilled between 2010 and 2014. Most of the
production wells are for power generation. Wells with temperature classification of over 150℃ is
the largest category, and more than 200 production wells were drilled as shown in table 9.

1677
Table 9: Wells drilled for different uses in Kenya from 2010 to 2014 [122].

Table 10 shows wells drilled for geothermal use between 2005 and 2008 in Japan. There was no
construction of a large power station during this period, but 13 production wells were over
150°C.

Table 10: Geothermal wells drilled for power generation in Japan from April 1, 2005, to December 31, 2008
[123].

5.4 By Direct Utilization


Geothermal energy utilization is commonly divided into two categories, i.e., electric production
and direct application. Direct use is typically associated with lower-temperature geothermal
resources, though some applications may require higher temperatures. In this method, the famous
Lindal diagram [124] and its cascade utilization method is being used in order to maximize the
optimum utilization of the resources. It is a similar scheme to the temperature classification
method, but with more focus on applications such as bathing and swimming, space heating,
agriculture and aquaculture, Industrial applications [125], and Ground Source Heat Pumps
(GSHP) [126].
In order to assess the technical and economic potential of shallow geothermal energy in
individual and district heating systems, a framework was proposed with a case study of Slovenia

1678
[127]. An open heat exchanger was proposed to increase the sustainability of geothermal
utilization in Icelandic hot spas [128].
To maximize the multiplication of effectiveness and capacity ratio to maximize heat transfer, a
new model was suggested for the determination of the heat capacity ratio that provides the
maximum heat transfer and the best type of heat exchanger [129].
Geothermal heat exchanger energy prediction based on times series and monitoring sensors
optimization was performed by Barugue et al., [130]. A proposed method by Georgiou [131] can
be applied for simultaneous minimization of a building’s import and export (surplus) grid energy
of geothermal district heating system.

6. New Geothermal Terms and Definitions, by the Geothermal Energy Association (GEA)
[132].
GEA issued a guideline requesting geothermal project developers to define the resource type
under one of the following categories: conventional hydrothermal (un-produced resource),
conventional hydrothermal (produced resource), conventional hydrothermal expansion,
geothermal energy, and hydrocarbon co-production, geopressured systems, and enhanced
geothermal systems.
GEA requests developers to indicate at what stage of development each separate geothermal
project falls under. Stages of development have been divided into four different phases labeled
Phase I, II, III, and IV. Each project must meet certain criteria to be considered and reported by
the GEA as being in a respective phase of development. The separate phases are specified below:
Phase I: Resource procurement and identification
Phase II: Resource Exploration and Confirmation
Phase III: Permitting and initial development
Phase IV: Resource Production and Power Plant Construction

Conclusions
On a global scale, geothermal energy is tied to the geologic environment and the large-scale
features specific to that environment. In defining geothermal prospects, resources, and reserves,
clear terms and definitions are required to provide reliable and comparable reserve estimation
analogous to the classifications schemes developed for petroleum resources. Various
classifications face various challenges in accessing, quantifying, and classification of geothermal
resources.
By stored heat- Classifying geothermal resources based on the heat in place only leads to large
figures that may be misunderstood by non-specialists, who may wrongly interpret them as
recoverable energy. On the other hand, when the process to be implemented to recover given
resources is still unknown or highly uncertain, the heat in place may represent the only reference
estimate.

1679
By electric power generation potential- Classification by electric power generation potential
requires reliable values for the recovery factor in the conversion of heat in place into power
potential. The use of term potential, applied to power generation that could be obtained from a
given geothermal occurrence, may generate confusion in the classification system.
By accessibility and discovery status- Impressive figures are usually stated for unconventional,
undiscovered, and currently unrecoverable geothermal resources worldwide. There is also the
risk of confusing resources with reserves, which increases the risk to the investor and decreases
confidence in geothermal development. There is therefore a need to apply a common framework
from in-situ potential to confirmed economic production.
Potential- Reliable values for the recovery factor are needed to convert theoretical potentials
into technical potentials which are hard to get. There is hardly any solid data about recovery
factors, not even for hydrothermal systems, let alone for petrothermal/EGS [133] & [134].
Therefore, the quality of geothermal fluid may be different even in the same class. Hard to
differentiate the potential referred to e.g theoretical, technical, economic, sustainable, or
developable.
Temperature- Temperature classification is a simple and easy to understand classification
method, but it does not represent the enthalpy and pressure of the geothermal fluid, so it isn't
enough to accurately evaluate the value of geothermal resources. Temperature cut-off gives
information on heat content regardless of other physical properties (e.g. permeability, porosity,
geochemistry, thermal capacity, and conductivity). Pressure, quality, and enthalpy are important
parameters for geothermal power generation. Classification by temperature can represent the
approximate usefulness of geothermal resources, but it isn't enough to properly evaluate the
value of geothermal resources, and other classifications need to be used to make an accurate
assessment.
Exergy- This method is considered more accurate as compared to the temperature and enthalpy
methods because of the advantage of comparison according to the ability to do work. However,
the main disadvantage is the difficulty in acceptance by the non-specialists (less accustomed to
thermodynamics terminology), and also dependent on the availability of wellhead pressure and
temperature data.
Geological setting- Effectively volcanic and fault-controlled extensional domain plays may
coexist in the same geologic system e.g the East African Rift System.
Currently, there is no universally and globally recognized standard level of confidence for
classifying and reporting geothermal resources. Inconsistent estimates which do not translate into
sustainable production may damage the perception of geothermal energy as an investment
option. The current classification systems have drawbacks and limitations on their own leaving
the door open for ambiguity and subjectivity.
A combinational method should be therefore be developed based on economic and social
viability, project status and feasibility, thermodynamic properties, and geological knowledge to
address these challenges. Table 11 summarizes the key classification methods discussed in this
paper and their advantages and disadvantages.

