You are on page 1of 24

LITERATURE REVIEW

Varma et al. studied out the possible heat recovery potential from a cement plant
and design of cogeneration plant. They carried out energy and exergy
approaches with possible improvements. Their case study is based on 7100 TPD
capacity cement plant at Sagar cement. The designed thermal power plant is
using simple Rankine cycle with deaerator. The steam turbine inlet condition is
maintained at 15 bar and 390 C. Their result conclude that increase in steam
pressure decreases the thermal efficiency of the cogeneration plant. The
increase in steam pressure increases the hot gas exit temperature at the
evaporator at a fixed pinch point which drops the heat recovery in superheater
and evaporator. Pinch point temperature is kept 30 K. They also shown the effect
of increase in plant capacity on the coal requirement, process heat and heat
recovery. The implementation of cogeneration plant in this case study increase
the energy utilization factor of existing plant from 0.60 to 0.63. A typical 15 MW
(before Cogeneration) cement plant in Andhra Pradesh was selected as the case
study . Plant runs on dry process with 6 stage preheater and a calciner. The
production capacity is 1600 TPD and 5500 TPD from both the kilns. It is the
pyroprocessing unit that includes the preheater, calciner,kiln and the grate
cooler. The streams into the system are the raw material, air into the cooler and
the coal fired into the kiln and the calciner. The streams leaving the system are
clinker out from the cooler, the exhaust gas from the preheater and the hot air
out from the cooler. In the preheater, the raw meal undergoes a series of
concurrent heat exchanges with the hot exhaust gas from the kiln system. The
gas and the material stream are separated by cyclones, after each heat
exchange process. The raw meal temperature increases from 80 0C to 1000 0C
within 40 seconds. The calcinated material entering the kiln, then undergoes a
long heating process.The material temperature rises from 10000C to14500C.
Mineral matrixes of raw material are totally destroyed and cement minerals are
formed at the sintering temperatures. A semi product called clinker is formed.
Coal and other alternate fuels are used as energy sources for the process. The
ash fromfuels is absorbed into the clinker matrix. The residual heat from the
clinker leaving the kiln is recovered by grate cooler to reduce the energy
requirement. The clinker cooler cools the hot granular mass by quenching air
into it bringing temperature down to <1100C. So air become hot and clinker
cold. This hot air is then utilized as combustion air for the firing system of the
kiln. The relevant data is collected from a cement factory in Andhra Pradesh. The
requirement of the plant is 15 MW. Hence, a target of 15 MW power from waste
heat recovery system to be explored. Possibility of increasing the power by
additional fuel firing (coal/gas/oil/other fuels) can also be explored.Considering
necessary data, preheater gas at 1.6Nm/KgCl at 3300C and cooling down to
2260C, Taking mid air takeoff of 0.7 kg/kgCl at 4200C. Balance gas at cooler
vent (after mid tap) 0.55 kg/kgCl at 176 0C will be heated up to 420 0C by using
HAG and the mixed gas(0.7+0.55 = 1.25 lg/kgCl at 420 0C) will be fed to the
cooler boiler. Necessary pre duster to reduce the dust content in the cooler gas
will be used. The combined cooler gas will be cooled to maximum possible limit.
The steam from preheater will be routed through the final superheater in the

cooler boiler and the superheated steam will be expanded into the turbine. The
above points are common for both the lines. To calculate the gas volume, 1600
TPD for line 1 and 5500 TPD for line 2, will be considered as Kiln operating
capacity as provided. The steam from both the lines will be fed into a common
turbine. Balance of plant will be common for boththe lines. Individual boiler
considered for each line to ensure easy operation of pyro system. In addition to
above an optional boiler at raw mill baghouse has been considered. All the
baghouse gases will be tapped after baghouse but before fan and this gas will be
heated upto 3300C with a dedicated HAG.
As the plant requires 15 MW of power which is purchased from APGENCO can be
completely produced inside the plant using the waste heat from preheater gases
and cooler vent gases. The steam from the preheater boiler will be routed
through the final super heater in the cooler boiler and the superheated steam will
be expanded in the turbine. From the theoretical calculations it is observed that a
total power of 12.28 MW can be produced by the plant itself when considering
the condenser operating pressure at 0.4 bar, 12.28 MW of power can be
produced. Additionally the industry requires 2.72 MW of power to run the plant.
So solar power is utilized to generate the additional 2.72 MW of power. Finally,
using cogeneration the waste heat can be effectively utilized to generate power
as large amounts of heat are simply released to the atmosphere resulting in
global warming. Hence, using these waste heat large amounts of power can be
generated keeping the environment safe from radiation and global warming
Less fuel is required to produce a given amount of electrical and thermal energy
in a single cogeneration unit than is needed to generate the same quantities of
both types of energy with separate, conventional technologies. fuel etc. has
enabled drastic reduction in energy consumption and increased productivity.
However, the Indian cement industry is yet to make a beginning for the adoption
of cogeneration technology due to existence of various technical and financial
barriers [1]. A cogeneration or combined heat and power (CHP) system produces
steam that provides thermal energy to heat
Production of Cement is one of the most energy intensive industries in the world.
In cement plants, clinker and rotary kilns are widely used for raw material
production. In the dry process cement plants nearly 40 % of the total heat input
is rejected as waste heat from exit gases of
preheater and grate cooler. This
waste heat can be effectively utilized for electric power generation. Cogeneration
of power besides reducing the problem of power shortage also helps in energy
conservation as well as reducing green house gas Emissions. Cogeneration
systems have been successfully operating in Cement plants in India and Asian
Countries. In existing plants cogeneration technologies based on bottoming
cycles have potential to generate up to 25-30 %of the power requirement of a
plant. However, the Indian cement industry is not utilizing the full potential of
Cogeneration due to existence of various technical and financial constraints. An
investigation on a cement industry is conducted and noticed that, industry
requires 15 MW of power for running the plant. So they are completely
depending on APGENCO. One solution that continues to gain profits is to produce

their own power using cogeneration. In this paper, using waste heat recovery
systems and supplementary firing, around 70-75% of power is supplied to the
cement industry. By theoretical approach, for the given input data of the plant,
12.28 MW of power can be produced using cogeneration.exchangers and
mechanical energy through expansion to turbine units. The turbine units then
transfer the mechanical energy to generators, which in turn produce electricity
[2]. World demand for cement was 2283 million tonnes in 2005and China
accounted for about 47% of the total demand. It is predicted that the demand
will be about 2836 MT in the year 2010. China will increase its demand by 250
million tons during this period. This increase will be higher than the total annual
demand for European Union. It was reported that Japan and the US, India is the
fourth largest
Karellas et al. carried out the comparative study of water steam Rankine cycle
and organic Rankine cycle (ORC) used in cement industries based on pinch point
analysis. They concluded that for high temperature of the exhaust gases watersteam Rankine cycle is efficient solution but in modern cement plant
temperature of exhaust gas is low due to high efficiency of the plant. In such a
case use of organic Rankine cycle (ORC) is advantageous. In their study they had
taken isopentane as a working fluid for ORC cycle. Increasing the number of
stages of suspension preheater reduces the gas temperature. Four stage
preheater provides the exhaust gases at 300-380 C, while 5-6 stage preheater
provides exhaust gases at 200-300 C. But it is limited upto six due to structural
limitations. Their research conclude that the maximum temperature of
superheated steam is set up 10 C lower than the inlet temperature of hot air in
Air quench cooling (AQC) boiler. The hot air from clinker cooler is available at a
temperature of about 360 C. The tapping of air from different point of cooler
gives the different temperature. The point closer to the kiln exhaust can be at
500 C than point at the center of clinker cooler that is at 300 C. On the other
side the mass flow rate of hot air is higher at the center of the cooler than point
nearer to exist of the kiln.
This paper presents waste heat recovery as a way to gain energy from the
exhaust gases in a cement plant. In a typical cement producing procedure, 25%
of the total energy used is electricity and 75% is thermal energy. However, the
process is characterized by significant heat losses mainly by the flue gases and
the ambient air stream used for cooling down the clinker. About 35% - 40% of the
process heat is lost by those waste heat streams [8]. Approximately 26% of the
heat input to the system is lost by dust, clinker discharge, radiation from the kiln
and pre-heater surfaces, and convection from the kiln and pre-heaters. A heat
recovery system could be used to increase the efficiency of the cement plant and
thus lower the CO2 emissions. Moreover, it would reduce the amount of waste
heat to the environment and lower the temperature of the exhaust gases. Waste
heat can be captured from combustion exhaust gases, heated products, or heat
losses from systems. This study aims at the identification of a best practice
example for energy utilization in an existing commercial cement production plant
with a waste heat recovery system as a new component. Two different methods
will be examined, using the commercial software IPSEpro by Simtech. Firstly, a

