Download as pdf or txt
Download as pdf or txt
You are on page 1of 35

Newcastle University

EPSRC: Thermal Management of Industrial Processes

National sources of low grade heat available from the process industry
Progress Report
(Feb. 2011)

List of Abbreviations
AEA: Energy and Environment
BERR: Department for Business Enterprise and Regulatory Reform (previously Department
for Trade and Industry)
BF: Blast Furnace
BFG: Blast Furnace Gas
BOF: Basic Oxygen Furnace
EA PAS: Environment Agencies Public registers
EAF: Electric Arc Furnace
NAP: National Allocation Plan
NEPIC: North East of England Process Industry Cluster
UNEP: United Nations Environment Programme
Unit
PJ: Peta Joule (=1015J)
TWh: Tera Watt-hour (=1012Wh)
Mt/yr: Million tons per year
List of figures
Figure 1: Schematic of a heat exchanger................................................................................ 7
Figure 2: Heat Transfer Efficiency versus Source Temperature .............................................. 8
Figure 3: Map of industrial heat [3] ....................................................................................... 9
Figure 4: Industrial heat load by industrial sector [3] ........................................................... 10
Figure 5: Schematic representation of a steel production plant ............................................. 14
List of tables
Table 1: Waste heat sources in major industrial processes (cf. Table 11.7 of [7]) .................... 6
Table 2: Steel capacity ......................................................................................................... 14
Table 3: Specific energy consumption and energy consumption splits within Steel industry
processes ............................................................................................................................. 14
Table 4: gas composition in Steel processes ......................................................................... 15
Table 5: Characterization and classification of potentially recoverable low grade heat gas
streams in the steel industry................................................................................................. 25
Table 6: Characterization and classification of potentially recoverable low grade heat cooling
water streams in the steel industry ....................................................................................... 25
Table 7: Gas waste heat sources and potential for recovery .................................................. 26
Table 8: Cooling water waste heat sources and potential for recovery .................................. 27

Contents
Research context ................................................................................................................... 4
Introduction .......................................................................................................................... 5
1. Developing an Industrial heat load map .......................................................................... 8
2. Potential test case studies ............................................................................................. 10
3. Waste heat survey guidelines ........................................................................................ 11
4. Case Study: Steel production process ........................................................................... 12
4.1. Coke production process........................................................................................... 16
4.2. Sinter process ........................................................................................................... 17
4.3. Blast Furnace (BF) process ....................................................................................... 18
4.4. Basic Oxygen Steelmaking (BOS) process................................................................ 19
4.5. Continuous casting process ....................................................................................... 20
4.6. Hot mill process ....................................................................................................... 21
4.7. Cold mill process ...................................................................................................... 22
4.8. Annealing process .................................................................................................... 23
4.9. Power plant .............................................................................................................. 24
4.10.
Low grade heat classification-Summary for Steel production process case study ... 26
4.10.1.
Gas .................................................................................................................... 26
4.10.2.
Cooling water .................................................................................................... 26
4.11.
Potential uses ........................................................................................................ 26
4.11.1. Gas ....................................................................................................................... 26
4.11.2. Cooling water ....................................................................................................... 27
4.12.
Concluding remarks .............................................................................................. 27

Research context
This report presents the progress achieved in the first six months of work of the EPSRC:
Thermal Management of Industrial Processes project.
The first stage of Newcastle Universitys part of this project is to identify the sources of low
grade heat available from the process industry across the UK. Once identified, as many of
these sources as possible will be quantified and characterized.
The objectives of this report are to:
- Identify and characterize opportunities for low grade heat recovery in the UK process
industry.
- Present the first test case study which is that of a steelworks
- Classify waste heat sources in the Steel Industry
The main available sources of information which have been used so far are:
1. UNEP
2. AEA
3. EA PAS database (http://www2.environment-agency.gov.uk/epr/)
4. NAP allocation database
5. NEPIC
6. Trade organisations
7. Steel Industry partner

Introduction
Waste heat refers to the heat absorbed by the environment. According to the Energy
Management Handbook [1],Waste heat is that energy which is rejected from a process at a
temperature high enough above the ambient temperature to permit the economic recovery of
some fraction of that energy for useful purposes. Heat recovery is a generic term used for a
large range of procedures involved with reusing heat otherwise wasted in the environment.
The importance of low-grade heat recovery projects has inevitably increased over the last
couple of years with the current concern for environmental issues and the associated political
policy requiring carbon dioxide emission reduction, as well as general concerns about fuel
security. The Climate Change Act 2008 [2] sets reduction targets, based on 1990 levels, of a
reduction of 80% by 2050 and an interim target of at least 34% by 2020. Given that the
industrial sector represents 40% of the overall CO2 emissions in the UK [3], pressure has
been put on it. For instance, the Climate Change Levy (CCL) [4], a levy on energy use was
applied to industry with a dispensation of 80% available to certain energy-intensive industries
in the form of Climate Change Agreements (CCAs) in return for undertaking energy saving
measures towards predefined goals. The Government have announced that this will be
reduced to 65% by April 2011 [5] thus raising energy costs further. Therefore, the new
priorities in industrys agenda have become to invest more and more in sustainable
technology.
Although a lot has already been done in the past to use energy more efficiently, the industrial
potential for waste heat recovery still represents a thermal energy market potential of some
144 PJ currently lost from industrial processes [6].
This project is particularly interested in low grade heat. The widely accepted definition of this
is, typically, ~250C or less [7].
However, as shown in Table 1, this can vary over the process industries. For example the
exhaust gas temperature of reheat furnaces can reach up to 600C in the Steel industry, and in
the Chemical and Oil industries processes can reject warm gas into the environment with
temperatures up to 340C.
Therefore, the temperature range to consider for the identification of low grade heat sources
will depend on the process industry considered for the analysis.
The main problem in most potential heat recovery applications is how to make effective use
of any recovered heat. There are usually technical solutions to the actual process and the
decision is not normally can it be done? but is it worth it? A key factor in this decision is
the quality of the available heat. The temperature of the heat source is the overriding limiting
factor since, clearly, it must be higher than the required sink temperature. However, the
concept of exergy, which is more normally concerned with the efficiency of heat engines, is
also useful in the comparison of different heat sources. Exergy is the maximum quantity of
technical work one can get from a given low grade heat source.