1680
References
1. Dickson, M.; Fenelli, M. What is Geothermal Energy? Ist. deGeoscienze eGeorisorse
2004, Pisa:Italy.
2. Li, J.; Zhang, Y. GIS-supported certainty factor (CF) models for assessment of geothermal
potential: A case study of Tengchong County, southwest China. 2017,
doi:10.1016/j.energy.2017.09.012.
3. Tü, N.; Lü, M.; Zen, S.; Leç, G. GIS based geothermal potential assessment: A case study
from Western Anatolia, Turkey. Energy 2009, 35, 246–261,
doi:10.1016/j.energy.2009.09.016.
4. Procesi, M.; Buttinelli, M.; Pignone, M. Geothermal favourability mapping by advanced
geospatial overlay analysis: Tuscany case study (Italy). 2015,
doi:10.1016/j.energy.2015.06.077.
5. Casasso, A.; Sethi, R. G.POT: A quantitative method for the assessment and mapping of
the shallow geothermal potential. 2016, doi:10.1016/j.energy.2016.03.091.
6. Kalogirou, S.A.; Florides, G.A.; Pouloupatis, P.D.; Panayides, I.; Joseph-Stylianou, J.;
Zomeni, Z. Artificial neural networks for the generation of geothermal maps of ground
temperature at various depths by considering land configuration.,
doi:10.1016/j.energy.2012.06.045.
7. Pan, C.; Chávez, O.; Romero, C.E.; Levy, E.K.; Aguilar Corona, A.; Rubio-Maya, C. Heat
mining assessment for geothermal reservoirs in Mexico using supercritical CO2 injection.
Energy 2016, 102, 148–160, doi:10.1016/j.energy.2016.02.072.
8. Falcone, G.; Liu, X.; Radido Okech, R.; Seyidov, F.; Teodoriu, C. Assessment of deep
geothermal energy exploitation methods: The need for novel single-well solutions. 2018,
doi:10.1016/j.energy.2018.06.144.
9. Huang, W.; Cao, W.; Jiang, F. A novel single-well geothermal system for hot dry rock
geothermal energy exploitation. 2018, doi:10.1016/j.energy.2018.08.055.
10. Williams, Colin F., Reed, M.J. and Anderson, A.F. Updating the Classification of
Geothermal Resources. Thirty-Sixth Work. Geotherm. Reserv. Eng. 2011.
11. MJ., R. Assessment of Low-Temperature Geothermal Resources of the United States -
1982.; 1983;
12. J.N. Schroeder, C.B Harto, R.M. Horner, and C.E.C. Geothermal Water Use: Life Cycle
Water Consumption, Water Resources Assessment, and Water Policy Framework,
ANL/EVS-14/2, Argonne National Laboratory, Environmental Science Division; 2014;
13. National Renewable Energy Laboratory (NREL). Available online:
https://www.nrel.gov/analysis/index.html.
14. International Geothermal Association (IGA). Available online: https://www.geothermal-

