You are on page 1of 144

POTENCIAL DE RECUPERACIÓN DE CALOR RESIDUAL

DE PROCESOS INDUSTRIALES EN ALEMANIA CON TEC-


NOLOGÍA DE CICLO ORGÁNICO DE RANKINE

Laura García Martínez

Máster Habilitante en Ingeniería Industrial

Especialidad Técnicas Energéticas

Technische Universität München ETSII-UPM

Master Thesis Trabajo de Fin de Máster

October 2019 Febrero 2020

Supervisor: Tutor:

M. Sc. Roberto Pili Rubén Abbas Cámara


Abstract I

Abstract
Direct CO2 emissions from industry accounted for 24% of global emissions in 2017, in-
cluding those from energy consumption and processing [1]. Waste heat recovery with or-
ganic Rankine cycle (ORC) power systems is an increasingly attractive option for a less
intensive energy consumption of industrial processes. ORC power systems can convert
available waste heat from these processes and produce electricity or electricity and district
heat, which can, for example, be used in the same plant reducing its demand.
The high waste heat potential from the industrial sector in Germany [2] has not been fully
exploited yet [3]. This, together with the acknowledged suitability and good performance of
ORC systems makes it interesting to estimate the theoretical, technical and economic po-
tential of this technology recovering heat from energy-intensive industry in Germany. This
is the goal of this thesis.
There is a lack of studies on the economic potential of waste heat recovery with ORC in
the main energy-intensive industries in Germany. Only some estimations on the technical
potential are available, but most of them are derived from studies in other countries. The
aim of this thesis is, therefore, to provide the researchers and industry with more knowledge
about the potential of the WHR technique with ORC power systems and to support its de-
velopment.
The most suitable industries for the installation of ORC systems are steel, cement and
glass [4]. Regarding the steel industry, the waste heat from the electric arc furnaces (EAF),
the basic oxygen furnaces (BOF) and the reheating furnaces (RHF) is considered. For the
cement industry, the heat in the off-gas from preheaters and hot air from clinker coolers at
cement dry kilns is regarded. For the glass industry, the melting furnaces for float and con-
tainer glass are included.
To estimate the recovery potential, a combination of the bottom-up and top-down ap-
proaches is used. The bottom-up method is used to gather data on the capacity of each
plant and to assess the specific waste heat available from each process susceptible of
waste heat recovery with an ORC unit. The top-down method is applied for estimating the
total waste heat released by each plant, where a common specific waste heat value (which
is different depending on the process) is multiplied by the production of each plant.
The uncertainty and high variability of the specific waste heat values for the same indus-
trial process led to considering a worst-, average- and best-case scenario. Also, for the
estimation of the technical potential, the ORC availability needs to be considered, as well
as the ORC unit efficiency. To estimate this efficiency, Turboden WHR-ORC units are con-
sidered because they are the global leading company in waste heat recovery applications
from industrial processes in both number of installations and capacity. The ORC availability
is considered to be 95% and the ORC efficiency varies in the approximate range of 15-21%.
For the analysis, a conservative case with an efficiency of 15% and another with 19% are
studied.
The economic potential is the accumulated electricity that is produced by the ORCs hav-
ing a lower Levelized Cost of Electricity (LCOE) than the electricity price. The range of elec-
tricity prices in 2018 for energy-intensive industries in Germany is in the range of 5.10-17.00
ct/kWh, with average price of 8.84 ct/kWh [5].
II

For the LCOE calculation, the ORC units’ specific investment costs need to be known.
To estimate these costs, a correlation by Turboden which depends on the unit size is used.
Also, assumptions regarding the O&M costs are made. The LCOE depends also on the
number of amortization years and the interest rate. As an initial assumption, an amortization
time of 10 years and an interest rate of 4% are considered for the economic analysis. Then,
a sensitivity analysis is made where these parameters are varied in order to evaluate their
influence on the economic potential.
German industry releases more than 200 TWh in waste heat every year, which is more
than the total annual energy consumption of Denmark [6]. The large number of steel, ce-
ment and glass factories in Germany produce a significant amount of high-temperature pro-
cess heat. In addition, these three energy-intensive industries are the ones showing the
most developed state-of-the-art on waste heat recovery with ORC in the world, with room
for a significative increase in the case of Germany, as shown in this work.
The optimistic results obtained from this work, together with the incentives and support
programs from the European Union and the German Energy Efficiency Fund (EEF) [7],
make the ORC technology a very attractive option for industrial waste heat recovery. The
economic potential grows with the amortization time and decreases with higher interest
rates. The economic potential results show promising for all the sectors studied (steel, ce-
ment and glass production) for amortization times between 10 and 5 years (except for glass,
which is longer) and interest rates smaller than 15% (except for glass, which is lower). The
results are especially optimistic for BOF, EAF and cement production. The installation of
ORC units for waste heat recovery in the plants where it is economically feasible could save
up to 149.22 M€/a and avoid 1.3873 Mt of CO 2 emissions per year, which corresponds to
2.13% of the GHG emissions from the industry in Germany in 2018.

Key Words: Waste heat recovery, Organic Rankine cycle, energy-intensive industries, Ger-
many
Table of Contents III

Table of Contents
1 Introduction ............................................................................................ 1
1.1 Motivation........................................................................................................ 1

1.2 Task ................................................................................................................ 2

2 Theoretical background ......................................................................... 3


2.1 Energy in the industrial sector ......................................................................... 3

2.1.1 Impact of the industrial sector growth worldwide ..................................... 3

2.1.2 Energy efficiency as best alternative for CO 2 emissions reduction .......... 4

2.1.3 Industrial waste heat recovery as a solution to energy efficiency ............. 4

2.1.4 Industrial waste heat potential in Germany .............................................. 5

2.2 Technology for waste heat recovery: state-of-the-art ....................................... 7

2.2.1 Overview ................................................................................................. 7

2.2.2 Electricity generation ............................................................................... 8

2.2.3 Waste heat recovery system with ORC and intermediate thermal circuit . 11

2.2.4 Advantages of an ORC waste heat recovery system ............................... 12

2.3 ORC power systems: state-of-the-art .............................................................. 13

2.3.1 Overview ................................................................................................. 13

2.3.2 ORC installed capacity worldwide ........................................................... 14

2.3.3 European projects for waste heat recovery with ORC systems in energy-
intensive industries.................................................................................. 14

2.4 Waste heat recovery potential ......................................................................... 16

2.4.1 Definitions of waste heat potential ........................................................... 16

2.4.2 Types of waste heat recovery potential estimation methods .................... 16

2.4.3 Waste heat recovery potential estimation for Germany approaches: state-
of-the-art ................................................................................................. 17

2.4.4 Notes on the technical potential .............................................................. 18

2.4.5 Notes on the economic potential ............................................................. 19

2.5 Steel industry .................................................................................................. 22

2.5.1 Description of the production process ..................................................... 22

2.5.2 Waste heat recovery with ORC from steel industry: state-of-the-art ........ 26
IV

2.6 Cement industry...............................................................................................28

2.6.1 Description of the production process ......................................................28

2.6.2 Waste heat recovery with ORC in the cement industry: state-of-the-art ...32

2.7 Glass industry ..................................................................................................33

2.7.1 Description of the production process ......................................................33

2.7.2 Waste heat recovery with ORC in the glass industry: state-of-the-art.......34

3 Methodology........................................................................................... 37
3.1 Overview of the method ...................................................................................37

3.2 Description of the method ................................................................................37

3.3 Definition of efficiencies for net electric power estimation ................................39

3.3.1 Utilization rate (𝜂𝑢) ..................................................................................39

3.3.2 Net electric efficiency of the ORC unit (𝜂𝑒𝑙, 𝑛𝑒𝑡, 𝑂𝑅𝐶) ..............................40

3.3.3 ORC unit availability ................................................................................40

3.4 Specific waste heat estimation for the different industrial processes ................41

3.4.1 Steel industry ...........................................................................................41

3.4.2 Cement industry.......................................................................................44

3.4.3 Glass industry ..........................................................................................45

3.5 Plants in Germany ...........................................................................................46

3.5.1 Steel ........................................................................................................46

3.5.2 Cement ....................................................................................................47

4 Results ................................................................................................... 49
4.1 Theoretical potential ........................................................................................49

4.1.1 Results ....................................................................................................49

4.1.2 Comparison to previous studies ...............................................................49

4.2 Technical potential ...........................................................................................50

4.2.1 Comparison of estimated ORC power to the power from ORCs in operation
................................................................................................................51

4.2.2 Comparison of the results to other studies ...............................................52

4.3 Economic potential ..........................................................................................52

4.4 Sensitivity analysis of the results .....................................................................56


Table of Contents V

4.5 Electricity savings............................................................................................ 58

4.6 CO2 emissions avoided ................................................................................... 59

4.7 Cost savings ................................................................................................... 60

5 Summary, conclusions and outlook ....................................................... 61


5.1 Summary and conclusions .............................................................................. 61

5.2 Future lines ..................................................................................................... 62

Bibliography
............................................................................................................................. lxii
i
6 Appendix ................................................................................................ i
A Resumen en español ...................................................................................... i

B WHR projects with ORC from ORMAT® and TURBODEN®


........................................................................................................................ xx
xviii

C Main steel production plants in Germany


........................................................................................................................ xlii
i

D Main cement production plants in Germany


........................................................................................................................ xlv
iii

E Main glass production plants in Germany ........................................................ l

F Economic potential graphs varying parameters ............................................... liii


VI
List of Figures VII

List of Figures
Figure 2.1. Industry direct CO2 emissions by IEA [1] .................................................................. 4

Figure 2.2. Final energy consumption in Germany in 2017 by sector [13]................................... 6

Figure 2.3. Locations of major EII with considerable volumes of excess heat. Source: The E-
PRTR database at EEA in Copenhagen [14] ................................................................. 6

Figure 2.4. Waste heat potential in each EU country per temperature level in all industries [2] ... 7

Figure 2.5. Technologies for industrial waste heat recovery/re-use [15] ..................................... 8

Figure 2.6. T–s diagram of water and various typical ORC fluids [20]....................................... 10

Figure 2.7. The Rankine cycle [22] .......................................................................................... 11

Figure 2.8. WHR system with ORC and intermediate thermal circuit scheme [23] .................... 11

Figure 2.9. ORC part load efficiency [22] ................................................................................. 12

Figure 2.10. Total installed capacity per application [4] ............................................................ 13

Figure 2.11. Shares of ORC installed capacity per heat recovery application [4] ...................... 14

Figure 2.12. Types of potential ................................................................................................ 16

Figure 2.13. Classification of methods [28] .............................................................................. 17

Several works tried to estimate the industrial waste heat potential in Germany (Figure 2.14) .
So far, almost only top-down analyses of the waste heat potential exist and they are
using key figures derived from studies in other countries [11] [2]. ................................ 17

Figure 2.15. Estimations of industrial waste heat in Germany in increasing order ..................... 18

Figure 2.16. SIC function [34] .................................................................................................. 20

Figure 2.17. Minimum and maximum electricity prices for EII (annual energy
consumption>10GWh) and average electricity price [5] ............................................... 22

Figure 2.18. Steel production routes: a) Oxygen steelmaking along BF and BOF converter, b)
Electric steelmaking in EAF [37].................................................................................. 23

Figure 2.19. General view of converter for BOF steelmaking process [40] ............................... 24

Figure 2.20. Illustration of EAF [44] ......................................................................................... 25

Figure 2.21. General scheme of the energy recovery system installed to Elbe-Stahlwerke
Feralpi [54] ................................................................................................................. 27

Figure 2.22. Schematic representation of the cement manufacturing process from quarry to
dispatch [58] ............................................................................................................... 29

Figure 2.23. Clinker production structure [61] .......................................................................... 30

Figure 2.24. Energy flow diagram of a cement plant clinker burning kiln [64] ............................ 32

Figure 2.25. Typical WHR with ORC installation in cement kilns [64] ....................................... 33
VIII

Figure 2.26. Schematic diagram of the production process of float glass [68]........................... 34

Figure 2.27. Heat recovery system with ORC technology in the glass industry typical layout [8]35

Figure 3.1. General scheme of the steps followed in this work ................................................. 37

Figure 3.2. Transferred heat depending on temperature difference .......................................... 39

Figure 3.3Turboden’s Standard HR (Heat Recovery) ORC units [72] ....................................... 40

Figure 3.4. Energy balance in an EAF [76] .............................................................................. 42

Figure 3.5. Tapping weight vs. energy consumption [76] ......................................................... 42

Figure 3.6. The Sankey Diagram of energy balance for a typical RHF in case study [57] ......... 43

Figure 3.7. Location of the main steel plants in Germany [88] .................................................. 46

Figure 3.8. Biggest steelmakers in Germany and their production routes [88] .......................... 47

Figure 3.9. Location of the German cement plants which are part of the VDZ [91] ................... 47

Figure 3.10. Biggest cement producers in Germany and capacity of plants with a kiln [91] ....... 48

Figure 4.1. Economic potential for steel production with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10
years: (a) BOF; (b) EAF and (c) RHF .......................................................................... 54

Figure 4.2. Economic potential for production with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 years:
(a) cement and (b) glass ............................................................................................. 55

Figure 4.3. Sensitivity analysis for: (a) BOF; (b) EAF and (c) RHF ........................................... 57

Figure 4.4. Sensitivity analysis for: (a) cement and (b) glass .................................................... 58

Figure 4.5. Development of the CO2 emission factor for the electricity mix in Germany in the
years 1990 to 2018 (in g/kWh) [96] ............................................................................. 59

Figure 4.6. GHG emission trends in Germany by sector 1990-2018. Data from: UBA 2019,
BMU 2019 [97] ........................................................................................................... 60

Figure 6.1. Economic potential for steel production with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟗 %, i=4%, n=10
years: (a) BOF; (b) EAF and (c) RHF ........................................................................... liii

Figure 6.2. Economic potential for production with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟗 %, i=4%, n=10 years:
(a) cement; (b) glass. ...................................................................................................liv

Figure 6.3. Economic potential for steel production with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=5
years: (a) BOF; (b) EAF and (c) RHF ........................................................................... lv

Figure 6.4. Economic potential for production with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=5 years: (a)
cement; (b) glass. ........................................................................................................lvi

Figure 6.5. Economic potential for steel production with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟓 %, i=6%, n=10
years: (a) BOF; (b) EAF and (c) RHF .......................................................................... lvii

Figure 6.6. Economic potential for production with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟓 %, i=6%, n=10 years:
(a) cement; (b) glass. ................................................................................................. lviii
List of Tables IX

List of Tables
Table 2.1. Waste heat temperature grades depending on their range [12] ................................. 5

Table 2.2. Options for Heat Recovery via Power Generation [17].............................................. 8

Table 3.1. Assumed utilization rate of the heat source for different processes.......................... 39

Table 3.2. Waste heat available from steel production, referred to 15°C [77], [78], [79], [80],
[81], [82], [83], [84], [85], [86], [87]............................................................................... 44

Table 3.3. Waste heat available from cement production, referred to 15°C [83], [78], [65] ........ 45

Table 3.4. Waste heat available from glass production, referred to 15°C [78], [8], [83] ............. 45

Table 4.1. Theoretical potential of waste heat available from the steel, cement and glass
industries in Germany at 15°C .................................................................................... 49

Table 4.2. Technical potential of waste heat potential available from the steel, cement and
glass industries in Germany with ORC availability at 95% ........................................... 50

Table 4.3. Cumulative capacity of technical potential from energy-intensive industry with ORC
availability at 95% ....................................................................................................... 51

Table 4.4. Estimated and actual ORC power for Feralpi and Lengfurt plants ............................ 52

Table 4.5. Economic potential of waste heat available from energy-intensive industry with
ORC availability at 95%, interest rate at 4% and 10 years of amortization ................... 55

Table 4.6. Cumulative capacity of economic potential from energy-intensive industry with ORC
availability at 95%, interest rate at 4% and 10 years.................................................... 56

Table 4.7. Saved electricity with ORC for steel, cement and glass with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟓 %,
i=4%, n=10 years........................................................................................................ 59

Table 4.8. CO2 emissions avoided with the economic potential of ORCs for steel, cement and
glass with 𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 years .................................................... 59

Table 4.9. Cost savings with the economic potential of ORCs for steel, cement and glass with
𝜼𝒆𝒍, 𝒏𝒆𝒕, 𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 years..................................................................... 60

Table 6.1. ORMAT industrial WHR projects globally [102] ..................................................xxxviii

Table 6.2. Turboden industrial WHR for cement industry projects globally [104]................... xxxix

Table 6.3. Turboden industrial WHR for steel industry projects globally [104]............................ xl

Table 6.4. Turboden industrial WHR for glass industry projects globally [104] ...........................xli

Table 6.5. Turboden industrial WHR for other industries projects globally [104] ....................... xlii

Table 6.6. Main German steel plants with BOF ....................................................................... xliii

Table 6.7. Main German steel plants with EAF ....................................................................... xliv
X

Table 6.8. Main German steel plants with RHF ....................................................................... xlvi

Table 6.9. Main German cement production plants ............................................................... xlviii

Table 6.10. Main glass production plants in Germany ................................................................. l


Abbreviations XI

Abbreviations
Abbrev. Meaning

AS Average Scenario
BCS Best-case Scenario
BF Blast Furnace
BOF Basic Oxygen Furnace
CC Clinker Cooler
CEPCI Chemical Plant Cost Index
CHP Combined Heat and Power
EAF Electric Arc Furnace
EIA U.S. Energy Information Administration
EII Energy-Intensive Industry
ESF Elbe-Stahlwerke Feralpi
GHG Greenhouse Gases
GWP Global Warming Potential
H-REII Heat Recovery in Energy-Intensive Industries
LCOE Levelized Cost of Energy
LD Long Dry Kiln
NSP New Suspension Preheater
ODP Ozone Depletion Potential
OPC Ordinary Portland Cement
ORC Organic Rankine Cycle
PH Preheater
RHF Reheating Furnace
SDS Sustainable Development Scenario
SIC Specific Investment Costs
SP Suspension Preheater
SWH Specific Investment Cost
WCS Worst-case Scenario
WHR Waste Heat Recovery
XII
Introduction 1

1 Introduction

1.1 Motivation
Direct CO2 emissions from the industrial sector (including energy and process emissions)
accounted for 24% of global emissions in 2017 [1]. Due to the limited use of renewable
energies for high-temperature processes as required by Energy-Intensive Industries (EII),
a reduction in energy consumption appears as the best way to reduce the emissions of this
field.
Energy consumption can be reduced by an efficient energy use, e.g., avoiding releasing
large amount of energy unused to the environment as waste heat. The waste heat from
industrial processes can be recovered by means of Organic Rankine cycle (ORC) technol-
ogy. These systems can convert the available waste heat from these processes and pro-
duce electricity or Combined Heat and Power (CHP). The electricity produced can, for ex-
ample, be reused internally in the same plant reducing its electricity demand. Moreover, the
heat recovery reduces the temperature of the exhaust gases. That is useful to avoid the
additional cooling of the gases before entering the gas treatment unit. Generating electricity
from recovered thermal energy can cover the electricity needs of the gas treatment unit as
well [8].
The high waste heat potential from the industrial sector in Germany [2] has not been fully
exploited yet [3]. This, together with the acknowledged suitability and good performance of
the ORC systems already installed in several industries worldwide, makes it interesting to
estimate both the technical and economic potential of the ORC integration for WHR from
the main industries in Germany.
Thus, in the context of the European Commission’s Energy Efficiency Directive, which
establishes a set of measures to help the EU reach its 20% energy efficiency target by
2020, the WHR is fundamental. The 20% energy efficiency target consists of saving 20%
of primary and final energy consumption by 2020 compared to business-as-usual projec-
tions [9].
In Germany, the Energy Efficiency Fund (EEF) is part of the so-called “Energiewende”
which aims to reduce primary energy consumption by 20% by 2020 and by 50% by 2050
(compared to 2008). In addition, GHG emissions are to be reduced by 40% until 2020 and
by 80 to 95% until 2050 (base year 1990). The fund currently consists of 23 policy
measures, including support programs and educational activities. The funds come from the
Federal Ministry for Economic Affairs and Energy (BMWi). The programs are managed by
various federal authorities or the development bank KfW. The beneficiaries are businesses,
households and municipalities [7].
There is a second program in Germany called the "Climate Action Program 2020". This
program aims to support the achievement of the goal of reducing the greenhouse gas emis-
sions in Germany by 40 percent by 2020 compared to 1990 [7].
2 1.2 Task

There is a lack of studies on the WHR with ORC potential in the main EII in Germany,
especially on the economic potential. Most of the current estimations are derived from stud-
ies in other countries. The aim of this thesis is, therefore, to provide the researchers and
the German EII with more knowledge about the potential of this WHR technique with ORC
power systems and to support its development. The results in this work are, thus, of great
help to overcome some of the current technological and information barriers.

1.2 Task
To assess to which extent ORC systems can be used to recover waste heat from indus-
trial processes, the technical and economic potential for WHR must be estimated. Because
of the impact of scale on costs, an analysis of the industrial sites in the country is required
to characterize the size of the plants. The results are of large interest to assess the feasibility
and importance of the application.
Following steps are required to assess the potential of WHR with ORC in industry:
• Literature review of the industrial processes from the energy-intensive industries,
the waste heat potential and ORC technology.
• Analysis of the main industrial plants in the country (for the production of steel, ce-
ment and glass) and characterisation of their dimensions.
• Estimation of the performance of the waste heat recovery process and the technical
potential.
• Estimation of the economic potential considering costs, amortization time and other
economic aspects.
Theoretical background 3

2 Theoretical background

2.1 Energy in the industrial sector

2.1.1 Impact of the industrial sector growth worldwide


According to the U.S. Energy Information Administration (EIA), the increase in the de-
mand of industrial products has led to a considerable growth of the industrial sector in the
recent years worldwide. In 2017, this sector accounted for 37% of the global final energy
use [1].
The reason for this growth is the current long-term rising trend in production of energy-
intensive industries, such as iron and steel, cement, chemicals, pulp and paper and alumin-
ium. According to Fernandez-Pales, et al. (2019) [1], EIA analysts, India and China have
experienced the highest growth, whereas the industry energy use has slightly declined in
Europe and the Americas.
The energy mix of the industrial sector has barely experienced changes since 2010. The
fossil fuel contribution is still strong, although it decreased from 73% to 70%, and the elec-
tricity use rose from 18% to 21%, mostly due to non-energy-intensive industry. Solar thermal
and geothermal energy accounted for less than a 0.05% of the total final energy use in 2017
because, although they continue to expand, they cannot provide high enough temperatures
for medium- and high- temperature heat processes. There is a lack of high-temperature
renewable fuels [1].

The fact that the energy mix remains almost unchanged and the current and expected
industrial production growth has led to a rise in CO2 emissions, which reached 8.5 GtCO2
in 2017, accounting for 24% of global emissions (Figure 2.1). In a Sustainable Development
Scenario (SDS), defined by the IEA, emissions must decline to 8.3 GtCO2 by 2030 [1].

To get on track with the SDS, efforts need to be done. According to Fernandez-Pales et
al. [1], material and energy efficiency, policy coverage (not only for energy efficiency and
process optimization, but also other factors related to industrial emissions such as process
emissions and technological shifts) and innovation efforts are key factors for medium- to
long-term CO2 emissions reductions in industry.
4 2.1 Energy in the industrial sector

Figure 2.1. Industry direct CO2 emissions by IEA [1]

2.1.2 Energy efficiency as best alternative for CO2 emissions re-


duction
Apart from the environmental impact, the increasing energy prices and uncertainties in
supply make a reduction of the reliance on fossil fuels desirable. To counteract this, there
are two options:
• The use of renewable energy systems.
• The reduction of energy consumption.
The use of renewable energy systems is becoming a more favorable option as prices
drop, but they cannot be extensively installed in every location and the investment costs are
still too high. Therefore, the reduction of energy consumption has the potential to be the
best alternative. There are three ways of reducing the energy demand of industry [10]:
• A reduction in total activity.
• Better energy management.
• Recovery and use of waste energy.
As stated by Woolley et al. (2018) [10], the first option requires an important change in
the company’s business model, and it is not suitable for all types of company. However,
energy management and recovery have been proven to be suitable for consumption im-
provement. Nevertheless, before it is used, waste heat should be avoided at first through
the energy management and optimization of processes.

2.1.3 Industrial waste heat recovery as a solution to energy effi-


ciency
Industrial waste heat is the heat that is rejected from industrial processes, in which en-
ergy (mostly heat and electricity) is used to produce high-added value products [2]. In other
Theoretical background 5

words, it is the heat that escapes a system and is released into the environment. As per
Papapetrou et al. (2018) [2], this waste heat is in a thermal carrier, which can be gaseous
(e.g. exhaust gas, steam, cooling air, etc.), liquid (e.g. cooling water, hot oil, etc.) and solids
(e.g. products, such as hot steel). In industrial facilities, sources of waste heat can be fur-
naces, combustion engines, refrigeration systems, waste water from washing, drying or
cooling processes, etc.
Waste heat is then calculated as the heat of a particular carrier, once cooled down from
its initial temperature. The larger the difference between the initial and final temperature,
the more heat can be obtained. The amount of waste heat can be calculated as follows:

𝑄̇ = 𝑐 ∙ 𝑚̇ ∙ ∆𝑇 ( 2.1)

where 𝑐 is the heat capacity of the heat carrying medium, 𝑚̇ is its mass flow rate and ∆𝑇
the difference between the initial and final temperature.
The ideal case where more heat is obtained would be to cool down to the ambient tem-
perature. However, the cooling is technically limited if "conventional" and cheaper materials
are used. Most flue gases produced by the combustion of fuels contain contaminants that
can condense into sulfuric, sulphurous or hydrochloric acids. Thus, it is recommended to
keep the flue gases above the dew point in order to avoid condensation of water vapor and
the production of corrosive acids [11]. This implies off-gas temperatures to the atmosphere
of 100-200oC.
Another important factor is the temporal distribution of the process occurrence over the
day, week and year. A more continuous waste heat flow is more favorable.
The heat demand varies from one industry to another and also in the same industry
depending on the production processes. They can vary between 60°C (cleaning processes)
and more than 1000°C for products of EII [11].
Waste heat can be classified into high, medium and low temperature grades (Table 2.1).
Waste Heat Recovery (WHR) systems are specific for each range of waste heat in order to
obtain the most efficient recovery.

Table 2.1. Waste heat temperature grades depending on their range [12]

Temperature grades Temperature range [°C]

High >400

Medium 100-400

Low <100

2.1.4 Industrial waste heat potential in Germany


In Germany, 28.9% of the total energy is consumed by the industrial sector, with an
amount of 750 TWh/a (Figure 2.2). Two thirds of this energy are used as process heat.
6 2.1 Energy in the industrial sector

Industry Transport
750 TWh 765 TWh
28.9% 29.5%

Trade and
services
Households 401 TWh
675 TWh
15.5%
26%

Figure 2.2. Final energy consumption in Germany in 2017 by sector [13]

A study performed by Aalborg University in the context of the Heat Roadmap Europe
2050 project [14] developed the first industrial waste heat estimation assessment at EU27
scale. This included the main industrial plants within the energy-intensive sectors (chemical
and petrochemical, food and beverage, iron and steel, non-ferrous metals, non-metallic
minerals, and paper, pulp and printing) and refineries. The results showed that Germany
has the highest industrial waste heat potential in Europe (19.39%) (Figure 2.3).