water-steam Rankine cycle will be analyzed and then an Organic Rankine Cycle
(ORC) with an intermediate pressurized water circuit will also be investigated.
Another aim of this paper is the optimization of the working fluid, the maximum
pressure and temperature of the two cycles as well as the components
arrangement, in terms of system efficiency and output power. Finally, an
exergetic analysis is done for both cycles.
Khurana et al. reviewed the utilization of waste heat and carried out energy
balance using Sankey diagram. They conclude that the use of waste heat of
exhaust gases increase the overall thermal efficiency of the plant by 10%. They
used 1 Mt per annum working cement plant runs on dry process with five stage
suspension preheater to carry at an energy balance for pyro-processing (clinker
formation) unit. The gases coming from preheater exhaust are highly dust laden
due to that separate boilers are required for both the streams. The formation of
sulphur oxides is mainly due to the coal combustion. The quality of coal in India
very low so burning of coal produce the acidic gases which causes corrosion
outside the tubes. The predominant result was that the temperature at the
outlet in both the boiler gas stream must be above the 120 C (Above acid due
point temperature) to avoid the corrosion problems over the tubes. The specific
energy consumption of this plant is 3.7 GJ/t of clinker and it is one of the efficient
plant in India suitable as a reference case for study. Prediction of fouling
characteristics of the dust laden gases only be done by experimentation.
The cement clinker production sector is a substantially energy intensive industry
accounting for 5060% of the production costs while is currently contributing
about 5% to the global anthropogenic emissions.Thermal energy demands
depend on the age of the plant and on the specific process but ranges between
3000 and 6500 MJ/tone clinker. The average specific energy consumption is
about 2.95 GJ per ton of cement produced for well-equipped advanced kilns,
while in some countries the consumption exceeds 5 GJ/ton. The electric energy
demand ranges from 90 to 150 kWh per cement ton . In a typical cement plant,
25% of the total energy used is electricity and 75% is thermal energy. However,
the process is characterized by significant heat losses mainly by the flue gases
and the ambient air stream used for cooling down the clinker. About 35% - 40%
of the process heat is lost by those waste heat streams . Approximately 26% of
the heat input to the system is lost by dust, clinker discharge, radiation from the
kiln and pre-heater surfaces, and convection from the kiln and pre-heaters . A
heat recovery system could be used to increase the efficiency of the cement
plant and thus lower the CO2 emissions. Moreover, it would reduce the amount
of waste heat to the environment and lower the temperature of the exhaust
gases . Waste heat can be captured from combustion exhaust gases, heated
products, or heat losses from systems . Waste heat recovery systems are already
in operation in various industries with success. In China, Canada, the Gold Creek
Power Plant has a heat recovery system that produces 6.5MW power using ORC
technology. In India, the A.P. Cement Works with 4 MW and ORC technology.
Another cement industry that uses waste heat recovery is Heidelberger Zement
AG Plant in Lengfurt (Germany) with 1.5 MW power and ORC technology. In
addition to these industries, a new waste heat recovery system is under

construction in Rohrdorf (Germany) with 6.8 MW power and water-steam cycle


technology. This study aims at the identification of a best practice example for
energy optimization in an existing commercial cement production plant with
waste heat utilization as a new component. Two different methods will be
examined in order to find which is more beneficial and more efficient for a
Cement industry. Firstly, a water-steam Rankine cycle will be analyzed. The basic
characteristics of this cycle are the two drums with 19 bar pressure and a
maximum temperature of 350 C. The other is an Organic Rankine Cycle (ORC) in
an indirect cycle with pressurized water at 30 bar. Part of this study was the
evaluation of several organic fluids. It was concluded that isopentane has the
optimum performance. Thus, any further analysis was carried out considering
isopentane as the organic working fluid of the ORC. Many parameters were
optimized in order to design the optimum thermodynamic cycle, in terms of
energetic and exergetic efficiency. Pressure and temperature are the most
important parameters regarding the efficiency of those systems. Also changes
have been made in the arrangement of the cycle and its different components in
order to improve the efficiency and design an optimum system. Aiming to define
the cycle with the best performance, energy and exergy analysis will be done in
order to find the cycle with the highest thermal and exergetic efficiency.
The first cycle that will be examined in this paper is a water-steam Rankine cycle
in order to recover the waste heat from the cement industry process. The cycles
maximum pressure is 19 bar at a maximum max temperature of 350 C at the
inlet of the turbine and 0.06 bar at the exit of the turbine. The exhaust steam of
the turbine is condensed in the condenser and then pumped to the deaerator
tank. Simultaneously, some of the steam is extracted from the turbine at 1 bar in
order to be used in the deaeration process. After that, the condensate goes
through the feed pump and enters the air preheater where it is preheated to 200
C. From that point, the feed water is separated into two streams. The first
stream is preheated, evaporated and superheated utilizing the energy from the
cooling air heat source. The other stream follows the same process utilizing the
exhaust gas heat source. Each steam generator system consists of a drum and
two heat exchangers. Finally, the two streams of superheated steam enter the
steam turbine and the process is repeated
In an ORC heat recovery system there is an intermediate heat transfer fluid in
order to transfer the heat from the heat sources to the working fluid through heat
exchangers. This is necessary for safety reasons, as many organic fluids are
inflammable and in case of failure of the heat exchanger the hot medium of the
heat source and the organic fluid would get in contact resulting in an explosion.
The heat transfer fluid should remain in liquid state and thus pressurized water
at 30 bar is ideal for this use. It is important not to have steam, because steam is
not able to transfer the heat to the organic fluid as effectively as water.
The two cycles were simulated with IPSEpro .The water-steam cycle has a
system efficiency of 23.58% producing 6.26 MW electric power, whilst the ORC
has a thermal efficiency of 17.56% producing 4.66MW electric power. More
specifically, the system with water-steam cycle has the thermal efficiency and