1681
energy.org/explore/our-databases/conference-paper-database/.
15. Geothermal Energy-Oregon Institute of Technology. Available online:
https://urbanecologycmu.wordpress.com/2016/11/01/geothermal-energy-oregon-institute-
of-technology/.
16. Anderson, D.N.; Lund, J.W. Direct Utilization of Geothermal Energy: A Technical
Handbook, Geothermal Resource Council Special Report No.7; Geothermal.; 1979; ISBN
0-934412-07-3.
17. Yuafrinaldi; Furqan, T.; Kartadjoemena, J.; Waren, R. Geothermal Resource
Classifications: Can We Talk the Same Language? In Proceedings of the 38th New
Zealand Geothermal Workshop; Auckland, New Zealand.
18. Cumming, W. Resource Capacity Estimation Using Lognormal Power Density from
Producing Fields and Area from Resource Conceptual Models; Advantages, Pitfalls and
Remedies. In Proceedings of the 41st Workshop on Geothermal Reservoir Engineering
Stanford University; Stanford, California, 2016.
19. Wilmarth, M.; Stimac, J. Power Density in Geothermal Fields. In Proceedings of the
World Geothermal Congress 2015; Melbourne, Australia, 2015.
20. Breede, K.; Dzebisashvili, K.; Liu, X.; Falcone, G. A systematic review of enhanced (or
engineered) geothermal systems: past, present and future. Geotherm. Energy 2013, 1, 4,
doi:10.1186/2195-9706-1-4.
21. Williams, C.F.; Reed, M.J.; Anderson, A.F. Updating the Classification of Geothermal
Resources. Proc. 36th Work. Geotherm. Reserv. Eng. 2011.
22. Saemundsson, K. Geothermal systems in global perspective. Short course IV on
exploration for geothermal resources, organized by UNU-GTP, KenGen and GDC, at
Lake Naivasha, Kenya, November 1-22.; 2009;
23. White, D. Geothermal energy. U.S. Geol. Surv. 1965, Circular ;, 519;17.
24. White, D.E.; Muffler, L.J.P.; Truesdell, A.H. Vapor-dominated hydrothermal systems
compared with hot-water systems. Econ. Geol. 1971, 66, 75–97,
doi:10.2113/gsecongeo.66.1.75.
25. Kruger, P.; Otte, C. 1973, p. 360.
26. Muffler, P.; Cataldi, R. Methods for regional assessment of geothermal resources.
Geothermics 1978, 7, 53–89, doi:10.1016/0375-6505(78)90002-0.
27. Williams, C.; Reed, M.; Mariner, R.; DeAngelo, J.; Galanis, S. Assessment ofmoderate-
and high-temperature geothermal resources of the United States. U.S. Geol. Surv. Fact
Sheet 2008, 3082,4.
28. Williams, C.; Reed, M.; Mariner, R. A review of methods applied by the U.S. Geological
Survey in the assessment of identified geothermal resources. U.S. Geol. Surv. Open-File

1682
Rep. 2008-1296, 27. 2008.
29. Armstead, H. Geothermal Energy. Its past, Present and Future Contribution to the Energy
Needs of Man (2nd). E. & F.N. Spon: London. 1983, 404.
30. Jalilinasrabady, S.; Itoi, R. Classification of geothermal energy resources in Japan
applying exergy concept. Int. J. Energy Res. 2013, 37, doi:10.1002/er.3002.
31. Falcone, G.; Antics, M.; Baria, R.; Bayrante, L.; Conti, P.; Grant, M.; Hogarth, R.;
Juiliusson, E.; Mijnlieff, H.; Nador, A.; et al. Application of the United Nations
Framework Classi cation for Resources (UNFC) to Geothermal Energy Resources -
Selected case studies; 2017; ISBN 9789211171365.
32. Minder, R.; Siddigi, G. Stakeholder Analysis on a National Level, Swiss Federal Office of
Energy SFOE.; 2013;
33. Rybach, L. The Future of Geothermal Energy” and Its Challenges. World Geotherm.
Congr. 2010, 1–4.
34. Rybach, L. Classification of geothermal resources by potential. Geotherm. Energy Sci.
2015, 3, 13–17, doi:10.5194/gtes-3-13-2015.
35. Chamorro, C.R.; García-Cuesta, J.L.; Mondéjar, M.E.; Pérez-Madrazo, A. Enhanced
geothermal systems in Europe: An estimation and comparison of the technical and
sustainable potentials. 2013, doi:10.1016/j.energy.2013.11.078.
36. Goldstein, B.A.; Hiriart, G.; Tester, J., B.; Bertani, R.; Bromley Great expectations for
geothermal energy to 2100. In Proceedings of the ,36th Workshop on Geothermal
Reservoir Engineering Stanford University, Stanford, California.
37. Beardsmore, G.R.; Rybach, L.; Blackwell, D.; Baron, C. A protocol for estimating and
mapping global EGS potential. Trans. - Geotherm. Resour. Counc. 2010, 34 1, 271–282.
38. Wees, J.-D. van; (TNO), T.B.; Calcagno, P.; (BRGM), C.D.; (Mannvitt), C.L.; (CNR),
A.M. A Methodology for Resource assessment and application to core countries Table of
Contents; 2013;
39. Oliver, J. Australian Code for Reporting Exploration Results, Geothermal Resources and
Geothermal Reserves Second Edition (2010). J. Chem. Inf. Model. 2013,
doi:10.1017/CBO9781107415324.004.
40. Resources, G.; Reserves, G.; Engineering, M.; Securities, J. the Canadian Geothermal
Code. 2010.
41. Moeck, I.S. Catalog of geothermal play types based on geologic controls. Renew. Sustain.
Energy Rev. 2014, 37, 867–882, doi:10.1016/j.rser.2014.05.032.
42. Taraszki, M.D. Book review: Gravitational Systems of Groundwater Flow: Theory,
Evaluation, Utilization, by József Tóth (Cambridge University Press, 2009). Hydrogeol. J.
2010, 18, 1971–1973, doi:10.1007/s10040-010-0649-2.