Figure 2.3. Locations of major EII with considerable volumes of excess heat. Source: The E-
PRTR database at EEA in Copenhagen [14]

Papapetrou et. al [2] estimates the available excess heat from the German industrial
sector in approximately 75 TWh/a (Figure 2.4). This is considerably higher than the rest of
the U-27 countries.
Theoretical background 7

Figure 2.4. Waste heat potential in each EU country per temperature level in all industries [2]

2.2 Technology for waste heat recovery: state-of-the-art

2.2.1 Overview
The choice of the technology to recover waste heat is highly affected by the temperature
of the waste heat, as well as the phase and chemical composition of the exhaust stream.
According to Miró (2016) [15], the main different options to recover the waste heat are:
• Direct use without upgrading.
• Use after upgrading, through heat pumping.
• Power generation.
• A combination of the previous.
Depending on these different uses of the waste heat, technologies can be categorized
as passive or active technologies (Figure 2.5).
The two main passive technologies are heat exchangers and thermal energy storages
(TES). TES are used in systems with fluctuating waste heat streams. This technology can
be used for reusing waste heat at later time within an industry to heat or preheat other
processes.
If the heat downgrade is below ambient temperature, then a cooling cycle is required
(active system). Active technologies include heat pumps, chillers and heat engines. Active
applications of waste heat provide heat at higher temperature level, cooling (at temperature
lower than ambient), or electricity. Technologies to provide heat or cold can be also called
heat transformation technologies as they modify the inlet temperature upgrading or down-
grading it.
For power generation, variants of the basic Rankine thermodynamic cycles (steam cy-
cles, ORCs, etc.) are used, apart from the different thermoelectric devices.
8 2.2 Technology for waste heat recovery: state-of-the-art

Thermal energy storage


Direct use
Heat exchanger

Technologies for industrial Heat Pump


waste heat recovery Upgrade/downgrade
Chiller

Power generation Heat engines

Figure 2.5. Technologies for industrial waste heat recovery/re-use [15]

2.2.2 Electricity generation


When there are no other economical uses for the waste heat, its conversion to a more
transportable form of energy such as electricity must be evaluated. Power production from
waste heat is nowadays not as widespread as the other technologies. However, it can be
the most economic option, especially for high temperature waste heat. At low or medium
levels of excess heat, the economics can be boosted by high power prices and/or powerful
policy instruments for GHG emissions [16].
Table 2.2 summarizes the options for heat recovery via power generation.
Table 2.2. Options for Heat Recovery via Power Generation [17]

Thermal Conversion Tech- Temperature


Typical Sources of Waste Heat
nology Range

Gas turbine exhaust, reciprocating engines, incinera-


Traditional Steam Cycle Medium, High
tors and furnaces

Kalina Cycle Low, Medium Gas turbine exhaust, boiler exhaust, cement kilns

Gas turbine exhaust, boiler exhaust, heated water,


Organic Rankine cycle Low, Medium
cement kilns

Thermoelectric generation Medium, High Not yet demonstrated in industrial applications

Piezoelectric generation Low Not yet demonstrated in industrial applications

Thermal Photovoltaic Medium, High Not yet demonstrated in industrial applications

A brief description of the less conventional technologies including ORC is made in the
following paragraphs based on [17].
Thermoelectric generation
Thermoelectric generators are based on the Seebeck effect, a phenomenon in which a
temperature difference between two dissimilar electrical conductors or semiconductors pro-
duces a voltage difference between the two materials. They require a large temperature
Theoretical background 9

difference between hot and cold side. The efficiencies range between 2 and 5%. New ma-
terials for thermoelectric generators are being developed in countries such as Germany,
USA and China. Variable costs are practically zero, so the SIC [€/kW] needs to be reduced
by means of increasing the efficiency or decreasing the material and production costs.
Piezoelectric power generation
This is a good option for converting low temperature excess heat (100-150°C). These
devices convert mechanical energy in form of ambient vibrations into electrical energy. Their
efficiency is, however, low (1% approximately) and they have high investment costs.
Thermo photovoltaic generator
These generators convert radiant energy to electricity and could enable new methods
for WHR, although only a small number of prototype systems have been built for small
burner applications and gas turbines.
Organic Rankine Cycle
There is a growing interest and development of ORC nowadays. In many industrial pro-
cesses, waste heat flows are discontinuous and therefore the cycle needs to be flexible.
ORC systems, which are already in operation since decades, can work at part-load condi-
tions up to 10%, while for example steam-cycle need more constant conditions [18].
Normally, the working fluid in a Rankine cycle is water, but when the heat source is at
lower temperature, it becomes of interest to use new working fluids that are performant in
this range. The ORC has the same elements and same working principle as a conventional
Rankine cycle. The only difference is the working fluid, which is organic, characterized by a
molecular mass higher than that of water, which leads to a slower rotation of the turbine.
However, environmental issues need to be considered when choosing.
The use of this organic fluids reports some advantages compared to the use of water
[19]. For a better understanding of the advantages explanation, see Figure 2.6 and Figure
2.7. The advantages are the following:

• Critical temperature and pressure are reduced. This makes possible to recover
heat at lower temperature than in a conventional cycle.
• Condensation pressure at ambient temperature higher than the atmospheric:
therefore, the leakage of air into the circuit is avoided and thus, extraction sys-
tems are no longer needed, simplifying the system.
• Evaporation pressure lower than in conventional Rankine cycle: complexity and
costs are therefore reduced.
• Null or positive slope of saturated vapor line in the T-s diagram: some organic
fluids show this, which makes possible the inlet into the turbine with saturated
vapor without the risk of having moisture at the outlet. This reduces the blades’
erosion, increasing their life span.
10 2.2 Technology for waste heat recovery: state-of-the-art

Figure 2.6. T–s diagram of water and various typical ORC fluids [20]

All this makes possible for ORCs to take advantage of low temperature or high variability
waste heat streams. However, there are also some disadvantages:
• The latent heat of organic fluids is way lower than of water: therefore, they need
a higher mass flow in order to capture the same thermal power in the evaporator
and thus, the pump’s consumption is higher.
• Water has very favorable properties: it is neither toxic nor flammable, it is very
stable, and the cost is lower.
The overall efficiency of ORC units is around 10-20%, depending on the temperature of
the condenser and evaporator. A Carnot engine (maximum possible efficiency) operating
with a heat source at 150°C and rejecting it at 25°C is only about 30% efficient, so an effi-
ciency of 10-20% is a relatively high one, especially compared to other low temperature
heat recovery options [21].
The organic fluid must be environmentally friendly (low Global Warming Potential, GWP,
and Ozone Depletion Potential, ODP), non-flammable, non-corrosive and chemically sta-
ble). Among the options for this fluid there are silicon oil (siloxanes), propane, isopentane,
isobutane, haloalkanes or p-xylene.
A short explanation of the thermodynamic cycle taking place in an ORC turbogenerator
by Turboden (2019) is presented here.
The ORC turbogenerator uses medium-to-high-temperature heat source to preheat and
vaporize a suitable organic working fluid in the evaporator (4>5). The organic fluid vapor
makes the turbine rotate (5>6). This turbine is directly coupled to the electric generator,
which produces electric power.
The exhaust vapor from the turbine goes through the recuperator (6>7), where it heats
the organic liquid (2>3) and is then condensed in the condenser and cooled by the cooling
circuit (7>8>1). The organic working fluid is then pumped (1>2) into the recuperator and
evaporator, completing the closed-cycle operation.
Theoretical background 11

Figure 2.7. The Rankine cycle [22]

2.2.3 Waste heat recovery system with ORC and intermediate


thermal circuit
Figure 2.8 depicts a general WHR system where it can be seen that the waste heat that
arrives from the industrial process into the heat exchanger for the heat transfer to the inter-
mediate loop is at 200-500oC and cooled down to 100-200°C, but not further to avoid con-
densation of acids in the flue gas and therefore corrosion.

Figure 2.8. WHR system with ORC and intermediate thermal circuit scheme [23]
12 2.2 Technology for waste heat recovery: state-of-the-art

The off-gas temperature varies from one industrial process to another, and so the differ-
ent layouts can be conceived to optimize the heat recovery. Some processes, such as the
EAF, have unsteady heat flux. This high variation in temperature and velocity needs a steam
accumulator in an intermediate thermal circuit or a secondary thermal oil loop which acts
like a buffer. This way, the ORC can operate at steady conditions and the risk of hotspots
like, for example, in a directly heated evaporator, is reduced [24].

2.2.4 Advantages of an ORC waste heat recovery system


ORC systems present several advantages that make them very adequate for WHR from
industrial processes and that can be summed up as follows [25]:
• No erosion of blades, low mechanical stress of the turbine.
• No corrosion issues on the working fluid side (use of non-aggressive working fluid).
• Ease of integration in main process: do not affect plant operations.
• Demonstrate very good part-load behavior (Figure 2.9).
• Flexible and modulable operation (good for variable or intermittent working regimes).
• Do not consume water and no need of water treatment.
• Low O&M requirements.
• ORC plants remotely monitored.

Figure 2.9. ORC part load efficiency [22]


Theoretical background 13

2.3 ORC power systems: state-of-the-art

2.3.1 Overview
ORC technology has been used for years in power generation applications such as solar,
geothermal, biomass and WHR. As of 31st December 2016, the ORC technology repre-
sented a total installed capacity around 2701 MW worldwide, distributed over 705 projects
and 1754 ORC units. The data comes from the websites and list of references of ORC
manufacturers [4].

Figure 2.10. Total installed capacity per application [4]

Power generation from geothermal brines is the main field of application with 74.8% of
all ORC installed capacity in the world (Figure 2.10).
WHR is an emerging field for ORC with an interesting potential for all unit sizes. ORC
power for heat recovery counted with 376 MW of installed capacity in the world by 2017,
and 39 MW of new capacity in construction (16 projects). The heat recovery market is still
at an early stage but has already passed the demo/prototype phase.
The main application is, by far, heat recovery from Diesel or gas engines and turbines
(65% of the total installed capacity) as can be seen in Figure 2.11. The reason for this
advantage in installed capacity is the fact that using exhaust heat from combustion engines
or turbines is easier than industrial heat recovery. However, engines are becoming more
efficient and this application is not considered as renewable in many countries developing
energy transition roadmaps [4].
Recovering the waste heat from the metal industry, which is mainly done by China and
Italy, represents the 7.5% of the applications. Cement and lime (9 projects) and glass (8
projects) industries follow.
14 2.3 ORC power systems: state-of-the-art

Figure 2.11. Shares of ORC installed capacity per heat recovery application [4]

2.3.2 ORC installed capacity worldwide


Being the North American company Ormat the world leader for the total ORC installed
capacity worldwide, a research on their industrial WHR with ORC has been carried out.
Their main WHR applications are done in the USA and Canada for gas pipeline compressor
stations and to recover waste heat from gas turbine exhaust from gas processing plants
into power. Some of the WHR ORC from Ormat are listed in Appendix B.
The Italian company Turboden has a larger number of units recovering waste heat in
industrial processes (Appendix B. The applications are mainly WHR from cement, steel and
glass production processes.
Furthermore, one of the latest innovative projects from Turboden in Germany is a 2.7
MW ORC unit that uses the heat from a portion of the saturated steam recovered from the
exhaust gases of an EAF. This plant is Elbe-Stahlwerke Feralpi (ESF) GmbH, in Riesa. This
company is one of Europe's most qualified iron and steel manufacturers and was the first
to introduce an EAF WHR ORC-based system and is considered a reference in this work.
The plant is a part of the H-REII DEMO Project (Heat Recovery in EII) and was co-financed
by the Directorate-General for Environment of the European Commission under the LIFE
Program for its technological innovation and environmental value [26]. Both this project and
program have been essential for the WHR with ORC systems development and are de-
scribed below.

2.3.3 European projects for waste heat recovery with ORC sys-
tems in energy-intensive industries
The H-REII DEMO Project
The H-REII DEMO Project is the first European project on mapping the potential for heat
recovery with ORC systems in EII (in a pilot area). It started in 2010 and it was born to
promote policy and governance actions to support WHR for power generation in EII and
Theoretical background 15

quantify the potential CO2 savings. The project confirmed the potential of EII heat recovery
systems as effective energy-saving tools and thus contributed to the Europe 2020 strategy
goals. Positive results concerning application of heat recovery systems in 22 of the Italian
glass, cement, steel and other EIIs were recorded. An additional 48 audits were carried out
in Austria and the methodology’s transferability was then confirmed [27].
The H-REII DEMO Project is the continuation and implementation of the H-REII Project
and had two main goals [25]:
• Developing the first prototype of ORC heat recovery plant from EAF in the steel
industry completely integrated in a fumes treatment cleaning system by using
water in a closed loop for cooling waste fumes and operating at a higher temper-
ature and pressure than traditional methods. Thanks to this, the total power con-
sumption decreased and there was an improvement in the performance of the
fume depuration plant in energy-intensive industrial applications.
• Promoting EU policy and governance actions for incentivizing WHR for power
generation, reducing CO2 emissions by the valorization of process effluents in
EII.
Both H-REII and H-REII DEMO are participating projects in the LIFE Program.
The LIFE Program
Over the last 20 years, the European Union has set up a series of legislative measures
to help decoupling industrial production from CO2 emissions. These are: the Emissions
Trading System, Directive on Energy Efficiency, Industrial Emission Directive, Ecodesign
Directive, Energy Performance of Buildings, Electricity Market Design, Renewable Energy
Directive, the 2020 Climate and Energy Package, the 2030 Climate and Energy Framework,
and the 2050 Low Carbon Road Map [27].
According to the European Commission [27], the LIFE Program and its Climate Action
sub-program is a European Union’s financial instrument that contributes to the development
of low-carbon technologies, the uptake of Best Available Techniques and the validation of
good practices in the energy-intensive sector.
As the first platform meeting to cover EII since the start of the LIFE Program, the Utrecht
meeting attracted some seventy participants including beneficiaries of the LIFE and Horizon
2020 programs, policy-makers from the EU and national level, civil society and private sec-
tor representatives with the goal to discuss how EEIs can contribute to emissions reduction
and, eventually, decarbonization, following the European Union’s roadmap. The platform
meeting covered the glass, ceramics, cement, steel and other metals sectors.
The platform meeting led to the conclusion that European industry has the skills, ideas
and willingness to innovate and contribute to achieving the EU emission mitigation targets.
However, there are barriers, and they are not only technical. They also have very much
to do with the framework of policies, funding mechanisms, incentives and disincentives.
Therefore, policy-makers and financial institutions, as well as the overall functioning of the
economy, play a crucial role in enabling the transformation of industrial activities towards a
decarbonized future [27].
16 2.4 Waste heat recovery potential

2.4 Waste heat recovery potential

2.4.1 Definitions of waste heat potential


In order to estimate the potential of industrial WHR with ORC technology, some infor-
mation on the different kinds of potentials must be given. Three different types of potentials
(Figure 2.12) are distinguished [15]:
• Theoretical or physical potential: it considers physical constraints. For example,
that only waste heat above ambient temperature is considered.
• Technical potential: it takes into account constraints related to technical aspects.
For instance, the minimum off-gas temperature to avoid acid condensation and,
therefore, corrosion. This potential depends on the technologies used.
• Economic potential: it considers financial parameters like energy prices, interest
rates, etc. to see if the project is economically feasible and profitable.

Theoretical
potential

Technical
potential

Economic
potential

Figure 2.12. Types of potential

2.4.2 Types of waste heat recovery potential estimation methods


Normally, methods for identifying excess heat are classified as either top-down or bot-
tom-up methods, which can also be combined [16]:
• The top-down approach: it starts from the primary energy use. Assumptions
about efficiencies and distribution of the energy use are made in order to estimate
the excess heat potential for different sectors. In some cases, the waste heat is
estimated according to key figures available for other countries and adapted to
the specific case.
• The bottom-up approach: using questionnaires or measurements, specific data
from representative companies and site is surveyed. Starting from data of the
single facilities, considerations on the total waste heat are made. This is a more
Theoretical background 17

complex and more accurate method, yet more resources (economic and time)
consuming.

Figure 2.13 summarizes the different approaches.

Waste heat
estimation
methods

Estimation
Survey
Bottom-
Bottom-up
up/Top-down

Field Existing Process,


measurement company data Sector, Region

Waste heat
Mandatory Online
Questionnaire Efficiency per company
reports database
size parameter

Heat or energy Company size


demand parameter

Figure 2.13. Classification of methods [28]

In addition, in 2008, Blesl et al. [29] differentiated three methods to estimate waste heat
by accuracy: a rough method (using few statistical data), a medium precise estimate (with
more detailed literature data and coefficients), and a high precision method (with measured
data).

2.4.3 Waste heat recovery potential estimation for Germany ap-


proaches: state-of-the-art
Several works tried to estimate the industrial waste heat potential in Germany (Figure
2.14) . So far, almost only top-down analyses of the waste heat potential exist and they are
using key figures derived from studies in other countries [11] [2].
Pehnt et al. (2011) [11] used the results of a Norwegian study to determine key figures
for energy efficiency in different industrial processes and adapted the findings to the Ger-
man industrial structure. Using a top-down approach, a potential of 476 PJ/a was assessed,
66% of it above 140 °C.
18 2.4 Waste heat recovery potential

Connolly et al. (2013) [14] applied emission facility-based data and geographical infor-
mation systems (GIS) to determine the excess heat available in view of possible integration
into district heating systems. They estimated a potential amounting to 525 PJ/a in Germany
and 2708 PJ/a in the EU-27 [30]. These values were slightly corrected by Persson et al.,
according to whom the assessed industrial excess heat in the EU-27 amounted 2,924 PJ/a,
566 PJ/a of which in Germany [31].
Campana et al. (2013) [18] estimate the ORC WHR potential in European (EU-27) EII.
Savings in electricity costs and GHG are quantified. The study is based on the analysis of
available energy audits. Energy audits provide the ORC power installed, or potentially
installable, in a considered plant analyzed during a feasibility study.
Recently, Papapetrou et al. (2018) [2] based their methodology on the estimation of
waste heat fractions per industry, EU-28 country and temperature level. The starting point
is a study of the UK industry where the technically available waste heat of each industrial
sector in the UK was estimated using data from 425 industrial sites from the period 2000–
2003. The results using a top-down approach estimated an industrial waste heat of 269.5
PJ/a for Germany.
However, Brueckner et al. (2017) [32] presented the first bottom-up approach for esti-
mating the industrial waste heat potential in Germany. This approach uses the CO2 emis-
sions report data from German production companies to calculate a conservative and lower
boundary value for the industrial waste heat. The industrial waste heat volume was evalu-
ated as 127 PJ/a.

Figure 2.15. Estimations of industrial waste heat in Germany in increasing order

2.4.4 Notes on the technical potential


For the understanding of the methodology, some notes in relation to the technical poten-
tial estimation are made.
• In the first place, the first idea for obtaining the waste heat available in each plant
was to access the waste heat source maps for the different states in Germany,
Theoretical background 19

where particular waste heat sources are made visible. Unfortunately, only Ba-
varia and Thuringia have developed these maps and only Thuringia’s map had
useful information about the industrial sites studied in this work.
Therefore, a method for estimating the waste heat from each plant was de-
veloped. However, the small amount of accurate and free access information on
the energy audits on the different plants made the task complicated.
• In the second place, for the estimation of the net electric energy that can be pro-
duced with an ORC power unit, the most suitable unit must be selected.
Turboden produces a wide range of ORC modules which can suit the varying
requirements of heat recovery projects. Turboden’s HR units are for standard
heat recovery applications. There are different types of HR units depending on
the characteristics of the primary heat source and on whether there is require-
ment for heat or not.
The nominal net electric efficiency depends on each HR model which de-
pends on the thermal power input.

2.4.5 Notes on the economic potential


The attractiveness of ORC is strongly dependent on economic feasibility. Once the tech-
nical potential has been estimated, the economic potential can be calculated based on the
method used by Eyerer, et al. (2017) [33] for geothermal energy on their work “Potential of
hydrothermal geothermal energy for power generation in Germany”.

Specific investment cost


In order to carry out a financial-economic analysis of a project, the annual cash flows,
composed of the capital investment and annual expenses and revenues, must be consid-
ered. The capital investment takes place at the beginning of the project. It includes the costs
directly associated with the system (materials, equipment, labor etc.) and indirect costs (en-
gineering, construction costs and contingencies). The main part of an ORC project’s capital
investment are the components of the ORC module itself: evaporator, expander and gen-
erator, condenser and pump. Another important part are the costs for integration of the ORC
module into an existing plant (e.g., for heat recovery applications). The annual expenses
for ORC projects depend on the type of application and the location, but their importance is
usually less than the investment costs.
To calculate the SIC [€2018/kWe], a correlation between the ORC size and the specific
investment cost (SIC) is used (Equation 2.2) based on Turboden products (both direct heat
exchange and intermediate heat transfer fluid configurations) from [34]. In this work, all
considered ORC units are equipped with an intermediate oil loop. Specific investment costs
can vary between 4000 down to 2000 USD/kWe for larger units [34]. In Figure 2.16, it can
be observed that the larger the ORC module the lower the SIC.
20 2.4 Waste heat recovery potential

€2018 ( 2.3)
𝑆𝐼𝐶 [ ] = 19774 · 𝑃𝑂𝑅𝐶 [𝑘𝑊𝑒 ]−0.277
𝑘𝑊

Figure 2.16. SIC function [34]

Levelized Cost of Energy


For the analysis of the economic potential, the Levelized Cost of Energy (LCOE) will be
calculated for each ORC unit. The LCOE measures lifetime costs divided by total energy
production and calculates the present value of the total cost of building and operating a
power plant over an assumed lifetime. This indicator is crucial for making the decision to
proceed with development of a project or not.
𝐶𝑀𝑙
𝑆𝐼𝐶 · 𝑃𝑛 + ∑𝑛𝑙=1
(1 + 𝑖)𝑙 ( 2.4)
𝐿𝐶𝑂𝐸 =
𝐸𝑙
∑𝑛𝑙=1
(1 + 𝑖)𝑙

“SIC” is the Specific investment cost, ”Pn“ the nominal power, “n” the assumed lifetime,
“l” the present year, “i” the discount rate, ”CMl“ the maintenance costs and “El“ the energy
production in year “l”. The SIC multiplied by the nominal power is the initial investment.
For this study, the following assumption has been made based on the work by [35]:

𝐶𝑀𝑙 = 2% · 𝑆𝐼𝐶 · 𝑃𝑛 ( 2.5)

The interest rate (i) and the amortization years (n) will be varied for the economic poten-
tial analysis.
ORC LCOE vs. electricity prices for EII
Once the LCOE for each ORC unit is obtained, it must be compared to the Levelized
Cost of the Electricity for the industry in Germany [5]. This electricity price is different for
households and for industry, as exemptions from fees, taxes and levies are applied [36].
These exemptions are different depending on the size of the site.
1. Exemption from grid access fees
Theoretical background 21

Energy-intensive companies can exempt themselves from paying grid access fees if they
meet certain criteria. The legal foundations for this are provided in the German Electricity
Network Charges Ordinance (StromNEV). For example, partial exemptions from this fee is
possible if a company’s peak load can be expected to deviate from the peak load situation
of the grid. These companies may be given discounts of up to a maximum of 80% of the
grid access fees. In the case of steelworks, for example, these discounts can be from 30%
to 40% [36].
Also, companies can negotiate individual grid access fees and exempt themselves from
up to 90% of the regular fee. If the electricity consumption by a company sums up to at least
7,000 hours of use from a single consumption point within the general supply grid and it
exceeds 10 GWh, an 80% deduction is possible. If the own consumption amounts to 7,500
hours, the deduction is the 85%, and in the case the consumption is 8,000 hours, the 90%
[36].
2. Exemption from taxes and levies
The electricity price in Germany comprises, not only the wholesale market prices (sup-
plier’s cost, grid charges) and the grid access fees, but also a series of taxes (sales tax,
electricity tax), levies (concession levy, offshore liability levy) and surcharges (surcharge for
Combined Heat and Power plants, renewable energy surcharge). Electricity intensive in-
dustries, however, benefit from important deductions on these. As by [36]:
Electricity supply to end consumers is subject to an electricity tax, also known as the
“ecological tax” in Germany. However, electricity intensive industrial production processes
are exempted from this tax. All other industries can benefit from a tax deduction of up to
90%.
The remuneration scheme for electricity generation from renewable energies sets a sur-
charge to cover its costs. However, electricity intensive industries can benefit from signifi-
cant reductions on this surcharge.
The support scheme for Combined Heat and Power plants (CHP) is also financed with a
surcharge on electricity deliveries to end consumers. Operators of CHP plants receive a
guaranteed price on the electricity they sell. The difference between the guaranteed price
and the actual price they receive on the market is financed through this surcharge. Again,
for electricity intensive industries this rate is reduced.
Grid operators must pay damages if they fail to connect offshore wind farms in a timely
manner in order to sell the power they produce. However, operators can pass these costs
on to consumers through a levy. The rates are also reduced for electricity intensive indus-
tries.
Considering all these taxes and levies reductions and exemptions for electricity intensive
industries, the average price for 2018 and for large industrial customers (annual energy
consumption superior to 10 GWh/a) according to the BDEW [5] is 8.84 ct/kWh. The mini-
mum price goes from 5.10 to 5.80 ct/kWh and the maximum price ranges from 14.40 to
17.00 ct/kWh (Figure 2.17).
22 2.5 Steel industry

Figure 2.17. Minimum and maximum electricity prices for EII (annual energy consump-
tion>10GWh) and average electricity price [5]

2.5 Steel industry

2.5.1 Description of the production process


Steel is an alloy made by combining iron with carbon and other chemical elements. There
are two main routes to produce steel: the integrated and the recycling route. The integrated
route is based on the production of iron from iron ore. The recycling route uses scrap iron
as the main raw material. In both cases, the energy consumed comes from fuel (mainly coal
and coke) and electricity. The recycling route consumes about 80% less energy than the
integrated route. Around two-thirds of crude steel worldwide are melted down as oxygen
refined steel in integrated steelmaking plants. One-third is melted from scrap in EAFs [37].
Steel making involves different activities such as heating, cooling, melting and
solidification. The steel industry is a particularly energy-intensive sector. The reduction of
energy consumption is of a special concern.
The steel industry uses about 22 TWh of electricity every year in Germany (for powering
EAFs that melt scrap to new steel and for operating rolling mills), which is 9% of the indus-
trial consumption and 4% of total consumption in the country [38].
Most of the plants are integrated steel mills, which have all the functions for the primary
steel production iron production: iron making (blast furnace, if necessary), steel making
(converter: BOF or EAF) and the rolling mill with casting (solidification of the liquid steel),
roughing rolling/billet rolling (reducing size of blocks) and product rolling (finished shapes).
Theoretical background 23

Figure 2.18. Steel production routes: a) Oxygen steelmaking along BF and BOF converter, b)
Electric steelmaking in EAF [37]

Based on the state-of-the-art for WHR with ORC, the most interesting processes for off-
gas energy recovery are described below. The blast furnace flue gas is not used for power
obtention with ORC (mostly because of the high dust content in the off-gas). Only the BOF,
EAF and reheating furnace (RHF) have been shown to be adequate for the ORC system
installation for power production.
Blast furnace
The blast furnace produces pig iron and slag from pretreated charging materials
(ore, sinter, pellets), reducing agents (coke, coal, oil, gas), and additives (lime, lime-
stone). The charged iron normally comes in the form of hematite (Fe2O3) and is reduced
to metallic iron. Coke and coal are used as reducing agents for iron oxide and as source of
energy from the combustion of the inherent carbon to create CO 2 [37].
Basic oxygen furnace
The molten pig iron (~1500°C) is transported from the BF to the BOF converter in a steel
plant in torpedo ladle cars. Normally, according to the World Steel Association (2019), the
residence time of the molten iron in the ladle is two to four hours. The pig iron still contains
tramp elements (C, Si, Mn, S, and P) which are typically removed in the combined-blowing
BOF by oxygen top-blowing and inert gas bottom stirring [37].
The BOF, then, refines a charge of molten pig iron and ambient scrap into steel of a
desired carbon and temperature using high purity oxygen. Inside a BOF, oxygen (O2) is
blown through the molten iron. It combines with carbon to form carbon monoxide (CO). The
CO is burned as it is generated. Steel is made in discrete batches called heats (every 30 to
65 min). The furnace or converter is a barrel shaped, open topped, refractory lined vessel
24 2.5 Steel industry

that can rotate on a horizontal axis. The basic steps of the process are shown in Figure
2.19.
The overall purpose of this process is to reduce the carbon content from about 4% to
less than 1% (and usually less than 0.1%), to reduce or control the sulfur and phosphorus,
and to raise the temperature of the liquid steel made to more than 1600°C [39].

