phd thesis, yuanjing zheng, endelig version, klar til print

288
TECHNICAL UNIVERSITY OF DENMARK DEPARTMENT OF CHEMICAL AND BIOCHEMICAL ENGINEERING Ph.D. Thesis, June 2011 Mercury Removal from Cement Plants by Sorbent Injection upstream of a Pulse Jet Fabric Filter Yuanjing Zheng

Upload: mjbotelho

Post on 02-Jan-2016

51 views

Category:

Documents


0 download

TRANSCRIPT

TECHNICAL UNIVERSITY OF DENMARK

DEPARTMENT OF CHEMICAL AND BIOCHEMICAL ENGINEERING

Ph.D. Thesis, June 2011

Mercury Removal from Cement Plants by Sorbent Injection

upstream of a Pulse Jet Fabric Filter

Yuanjing Zheng

I

Preface

This thesis is written for partial fulfillment of the requirements to obtain the

Ph.D. degree at the Technical University of Denmark. The work has been carried out

at the CHEC (Combustion and Harmful Emission Control) Research Centre at the

Department of Chemical and Biochemical Engineering under the supervision of Prof.

Anker Degn Jensen from CHEC, Department Manager Christian Windelin and

Flemming Jensen from FLSmidth A/S. The project is financially supported by the

Industrial PhD programme of the Danish Ministry of Science, Technology and

Innovation, Danish Advanced Technology Foundation as part of the Research

Platform on New Cement Production Technology.

I would like to thank my supervisors, particularly Anker Degn Jensen, for

their support, fruitful discussion and comments. Technician Thomas Wolfe and

department workshop are gratefully acknowledged for help building the fixed-bed

reactor system. I am very grateful to Mr. Peter Paone from FLSmidth A/S for reading

part of the manuscript. Student Jacob Clement Nielsen is acknowledged for

performing some of the screening tests. Thanks to all the other people at CHEC and

FLSmidth, not mentioned here, for their great help received during my study.

Finally, I would like to thank my family and friends for their support and

encouragement.

Yuanjing Zheng

Kgs. Lyngby, June 2011

II

Abstract

There are growing concerns over mercury emissions due to their toxicity,

volatility, persistence, and bioaccumulation in the environment. Mercury emissions

from cement plants are being regulated by environmental agencies in most countries.

Among the available technologies for mercury removal from flue gas, sorbent

injection upstream of a polishing fabric filter is considered as the most promising and

suitable technology for cement plant application. Cement plants are quite different

from power plants and waste incinerators regarding the flue gas composition,

temperature, gas and solid residence time, and inherent material circulation. Thus

knowledge obtained from mercury removal in power plants and incinerators might

not be applied to cement plants directly and fundamental investigation under well

controlled cement kiln condition is imperative.

Tests in simulated cement kiln flue gas show that the red brass converter

developed for waste incinerator application does not work properly for either

elemental or total mercury measurement. Sodium sulfite converter is developed and

optimized for oxidized mercury reduction and total mercury measurement. The

response time of the sulfite converter is short, which makes it appropriate for

dynamic measurement of mercury adsorption and oxidation by sorbents.

Screening tests of sorbents for mercury removal from cement plants have

been conducted in the fixed-bed reactor system using simulated cement kiln flue gas

with elemental mercury and mercury chloride sources. The tested sorbents include

commercial activated carbons, commercial non-carbon sorbents, and cement

materials. With elemental mercury present in the flue gas, no mercury adsorption or

oxidation by non-carbon based sorbents and cement materials is observed. Generally

larger amount of adsorbed mercury is obtained with sorbents that have larger mercury

oxidation capacity. While all the non-carbon based sorbents and cement materials

show some adsorption of mercury chloride. Among the tested sorbents the Darco Hg

III

activated shows the best performance of adsorption of both elemental and oxidized

mercury and is recommended as the reference sorbent for fundamental investigation.