Table 1: Waste heat sources in major industrial processes (cf. Table 11.7 of [7])

Industry

Plant source

Heat content
(GJ/annum)

Temperature
(C)

Nature

Steel

Coke oven stack gas

3.88E+06

190

Gas

Steel

Sinter from sinter plant

5.52E+06

250

Radiant heat

Steel

Blast furnace stoves

5.50E+06

250

Pressure Energy Gas

Steel

3.75E+06

300

Pressure Energy Gas

1.49E+07

200-600/300-400

gas

Steel

Blast furnace stoves


Finishing soaking pit
reheat furnaces
Cooling water from
reheating furnaces

1.72E+07

20-40

Water

Glass

Container glass melting

2.02E+06

160-200

Gas

Glass

Container glass melting

2.02E+06

140-160

Gas

Glass

Flat glass melting

1.27E+06

160-200

Gas

Glass

Fibre glass melting

140-160

Gas

Glass

Domestic glass melting

Glass

Other glass melting

Steel

1.80E+06

Gas
Gas

Oil

Processing furnaces exhaust

6.56E+07

340

Gas

Oil

Boiler exhaust

1.94E+07

230

Gas

Oil

Condensate

4.80E+06

82

Water

Oil

Process water

2.92E+07

50

Water

Oil

Condenser cooling water

7.30E+06

45

Water

Chemical

Processing furnace exhaust

2.10E+07

340

Gas

Chemical

Boiler exhaust

2.30E+07

230

Gas

Chemical

Condensate

4.00E+06

82

Water

Chemical

Process water

1.00E+07

50

Water

Chemical

Condenser cooling water

2.10E+07

45

Water

Electricity

Flue gases

1.80E+08

130

Gas

Electricity

Cooling water

1.00E+09

25

Water

For any given state defined by the temperature T and the entropy S, with respect to the
ambient standard state , , the specific exergy is defined as, for a continuous flow [8]:
[1],
where h is the enthalpy.
For a fluid stream, the exergy may be written as [8]:
[2],
where
is the fluid specific heat capacity in J/(kg.K) and
is the fluid stream mass flow
rate in kg/s.
The efficiency of a system producing work from a supply of heat is normally considered in
terms of the first law of thermodynamics which considers that the energy within a process is
conserved. The efficiency is often defined according to the first law in terms of the net
6

work output and the energy input. However, this analysis provides no indication of how the
efficiency compares to the maximum efficiency possible, which is not 100%. This is due to
the second law of thermodynamics (which, as was stated by Lord Kelvin, says that it is
impossible to convert heat completely into work). This current study takes this into account
by using the concept of exergy. More precisely, according to Equation (1), exergy accounts
for the irreversibility of the process due to the increase in entropy. Consequences of the
second law of thermodynamics create therefore fundamental constraints on the efficiency of a
heat engine related to the operating temperatures.
In the context of low grade heat recovery, exergy refers to the maximum amount of work
which can be delivered from a system operating between high source temperature and
ambient temperature. It is clearly a measure of the quality of the heat source and, hence, its
usefulness in the consideration of heat recovery applications. Heat recovery necessarily
involves heat transfer and this results in a loss of exergy, i.e. the exergy of the recovered
stream is less than that of the source stream. In other words, the efficiency of the transfer is
temperature dependent and according to Equation (2), exergy is always destroyed when the
process involves a temperature gradient. The exergy efficiency needs to be considered instead
of the energy efficiency. The importance of the exergy efficiency was clearly underlined by
Winter [9] for the design of future industrial processes constrained by energy savings and
CO2 footprint reductions.
In order to define the exergy efficiency of the heat transfer between the source and the
environment, the heat transfer is approximated as a counter flow heat exchanger (cf. Figure
1).
From this approximation, the exergy efficiency, between the heat source and the
environment can be defined as follows:
= Exc / Exh
[3],
In effect, the efficiency given by Equation (3) is the actual increase in the exergy of the cold
stream Exc compared to the maximum possible exergy available for transfer given by the
difference Exh.

Figure 1: Schematic of a heat exchanger

As an example of the effect of temperature on the exergy efficiency, consider a counter-flow


heat exchanger with a 10C approach temperature, the approach temperature being the
minimum temperature difference between the hot fluid and the cold fluid. Th represents the
temperature of the source stream with
while Tc represents the temperature of the
7

Exergy Efficiency (%)

environment sink with


.
Hot and cold streams are supposed to have same heat capacity. Here, T0 = 5C and

= 20C.

100
95
90
85
80
75
70
65
60
35

50

70
90
110 130 150 170
Hot Fluid source temperature (C)

190

Figure 2: Heat Transfer Efficiency versus Source Temperature

Figure 2 shows the efficiency variation with the hot fluid source temperature,
. The heat
transfer efficiency from heat to power decreases as the temperature goes down.
Thus, a high source temperature provides greater choice of applications but also a more
efficient transfer process.
The temperature of the source is of primary importance but the actual energy conversion
depends on many other factors. In fact, the usefulness of the source will also depend on the
quantity, the reliability of supply, the form (gas or liquid, corrosive/non-corrosive) and the
ease of access. Ultimately, a source is not useful unless potential users are found. Users must
be located within a certain distance of the plant, this distance depends on the distance the heat
can be economically transported.
Heat is usually transported via water or steam. According to the report by Terra Infirma [10] ,
steam with temperature in the range of 120-250C can be transported over ~3 to 5 km while
water with temperature in the range of 90-175C can be transported over 30 km. Other
sources cited in that same report mentioned that 9 miles (~15km) is the economic limit for
low-grade heat.
In fact, how far heat can be transported depends on several factors. If heat is assumed to be
transported via a pipe, the heat loss factor which is the ratio between heat loss and the
quantity of heat supplied by the source, depends on the efficiency of pipe insulation but also
on the average size of the pipe and the temperature of the fluid circulating in the pipe relative
to the annual average of the outdoor temperature. The profitability of the heat recovery
project also depends on the cost invested in heat transportation, the total cost being the sum
of the cost for pipeline installation, for heat losses and for pumping power [11].
Hence, heat transport is case specific and further research will include the definition of a
methodology for determining the distance threshold above which no potential users can be
found for economic heat recovery solution.