1683
43. CRAW, D.; KOONS, P.O.; ZEITLER, P.K.; KIDD, W.S.F. Fluid evolution and thermal
structure in the rapidly exhuming gneiss complex of Namche Barwa-Gyala Peri, eastern
Himalayan syntaxis. J. Metamorph. Geol. 2005, 051031032640003, doi:10.1111/j.1525-
1314.2005.00612.x.
44. Majorowicz, J.A.; Garven, G.; Jessop, A.; Jessop, C. Present Heat Flow Along a Profile
Across the Western Canada Sedimentary Basin: The Extent of Hydrodynamic Influence.
In; 1999; pp. 61–79.
45. Grasby, S.E.; Hutcheon, I. Controls on the distribution of thermal springs in the southern
Canadian Cordillera. Can. J. Earth Sci. 2001, 38, 427–440, doi:10.1139/e00-091.
46. Nukman, M.; Moeck, I. Structural controls on a geothermal system in the Tarutung Basin,
north central Sumatra. J. Asian Earth Sci. 2013, 74, 86–96,
doi:10.1016/j.jseaes.2013.06.012.
47. Deon, F., Moeck, I., Scheytt, T., Jaya, M.S. Preliminary assessment of the geothermal
system of the Tiris volcanic area, East Java, Indonesia,. In Proceedings of the Geophysical
Research Abstracts Vol. 14, EGU2012-9364, 2012; General Assembly European
Geosciences Union: Vienna, Austria, 2012.
48. Frisch Wolfgang, M.M. Plattentektonik; WBG (Wiss. Buchgesell.): Darmstadt, 2013;
ISBN 978-3-534-26246-5.
49. Moeck, I.; Schandelmeier, H.; Holl, H.-G. The stress regime in a Rotliegend reservoir of
the Northeast German Basin. Int. J. Earth Sci. 2009, 98, 1643–1654, doi:10.1007/s00531-
008-0316-1.
50. Bertini, G.; Casini, M.; Gianelli, G.; Pandeli, E. Geological structure of a long-living
geothermal system, Larderello, Italy. Terra Nov. 2006, 18, 163–169, doi:10.1111/j.1365-
3121.2006.00676.x.
51. Blackwell, D.D.; Waibel, A.F.; Richards, M. Why basin and range systems are hard to
find: The moral of the story is they get smaller with depth! Trans. - Geotherm. Resour.
Counc. 2012, 36 2, 1321–1326.
52. Moeck, I.S.; Dussel, M.; Weber, J.; Schintgen, T.; Wolfgramm, M. Geothermal play
typing in Germany, case study Molasse Basin: A modern concept to categorise geothermal
resources related to crustal permeability. Geol. en Mijnbouw/Netherlands J. Geosci. 2020,
2019, doi:10.1017/njg.2019.12.
53. Cacace, M.; Blocher, G.; Watanabe, N.; Moeck, I.; Borsing, N.; Kolditz, O. Modelling of
fractured carbonate reservoirs: Outline of a novel technique via a case study from the
Molasse Basin, southern Bavaria (Germany). Env. Earth Sci E pub 2013, 21, 18.
54. Sigrudsson, O. The Reykjanes Sea water geothermal system – Its exploitation under
regulatory constraints. In Proceedings of the WGC 2010; Bali, Indonesia; p. 1148,7p.
55. Bignall, G.; Rae, A.; Rosenberg, M. Rationale for targeting fault versus formation- hosted
permeability in high-temperature geothermal systems of the Taupo volcanic zone,