the heat exchangers efficiency 28.53% and 82.7% respectively and for the
system with ORC has 21.23% and 82.7%. The other thermodynamic tool that is
used for the comparison of the cycles is the exergy analysis. After the analysis
with IPSEpro, exergy efficiencies were calculated at 32.56% for the
watersteam cycle and at 24.00% for the ORC. Once more, it is shown that the
water-steam cycle has a better performance in these conditions. The main
reason for this result is the higher maximum temperature of the water-steam
cycle which results in more work produced by the turbine and the existence of
the intermediate pressurized water circuit in the ORC, which causes additional
exergy losses. Another way to determine the most efficient system is to
compare the Q-T diagrams (Fig. 3, 8). This can be done by comparing the area
between the lines of the heat source and the working fluid. In the ORC, this area
is bigger, which means that the exergy destruction is higher and the
performance of the system deteriorates. The reasons for this are the use of
pressurized water and the lower critical point of the isopentane in comparison to
waters. All the results of the exergy analysis are summarized in Table 4. For the
water-steam cycle it is clear that the main exergy loss is the gas exhaust (Table
4a) because of the high exit temperature (270 C) of the exhaust gas, which is a
restriction imposed by the production process of the cement plant. The next
major exergy waste is located in the two heat exchangers that operate as
evaporators. For the ORC, the main exergy loss is also located in the gas exhaust
(Table 4b) and in the two heat exchangers. Another useful tool for the exergetic
comparison of the two systems is the Grassmann diagrams which are presented
in Fig. for the water-steam and the ORC system respectively. From these
diagrams it is clear that the main reason which makes the water-steam system
more efficient is the lower exergy losses in the heat exchangers. More
specifically, the water-steam system has 21.8% losses compared to 26.7% of the
ORC system.
After analyzing the performance of the two systems, it is possible to make an
estimation of the avoided CO2 emissions. According to the energy mix of Greece,
an average of 0.85 tones of CO2 are emitted per produced MWhe. In addition it is
assumed that the cement plant has an annual operation time of 7000 hours and
that all the energy produced by the heat recovery system is either consumed by
the plant itself or delivered to the national power grid
Hasanbeigi et al. studied out 16 cement plants with new suspension preheater
and pre-calciner kiln technology. They compared 16 cement plant in Shandong
province, China with international and domestic Best practice cement plants in
terms of the use of different raw materials and the type of cement produced.
They done benchmarking of all plants by comparing Energy Intensity Index
(EII) of these plants with suitable benchmarked plant on both domestic and
international level. Comparison with the international best practice, none of
cement plants were at or near this benchmark. The range of EII compared to
international best practice is 118 to 158. Comparison with the domestic best
practice, 5 of 16 surveyed plants were quite close to best practice. The range of
EII compared to domestic best practice is 102 to 132. Average primary savings of
12% and 23% is available when compared to domestic and international best

practices respectively. They developed 34 energy efficient technologies for their


case study and found out that 29 measures were applicable to cement plants in
which 23 measures are electricity saving measures and 6 of which are fuelsaving measures. Their study also shown the CO2 emission reduction potential
as the result of implementation of the energy efficient technologies or
benchmark applied to the plants.
There are three basic types of VSKs: ordinary, mechanized, and improved. In
ordinary VSKs, high-ash anthracite coal and raw materials are layered in the kiln,
consuming high amounts of energy while producing cement of inferior quality
and severe environmental pollution. Mechanized VSKs use a manually operated
feed chute to deliver mixed raw materials and fuel to the top of the kiln.
Improved VSKs been upgraded and produce higher quality cement with lower
environmental impacts (Sinton, 1996; ITIBMIC, 2004).
Rotary kilns can be either wet or dry process kilns. Wet process rotary kilns are
more energyintensive. Energy-efficient dry process rotary kilns can be equipped
with grate or suspension preheaters to heat the raw materials using kiln exhaust
gases prior to their entry into the kiln. In addition, the most efficient dry process
rotary kilns use precalciner to calcine the raw materials after they have passed
through the pre-heater but before they enter the rotary kiln (WBCSD, 2004).
Construction of these modern NSP kilns has been growing rapidly in China since
about 2000. Large and medium sized NSP kilns produced 56 Mt (10%) of cement
in China in 2000, increasing to 623 Mt (50%) by 2006 (ITIBMIC, 2004; CCATC,
2008).
Globally, coal is the primary fuel burned in cement kilns, but petroleum coke,
natural gas, and oil can also be combusted in the kiln. Waste fuels, such as
hazardous wastes from painting operations, metal cleaning fluids, electronic
industry solvents, as well as tires, are often used as fuels in cement kilns as a
replacement for more traditional fossil fuels. In China, coal is used almost
exclusively as the fuel for the cement kilns, while electricity both provided by
the grid and through the generation of electricity on-site using waste heat is
used to power the various grinding mills, conveyers, and other auxiliary
equipment. In 2007, Chinese cement kilns used 174 Mt of mostly raw coal and
119 terawatt-hours (TWh) of electricity (CCA, 2009). There is very little use of
alternative fuels (defined as waste materials with heat value more than
4000kcal/kg for cement clinker burning) or co-processing of waste materials
(defined as the incineration of wastes for disposal purposes even if the calorific
value of the waste can be used as a fuel) in cement production in China (Wang,
L., 2008). Less than 20 cement facilities either burn alternative fuels or coprocess waste materials as demonstration or pilot projects, but Chinese laws and
industrial policies now encourage the use of alternative fuels and the National
Development and Reform Commission (NDRC) has begun efforts to develop a
Cement Kiln Alternative Fuel Program that will expand the demonstration
projects, prepare regulations, develop a permitting-type system, and establish
financing mechanisms (Wang, S., 2007).

Once clinker has been produced in either a shaft or rotary kiln, it is inter-ground
with additives to form cement. Common Portland cement is comprised of 95%
clinker and 5% additives. Blended cement is the term applied to cement that
made from clinker that has been interground with a larger share of one or more
additives. These additives can include such materials as fly ash from electric
power plants, blast furnace slag from iron-making facilities, volcanic ash, and
pozzolans. Blended cements may have a lower short-term strength (measures
after less than 7 days), but have a higher long-term strength, as well as
improved resistance to acids and sulfates. In 2007, 5.4% of the cement produced
in China was Pure Portland Cement, which is defined as either being comprised of
100% clinker and gypsum or >95% clinker and gypsum with <5% of either
granulated blast furnace slag (GGBS) or limestone. Common Portland Cement,
comprised of >80% and <95% of clinker and gypsum combined with >5% and
<20% of additives (GGBS, pozzolana, fly ash, or limestone), made up 54% of the
cement produced in China that year. Slag Portland Cement, that blends
anywhere from >20% to <70% GGBS with clinker and gypsum, constituted 36%
of 2007 cement production. The remaining 5% of cement was Pozzolana (>20%
to <40% pozzolan additives), fly ash (>20% to <40% fly ash), or other blended
cement (>20% to <50% other additives) (Wang, L., 2009).
Given its large size, complexity, and global importance in terms of both energy
consumption and greenhouse gas (GHG) emissions, the cement sector in China is
receiving increasing attention among analysts, policy-makers, and others around
the world. Early analyses of the industry in the 1990s focused on improvements
that could be made to VSKs as well as scenarios exploring the energy savings
possible with increased adoption of more modern pre-calciner kilns (Liu et al.,
1995) and developments related to mechanized VSKs which at the time were
less energyintensive than both non-mechanized VSKs and the currently-used
rotary kilns (Sinton, 1996). However, future efforts and studies focused more on
the NSP rotary kilns as the structure of the Chinese cement industry changes
towards this type of kiln (Hasanbeigi et al. 2010c, Ziwei Mao, 2009, and Liu et al.
2007).
The study presented in this report is unique for China as it provides a detailed
analysis of energy efficiency improvement opportunities for the entire cement
industry in China. In addition, compared with other international studies, the
potential application of a larger number of energyefficiency technologies is
assessed. The objective of this study is to assess the potential for energy saving
in the Chinese cement industry using a technology-level, bottom-up approach
and to estimate the cost associated with this potential. These results can guide
policy makers in designing better sector-specific energy efficiency policy
programs.
In this report, we first briefly present an overview of the cement industry in
China. In the next section, the methodology will be presented. After that, we
present the technologies and measures available for energy-efficiency
improvement and greenhouse gas (GHG) emission reduction in the cement
industry, and conduct the technical and cost assessment for implementing those

measures in China. We use the concept of a Conservation Supply Curve (CSC)