1. Developing an Industrial heat load map


In July 2006, the market potential for surplus heat from industrial processes in the UK was
8

estimated at 144PJ (40TWh) by the Carbon Trust [6] and more recently at, 65 PJ (18 TWh)
by the Governments Office of Climate Change [12] and 36-71 PJ (10-20 TWh) in a report by
McKenna [3].
These figures reflect a great potential which remains unexploited until now.
It should be noted that obtaining exact national data on waste heat is difficult, so most of
those investigations extrapolated from industrial CO2 emissions, probably due to the relative
ease of obtaining emissions data. This goes some way towards explaining the large variations
between the figures given above.

Figure 3: Map of industrial heat [3]

McKenna [3] developed a procedure to determine the quantity of thermal energy released into
the environment, based on CO2 emission and energy consumed by industries. Emission
factors which give the average emission rate of Carbon released per unit of energy produced
were taken from [13]and energy consumption split by industrial sector was taken from [14].
The emission factors and energy consumption split by industrial sector were then used to
calculate the fuel consumption from combustion. The latter was then used to determine the
heat load for this fuel consumption and the conversion efficiency from fuel to heat estimated
from Carnot cycle efficiency. The actual heat load was weighed against heat recovery factor
for each sector. The recoverable part of the heat load was assumed to be half of the heat
exhaust fraction. In this analysis, temporal variation in heat load was neglected. The head
loads were assumed to be constant over the year.
The results of this investigation are presented in Figure 3. The distribution of industrial heat
loads is represented by the empty circles while the potential for heat recovery is represented
by the solid circles. It is clear from this map that the high temperature recovery potential
concentrates into 3 main centres corresponding to iron and steel plants. This map [3] has
9

already been overtaken by an economic downturn in the industry, resulting in the mothballing
of part of one of the sites. As part of this project, information presentation will be
investigated so large singular sites do not obscure smaller more diffuse, although still
valuable, recovery opportunities.
The energy use for heat is plotted for different types of industry in Figure 4. Low recovery
represents the heat recovery potential of source temperature in the range 100-500C and high
recovery the heat recovery potential of source temperature higher than 500C. This
investigation does not include the potential for temperature lower than 100C. While the heat
recovery potential was underestimated, Figure 4 can be used to determine the sectors with
highest heat loads.

Figure 4: Industrial heat load by industrial sector [3]

The largest heat user is the Iron and Steel sector with a heat load around 213 PJ followed by
the chemical sector with 167 PJ. The Food and Drink, Pulp and Paper, Cement, Glass,
Aluminium and Ceramics sectors are also significant heat users.
With regards to Figure 4, test case studies will consider most of these industrial sectors. Other
important key sectors in the UK process industry are Pharmaceutical and Biotechnology and
Oil/Biofuel as illustrated by the large number of companies in this sector which are members
of the North East of England Process Industry Cluster (NEPIC, website:
http://www.nepic.co.uk/).

2. Potential test case studies


Following the identification of the sectors with the largest heat demand, companies who are
influential in these sectors will be approached to provide case study material. These are the
sectors being considered:
10

- Iron and Steel


- Chemical/ Petrochemical
- Food and Drink
- Pulp and Paper
- Cement
- Pharmaceutical and Biotechnology
- Oil and Biofuel
Given the importance of Small and Medium-sized Enterprise (SME) in the UK economy and
given the large number of SME working in different industrial sectors, one of the test case
studies can be based on the selection of a cluster of companies in the Northeast ( in
collaboration with NEPIC) within which both heat sources and potential users would be
identified. This case study would demonstrate industrial ecology for heat transfer, allowing
the investigation of the potential for synergistic relationships across sectors.

3. Waste heat survey guidelines


It is important to define a methodology while processing industrial partner data for
identifying waste heat.
First, all the thermal energy streams containing sensible and/ or latent heat that flow from the
plant into the environment are identified.
For each thermal energy stream, the following information is given:
Waste heat source composition
Gas moisture content
Flow rate
Temperature
Ease of access for utilisation
Heat content or specific enthalpy
Reliability of supply
Secondly, the accuracy of the collected data will be established by determining the total heat
balance of the system under investigation. This will be verified by double checking with the
industrial partner.
There are a large number of types of plant and equipment from which waste heat is available.
The following three basic heat sources can be identified:
- Gases and vapour
- Liquids
- Solids (the least common category)
Potential for waste heat recovery exists in items common to many sites in the process
industries such as [1]:
- Air compressors
- Boilers
11

- Prime movers
- Refrigeration plants
- Distillation plants
- Dryers
- Dyeing and finishing plants
- Evaporators
- Furnaces
- Gas turbines
- Kilns
- Ovens
- Pasteurisers
- Process coolers
- Process heaters
- Spinning and weaving equipment
- Sterilisation equipment
- Ventilation equipment
- Washers
The standard gas volume reference conditions used in this study will be those of ISO 13443
which defines a temperature of 15C and a pressure of 101325 kPa.

4. Case Study: Steel production process


The steel production consists of 3 main steps:
- Iron making
- Steel making
- Steel casting and rolling
Figure 5 shows a schematic diagram of the steel production process.
Steel production in the UK is concentrated in the Blast Furnace (BF) / Basic Oxygen Furnace
(BOF) route (for primary steel) and the Electric Arc Furnace route (for secondary steel).

4.1. Primary steel production process


Data from a thermal energy audit in a steelworks has been provided by Corus. Data is
averaged over the year. The steelworks produces nearly 5 million tons of steel slabs per
annum. The steel capacity of the plant is given in Table 2 at different stages of the production
process.