Figure 2.19. General view of converter for BOF steelmaking process [40]

The flue gas from the BOF converter consists mainly of CO, CO 2, H2 and N2 being un-
combusted CO at 85-90%. The composition of this gas varies in time because the reaction
which releases the converter gas is discontinuous. For example, during oxygen blowing
periods CO flow is high. Within the last 25 % of the blowing time the CO content decreases,
since the carbon content of the hot metal has already been converted into CO. Hence,
heating values of converter gas vary depending on the gas composition. In addition, the
amount of converter gas which can be recovered from the BOFs can vary [41].
Electric arc furnace
The primary raw material for the EAF is normally scrap metal. The scrap metal is melted
and refined using electrical energy. During the melting, oxidation of P, Si, Mn, C, and other
materials occurs, and a slag containing some of these oxidation products forms on top of
the molten metal. To decarburize the molten steel and provide thermal energy oxygen is
used. This is a batch process with a cycle time of about 45 min in the most modern furnaces
(this is called tap-to-tap time). Since scrap metal is used instead of molten iron, no blast
furnace is required with an EAF [42].
This technology results in the production of metal dusts, slag, and gaseous products.
Emissions consisting of particulate matter and gases are conveyed into a gas cleaning sys-
tem (wet or dry) [42]. The off-gas composition is (time-averaged analysis approx.) 30% CO,
10% CO2, 55% N2, 5% H2 at temperatures from 1200-1750oC [43].
Theoretical background 25

Figure 2.20. Illustration of EAF [44]

Reheating furnace
The liquid steel produced in BOFs and EAFs is cast and solidified for subsequent further
processing. About 96% of the liquid steel in Germany undergoes a continuous casting pro-
cess [45]. The rest is cast in ingots. Casting operations are located at the end of integrated
steel plants (BF and BOF) or EAFs and before further processing operations, the rolling mill
or forging plant.
Steel temperature needs to be increased up to a suitable rolling temperature (300-1050°C)
before being processed by the hot rolling mills through a RHF, usually fueled with natural
gas.
The most common types of continuous RHFs include [46]:
• Pusher-type furnaces (e.g. for blooms, billets, slabs)
• Walking-beam furnaces (e.g. for blooms, billets, tubes, slabs)
• Rotary-hearth furnaces (e.g. for round bars)
• Inductive furnaces (e.g. for soaking of near-net-shape products)
Most of the plants studied have either pusher-type or walking-beam furnaces. Pusher-
type furnaces are an inexpensive alternative to more sophisticated walking-hearth or walk-
ing-beam types.
In Elbe-Stahlwerke Feralpi, for instance, in the rolling section (after the molten steel is
tapped into the ladle, analyzed and cast into intermediate billets on the five-line continuous
casting plant) the billets are reheated from 600 to 850°C to the rolling temperature of 1150
to 1200°C in a gas-fired walking-beam furnace [47].
At Stahlwerk Thüringen beam blanks are about 500°C when charged into the pusher-
type furnace and are reheated to 1200°C [48].
26 2.5 Steel industry

2.5.2 Waste heat recovery with ORC from steel industry: state-
of-the-art
Basic oxygen furnace
As an alternative to flaring this gas, energy can be recovered from it. There are two main
methods to recover energy from the BOF’s off-gas before it is cleaned [49]:
• “Open combustion” systems:
CO in the gas is fully or partially combusted by air. The heat generated is recovered in a
waste heat boiler to produce steam, which can be used to generate electricity. About 80%
of the total outgoing heat can be recovered. CO leaving the furnace is allowed to combust
by letting large amounts of air to enter the exhaust hood. The resulting hot gas from the
combustion is then used in a heat recovery boiler to produce high pressure steam. The
amount of steam generation can be regulated by the amount of CO burnt in the boiler.
However, as the steam is produced intermittently (BOF is a batch process), steam accumu-
lators can be installed to ensure a very constant steam flow rate. After the heat has been
recovered, the off-gas is flared, releasing CO2.
• “Suppressed combustion” (or “non-combustion”) systems:
These systems are the best option for both heat and fuel recovery. Air infiltration is re-
duced during the oxygen blowing in order to obtain a CO-rich gas, which is then collected,
cleaned and stored for subsequent use for power generation, as fuel gas or for other uses.
A waste heat boiler, generating steam, can recover the sensible heat of the gas before it is
cleaned and stored. This recovers around the 10-30% of the total energy output. Another
50-80% is recovered as chemical energy (CO) [49]. The advanced non-combustion closed
method has become the main stream. The facilities are designed to recover about 70% of
the latent heat and sensible heat [50].
Steam produced with the recovered chemical and sensible waste heat can then be in-
troduced into an ORC unit in order to produce electricity.
Electric arc furnace
The high amount of latent and sensible enthalpy in the off-gas after the steelmaking
process in the EAF offers high potential for WHR. Evaporative cooling systems installed at
dedusting systems of some EAFs are utilizing this waste heat for steam generation and
subsequent usage of steam for further applications [51].
In 2013, Turboden started up the first ORC that recovers heat from exhausted gases of
an EAF in the Feralpi Group plant of Riesa, Germany. Elbe-Stahlwerke Feralpi GmbH is
one of Europe's most qualified iron and steel manufacturers and the first to introduce an
EAF WHR ORC-based system. The plant is a part of the H-REII DEMO Project and was
co-financed by the DG Environment of the European Commission under the LIFE program
[26]. However, the first iRecovery project was commissioned in 2009 at the EAF of
Georgsmarienhütte GmbH, Germany.
This installation is based on the Evaporating Cooling System (ECS) of Tenova (iRecov-
ery®) and on the ORC technology of Turboden. Currently, evaporative cooling represents
the best solution for off-gas heat recovery because of its flexibility [52]. The benefits of this
Theoretical background 27

system include also less inner corrosion, lower water consumption, lower water volume and
higher safety in different emergency situations [53].
The system has the following parts ( [54], [8] ) that can be observed in Figure 2.21:
• Energy recovery section at High Temperature (HT section): The energy from the off-
gas is recovered through a first radiation heat exchanger that produces saturated
steam at the temperature of 245°C and at the pressure of 27 bar. This steam passes
through a drum that stores its energy. When the off-gas exits the radiation heat ex-
changer, it has a temperature of almost 600°C and is filtered before exiting the sys-
tem with a temperature of around 150°C. This section also includes a combustion
chamber. Post combustion of waste gases can be performed by additional burners
as well as by the supply of secondary air, both in the duct or in a combustion cham-
ber, or also in the furnace itself [46]. Post combustion in the furnace allows for a
partial utilization of the chemical energy of the CO contained in the off-gas. For this,
burners or lances are used to add oxygen in an excess stoichiometric relation.
Georgsmarienhütte EAF makes only use of this section.
• Energy recovery section at Low Temperature (LT section): Residual energy is re-
covered by a convective heat exchanger (waste heat boiler) that fills the steam
drum. Thus the thermal power made available is quite stable during the processes.
• Steam Drum: The thermal flow in the off-gas from an EAF varies during the melting
cycle (as it is a batch flow) and while the scrap material is loaded into the basket
there is no thermal power available. Due to this, and in order to storage thermal
energy, a steam drum is necessary.
• Accumulator: tank for steam accumulation used to satisfy the steam request when
not available by EAF batch process.
• ORC plant: plant for electrical energy generation.
• Steam user: In the Riesa project, some of the steam stored in the steam drum is
sold to another industrial process through the district heating grid.

Figure 2.21. General scheme of the energy recovery system installed to Elbe-Stahlwerke Fe-
ralpi [54]
28 2.6 Cement industry

It is to note that having both these HT and LT levels improves the efficiency of the WHR.
Whereas Georgsmarienhütte uses heat recovery down to 600°C, at Feralpi heat recovery
is designed down to 200°C. For any potential customer, the decision to install the iRecovery
HT or LT level system depends basically on steam demand and supply.
Reheating furnace
In an integrated iron and steel plant, one of the most intensive energy consumers is the
RHF. Due to working at high temperatures, RHFs need to operate as efficiently as possible
to reduce energy consumption and the environmental pollution caused by waste gases.
ORC technology is especially attractive for WHR in RHFs due to some reasons [8]:
• The fumes produced by methane combustion are almost free of dust and they
need no special treatment and filtration.
• The recoverable exhaust heat output of RHFs is ideal for ORC cycles.
• The rolling mills operate typically on continuous cycles over 24 hours, and do not
require frequent stops for maintenance.
The exhausted gases are clean enough to allow the direct exchange with the organic
working fluid: avoiding the installation of a secondary loop decreases investment costs. Di-
rect heat exchange between the source and the organic working fluid in the ORC is only
possible if the heat source is not corrosive and its temperature peaks do not exceed working
fluid limits [8].
The ORC unit generates electricity that is self-consumed by the industrial plant [8]. At
the end of the heating process, the waste gas which contain high heat energy are being
channel out to the chimney stack before releasing to the atmosphere, ranging between 300-
425°C [55]. This is a simpler solution with a more compact layout, less components and a
greater efficiency (the organic fluid evaporates at a higher temperature) [56].
This system has already been applied in NatSteel (Singapore), Turboden’s only ORC
installation for electric power production from waste heat in a rolling mill. This first plant can
be replicated for all hot rolling mills, both those located at the end of integrated steel plants
(blast furnace and converter shop) and those at the exit of EAFs [8].

2.6 Cement industry

2.6.1 Description of the production process


In 2019, Watts from The Guardian stated that “After water, concrete is the most widely
used substance on Earth. If the cement industry were a country, it would be the third largest
carbon dioxide emitter in the world with up to 2.8 billion tonnes, surpassed only by China
and the US“.
Theoretical background 29

Cement is mixed with an aggregate such as sand or gravel and water to form concrete.
Over three tons of concrete are produced each year per person for the entire global popu-
lation. Ground raw materials (mainly limestone and clay or shale) are burnt finely to obtain
clinker as an intermediate product. A large rotary kiln (cylindrical furnace) is used for fusing
these materials into new mineralogical phases when heated to around 1450°C. This burned
product is called clinker. Then, clinker is ground into a fine powder with small quantities of
gypsum and other components to become cement. Ordinary Portland Cement (OPC) gen-
erally contains at least 90% clinker [57].

Figure 2.22. Schematic representation of the cement manufacturing process from quarry to
dispatch [58]

Figure 2.22 depicts the cement production process from the raw material extraction to
the dispatch. The waste heat is usually recovered from the rotary kiln and, for that reason,
this work is focused on this part of the production process.
Cement manufacturing is energy-intensive consuming around 4 GJ per tonne [59]. The
average specific energy consumption is about 2.95 GJ per tonne of cement produced for
well-equipped advanced kilns. However, in some countries, the consumption exceeds 5
GJ/t [60]. A typical European dry type clinker kiln requires approximately 3.3 GJ/t-clinker of
input thermal energy [57]. These values can vary greatly depending on the age and config-
uration of clinker kilns.
The kilns require thermal energy. This is provided mainly by fossil fuels (mainly hard
coal, lignite, to a lesser extend oil and gas). Combustible waste material (mainly in mature
countries) or biomass derived fuels from wood processing and agricultural activities (e.g.
crop residues) can be also used [57].
Due to this fossil fuels consumption, cement manufacture releases a great amount of
CO2. The CO2 emissions. Normally, 40% of direct CO2 emissions for OPC come from com-
busting fuel required to drive the reactions necessary to make clinker, 60% comes from the
30 2.6 Cement industry

de-carbonation of limestone to produce CaO [57]. Indirect emissions from electric power
consumption and internal transport contribute another 10% to overall CO 2 emissions.
Clinker burning kiln system
The raw material (lime and clay) is burned to clinker using fossil fuel fired rotating kilns.
The raw material is first preheated. Then, it passes through the rotary kiln. The hot clinker
leaving the kiln is cooled down from 1450°C to about 100°C by a grate type air cooler. Half
of the cooling air is used as pre-heated combustion air for the kiln burner. The other half of
the cooling air is vented [57].

Figure 2.23. Clinker production structure [61]

Rotary kiln process types


Rotary kilns are either dry-process or wet-process, depending on how the raw materials
are prepared.
• Wet process
The originary rotary kilns were wet-process kilns. Here, raw materials are fed into
the kiln as slurry with a moisture content of 30-40%. The wet process has some
advantages. Wet grinding of hard minerals is normally much more efficient than
dry grinding. When slurry is dried in the kiln, it forms a granular crumble that is
perfect for subsequent heating. In the dry process it is very difficult to keep the
fine powder rawmix in the kiln, as the fast blowing combustion gases tend to blow
it back out [62]. However, wet process has much higher energy requirements due
to the amount of water that must be evaporated before calcination can take place
(nearly 100% more kiln thermal energy compared to an efficient dry kiln).
• Dry process
In modern works, the blended raw material enters the kiln via the pre-heater
tower. Here, hot gases from the kiln are used to heat the raw material. The dry
process is thermally much more efficient than the wet process. Firstly, this is be-
cause there is little or no water that has to be evaporated. Secondly, the process
of transferring heat is much more efficient in a dry process kiln because the meal
Theoretical background 31

particles have a very high surface area in relation to their size and because of
the large difference in temperature between the hot gas and the cooler meal.
There are three kinds of dry-process kilns in operation: long dry kilns without pre-
heaters (LD), suspension preheater (SP) kilns, and preheater/precalciner or new
suspension preheater (NSP) kilns.

In SP and NSP kilns, the pyro-processing starts in the preheater sections (before mate-
rials enter the rotary kiln). The preheater consists of a series of vertical cyclones. As the
raw material is passed down through these cyclones it comes into contact with hot kiln
exhaust gases moving in the opposite direction and thus, heat is transferred from the gas
to material. This preheats and partially calcines the material before it enters the kiln, which
makes the necessary chemical reactions occur more quickly and efficiently. A kiln can have
from three to six stages of cyclones (depending on the moisture content of the raw material)
with increasing heat recovery with each extra stage. As a result, SP and NSP kilns tend to
have higher production capacities and greater fuel efficiency compared to other types of
cement kilns [63].
While the energy performance of kilns has remained relatively consistent for the last
twenty years, overall energy intensity and CO2 emissions intensity of cement production
worldwide have declined. This is due to the phase out of wet-process and inefficient long
dry-process kilns and new capacity additions [63].
After the clinker is formed in the rotary kiln, it is cooled rapidly to minimize glass phase
formation and ensure maximum yield of alite (tricalcium silicate) formation, an important
component for cement hardening properties [63].The main technologies are a grate cooler
or a tube (planetary cooler). In the grate cooler, the clinker is transported over a grate
through which air flows perpendicular to the clinker flow. In the planetary cooler (a series of
tubes surrounding the discharge end of the rotary kiln), the clinker is cooled in a counter-
current air stream [63].
The cooling air is partially used as secondary combustion air for the kiln. After cooling,
clinker can be stored in domes, silos or bins and then transported to the finish mill.
Figure 2.24 shows the typical energy output flows of a modern dry type clinker kiln [57].
Typically, the clinker coolers release large amounts of heated air at 250-350°C into the
atmosphere (the part which is not used as secondary combustion air for the kiln). The kiln
off-gas from the preheaters at 290-380°C is typically used to dry material in the raw mill
and/or the coal mill and then sent to electrostatic precipitators or bag filter houses to remove
dust, before finally being vented to the atmosphere [57].
The energy cost is one of the main of cement manufacturing. Therefore, the cement
industry permanently seeks for technologies for improving the energy efficiency of the burn-
ing process.
The remaining waste heat from the preheater and clinker cooler exhausts can be recov-
ered and used for two different things: either to provide low temperature heating needs in
the plant or to generate power to offset a portion of power purchased from the grid, or cap-
tive power generated by fuel consumption at the site [57]. Typically, cement plants hardly
require low temperature heating, so most of WHR projects have been focusing on power
generation.
32 2.6 Cement industry

Figure 2.24. Energy flow diagram of a cement plant clinker burning kiln [64]

2.6.2 Waste heat recovery with ORC in the cement industry:


state-of-the-art
In a typical WHR system installation for a cement plant the following elements can be
identified [64] and seen in Figure 2.25:
• Two heat-exchangers designed for high dust load (one for the preheater and the
other for the cooler).
• Two intermediate heat transfer loops (either thermal oil or pressured water).
• ORC cycle with recuperator (hydrocarbon, siloxane or refrigerator fluid).
• Air or water (evaporation) cooled condenser.
The amount of waste heat available for recovery from a kiln depends on several factors
including the following [65]:
• Moisture content of the raw material feed: it determines heat requirement for the
kiln and the amount of preheater exhaust needed for drying.
• Amount of excess air in the kiln and air infiltration.
• Number and efficiency of preheater/precalciner stages: the higher the number of
stages, the higher the overall thermal energy efficiency and the lower the WHR
potential.
• Configuration of the clinker cooler system.
Theoretical background 33

Figure 2.25. Typical WHR with ORC installation in cement kilns [64]

The first ORC used in a cement plant is located in Lengfurt (Germany) and belongs to
Heidelberger Zement AG. A 1.5 MW ORC recovers the heat available from the grate cooler
and generates heat on a continuous basis (ORC availability of 97% of the operation time of
the cement kiln). It has been operating for more than 10 years [63].
It is to note that for the kilns using a wet process, it is not convenient to install ORC
systems, due to technical problems for the heat exchanges. However, according to
Schorcht et al. (2013) [66], wet kilns will be replaced by dry ones, thus it is possible to
consider them as dry kilns, in a future perspective.

2.7 Glass industry

2.7.1 Description of the production process


There a different types of glass depending on their chemical composition, their produc-
tion method or their processing behavior. If the classification is made regarding their chem-
ical composition, there are three types: soda-lime glass, lead glass and borosilicate glass.
However, soda-lime glass is the one which is produced in the largest quantities by far [67].
This type, which contains 71-75% SiO2, 12-16% Na2O, 10-15% CaO and small quantities
of other substances such as dyes, is used to make containers and sheet glass products
[67]. These substances are added to the sand (SiO2) to reduce its melting temperature
(which is higher than 2000oC) and to improve the material properties. These two types of
glass manufacture are considered for the waste heat recovery in this work: float and con-
tainer glass. These types are, by far, the most produced [67]. The production processes are
now shortly described. As observed, the two processes are quite similar up to the shaping
part.
34 2.7 Glass industry

Float glass production


In the float glass production process, the ingredients are first mixed together and also
with cullet (crushed recycled glass) and then heated in a furnace to around 1600 oC to form
molten glass. Then, it undergoes a refining process. The molten glass is then fed onto the
top of a molten tin bath. A flat glass ribbon of uniform thickness is produced by flowing
molten glass on the tin bath under controlled heating. After the tin bath, the glass is gradu-
ally cooled down, controlled and cut into sheets.

Figure 2.26. Schematic diagram of the production process of float glass [68]

Container glass production


Once the materials required (including chemicals for dying and cullet) are mixed, the
batch is fed continuously and steadily into the furnace, where it is heated by electrical, gas
or oil-burning systems to a temperature of 1550-1600oC. In order to save energy and reduce
the environmental impact, the waste gases are used to preheat the air inside the furnace
[69]. When the melted glass comes out of the furnace, it is distributed to the forehearth,
which ensures that the molten glass is heated evenly throughout. This process is called
glass conditioning. After that, the molten glass will be turned into shapes that can be molded
into containers [69].

2.7.2 Waste heat recovery with ORC in the glass industry: state-
of-the-art
Glass is one of the most sustainable existing packages, since it is made by natural ele-
ments, it is largely reusable and 100% recyclable. However, the production is high energy
demanding. Furnaces operate at temperatures of over 1500 oC, 24 hours a day and 7 days
a week, making glass production a significant source of waste heat [70].
As it can be seen in Table 6.4, there are several projects in operation around the world
which recover waste heat from either float glass or container glass melting furnaces with an
ORC unit for electricity production.
The Düzcecam Glass Plant in Düzce (Turkey) has the largest ORC unit for glass, in-
stalled by the group leader in ORC turbogenerators for WHR, Turboden. The Turboden 65
HRS ORC unit, designed for 6.2 MW nominal capacity, converts the off-gas waste heat
from the two float glass production lines into electric power [71].
Theoretical background 35

The heat recovery system configuration includes thermal oil boilers (one for each line)
with a heat carrier circuit to convey the heat from the exhaust gas to the ORC unit. The two
waste heat boilers are installed at the bottom of the float glass production lines, in parallel
to the existing quenching tower system [71].
In Figure 2.27, the typical layout of an ORC waste heat recovery system from a glass
furnace is presented. It counts with an intermediate thermal oil heat exchanger.

Figure 2.27. Heat recovery system with ORC technology in the glass industry typical layout
[8]
36 2.7 Glass industry
Methodology 37

3 Methodology

3.1 Overview of the method

Figure 3.1 sums up the steps followed in this work in order to estimate the technical and
economic potential of the implementation of ORC WHR technology in the most important
EII in Germany.
Firstly, the waste heat leaving each process from each plant is estimated. Then, taking
into account the recovery efficiency, the net electric power output is estimated. With that,
the technical potential is determined. For the economic potential, the ORC costs must be
considered.

Waste heat Net electric


Technical
ηORC power output
𝑄̇ 𝑇→𝑇 potential
𝑚𝑖𝑛 ORC

Economic
Costs
potential

Figure 3.1. General scheme of the steps followed in this work

3.2 Description of the method


In this thesis, a combination of the bottom-up and top-down methods is used. The bot-
tom-up method is used for the obtention of the SWH data from different authors in order to
estimate a SWH associated to each process. The top-down method is applied when this
SWH value common to every process is multiplied by the size of each plant (in tonnes of
produced goods) to obtain the total waste heat leaving each plant. Also, a “medium precise”
method is used [29].
The processes in the different EII are studied and analyzed in order to identify the ones
which are more interesting for the WHR. Therefore, for each industry, a high energy con-
suming process is chosen for calculating the waste heat potential. The technical feasibility
based on the already installed and in operation ORCs in other WHR projects worldwide is
also taken into account for the selection of the process subject to heat recovery.
In the steel industry, for example, the EAF, the BOF and the RHF in the rolling mill have
been considered. In the cement industry, the rotary kiln preheater and the clinker cooler are
the processes from where waste heat can be and is being recovered. In the glass industry,
the glass melting furnace for container and flat glass.
38 3.2 Description of the method

The waste heat and then the electric power that can be produced with the ORC are
estimated for each one of the plants from the different industries studied for a worst-case
scenario, an average scenario and a best-case scenario. This is required because the data
on available waste heat from the processes are considerably variable depending on local
factors that cannot be generalized.
The worst-case scenario (WCS) is the one with the lowest SWH value found in different
literature sources. The best-case scenario (BCS) uses the highest waste heat value ob-
tained among publications of various authors. The average scenario (AS) takes an average
value between the lowest and the highest SWH. In this way, the economic potential can be
determined based on the aggregated power capacity from the sites in which the LCOE is
lower or equal to electricity prices and thus economically feasible.
An overview of the method with the steps to be followed is presented here.

1. Selection of EII in Germany.


2. Identification of the high energy consuming processes in each industry which can
be used for WHR.
3. Identification of plants from each industry. Information gathering on the process, ca-
pacity and ORC WHR units already installed.
Theoretical potential
4. Estimation of the SWH that can be obtained for each process.
a. Research on the processes waste heat and energy consumption.
b. Calculation of the worst-, average- and best-case scenario‘s SWH for each
process using these data.
5. Calculation of the total theoretical waste heat per year from each plant for the three
scenarios, multiplying the SWH by the plant production per year.
Technical potential
6. Calculation of the estimated electric power generation with ORC potential for each
plant taking into account technical limitations such as the ORC net electric efficiency,
the utilization rate and the annual availability of the ORC unit.
Economic potential
7. Calculation of the SIC in function of the size of the ORC with Bianchi et al. correlation
(2019) and the LCOE associated to each ORC unit.
8. Elaboration of graphs which relate the aggregated electric energy produced by every
ORC that can be potentially installed in Germany versus the LCOE from those
ORCs.
9. Determine the electricity price range for the industrial sector in Germany.
10. Elaboration of a sensitivity analysis to determine how many ORC units are econom-
ically feasible installing.
Methodology 39

3.3 Definition of efficiencies for net electric power estimation

3.3.1 Utilization rate (𝜂𝑢 )


The output temperature (the temperature of the flue gas when it exits the heat recovery
system) is considered to be 150°C, which ensures avoiding the condensation of the flue
gas and therefore corrosion. The utilization rate is a function to the temperature of the heat
source leaving the recovery system, and is defined as:
𝑇ℎ𝑠,𝑖𝑛 − 𝑇ℎ𝑠,𝑜𝑢𝑡
𝜂𝑢 = ( 3.1)
𝑇ℎ𝑠,𝑖𝑛 − 𝑇𝑟𝑒𝑓

This rate is needed because the theoretical waste heat is often referred to the ambient
temperature. This ambient temperature can vary depending on literature. The heat that can
be obtained with a cooling to 15°C is higher than the one obtained for a cooling to 150°C,
as can be seen in Figure 3.2.

Figure 3.2. Transferred heat depending on temperature difference


Table 3.1. Assumed utilization rate of the heat source for different processes

Inlet Minimal
Waste heat Utilization rate
temperature of temperature of
source (-)
heat source (°𝑪) heat source (°𝑪)

BOF off-gas 1200 150 0.89

EAF off-gas 1200 150 0.89

RHF off-gas 400 150 0.65

Clinker 300 100 0.70


cooling air

Clinker preheater 390 150 0.64


(TPD 1000-2000)

Clinker preheater 300 150 0.53


(TPD 2000-8000)

Glass furnace 400 200 0.52


40 3.3 Definition of efficiencies for net electric power estimation

Table 3.1 shows the assumed utilization rates for the steel, cement and glass production.
For the clinker preheater, the inlet temperature is different depending on the size.

3.3.2 Net electric efficiency of the ORC unit (𝜂𝑒𝑙,𝑛𝑒𝑡,𝑂𝑅𝐶 )


The net electrical efficiency of the ORC units is retrieved from the units of the company
Turboden, which is the leading company in WHR with ORC from industrial processes. All
considered units are equipped with an intermediate oil loop. The net electrical efficiency is
in general a function of the heat source temperature and size of the plant.
Because of system standardization and limitations in the temperature of the intermediate
loop, it can be seen in Figure 3.3 that the nominal net efficiency is for all the Turboden units
at 16-21%. In real operation, especially for air-cooled units, this value drops by some per-
cent point.

Figure 3.3Turboden’s Standard HR (Heat Recovery) ORC units [72]

3.3.3 ORC unit availability


The ORC availability is usually of a 95% with respect to the number of hours of waste
heat flux availability for power production. This is high, due to the fact that the maintenance
is done mainly when the waste heat is not available.
Methodology 41

3.4 Specific waste heat estimation for the different industrial pro-
cesses

3.4.1 Steel industry


Basic oxygen furnace
The BOF specific off-gas flow and the heating value of the off-gas are based on the
energy content of each raw material and process product with respect to their chemical
analysis and temperature, and also from the type of recovery, volume and weight of the
furnace, time between “heats” or steel temperature. For this reason, it is relatively variable
and its generalization leads to deviations with respect to the individual plants.
However, no accurate data of this kind is available for each German steel plant. There-
fore, only an estimation can be done based on literature from different sources.
Data from various sources have been collected in order to estimate the WHR potential
per tonne of steel produced with the BOF (Table 3.2). These data are the sensible heat, the
chemical heat, the temperature range of the off-gas and the reference temperature.
The total available waste heat includes both the sensible and chemical heat potential. In
BOFs, the chemical heat is recoverable from the burning of the CO in a post-combustion
chamber just after the exit of the flue gas from the furnace. It has a higher potential than the
sensible heat.

Electric arc furnace


The values of the EAF flue gas energy content are highly dependent on the individual
operation and vary significantly from one facility to another. Factors such as raw material
composition, power input rates and operating processes such as scrap preheating can alter
the balance [73].
The temperature of the off-gas from the EAF is 1200-1800oC. The off-gas can be from
less than 10% of the total energy to more than the 40% depending on the plant (Kirschen
et al., 2009). The relation between sensible and chemical heat of non-combusted CO and
hydrogen also varies from one author to another: 15.5% as sensible heat and 16.9% as
chemical energy as by [74] and 16.7% as sensible and 21.4% as chemical [75].
A general energy balance is presented in Figure 3.4.
42 3.4 Specific waste heat estimation for the different industrial processes

Figure 3.4. Energy balance in an EAF [76]

Around a 70% of the energy losses are in the off-gas. Focusing on this off-gas energy
can therefore considerably increase the energy efficiency of EAFs. Also, as seen previously,
there is approximately a 50/50 split between the sensible (thermal) and chemical energy in
the off-gas.
Contrarily to expected, there is no relation between the size of the plant represented by
the tapping weight of the EAF and the energy consumption (Figure 3.5). Due to this, an
estimated SWH can be used for all the sites for each scenario (Table 3.2).