(Meier 1982) to construct a bottom-up model in order to capture the costeffective potential as well as the technical potential for energy efficiency
improvement and CO2 emission reduction. Finally, we illustrate the results of the
analysis and some of the barriers to the implementation of the efficiency
measures in the cement industry.
The data collection in this report draws upon work done by Lawrence Berkeley
National Laboratory (LBNL) on the assessment of energy efficiency and CO2
emission reduction potentials of the cement industry in the U.S. (Worrell et al.
2000; Worrell et al., 2008; LBNL & ERI, 2008) and in Shandong province of China
(Hasanbeigi et al. 2010), as well as other references. In this report, we have
included two categories of the technologies:
1) International Technologies
International technologies are defined in our study as technologies that are
manufactured outside of China. The data on the energy saving, cost, lifetime,
and other details on each technology were obtained from these LBNL reports,
which are based on case-studies around the world.
2) Chinese Domestic Technologies
In addition to international technologies, we also obtained data on the energy
saving, costs, penetration rate of Chinese domestic technologies for the cement
industry.
Because we could not obtain Chinese domestic technology information for all the
energy efficiency measures/technologies in our list, the base case analysis in this
report is done based on international technologies only. Then, we developed an
alternative scenario in which we applied the penetration of Chinese domestic
technologies which is further explained.
Many of the international energy-efficient technologies examined in LBNL
publications and reports are used in this analysis because other studies on
energy efficiency in the cement industry do not provide consistent and
comprehensive data on energy savings, CO2 emission reductions, and the cost of
different technologies. Information on some of the technologies examined,
however, was presented in other studies (e.g. CSI/ECRA, 2009). Furthermore, the
methodology used for this analysis, i.e. construction of the energy CSC and
abatement cost curve, is also used by LBNL for the cement industry (Sathaye et
al. 2010, Worrell et al., 2000, Hasanbeigi et al. 2010d).
The national level data for the production of different products for Chinas
cement industry was obtained from China Cement Almanac (China Cement
Association, 2009) and from the China Cement Association. For the penetration
rate of the energy efficient measures, a questionnaire was developed and sent to
individual experts in China (see Appendix 4 for a copy of the questionnaire used).
In addition, we also benefited from the penetration rates published in Hasanbeigi
et al. (2010c) which is based on the detailed survey of sixteen NSP kilns cement

plants in Shandong province in China. Finally, we obtained some data from the
National Key Energy Conservation Technologies Promotion Catalogue published
by National Development and Reform Council (NDRC, 2008, 2009, 2010). The
methodology used for the analysis consists of five main steps as follows:
1. Establish 2009 as the base year for energy, material use, and production in
the cement industry. The base year is also used to calculate the costs in constant
base year dollar. The study period for which the CSC was developed is 20102030. Thus, the implementation of the measure starts in 2010. This is different
from some other studies such Sathaye et al. (2010) where the application of
energy efficiency technologies and the cost-effectiveness is assessed only for the
base year.
2. Develop a list of commercially available energy-efficiency technologies and
measures in the cement industry to include in the construction of the
conservation supply curves. We assumed that the energy efficiency measures
are mutually exclusive and there is no interaction between them. The 23 energy
efficiency measures/technologies are used in this study based on their
applicability to the Chinese cement industry as well as the significant energy
saving that can be achieved by the implementation of them. This information
was obtained from previous LBNL study for a group of 16 cement plants in China
(Hasanbeigi et al. 2010c).
3. Determine the potential application of energy-efficiency technologies and
measures in the Chinese cement industry in the base year based on information
collected from several sources. We assumed 70% of the potential for energy
efficiency measures will be realized by the end of 2030 (3.5% per year) (except
for a two measures, replacing a ball mill with vertical roller mill in finish grinding
and the use of alternative fuels, which were treated differently), with a linear
deployment rate assumed between the start year (2010) and end year (2030)
. 4. Obtain the annual forecast data for clinker and cement demand up to 2030.
The adoption rate explained in step 3 was based on the base years production
capacity. However, there will be new capacity installed between 2010 and 2030
to meet increased demand. Additionally, there will be plant retirements in the
existing capacity that will be replaced with new capacity. To define the potential
application of the measures to the new production capacity, we used the new
capacity with EE implementation indicator. By defining this indicator, we take
into consideration how much of the new capacity will have already implemented
the energy efficiency measures from the start and how much potential will still
exist in each subsequent year. We apply the same adoption assumptions to the
retired and replaced capacity as we do to the new capacity.
5. Construct an Electricity Conservation Supply Curve (ECSC) and a Fuel
Conservation Supply Curve (FCSC) separately in order to capture the
accumulated cost-effectiveness and total technical potential for electricity and
fuel efficiency improvements in the cement industry from 2010 to 2030. For this
purpose, the Cost of Conserved Electricity (CCE) and Cost of Conserved Fuel
(CCF) were calculated separately for respective technologies in order to

construct the CSCs. After calculating the CCE or CCF for all energy-efficiency
measures, rank the measures in ascending order of CCE or CCF to construct an
ECSC and a FCSC, respectively. Two separate curves for electricity and fuel are
constructed because the cost-effectiveness of energy-efficiency measures is
highly dependent on the price of energy. Since average electricity and fuel prices
are different and because many technologies save either solely electricity or fuel,
it is appropriate to separate electricity and fuel saving measures. Hence, the
ECSC with discounted average unit price of electricity only plots technologies
that save electrical energy while the FCSC with discounted average unit price of
fuel only plots technologies that save fuel.
In addition, industrial production capacity may grow between 2010 and 2030. To
determine the implementation potential of efficiency measures in the new
additional capacity, we did the following. First, we used estimated production
capacity growth from (Ke et al. 2012) and assumed that a certain proportion of
the new capacity will adopt the efficiency measures autonomously each year (4%
per year between 2010 and 2030) as a result of the installation of new efficient
technology in the new stock (gray angular striped area in Figure 2Error!
Reference source not found.). Since the autonomous implementation of the
measure in some of the new capacity will occur regardless of new policies, the
savings potential of the autonomous implementation is excluded from the supply
curves calculation. Second, the new capacity with additional potential for
implementing the efficiency measures (not captured in autonomous
improvement) is determined for each year (blue angular striped area in Figure 2).
We assumed that a certain portion of the new capacity with additional potential
for implementing the efficiency measures adopts the measures each year (2%
per year between 2010 and 2030) (the red angular striped area in Figure 2). We
treat the retired and replacement capacity the same as new capacity expansions
by assuming the same rates for autonomous adoption of energy efficiency
measures and adoption rates within the additional potential for implementing the
efficiency measures (the horizontal striped area in Figure 2). Because the new
capacity and retired and replaced capacity are both calculated as the product of
growth rates and the adoption rates, the resulting wedges are not always
straight lines (e.g., gray stripped areas both horizontal and angular). To sum up,
the red solid and red stripped areas in Figure 2 is the total source of energy
saving potentials captured on the supply curves.
In forecasted years when the demand for cement declines either relative to the
previous year or even relative to the base year, which is the case for the Chinese
cement demand forecast, we assumed that new capacity added after 2009
remains in production. Thus, we assumed that reduced demand results in a
reduced production at inefficient plants. However, we first estimated energy
efficiency adoptions in the existing capacity regardless of reduced demand.
Therefore, if the demand decline between 2010 and 2030 is large enough, the
entire inefficient capacity can reach the decommissioning or zero production
point within this period. This results in saturated adoption in the remaining
existing capacity and no additional adoptions are possible since the entire
existing capacity has either adopted the measures or been decommissioned by