1684
NewZealand. In Proceedings of the WGC 2010; Bali, Indonesia; p. 1148,7p.
56. Bertani, R.; Bertini, G.; Cappetti, G.; Fiordelisi, A.; Marocco, B. An update of the
larderello-travale/radicondoli deep geothermal system. In Proceedings of the WGC 2005;
Antalya/Turkey; p. 0936,6p.
57. Harvey, C.; Harvey, M. The prospectivity of hot spot volcanic islands for geothermal
exploration. In Proceedings of the WGC 2010; Bali, Indonesia; p. 1182,6p.
58. Jr, R.J.E. ABSTRACT A Geologic Systems-Based Classification For Geothermal Energy,
#90128 (2011). 2011.
59. Rajaobelison, M.; Raymond, J.; Malo, M.; Dezayes, C. Classification of geothermal
systems in Madagascar. Geotherm. Energy 2020, 8, 22, doi:10.1186/s40517-020-00176-7.
60. Agemar, T.; Weber, J.; Moeck, I.S. Assessment and public reporting of geothermal
resources in Germany: Review and outlook. Energies 2018, 11, doi:10.3390/en11020332.
61. Moeck, I.; Beardsmore, G.C. Cataloging worldwide developed geothermal systems by
geothermal play type. In Proceedings of the WGC2015; Melbourne, Australia.
62. Muffler, P.; Cataldi, R. Methods for regional assessment of geothermal resources.
Geothermics 1978, 7, 53–89, doi:10.1016/0375-6505(78)90002-0.
63. Hochstein, M. Geothermics,.Vol. 17,. Geothermics 1988, 17, 15–49.
64. Bendritter, Y.; Cormy, G. Possible approach to geothermal research and relative costs, in
Dickson, M.H., and Fanelli, M., eds., Small Geothermal Resources: A Guide to
Development and Utilization; New York, 1990;
65. Sanyal, S.K. Classification of Geothermal Systems-a Possible Scheme. Thirtieth Work.
Geotherm. Reserv. Eng. 2005, 1–8.
66. Kotas, T.J. The Exergy Method of Thermal Plant Analysis; Krieger Publishing Co. Ltd:
Florida, USA, 1995; ISBN 9781483100364.
67. Bejan, A. Entropy Generation through Heat and Fluid Flow, John Wiley and Sons Inc.,
New York.; 1982;
68. Kuzgunkaya, E.; Hepbasli, A. Exergetic performance assessment of a ground-source heat
pump drying system. Int. J. Energ. Res. 2007, 31, 760–777.
69. Jalilinasrabady, S.; Itoi, R. Classification of geothermal energy resources in Japan
applying exergy concept. Int. J. Energy Res. 2013, doi:10.1002/er.3002.
70. Jalilinasrabady, S.; Itoi, R.; Valdimarsson, P.; Saevarsdottir, G.; Fujii, H. Flash cycle
optimization of Sabalan geothermal power plant employing exergy concept. Geothermics
2012, 43, doi:10.1016/j.geothermics.2012.02.003.
71. Chamorro, C.R.; Mondéjar, M.E.; Ramos, R.; Segovia, J.J.; Martín, M.C.; Villamañán,
M.A. World geothermal power production status: Energy, environmental and economic

1685
study of high enthalpy technologies. Energy 2012, 42, 10–18,
doi:10.1016/j.energy.2011.06.005.
72. Lee, K.C. Classification of geothermal resources by exergy. Geothermics 2001, 30, 431–
442, doi:10.1016/S0375-6505(00)00056-0.
73. Bodvarsson, G.; Eggers, D.E. The exergy of thermal water. Geothermics 1972, 1, 93–95,
doi:10.1016/0375-6505(72)90033-8.
74. Brook, C.; Mariner, R.; Makey, D.; Swanson, J.; Guffanti, M.; Muffler, L. Hydrothermal
convection systems with reservoir temperature. In Assessment of Geothermal Resources
of the United States-1978, Library of Congress Card No:79-600006, Muffler LJP (ed.) US
Geological Survey Circular 790, Arlington, VA, 18–85. 1979.
75. Hipólita, R.; Jordi, T.; Gilles, L.; Enrique, T.; Germán, R., A.; Héctor, P. New SExI tools
to evaluate the evolution and anthropic disturbance in geothermal fields: The case of Los
Azufres geothermal field, México. Revista Mexicana de Ciencias Geológicas. Rev. Mex.
Ciencias Geológicas 2010, 27, 520–529.
76. Lee, K.C. Classification of Geothermal Resources-An Engnineering Approach. In
Proceedings of the PROCEEDINGS, Twenty-First Workshop on Geothermal Reservoir
Engineering Stanford University, Stanford, California, January 22-24,; 1996.
77. TEİAŞ 2019. https://www.teias.gov.tr/sites/default/files/2019-
01/kurulu_guc_aralik_2018.pdf.
78. Mertoglu, O.; Simsek, S.; Basarir, N. Geothermal Country Update Report of Turkey
(2010-2015). In Proceedings of the WGC2015.
79. Mertogl, O.; Simsek, S.; Nilgun, B.; Paksoy, H. Geothermal Energy Use, Country Update
for Turkey. In Proceedings of the European Geothermal Congress 2019; Den Haag, The
Netherlands.
80. Hepbasli, A.; Ozgener, L. Development of geothermal energy utilization in Turkey: A
review. Renew. Sustain. Energy Rev. 2004, 8, 433–460, doi:10.1016/j.rser.2003.12.004.
81. Daǧdaş, A.; Öztürk, R.; Bekdemir, Ş. Thermodynamic evaluation of Denizli Kizildere
geothermal power plant and its performance improvement. Energy Convers. Manag. 2005,
46, 245–256, doi:10.1016/j.enconman.2004.02.021.
82. Kuzgunkaya, E.H.; Hepbasli, A. Exergetic evaluation of drying of laurel leaves in a
vertical ground-source heat pump drying cabinet. Int. J. Energy Res. 2007, 31, 245–258,
doi:10.1002/er.1245.
83. Rosen, M.A.; Dincer, I.; Kanoglu, M. Role of exergy in increasing efficiency and
sustainability and reducing environmental impact. Energy Policy 2008, 36, 128–137,
doi:10.1016/j.enpol.2007.09.006.
84. Ozgur, M.A. Review of Turkey’s renewable energy potential. Renew. Energy 2008, 33,
2345–2356, doi:10.1016/j.renene.2008.02.003.