1000
ENERGY CONSUMPTION [KWH/T]

900
800
700
600
500
400
300
200
100
0
0 20 40 60 80 100 120 140 160
TAPPING WEIGHT [T]

Figure 3.5. Tapping weight vs. energy consumption [76]


Methodology 43

Reheating furnace
For a typical RHF, the energy in the exhaust gas is slightly less than 30% of the heat
from the combustion of natural gas, which can be estimated, on average, about 430 kWh/t
(1.548 GJ/t) [8].
For the case study at Sidenor Basauri Works in the context of the project funded by the
European Commission within the H2020 Programme (2014-2020) „WHR for Power Valori-
sation with Organic Rankine Cycle Technology in EII” [57] this value is 405 kWh/t (1.458
GJ/t). This input energy depends on the charging temperature.
If preheating of the combustion air is considered, like in Basauri plant [57], which is highly
recommendable, only a part of the waste heat would be available for electricity production.
The rest of the heat would go into a recuperator. A Sankey diagram on this plant can be
found in Figure 3.6.
According to different sources, the total waste heat from the flue gas that can be recov-
ered by an ORC system from a RHF has been calculated based on the sensible energy
(Table 3.2).

Figure 3.6. The Sankey Diagram of energy balance for a typical RHF in case study [57]
44 3.4 Specific waste heat estimation for the different industrial processes

Table 3.2. Waste heat available from steel production, referred to 15°C [77], [78], [79], [80],
[81], [82], [83], [84], [85], [86], [87]

Waste heat available


Waste Temperature
heat range Characteristics Total (𝑮𝑱/𝒕)
Sensible Chemical
source (°𝑪)
(𝑮𝑱/𝒕) (𝑮𝑱/𝒕)
WCS AS BCS

BF off- Contain combustibles,


120-320 0.48-0.54 3.40-6.00 3.88 5.21 6.54
gas particulates, etc.

BOF Contain combustibles,


1200-1500 0.15-0.18 0.84-0.86 0.99 1.02 1.04
off-gas particulates, etc.

EAF Contain combustibles,


1200-1800 0.40-0.51 0.44-0.57 0.44 0.73 1.02
off-gas particulates, etc.

RHF
400-550 Clean gas 0.15-0.30 0.00 0.15 0.23 0.30
off-gas

3.4.2 Cement industry


Approximately 20-30% of the required electricity for the cement production process can
be generated from the waste heat [65]. Around 70% can be obtained from the preheater
and 30% from the cooler [64].
The number of preheater stages in a cement plant has significant influence on the ther-
mal energy consumption and waste heat recovery potential. The higher the number of
stages, the higher the thermal energy efficiency of the kiln and the lower the potential for
waste heat recovery. Selection of the number of preheater stages is based several factors
such as cooler efficiency, restrictions on preheater tower height, or heat requirements for
the mill itself. Preheater exhaust temperatures range from 390°C for small kilns with four
preheater stages, to below 300°C for large kilns with six preheater stages [65].
The clinker preheaters without precalciner have 4 stages and a kiln capacity of 1000-
2000 TPD. The preheaters with precalciner can have from 4 to 6 stages and a capacity of
2000-8000 TPD [65].
The clinker cooler design also impacts waste heat availability. Typically, state-of-the-art
coolers are grate coolers, which have various stages of development. Exhaust air temper-
atures from the clinker cooler range from 250 to 330°C depending on cooler configuration
and recuperation efficiency (the higher the generation, the higher the efficiency and thus
the smaller the waste heat availability). First, second and third generation grate coolers
have been considered (Table 3.3).
Methodology 45

Table 3.3. Waste heat available from cement production, referred to 15°C [83], [78], [65]

Waste heat available


Temperature (𝑮𝑱/𝒕)
Waste heat source Characteristics
range (°𝑪)
WCS AS BCS

Clinker cooling air 250-350 Clean gas 0.30 0.41 0. 52

Clinker preheater Contain combustibles, partic-


390 0.92 0.92 0.92
(TPD 1000-2000) ulates, etc.

Contain particulates, etc.


Clinker preheater Relatively easy to handle
200-400 0.60 0.71 0.81
(TPD 2000-8000)

3.4.3 Glass industry


For the waste heat that can be recovered from the glass manufacture, both flat and con-
tainer glass are considered. Different types of glass melting furnace are included in the
analysis to determine the waste heat available in the worst-, average and best-case sce-
narios. These types are, for instance, end fired glass furnace, furnace with regenerator or
furnaces with electric boosting equipment.
The temperature range of the waste heat that comes out of the glass melting furnace is
400-500oC (Table 3.4).
Table 3.4. Waste heat available from glass production, referred to 15°C [78], [8], [83]

Waste heat available


Temperature (𝑮𝑱/𝒕)
Waste heat source Characteristics
range (°𝑪)
WCS ACS BCS

Contain particulates, con-


Glass melting furnace 400-500 1.05 1.85 2.64
densable vapors, etc.
46 3.5 Plants in Germany

3.5 Plants in Germany

3.5.1 Steel
The main steel plants in Germany in operation today are obtained from [88] and are
shown in Figure 3.7. Data from [89] are used to determine the capacity of each plant. A total
of 9 BOF plants, 31 EAFs and 31 RHFs for hot rolling mills is considered. The total produc-
tion amounted 43.3 Mio. tons of steel in 2017 [90]. From the plant capacity and the actual
total production, an average load factor of 82.2 %, 82.9 % and 89.4 % for respectively BOF,
EAF and RHF is found.

Integrated steel mills Electric arc furnaces


(Blast furnace, steelmaking and
rolling mills)
Year: 2014

Figure 3.7. Location of the main steel plants in Germany [88]

Figure 3.8 shows the main steelmakers in Germany and the capacity installed for the
different production routes.
Methodology 47

10
9
8

Capacity [Mt]
7
6
5
4
3
2
1
0

EAF BOF

Figure 3.8. Biggest steelmakers in Germany and their production routes [88]

3.5.2 Cement
The cement plants in Germany can be obtained from the Verein Deutscher Zementwerke
(VDZ) which means Association of German Cement Plants. The members of VDZ operate
46 of the 53 cement plants in Germany.

Figure 3.9. Location of the German cement plants which are part of the VDZ [91]
48 3.5 Plants in Germany

It is to note that 35 out of 46 (76%) of the cement plants in Germany include clinker
production and, thus, a kiln from where waste heat can be recovered. The rest are mainly
grinding mills. Out of these, only 76 % actually produce clinker (and have relevant waste
heat available), whereas the remaining ones are mainly grinding mills. The production of
these plants reached 31 Mio. tonnes of cement in 2015 [91].
Figure 3.10 shows the main cement companies in Germany and the capacity of the
plants with a kiln that there exist in this country.
Brunke [92] reported the capacity and production of each plant in 2013, where a load
factor of 85.9 % is estimated. These data (34 plants are considered) are used for the present
analysis.

10
9
8
Capacity [Mt/a]

7
6
5
4
3
2
1
0

Figure 3.10. Biggest cement producers in Germany and capacity of plants with a kiln [91]
Results 49

4 Results

In this chapter, the results for the estimation on the theoretical, technical and economic
WHR with ORC potential for the energy-intensive steel, cement and glass German indus-
tries are presented and analyzed.

4.1 Theoretical potential

4.1.1 Results
The theoretical waste heat (energy in the off-gas) available from the steel, cement and
glass industries in Germany has been estimated based on the method described in this
work.
The results for WCS, AS and BCS are shown in Table 4.1 for each industry. The total
theoretical potential referred to 15 °C is between for 11.5-15.6 TWh/a for steel, 7.9-9.3
TWh/a for cement and 1.7-4.3 TWh/a for glass.

Table 4.1. Theoretical potential of waste heat available from the steel, cement and glass in-
dustries in Germany at 15°C

Waste heat available at 15°C (𝑻𝑾𝒉/𝒂)


Waste heat source
WCS AS BCS

BOF off-gas 8.33 8.58 8.75

EAF off-gas 1.59 2.65 3.70

RHF off-gas 1.56 2.40 3.13

Steel (total) 11.48 13.63 15.58

Cement 7.87 8.56 9.27

Glass 1.73 3.04 4.34

4.1.2 Comparison to previous studies


Despite the fact that very few studies exist that estimate a value for the waste heat
potential for the steel, cement and glass industries in Germany, two have been found and
will be used to compare the results obtained in this work with the results of these previous
works which used different methods. The methods are described shortly in section 2.4.3.
50 4.2 Technical potential

Papapetrou et al. (2018) [2] estimates the waste heat potential in Germany from iron and
steel industry in approximately 45 TWh/a.
Miró (2016) [15] gives a value of the industrial heat potential for the cement and glass
production in Germany based on the CO2 production from the E-PRTR database. These
estimations are closer to the lower bound for cement (7.8 TWh/a from Miró vs. 7.9 TWh/a
from this work), whereas the estimation for glass is much lower than the estimated value in
this work (0.59 vs. 1.73 TWh/a). The main reason for this is the lower amount of glass
manufacturing plants included in the E-PRTR database and considered by Miró et al. [15]
(7 vs. 71 plants).

4.2 Technical potential


The technical potential can be estimated as the net electric energy that can be obtained
from the waste heat source with an ORC unit. It takes into account for the limited heat
source utilization and the net electrical efficiency of the ORC unit. Also, the annual availa-
bility of the ORC unit is multiplied to the total conversion efficiency to find the annual elec-
tricity produced.
The total net electric energy that can be obtained from the waste heat for the steel, ce-
ment and glass industries with ORC systems is summed up in Table 4.2. An ORC efficiency
of 15% is taken for a more conservative case and 19% for the more optimistic scenario. In
Table 4.3 the cumulative power output from the technical potential is shown for each waste
heat source, ORC efficiency and scenario.

Table 4.2. Technical potential of waste heat potential available from the steel, cement and
glass industries in Germany with ORC availability at 95%

Technical potential (𝑻𝑾𝒉/𝒂)

ORC efficiency ORC efficiency


Waste heat source
15% 19%

WCS AS BCS WCS AS BCS

BOF off-gas 1.05 1.08 1.11 1.33 1.37 1.40

EAF off-gas 0.20 0.33 0.47 0.26 0.42 0.59

RHF off-gas 0.14 0.22 0.29 0.18 0.28 0.37

Steel (total) 1.39 1.63 1.87 1.77 2.07 2.36

Cement 0.70 0.86 1.02 0.88 1.09 1.30

Glass 0.13 0.23 0.32 0.16 0.29 0.41


Results 51

Table 4.3. Cumulative capacity of technical potential from energy-intensive industry with
ORC availability at 95%

ORC cumulative power output (𝑴𝑾𝒆 )

ORC efficiency ORC efficiency


Waste heat source
15 % 19 %

WCS AS BCS WCS AS BCS

BOF off-gas 120.1 123.7 126.2 152.1 156.7 159.8

EAF off-gas 23.0 38.1 53.3 29.1 48.3 67.5

RHF off-gas 16.5 25.3 33.0 20.9 32.1 41.9

Steel (total) 159.6 187.1 212.5 202.1 237.1 269.2

Cement 79.6 98.3 116.9 100.8 124.5 148.1

Glass 14.6 25.7 36.7 18.5 32.5 46.4

4.2.1 Comparison of estimated ORC power to the power from


ORCs in operation
For the steel industry in Germany, only Elbe-Stahlwerke Feralpi in Riesa counts with an
ORC system for recovering the heat from the EAF off-gas. For the cement industry, Lengfurt
plant is the only one with an ORC.
It is of great interest to compare the actual power from these ORCs and the estimated
power.
Feralpi’s ORC produces 2.7 MWe, but one third of the steam stored in the steam drum is
sold to another industrial process (Goodyear Dunlop tire plant, which is nearby) through the
district heating grid [93]. Thus, only two thirds of the technically available heat are used for
electricity generation. That means that the potential electric power that could be produced
is 4.05 MWe.
The Heidelberg Cement Group’s plant in Lengfurt has a 1.5 MW e ORC and uses the
waste heat only from the grate clinker cooler.
As it can be observed in Table 4.4, the nominal electric power from the ORCs installed
at Feralpi’s EAF and Lengfurt’s clinker cooler is, for both cases, within the range of possible
power values for the worst-, average- and best-case scenarios. For the EAF at Feralpi, the
potential electric power that could be obtained is slightly higher than the BCS with a 19% of
ORC efficiency.
52 4.3 Economic potential

Table 4.4. Estimated and actual ORC power for Feralpi and Lengfurt plants

Real nominal Potential


Estimated power [MW]
power [MW] power [MW]

Site ORC efficiency ORC efficiency


15% 19%

WCS AS BCS WCS AS BCS

Feralpi 1.32 2.18 3.05 1.67 2.76 3.86 2.70 4.05

Lengfurt 0.72 0.97 1.23 0.91 1.23 1.56 1.50 -

4.2.2 Comparison of the results to other studies


The results of the technical potential estimated in this work are compared to the study
by Campana et al. (2013) [18], which is within the H-REII Demo project for the research on
WHR in EII.
This study considers for the steel industry only EAFs and RHFs. Thus, for the compari-
son, only the EAFs together with the RHFs will be considered. The cumulative ORC power
by Campana et al. [18] is 74.0 MWe for EAF and 82.2 MWe for RHF. These values are
significantly larger than the estimated 23.0-67.5 MWe for EAF and 16.5-41.9 MWe for RHF
from Table 4.3. The main difference is due to the difference in the steel production consid-
ered in this work and the one considered by Campana et al. (16.7 Mt of steel for EAF and
50.8 Mt for RHF instead of the 13.0 and 37.5 Mt of steel used in this work).
For cement, Campana et al. [18] estimated a cumulative ORC power of 70.3 MWe,
whereas a larger potential is found in the present analysis, in the range of 79.6-148.1 MWe.
The difference is caused by a different specific ORC power per unit production of cement.
For glass, only the total EU-27 capacity of flat glass is considered by Campana et al.
[18], therefore a direct comparison for Germany is not possible.

4.3 Economic potential


For the evaluation of the economic potential, a method inspired in [33] is used. This
method consists on the elaboration of a supply-cost curve for electricity generation with
ORC from waste heat. On the y-axis, the accumulated electrical energy is presented versus
the Levelized Cost of Energy (LCOE) of each ORC for 2018 that is added until the technical
potential is reached.
Once the curve is presented, the economic feasibility will be determined by the total
electrical energy of those ORC units that have a lower LCOE than the Levelized Cost of the
Electricity for the industry in Germany in 2018. As seen previously in this thesis, there is a
minimum and maximum electricity price for EII depending on the exemptions from fees,
Results 53

taxes and levies which is different for each site depending on its size. For energy-intensive
industries, the price in 2018 was 5.1-17.0 ct/kWhe in Germany [5]. In the graphs presented
in this work, the minimum, average and maximum electricity prices for EII appear.

Figure 4.1 shows the results for steel BOF, EAF and RHF at the most conservative ORC
efficiency 𝜂𝑒𝑙,𝑛𝑒𝑡,𝑂𝑅𝐶 = 15 %, an ORC availability of 95%, the interest rate at 4% and 10
years of amortization. In Appendix F, other figures are shown varying some of the parame-
ters (𝜂𝑒𝑙,𝑛𝑒𝑡,𝑂𝑅𝐶 = 19 %, interest rate at 6%, 5 years of amortization) in order to create dif-
ferent combinations and analyze different possible scenarios. However, in the case of the
amortization years, for instance, 10 years is more realistic than 5 years.
If these figures for steel are analyzed (Figure 4.1) it can be seen that the LCOE for BOF
plants is significantly lower than the lowest electricity price (5.1 ct/kWh e). This means that
installing an ORC unit is economic for every BOF plant and the economic potential corre-
sponds to the technical potential (1052-1105 GWhe/a).
Most of the ORC units potentially installed at plants with an EAF have a LCOE below the
lowest electricity price. The economic potential is 164-455 GWhe/a depending on the sce-
nario considered and it is relatively close to the technical potential.
For the plants with RHF, the feasibility is still positive but less favorable than BOF and
EAF. The difference between the technical and the economic potential is greater for the
lowest electricity price than in the case of the EAF and the BOF. In the case of the RHF,
this potential is between 85 and 222 GWhe/a. Also, in Figure 4.1, it can be observed that,
for all BOF, EAF and RHF, the economic potential for the average and highest electricity
prices virtually matches the technical one.
For cement (Figure 4.2), the economic potential is almost the same as the technical for
every scenario in the whole range of electricity prices, even for the lowest. In this case, the
economic potential is 685-1025 GWhe/a. In the case of glass, it is interesting to note that in
the worst-case scenario, the installation of an ORC unit for WHR would not be economic
for any plant for the lowest electricity price of 5.1 ct/kWhe. In the other two scenarios, the
economic potential for the lowest price is far from the technical, and it grows progressively
until they are virtually equal for an average electricity price of 8.84 ct/kWh e.
The economic potentials for the lowest electricity price, i=4%, n=10 years, ORC availa-
bility of 95% and an ORC efficiency of 15% and 19% are presented in Table 4.5. In Table
4.6, the corresponding cumulative power output is shown.
54 4.3 Economic potential

(a) (b)

(c)

Figure 4.1. Economic potential for steel production with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 years:
(a) BOF; (b) EAF and (c) RHF
Results 55

(a) (b)
Figure 4.2. Economic potential for production with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 years: (a)
cement and (b) glass

Table 4.5. Economic potential of waste heat available from energy-intensive industry with
ORC availability at 95%, interest rate at 4% and 10 years of amortization

Economic potential (G𝑾𝒉/𝒂) referred to 5.1 ct/kWhe

ORC efficiency ORC efficiency


Waste heat source
15 % 19 %

WCS AS BCS WCS AS BCS

BOF off-gas 1052.1 1084.0 1105.2 1332.6 1373.0 1399.9

EAF off-gas 163.8 319.0 454.7 231.4 412.2 575.9

RHF off-gas 84.8 156.4 221.9 116.1 215.5 310.9

Steel (total) 1300.7 1559.4 1781.8 1680.1 2000.7 2286.7

Cement 684.9 854.7 1024.5 883.0 1090.5 1297.7

Glass 0.0 21.3 120.5 0.0 94.5 227.8


56 4.4 Sensitivity analysis of the results

Table 4.6. Cumulative capacity of economic potential from energy-intensive industry with
ORC availability at 95%, interest rate at 4% and 10 years

ORC cumulative power output of economic potential (𝑴𝑾𝒆 )

ORC efficiency ORC efficiency


Waste heat source
15% 19%

WCS AS BCS WCS AS BCS

BOF off-gas 120.1 123.7 126.2 152.1 156.7 159.8

EAF off-gas 18.7 36.4 51.9 26.4 47.1 65.7

RHF off-gas 9.7 17.9 25.3 13.3 24.6 35.5

Steel (total) 148.5 178.0 203.4 191.8 228.4 261.0

Cement 78.2 97.6 116.9 100.0 124.5 148.1

Glass 0.0 2.4 13.8 0.0 10.8 26.0

In the light of these results, it seems like the economic potential for WHR from industrial
processes with ORC technology is remarkable, if an amortization time of 10 years is con-
sidered. It can be up to 435 MWe summing up the considered sectors.

4.4 Sensitivity analysis of the results


The previous results are based on specific assumptions. However, the economic poten-
tial may and will change when these parameters are varied. In order to gain a better under-
standing of the influence of each parameter on the results, a sensitivity analysis is carried
out, where the interest rate 𝑖, the amortization time 𝑛, the net electrical efficiency of the ORC
are varied. For simplifying the understanding, only the average scenario and the lowest
electricity price are considered.
Figure 4.3 and Figure 4.4 show how the economic potential varies when the amortization
time in years and the interest rate change within a range. Regarding the amortization time,
5 to 15 years have been considered for the analysis, and the interest rate varies from 2%
to 15%. The ORC efficiency and the ORC availability have an influence only on the technical
potential. The technical potential does not depend on the interest rate and amortization
time.
For BOF, due to the extreme favorability of the ORC installation, the economic potential
does not vary with changes in the parameters in these ranges. In the case of the EAF, the
economic potential starts decreasing when the amortization time is less than 10 years and
becomes zero before reaching 5 years. The increasing interest rate makes the economic
potential drop until zero at a little less than 15%. For RHF, the economic potential is zero
below 5 years of amortization and for more than 15% of interest rate.
Results 57

For the case of cement (Figure 4.4), the situation is better. The economic potential drops
to around 60 GWh/a at 5 amortization years and it does not become zero for an interest
rate of 15%. Glass is the least favorable sector, because of the lower utilization rate and
smaller size of the plants. No ORC in no plant is economical below 10 years of amortization
time or above 4.3 % of interest rate (Figure 4.4).

Economic potential [GWh/a]


Economic potential [GWh/a]

(a) (b)
Economic potential [GWh/a]

(c)
Figure 4.3. Sensitivity analysis for: (a) BOF; (b) EAF and (c) RHF
58 4.5 Electricity savings

Economic potential [GWh/a]


Economic potential [GWh/a]

(a) (b)
Figure 4.4. Sensitivity analysis for: (a) cement and (b) glass

4.5 Electricity savings


The main benefit of the installation of these ORC units in the plants where it results
economically interesting is the electricity that is saved, since it is produced with the excess
heat which would otherwise be useless.
The steel industry uses about 22 TWh of electricity every year. That is 9% of Germany’s
industrial consumption and 4% of total consumption in the country. Electricity is required for
powering EAFs that melt scrap to new steel and for operating rolling mills [94].
For the cement industry, each tonne of cement requires approximately 110 kWh of elec-
tricity [91]. That, multiplied by the cement production in Germany of 23.17 Mt makes 2.55
TWh/a. The production of glass requires approximately 203 kWh/t for flat glass and 372
kWh/t for container glass [95]. Taking into account the approximate glass production in Ger-
many per year (1.98 Mt of flat glass and 3.93 Mt of container glass) [92], a electricity re-
quirement of approximately 1.86 TWh/a is obtained.
Table 4.7 shows the percentage of the total electricity consumed in the German steel,
cement and glass production that could be saved by the installation of the profitable ORCs
for the lowest electricity price and in the case of 10 years of amortization time, interest rate
of 4%, ORC availability of 95% and ORC efficiency of 15%. An interval from worst- to best-
case scenario is calculated.
Results 59

Table 4.7. Saved electricity with ORC for steel, cement and glass with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %,
i=4%, n=10 years

Sector Electricity saved Electricity con- 𝑬𝒍𝒆𝒄𝒕𝒓𝒊𝒄𝒊𝒕𝒚 𝒔𝒂𝒗𝒆𝒅


[%]
[TWh/a] sumed [TWh/a] 𝑬𝒍𝒆𝒄𝒕𝒓𝒊𝒄𝒊𝒕𝒚 𝒄𝒐𝒏𝒔𝒖𝒎𝒆𝒅

Steel 1.30-1.78 22.00 5.91-8.09%

Cement 0.68-1.02 2.55 26.67-40%

Glass 0-0.120 1.86 0-6.45%

4.6 CO2 emissions avoided


In 2018, the CO2 emission factor for the electricity mix in Germany is estimated at 474
g/kWh on average (Figure 4.5).

Figure 4.5. Development of the CO2 emission factor for the electricity mix in Germany in the
years 1990 to 2018 (in g/kWh) [96]
Table 4.8. CO2 emissions avoided with the economic potential of ORCs for steel, cement and
glass with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 years

Sector CO2 emissions 𝐂𝐎𝟐 𝐞𝐦𝐢𝐬𝐬𝐢𝐨𝐧𝐬 𝐚𝐯𝐨𝐢𝐝𝐞𝐝


[%]
avoided [103 tCO2/a] 𝑮𝑯𝑮 𝒆𝒎𝒊𝒔𝒔𝒊𝒐𝒏𝒔 𝒊𝒏𝒅𝒖𝒔𝒕𝒓𝒚 𝑮𝒆𝒓𝒎𝒂𝒏𝒚 𝟐𝟎𝟏𝟖

Steel 616.53-844.57 0.95-1.30%

Cement 324.64-485.61 0.50-0.75%

Glass 0-57.12 0-0.09%


60 4.7 Cost savings

With the implementation of these WHR systems, a part of the electricity generation can
be avoided because the plant can use the energy generated with the ORC instead of taking
it from the grid. Therefore, the CO 2 emission associated to the production of this electricity
is reduced.
The GHG emissions from the industry sector in Germany in 2018 measured in CO2
equivalents in million tonnes are 65 MtCO 2eq (Figure 4.6). The installation of the economic
potential of ORC units for the assumptions in Table 4.8 would suppose GHG reductions
which can be read in the table. These GHG reduction is in the form of CO 2 emissions
avoided from savings in electricity production.

Figure 4.6. GHG emission trends in Germany by sector 1990-2018. Data from: UBA 2019,
BMU 2019 [97]

4.7 Cost savings


Not only is the installation of ORC units good for sustainability goals, but also it can save
money by reducing the electricity purchase cost.
The total cost savings are estimated for the case where the ORC efficiency is 15%, i=4%
and n=10. The electricity price is 5.1 ct/kWh (Table 4.9).
Table 4.9. Cost savings with the economic potential of ORCs for steel, cement and glass
with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 years

Sector Cost saving [M€/a]

Steel 66.34-90.87

Cement 34.93-52.25

Glass 0-6.1
Summary, conclusions and outlook 61

5 Summary, conclusions and outlook

5.1 Summary and conclusions


The goal of this work is to estimate the technical and economic potential of the installation
of ORC units for the recovery of waste heat from industrial processes in Germany. The most
interesting industries for WHR are steel, cement and glass [4].
To estimate the recovery potential, a combination of the bottom-up and top-down ap-
proaches is used. The bottom-up method is used to gather data on the capacity of each
plant and to assess the specific waste heat available from each process susceptible of
waste heat recovery with an ORC unit. These production process parts are BOF, EAF and
RHF for steel, preheater and clinker cooler for cement and melting furnace for glass. The
top-down method is applied for estimating the total waste heat released by each plant,
where a common specific waste heat value (which is different depending on the process) is
multiplied by the production of each plant.
The uncertainty and high variability of the specific waste heat values for the same indus-
trial process led to considering a worst-, average- and best-case scenario. Also, for the
estimation of the technical potential, the ORC availability needs to be considered, as well
as the ORC unit efficiency. To estimate this efficiency, Turboden WHR-ORC units are con-
sidered because they are the global leading company in waste heat recovery applications
from industrial processes in both number of installations and capacity. The ORC availability
is considered to be 95% and the ORC efficiency varies in the approximate range of 15-21%.
For the analysis, a conservative case with an efficiency of 15% and another with 19% are
studied.
The economic potential is the accumulated electricity that is produced by the ORCs hav-
ing a lower LCOE than the electricity price. The range of electricity prices in 2018 for EII in
Germany is in the range of 5.10-17.00 ct/kWh, with average price of 8.84 ct/kWh [5].
For the LCOE calculation, the ORC units’ specific investment costs need to be known.
To estimate these costs, a correlation by Turboden which depends on the unit size is used.
Also, assumptions regarding the O&M costs are made. The LCOE depends also on the
number of amortization years and the interest rate. As an initial assumption, an amortization
time of 10 years and an interest rate of 4% are considered for the economic analysis. Then,
a sensitivity analysis is made where these parameters are varied in order to evaluate their
influence on the economic potential.
German industry releases more than 200 TWh in waste heat every year, which is more
than the total annual energy consumption of Denmark [6]. The large number of steel, ce-
ment and glass factories in Germany produce a significant amount of high-temperature pro-
cess heat. In addition, these three energy-intensive industries are the ones showing the
most developed state-of-the-art on waste heat recovery with ORC in the world, with room
for a significative increase in the case of Germany, as shown in this work.
The optimistic results obtained from this work, together with the incentives and support
programs from the European Union and the German Energy Efficiency Fund (EEF) [7],
make the ORC technology a very attractive option for industrial waste heat recovery. The
62 5.2 Future lines

economic potential grows with the amortization time and decreases with higher interest
rates. The economic potential results show promising for all the sectors studied (steel, ce-
ment and glass production) for amortization times between 10 and 5 years (except for glass,
which is longer) and interest rates smaller than 15% (except for glass, which is lower). The
results are especially optimistic for BOF, EAF and cement production. The installation of
ORC units for waste heat recovery in the plants where it is economically feasible could save
up to 149.22 M€/a and avoid 1.3873 Mt of CO 2 emissions per year, which corresponds to
2.13% of the GHG emissions from the industry in Germany in 2018.

5.2 Future lines


The methodology and results presented in this thesis yield some guidelines for future
work and further research on the following topics:
• Application of the developed method for the evaluation of the WHR with ORC
viability for other processes (ceramics, oil & gas, chemical) from energy-intensive
industry in Germany.
• Extension of this study to other countries of interest for the installation of this
technology.
• The specific waste heat from each process in each particular plant could be gath-
ered in order to improve the accuracy of the results instead of having a worst-,
average and best-case scenario. Access to this information for the German in-
dustries would contribute to a better estimation.
Bibliography 63

Bibliography

[1] A. Fernandez-Pales, P. Levi and T. Vass, "Tracking Clean Energy Progress:


Industry," 24 May 2019. [Online]. Available: https://www.iea.org/tcep/industry/#fi-
nal-energy-consumption. [Accessed 9 May 2019].