the saturation year. This represents one approach to deal with the sharp
decreased cement demand in the future. An extreme case in the opposite
direction is that production never falls despite domestic demand reductions and
instead excess production is exported resulting in the same energy consumption,
emissions, and energy efficiency adoption potential as would be the case if
demand kept rising. Because of the transportation costs, exporting cement is
not a highly profitable trade and Chinese companies are not exporting a high
volume of cement either. However, a large domestic demand reduction could put
considerable downward pricing pressure on the cement industry and could result
in significant exports in the future. Another case could be the export of old yet
not retired equipment to another country when Chinese domestic demands fall
considerably and exporting cement would not be attractive. We have no way of
modeling exported equipment and therefore made a conservative assumption
that inefficient capacity will no longer be available within China to adopt energy
conservation measures.
Although the CSC model developed is a good screening tool for evaluating the
potentials of energy-efficiency measures, the actual energy savings potential
and cost of each energyefficiency measure and technology may vary and depend
on various conditions such as raw material quality (e.g. moisture content of raw
materials, hardness of the limestone, etc.), technology provider, production
capacity, size of the kiln, fineness of the final product and byproducts, time of
the analysis, and other factors. Moreover, it should be noted that some energy
efficiency measures also provide additional productivity and environmental
benefits which are difficult and sometimes impossible to quantify. However,
including quantified estimates of other benefits could significantly reduce the
CCE for the energy-efficiency measures (Worrell et al., 2003; Lung et al., 2005).
Based on the methodology explained above and the information from Table 2,
the FCSC and ECSC were constructed separately to estimate the cost-effective
and total technical potential for electricity and fuel efficiency improvement in the
Chinese cement industry from 2010 to 2030. In addition, the CO2 emission
reduction potential from implementing efficiency measures was also calculated.
Out of 23 energy-efficiency measures that were included in our questionnaire, 22
measures were applicable to the cement industry in China, 17 of which are
electricity-saving measures that are included in ECSC and 5 of which are fuelsaving measures used to derive the FCSC.
However, it should be noted that there are a few technologies such as energy
management and process control systems in clinker making, optimize heat
recovery/upgrade clinker cooler, and blended cement production that either save
both electricity and fuels, or increase electricity consumption as a result of
saving fuel. These technologies with fuel savings accounting for a significant
portion of their total primary energy savings are included in the FCSC.
Five energy-efficiency measures were used to construct the FCSC. Figure 6 shows
that all five energy-efficiency measures fall below the discounted average unit
price of fuel (coal) in the cement industry from 2010 to 2030 (1.4US$/GJ),

indicating that the CCF is less than the discounted average unit price of fuel for
these measures. In other words, the cost of investing in these five energyefficiency measures to save one GJ of energy in the period of 2010 - 2030 is less
than purchasing one GJ of fuel at the given price.
The two most cost-effective measures are installation of high efficiency motors
and High efficiency fan for raw mill vent fan with inverter. The largest electricity
saving potential is from replacing a ball mill with vertical roller mill in finish
grinding (ranked 7 on the curve) and low temperature waste heat recovery
power generation, which is saving in purchased electricity by generating
electricity from the waste heat onsite
A similar study that investigated barriers to the implementation of cost-effective,
energyefficiency technologies and measures in Thailand (Hasanbeigi, 2009)
found the following key barriers: Management concerns about the high
investment costs of energy efficiency measures: Even though the payback period
of efficiency measures might be short, some cement plants still have difficulty
acquiring the high initial investment needed to purchase energy efficiency
measures.
Management considers production more important: In many industrial
production plants, upper management is focused solely on production output,
final product quality and sales, with little or no attention to energy efficiency. This
is also the case for some cement plants, although energy costs high share of
cement production cost makes it less of a barrier when compared to less energyintensive industries .
Management concerns about time required to improve energy efficiency: The
high cost of disrupting industrial production may raise concerns about the time
requirements for implementing energy efficiency measures.
Lack of coordination between external organizations: The implementation of
energy and environmental regulations lacks proper execution and enforcement
as a result of the lack of coordination between different ministries and
government institutions responsible for energy and environmental issues.
Current installations are already considered efficient: This is especially true for
newly-installed cement production lines, although they may not be as efficient as
the best commercially available technologies.
Given the importance of the cement industry in China as one of the highest
energy-consuming and CO2-emitting industry, this study aims to understand the
potential for energy-efficiency improvement and CO2 emission reductions using
a bottom-up model. Specifically, bottom-up Energy Conservation Supply Curves
(i.e. ECSC and FCSC) were constructed for the Chinese cement industry to
estimate the savings potential and costs of energy-efficiency improvements by
taking into account the costs and energy savings of different technologies.
We analyzed 23 energy efficiency technologies and measures for the cement
industry. Using a bottom-up CSC models, the cumulative cost-effective and

technical electricity and fuel savings as well as the CO2 emissions reduction
potentials for the Chinese cement industry for 2010-2030 are estimated. By
comparison, the total technical primary2 energy saving achieved by the
implementation of the studied efficiency measures in the Chinese cement
industry over 20 years (2010-2030) is equal to around 32% of total primary
energy supply of Latin America or the Middle East or around 74% of primary
energy supply of Brazil in 2007 (IEA 2009).
Doheim et al. studied out the savings associated with the waste heat and losses
from different equipment. Around 2% fuel saving can be achieved by utilizing
insulation of mineral wool around preheater unit. Thermal flow diagram is used
for fuel saving analysis. They carried out the study of heat losses from various
equipment that can be utilized as a technical savings.

Madlool et al. studied out the cost saving by utilizing the waste heat streams
and payback period of waste heat recovery boilers. For the temperature of 220
C of hot air and 325 C of exhaust gases which leave the suspension preheater,
net power of cogeneration plant is 834.12 KJ/kg which accounts the thermal
efficiency of 26.6%.The payback period of their system is 20 months roughly.
optimized the operational parameters such as masses of cooling air and clinker,
cooling air temperature, and grate speed to improve the energy, exergy and
recovery efficiencies of a grate cooling system. Using heat recovery from the
exhaust air, energy and exergy recovery efficiencies of the cooling system were
increased by 21.5% and 9.4%, respectively. About 38.10% and 30.86% energy
cost can be saved by changing mass flow rate of clinker and mass flow rate of
cooling air, respectively. Besides these studies, the exergy analysis for complete
system in the cement production was demonstrated by Koroneos et al. .
Whereas, the exergy analysis for cement industries was reviewed by Madlool et
al. . At the range of temperature of 200 to 300oC, almost 40% of total input of
heat is emitted from the exit gases of pre-heater and clinker cooler. The waste
heat is used in different applications, such as drying of raw materials, air
preheating which is required for the coal combustion cogeneration . Al-Rabghi et
al. reviewed the utilization of waste heat recovery in diverse industry sectors.
Saneipoor et al. studies the performance of a new Marnoch Heat Engine (MHE)
in a typical cement plant. Sogut et al. examined rotary kiln heat recovery for a
cement plant in Turkey. Jiangfeng Wang et al. used four kinds of power plant to
recover the waste heat from the exit gases of pre-heater and grate cooler in
order to generate the power in a cement plant. This work determines the
electrical energy saving which is led to reduction in energy consumption, and
thus, reduction in cost saving. As well as, estimation of the simple payback
period was accomplished.
The energy consumption is ranged from 4 to 5 GJ/tonne of cement was indicated
by studies. A share of energy consumption of cement industry in the industrial
field is between 12% to 15%. And it represents 2% to 6% of total energy
consumption in terms of countries. Continually, to improve energy use