1686
85. Kuzgunkaya, E.H. Evaluation of Turkey ’ s Geothermal Energy Resources in terms of
Exergy Analysis. World Geotherm. Congr. 2015 2015, 8, 19–25.
86. Kuzgunkaya, E.H. Evaluation of Geothermal Energy Resources in Terms of Exergy
Analysis. Karaelmas Sci. Eng. J. 2018, 8, 63–72.
87. Kępińska, B. Geothermal Energy Country Update Report from Poland, 2010–2014. In
Proceedings of the WGC2015; Melbourne, Australia.
88. Barbacki, A. Classification of geothermal resources in Poland by exergy analysis -
Comparative study. Renew. Sustain. Energy Rev. 2012, 16, 123–128,
doi:10.1016/j.rser.2011.07.141.
89. N, B.; Bidini G. Larderello-Farinello-Valle Secolo geothermal area: exergy analysis of the
transportation network and of the electric power plants. Geothermics 1996, 1, 3–16.
90. Etemoglu, A.; Can, M. Classification of geothermal resources in Turkey by exergy
analysis. RenewableandSustainableEnergyReview 2007, 11, 1596–1606.
91. Reykjavik Energy. Nesjavellir power plant( brochure); 2006;
92. Fričovský, B.; Černák, R.; Marcin, D.; Benková, K.; Remšík, A.; Fendek, M. Engineering
Approach in Classification of Geothermal Resources of the Slovak Republic ( Western
Carpathians ). PROCEEDINGS, 41st Work. Geotherm. Reserv. Eng. 22-24, 2016 2016, 1–
10.
93. Fendek, M.; Fendekova, M. Country Update of the Slovak Republic. Proc. World
Geotherm. Congr. 2015, 25–29.
94. Yasukawa, K.; Sasada, M. Country Update of Japan: Renewed Opportunities. In
Proceedings of the WGC 2005; Melbourne, Australia.
95. Jalilinasrabady, S.; Itoi, R.; Valdimarsson, P.; Fujii, H.; Tanaka, T. Energy and exergy
analysis of Takigami Geothermal Power Plant. J. Geotherm. Res. Soc. Japan 2011.
96. Jalilinasrabady, S.; Itoi, R. Flash cycle and binary geothermal power plant optimization. In
Proceedings of the Transactions - Geothermal Resources Council; 2012; Vol. 36 2.
97. Jalilinasrabady, S.; Itoi, R.; Valdimarsson, P.; Saevarsdottir, G.; Fujii, H. Flash cycle
optimization of Sabalan geothermal power plant employing exergy concept. Geothermics
2012, doi:10.1016/j.geothermics.2012.02.003.
98. Jalilinasrabady, S.; Itoi, R.; Gotoh, H.; Kamenosono, H. Energy and exergy analysis of
Takigami Geothermal Power Plant, Oita, Japan. In Proceedings of the Transactions -
Geothermal Resources Council; 2010; Vol. 34 2, pp. 966–971.
99. Amoo, O.M.A. Thermodynamic based resource classification of renewable geothermal
energy in Nigeria. J. Renew. Sustain. Energy 2014, 6, 033129, doi:10.1063/1.4881687.
100. Mohammadzadeh Bina, S.; Jalilinasrabady, S.; Fujii, H.; Pambudi, N.A. Classification of
geothermal resources in Indonesia by applying exergy concept. Renew. Sustain. Energy