Parametric studies of mercury adsorption by activated carbon have been

conducted in the fixed-bed reactor regarding the effects of adsorption temperature,

flue gas rate, mercury level, carbon particle size, carbon load, and flue gas

composition. The mercury adsorption isotherm follows Henry’s law for the applied

mercury inlet levels in this project. Henry’s constant and heat of adsorption are

derived for model input. The mercury adsorption capacity does not change with O2,

CO, and NO levels in the flue gas, but decreases when CO2, H2O, SO2, and NO2

concentrations increase. Slight promoting effects of HCl on mercury adsorption are

observed with HCl in the flue gas up to 20 ppmv. Larger mercury adsorption capacity

is obtained when HCl is removed from the gas. Similar adsorption behaviors of

mercury chloride and elemental mercury by Darco Hg activated carbon are observed

using simulated cement kiln flue gas, due to the effective catalytic oxidation of

elemental mercury by the activated carbon.

Mathematical models are developed to simulate mercury adsorption by a

single carbon particle, fixed carbon bed, in the duct and fabric filter. The developed

fixed bed model can reasonably simulate the mercury breakthrough curve of the fixed

carbon bed. Comparison with fabric filter model simulations and experimental data

from slipstream tests at a cement plant shows that the developed two-stage model is a

valuable tool and can reasonably predict the mercury removal from cement plants by

carbon injection upstream of a fabric filter.

IV

Resumé (summary in Danish)

Der er voksende bekymringer over kviksølvemissioner grundet disses

giftighed, flygtighed, bestandighed og biologisk akkumulation i miljøet.

Kviksølvemissioner fra cementfabrikker reguleres i de fleste lande af miljøorganer.

Blandt de tilgængelige teknologier til fjernelse af kviksølv fra røggas anses

sorbentinjektion opstrøms for et posefilter for den mest lovende og velegnede

teknologi til anvendelse på cementfabrikker. Cementfabrikker er temmelig forskellige

fra kraftværker og affaldsforbrændingsanlæg med hensyn til røggassammensætningen,

temperatur, opholdstid af gas og faststof samt iboende materialecirkulation. Derfor

kan viden opnået fra kviksølvfjernelse i kraftværker og affaldsforbrændingsanlæg

ikke anvendes direkte på cementfabrikker og fundamental undersøgelse under

velkontrollerede forhold svarende til cementfremstilling brændingsovn er essentielt.

Test i simuleret røggas fra cementbrændingsovn viser, at en kommerciel

konverter udviklet til anvendelse på affaldsforbrændingsanlægs ikke virker godt for

hverken elementær kviksølvmåling eller total kviksølvmåling. Som en del af

projektet er der udviklet en natriumsulfitkonverter til reduktion af oxyderet kviksølv

samt total kviksølvmåling. Sulfit konverterens responstid er kort hvilket gør den

velegnet til dynamisk måling af kviksølv adsorption og oxidation med sorbenter.

Screeningsforsøg af sorbenter til fjernelse af kviksølv fra cementfabrikker er

udført i et fixed bed reaktorsystem ved brug af simuleret røggas fra cementsovne med

både elementært kviksølv samt kviksølvklorid. De testede sorbenter inkluderer

kommercielle aktivt kul- og kommercielle ikke-kulstofsorbenter samt

cementmaterialer. Med elementært kviksølv tilstede i røggassen blev hverken

kviksølvadsorption eller -oxidation observeret med de ikke kulstofbaserede sorbenter

og cementmaterialer. Generelt opnås større adsorberet mængde kviksølv med

sorbenter der har større kviksølvoxidationskapacitet. Alle de ikke-kulstofbaserede

sorbenter og cementmaterialer viser nogen adsorption af kviksølvklorid. Blandt de

testede sorbenter udviser Darco Hg aktivt kul den bedste evne til adsorption af både

V

elementært og oxideret kviksølv og anbefales som referencesorbent i den

fundamentale undersøgelse.

Parameterstudier af kviksølvadsorption med aktivt kul er blevet udført i en

fixed bed reaktor med hensyn til effekter af adsorptionstemperatur, røggasmængde,

kviksølvniveau, kulstofpartikelstørrelse, kulstofbelastning og røggassammensætning.

I dette projekt følger kviksølvadsorptionsisotermen Henrys lov for den anvendte

koncentration af kviksølv. Henrys konstant og adsorptionsvarmen er fundet til

indsættelse i model. Kviksølvadsorptionskapaciteten ændres ikke som følge af O2,

CO og NO niveauer i røggassen, men falder når CO2, H2O, SO2, og NO2

koncentrationerne stiger. En mindre positiv effekt af HCl på kviksølvadsorption er

observeret med HCl i røggassen op til 20 ppmv. Større kviksølv adsorptionskapacitet

opnås når HCl fjernes fra gassen. Lignende adsorptionsmønster for kviksølvklorid og

elementært kviksølv med Darco Hg aktivt kul er observeret ved brug af simuleret

røggas fra cementsovne, på grund af den effektive katalytiske oxidation af elementært

kviksølv med det aktive kul.