[2] M. Papapetrou, G. Kosmadakis, A. Cipollina, U. L. Commare and G. Micale,


"Industrial waste heat: Estimation of the technically available resource in the EU
per industrial sector, temperature level and country," Applied Thermal Engineer-
ing, vol. 138, pp. 207-216, 2018.

[3] A. Widrat, "Sustainably digital: Waste heat recovery as the key to green data
centres," Enviro Sustain, 30 July 2019.

[4] T. Tartière and M. Astolfi, "A World Overview of the Organic Rankine Cycle
Market. IV International Seminar on ORC Power Systems," Energy Procedia 129,
pp. 2-9, 2017.

[5] BDEW, "BDEW (Bundesverband der Energie- und Wasserwirtschaft e.V.)-


Strompreisanalyse Mai 2018. Haushalte und Industrie," Berlin, 2018.

[6] Germany Trade and Industry, "Germany Trade and Industry. Invest. Industries
in Germany. Energy Efficiency in Industry," 2019. [Online]. Available:
https://www.gtai.de/gtai-en/invest/industries/energy/energy-efficiency-in-industry.

[7] Odyssee-Mure, "Germany, energy profile," 2018.

[8] D. Forni, F. Campana and D. D. Santo, "Innovative system for electricity gen-
eration from waste heat recovery," in ECEEE 2014 Industrial Summer Study- Re-
tool for a Competitive and Sustainable Industry, 2014.

[9] European Commission, "European Commission. Energy efficiency – targets,


directive and rules," [Online]. Available: https://ec.europa.eu/energy/en/topics/en-
ergy-efficiency/targets-directive-and-rules/eu-targets-energy-efficiency.

[10] E. Woolley, Y. Luo and A. Simeone, "Industrial waste heat recovery: A system-
atic approach," Sustainable Energy Technologies and Assessments, vol. 29, pp.
50-59, 2018.

[11] M. Pehnt, J. Bödeker, M. Arens, E. Jochem and F. Idrissova, "Industrial waste


heat – tapping into a neglected efficiency potential," in eceee 2011 Summer Study
64

on energy efficiency: Energy efficiency first: The foundation of a low-carbon soci-


ety. Panel 3. Energy use in industry: The road from policy to action, Presqu’île de
Giens, France, 2011.

[12] S. Bruckner, S. Liu, M. Laia, M. Radspieler, L. Cabeza and L. Eberhard, "In-


dustrial waste heat recovery technologies: an economic analysis of heat transfor-
mation technologies," Appl. Energy, no. 151 (1), pp. 157-167, 2015.

[13] Umwelt Bundesamt auf Basis Arbeitsgemeinschaft Energiebilanzen, "Enden-


ergieverbrauch 2017 nach Sektoren und Energieträgern. Energieverbrauch nach
Energieträgern, Sektoren und Anwendungen," 15 March 2019. [Online]. Available:
https://www.umweltbundesamt.de/sites/default/files/me-
dien/384/bilder/dateien/4_abb_eev-sektoren-et_2019-02-26.pdf.

[14] D. Connolly, B. V. Mathiesen, P. A. Østergaard, B. Möller, S. Nielsen, H. Lund,


U. Persson, S. Werner, J. Grözinger, T. Boermans, M. Bosquet and D. Trier, "Heat
Roadmap Europe 2: Second Pre-Study for the EU27," 2013.

[15] L. Miró, "Industrial waste heat: mapping, estimations and recovery by means
of TES," 2016.

[16] IEA, "Annex XV: Industrial Excess Heat Recovery- Technologies and Applica-
tions," 2015.

[17] U.S. Department of Energy (prepared by BCS Inc.), "Waste Heat Recovery:
Technology and Opportunities in the U.S. Industry," 2008.

[18] F. Campana, M. Bianchi, L. Branchini, A. D. Pascale, A. Peretto, M. Baresi, A.


Fermi, N. Rossetti and R. Vescovo, "ORC waste heat recovery in European en-
ergy intensive industries: Energy and GHG savings," Energy Conversion and
Management 76, pp. 244-252, 2013.

[19] M. Amat, "Desarrollo de un modelo de ciclo orgánico Rankine. Ejemplo de apli-


cación para análisis de fluidos de trabajo de bajo potencial de efecto invernadero,"
Castellón de la Plana, 2017.

[20] y. Quoilina, M. V. D. Broekb, S. Declayea, P. Dewallefa and V. Lemort,


"Techno-economic survey of Organic Rankine Cycle (ORC) systems," Renewable
and Sustainable Energy Reviews 22, pp. 168-186, 2013.

[21] Y. T. Shah, Thermal Energy: Sources, Recovery, and Applications, CRC Press.
taylor & Francis Group, 2018.

[22] Turboden, "Turboden ORC technology," 2019. [Online]. Available:


https://www.turboden.com/turboden-orc-technology/1062/the-orc-technology.
Bibliography 65

[23] Exergy S.p.A, "Exergy. Heat recovery from industrial processes," 2019.
[Online].

[24] S. Lecompte, O. A. Oyewunmi, C. N. Markides, M. Lazova, A. Kaya, M. v. d.


Broek and M. D. Paepe, "Case Study of an Organic Rankine Cycle (ORC) for
Waste Heat Recovery from an Electric Arc Furnace (EAF)," Energies 10, 649,
2017.

[25] Turboden, "H-REII Demo. Project supported by LIFE EU Programme," 2014.

[26] Turboden, "The first Electric Arc Furnace waste heat recovery ORC-based sys-
tem is in commercial operation at Feralpi steel plant in Germany," 2014.

[27] Executive Agency for Small and Medium-sized Enterprises (European Com-
mision), "Climate Change Mitigation in Energy Intensive Industries (EII)," in LIFE
Platform Meeting, Utrecht, 2018.

[28] S. Brueckner, L. Miró, L. F. Cabeza, M. Pehnt and E. Laevemann, "Methods to


estimate the industrial waste heat potential of regions – A categorization and liter-
ature review," Renewable and Sustainable Energy Reviews, vol. 38, pp. 164-171,
October 2014.

[29] M. Blesl, S. Kempe, M. Ohl, U. Fahl, A. König, T. Jenssen and L. Eltrop, "Wär-
meatlas Baden-Württemberg: Erstellung eines Leitfadens und Umsetzung für
Modellregionen," 2008.

[30] D. Connolly, B. Mathiesen, P. Østergaard, B. Möller and S. Nielsen, "Heat


Roadmap Europe 2050. Second Pre-Study for the EU-27," 2013.

[31] U. Persson, B. Möller and S. Werner, "Heat Roadmap Europe: Identifying stra-
tegic heat synergic regions," Energy Policy, vol. 74, pp. 663-681, 2014.

[32] S. Brueckner, R. Arbter, M. Pehnt and E. Laevemann, "Industrial waste heat


potential in Germany—a bottom-up analysis," Energy Efficiency, vol. 10, no. 2, p.
513–525, 2017.

[33] S. Eyerer, C. Schifflechner, S. Hofbauer, C. Wieland, K. Zosseder, W. Bauer,


T. Baumann, F. Heberle, C. Hackl, M. Irl and H. Spliethoff, "Potential der hydro-
thermalen Geothermie zur Stromerzeugung in Deutschland," 2017.

[34] M. Bianchi, L. Branchini, A. D. Pascale, F. Melino, A. Peretto, D. Archetti, F.


Campana, T. Ferrari and N. Rossetti, "Feasibility of ORC application in natural
gas compressor stations," Energy, vol. 173, pp. 1-15, 2019.
66

[35] R. Pili, A. Romagnoli, H. Spliethoff and C. Wieland, "Techno-Economic Analy-


sis of Waste Heat Recovery with ORC from Fluctuating Industrial Sources," En-
ergy Procedia, vol. 129, pp. 503-510, 2017.

[36] F. C. Matthes, "The current electricity costs of energy-intensive industries in


Germany," 2013.

[37] H.-J. Odenthal, J. Alken, A. Kemminger, F. Krause, N. Vogl and S. Rühl, "Ap-
plied numerical simulation for safe and efficient process conditions of metallurgical
plants," in 8th European Oxygen Steelmaking Conference (EOSC), Taranto,
2018.

[38] Stahl, "The steel industry as an energy-intensive sector," 2019. [Online]. Avail-
able: https://en.stahl-online.de/index.php/topics/energy-and-environment/en-
ergy/.

[39] Bethlehem Steel Corp, "Oxygen Steelmaking Processes," in The making,


shaping, and treating of steel, Pittsburgh, PA : AISE Steel Foundation, 1998, pp.
475-524.

[40] T. Park, B. Kim, T. Kim, I. Jin and Y. Yeo, "Comparative Study of Estimation
Methods of the Endpoint Temperature in Basic Oxygen Furnace Steelmaking Pro-
cess with Selection of Input Parameters," Korean Journal of Metals and Materials,
no. 56 (11), pp. 813-821, 2018.

[41] M. Arens and E. Worrell, "Energy efficient technologies in the German steel
industry – low hanging fruits?," in eceee 2014 Industrial Summer Study: Retool
for a competitive and sustainable industry. Panel 4. Undertaking high impact ac-
tions: The role of technology and systems optimisation, 2014.

[42] N. P. Cheremisinoff, P. Rosenfel and A. R. Davletshin, Responsible Care: A


New Strategy for Pollution Prevention and Waste Reduction Through Environment
Management, G. P. Company, Ed., 2008, pp. 435-476.

[43] M. Kirschen, Energieeffizienz und Emissionender Lichtbogenöfen in der


Stahlindustrie, Aachen: Verlag Stahleisen GmbH, 2007.

[44] K. El-Akruti, T. Zhang and R. Dwight, "Developing an optimum maintenance


policy by life cycle cost analysis – a case study," International Journal of Produc-
tion Research, no. 54 (19), pp. 1-17, 2016.

[45] Stahl, "Stahl web site. Technology. Steelmaking," 2019. [Online]. Available:
https://en.stahl-online.de/index.php/topics/technology/steelmaking/3/.
Bibliography 67

[46] O. Rentz, R. Jochum and F. Schultmann, "Report on Best Available Tech-


niques (BAT) in the German Ferrous Metals Processing Industry," Karlsruhe,
1999.

[47] Feralpi Stahl, "Elbe-Stahlwerke Feralpi GmbH. Production. From scrap to high-
quality steel," 2019. [Online]. Available: https://www.feralpi.de/en/company/elbe-
stahlwerke-feralpi-gmbh/production.html.

[48] Stahlwerk Thüringen GmbH, "A modern steel production site with tradition".

[49] P. Cavaliere, Clean Ironmaking and Steelmaking processes. Efficient technol-


ogies for Greenhouse Emissions Abatement, Springer, 2019.

[50] NEDO, "Global Warming Countermeasures: Japanese Technologies for En-


ergy Savings / GHG Emissions Reduction," 2008.

[51] K. Gandt, T. Meier, T. Echterhof and H. Pfeifer, "Heat recovery from EAF off-
gas for steam generation: analytical exergy study of a sample EAF batch," Iron-
making & Steelmaking: Processes, Products and Applications, vol. 43, no. 8, pp.
581-587, 2016.

[52] G. Nardin, G. Ciotti, F. D. Magro, A. Meneghetti and P. Simeoni, "Waste heat


recovery in the steel industry: better internal use or external integration?," in XXIII
Summer School “Francesco Turco” – Industrial Systems Engineering, Palermo,
2018.

[53] C. Born and R. Granderath, "Benchmark for heat recovery from the offgas duct
of electric arc furnaces," MPT Metallurgical Plant and Technology International,
vol. 36, no. 1, pp. 32-35, 2013.

[54] P. Frittella, A. Ventura, S. Galassi, L. Angelini, T. Bause, M. Baresi, D. Forni


and R. Granderath, "Modeling Approach for the Analysis of Energy Recovery Ben-
efits Applied in EAF Process for the Case of Elbe Stahlwerke Feralpi GmbH,"
2015.

[55] C. Weikang, "Organic Rankine Cycle (ORC) waste heat recovery system at
rolling mill reheating Furnace," 2015.

[56] Turboden, "ORC: direct exchange," 2017.

[57] Sidenor, "Energetic flowchart for cement, glass, steel industries and petro-
chemical sectors," 2016.

[58] VDZ, "Environmental Data of the German Cement Industry," Düsseldorf, 2018.
68

[59] S. Khurana, R. Banerjee and U. Gaitonde, "Energy balance and cogeneration


for a cement plant," Applied Thermal Engineering, vol. 22, no. 5, pp. 485-494,
2002.

[60] T. Engin and V. Ari, "Energy auditing and recovery for dry type cement rotary
kiln systems––A case study," Energy Conversion and Management, vol. 46, p.
551–562, 2005.

[61] M. Renó, L. Alves, J. Palacio, L. Souza, F. Centeno González and P. Torres,


"Environmental analyze of cement production with application of wastes," Enge-
vista, vol. 19, 2017.

[62] G. Singh, Applied Chemistry, New Deli: Discovery Publishing House Pvt. Ltd.,
2009.

[63] R. Altieri, F. Campana and R. Vescovo, "D2.5 A - Report of data for the cement
plant complete with leaflet for dissemination," 2015.

[64] Lafarge Holcim, "Technical and economical experiences with large ORC sys-
tems using industrial waste heat streams of cement plants," in 2. ORC-Sympo-
sium. Hochschule Luzern - Technik & Architektur Horw, Luzern, 2015.

[65] International Finance Corporation, "Waste Heat Recovery for the Cement Sec-
tor: Market and Supplier Analysis," 2014.

[66] F. Schorcht, I. Kourti, B. M. Scalet, S. Roudier and L. D. Sancho, "Best Availa-


ble Techniques (BAT) Reference Document for the Production of Cement, Lime
and Magnesium Oxide," Luxembourg, 2013.

[67] BV Glas, "BV Glas. Types of glass," [Online]. Available:


https://www.bvglas.de/en/about-glass/glass-is-a-multitalented-material/types-of-
glass/.

[68] M. Achintha, "Sustainability of glass in construction," in Sustainability of Con-


struction Materials, 2016, pp. 79-104.

[69] Emhart Glass, "Bucher Emhart Glass," [Online]. Available: https://www.em-


hartglass.com/node/22765.

[70] Interreg Central Europe, "Interreg Central Europe. CE-Heat," 20 February


2019. [Online]. Available: https://www.interreg-central.eu/Content.Node/CE-
HEAT/Waste-heat-recovery-in-the-glass-industry.html.
Bibliography 69

[71] Turboden, "Turboden Press Releases. Turboden to supply a 6 MW ORC unit


for a glass factory in Turkey," 5 April 2017. [Online]. Available: https://www.turbo-
den.com/press/press-releases/1672/turboden-to-supply-a-6-mw-orc-unit-for-a-
glass-factory-in-turkey.

[72] INMIS Energy, "Heat Recovery. Turboden ORC units. HR units. STANDARD
HR UNITS. Data sheet," 2019. [Online]. Available: http://www.inmis-
energy.com/5-0-heat-recovery/turboden-orc-for-hr/5-5-1-hr-units.

[73] E. Pretorius and H. Oltmann, "EAF Fundamentals," 2017.

[74] D. J. Zuliani, V. Scipolo and C. Born, "Opportunities for increasing productivity,


lowering operating costs and reducing greenhouse gas emissions in EAF and
BOF steelmaking," in Steelmaking and casting. MILLENNIUM STEEL INDIA
2010, 2010.

[75] S. U. Nimbalkar, A. C. Thekdi, B. M. Rogers, O. L. Kafka and T. J. Wenning,


"Technologies and Materials for Recovering Waste Heat in Harsh Environments,"
2014.

[76] M. Kirschen, V. Risonarta and H. Pfeifer, "Energy efficiency and the influence
of gas burners to the energy related carbon dioxide emissions of electric arc fur-
naces in steel industry," Energy, vol. 34, no. 9, pp. 1065-1072, 2009.

[77] G. F. E. Agency, "Abwärmenutzungs-potenziale in Anlagen integrierter


Hüttenwerke der Stahlindustrie. [English: Potential for Waste Heat Recovery from
Integrated Steelmaking Plants in the Steel Industry]," 2019.

[78] Gerdau, "TASIO project, Deliverable D2.1. Energetic flowchart for cement,
glass, steel industries and petrochemical sectors.," 2016.

[79] D. Zuliani, V. Scipolo, J. Maiolo and C. Born, "Opportunities for Increasing


Productivity and Lowering Operating Costs While Reducing GHG Emissions in
Steelmaking," in Proceedings of the Iron & Steel Technology Conference: Assoc.
for Iron and Steel Tech., Pitts-burgh, Pennsylvania, U.S.A, 2010.

[80] L.-Z. Yang, R. Zhu and G.-H. Ma, "EAF Gas Waste Heat Utilization and Dis-
cussion of the Energy Conservation and CO2 Emissions Reduction p. 881," High
Temperature Materials and Processes, vol. 35, no. 2, p. 881, 2016.

[81] B. Lee and I. Sohn, "Review of Innovative Energy Savings Technology for the
Electric Arc Furnace," JOM: The Journal of The Minerals, Metals & Materials So-
ciety, vol. 66, no. 9, p. 1581–1594, 2014.
70

[82] T. Keplinger, M. Haider, T. Steinparzer, A. Patrejko, P. Trunner and M.


Haselgrübler, "Dynamic simulation of an electric arc furnace waste heat recovery
system for steam production," vol. 135, pp. 188-196, 2018.

[83] S. U. Nimbalkar, A. C. Thekdi, B. M. Rogers, O. L. Kafka and T. J. Wenning,


"Technologies and Materials for Recovering Waste Heat in Harsh Emvironments,"
Oak Ridge National Laboratory (US Department of Energy), 2014.

[84] K. Manatura and M. Tangtrakul, "Energy Conservation in a Reheating Furnace


Using Regenerative burners Combined with Recuperator," Silpakorn U Science &
Tech J, vol. 4, no. 2, pp. 7-13, 2010.

[85] K. Kangvanskol and C. Tangthieng, "An Energy Analysis of a Slab Preheating


Chamber for a Reheating Furnace," Engineering Journal, vol. 18, no. 2, 2014.

[86] R. Ortwein, "Elektrostahlwerke. Erschließung ungenutzter Potentiale für Indus-


trie und anschließende Gemeinden. [English: Electric steel plants. Analysis of un-
used Potentials for Industry and connected Communities]Stuttgart," Stuttgart,
2014.

[87] H. Pfeifer, "Trends im Industrieofenbau (Hybrid Heating). [English: Trends in


the Construction of Industrial Furnaces (Hybrid Heating)]," 2014.

[88] Stahl, "Fakten zur Stahlindustrie in Deutschland," 2014.

[89] VDEh, "PLANTFACTS Database," 2019.

[90] Wordsteel Association, "Steel Statistical Yearbook 2018," [Online]. Available:


https://www.worldsteel.org/steel-by-topic/statistics/steel-statistical-yearbook.html.

[91] VDZ, "VDZ. Cement industry. German cement sector. Cement works," 2017.
[Online]. Available: https://www.vdz-online.de/en/cement-industry/cement-sec-
tor/cement-works/.

[92] J. Brunke, "Energieeinsparpotenziale von energieintensiven Produk-


tionsprozessen in Deutschland : eine Analyse mit Hilfe von Energieeinspar-
kostenkurven".

[93] A. Foresti and D. Archetti, "Waste Heat Recovery Valorization with ORC Tech-
nology," 2018.

[94] Stahl, "Stahl online. Energy and Environment. The steel industry as an energy-
intensive sector," 2019. [Online]. Available: https://en.stahl-online.de/in-
dex.php/topics/energy-and-environment/energy/.
Bibliography 71

[95] Ecofys, "Methodology for the free allocation of emission allowances in the EU
ETS post 2012. Sector report for the glass industry," 2009.

[96] Umwelt Bundesamt, "Entwicklung der spezifischen Kohlendioxid Emissionen


des deutschen Strommix in den Jahren 1990 - 2018," 2019.

[97] K. Appunn and J. Wettengel, "Germany’s greenhouse gas emissions and cli-
mate targets," Clean Energy Wire. Journalism for the energy transition, 11 Octo-
ber 2019.

[98] M. Emre Ertem e S. Gürgen, «Energy balance analysis for Erdemir blast fur-
nace number one,» Applied Thermal Engineering, vol. 26, n. 11-12, pp. 1139-
1148, 2006.

[99] Markewitz, Zhao, Ryssel, Moumin and Wang, "Carbon Capture for CO2 Emis-
sion Reduction in the Cement Industry in Germany," Energies, vol. 12, no. 12, p.
2432, 2019.

[100] Tenova, "Innovative solutions for metals & mining. Metal and Mining industry.
Metals. Metal making. Fume treatment. Heat Recovery/Cogeneration Systems.
Product: Production of Steam / Hot Water," 2019. [Online]. Available:
https://www.tenova.com/product/production-of-steam-hot-water/.

[101] E. Spagnoli, "Waste heat recovery opportunity in Energy Intensive Industry with
ORC technology," in Part of: Webinar on “TASIO project: Waste Heat Recovery
Systems based on the Organic Rankine Cycle (ORC) technology: opportunities
for sustainable energy intensive industries”, 2019.

[102] ORMAT, "ORMAT. Global projects," 2019. [Online]. Available: https://www.or-


mat.com/en/projects/all/main/?Country=0&Seg=0&Tech=8.

[103] VDZ, "VDZ. Cement industry. Energy consumption. Energy input in the cement
industry," 2019. [Online]. Available: https://www.vdz-online.de/en/cement-in-
dustry/energy-consumption/.

[104] Turboden, «Turboden. Applications. Waste Heat Recovery,» 2019. [Online].


Available: https://www.turboden.com/applications/1053/waste-heat-recovery.

[105] H. Spliethoff, Combustion of solid fuels, Berlin: Springer, 2010.

[106] M. Kaltschmitt, W. Streicher e A. Wiese, Erneuerbare Energien -


Systemtechnik, Wirtschaftlichkeit, Umweltaspekte, Berlin Heidelberg: Springer
Vieweg, 2013.
72

[107] A. Rettig, M. Lagler, T. Lamare, S. Li, V. Mahadea, S. McCallion and J. Cher-


nushevich, "Application of Organic Rankine Cycles (ORC)," in World Engineers'
Convention, Geneva, 2011.

[108] G. Nardin, A. Meneghetti, F. D. Magro and N. Benedetti, "PCM-based energy


recovery from electric arc furnaces," Applied Energy, vol. 136, pp. 947-955, 2014.

[109] M. Jha and V. Singh, "Assessment of energy efficiency in reheating furnace of


a steel plant by using process heating assessment and survey tool (PHAST),"
Recent Research in Science and Technology, vol. 5, no. 5, pp. 33-36, 2013.

[110] N. Pardo, J. Moya and K. Vatopoulos, "Prospective Scenarios on Energy Effi-


ciency and CO2 Emissions in the EU Iron & Steel Industry," 2012.

[111] R. Remus, M. A. Aguado-Monsonet, S. Roudier and L. D. Sancho, "Best Avail-


able Techniques (BAT) Reference Document for Iron and Steel Production," 2013.

[112] H. Schliephake, C. Born, R. Granderath, F. Memoli and J. Simmons, " Heat


recovery for the EAF of Georgsmarienhütte, Germany," Iron and Steel Technol-
ogy, vol. 8, no. 5, pp. 330-335, 2011.

[113] B. Kleimt, "Valorisation and Dissemination of EAF Technology- VALEAF," Düs-


seldorf, 2015.

[114] J. Goldemberg and O. Lucon, Energy, Environment and Development, London:


Earthscan, 2010.

[115] L.-z. Yang, R. Zhu and G.-h. Ma, "EAF Gas Waste Heat Utilization and Discus-
sion of the Energy Conservation and CO2 Emissions Reduction," High Tempera-
ture Materials and Processes, vol. 35, no. 2, p. 195–200, 2016.

[116] Chemical Engineering, "Business and Economics. CEPCI Updates," 2019.


[Online]. Available: https://www.chemengonline.com/cepci-updates-january-
2018-prelim-and-december-2017-final/.

[117] Institute for Energy and Environmental Research (IFEU), «The use of industrial
waste heat - techno-economic potential and energy policy,» Heidelberg, Karls-
ruhe, 2010.
Appendix i

6 Appendix

A Resumen en español

Introducción: motivación y justificación


En la actualidad, la producción industrial no cesa de aumentar mientras que el mix ener-
gético del sector se mantiene casi inalterado. Esto ha provocado un aumento de las emi-
siones de CO2, que alcanzaron las 8,5 GtCO2 en 2017, lo que supuso un 24% de las emi-
siones globales. En un Escenario de Desarrollo Sostenible (SDS), definido por la AIE, las
emisiones deben disminuir hasta 8,3 GtCO 2 para 2030 [1]. La eficiencia energética juega
un papel importante para alcanzar este objetivo.
El consumo de energía puede reducirse mediante su uso eficiente, por ejemplo, evitando
liberar grandes cantidades de energía no utilizada en forma de calor residual en los gases
de escape. Este calor excedente de los procesos industriales puede recuperarse mediante
turbogeneradores que utilizan la tecnología del ciclo orgánico de Rankine (ORC). Estos
sistemas pueden utilizar este calor para producir solo electricidad o una combinación de
calor y electricidad (CHP). La electricidad producida puede, por ejemplo, reutilizarse inter-
namente en la misma planta reduciendo su demanda.
Alemania tiene un alto potencial para la utilización del calor residual procedente del sec-
tor industrial (el mayor de Europa) [2] que aún no ha sido completamente explotado [3].
Esto, junto con el buen funcionamiento para esta aplicación de los turbogeneradores ORC
ya instalados en varias industrias en todo el mundo, hace que resulte interesante estimar
tanto el potencial técnico como el económico de la integración de unidades ORC para la
recuperación del calor residual de las principales industrias en Alemania.
Por lo tanto, en el contexto de la Directiva de Eficiencia Energética de la Comisión Eu-
ropea, que establece un conjunto de medidas para ayudar a la UE a alcanzar su objetivo
de 20% de eficiencia energética para el año 2020, la recuperación del calor residual es
fundamental. El objetivo del 20% de eficiencia energética consiste en ahorrar un 20% del
consumo de energía primaria y final para el año 2020, en comparación con las proyeccio-
nes del "business-as-usual" [9].
En Alemania, el Fondo de Eficiencia Energética (FEE) forma parte de la llamada "Ener-
giewende" (transición energética) que tiene como objetivo reducir el consumo de energía
primaria en un 20% para el año 2020 y en un 50% para el 2050 (en comparación con el
año 2008). Además, las emisiones de gases de efecto invernadero deben reducirse en un
40% hasta el 2020 y en un 80 a 95% hasta el 2050 (año base 1990). Actualmente, el fondo
consta de 23 medidas políticas, incluyendo programas de apoyo y actividades educativas.
Los fondos provienen del Ministerio Federal de Economía y Energía (BMWi). Los progra-
mas son administrados por diversas autoridades federales o por el banco de desarrollo
KfW. Los beneficiarios son empresas, hogares y municipios [7].
2

Existen escasos estudios sobre el potencial de recuperación del calor residual con ORC
en las principales industrias de Alemania (energéticamente intensivas). Concretamente, no
existen estudios sobre el potencial económico de la instalación de unidades ORC para la
producción de electricidad a partir de este calor. Por otra parte, la mayoría de las estima-
ciones del potencial técnico actuales se derivan de estudios en otros países.

Objetivos
El objetivo de esta tesis es, por lo tanto, proporcionar a los investigadores y a la industria
alemana un mayor conocimiento sobre el potencial, tanto técnico como económico, de esta
técnica de recuperación de calor residual con la tecnología ORC.
Para evaluar en qué medida los sistemas ORC pueden ser utilizados para recuperar el
calor residual de los gases de escape de procesos industriales, se procede a estimar el
potencial técnico y económico. Los siguientes pasos son necesarios para la evaluación del
potencial:
• Revisión de la literatura sobre los procesos de las industrias de alto consumo
energético, el calor residual que se obtiene de los gases de escape de cada uno
y la tecnología ORC.
• Caracterización de la distribución de los sectores industriales en Alemania y el
calor residual disponible.
• Análisis de las principales plantas industriales en el país (para la producción de
acero, cemento y vidrio) y caracterización de sus dimensiones.
• Estimación del rendimiento del proceso de recuperación del calor residual y del
potencial técnico.
• Estimación del potencial económico considerando costos, tiempo de amortiza-
ción y otros aspectos económicos.