12

Iron Ore

Coal

CO gas

Power plant

BF gas

Coke oven

Sinter

Powder

Coke

Blast furnace

Molten Iron

Basic oxygen furnace

Low carbon steel

Continuous casting (Concast)

Slab

Hot mill

Cold mill

Annealing processing line

Coil

13

Figure 5: Schematic representation of a steel production plant


Table 2: Steel capacity

Total BF
capacity
(Mt/yr)

Total sinter
capacity
(Mt/yr)

Total coke
capacity
(Mt/yr)

Total liquid
steel capacity
(Mt/yr)

Capacity
as cast
(Mt/yr)

4.3

4.7

0.9

4.9

4.7

Table 3: Specific energy consumption and energy consumption splits within Steel industry processes

Operation
Coke ovens
Sinter strands
Blast furnace
Basic oxygen
furnace
Continuous
casting

SEC
(GJ/t)
2.95
1.64
14.7
1.44

COG/BFG/
Solid fuel Electricity
natural gas
0.93
0.02
0.08
0.85
0.07
0.75
0.01
0.19

0.39

0.31

Steam

Other

0.05
0.24
0.42

1.00
0.36

Slab mill

2.87

0.64

Hot rolling
Cold rolling
Pickling
Electric Arc
furnace

2.43
1.69
1.27

0.35
0.56
0.67

0.65
0.44
0.33

2.50

0.75

0.25

McKenna [3] produced an approximation of the Specific Energy Consumption (SEC) and
energy consumption splits by process for a steelworks (cf. Table 3). The information
contained in Table 3 can be used in order to determine the total heat balance for each process
of the steelworks under investigation once waste heat sources are identified.
The Steel production plant is composed of the following individual processes:
- Coke oven
- Sinter
- Blast Furnace (BF)
- Basic Oxygen Furnace (BOF)
- Continuous casting
- Hot mill
- Cold mill
- Annealing processing line
- Power plant
Steel production is a continuous process and therefore the waste heat sources are highly
14

consistent over time.


Table 4: gas composition in Steel processes

GAS BLEND
Coke Oven Gas
Blast Furnace Gas
BOS gas
Sinter Gas
Fume
Coke oven flue gas 1
Blast Furnace flue gas 1
Ammonia incinerator gas
Underfiring gas
at the coke oven 3

H2
0.61
0.035
0.02
0
0
0
0
0
0

O2
N2
CO2
CO
SO2
NO2
0.002
0.03
0.017
0
0.07
0
0
0.465
0.25
0
0.25
0
0
0.13
0.15
0
0.7
0
0.1667 0.7562 0.0415
0
0
0
0.21
0.79
0
0
0
0
0.068 0.725 0.052
0
0
0
0.079 0.705 0.202
0
0
0
0.18
0.68
0.01 0.0049 0.00589 0.0665
0.0735

0.715

0.127

CH4
0.245
0
0
0
0
0
0
0

H2O
0
0
0
0.0345
0
0.0156
0.014
0.0526

0.085

Data provided by Corus for this study has been drawn together from various sources.
Information on gases came from the environmental department of the participating site.
Information on cooling water was obtained from cooling tower manufacturers.
Various types of information were collated from the plants control rooms and from contacts
on site.
Data error margin is reported to be 10%. Data are averaged over approximately a year.
Steam waste heat has not been quantified by the thermal energy audit used for this study but
according to Corus, the steam energy waste from the 11 bar system is estimated at 0.83 PJ/yr,
which is equivalent to ~ 5millions of natural gas utilisation. Cardiff University is in the
process of assessing the steam thermal energy losses and redesigning the steam distribution
system.
A description of each process is given in the following sections. For each process, the
low-grade heat sources are identified in orange. Intermediate heat streams are given in red
while incoming gas streams are given in blue and incoming products are given in green.
Each stream is characterised by the temperature, the mass flow rate and the specific enthalpy
calculated at a temperature of 15C and a pressure of 101325 kPa.
For gas heat source analysis, composition is also necessary. Gas molar fractions and molar
masses are given in Table 4.
The same properties are unknown for the input streams but would be necessary for a full
energy balance check as recommended in Section 3.

The composition was determined for 8% oxygen combustion


The ammonia stream is combusted at ~1000C. It was assumed that at such high temperature only NO2 was
formed. The waste gas was then diluted with air to reach 210C temperature.
3
The composition of underfiring coke oven is derived from the mixture of 50% coke oven flue gas and 50%
blast furnace flue gas.
2

15

4.1.1. Coke production process


Coke ovens produce coke from coal for use in the Blast Furnace as a reducing agent and fuel.
Within the oven coal is heated for several hours or days to produce coke through pyrolysis.
The main sources available for thermal energy recovery (cf. orange stream) are gas
underfiring at maximum temperature of 220C and cooling water at 40C. Most of the gas
from the coke oven is reused in the plant and therefore is not available for recovery.
Heat not quantified
(No data available)

Heat available: 46 MW

NH3 gas prior


to dilution

Coal
Cooling water

Raw gas

Coke oven

Air

Cooling + Chemical recovery

Heat available: 15 MW
Coke for quenching

Gas underfiring

Lean gas

Heat available: 21 MW

Heat available: 42 MW

Heat available: 0.9 MW

External Quench of hot coke

Quenched coke

Steam

16

NH3
combustion
gas

Heat available: 3.6 MW

4.1.2. Sinter process


Sinter plants produce the fine powder of iron ore for injection into the BF.
Sinter gas at a maximum temperature of 180C and cooling water at 50C are the main
streams available for recovery.
Iron ore

Mixing

;
Heat available: 0.2 MW

Cooling water
Furnace

Breaker bar + Cooler

Air

Sinter bed

Fan

Powder to BF

EPS

Heat available: 72 MW

Combustion gas

Heat available: 44 MW

Stack 1

Stack 2

End of sinter
strand

End of sinter gas

De-dust gas

Heat available: 7.5 MW


Heat available: 3.6 MW

17

4.1.3. Blast Furnace (BF) process


The blast furnace is the vessel within which iron ore is reduced by coke at high temperatures
to yield pig iron. The main sources available for recovery are cooling water at maximum
temperature of 40C, fume (air) at 50C and BF flare gas at 200C. Combustion gas is reused
in the power plant.
Powder

Limestone

Blast Furnace vessel (BF a/BF b)

Cooling water

Tuyere

Coke

Copperwork
Hot stoves

Gas

Air

Heat available: 29 MW Heat available: 11 MW


Combustion gas (CO, BF)