profitability and competitiveness, many effectual technologies in energy use by


industry were adopted. A significant number of studies have been concentrated
on the analysis of energy and its utilization in cement industry. Among them,
there are very important and constructive papers. Madlool et al. reviewed the
energy use in cement industry. Doheim et al. examined the thermal energy
consumption, losses and the heat saving potentials. For dry process in cement
industry, Engin et al. and Kabir et al. applied an analysis of energy audit.
Khurana et al. presented the analysis of thermodynamic and cogeneration
WASTE HEAT RECOVERY In addition to the plan of reducing of energy
consumption in cement production process, the recovery waste heats can be
achieved in order to produce the electrical energy by utilization cogeneration
power plant. This means no additional fuel consumption and thus, reducing the
high cost of electrical energy and the emissions of greenhouse gases. The waste
heats can be classified as waste heats of middle and low temperatures. Some
power plants are available and suitable to recover the waste heats [6]. The waste
heat sources in the cement plant include the exit gases from the pre-heater and
the clinker cooler ejection hot air. And for cogeneration power, these sources
which have diverse level of temperature can be used separately or together. The
temperature of ejection hot air from the cooler is 220oC and the temperature of
gases which leave the suspension pre-heater is 325oC. As shown in Fig1, the
steam which is generated via WHRSG by utilizing these two sources would be
used to drive a steam turbine. A steam turbine will drive the electric generator to
produce the electricity. This will reflect in reduction of electricity demand.
The
net power of cogeneration power plant was 834.12 kJ/kg. If we consider the
operation hours in 1 year is 8000hrs. Then the electric saving will be 6.673103
MWh/year.
The results can be written in some points as following: 1) The net power output
was 834.12 kJ/kg and the generated electricity was 6.673103 MWh/yr in
cogeneration power plant.
2) The cost saving was 0.233 USD/tonne for the cogeneration power system.
3) Payback period for this system will be roughly 20 months.

Madlool et al. studied out the energy efficient technologies for energy saving and
to reduce the greenhouse gases. Their study shown that increasing the stages of
suspension preheater reduce the exhaust gases temperature and thermal energy
consumption for preheating the raw materials. The use of suspension preheater
reduce the pressure drop and power requirement in fans. They studied out that
suspension preheaters are more efficient in recovering the heat. The energy from
waste gas can be converted into power which shows thermal efficiency upto
0.21-0.22 GJ/t and reduction of greenhouse gas by 3.68-9.25 kgCO2/t.
Mehul Shah et al. propounded India should increase its cement production
significantly. However, cement industry is considered as one of the large energy

consuming industry. Cement industry has been consuming high amounts of


energy for many years. For producing clinker normally rotary kilns are widely
used in cement plants. Significant amount of coal is burnt to carry out the kiln
processing. Each tone of cement produced releases approximately 1 tone of
CO2.The major part of the CO2 emissions from the production of cement is
released from the calcination of limestone (50%) and from the combustion of
fuels (40%). About 35% to 40% of the process heat is lost from those waste heat
streams. 1) Cement Production Process: A typical process of manufacture
consists of three stages:- Preparation of raw-mix by grinding the mixture of
limestone and clay. Clinker formation by heating the raw-mix to sintering
temperature (up to 1450 C) in a cement kiln. Grinding the resulting clinker to
make cement. In the second stage, the raw-mix is fed into the kiln and gradually
heated by contact with the hot gases from combustion of the kiln fuel.
Successive chemical reactions take place as the temperature of the raw-mix
rises.
They studied out the possible heat recovery potential from a cement plant and
design of cogeneration plant. They carried out energy and exergy approaches
with possible improvements. Their case study is based on 7100 TPD capacity
cement plant at Sagar cement. The designed thermal power plant is using simple
Rankine cycle with deaerator. The steam turbine inlet condition is maintained at
15 bar and 390 C. Their results conclude that increase in steam pressure
decreases the thermal efficiency of the cogeneration plant. The increase in
steam pressure increases the hot gas exit temperature at the evaporator at a
fixed pinch point which drops the heat recovery in superheater and evaporator.
Pinch point temperature is kept 30 K.
They carried out the comparative study of water steam Rankine cycle and
organic Rankine cycle (ORC) used in cement industries based on pinch point
analysis. They concluded that for high temperature of the exhaust gases watersteam Rankine cycle is efficient solution but in modern cement plant
temperature of exhaust gas is low due to high efficiency of the plant. In such a
case use of organic Rankine cycle (ORC) is advantageous. In their study they had
taken isopentane as a working fluid for ORC cycle. Increasing the number of
stages of suspension preheater reduces the gas temperature. Four stage
preheater provides the exhaust gases at 300-380 C, while 5-6 stage preheater
provides exhaust gases at 200-300 C. But it is limited upto six due to structural
limitations. Their research conclude that the maximum temperature of
superheated steam is set up 10 C lower than the inlet temperature of hot air in
Air quench cooling (AQC) boiler.
This paper presents the case study of Van cement industry, Turkey. The plant
having the capacity of 600 TPD running on dry process. They suggested three
possible ways to recover the waste heat from existing cement plant. Those are
as. Use of waste heat recovery steam generator (WHRSG) Use of waste heat
to preheat the raw material Heat recovery from kiln surface by providing
secondary shell They carried out detail audit of cement plant along with the
possibility to recover waste heat. The plant having overall efficiency of 48.7%.

The major heat loss sources have been determined as kiln exhaust (19.15% of
total input), cooler exhaust (5.61% of total input) and combined radiative and
convective heat transfer from kiln surfaces (15.11% of total input). They
suggested the use of WHRSG for the first two losses and use of secondary shell
around the kiln to recover the heat loss from kiln surface. The total saving for the
whole system has been estimated, which indicates an energy recovery of 15.6%
of the total input energy. The specific energy consumption of the plant is 3.68
GJ/t of clinker produced.
They carried out the comparative study of power generation through waste heat
recovery by different cogeneration cycles. Their project is done on Askari Cement
Limited (ACL), situated near Hassan-Abdal, Punjab province in Pakistan. Plant
having the capacity of 3700 TPD. They studied out the waste heat recovery
power plant (WHRPP) based on different technologies. Their study is based on
comparison of single flash based WHRPP, Dual pressure based WHRPP, ORC
(organic rankine cycle) based WHRPP and Kalina cycle based WHRPP. The
auxiliary power required for different cycles based on net power output is
calculated and efficiency of the system based on cycle adopted is also calculated
They have stated that maximum power is produced by Kalina based WHRPP as
compare to steam based WHRPP with maximum process efficiency but has
comparatively high auxiliary equipment and usage consumption. Steam based
WHRPP produce large amount of power by using average amount of auxiliary
equipment and the system is environment friendly as it works on water as fluid
and also requirement and availability of trained man power is not an issue and
this technology only works efficiently on high waste gas temperature. Another
result shows that the more power is produced, the more CO2 is reduced and the
more carbon credits are produce. The Kalina based WHRPP produces maximum
number of carbon credits
They studied out the energy efficient technologies for energy saving and to
reduce the greenhouse gases. Their study shown that increasing the stages of
suspension preheater reduce the exhaust gases temperature and thermal energy
consumption for preheating the raw materials. The use of suspension preheater
reduce the pressure drop and power requirement in fans. They studied out that
suspension preheaters are more efficient in recovering the heat. The energy from
waste gas can be converted into power which shows thermal efficiency upto
0.21-0.22 GJ/t and reduction of greenhouse gas by 3.68-9.25 kgCO2/t.
A heat recovery unit needs to be install to recover the heat potential from this
streams so that power can be produced. Design of such equipment to recover
this heat and power savings through that is must be carried out to improve the
efficiency of cement plants. There are many ways to recover the waste heat from
the cement plant. Mainly three ways to recover the waste heat. They are as..
Use of waste heat recovery system Use of waste heat to preheat the raw
material Heat recovery from kiln surface The most efficient way is to use of
waste heat recovery boiler to recover the waste heat. Preheater boiler and Air

quench cooling boiler can be used to recover the waste heat from exhaust gases
and hot air respectively.
H. Dipak Sen, Rajsekhar Panua, Pulak Sen et al. They have done the case study
on BVCL (Barak Valley Cement Ltd.), Assam running on dry process with four
stage suspension preheater having the capacity of 500 TPD. They have taken a
minimum pinch point of 20 C. They have carried out the first law efficiency of
kiln system and that is 52.04%.They proposed the heat recovery power plant and
estimated its capacity that is around 987 KW. The efficiency of proposed plant is
38.42%.