1687
Rev. 2018, doi:10.1016/j.rser.2018.05.018.
101. Rezaie, M.; Aghajani, H. A New Combinational Terminology for Geothermal Systems.
Int. J. Geosci. 2013, 04, 43–48, doi:10.4236/ijg.2013.41005.
102. Nathenson Marianne, M.G. Compilation of geothermal-gradient data in the conterminous
United States; 1987; Vol. 87–592;.
103. Axelsson, G.; Gunnlaugsson, E. Geothermal Utilization, Management and Monitoring, in
Long-Term Monitoring of High and Low-Enthalpy Fields under Exploitation. In
Proceedings of the WGC 2000; 2000; pp. 3–10.
104. Nathenson, M. Physical factors determining the fraction of stored energy recoverable
from hydrothermal convection systems and conduction-dominated areas; 1975;
105. D. E. White and D. l. Williams, E. Assessment of Geothermal Resources of the United
States-1975; 1975;
106. Muffler, P.; Cataldi, R. METHODS FOR REGIONAL ASSESSMENT RESOURCES OF
GEOTHERMAL The critical dependence of modern society on minerals and fuels has
fostered an increasing awareness of the need to estimate not only the quantities that could
be produced under present economic cond. Geothermics 1978, 7, 53–89.
107. Muffler, L.J.P. Assessment of geothermal resources of the United States; 1978; Vol.
Geothermal;
108. Zarrouk, S.; Simiyu, F. A REVIEW OF GEOTHERMAL RESOURCE ESTIMATION
METHODOLOGY. In Proceedings of the 35th New Zealand Geothermal Workshop;
Rotorua, New Zealand, 2013.
109. Pezzuolo, A.; Benato, A.; Stoppato, A.; Mirandola, A. The ORC-PD: A versatile tool for
fluid selection and Organic Rankine Cycle unit design. 2016,
doi:10.1016/j.energy.2016.02.128.
110. Kang, S.H. Design and preliminary tests of ORC (organic Rankine cycle) with two-stage
radial turbine. 2015, doi:10.1016/j.energy.2015.09.040.
111. Georgiou, G.S.; Christodoulides, P.; Kalogirou, S.A. Real-time energy convex
optimization, via electrical storage, in buildings e A review. 2019,
doi:10.1016/j.renene.2019.03.003.
112. Franco, A.; Vaccaro, M. An integrated “Reservoir-Plant” strategy for a sustainable and
efficient use of geothermal resources. 2011, doi:10.1016/j.energy.2011.11.029.
113. Astolfi, M.; Romano, M.C.; Bombarda, P.; Macchi, E. Binary ORC (Organic Rankine
Cycles) power plants for the exploitation of mediumelow temperature geothermal sources
e Part B: Techno-economic optimization. 2014, doi:10.1016/j.energy.2013.11.057.
114. Guzovi, Z.; Lon Car, D.; Ferdelji, N. Possibilities of electricity generation in the Republic
of Croatia by means of geothermal energy., doi:10.1016/j.energy.2010.04.036.

1688
115. Ra Skovi C, P.; Guzovi, Z.; Cvetkovi, S. Performance analysis of electricity generation by
the medium temperature geothermal resources: Velika Ciglena case study. 2013,
doi:10.1016/j.energy.2013.03.009.
116. Abor Gy, G.; Orke, €; Deiters, U.K.; Groniewsky, A.; Lassu, I.; Imre, A.R. Novel
classification of pure working fluids for Organic Rankine Cycle. 2017,
doi:10.1016/j.energy.2017.12.135.
117. Usman, M.; Imran, M.; Yang, Y.; Lee, D.H.; Park, B.-S. Thermo-economic comparison of
air-cooled and cooling tower based Organic Rankine Cycle (ORC) with R245fa and
R1233zde as candidate working fluids for different geographical climate conditions. 2017,
doi:10.1016/j.energy.2017.01.134.
118. JOGMEC Japan Oil, Gas and Metals National Corporation (JOGMEC).
119. Boyd, T.L.; Sifford, A.; Lund, J.W. The United States of America country update 2015.
Trans. - Geotherm. Resour. Counc. 2015, 39, 65–74.
120. Tonya, L.; Boyd, A.S.; W.L., J. The United States of America Country Update 2015. In
Proceedings of the WGC2015; Melbourne, Australia.
121. Carey, B.; Dunstall, M.; Mcclintock, S.; White, B.; Bignall, G.; Luketina, K.; Zarrouk, S.;
Seward, A. 2015 New Zealand Country Update. World Geotherm. Congr. 2015, 19–25.
122. Omenda, P.; Simiyu, S. Country Update Report for Kenya 2010-2014. In Proceedings of
the WGC2015.
123. Sugino, H.; Akeno, T. 2010 Country Update for Japan Generation of Geothermal Energy.
Proc. World Geotherm. Congr. 2010, 25–29.
124. Lindal, B. Industrial and other applications of geothermal energy. geothermal energy. In
Earth Science, v. 12, Armstead HCH (ed.), UNESCO: Paris; 135–148. 1973.
125. Lund, J.W.; Rangel, M.A. Pilot Fruit Drier for the Los Azufres Geothermal Field, Mexico,
pp. 2335-2338. In Proceedings of the Proc. of the World Geothermal Congress; 1995.
126. Lund, J.W. Direct heat utilization of geothermal resources. Renew. Energy 1997, 10, 403–
408, doi:10.1016/0960-1481(96)00097-3.
127. Sper Stegnar, G.; Stani Ci C A, D.; Cesen, M.; Ci Zman, J.; Pestotnik, S.; Prestor, J.;
Urban Ci C A, A.; Mer Se, S. A framework for assessing the technical and economic
potential of shallow geothermal energy in individual and district heating systems: A case
study of Slovenia. 2019, doi:10.1016/j.energy.2019.05.121.
128. Jalilinasrabady, S.; Palsson, H.; Saevarsdottir, G.; Itoi, R.; Valdimarsson, P. Experimental
and CFD simulation of heat efficiency improvement ingeothermal spas. Energy 2013, 56,
doi:10.1016/j.energy.2013.04.057.
129. gra, O.A.; Hüseyin Erdem, H.; Demir, H.; Atayılmaz, O.; Teke, I. Heat capacity ratio and
the best type of heat exchanger for geothermal water providing maximum heat transfer.