BF a

BF b

Dust catcher

Heat available: 12 MW

Heat available: 15 MW

Cooling water

;
Heat available: 18 MW

Venturi scrubber

Cast house

Cooler

Fume (Air)
Convector hood

Heat available: 5.6 MW

Skimmer

Gas

Liquid iron

Slag

Heat available: 14 MW

Stack

Heat available: 45 MW
Flare BF gas

18

Heat available: 2.7 MW

4.1.4. Basic Oxygen Steelmaking (BOS) process


The BOF converts pig iron into steel by adding oxygen to remove the carbon, as well as
amounts of silicon, manganese and phosphorous.
The waste heat sources are from the fume (air) at a maximum temperature of 50C, BOS gas
at a maximum temperature of 150C and cooling water at 35C.
Molten iron

Ladle

Heat available: 3 MW
Slag

Desulphurization

Cooling water
BOS gas

Burnt lime

Heat available: 33 MW
Heat available: 1.2 MW

Oxygen

BOS gas

Primary BOS

Extraction A10A

Extraction

Heat available: 1.6 MW


Slag

Fume

Heat available: 22 MW

Heat available: 3.8 MW

Fume

Heat available: 3.8 MW

Secondary BOS

Fluxes

Steel

Secondary cooler

Fume

NG

Laddle preheaters 1 to 4

Combustion gas

Steel for concast

19

; Heat available: 2.5 MW

4.1.5. Continuous casting process


The continuous casting process is a batch wise process in moulds before reheating for rolling.
Water is the only waste heat source available in continuous casting with a maximum
temperature of 42C.
Low carbon steel + Alloy

Heat available: 28 MW

Water spray

Heat available: 16 MW Heat available: 40 MW

Heat available: 29 MW

Heat available: 40 MW

Caster 1

Slab

Water

Water

Water

Caster 2

Steam

Slab

Water spray

Heat available: 9 MW Heat available: 39 MW

20

Caster 3

Steam

Slab

Heat available: 14 MW

Water spray

Heat available: 16 MW

Steam

4.1.6. Hot mill process


The steel still needs to be reheated in order to be malleable enough to roll.
In this process the waste sources are available as water at a maximum temperature of 38C.
Cooling water
Cooling water

Slab yard

Air

Heat available: 8 MW

Cooler

Heat available: 7 MW
Re-heat furnace a

Slab for rolling

Re-heat furnace b

Gas

Heat available: 45 MW

Heat available: 10 MW

Roughing mills

Water return

;
Run-out table

;
Cooling water

Coiler
Heat available: 8 MW

Coils

21

; Heat available: 74 MW

4.1.7. Cold mill process


The metal passes through rollers at a temperature below its recrystallization temperature in
order to increase metal yield strength and hardness.
Note that the fume at 30C and extraction gas (air) at 40C provide less than 1 megawatt of
energy. No data is currently available for the cooling water from this process.
Coil from hot mill

Quenching tank

Cooling water

Not quantified (No data available)

Stretch leveler

Fume (air)

Pickling line

; Heat available: 0.1 MW

Extraction gas

Cold rolling

Coils

Heat available: 0.9 MW

22

; Heat available: 0.4 MW

4.1.8. Annealing process


This process induces metal ductility in steel from the cold mill.
Waste heat source which has been identified and available for recovery is in the exhaust gas
stream from heat treatment process at 600C.
Coil from cold mill

Heat treatment

Exhaust gas

; Heat available: 10 MW

Cooling water from entry


High cooling rate Gas Jet
Cooler (HGJC)

; Heat available: 17.7 MW

Cooling water from exhaust

;
Quench tank 1

Cooling water

;
Quench tank 2

; Heat available: 17.7 MW

; Heat available: 9 MW

; Heat available: 2.9 MW

Cooling water

Temper mill

Accumulator

Electrostatic oiler

Coils

23

4.1.9. Power plant


The power plant recovers (Blast Furnace and Coke Oven) combustion gas heat in order to
convert it into electricity. It corresponds to a Rankine cycle for steam. The excess of steam or
steam bleed-off is drawn from the boiler through the continuous blow down system, the latter
being the only low grade heat source available in the power plant unit.
Steam represents a potential source for recovery. No information is available for steam
bleed-off in the power plant but the total steam loss within the site is estimated to be ~26
MW.
Water to condense steam
(Dock supply)

CO and BF combustion gas


Boiler A

; Heat available: 26.6 MW

Boilers

CO and BF combustion gas


Boiler B

Cooling water to
Condenser

; Heat available: 17 MW

Heat available: 104 MW

; Heat available: 11 MW

CO(Margam
and BF combustion
tion gas
B, Mitchell gas
6&7)
Boiler C

Pump

CO and BF combustion gas


Boiler D

;
Turbine

Heat available: 37 MW

Steam

Continuous blow
down system

Steam (Bleed off)

24

; Heat available: 7 MW

Table 5: Characterization and classification of potentially recoverable low grade heat gas streams in the steel industry
Composition
H2 O

Tout
(C)

Quantity
(kg/s)

Exergy
(MW)

30

12

0.002

40

22

0.014

0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0

0
0
0
0
0
0
0
0
0
0
0

50
50
70
40
200
150
50
50
50
200
180

60
86
32
191
3
10
185
185
245
10
36

0.088
0.125
0.125
0.126
0.148
0.229
0.27
0.27
0.36
0.443
0.734

0.0049

0.0526

210

10.75

0.827

0.2

0.13

220

100

5.128

0.0415

0.0345

130

388

6.666

N/A

N/A

N/A

Location

Type

H2

O2

N2

CO2

CO

CH4

NO2

Cold mill stretch


leveller

Stretch leveller
extraction fume

0.21

0.79

Cold mill Pickle line

extraction gas

0.21

0.79

BOS Primary
BOS Secondary
BOS Primary
BOS primary
BF a
BOS primary
Casthouse (north)
Casthouse (south)
Sinter Dedust
BF b
End of sinter Strand
Ammonia
incinerator
Coke oven gas
underfiring
Main stack
Power plant bleed
off

Hot metal pouring fume


Fume
BOS gas
Hot metal pouring fume
flare BF gas
Dephulsurisation gas
Fume
Fume
Sinter gas
Flare BF gas
Sinter gas

0
0
0.02
0
0.03
0.02
0
0
0
0.03
0

0.21
0.21
0
0.21
0
0
0.21
0.21
0.21
0
0.21

0.79
0.79
0.13
0.79
0.585
0.13
0.79
0.79
0.79
0.585
0.79

0
0
0.15
0
0.128
0.15
0
0
0
0.128
0

NH3 combustion gas

0.18

0.68

0.01

0
0
0.7
0
0.257
0.7
0
0
0
0.257
0
0.005
89

0.3

0.001

0.3

0.075

0.1669

0.7562

mixture of FB and Coke


Oven gas
Sinter gas
Water vapour

Table 6: Characterization and classification of potentially recoverable low grade heat cooling water streams in the
steel industry
Type