State-of-the-art new suspension process (NSP) kilns include multi-stage


preheaters and pre-calciners to preprocess raw materials before they enter the
kiln, and an air-quench system to cool the clinker product. Kiln exhaust streams,
from the clinker cooler and the kiln preheater system, contain useful thermal
energy that can be converted into power. Typically, the clinker coolers release
large amounts of heated air at 250 to 340 C (480 to 645 F) directly into the
atmosphere. At the kiln charging side, the 300 to 400 C (570 to 750 F) kiln gas
coming off the preheaters 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. If the raw mill is
down, the exhaust gas would be cooled with a water spray or cold air before it
entered the dust collectors. Maximizing overall kiln process efficiency is
paramount for efficient plant operation, but remaining waste heat from the
preheater exhausts and clinker coolers can be recovered and used to provide low
temperature heating needs in the plant, or used to generate power to offset a
portion of power purchased from the grid, or captive power generated by fuel
consumption at the site. Typically, cement plants do not have significant lowtemperature heating requirements, so most waste heat recovery projects have
been for power generation. The amount of waste heat available for recovery
depends on kiln system design and production, the moisture content of the raw
materials, and the amount of heat required for drying in the raw mill system,
solid fuel system and cement mill. Waste heat recovery can provide up to 30
percent of a cement plants overall electricity needs and offers the following
advantages (LBNL 2008, EPA 2010):

Reduces purchased power consumption (or reduces reliance on captive


power plants), which in turn reduces operating costs

Mitigates the impact of future electric price increases

Enhances plant power reliability

Improves plant competitive position in the market

Lowers plant specific energy consumption, reducing greenhouse gas


emissions (based on credit for reduced central station power generation or
reduced fossil-fired captive power generation at the cement plant)

Waste Heat Recovery Power Systems- Waste heat recovery power systems used
for cement kilns operate on the Rankine Cycle. This thermodynamic cycle is the
basis for conventional thermal power generating stations and consists of a heat
source (boiler) that converts a liquid working fluid to high-pressure vapor (steam,
in a power station) that is then expanded through a turbogenerator producing
power. Low-pressure vapor exhausted from the turbogenerator is condensed
back to a liquid state, with condensate from the condenser returned to the boiler
feedwater pump to continue the cycle. Waste heat recovery systems consist of
heat exchangers or heat recovery steam generators (HRSGs) that transfer heat
from the exhaust gases to the working fluid inside, turbines, electric generators,
condensers, and a working fluid cooling system. Three primary waste heat
recovery power generation systems are available, differentiated by the type of
working fluid (Gibbon 2013, EPA 2012, CII 2009), as follows:
Steam rankine Cycle (SrC) The most commonly used Rankine cycle system for
waste heat recovery power generation uses water as the working fluid and
involves generating steam in a waste heat boiler, which then drives a steam
turbine. Steam turbines are one of the oldest and most versatile power
generation technologies in use. As shown in Figure 4, in the steam waste heat
recovery steam cycle, the working fluidwateris first pumped to elevated
pressure before entering a waste heat recovery boiler. The water is vaporized
into high-pressure steam by the hot exhaust from the process and then
expanded to lower temperature and pressure in a turbine, generating mechanical
power that drives an electric generator. The low-pressure steam is then
exhausted to a condenser at vacuum conditions, where the expanded vapor is
condensed to low-pressure liquid and returned to the feedwater pump and boiler.
Organic rankine Cycles (OrC) Other types of working fluids with better
generation efficiencies at lower heat source temperatures are used in organic
Rankine cycle (ORC) systems. The ORCs typically use a high molecular mass
organic working fluid such as butane or pentane that has a lower boiling point,
higher vapor pressure, higher molecular mass, and higher mass flow compared
to water. Together, these features enable higher turbine efficiencies than those
offered by a steam system. The ORC systems can be utilized for waste heat
sources as low as 150 C (300 F), whereas steam systems are limited to heat
sources greater than 260 C (500 F). The ORC systems are typically designed
with two heat transfer stages. The first stage transfers heat from the waste
gases to an intermediate heat transfer fluid (e.g., thermal transfer oil). The
second stage transfers heat from the intermediate heat transfer fluid to the
organic working fluid. The ORCs have commonly been used to generate power in
geothermal power plants, and more recently, in pipeline compressor heat
recovery applications in the United States. The ORC systems have been widely
used to generate power from biomass systems in Europe. A few ORC systems
have been installed on cement kilns.10 The ORCs specific features include the
following (Turboden 2012, Holcim 2011, Ormat 2012, Gibbon 2013):

Can recover heat from gases at lower temperatures than is possible with
conventional steam systems, enabling ORCs to utilize all recoverable heat from

the air cooler


Operate with condensing systems above atmospheric
pressure, reducing risk of air leakage into the system and eliminating the need
for a de-aerator Not susceptible to freezing
Because ORCs operate at
relatively low pressure, they can operate unattended and fully automated in
many locations depending on local regulations
The organic fluid properties
result in the working fluid remaining dry (no partial condensation) throughout the
turbine, avoiding blade erosion
Can utilize air-cooled condensers without
negatively impacting performance Lower-speed (rpm) ORC turbine allows
generator direct drive without the need for and inefficiency of a reduction gear
ORC equipment (turbines, piping, condensers, heat exchanger surface) is
typically smaller than that required for steam systems, and the turbine generally
consists of fewer stages

Although ORCs can provide generation efficiencies comparable to a steam


Rankine system, ORCs are typically applied to lower temperature exhaust
streams, and limited in sizing and scalability, and generally are smaller in
capacity that steam systems. Depending on the application, ORC systems
often have a higher specific cost (US$/kW) than steam systems The two-stage
heat transfer process creates some system inefficiencies The heat transfer
fluids and organic fluids normally used in ORCs are combustible, requiring fire
protection measures and periodic replacement over time. Also, there may be
environmental concerns over potential system leaks.
In general, ORC
systems are well-matched with small- to medium-size, high-efficiency kilns or
kilns with elevated raw material moisture content
The kalina Cycle is another Rankine cycle that uses a binary mixture of water and
ammonia as the working fluid, which allows for a more efficient energy extraction
from the heat source. The Kalina cycle takes advantage of the ability of
ammonia-water mixtures to utilize variable and lower temperature heat sources.
The Kalina cycle has an operating temperature range that can accept waste heat
at temperatures of 95 C (200 F) to 535 C (1,000 F) and is claimed to be 15 to
25 percent more efficient than ORCs at the same temperature level. The Kalina
cycle is in market introduction, with a total of nine operating systems in diverse
industries such as steel and refining, and in geothermal power plants where the
hot fluid is very often a liquid below 150 C (300 F).11 Kalina cycle systems are
now being piloted in the cement industry.12 Key features of the Kalina cycle
include the following (Gibbon 2013, Mirolli 2012).
The amount of recoverable waste heat from an NSP kiln depends on several
factors including the following:

Moisture content of the raw material feed (i.e., determines heat


requirement for the kiln and the amount of preheater exhaust needed for drying)