Estado del arte: estudios previos


En Alemania, el sector industrial supuso el 28,9% de la energía total consumida, con
una cantidad de 750 TWh/a, lo que lo convierte en el mayor sector consumidor [13]. De
estas cifras se deduce que las medidas de eficiencia energética y la recuperación de calor
residual de los procesos industriales pueden tener una gran contribución al total de la ener-
gía consumida y de las emisiones del país.
Varios trabajos han intentado estimar el potencial de recuperación del calor residual
industrial en Alemania. Estos estudios utilizan, en principio, un enfoque “top-down” o “bot-
tom-up”. En el primero, el calor residual se estima según las cifras clave disponibles para
otros países y se adapta al caso específico. En el segundo, a partir de los datos de las
instalaciones individuales, se calcula el calor residual total.
Appendix iii

Pehnt et al. (2011) [11] utilizan los resultados de un estudio noruego para determinar la
eficiencia energética en diferentes procesos industriales y adaptan los resultados a la in-
dustria alemana. Utilizando un enfoque “top-down”, se estima un potencial de 476 PJ/a, el
66% del cual es superior a 140°C.
Connolly et al. (2013) [30] aplican datos basados en instalaciones de emisión y sistemas
de información geográfica (GIS) para determinar el exceso de calor disponible con vistas a
una posible integración en los sistemas de calefacción urbana. Estiman un potencial de
525 PJ/a en Alemania y 2.708 PJ/a en la UE-27. Estos valores fueron ligeramente corregi-
dos por Persson et al. [31], según los cuales el exceso de calor industrial evaluado en la
UE-27 ascendía a 2.924 PJ/a, 566 PJ/a de los cuales en Alemania.
Campana et al. (2013) [18] estiman el potencial de la recuperación de calor residual con
ORC en la industria europea (UE-27) energéticamente intensiva. Se cuantifican los ahorros
en los costos de electricidad y la emisión de GEI. El estudio se basa en el análisis de las
auditorías energéticas disponibles. Las auditorías energéticas proporcionan la potencia
ORC instalada, o potencialmente instalable, en una planta considerada analizada durante
un estudio de viabilidad.
Recientemente, Papapetrou et al. (2018) [2] basaron su metodología en la estimación
de las fracciones de calor residual por industria, país de la UE-28 y nivel de temperatura.
El punto de partida es un estudio de la industria del Reino Unido en el que se estimó el
calor residual técnicamente disponible de cada sector industrial en el Reino Unido utili-
zando datos de 425 emplazamientos industriales del período 2000-2003. Los resultados,
utilizando un enfoque “top-down”, estimaron un calor residual industrial para Alemania de
269,5 PJ/a.
Sin embargo, Brueckner et al. (2017) [32] presentaron el primer enfoque “bottom-up”
para estimar el potencial de calor residual industrial en Alemania. Este enfoque utiliza los
datos del informe de emisiones de CO2 de las empresas productoras alemanas para cal-
cular un valor límite conservador para el calor residual industrial. El volumen de calor resi-
dual industrial fue evaluado como 127 PJ/a. Asumiendo una proporción similar de calor
residual en el resto de la industria, se evalúa un potencial de 223 PJ/a.
En la Figura 6.1 se presenta un resumen del potencial de calor residual industrial en Ale-
mania. Las estimaciones varían entre 127 PJ/a y 566 PJ/a.

Figura 6.1. Estimaciones del potencial de calor residual de la industria en Alemania por va-
rios autores en orden ascendente
4

Potencial de la recuperación de calor residual con


ORC en las industrias con uso intensivo de energía
en Alemania
Para cuantificar el impacto de la recuperación de calor residual de industrias intensivas
en energía con sistemas ORC en Alemania, se hace un análisis del potencial enfocado a
las industrias del acero, cemento y vidrio. Se incluye un análisis de sensibilidad para tener
en cuenta las incertidumbres de los datos. Los resultados resaltan la importancia y la ren-
tabilidad de esta opción para mejorar la eficiencia energética de la industria.
Descripción de los procesos y estimación del calor residual disponible en las plantas de
producción de acero, cemento y vidrio
Según [12], el 53 % del calor residual está disponible en la industria de fabricación de
metales básicos (por ejemplo, acero y aluminio) y en la industria de minerales no metálicos
(por ejemplo, vidrio, cerámica y cal). Como representante de estas categorías, se investi-
gan los principales procesos de fabricación de acero, cemento y vidrio, y se destacan las
principales fuentes de calor residual.

Acero
El acero se produce principalmente a través de dos rutas: a partir de mineral de hierro
(siderurgia de oxígeno) o a partir de chatarra metálica (siderurgia eléctrica). A nivel mundial,
la primera ruta supone dos tercios de los procesos de producción y la segunda el tercio
restante. La Figura 6.2 muestra los principales pasos necesarios para convertir la materia
prima en producto final de acero. En la ruta de fabricación de acero al oxígeno, el mineral
de hierro se alimenta típicamente en forma de (hermatita) Fe 2O3 al alto horno (BF) junto
con agentes reductores (coque, carbón, petróleo y gas) y aditivos (cal, piedra caliza). De
esta manera, el mineral de hierro se reduce a hierro metálico (también llamado arrabio). El
arrabio caliente rico en carbono (1500°C) es transportado por cucharas torpedo al alto
horno de oxígeno (BOF, también conocido como convertidor de oxígeno), donde las impu-
rezas (C, Si, Mn, S y P) se eliminan mediante el soplado de oxígeno y gas inerte. En la ruta
de la siderurgia eléctrica, la chatarra metálica se funde mediante electrodos y carbón de
carga en un horno de arco eléctrico (EAF). También se utilizan quemadores de gas para
apoyar el proceso de fusión.
El acero líquido producido en los BOF y EAF se funde y solidifica para su posterior
procesamiento. Alrededor del 96% del acero líquido en Alemania se somete a un proceso
de colada continua [45]. El resto se funde en lingotes. Las operaciones de fundición se
encuentran al final de las plantas siderúrgicas integradas (BF y BOF) o EAF y justo antes
de las operaciones de procesamiento posterior (en el tren de laminación o en la planta de
forja). La temperatura del acero debe ser aumentada hasta una temperatura de laminación
adecuada (300-1050°C) antes de ser procesado por los laminadores en caliente a través
de un horno de recalentamiento (RHF), normalmente alimentado con gas natural. La ma-
yoría de las plantas estudiadas tienen o bien hornos de empuje o tipo “walking-beam”. Los
Appendix v

hornos de empuje son una alternativa barata a los más sofisticados de tipo "walking-hearth"
o "walking-beam".

Figura 6.2. Rutas de producción de acero: acería al oxígeno (BF y BOF) y acería eléctrica
(EAF). También se observa el horno de recalentamiento (pusher-type furnace) [37]

Los procesos de los que se puede obtener calor residual para su recuperación son los
que ocurren en el BOF, EAF y RHF, donde los gases salen de las plantas a temperatura
media/alta (>400°C). Los gases calientes también salen de los BF, pero dado su conside-
rable valor calorífico (3-3,5 MJ/Nm3, [92]), típicamente se recuperan, se limpian y se utilizan
como combustible en otras partes de la planta [98]. Varias unidades ORC ya han sido ins-
taladas en EAFs y RHFs, pero en BOFs todavía no. En estos últimos, el gas de combustión
rico en CO puede ser quemado en el canal de escape o puede ser enfriado, limpiado y
almacenado para su uso posterior. Para este trabajo, se considera un total de 9 plantas
con BOF, 31 con EAF y 31 con RHF.
En la Figura 6.3 se muestra la distribución de las fábricas de acero más importantes de
Alemania. Los datos de [22] se utilizan para determinar la capacidad de cada planta. La
diferencia entre la capacidad de la planta y la producción anual se define mediante un factor
de carga que es, en promedio, de 82,2%, 82,9% y 89,4% para el BOF, EAF y RHF respec-
tivamente. La producción total ascendió a 43,3 millones de toneladas de acero en 2017
[90].
La Tabla 6.1 resume la información sobre el calor residual disponible en los gases de
escape en varias etapas de la producción de acero. Debido a la dispersión del calor residual
específico disponible en la literatura, se definen tres escenarios: el más pesimista (worst-
case scenario o WCS), el promedio (average scenario) y el más optimista (best-case sce-
nario o BCS). Dada la intermitencia del gas de escape y la fluctuación del contenido de CO,
se requiere un almacenamiento intermedio de gas, lo cual puede causar altos costos [26].
Por este motivo, la combustión directa de gas y la recuperación de calor pueden ser una
mejor opción.
6

Integrated steel mills Electric arc furnaces


(Blast furnace, steelmaking and
rolling mills)
Year: 2014

Figura 6.3. Acerías en Alemania [29]


Tabla 6.1. Calor residual disponible de la producción de acero, referido a 15°C [77], [78],
[79], [80], [81], [82], [83], [84], [85], [86], [87]

Calor residual disponible


Fuente de ca-
Rango de
lor residual
temperaturas Características Total (𝑮𝑱/𝒕)
(gases de es- Sensible Químico
(°𝑪)
cape) (𝑮𝑱/𝒕) (𝑮𝑱/𝒕)
WCS AS BCS

Contienen com-
0.48- 3.40-
BF 120-320 bustibles, par- 3.88 5.21 6.54
0.54 6.00
tículas, etc.

Contienen com-
0.15- 0.84-
BOF 1200-1500 bustibles, par- 0.99 1.02 1.04
0.18 0.86
tículas, etc.

Contienen com-
0.40- 0.44-
EAF 1200-1800 bustibles, par- 0.44 0.73 1.02
0.51 0.57
tículas, etc.

0.15-
RHF 400-550 Gas limpio 0.00 0.15 0.23 0.30
0.30
Appendix vii

Cemento
El sector del cemento es particularmente intensivo en emisiones de CO2, contribuyendo
al 2,5 % de las emisiones totales en Alemania en 2017 [99]. Las materias primas utilizadas
son piedra caliza, tiza, pizarra, arcilla, arena y aditivos menores. El proceso de fabricación
se resume en el esquema de la Figura 6.4. El tamaño de las partículas sólidas se reduce
mediante la trituración y la molienda. Los diferentes materiales son posteriormente mezcla-
dos para obtener la composición química necesaria. La materia prima (principalmente cal
y arcilla) se quema y reacciona para transformarse en clínker utilizando hornos rotativos
alimentados con combustibles fósiles. La materia prima se precalienta primero. El clínker
caliente que sale del horno se enfría de 1450°C a unos 100°C mediante un enfriador de
aire de tipo rejilla, molido con yeso y otros aditivos menores para producir el cemento. La
mitad del aire de enfriamiento se utiliza como aire de combustión precalentado para el
quemador del horno. La otra mitad del aire de enfriamiento se ventila [57].
El calor residual restante de los escapes del precalentador y del enfriador de clínker
puede ser recuperado y utilizado para dos cosas diferentes: o bien para proporcionar las
necesidades de calefacción a baja temperatura en la planta o para generar energía para
compensar una parte de la energía comprada a la red, o energía cautiva generada por el
consumo de combustible en el sitio [57].Típicamente, las plantas de cemento apenas re-
quieren calefacción a baja temperatura, por lo que la mayoría de los proyectos de recupe-
ración de calor se han centrado en la generación de energía.

Figura 6.4. Representación esquemática del proceso de fabricación de cemento desde la


cantera hasta el envío [57]

Las plantas de cemento que forman parte de la Asociación Alemana de las Plantas de
Cemento (VDZ e.V.) se representan en la Figura 6.5, que incluye 46 de las 53 plantas de
cemento en Alemania [31]. De éstas, sólo el 76% producen realmente clínker (y tienen
disponible el calor residual correspondiente), mientras que las restantes son principalmente
molinos. La producción de estas plantas alcanzó los 31 millones de toneladas de cemento
en el año 2015 [91].
8

Brunke [92] recopiló datos sobre la capacidad y producción de cada planta en 2013,
donde se estima un factor de carga del 85,9%. Estos datos (se consideran 34 plantas) son
los que se utilizan para el presente análisis. La Tabla 6.2 resume las cifras cuantitativas
sobre el calor residual disponible.

Figura 6.5. Ubicación de las plantas de cemento alemanas que forman parte de la VDZ [91]

Tabla 6.2. Calor residual disponible de la producción de cemento, referido a 15°C [83], [78],
[65]

Calor residual dispo-


Rango de
Fuente de calor re- nible (𝑮𝑱/𝒕)
temperaturas Características
sidual
(°𝑪)
WCS AS BCS

Aire del enfriador de


250-350 Gas limpio 0.30 0.41 0. 52
clínker

Precalentador de
Contiene combustibles, par-
clínker (TPD 1000- 390 0.92 0.92 0.92
tículas, etc.
2000)

Precalentador de Contiene partículas, etc. Re-


clínker (TPD 2000- 200-400 lativamente fácil de manejar. 0.60 0.71 0.81
8000)
Appendix ix

Vidrio
Existen diferentes tipos de vidrio en función de su composición química, su método de
producción o su comportamiento ante el procesamiento. Si se hace una clasificación en
función de su composición química, existen tres tipos: vidrio sodocálcico, vidrio de plomo y
vidrio borosilicato. Sin embargo, el vidrio sodocálcico es el que se produce en mayor can-
tidad con diferencia [67]. Este tipo, que contiene 71-75% SiO2, 12-16% Na2O, 10-15% CaO
y pequeñas cantidades de otras sustancias como colorantes, se utiliza para fabricar enva-
ses y productos de vidrio laminado [67]. Estas sustancias se añaden a la arena (SiO2) para
reducir su temperatura de fusión (que es superior a 2000ºC) y para mejorar las propiedades
del material. Para la recuperación del calor residual, en este trabajo se consideran dos tipos
de fabricación de vidrio: el vidrio flotado y el vidrio para envases. Estos tipos son, con dife-
rencia, los más producidos [67]. A continuación, se describe brevemente el proceso de
producción. Como se ha observado, los procesos para el vidrio flotado y el vidrio para en-
vases son bastante similares hasta la parte de conformación.
En el proceso de producción de vidrio flotado (Figura 6.6), los ingredientes se mezclan
primero entre sí y también con cullet (vidrio reciclado triturado) y luego se calientan en un
horno mediante sistemas eléctricos, de gas o de combustión de aceite a unos 1600ºC para
formar el vidrio fundido. A continuación, se somete a un proceso de refinado. El vidrio fun-
dido se introduce en la parte superior de un baño de estaño fundido. Se produce entonces
una cinta de vidrio plano de espesor uniforme haciendo fluir el vidrio fundido en el baño de
estaño bajo un calentamiento controlado. Después del baño de estaño, el vidrio se enfría
gradualmente, se controla y se corta en láminas o se convierte en formas que puedan ser
moldeadas en contenedores [69].

Figura 6.6. Esquema del proceso de producción del vidrio flotado [68]

Parte de la energía térmica de los gases de combustión se utiliza para precalentar el


aire alimentado a los quemadores del horno. El resto puede ser recuperado, por ejemplo,
por una unidad ORC. El calor residual disponible de un horno de fusión de vidrio puede
variar entre 1,05 y 2,64 GJ/t (a 15ºC), como se resume en la Tabla 6.3. Los datos de 56
plantas de vidrio para envases y 15 plantas de vidrio plano se han tomado de [92]. Los
datos se refieren al año 2013 y resultan en un factor de carga de 87,9% para el vidrio para
envases y 71,4% para el vidrio plano.
10

Tabla 6.3. Calor residual disponible de la producción de vidrio, referido a 15°C [78], [8], [83]

Calor residual dispo-


Rango de
Fuente de calor re- nible (𝑮𝑱/𝒕)
temperaturas Características
sidual
(°𝑪)
WCS ACS BCS

Horno de fusión de Contiene partículas, vapores


400-500 1.05 1.85 2.64
vidrio condensables, etc.

En los sistemas con recuperador, el gas de combustión sale de este componente a 400-
500ºC. Se debe prestar atención para evitar la condensación ácida de los gases de com-
bustión (<200ºC) y las partículas contenidas en los gases de combustión que pueden cau-
sar el ensuciamiento de los intercambiadores de calor [83].

A.5 Conversión del calor residual con la tecnología ORC


La Figura 6.7 muestra el esquema básico de un sistema de recuperación de calor resi-
dual con la tecnología de ORC. Un lazo intermedio se usa muy a menudo para desacoplar
la operación del proceso industrial de la operación de la unidad ORC. Para el calor residual
de BOF y EAF, se utiliza un sistema de recuperación de agua/vapor como bucle intermedio,
ya que la industria siderúrgica no acepta el aceite térmico como portador de calor debido a
los problemas de inflamabilidad. Para limitar las fluctuaciones del calor residual, se puede
incluir un acumulador de vapor. En particular, la empresa TENOVA ha desarrollado este
tipo de sistemas [100]. En la planta de acero Feralpi EAF en Riesa (Alemania), se ha aco-
plado un acumulador de vapor a un sistema ORC de 2,6 MW e [93]. Más plantas de este
tipo han sido construidas en los últimos años. Sin embargo, para la recuperación del calor
de los hornos de recalentamiento de los trenes de laminación en caliente, se ha instalado
un sistema de evaporación directa en una planta de producción de acero en Singapur, dado
que los gases no fluctúan tanto y la temperatura es más baja. La evaporación directa es
menos costosa y más atractiva siempre que sea posible.
Las plantas de cemento suelen transportar el calor residual de los precalentadores de
clínker y de los enfriadores de parrilla por medio de aceite térmico. En el marco del proyecto
TASIO [101] se ha realizado un primer intento de conseguir la evaporación directa en esta
industria. Dado que la evaporación directa se encuentra sus primeras fases de desarrollo,
las plantas de vidrio tienen en general un circuito intermedio de aceite térmico [8].
Appendix xi

Figura 6.7. Esquema general de un sistema de recuperación de calor industrial con ORC y
circuito intermedio [23]

A.6 Metodología
En este trabajo se estiman los potenciales teórico, técnico y económico de la implemen-
tación de la tecnología ORC para la recuperación del calor residual de los gases de escape
de algunos procesos de las industrias energéticamente intensivas más importantes de Ale-
mania. Para ello, en primer lugar, se estima el calor residual de cada proceso en cada
planta. Luego, teniendo en cuenta la eficiencia de la recuperación, se estima la producción
neta de energía eléctrica. Con ello se determina el potencial técnico. Para el potencial eco-
nómico, se deben considerar los costos de ORC.
Se deben definir estos tres tipos de potenciales [15]:
• Potencial teórico o físico: considera las limitaciones físicas. Por ejemplo, solo se
considera el calor residual por encima de la temperatura ambiente.
• Potencial técnico: tiene en cuenta las limitaciones relacionadas con los aspectos
técnicos. Por ejemplo, la temperatura mínima de salida del gas para evitar la
condensación de ácidos y, por lo tanto, la corrosión. Este potencial depende de
las tecnologías utilizadas.
• Potencial económico: considera parámetros financieros como los precios de la
energía, los tipos de interés, etc. para ver si el proyecto es económicamente
viable y rentable.
En esta tesis se utiliza una combinación de los métodos “bottom-up” y “top-down”. El
método “bottom-up” se utiliza para la obtención de los datos de los calores residuales es-
pecíficos de diferentes fuentes bibliográficas con el fin de estimar un valor asociado a cada
12

proceso. El método “top-down” se aplica cuando este valor común a cada proceso se mul-
tiplica por el tamaño de cada planta (en toneladas producidas) para obtener el calor residual
total que se obtiene de cada planta.

Potencial técnico
El potencial técnico de esta tecnología depende de la eficiencia de la conversión en
energía eléctrica del calor residual que llega de la industria y de las horas en las que está
disponible la unidad ORC. La eficiencia de conversión total (𝜂𝑡𝑜𝑡 ) está relacionada con la
tasa de utilización de la fuente de calor (𝜂𝑢 ) y la eficiencia eléctrica neta de la unidad ORC
(𝜂𝑒𝑙,𝑛𝑒𝑡,𝑂𝑅𝐶 ):

𝜂𝑡𝑜𝑡 = 𝜂𝑢 𝜂𝑒𝑙,𝑛𝑒𝑡,𝑂𝑅𝐶 ( 6.1)

La temperatura de salida (la temperatura de los gases de combustión a la salida del


sistema de recuperación de calor) se considera en torno a 150°C, lo que garantiza que se
evite la condensación de los gases de combustión y, por lo tanto, la corrosión. Esta tasa
es necesaria porque el calor residual teórico se refiere a menudo a la temperatura am-
biente. Esta temperatura ambiente puede variar dependiendo de la literatura. El índice de
utilización es una función de la temperatura de la fuente de calor que sale del sistema de
recuperación, y se define como:

𝑇ℎ𝑠,𝑖𝑛 − 𝑇ℎ𝑠,𝑜𝑢𝑡
𝜂𝑢 = ( 6.2)
𝑇ℎ𝑠,𝑖𝑛 − 𝑇𝑟𝑒𝑓

Siendo 𝑇ℎ𝑠,𝑖𝑛 la temperatura de entrada de los gases que contienen el calor residual de
la fuente (proceso industrial) en el sistema de recuperación, 𝑇ℎ𝑠,𝑜𝑢𝑡 la de salida (depen-
diendo del proceso será diferente para que los gases de combustión no condensen) y 𝑇𝑟𝑒𝑓
la temperatura de referencia, que es a la que salen los gases según las diversas fuentes y
que suele ser la ambiente.
La Tabla 6.4 muestra las tasas de utilización supuestas para la producción de acero,
cemento y vidrio. La eficiencia eléctrica neta de las unidades ORC se basa en las tablas
de especificaciones de los turbogeneradores ORC de la compañía Turboden, que es la
empresa líder en recuperación de calor residual con ORC de procesos industriales [72].
Todas las unidades consideradas están equipadas con un bucle de aceite intermedio. La
eficiencia eléctrica neta es en general una función de la temperatura de la fuente de calor
y el tamaño de la planta. Debido a la estandarización del sistema y a las limitaciones en la
temperatura del lazo intermedio, la eficiencia neta nominal es para todas las unidades Tur-
boden de 16-21%. En la operación real, especialmente para las unidades enfriadas por
aire, este valor cae en algún punto porcentual.
Appendix xiii

Tabla 6.4. Tasa de utilización de la fuente de calor para los diferentes procesos
Temperatura de los
gases de escape a Temperatura de sa-
Fuente de calor re- Tasa de utilización
su entrada en el sis- lida de los gases a la
sidual (-)
tema de recupera- atmósfera (°𝑪)
ción (°𝑪)
BOF 1200 150 0.89

EAF 1200 150 0.89

RHF 400 150 0.65

Aire del enfriador de


300 100 0.70
clínker

Precalentador de
clínker (TPD 1000- 390 150 0.64
2000)
Precalentador de
clínker (TPD 2000- 300 150 0.53
8000)
Horno de fusion de vi- 400 200 0.52
drio

Además, también se considera la disponibilidad del ORC (𝜏) suele ser de un 95% con
respecto al número de horas de disponibilidad del flujo de calor residual para la producción
de energía. De esta manera, la energía eléctrica que podría ser generada con una unidad
de ORC (E) a partir del calor residual disponible (Qresidual) pasaría a calcularse como:

𝐸 = 𝑄𝑟𝑒𝑠𝑖𝑑𝑢𝑎𝑙 ∙ 𝜂𝑡𝑜𝑡 ∙ 𝜏 ( 6.3)

Potencial económico
El atractivo de esta tecnología depende en gran medida de su viabilidad económica.
Una vez que el potencial técnico es estimado, el potencial económico puede ser calculado
en base al método utilizado por Eyerer, et al. (2017) [33] para la energía geotérmica en su
trabajo "Potencial de la geotermia hidrotermal para la generación de energía en Alemania".

Costo de inversión específico


Para realizar un análisis económico-financiero de un proyecto, se deben considerar los
flujos de caja anuales, compuestos por la inversión de capital y los gastos e ingresos anua-
les. La inversión de capital tiene lugar al inicio del proyecto. Incluye los costes directamente
asociados al sistema (materiales, equipos, mano de obra, etc.) y los costes indirectos (in-
geniería, costes de construcción y contingencias). La parte principal de la inversión de ca-
14

pital de un proyecto ORC son los componentes del propio módulo ORC: evaporador, ex-
pansor y generador, condensador y bomba. Otra parte importante son los costes de inte-
gración del módulo ORC en una planta existente (por ejemplo, para aplicaciones de recu-
peración de calor). Los gastos anuales de los proyectos ORC dependen del tipo de aplica-
ción y de la ubicación, pero su importancia suele ser menor que los costes de inversión.
Para calcular el coste de inversión específico (SIC) [€ 2018/kWe], se utiliza una correlación
entre el tamaño del ORC y el SIC (Ecuación ( 6.4) basado en los productos de Turboden
obtenida de [34]. En este trabajo, todas las unidades ORC consideradas están equipadas
con un bucle de aceite intermedio. Los costos de inversión específicos pueden variar entre
4000 y 2000 USD/kWe para las unidades más grandes [34]. En la Figura 6.8, se puede
observar que cuanto más grande es el módulo ORC, más bajo es el SIC. Una vez se co-
nozca la potencia de la unidad ORC estimada a partir del calor residual y considerando las
limitaciones técnicas, se podrá calcular el coste de inversión específico para cada unidad.

Figura 6.8. SIC function [34]

€2018
𝑆𝐼𝐶 [ ] = 19774 · 𝑃𝑂𝑅𝐶 [𝑘𝑊𝑒 ]−0.277 ( 6.4)
𝑘𝑊

Costo nivelado de la energía


Para el análisis del potencial económico, se calculará el Costo Nivelado de Energía
(LCOE) para cada unidad ORC. El LCOE mide los costos de la vida útil divididos por la
producción total de energía y calcula el valor actual del costo total de la construcción y
operación de una planta de energía durante una vida útil supuesta. Este indicador es crucial
para tomar la decisión de proceder o no al desarrollo de un proyecto.

𝐶𝑀𝑙
𝑆𝐼𝐶 · 𝑃𝑛 + ∑𝑛𝑙=1
(1 + 𝑖)𝑙
𝐿𝐶𝑂𝐸 = ( 6.5)
𝐸𝑙
∑𝑛𝑙=1
(1 + 𝑖)𝑙

"SIC" es el costo específico de inversión, ”P n“ la potencia nominal, "n" la vida útil su-
puesta, "l" el año en curso, "i" la tasa de descuento, ”CMl“ los costos de mantenimiento y
Appendix xv

“El“ la producción de energía en el año "l". El SIC multiplicado por la potencia nominal es
la inversión inicial.
El coste de inversión es, pues, el producto del costo específico de inversión (SIC) y la
potencia nominal de cada unidad ORC. Esta potencia nominal es calculada partiendo del
calor residual disponible según el potencial técnico, es decir, teniendo en cuenta la eficien-
cia de conversión y la disponibilidad de la unidad ORC. Una vez conocida la potencia, se
puede conocer la producción anual de energía “El“. Para calcular el LCOE solo faltaría
conocer cuáles son los costes de operación y mantenimiento (”CMl“). Para ello, se hace e
se ha hecho el siguiente supuesto basado en [35]:

𝐶𝑀𝑙 = 2% · 𝑆𝐼𝐶 · 𝑃𝑛 ( 6.6)

La tasa de interés (i) y los años de amortización (n) serán variados para el análisis del
potencial económico.

ORC LCOE vs. precios de electricidad para EII


Una vez que se obtiene el LCOE para cada unidad ORC, debe ser comparado con el
Costo Nivelado de la Electricidad para la industria en Alemania [5]. Este precio de la elec-
tricidad es diferente para los hogares y para la industria, ya que se aplican exenciones de
tasas e impuestos [32]. Estas exenciones son diferentes dependiendo del tamaño del em-
plazamiento.
Considerando todos los impuestos y tasas, reducciones y exenciones para las industrias
intensivas en electricidad, el precio medio para el año 2018 y para los grandes clientes
industriales (consumo de energía anual superior a 10 GWh/a) según el BDEW [5] es de
8,84 ct/kWh. El precio mínimo va de 5,10 a 5,80 ct/kWh y el precio máximo oscila entre
14,40 y 17,00 ct/kWh (Figura 6.9).

Figura 6.9. Precios de la electricidad para las industrias intensivas en consumo de energía
(consumo anual>10GWh) [5]

Cuando el LCOE de una unidad ORC es inferior al precio de la electricidad que se con-
sidere para la planta en cuestión, resulta económicamente rentable instalarla. Si se suma
la electricidad que podría ser producida con las unidades que son rentables para cada
16

industria (acero, cemento y vidrio), se obtiene el potencial económico de esta tecnología.