Location

Cooling water

Breaker bar cooler (sinter)

Cooling water

Tout(C)

Quantity(kg/s)

Exergy (MW)

50

0.016

Gas wash (BF b)

41

257

0.311

Cooling water

Hot mill re-heat furnace B

38

233

0.337

Cooling water

Hot mill re-heat furnace A

38

218

0.353

Cooling water

Gas wash (BF a)

35

307

0.466

Cooling spray

Caster 3

40

200

0.535

cooling / quench water

Hot mill run out table

35

444

0.599

Cooling water

Caster 3

33

542

0.62

Cooling water

Open cooling (BF a)

35

665

0.651

Cooling water

Copperwork (BF b)

40

1405

0.701

Cooling water

Tuyere (BF b)

37

417

0.81

Cooling water

BOS primary (North & South)

35

565

0.824

Cooling water

Open cooling (BF b)

36

511

0.882

Cooling water

Caster 1

42

316

1.019

Cooling spray

Caster 2

40

486

1.296

Cooling spray

Caster 1

40

495

1.32

Cooling water
Main recirculating cooling
water

Caster 2

40

497

1.32

Coke oven

40

556

1.518

dirty water return

hot mill

35

1827

2.457

25

4.1.10.
Low grade heat classification-Summary for
Primary steel production process case study
As identified in previous sections, the sources of low grade heat come mainly from stacks and
cooling towers. In this section, gas and cooling water streams identified in previous sections
are classified in terms of their exergy (values provided by the steelworks) and characterised
with the properties defined in Section 4.

4.1)10.1. Gas
Table 5 gives the main properties of gas low grade heat streams classified as a function of
their exergy values. For gas, exergy is given by Equation (2). It depends on temperature, mass
flow rate and calorific capacity of the streams. The most exergetic stream has a temperature
of 130C but is available in higher quantity (388 kg/s).

4.1)10.2. Cooling water


Cooling water low grade heat streams are characterised in Table 6.
The exergy of the water streams depends on the temperature difference in the cooling towers.
Streams with highest mass flow rate present the highest exergy.

4.1.11.

Potential uses

Some initial recommendations for waste heat source utilisation are specified in this section.

4.1)11.1. Gas
Table 7 lists the main gas sources identified in the processes as a function of their temperature
range and gives an indication for potential recovery technology.
Table 7: Gas waste heat sources and potential for recovery
Source

Temperature

Potential uses

Typically 30-80 oC

Heat pipes, ORC, Kalina, Biomass


drying, coal drying

Combustion stacks

150-250 oC

Heat pipes, ORC, Kalina, Biomass


drying, coal drying

BF stoves

200 oC

Heat pipes, ORC, Kalina, Biomass


drying, coal drying

CO underfiring gases

220 oC

Heat pipes, ORC, Kalina, Biomass


drying, coal drying

Steam from casters and continuous


blow-down system

Not quantified

Steam boiler

Extraction systems

26

4.1)11.2. Cooling water


If the temperature is high enough (typically > 80C), the Kalina cycle can be used for
recovery of cooling water stream heat (cf. Table 8).
Table 8: Cooling water waste heat sources and potential for recovery

Source

Potential uses

Cooling water
systems with
T<25 C and
high flows

Heat pumps,
space/office/buildings heating

Cooling water
systems with
T>40 C and
high flows

Heat pumps and Kalina,


space/office/buildings heating

4.1)11.3. Concluding remarks


Streams have been characterised by their temperature, mass flow rate, composition (for gas
streams), specific enthalpy and they have been classified in terms of their exergy which
represents the maximum technical work one can get from the stream heat.
Low grade heat sources identified as potentially recoverable mainly come from cooling
towers and stacks throughout the plant and are available either as gas or cooling water. The
specific heat consumed by each Steel making process unit can be approximated as the sum of
the specific steam and natural gas consumptions given in Table 3. For a capacity of 4.9.10 6
tons of steel produced per annum and based on exergy values of the waste streams given in
Table 6 and Table 7, the percentage of low grade heat can be determined for each process unit
as shown in Table 9. This report shows that the recoverable potential for low grade heat
represents approximately 1.1% of the heat consumption. This is nevertheless equivalent to
approximately to 60 MW.
Table 9: Percentage of low grade heat in each unit of the primary steel making process

Operation
Coke ovens
Sinter strands
Blast furnace
Basic oxygen
furnace
Continuous casting
Hot rolling
Cold rolling
Total

Heat
consumption
(MW)
449.1977423
236.9824962
4522.431507
136.4840183
30.82699137

245.4195205
115.5390665
6193.366

Flue gas stream


(%)
1.33
3.27
0.03

Cooling water stream


(%)
0.34
0.01
0.08

0.51
0.00
0.00
0.00
0.25

0.60
19.82
1.53
0.01
0.5

27

Steam
stream (%)

0.42

Low grade heat


(%)
1.66
3.28
0.11
1.11
19.82
1.53
0.01
1.1%

It is worth mentioning that low temperature gas with recovery potential does not exceed
250C in the steel industry although gas streams with higher temperature are available within
the plant but either recovered internally or inaccessible for over-the-fence activities. Gas
streams with the highest exergy are not those with the highest temperature (130C) but those
available in higher quantity.
Potential for cooling water heat recovery is available throughout the steelworks in large
quantities with a maximum temperature of 50C. Despite low temperature range, cooling
water represents a recoverable potential of approximately 30 MW.
The next section attempts to extrapolate this finding to the steel sector.