Amount of excess air in the kiln

Amount of air infiltration

Number and efficiency of preheater/precalciner stages

Configuration of the clinker cooler system

he number of preheater stages in a cement plant has significant bearing on the


overall thermal energy consumption and waste heat recovery potential. The
higher the number of stages, the higher the overall 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.
Table 5 summarizes the quantity of waste heat recoverable from state-of-the-art
NSP kilns. Preheater exhaust temperatures range from 390 C (735 F) for small
kilns with four preheater stages, to below 300 C (570 F) for large kilns with six
preheater stages.
The clinker cooler design also impacts waste heat availability. The basic cooler
function is to remove heat from hot clinker discharged from the kiln so the clinker
can be handled by subsequent equipment. Rapid cooling also improves clinker
quality and grindability. Typically, state-of-the-art coolers are grate coolers, which
have various stages of development. Table 6 summarizes the heat available in
different generations of grate coolers. Exhaust air temperatures from the clinker
cooler range from 250 to 330 C (480 to 625 F) depending on cooler
configuration and recuperation efficiency.
Typically, the potential electrical power generation, depending on waste heat
losses and the number of preheater cyclone stages, ranges from 25-45 kWh/t of
clinker. Assuming an average plant electrical drive power requirement of 106
kWh/t of cement and a clinker factor of 0.75, approximately 20 to 30 percent of
the required electricity for the cement production process can be generated from
the waste heat. Figure 8 shows the band of expected power generation for a
range of kiln capacities.
An additional limiting factor in the heat available for effective recovery is the
moisture content of the raw material entering the kiln. Limestone deposits
moisture content can range from 2 to 15 percent depending on the limestone
origin. The amount of moisture present in the feed material entering the kiln
preheater influences specific heat consumption in the kiln and the kiln
production rate. Typical practice is to limit moisture content entering the kiln to
less than 1.0 percent (CII 2009). To achieve this level, raw feed material is
normally dried during grinding by utilizing preheater hot gas and/or cooler hot
gas as the heat source.
Theoretically about 2.26 GJ is required to evaporate or remove one tonne of
moisture from raw feed or limestone (540 kcal/kg water). However, in practice,
vertical roller mills require 3.77 to 4.61 GJ of heat per tonne of moisture removed
(900 to 1100 kcal/kg water), and ball mills require about 3.14 to 3.56 GJ of heat
per tonne of moisture due to losses in mill outlet gas, radiation losses, and air
infiltration (750 to 850 kcal/kg water). To illustrate the impact of moisture on
drying requirements, Table 8 gives the raw material flows and heat required in
Mkcal / hr for various kiln production rates at different moisture levels based on
the following assumptions:

Raw meal to clinker factor of 1.55

Heat requirement of 3.98 GJ / tonne of water for raw mill (950 kcal/kg)

Raw mill running hours per day 22

Project Economics of Waste Heat Recovery Power Generation The project


economics of waste heat power generation depend on several site-specific and
project-specific factors, including the following considerations: The amount of
heat available in (exhaust gas volume and temperature) and conditions of the
waste gases determine the size, potentially the technology (e.g., ORCs are more
applicable for lower-temperature exhaust streams and lower gas volumes), and
overall generation efficiency (e.g., amount of power that can be produced) of the
WHR system. The amount of heat available and at what temperature is a
function of the size and configuration of the kiln (i.e., tpd and number of
preheater/precalciner stages) and the raw material moisture level (determines
the percentage of hot exhaust gases need for drying). Capital cost of the
heat recovery system is generally a function of size, technology and, as
discussed below, supplier.
System installation costs (design, engineering,
construction, commissioning and training) are functions of the installation size,
technology, complexity, supplier and degree of local content. System
operating and maintenance costs are a function of size, technology, site-specific
operational constraints or requirements; costs are influenced by staffingwill the
system be handled by existing operating staff, new staff that require training, or
outsourcing?
Operating hours of the kiln and availability of the heat
recovery system Displaced power prices based on grid electricity no longer
purchased, or reduced dependence on captive power plants and associated costs

Net power output of the WHR system. Net output is more important in
determining project economics than gross power output. The impact of auxiliary
power consumption and process/booster fans must be included in efficiency and
economic calculations.
SK Gupta et al. states that, Size of WHRPP is influenced by the moisture content
present in raw material and fuel (coal). Even after considering heat for drying-off
nominal moisture content, around 30 kwh/ t clinker of power can still be
generated by WHRPP, i.e. say, 7.5 MW for a 6,000 tpd plant.
The cost of installation of WHRPP is around Rs. 10 crores per MW, and the
operating cost is less than Rs. 0.5/ unit (excluding interest & depreciation).
For a 6000 tpd plant, around 70,000 t/annum of CO2 can be reduced. Installation
of WHRPP also reduces carbon dioxide generation.
Based on the experience of the cement/ captive power plants, this paper deals
with the various issues involved in decision-making, as well as in implementation
of WHRPP in cement industry
The various avenues for the exploitation of this available waste heat energy are :
Drying of raw material(s) and coal , Generation of electricity Heating of

equipment/storage hoppers to facilitate easy handling of sticky material(s) ,


Heating of building(s) in cold countries , Heating of water , Generation of steam
for oil handling installation and driving some auxiliaries , Air conditioning/deep
freezing by adsorption process , Scavenging air for bag filters
Principally, direct utilization of the waste heat for drying of raw material and coal,
yields the maximum efficiency, and should be given the first preference.
However, even after meeting the nominal drying requirement, substantial heat is
available in waste gases. Electricity generation from surplus waste heat is the
next preferred option, out of the above-mentioned avenues.
As mentioned above, normally, the first priority for utilization of the waste heat
is, for the drying of raw materials and coal, since the efficiency of energy
conversion system, from thermal to electricity, like in WHRPP, is as low as 18 to
20% (conversion losses are very high). Second priority is given to the power
generation.
Though, incorporation WHRPP, increases complexity of the system, to some
extent, it is still worth considering, as power is generated at much lower cost,
than that available from Grid.
As a broad criteria, at places where un-interrupted power supply is available at
cheaper rates, it may not be economical to install WHRPP. Places where
electricity rates are high, or there are frequent power interruptions, use of
WHRPP is very lucrative.
3.2 Moisture in raw materials has an important influence on the sizing of the
power generation plant. This fact is depicted in Fig -1. It shows that the waste
heat recovery power plant is sized based on the different moisture conditions in
raw material / coal i.e. the outlet gas temp from waste heat recovery system is
decided, so as to cater to the moisture drying requirement.
The moisture content in the raw material increases during rainy season.
Quantum of moisture, and the duration, varies from place to place. The sizing of
WHRPP plant can be done for any of the following conditions. It can, therefore,
run on full load only, during the extremely dry season. Whenever moisture is
more than min. value, power generation would need to be curtailed, so that
waste heat is first utilized for drying.
Considerations for Installation of WHRPP
01. In the PH gases, dust load is high, and is sticky. In cooler, is very abrasive.
Proper care hence, has to be taken for selection of the various equipment
especially, dust dislodging system, dampers, velocity profile in boiler, wear
liners, configuration of the boiler, turbine and condenser etc.
02. While introducing WHRPP, the additional pressure drop is there in the system.
PH and cooler fans hence, have to be sized accordingly.

03. In actual practice, the kiln performance may vary from time to time, resulting
in fluctuating exhaust gas conditions. As such, the waste heat boiler and the
connected turbine generator set, should be suitably designed/selected, to take
care-of such fluctuations.
Cement Industry being a power intensive industry, one has to think in terms of
energy optimisation and savings. Waste Heat Recovery Power Plants, contribute
significantly, to the electrical energy saving (to the tune of 25%).
The reduction in CO2 emission, makes it environmental friendly.
Installation of the waste heat recovery plant has to be tackled as a system
approach, rather than considering cement plant and WHRPP operations,
independently.
The Waste Heat Recovery Technology, as any other technology, is in an
incessant phase, and many more innovations, in terms of equipment and
applications, may be expected in future.

You might also like