1689
2015, doi:10.1016/j.energy.2015.06.107.
130. Baruque, B.; Porras, S.; Jove, E.; Luis Calvo-Rolle, J. Geothermal heat exchanger energy
prediction based on time series and monitoring sensors optimization. 2019,
doi:10.1016/j.energy.2018.12.207.
131. Georgiou, G.S.; Christodoulides, P.; Kalogirou, S.A. Optimizing the energy storage
schedule of a battery in a PV grid-connected nZEB using linear programming. Energy
2020, 208, doi:10.1016/j.energy.2020.118177.
132. GEA New Geothermal Terms and Definitions. 2010.
133. Rybach, L. Geothermal Power Growth 1995–2013—A Comparison with Other
Renewables. Energies 2014, 7, 4802–4812, doi:10.3390/en7084802.
134. Bayer, P.; Rybach, L.; Blum, P.; Brauchler, R. Review on life cycle environmental effects
of geothermal power generation. Renew. Sustain. Energy Rev. 2013, 26, 446–463,
doi:10.1016/j.rser.2013.05.039.

1690
Table 11: Key classification methods discussed in this paper and their pros and cons.

Classification Pros Cons


Method
Temperature The range of temperature for geothermal fluid is used to classify the resources based on their potential to do work.
Simple and easy to understand. Different boundaries for classifying resources.
Can represent the approximate usefulness of geothermal resources There is no certain agreed temperature range for each category.
It could be misleading to conclude the nature of the fluids based on only
temperature or enthalpy.
Geological The initial data will be used to create geological models, and these models will provide a criteria to classify geothermal system based on likelihood of
discovery.
Can be developed parallel to conceptual model development and they can complement each other.
Based on heat transfer regimes and geological settings, the geothermal resources could be classified under two categories of conductive and convective
Systems.
Gives an idea of prospects for development scenarios. Requires comprehensive geological knowledge.
It is possible to perform from the planning phase of the development and Outcomes are not clear enough for development scenarios.
can be upgraded as the project progresses.
Stored Heat Mainly based on the reservoir’s temperature and enthalpy.
The reservoir size and temperature distribution have a very significant role in the calculation of the resource potential.
The stored thermal energy in a rock with estimated permeability and porosity are being estimated. Input parameters are the dimensions of the assumed
reservoir, temperature, rock properties (such as density, specific heat) and fluid properties.
Simple and widely used. Uncertainty in required input parameters increases the inaccuracy of
Well-recognized geothermal resources assessment method. the stored heat estimation.
The data is collected or calculated based on proven correlations. Doesn’t clearly address the sustainability of the resource.
Results might be misleading.
Accessibility Geothermal resources are categorized as resource and reserve. Geothermal resource, include all the thermal energy beneath a specific area judged from the
and discovery measured temperature. Geothermal reserve, include those resources that are feasible to recover using current technology.

Can give a general idea of the resources. Not enough to convince investors unless there is very obvious evidence.
Potential The resource is categorized into theoretical, technical, economic, sustainable, and developable.
Feasibility of the geothermal potential is in reverse proportion to its size.
Simple with sensible numbers in technical and economics. Needs plenty of data with uncertain utilization scenarios due to the wide
Utilization proposals in forms of potential could be prepared. range of parameters.
Exergy Using exergy for resource classification has the advantage of comparing according to the ability to do work.
Very accurate and sensible results. It is very difficult to obtain the required information for exergy classification
Too technical. at the early stages of a project.
Points out exergy losses and destructions.
Combinational Takes into consideration three parameters i.e geological setting, temperature, and physical state of the geothermal reservoir.
Terminology Easy to understand for technical and non-technical counterparts. Requires more input data and that increases the risk of uncertainty.
Utilization The geothermal resources utilization method can be decided based on conclusion from the stored heat estimation.

It can deliver numbers and sensible results for developers at the early Not an independent method, it relies on other methods such as stored heat.
stages of development. Results are not very accurate and their reliability might improve with
A development scenario can be proposed based on the estimated potential. additional data obtained from advancement of the development.

1691

View publication stats

You might also like