4.2. Secondary steel production process


Secondary Steel is produced in an Electric Arc Furnace (EAF) in which the scrap is
electrically heated (cf. Figure 6). Figure 7 gives the overview of the mass streams in an EAF.
An estimate of the energy consumed is given in Table 10. The total specific energy
consumption is in the range 2.3-2.7 GJ/t with 1.25-1.8 GJ/t for the electricity use. In the UK
and according to [3], the EAF specific consumption is 2.5 GJ/t and 0.75% of the energy
consumed is electricity which corresponds to about 1.8 GJ/t. As a first approximation, the
heat consumption can be considered as equivalent to the electricity consumption given that
the EAF is perfectly insulated. EAF emission mass streams are given in Figure 7. They
represent the main source of waste energy since the cooling water is not available for
recovery as it circulates in a closed cycle. The European Commission reported in [15] that
85-90% of the emissions from EAF are recoverable and that the primary off gas represents
95% of the emissions. It is therefore conservative to assume that the heat recovery fraction is
15% of the heat consumption (~1.8 GJ/t). According to [16], after cooling at about 200-300
C, the primary gas is mixed with the secondary gas at 50-70 C coming from the canopy
hood situated over the EAF in order to reach the filtering at a temperature typically below
130C. The temperature of the flue gas from the EAF is approximated to 140 C before the
filtering (typically between 130 and 200C). The potential for low grade heat in the
Secondary steel production process sector can be estimated from the calculation of the exergy
of the flue gas stream. Table 11 gives the list of the EAF in the UK with the heat consumption
and potential for low grade heat.
Filtering

Flue gas
for
recovery

Figure 6: Schematic presentation of a Electric Arc Furnace (EAF) [17]

28

Figure 7: EAF mass stream overview [18]

29

Table 10: Input and output mass stream from a EAF [18]

Table 11: EAF low grade heat potential in the UK

Location
Celsa UK (Cardiff)
Thamesteel (Sheerness)
Outokumpu (Sheffield)
Corus UK Ltd (Rotherham)
Forgemasters (Sheffield)

Capacity
(105 tons/annum)
1.2
0.72
0.54
1.25
1.30

Heat consumption
(MW)
71
43
32
74
7.8

Waste heat recovery potential


(MW)
1.92
1.15
0.86
2
0.2

4.3. Overview of the steel sector


The European Union Emissions Trading Scheme (EU ETS) [19] was launched in 2005 to
meet its GHG emissions reduction target under the Kyoto Protocol. The EU has to make an
eight per cent reduction on 1990 levels by 2012.
Under the EU ETS, energy intensive industries must monitor and annually report their GHG
emissions. The list of the energy intensive industries published by the European Commission
in 2007 [20] was used in order to identify the main heat emitters in the sector steel. It is worth
mentioning that some of the large heat emitters recently have or are in the process of closing
down which reduces significantly the potential compared with previous market potential
estimates such as in [21]. The potential for low grade heat as steam, water and flue gas was
estimated for each industry whose emissions were listed in [20] and whose heat consumption
and percentage for low grade heat were estimated, based on the results presented in the
previous sections. The results are summarized in Table 12. The total potential for low grade
30

heat recovery in the Steel sector is estimated to be approximately 137 MW, i.e. approximately
1.2 TWh. This represents less than 10% of the potential estimated in 1994 and reviewed in
[22]. Apart from the fact that some of the large heat emitters have closed down, another
reason for such a difference is that the potential estimated in this study considers the exergy
instead of the energy content which only represents the useful part of the heat sources
identified. This represents nevertheless a significant amount of energy with regards to the
potential uses as discussed in the next section. The low grade heat sources were located on a
map as shown in Figure 8 with the associated potential for low grade heat recovery for both
primary and secondary steel making processes.
Table 12: Low grade heat recoverable potential in the Steel sector in the UK

Type of products

Capacity
5
(10 tn/an)

Gas (MW)

Water (MW)

Steam (MW)

Total
low grade heat
recoverable
potential (MW)

EAF steel

6.21

8.03

8.03

Primary steel
Hot and cold
rolling for
automotive
steel making

9.6

31

35

59

125

3.76

3.76

NA

NA

43

35

59

137

Figure 8: mapping of low grade heat potential in the Steel sector

4.4. Uses for low grade heat


In order to harness the potential of the low grade heat identified in the Steel sector, there is a
31

need for matching the sources with potential end-users in the surrounding of the plant. The
review by Ammar et al. [22] have identified how low grade heat could be used in the future to
reply to the societal needs. The most attractive application is to produce electricity. The
feasibility of such application is however limited to temperature higher than typically
approximately 70C for the most advanced Rankine cycle derivatives. This is the case for
more than 50% of the flue gas streams identified. One of the most important challenges is to
harness the potential of the waste water streams abundantly available at lower temperatures
(between 35 and 55C). As underlined in Section 4.11, space heating/cooling can be an
interesting alternative for low temperature. Typical heat and cooling loads are provided in
Table 13 .
Table 13: Typical heat and cooling loads

Heat load

Maximum thermal
Minimum thermal energy
energy consumption (kW)
consumption (kW)

Sources

energy audit data


Middlebrough
Energy audit data
School (per pupil)
0.15
1.15
County Durham
2
Greenhouse (per m )
55
0.1
Bremer Energie Institut
Data centre (per rack) with COP ~ 2
60
30
[23]
2
Supermarket (per m ) with COP~4
0.16
0.04
[24]
The viability of a low grade heat recovery projectdepends on whether the heat available can
economically be transferred from the source to an identified sink. So far, however, there has
been little discussion about the economic distance from the source to the sink. Industrial heat
is usually transported via water or steam. According to the report by Terra Infirma [25], steam
with a temperature of 120-250C can be transported over approximately 3 to 5 km while
water with a temperature of 90-175C can be transported over 30 km. Other sources cited in
that same report mentioned that 9 miles (around 15km) is the economic limit for low-grade
heat. In fact, how far heat can be transported depends on several factors. If heat is assumed to
be transported via a pipe, the heat loss factor, which is defined as the ratio between heat loss
and the quantity of heat supplied by the source, depends on the pipe material and the
efficiency of its insulation, pipe diameter and the temperature of the fluid circulating in the
pipe. The profitability of any heat recovery project will also depend upon the cost invested in
heat transportation, the total cost being the sum of the pipeline installation, heat losses and
pumping cost [11]. For long distances (typically over 10 km), sorption processes are efficient
heat transportation systems [26] [27] Recently, Lin et al. [28] investigated the performance
and the economics of a 500 MW transportation system over 50 km, with the heat coming
from a nuclear plant. They showed a payback period of 3 years and 8 months for the whole
system. Less waste heat is available for free from the process industry and therefore the
economics of the low grade heat recovery project need to be revised accordingly. However
this is nevertheless evidence of a potentially economic method to transport low grade heat.
The economic distance can be defined as the limit for economically transferring low grade
heat from the source to the sink. The value of the economic distance is likely to increase over
time as the price of the fuel equivalent to the low grade heat recovery savings is likely to keep
Office (per m2)