En algunas ocasiones, cuando la situación es muy favorable, el potencial económico sería
muy cercano al técnico.

Resumen de la metodología
Aunque ya se han explicado un gran número de detalles sobre la metodología seguida
en este trabajo, a continuación se presenta una visión general del método con los pasos
de forma sintetizada y ordenada:
1. Identificación de las industrias intensivas en consumo de energía en Alemania.
2. Identificación de los procesos de alto consumo de energía en cada industria que
pueden ser utilizados para la recuperación del calor residual.
3. Identificación de las plantas dentro de cada industria.
4. Recopilación de información sobre los procesos, cómo se lleva a cabo la recupera-
ción de calor residual, la capacidad y producción anual de las plantas.
Potencial teórico

5. Estimación del calor residual específico que se puede obtener para cada proceso.
a. Investigación sobre el calor residual de los procesos y el consumo de ener-
gía.
b. Cálculo del calor residual específico en un escenario pesimista, intermedio
y optimista para cada proceso.
6. Cálculo del calor residual total teórico anual de cada planta para los tres escenarios,
multiplicando el específico por la producción anual de la planta.
Potencial técnico

7. Estimación de la generación de electricidad con ORC para cada planta teniendo en


cuenta las limitaciones técnicas como la eficiencia eléctrica neta de la unidad ORC,
la tasa de utilización del calor y la disponibilidad anual de la unidad ORC.
Potencial económico

8. Cálculo del SIC en función del tamaño de la unidad ORC con la correlación de [34]
y, posteriormente, el LCOE asociado a cada unidad de ORC.
9. Elaboración de gráficos que relacionan la energía eléctrica agregada producida por
cada ORC potencialmente instalable en Alemania con el LCOE de esas unidades
ORC.
10. Elaboración de un análisis de sensibilidad para determinar cuántas unidades ORC
son económicamente viables de instalar.

A.7 Resultados
En este apartado se presentan y analizan los resultados de la estimación sobre el po-
tencial teórico, técnico y económico de la recuperación de calor residual con ORC para las
industrias alemanas de acero, cemento y vidrio de uso intensivo de energía.
Appendix xvii

Potencial teórico
El calor residual teórico (energía en el gas residual) disponible en las industrias del
acero, el cemento y el vidrio en Alemania ha sido estimado en base al método descrito en
este trabajo.
Los resultados para los diferentes escenarios (WCS, AS y BCS) se muestran en la Fi-
gura 6.9. Precios de la electricidad para las industrias intensivas en consumo de energía
(consumo anual>10GWh) para cada industria. El potencial teórico total referido a 15ºC se
sitúa entre 11,5-15,6 TWh/a para el acero, 7,9-9,3 TWh/a para el cemento y 1,7-4,3 TWh/a
para el vidrio.
Tabla 6.5. Potencial teórico del calor residual disponible de las industrias del acero, el ce-
mento y el vidrio en Alemania referido a 15°C

Calor residual disponible a 15°C (𝑻𝑾𝒉/𝒂)


Fuente de calor residual
WCS AS BCS

BOF 8.33 8.58 8.75

EAF 1.59 2.65 3.70

RHF 1.56 2.40 3.13

Acero (total) 11.48 13.63 15.58

Cemento 7.87 8.56 9.27

Vidrio 1.73 3.04 4.34

Potencial técnico
El potencial técnico se define como la energía eléctrica neta que se puede obtener de
la fuente de calor residual con una unidad ORC teniendo en cuenta limitaciones técnicas
(uso limitado de la fuente de calor, eficiencia de recuperación, disponibilidad anual de la
unidad de ORC). Este se resume en la Tabla 6.6. Se estudian los casos en los que se
considera una eficiencia de la unidad ORC del 15% (más conservador) y del 19% (más
optimista). En la Tabla 6.7 se muestra la potencia acumulada del potencial técnico para
cada fuente de calor residual, según la eficiencia del ORC y el escenario.
18

Tabla 6.6. Potencial técnico de calor residual disponible de las industrias del acero, el ce-
mento y el vidrio en Alemania con una disponibilidad de ORC del 95%

Potencial técnico (𝑻𝑾𝒉/𝒂)

Fuente de calor Eficiencia del ORC del Eficiencia del ORC del
residual 15% 19%

WCS AS BCS WCS AS BCS

BOF 1.05 1.08 1.11 1.33 1.37 1.40

EAF 0.20 0.33 0.47 0.26 0.42 0.59

RHF 0.14 0.22 0.29 0.18 0.28 0.37

Acero (total) 1.39 1.63 1.87 1.77 2.07 2.36

Cemento 0.70 0.86 1.02 0.88 1.09 1.30

Vidrio 0.13 0.23 0.32 0.16 0.29 0.41

Tabla 6.7. Capacidad acumulativa del potencial técnico de la industria intensiva en energía
en Alemania con una disponibilidad de ORC del 95%

Potencia acumulada del ORC (𝑴𝑾𝒆 )

Fuente de calor Eficiencia del ORC del Eficiencia del ORC del
residual 15% 19%

WCS AS BCS WCS AS BCS

BOF 120.1 123.7 126.2 152.1 156.7 159.8

EAF 23.0 38.1 53.3 29.1 48.3 67.5

RHF 16.5 25.3 33.0 20.9 32.1 41.9

Acero (total) 159.6 187.1 212.5 202.1 237.1 269.2

Cemento 79.6 98.3 116.9 100.8 124.5 148.1

Vidrio 14.6 25.7 36.7 18.5 32.5 46.4


Appendix xix

Potencial económico
Para la evaluación del potencial económico se utiliza un método inspirado en [33]. Este
método consiste en la elaboración de una curva producción-coste para la generación de
electricidad con ORC a partir del calor residual. En el eje Y se representa la energía eléc-
trica acumulada. Esto es, la electricidad que podría generarse al año con cada unidad ORC
sumada de manera acumulativa empezando por aquella con un menor LCOE y añadiendo
según va aumentando el LCOE hasta alcanzar el potencial técnico. En el eje X se repre-
senta el Costo Nivelado de Energía (LCOE) de cada ORC para el año 2018.
Una vez presentada la curva, la rentabilidad económica se determinará por la energía
eléctrica total de aquellas unidades de ORC que tengan un LCOE inferior al precio de la
electricidad para la industria en Alemania en 2018. Como se ha visto anteriormente en esta
tesis, existe un precio mínimo y máximo de la electricidad para la industria intensiva en
energía dependiendo de las exenciones de tasas e impuestos que es diferente para cada
planta dependiendo de su tamaño. Para estas industrias, el precio en 2018 fue de 5,1-17,0
ct/kWhe en Alemania. En los gráficos presentados en este trabajo, aparecen los precios
mínimos, medios y máximos.
La Figura 6.10 muestra los resultados para el acero (BOF, EAF y RHF) para la eficiencia
más conservadora del ORC (15%), una disponibilidad del ORC del 95%, la tasa de interés
al 4% y 10 años de amortización. En el Apéndice F, se muestran otras gráficas variando
algunos de los parámetros (eficiencia del 19%, tasa de interés al 6%, 5 años de amortiza-
ción) con el fin de crear diferentes combinaciones y analizar diferentes escenarios posibles.
Sin embargo, en el caso de los años de amortización, por ejemplo, 10 años es más realista
que 5 años.
Si se analizan las gráficas para el acero, se puede ver que el LCOE para las plantas
BOF es significativamente menor que el precio más bajo de la electricidad (5,1 ct/kWh).
Esto significa que la instalación de una unidad ORC es económica para cada planta del
BOF y el potencial económico corresponde al potencial técnico (1052-1105 GWh/a).
La mayoría de las unidades ORC que podrían ser instaladas en plantas con un EAF
tienen un LCOE inferior al precio de la electricidad más bajo. El potencial económico es de
164-455 GWh/a dependiendo del escenario considerado y está relativamente cerca del
potencial técnico.
Para las plantas con RHF, la factibilidad es todavía positiva pero menos favorable que
la del BOF y la EAF. La diferencia entre el potencial técnico y el económico es mayor para
el precio más bajo de la electricidad que en el caso del EAF y el BOF. En el caso del RHF,
este potencial está entre 85 y 222 GWh/a. Además, en la Figura 6.10 se puede observar
que, para todos los BOF, EAF y RHF, el potencial económico para los precios promedio y
más altos de la electricidad prácticamente coincide con el técnico.
En el caso del cemento (Figura 6.11), el potencial económico es casi igual al técnico
para cada escenario en todo el rango de precios de la electricidad, incluso para el más
bajo. En este caso, el potencial económico es de 685-1025 GWh/a. En el caso del vidrio,
es interesante observar que, en el peor de los casos, la instalación de una unidad ORC
para recuperación del calor residual no sería económica para ninguna planta si el precio
de electricidad es inferior a 5,1 ct/kWh. En los otros dos escenarios, el potencial económico
para el precio más bajo está lejos de ser igual al técnico, aunque crece progresivamente
20

hasta que son prácticamente iguales para un precio promedio de electricidad de 8,84
ct/kWh.

(a) (b)

(c)

Figura 6.10. Potencial económico para la producción de acero con 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%,
n=10 años
Appendix xxi

(a) (b)
Figura 6.11. Potencial económico para la producción de: (a) cemento y (b) vidrio con
𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 años

Los potenciales económicos para el menor precio de la electricidad, i=4%, n=10 años,
disponibilidad del ORC del 95% y una eficiencia del ORC del 15% y 19% se presentan en
la Tabla 6.8. En la Tabla 6.9 se muestra la correspondiente potencia acumulada.

Tabla 6.8. Potencial económico del calor residual disponible de la industria intensiva en
energía con disponibilidad de ORC al 95%, tasa de interés al 4% y 10 años de amortización

Potencial económico (G𝑾𝒉/𝒂) referido a 5.1 ct/kWhe

Fuente de calor
Eficiencia del ORC del 15% Eficiencia del ORC del 19%
residual

WCS AS BCS WCS AS BCS

BOF 1052.1 1084.0 1105.2 1332.6 1373.0 1399.9

EAF 163.8 319.0 454.7 231.4 412.2 575.9

RHF 84.8 156.4 221.9 116.1 215.5 310.9

Acero (total) 1300.7 1559.4 1781.8 1680.1 2000.7 2286.7

Cemento 684.9 854.7 1024.5 883.0 1090.5 1297.7

Vidrio 0.0 21.3 120.5 0.0 94.5 227.8


22

Tabla 6.9. Capacidad acumulada del potencial económico de la industria intensiva en ener-
gía con disponibilidad de ORC al 95%, tasa de interés al 4% y 10 años de amortización

Potencia acumulada correspondiente al potencial económico (𝑴𝑾𝒆 )

Fuente de calor
Eficiencia del ORC de 15% Eficiencia del ORC de 15%
residual

WCS AS BCS WCS AS BCS

BOF 120.1 123.7 126.2 152.1 156.7 159.8

EAF 18.7 36.4 51.9 26.4 47.1 65.7

RHF 9.7 17.9 25.3 13.3 24.6 35.5

Acero (total) 148.5 178.0 203.4 191.8 228.4 261.0

Cemento 78.2 97.6 116.9 100.0 124.5 148.1

Vidrio 0.0 2.4 13.8 0.0 10.8 26.0

A.8 Análisis de los resultados


Para las hipótesis de las gráficas anteriores, las conclusiones sobre el potencial
económico son, para la mayor parte de los procesos, positivas. Se considera para ello el
menor precio de la electricidad (situación más conservadora).
Para la industria del acero, el potencial económico de la instalación de unidades ORC
en hornos BOF es igual al técnico bajo cualquier hipótesis y escenario, siendo este de en
torno a 1100 GWh/a para una eficiencia del ORC del 15% y entre 1300 y 1400 GWh/a para
una eficiencia del 19%. Los EAF son, también, bastante interesantes de considerar para la
instalación de sistemas ORC de recuperación de calor de sus gases de escape. El
potencial técnico varía según el escenario determinado por el calor residual disponible
(peor, intermedio y mejor caso) entre 200 y 470 GWh/a para una eficiencia del ORC del
15% y entre 250 y 600 para una eficiencia del 19%. Para una eficiencia del 15%, el
potencial económico es un poco mayor al 80% del potencial técnico para el peor escenario,
y un 97% para el mejor escenario. Para el caso con una eficiencia del 19%, el potencial
económico se encuentra ligeramente más cerca del técnico. Para los RHF, los resultados
no son tan favorables como para los hornos anteriores, siendo el potencial económico de
entre un 60% y un 77% del potencial técnico (dependiendo del escenario) para una
eficiencia del ORC del 15% y entre el 65% y el 84% para una eficiencia del ORC del 19%.
El potencial técnico para los RHF varía entre 140 y 290 GWh/a para una eficiencia del ORC
del 15% y entre 180 y 370 GWh/a para el 19%.
Para la industria del cemento, los resultados son muy positivos. El potencial teórico varía
entre 700 y 1020 GWh/a para una eficiencia del ORC del 15% y entre 880 y 1300 GWh/a
si es del 19%. El potencial económico es de más del 97% incluso para el peor de los casos
y la eficiencia más baja.
Appendix xxiii

Para la industria del vidrio, no resulta tan rentable la instalación de esta tecnología. El
potencial técnico está entre 130 y 320 GWh/a para una eficiencia del ORC del 15% y entre
160 y 410 GWh/a para una eficiencia del 19%. El potencial económico resulta cero para
los peores escenarios para ambas eficiencias del ORC y un precio de la electricidad de 5,1
ct/kWh como se ha considerado hasta ahora. Para el mejor de los casos y una eficiencia
del 19% se consigue un potencial económico del 55%. Sin embargo, para precios de la
electricidad de entorno a la media (8,84 ct/kWh), el potencial económico es muy próximo a
técnico (cercano al 100%).
A la luz de estos resultados, se puede concluir que el potencial económico es notable,
si se considera un tiempo de amortización de 10 años, la tasa de interés del 4% y la dispo-
nibilidad del ORC del 95%. Puede llegar a ser de hasta 435 MWe sumando los sectores
considerados (Tabla 6.9).

Comparación con estudios previos


A continuación, se lleva a cabo una comparación de los resultados obtenidos en este
trabajo con otros procedentes de otros autores para comprobar si están en el mismo orden
de magnitud o si varían entre sí.
Potencial teórico
El calor residual disponible a 15ºC estimado en este trabajo está entre 21,08 y 29,19
TWh/a teniendo en cuenta los tres sectores. Esto equivale a 75,89-105,08 PJ/a, lo que se
sitúa por debajo de la estimación más baja (Brueckner) que se observa en la Figura 6.1.
Esta figura muestra las estimaciones por parte de diversos autores para toda la industria
alemana. Por lo tanto, tiene sentido que la estimación de este trabajo para tres de los sec-
tores más importantes esté dentro del orden de magnitud, pero sea inferior a todas las
estimaciones. Esto garantiza, además, que este trabajo provee una estimación del poten-
cial teórico conservadora y que, si los resultados son positivos para la misma, lo serán en
caso de que el potencial sea mayor.
A pesar de que existen muy pocos estudios que estimen un valor del potencial calorífico
residual concretamente para las industrias del acero, del cemento o del vidrio en Alemania,
se han encontrado dos y se utilizarán para comparar los resultados obtenidos en este tra-
bajo con los resultados de estos trabajos anteriores que utilizaron métodos diferentes.
Papapetrou et al. (2018) [2] estiman el potencial de calor residual de la industria side-
rúrgica en Alemania en aproximadamente 45 TWh/a, lo cual es muy superior a la calculada
en este trabajo, que va de 11,48 a 15,58 TWh/a según el escenario.
Miró (2016) [15] da un valor del potencial de calor industrial de la producción de cemento
y vidrio en Alemania basado en la producción de CO 2 de la base de datos E-PRTR. Estas
estimaciones se acercan más al límite inferior de la estimación que se hace en este trabajo
para el caso del cemento (7,8 TWh/a de Miró vs. 7,9 TWh/a de este trabajo), mientras que
la estimación para el vidrio es mucho más baja que el valor estimado en este trabajo (0,59
vs. 1,73 TWh/a). La razón principal de esto es la menor cantidad de plantas de fabricación
de vidrio incluidas en la base de datos del E-PRTR y consideradas por Miró et al. [15] (7
vs. 71 plantas).
24

Potencial técnico
Los resultados del potencial técnico estimado en este trabajo se comparan con el estu-
dio de Campana et al. (2013) [18]. Este estudio considera para la industria siderúrgica so-
lamente EAFs y RHFs. Por lo tanto, para la comparación, solo se considerarán los EAF
junto con los RHF. La potencia ORC acumulada por Campana et al. [18] es de 74,0 MWe
para EAF y 82,2 MWe para RHF. Estos valores son significativamente mayores que los
estimados de 23,0-67,5 MWe para EAF y 16,5-41,9 MWe para RHF de la Tabla 6.7. La
principal divergencia se debe a la diferencia en la producción de acero considerada en este
trabajo y la considerada por Campana et al. (16,7 Mt de acero para EAF y 50,8 Mt para
RHF en lugar de los 13,0 y 37,5 Mt de acero utilizados en este trabajo).
Para el cemento, Campana et al. [18] estimaron una potencia ORC acumulada de 70,3
MWe, mientras que en el presente análisis se estima un potencial mayor, en el rango de
79,6-148,1 MWe. La discrepancia es debida a una diferencia en la potencia del ORC espe-
cífica por unidad de producción de cemento.
Para el vidrio, solo la producción total de la UE-27 de vidrio plano es considerada por
Campana et al. [18], por lo que no es posible una comparación directa para Alemania.
Comparación de la potencia estimada de los ORC con la potencia de los
ORC en funcionamiento
Para la industria siderúrgica en Alemania, solo Elbe-Stahlwerke Feralpi en Riesa cuenta
con un sistema ORC para recuperar el calor del gas de escape del EAF. Para la industria
del cemento, la planta de Lengfurt es la única con un sistema ORC.
Resulta de gran interés comparar la potencia real de estos ORC con la potencia esti-
mada en este trabajo. El ORC de Feralpi produce 2,7 MWe, pero un tercio del vapor alma-
cenado en el tambor de vapor se vende a otro proceso industrial (la planta de neumáticos
de Goodyear Dunlop, que está en las proximidades) a través de la red de calefacción del
distrito [93]. Por lo tanto, solo dos tercios del calor técnicamente disponible se utilizan para
la generación de electricidad. Esto significa que la energía eléctrica que podría producirse
es de 4,05 MWe.
La planta del Heidelberg Cement Group en Lengfurt tiene un ORC de 1,5 MWe y utiliza
el calor residual sólo del enfriador de clínker de parrilla [102].
Tabla 6.10. Potencia ORC estimada y real para las plantas de Feralpi y Lengfurt

Potencia Potencia
Potencia estimada [MW] nominal potencial
[MW] [MW]

Planta
Eficiencia del ORC de Eficiencia del ORC de
15% 19%

WCS AS BCS WCS AS BCS

Feralpi 1.32 2.18 3.05 1.67 2.76 3.86 2.70 4.05

Lengfurt 0.72 0.97 1.23 0.91 1.23 1.56 1.50 -


Appendix xxv

Como se puede observar en la Tabla 6.10, la potencia eléctrica nominal obtenida de los
ORC instalados en el EAF de Feralpi y en el enfriador de clínker de Lengfurt está, para
ambos casos, dentro del rango de valores de potencia posibles para los escenarios del
peor, el promedio y el mejor de los casos. Para el EAF de Feralpi, la potencia eléctrica
potencial que podría obtenerse es ligeramente superior a la del BCS con una eficiencia del
ORC del 19%.

Análisis de sensibilidad
Los resultados anteriores sobre el potencial económico están basados en hipótesis so-
bre el valor de la de tasa de interés, el tiempo de amortización y la eficiencia y disponibilidad
del ORC. Sin embargo, este puede cambiar con la variación de estos parámetros. Para
comprender mejor la influencia de cada parámetro en los resultados, se realiza un análisis
de sensibilidad en el que se modifican de forma independiente la tasa de interés i y el
tiempo de amortización n. Estos dos parámetros son los que influyen en el potencial eco-
nómico. La eficiencia no se incluye en el estudio de sensibilidad dado que su modificación
solo contribuye con un aumento o descenso lineal del potencial técnico. Por simplicidad,
solo se considera el escenario promedio (determinado por la magnitud del calor específico
de los gases de escape del proceso) y el precio de electricidad más bajo.
Las Figura 6.12 y Figura 6.13 muestran cómo varía el potencial económico cuando el
tiempo de amortización en años y la tasa de interés cambian dentro de un rango. En cuanto
al tiempo de amortización, se consideran de 5 a 15 años para el análisis, y la tasa de interés
varía entre el 2% y el 15%. Cuando se varía el tiempo de amortización, la tasa de interés
es i=4%, y cuando se varía esta tasa, el tiempo de amortización es n=10. La eficiencia es,
para ambos casos, del 15%. La eficiencia y la disponibilidad del ORC influyen solo en el
potencial técnico, que no depende del tipo de interés y del tiempo de amortización.
Para el BOF, debido al alto potencial económico (igual al técnico) de la instalación de
sistemas ORC, este no varía con los cambios en los parámetros en estos rangos. En el
caso del EAF, el potencial económico comienza a disminuir cuando el tiempo de amortiza-
ción es inferior a 10 años y se convierte en cero antes de llegar a los 5 años. Por otra parte,
el aumento del tipo de interés hace que el potencial económico descienda hasta cero a
poco menos del 15%. En el caso del RHF, el potencial económico es nulo a menos de 5
años de amortización y a más del 15% de la tasa de interés.
Para el caso del cemento (Figura 6.13), la situación es más favorable. El potencial eco-
nómico baja a alrededor de 60 GWh/a para 5 años de amortización y no llega a cero para
una tasa de interés del 15%. El vidrio es el sector menos favorable, por la menor tasa de
utilización y el menor tamaño de las plantas. Ningún ORC en ninguna planta es económico
por debajo de 10 años de tiempo de amortización o por encima del 4,3 % de la tasa de
interés (Figura 6.13).
26

Economic potential [GWh/a]


Economic potential [GWh/a]

(a) (b)
Economic potential [GWh/a]

(c)
Figura 6.12. Análisis de sensibilidad para: (a) BOF; (b) EAF y (c) RHF
Economic potential [GWh/a]
Economic potential [GWh/a]

(a) (b)
Figura 6.13. Análisis de sensibilidad para: (a) cemento y (b) vidrio
Appendix xxvii

Ahorro de electricidad
El principal beneficio de la instalación de estos equipos ORC en las plantas donde re-
sulta económicamente interesante es la electricidad que se ahorra, ya que se produce con
el exceso de calor que de otra manera sería desechado.
La industria siderúrgica utiliza unos 22 TWh de electricidad al año. Esto supone el 9%
del consumo industrial de Alemania y el 4% del consumo total del país. La electricidad es
necesaria para alimentar los EAF que funden la chatarra para convertirla en acero y para
el funcionamiento de los trenes de laminación [94].
Para la industria del cemento, cada tonelada de cemento requiere aproximadamente
110 kWh de electricidad [103]. Esto, multiplicado por la producción de cemento en Alema-
nia de 23,17 Mt, hacen 2,55 TWh/a. La producción de vidrio requiere aproximadamente
203 kWh/t para el vidrio plano y 372 kWh/t para el vidrio para envases [95]. Teniendo en
cuenta la producción aproximada de vidrio en Alemania por año (1,98 Mt de vidrio plano y
3,93 Mt de vidrio para envases) [92], se obtiene una necesidad de electricidad de aproxi-
madamente 1,86 TWh/a.
La Tabla 6.11 muestra el porcentaje de la electricidad total consumida en la producción
alemana de acero, cemento y vidrio que podría ahorrarse mediante la instalación de los
ORC rentables para el menor precio de la electricidad y en el caso de 10 años de tiempo
de amortización, tasa de interés del 4%, disponibilidad del ORC del 95% y eficiencia del
ORC del 15%. Se calcula un intervalo entre el peor y el mejor escenario.

Tabla 6.11. Ahorro de electricidad con ORC para acero, cemento y vidrio con 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 =
𝟏𝟓 %, i=4%, n=10 años

Sector Electricidad Electricidad con- 𝑬𝒍𝒆𝒄𝒕𝒓𝒊𝒄𝒊𝒅𝒂𝒅 𝒂𝒉𝒐𝒓𝒓𝒂𝒅𝒂


[%]
ahorrada [TWh/a] sumida [TWh/a] 𝑬𝒍𝒆𝒄𝒕𝒓𝒊𝒄𝒊𝒅𝒂𝒅 𝒄𝒐𝒏𝒔𝒖𝒎𝒊𝒅𝒂

Acero 1.30-1.78 22.00 5.91-8.09%

Cemento 0.68-1.02 2.55 26.67-40%

Vidrio 0-0.120 1.86 0-6.45%


28

Emisiones de CO2 evitadas


En el año 2018, el factor de emisión de CO 2 para el mix de energía eléctrica en Alemania
se estimó en una media de 474 g/kWh (Figura 6.14).

Figura 6.14. Desarrollo del factor de emisión de CO2 para el mix de electricidad en Alemania
de 1990 a 2018 (en g/kWh) [96]

Con la implementación de estos sistemas de recuperación de los gases de escape, se


puede evitar una parte de la generación de electricidad porque la planta puede utilizar la
energía generada con el ORC en lugar de tomarla de la red. Por lo tanto, se reduce la
emisión de CO2 asociada a la producción de esta electricidad.
Las emisiones de gases de efecto invernadero (GEI) del sector industrial en Alemania
en 2018, medidas millones de toneladas de CO 2 equivalente, fueron de 65 MtCO2eq (Figura
6.15). La instalación de las unidades de ORC según el potencial económico para los su-
puestos de la Tabla 6.12 supondría unas reducciones de GEI que se pueden leer en la
tabla. Estas reducciones de GEI se dan en forma de emisiones de CO2 evitadas por el
ahorro en la producción de electricidad.