0.1

0.001

32

increasing. Within the scope of this study, the feasibility of using low grade heat to provide
heating is examined for 3 characteristic transportation radii from the heat source; 1 kilometer,
9 kilometers and 25 kilometers. Port Talbot is chosen as a case study. The heat map produced
by the Department of Environment, Food and Rural Affairs (DEFRA) is used in order to
determine the heat potential from the demand side. The results are summarized in Table 12
for the different characteristic distances from Port Talbot steelworks.
Table 14: Heat consumers at different radii from Port Talbot Steelworks

Heat consumers
Public Buildings (MW)
Commercial Offices (MW)
Hotel and Catering (MW)
Other Services (MW)
Retail (MW)
Sport and Leisure (MW)
Small Scale Industrial (MW)
Domestic (MW)
Schools (MW)
Hospitals (MW)
Warehouses (MW)
Total (MW)

25 km
2.141
0.757
2.642
1.025
2.518
0.613
41.991
100.72
0.866
0.675
1.775
155.7

9 km
0
2
0
8
5
0
0
3
0
0
0
18

1 km
0
0
0
0
0
0
0
0.2
0
0
0
0.2

Most of the heat demand is located more than 9 km away from the heat sources identified in
the steelworks. Within a radius of 25 km, the potential for heat demand overcome the heat
supply with approximately 155 MW. Most of the heat demand comes from households.
Industrial low grade heat could therefore be integrated in a new district heating scheme to
retrofit community heating.

33

References:
1.
Turner C. W. and S. Doty, Waste-heat recovery, in Energy management handbook.
2006, Fairmont Press: Lilburn.
2.
HM Government, Climate Change Act 2008: Chapter 27. 11/26/2008.
3.
Mc Kenna, R.C., Industrial energy efficiency Interdisciplinary perspectives on the
thermodynamic, technical and economic constraints, in Department of Mechanical
Engineering. 2009, University of Bath.
4.
HM Treasury The climate change levy package. 2006: London.
5.
HM Treasury, Budget 2010: London.
6.
The Carbon Trust. About the Carbon Trust.
[cited 2010; Available from:
http://www.carbontrust.co.uk/about.
7.
Profiting from low-grade heat, The watt committee on Energy report No. 26. 1994,
London: The institution of Electrical Engineers.
8.
Perrot, P., A to Z of Thermodynamics. 1998: Oxford University Press.
9.
Winter, C.-J., Energy efficiency, no: It's exergy efficiency! International Journal of
Hydrogen Energy, 2007. 32(17): p. 4109-4111.
10.
Potential uses of waste heat from a proposed new power station at Blyth. Report by
Terra Infirma for the National Industrial Symbiosis Project (NISP). 2008.
11.
Hlebnikov, A. and A. Siirde, The major characteristic parameters of the Estonian
district heating networks and their efficiency increasing potential. Energetika, 2008.
54(4): p. 67-74.
12.
BERR, Heat Call for Evidence. 2008, Department for Business Enterprise &
Regulatory Reform: London.
13.
Choudrie, S.L., et al., UK Greenhouse Gas Inventory, 1990 to 2006. 2008, DEFRA:
London.
14.
BERR, ECUK Table 1.14: Overall energy consumption for heat and other end uses by
fuel 2006. 2008: London, [spreadsheet].
15.
Technical Note on the Best Available Technologies to Reduce Emissions of Pollutants
into the Air from Electric Arc Steel Production Plants. 1994, European Commission.
16. Griffini, G.P.a.N., De-dusting plants for electric arc furnaces, in Millinium Steel. 2005,
VAI Pomini SrI: Millan, Italy.
17.
Kirschen, M., L. Voj, and H. Pfeifer, NO2 emission from electric arc furnace in steel
industry: contribution from electric arc and co-combustion reactions. Clean
Technologies and Environmental Policy, 2005. 7(4): p. 236-244.
18.
Best Available Techniques Reference Document on the Production of Iron and Steel.
2001, European Commission.
19.
Ellerman, A.D. and B.K. Buchner, The European Union Emissions Trading Scheme:
Origins, Allocation, and Early Results. Review of Environmental Economics and
Policy 2007. 1(1): p. 66-87.
20.
EU ETS Phase II National Allocation Plan (2008-2012). Appendix E: NAP data. 2007,
European Commission.
21.
Boddy, J.H., Sources of heat, in Profiting from low-grade heat, Thermodynamic cycles
for low-temperature heat sources, The Watt Committee on Energy Report No. 26, A.W.
Crook, Editor. 1994, The institution of Electrical Engineers: London.
22.
Ammar, Y., et al., Review of low grade thermal energy sources and uses from the
process industry in the UK. Special issue, Applied Energy Journal, In press.
23.
Almoli, A., et al., Computational Fluid Dynamic Investigation of Liquid Rack Cooling
in data centres. Special issue. Applied Energy Journal, In press.
24.
Lazzarin, R.M. and F. Castellotti, A new heat pump desiccant dehumidifier for
supermarket application. Energy and Buildings, 2007. 39(1): p. 59-65.
34

25.
26.

27.
28.

Bujak, J., Energy savings and heat efficiency in the paper industry: A case study of a
corrugated board machine. Energy, 2008. 33(11): p. 1597-1608.
Mazet, N., et al., Feasibility of long-distance transport of thermal energy using solid
sorption processes. International Journal of Energy Research, 2010. 34(8): p.
673-687.
Kang, Y.T., et al., Absorption heat pump systems for solution transportation at
ambient temperature -- STA cycle. Energy, 2000. 25(4): p. 355-370.
Lin, P., R.Z. Wang, and Z.Z. Xia, Ammonia-water absorption cycle - a prospective
way to transport low grade heat energy over long distance, in SET2010 -9th
International Conference on Sustainable Energy Technologies. 2010: Shanghai,
China.

35

You might also like