Tabla 6.12. Emisiones de CO2 evitadas con la instalación del potencial económico de los
ORC para el acero, el cemento y el vidrio con 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 años

Sector Emisiones de CO2 evita- 𝑬𝒎𝒊𝒔𝒊𝒐𝒏𝒆𝒔 𝒅𝒆 𝑪𝑶𝟐 𝒆𝒗𝒊𝒕𝒂𝒅𝒂𝒔


[%]
das [103 tCO2/a] 𝑬𝒎𝒊𝒔𝒊𝒐𝒏𝒆𝒔 𝒅𝒆 𝑮𝑬𝑰 𝑨𝒍𝒆𝒎𝒂𝒏𝒊𝒂 𝟐𝟎𝟏𝟖

Acero 616.53-844.57 0.95-1.30%

Cemento 324.64-485.61 0.50-0.75%

Vidrio 0-57.12 0-0.09%


Appendix xxix

Figura 6.15. Tendencias de las emisiones de GEI en Alemania por sectores 1990-2018. Datos
de: UBA 2019, BMU 2019 [97]

Ahorro de costes
La instalación de unidades ORC no solo es buena para los objetivos de sostenibilidad,
sino que también puede significar un importante ahorro económico al reducir el costo de
compra de electricidad.
El ahorro total de costes se estima para el caso en que la eficiencia del ORC es del 15%,
con i=4% y n=10. El precio de la electricidad es de 5,1 ct/kWh (Figura 6.16).
Figura 6.16. Ahorro de costes con la instalación del potencial económico de ORC para
acero, cemento y vidrio con 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=10 años

Sector Ahorro de costes [M€/a]

Acero 66.34-90.87

Cemento 34.93-52.25

Vidrio 0-6.1
30

A.9 Conclusiones
La industria alemana libera más de 200 TWh de calor residual cada año, lo cual es más
que el consumo anual total de energía en Dinamarca [6]. El gran número de fábricas de
acero, cemento y vidrio en Alemania produce una cantidad significativa de calor de proceso
de alta temperatura. Además, estas tres industrias intensivas en energía son las que mues-
tran el estado más desarrollado del mundo en cuanto a recuperación de calor residual con
la tecnología ORC, aunque el potencial es mucho mayor en el caso de Alemania, como se
muestra en este trabajo.
Las estimaciones del potencial teórico y del técnico son, como se ha comprobado, del
mismo orden de magnitud aunque un tanto inferiores a las realizadas por otros autores
previamente. Esto garantiza que las estimaciones realizadas en este trabajo son
conservadoras. Además, las conclusiones sobre el potencial económico de esta tecnología
son, para la mayor parte de los procesos, positivas. Se considera para el análisis del
potencial económico el menor precio de la electricidad para estas industrias (situación más
conservadora).
Para la industria del acero, el potencial económico de la instalación de unidades ORC
en hornos BOF es igual al técnico bajo cualquier hipótesis y escenario, siendo este de en
torno a entre 1100 y 1400 GWh/a para eficiencias del ORC del 15% al 19% y considerando
los tres escenarios (de más pesimista a más optimista) según la magnitud del calor residual
disponible. Los EAF son, también, bastante interesantes de considerar para la instalación
de sistemas ORC de recuperación de calor de sus gases de escape. El potencial técnico
varía entre 200 y 600 GWh/a. Para una eficiencia del 15%, el potencial económico es un
poco mayor al 80% del potencial técnico para el peor escenario, y un 97% para el mejor
escenario. Para el caso con una eficiencia del 19%, el potencial económico se encuentra
ligeramente más cerca del técnico. Para los RHF, los resultados no son tan favorables
como para los hornos anteriores, siendo el potencial económico de entre un 60% y un 77%
del potencial técnico (dependiendo del escenario) para una eficiencia del ORC del 15% y
entre el 65% y el 84% para una eficiencia del ORC del 19%. El potencial técnico para los
RHF varía entre 140 y 370 GWh/a.
Para la industria del cemento, los resultados son muy positivos. El potencial teórico varía
entre 700 y 1300 GWh/a. El potencial económico es de más del 97% incluso para el peor
de los casos y la eficiencia más baja.
Para la industria del vidrio, no resulta tan rentable la instalación de esta tecnología. El
potencial técnico está entre 130 y 410 GWh/a. Sin embargo, el potencial económico resulta
cero para los peores escenarios y un precio de la electricidad de 5,1 ct/kWh. No obstante,
para el más optimista de los escenarios y una eficiencia del 19% se consigue un potencial
económico del 55%. Además, para precios de la electricidad de entorno a la media (8,84
ct/kWh), el potencial económico es muy próximo a técnico (cercano al 100%). Esto no solo
ocurre para el vidrio, sino que también para las demás industrias y sus diferentes procesos.
El potencial económico crece con el tiempo de amortización y disminuye con los tipos
de interés más altos. Los resultados del potencial económico son prometedores para la
recuperación de calor de los procesos estudiados de los sectores de producción de acero
y cemento para tiempos de amortización entre 10 y 5 años y tipos de interés inferiores al
15%. Para el vidrio, sin embargo, el tiempo de amortización deberá ser superior a 10 años
y la tasa de interés más baja que 4,3%.
Appendix xxxi

La instalación de unidades de ORC para la recuperación de calor residual en las plantas


donde es económicamente viable podría suponer un ahorro de hasta 149,22 millones de
euros anuales en total y evitar 1,39 millones de toneladas de emisiones de CO 2 por año, lo
que corresponde al 2,13% de las emisiones de GEI por parte de la industria alemana en
2018.
Como conclusión, los resultados sobre el potencial (técnico y económico) de la instala-
ción de sistemas de recuperación de calor residual con ORC obtenidos en este trabajo son
bastante optimistas. Esto, junto con los incentivos y programas de apoyo de la Unión Eu-
ropea y del Fondo Alemán de Eficiencia Energética (EEF) [7], hacen de la tecnología ORC
una opción muy atractiva para la recuperación de calor residual industrial.

A.10 Líneas futuras


La metodología y los resultados presentados en esta tesis proporcionan algunas pautas
para el trabajo futuro y la investigación posterior en los siguientes temas:
• Aplicación del método desarrollado para la evaluación del potencial de recupe-
ración de calor residual con ORC para otros procesos de otros sectores (cerá-
mica, petróleo y gas, química) de la industria intensiva en energía en Alemania.
• Extensión de este estudio a otros países de interés para la instalación de esta
tecnología.
• Con el fin de mejorar la precisión de los resultados, sería interesante conocer de
forma directa el calor residual específico de cada proceso en cada planta, en
lugar de tener un escenario de peor, promedio y mejor caso. El acceso a esta
información sobre las industrias alemanas contribuiría a una mejor estimación,
pero resulta complicado y requiere una inversión económica y de tiempo impor-
tante.
32

A.11 Planificación temporal y presupuesto

A.12 Planificación temporal


En este apartado se desarrolla la planificación temporal del proyecto. En primer lugar se
elabora la Estructura de Descomposición del Proyecto (EDP). La EDP (Figura 6.17) sirve
para mostrar de forma esquemática las actividades desarrolladas durante el proyecto. Este
gráfico tiene una estructura jerárquica de distribución de actividades, apareciendo en pri-
mer lugar las principales y, después, se van subdividiendo hasta completar el total de ta-
reas desarrolladas.
La primera tarea de este proyecto consiste en una documentación y un estudio sobre el
estado del arte. En primer lugar, sobre el marco económico y medioambiental de Alemania
y el papel de la industria y las industrias que son intensivas en consumo de energía (pro-
cesos de producción, balance de energía…). Luego, sobre los sistemas de recuperación
del calor residual de los procesos industriales y, en concreto, con la tecnología ORC. Por
último, se investiga sobre trabajos anteriores de estimaciones del potencial.
En segundo lugar, tras desarrollar un método, se lleva a cabo la recopilación de datos
necesarios. Esta parte es la más complicada, dada la falta de información de libre acceso
y su dispersión.
Una vez obtenidos todos los datos necesarios, se procede a calcular los potenciales
teórico, técnico y económico aplicando el método desarrollado en este trabajo. A su vez,
se elaboran en MATLAB® las gráficas del potencial económico. A continuación, para el
estudio de los resultados, se realiza un análisis de sensibilidad utilizando también
MATLAB®.
Durante el transcurso del trabajo se procede a la redacción de la memoria del mismo en
inglés para la Technische Universität München (TUM). Por último, se traduce el documento
para la creación de un resumen para el documento Trabajo de Fin de Grado para la Escuela
Técnica Superior de Ingenieros Industriales (ETSII) de Madrid, se redactan el presupuesto
y la planificación temporal y se hace una revisión del documento.
Appendix xxxiii

Figura 6.17. Estructura de Descomposición del Proyecto


34

Para completar la información presentada en la EDP se realiza un diagrama PERT so-


brevista de Gantt. En él aparece un esquema con información sobre la fecha de comienzo
y finalización de las distintas actividades y la relación entre ellas. Este diagrama debe tener
un único comienzo y final y se presenta en la Figura 6.18.

A.13 Presupuesto
En esta sección se pretende estimar el coste que ha supuesto la realización de este
trabajo.
Para evaluar dicho coste, el principal concepto a valorar son las horas de trabajo dedi-
cadas. Esto se debe a que no ha sido necesaria la adquisición de equipos ni de materiales.
Sin embargo, sí ha de tenerse en cuenta la amortización del ordenador personal utilizado,
el software MATLAB® y el paquete de Microsoft Office, que no son de libre distribución.
En primer lugar, se estima el número de horas dedicadas al trabajo. Para ello, se deberá
tener en cuenta el coste horario del alumno y del tutor. Se estima que estos costes son de
15 €/h para el alumno y 30 €/h para el tutor. Se estima que las horas trabajadas por el tutor
son de 12 horas al mes. Esto, durante seis meses, supone unas 72 horas. Para la estima-
ción de las horas trabajadas por el alumno se parte del número de horas dedicadas a cada
parte según la planificación temporal y considerando el número de horas diarias reflejado
en la Tabla 6.13.
Tabla 6.13. Horas de trabajo del alumno

Concepto Días Horas/día Horas total

Documentación y estudio del estado del arte 51 3 153

Obtención de datos 30 3 90

Desarrollo del método 20 4 80

Obtención de resultados y análisis 19 4 76

Redacción de documentos 115 4 460

Total 859

Se obtiene que el número total de horas trabajadas por el alumno son 859 horas. Te-
niendo el cuenta que este Trabajo de Fin de Máster con ampliación tiene asignados 30
créditos ECTS y que cada crédito se estima en de 25 a 30 horas de trabajo del alumno (lo
cual multiplicado por 30 son entre 750 y 900 horas), el número de horas dedicadas a esta
tesis entra dentro del rango previsible. La Tabla 6.14 muestra el coste total del alumno y
del profesor y la suma de ambos o coste del personal.
Appendix xxxv

Tabla 6.14. Coste del personal

Concepto Coste horario (€) Horas Coste total (€)

Alumno 15 859 12885

Tutor 30 72 2160

Total 15045

En segundo lugar, se calcula el coste de amortización del ordenador utilizado para la


realización del trabajo, cuyo precio es de en torno a 800 €. Estimando una vida útil de 5
años, una utilización diaria media de 4 horas y un valor residual nulo, y teniendo en cuenta
que se utiliza el 95% del tiempo empleado en la realización de este TFM, la amortización
es, aproximadamente:
0,95 ∙ 859
𝐶𝑎𝑚𝑜𝑟𝑡𝑖𝑧𝑎𝑐𝑖ó𝑛 = ∙ 800 = 89,43 €
4 ∙ 365 ∙ 5

Por último, se tiene en cuenta el coste de las licencias de MATLAB® y del paquete de
Microsoft Office. La licencia de MATLAB® tiene un coste de en torno a 100 € y un periodo
de validez de 1 año. La licencia de Microsoft Office tiene un coste de 100 € y un periodo de
validez de 1 año también. Teniendo en cuenta que se ha hecho uso de ambas durante 6
meses, el coste es de 50 € por cada licencia, es decir, 100 € en total.
Sumando todos los costes se obtiene un coste bruto de 15234 € y aplicando un 21% de
IVA se obtiene un coste neto de 18433,14 €.
Tabla 6.15. Coste total del proyecto

Concepto Horas de tra- Amortización Licencia Licencia Office


bajo PC MATLAB®

Subtotal (€) 15045 89 50 50

Total bruto (€) 15234

Total neto (€) 18433,14


36

Figura 6.18. Diagrama PERT sobre vista de Gantt


Appendix xxxvii

A.14 Valoración del impacto del trabajo


En esta sección se pretende analizar de forma breve el impacto social, económico y
medioambiental que supone la elaboración de este trabajo.
Las estimaciones llevadas a cabo son una primera aproximación al potencial de la tec-
nología ORC para la recuperación de calor residual industrial. Los resultados aquí presen-
tados proveen a la industria, a los inversores y a los gobiernos con información sobre la
viabilidad económica y la rentabilidad de estos sistemas. Esto supone una base científica
desde la cual considerar la inversión en estos proyectos con mayor seguridad económica.
Además, la producción de electricidad con las unidades ORC en las plantas donde re-
sultaría, en principio, rentable supondría un ahorro económico para las industrias que re-
percutiría en toda la sociedad. Estas, producirían electricidad a partir del calor en los gases
de escape de determinados procesos, reduciendo el consumo de electricidad de la red.
Por otra parte, de la mano del ahorro económico va la reducción de las emisiones de
CO2, al dejar de producirse la energía eléctrica a partir de fuentes no renovables y, en su
lugar, generarla a partir del calor de los gases de escape que de otra forma sería perdido.
Esta contribución a la eficiencia energética reporta beneficios a la industria y a toda la so-
ciedad, contribuyendo a la consecución de los objetivos de eficiencia recogidos en la Di-
rectiva de Eficiencia Energética de la Comisión Europea para 2020 y en el Fondo de Efi-
ciencia Energética (FEE), que forma parte de la llamada "Energiewende" (transición ener-
gética) que tiene como objetivo reducir el consumo de energía primaria en un 20% para el
año 2020 y en un 50% para el 2050 (en comparación con el año 2008).
38

B WHR projects with ORC from ORMAT® and TURBODEN®


Table 6.1. ORMAT industrial WHR projects globally [102]

ORMAT® INDUSTRIAL HEAT RECOVERY

ORC
Location Customer Status Notes
size

ACP (Waterval Smelter).


In operation First ORC on Metallurgical Con-
5 MW South Africa Eternity Power for Anglo
since 2015 verter in the World
Platinum

Tadipatri,
In operation Waste heat is recovered from
4 MW Andhra Ultratech Cement
since 2007 the clinker cooler vent air
Pradesh, India

1.2 Lengfurt, Heidelberg Cement In operation Waste heat is recovered from


MW Germany Group since 1999 the clinker cooler vent air

Electric power production from


Manfredonia, In operation
3 MW Vetrerie Sangalli waste heat in float glass produc-
Italy since 2008
tion process
Appendix xxxix

Table 6.2. Turboden industrial WHR for cement industry projects globally [104]

TURBODEN® INDUSTRIAL HEAT RECOVERY

Heat recovery from cement production process

ORC
Location Customer Status Notes
size

Electric power production


2 Ait Baha, In operation from waste heat in cement
Ciments Du Maroc
MWe Morocco since 2010 production process + CSP
(Hybrid plant)

Electric power production


4 In operation
Aleşd, Romania Holcim SA from waste heat in cement
MWe since 2012
production process

Electric power production


5 Rohožník, In operation
CRH from waste heat in cement
MWe Slovakia since 2014
production process

Fieni,
Electric power production
3.8 Dâmboviţa S.C. Carpatcement Holding S.A. In operation
from waste heat in cement
MWe County, (Heidelberg Group) since 2015
production process
Romania

Möriken- Electric power production


2.3 Jura-Cement-Fabriken AG (CRH In operation
Wildegg, from waste heat in cement
MWe Group) since 2016
Switzerland production process

Electric power production


2 Industria Cementi Giovanni Under
Piacenza, Italy from waste heat in cement
MWe Rossi construction
production process

Turkey CTP Team S.r.l. (Italy)


Electric power production
7 and CTN Makina Mühendislik In operation
Narli, Turkey from waste heat in cement
MWe Insaat Celik Konstrüksiyon since 2019
production process
Otomasyon (Turkey)

Electric power production


1.3 Eclépens, Cadcime SA / Holcim Suisse Under
from waste heat in cement
MWe Switzerland Eclépens – LafargeHolcim group construction
production process

Yumurtalik, Electric power production


7.3 CTP Team Srl / Sönmez Çi- Under
Erzin / Hatay, from waste heat in cement
MWe mento construction
Turkey production process
40

Table 6.3. Turboden industrial WHR for steel industry projects globally [104]

TURBODEN® INDUSTRIAL HEAT RECOVERY

Heat recovery from steel production process

ORC
Location Customer Status Notes
size

ESF Elbe-
2.7 In operation Electric power production from
Riesa, Germany Stahlwerke Feralpi
MWe since 2013 waste heat in steel industry (EAF)
GmbH

Electric power production from


0.7 NatSteel Holdings In operation
Singapore waste heat in steel industry (Rolling
MWe Pte Ltd since 2013
mill)

2.2 In operation Electric power production from


Brescia, Italy ORI Martin S.p.A.
MWe since 2016 waste heat in steel industry (EAF)

Mitsubishi Heavy In-


2.5 Under Electric power production from
Nagoya, Japan dustries / Aichi
MWe construction waste heat in steel industry (EAF)
Steel

10 In operation Electric power production from


Cremona, Italy Arvedi S.p.A.
MWe since 2018 waste heat in steel industry (EAF)

Electric power production from


1.2 Gwangyang City, POSCO ICT Under
waste heat in steel industry (Sub-
MWe South Korea Company construction
merged Arc Furnace)
Appendix xli

Table 6.4. Turboden industrial WHR for glass industry projects globally [104]

TURBODEN® INDUSTRIAL HEAT RECOVERY

Heat recovery from glass production process

ORC
Location Customer Status Notes
size

Electric power production from


1.3 In operation
Cuneo, Italy Gea Bischoff/AGC waste heat in float glass pro-
MWe since 2012
duction process

Electric power production from


0.5 Villotta di Chions In operation
BDF Industries waste heat in container glass
MWe (PN), Italy since 2015
production process

Electric power production from


6.2 Beyköy - Düzce, Çalbıyık Grup / Düzce In operation
waste heat in float glass pro-
MWe Turkey Cam since 2018
duction process

GEA Bischoff GmbH / Electric power production from


1.2 Sriperumbudur, Under
Saint-Gobain India waste heat in float glass pro-
MWe India construction
Pvt.Ltd. - Chennai duction process

Electric power and compressed


1.2 GEA Process Engineer- Under
Pisa, Italy air production from waste heat
MWe ing S.p.A. construction
in float glass production process
42

Table 6.5. Turboden industrial WHR for other industries projects globally [104]

TURBODEN® INDUSTRIAL HEAT RECOVERY

Heat recovery from other production processes

ORC
Location Customer Status Notes
size

0.8 Radenthein, Veitsch-Radex In operation Electric power production from waste


MWe Austria GmbH & Co since 2009 heat in refractory production process

1.7 Under Electric power production from waste


Germany Undisclosed
MWe construction heat in aluminum production process

0.7 Torbole Casaglia Fonderia di In operation Electric power production from waste
MWe (BS), Italy Torbole since 2016 heat from cast iron cupola furnace

Electric power production from waste


1.9 Negeri Sembilan, Under
Invest Energy heat from chemical vapor infiltration
MWe Malaysia construction
furnace

2.2 Under Electric power production from waste


Vercelli, Italy Sacal S.p.A.
MWe construction heat in aluminum production process
Appendix xliii

C Main steel production plants in Germany


The steel production plants used for the potential estimation in this work are presented
in Table 6.6, Table 6.7 and Table 6.8 [89].
Table 6.6. Main German steel plants with BOF

Company Location

ArcelorMittal Bremen GmbH Bremen

ArcelorMittal Eisenhüttenstadt GmbH Eisenhüttenstadt

Hüttenwerke Krupp Mannesmann (HKM) GmbH Duisburg

Salzgitter AG Salzgitter

ThyssenKrupp Steel Europe AG Duisburg

ThyssenKrupp Steel Europe AG Duisburg

Saarstahl AG Völklingen

AG der Dillinger Hüttenwerk Dillingen

ArcelorMittal Duisburg GmbH Duisburg


44

Table 6.7. Main German steel plants with EAF

Company Location

ArcelorMittal Hamburg GmbH Hamburg

B.E.S. Brandenburger Elektrostahlwerke GmbH Brandenburg

B.E.S. Brandenburger Elektrostahlwerke GmbH Brandenburg

Badische Stahlwerke GmbH Kehl

Badische Stahlwerke GmbH Kehl

Benteler Steel/Tube GmbH Lingen

BGH Edelstahl Freital GmbH Freital

BGH Edelstahl Freital GmbH Siegen

Buderus Edelstahl GmbH Wetzlar

Deutsche Edelstahlwerke Siegen

Deutsche Edelstahlwerke Witten

Dörrenberg Edelstahl Rönderoth

Eisenwerk Geweke Hagen

ESF Elbe-Stahlwerke Feralpi GmbH Riesa

GMH Georgsmarienhütte GmbH Georgsmarinenhütte

GMH Schmiedewerke / Elektrostahlwerke Gröditz GmbH Gröditz

Friedr. Wilhelms-Hütte Mülheim/ruhr

Gontermann Peipers Siegen

H.E.S. Hennigsdorfer Elektrostahlwerke GmbH Henningsdorf

H.E.S. Hennigsdorfer Elektrostahlwerke GmbH Henningsdorf

Lech-Stahlwerke GmbH Meitingen

Lech-Stahlwerke GmbH Meitingen

Peiner Träger GmbH (Salzgitter) Peine

Peiner Träger GmbH (Salzgitter) Peine


Appendix xlv

Pleissner Guss GmbH Herzberg am Harz

Saarstahl AG / Saarschmiede GmbH Völklingen

Sande Stahlguss Sande (Oldenburg)

Stahlguss Saar GmbH St.ingbert

Stahlwerk Bous GmbH (GMH) Bous

Stahlwerk Thüringen GmbH Unterwellenborn

ThyssenKrupp Nirosta GmbH Bochum

ThyssenKrupp Nirosta GmbH Krefeld

Trierer Stahlwerk Trier

Vdm Metals GmbH Unna


46

Table 6.8. Main German steel plants with RHF

Company Location

ArcelorMittal Bremen GmbH Bremen

ArcelorMittal Eisenhüttenstadt GmbH Eisenhüttenstadt

ArcelorMittal Hamburg GmbH Hamburg

ArcelorMittal Ruhrort GmbH Duisburg

ArcelorMittal Ruhrort GmbH Duisburg

B.E.S. Brandenburger Elektrostahlwerke GmbH Brandenburg

Badische Stahlwerke GmbH Kehl

Badische Stahlwerke GmbH Kehl

BGH Edelstahl Freital GmbH Freital

BGH Edelstahl Freital GmbH Freital

Buderus Edelstahl GmbH Wetzlar

Buderus Edelstahl GmbH Wetzlar

Buderus Edelstahl GmbH Wetzlar

Deutsche Edelstahlwerke Siegen

Deutsche Edelstahlwerke Witten

ESF Elbe-Stahlwerke Feralpi GmbH Riesa

GMH Georgsmarienhütte GmbH Georgsmarinenhütte

H.E.S. Hennigsdorfer Elektrostahlwerke GmbH Henningsdorf

Lech-Stahlwerke GmbH Meitingen

Lech-Stahlwerke GmbH Meitingen

Peiner Träger GmbH (Salzgitter) Peine

Peiner Träger GmbH (Salzgitter) Peine

Saarstahl AG / Saarschmiede GmbH Völklingen

Saarstahl AG / Saarschmiede GmbH Völklingen


Appendix xlvii

Salzgitter AG Salzgitter

Stahlwerk Thüringen GmbH Unterwellenborn

Thyssenkripp Hohenlimburg Hagen

ThyssenKrupp Steel Europe AG Bochum

ThyssenKrupp Steel Europe AG Duisburg

ThyssenKrupp Steel Europe AG Duisburg

ThyssenKrupp Steel Europe AG Duisburg

ThyssenKrupp Steel Europe AG Duisburg

Trierer Stahlwerk Trier


48

D Main cement production plants in Germany


The cement production plants used for the potential estimation in this work are presented
in Table 6.9 [89].
Table 6.9. Main German cement production plants

Company Location

CEMEX OstZement GmbH Rüdersdorf

CEMEX West-Zement GmbH Beckum /KollenbacH

Dyckerhoff AG Amöneburg

Dyckerhoff AG Deuna

Dyckerhoff AG Geseke

Dyckerhoff AG Göllheim

Dyckerhoff AG Lengerich

HeidelbergCement AG Burglengenfeld

HeidelbergCement AG Ennigerloh-Nord

HeidelbergCement AG Geseke

HeidelbergCement AG Hannover

HeidelbergCement AG Leimen

HeidelbergCement AG Lengfurt

HeidelbergCement AG Paderborn

HeidelbergCement AG Schelklingen

Holcim (Deutschland) AG Höver

Holcim (Deutschland) AG Lägerdorf

Holcim (Süddeutschland) AG Dotternhausen

Lafarge Zement Karsdorf

Lafarge Zement Wössingen Wössingen

Märker Zement GmbH Harburg


Appendix xlix

Phoenix Zementwerke Beckum

Portlandzementwerk "Wotan" H. Schneider KG Üxheim-Ahütte

Portland-Zementwerk Wittekind Erwitte

Portland-Zementwerke Gebr. Seibel GmbH & Co. KG Erwitte

Schwenk Zement KG Allmendingen

Schwenk Zement KG Bernburg

Schwenk Zement KG Karlstadt

Schwenk Zement KG Mergelstetten

Solnhofer Portland-Zementwerke GmbH & Co. KG Solnhofen

Spenner Zement GmbH & Co KG Erwitte

Südbayer. Portland-Zementwerk Gebr. Wiesböck & Co. GmbH Rohrdorf

Zement- und Kalkwerke Otterbein GmbH & Co. KG Großenlüder/Müs

Zementwerk Seibel und Söhne oHg Erwitte


50

E Main glass production plants in Germany


The cement production plants used for the potential estimation in this work are presented
in Table 6.10 [89].
Table 6.10. Main glass production plants in Germany

Glass type Company Location

Container glass Ardagh Glass Group Bad Münder

Ardagh Glass Group Drebkau

Ardagh Glass Group Germersheim

Ardagh Glass Group Germersheim

Ardagh Glass Group Lünen

Ardagh Glass Group Lünen

Ardagh Glass Group Neuenhagen

Ardagh Glass Group Nienburg

Ardagh Glass Group Nienburg

Ardagh Glass Group Nienburg

Ardagh Glass Group Nienburg

Ardagh Glass Group Obernkirchen

Ardagh Glass Group Wahlstedt

August Pohli GmbH & Co. KG Lauscha

August Pohli GmbH & Co. KG Lauscha

August Pohli GmbH & Co. KG Lauscha

Gerresheimer AG Steele-Horst

Gerresheimer AG Steele-Horst

Gerresheimer AG Lohr

Gerresheimer AG Lohr

Gerresheimer AG Tettau
Appendix li

Gerresheimer AG Tettau

Heinz-Glas GmbH & Co KG Piesau

Heinz-Glas GmbH & Co KG Piesau

Heinz-Glas GmbH & Co KG Tettau

Heinz-Glas GmbH & Co KG Tettau

Hindustan National Glass & Industries GmbH Gardelegen

Noelle + von Campe Glashütte GmbH Werk I/Boffzen

Noelle + von Campe Glashütte GmbH Werk II/Boffzen

O-I glasspack GmbH & Co. KG Achern

O-I glasspack GmbH & Co. KG Bernsdorf

O-I glasspack GmbH & Co. KG Holzminden

O-I glasspack GmbH & Co. KG Rinteln

O-I glasspack GmbH & Co. KG Rinteln

Saint-Gobain Oberland AG Bad Wurzach

Saint-Gobain Oberland AG Bad Wurzach

Saint-Gobain Oberland AG Bad Wurzach

Saint-Gobain Oberland AG Essen/Kanarp

Saint-Gobain Oberland AG Essen/Kanarp

Saint-Gobain Oberland AG Essen/Kanarp

Saint-Gobain Oberland AG Kipfenberg

Saint-Gobain Oberland AG Neuburg

Saint-Gobain Oberland AG Neuburg

Saint-Gobain Oberland AG Neuburg

Saint-Gobain Oberland AG Wirges

Saint-Gobain Oberland AG Wirges

Weck GmbH & Co. KG Bonn-Duisdorf


52

Weck GmbH & Co. KG Bonn-Duisdorf

Wiegand Bayerische Flaschen-Glashüttenwerke GmbH Steinbach a. W.

Wiegand Bayerische Flaschen-Glashüttenwerke GmbH Steinbach a. W.

Wiegand Bayerische Flaschen-Glashüttenwerke GmbH Steinbach a. W.

Wiegand GmbH Großbreitenbach

Wiegand GmbH Großbreitenbach

Wiegand GmbH Schleusingen

Wiegand GmbH Schleusingen

Wiegand GmbH Steinbach a. W.

Flat glass Euroglas AG Haldensleben

Euroglas AG Osterweddingen

f | glass GmbH Osterweddingen

Glasfabrik Lamberts GmbH & Co. KG Wunsiedel

GMB Glasmanufaktur Brandenburg GmbH Tschernitz

Guardian Flachglas GmbH Thalheim

Pilkington Deutschland AG Gladbeck/Rentfort

Pilkington Deutschland AG Gladbeck/Rentfort

Pilkington Deutschland AG Schmelz

Pilkington Deutschland AG Weiherhammer

Pilkington Deutschland AG Weiherhammer

Saint-Gobain AG Herzogenrath

Saint-Gobain AG Köln-Porz

Saint-Gobain AG Stolberg

Saint-Gobain AG Torgau
Appendix liii

F Economic potential graphs varying parameters

(a) (b)

(c)
Figure 6.1. Economic potential for steel production with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟗 %, i=4%, n=10 years:
(a) BOF; (b) EAF and (c) RHF
54

(a) (b)

Figure 6.2. Economic potential for production with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟗 %, i=4%, n=10 years: (a)
cement; (b) glass.
Appendix lv

(a) (b)

(c)
Figure 6.3. Economic potential for steel production with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=5 years:
(a) BOF; (b) EAF and (c) RHF
56

(a) (b)

Figure 6.4. Economic potential for production with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=4%, n=5 years: (a)
cement; (b) glass.
Appendix lvii

(a) (b)

(c)
Figure 6.5. Economic potential for steel production with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=6%, n=10 years:
(a) BOF; (b) EAF and (c) RHF
58

(a) (b)

Figure 6.6. Economic potential for production with 𝜼𝒆𝒍,𝒏𝒆𝒕,𝑶𝑹𝑪 = 𝟏𝟓 %, i=6%, n=10 years: (a)
cement; (b) glass.

You might also like