Matematiske modeller er udviklet til at simulere kviksølvadsorption på en

enkel kulpartikel, i en fixed bed af aktivt kul, i kanalen og i posefilteret. Den

udviklede fixed bed model med god nøjagtighed simulere kviksølv

gennembrydningskurven for fixed bed forsøgen. Sammenligning af posefiltermodel

simuleringer med eksperimentelle data fra slipstrømstests på en cementfabrik viser at

den udviklede to-trins model er et værdifuldt værktøj der på fornuftigvis kan

forudsige kviksølvfjernelsen fra cementfabrikker med kulstofinjektion opstrøms for et

posefilter.

VI

Table of contents

Preface ........................................................................................................................... I

Abstract.........................................................................................................................II

Resumé (summary in Danish)..................................................................................... IV

Table of contents......................................................................................................... VI

1. Introduction............................................................................................................... 1

1.1 Project background ............................................................................................. 1

1.2 Project objectives................................................................................................ 3

1.3 Outline of the thesis ............................................................................................ 3

1.4 References........................................................................................................... 4

2. Mercury emissions and transformations in cement plants........................................ 6

2.1 Cement production processes ............................................................................. 6

2.2 Mercury contents in fuels and cement raw materials ....................................... 12

2.3 Mercury emissions............................................................................................ 14

2.3 Mercury transformation during combustion ..................................................... 15

2.3.1 Mercury transformation in coal combustion flue gas ................................ 17

2.3.2 Mercury transformation within cement kiln system.................................. 23

2.4 Conclusions....................................................................................................... 27

2.5 Further work ..................................................................................................... 28

2.6 References......................................................................................................... 28

3. Review of technologies for mercury removal from flue gas .................................. 32

3.1 Introduction....................................................................................................... 32

3.2 Mercury avoidance technology......................................................................... 33

3.2.1 Coal cleaning ............................................................................................. 33

3.2.2 Cement raw material cleaning ................................................................... 33

3.2.3 Fuel switching............................................................................................ 34

3.3 Mercury removal by powdered activated carbon injection .............................. 35

3.3.1 Parameters affecting mercury removal by activated carbon injection....... 35

3.3.2 Tests of mercury sorbents in lab-scale fixed-bed reactors......................... 38

3.3.3 Sorbent injection in power plants .............................................................. 49

3.3.5 Carbon surface chemistry and mechanisms of mercury capture on carbons

............................................................................................................................ 58

3.3.6 Processing and reuse of mercury laden activated carbon .......................... 63

3.3.7 Applicability of sorbent injection in cement plants ................................... 65

VII

3.4 Mercury removal by activated carbon bed ....................................................... 65

3.5 Mercury control by flue gas desulphurization systems .................................... 67

3.6 Mercury removal by sodium tetrasulfide injection........................................... 68

3.7 Enhanced mercury removal by oxidation ......................................................... 69

3.8 Mercury removal by roaster process................................................................. 72

3.9 Conclusions....................................................................................................... 73

3.10 Further research requirement .......................................................................... 75

3.11 Abbreviations.................................................................................................. 75

3.12 References....................................................................................................... 76

4. Experimental methods and materials ...................................................................... 86

4.1 Description of the fixed-bed reactor system..................................................... 86

4.1.1 Gas mixing system..................................................................................... 88

4.1.2 Mercury vapor addition system ................................................................. 88

4.1.3 Humidifier for water vapor addition.......................................................... 90

4.1.4 Low temperature furnace and fixed-bed reactor........................................ 92

4.1.5 Mercury analysis system............................................................................ 93

4.2 Converter and sorbent materials ..................................................................... 100

4.3 Flue gas composition ...................................................................................... 103

4.4 Sorbent load in fixed-bed test ......................................................................... 103

4.5 Experimental procedure.................................................................................. 105

4.6 Sorbent characterization ................................................................................. 106

4.6.1 Scanning electron microscopy ................................................................. 106

4.6.2 Particle size distribution........................................................................... 107

4.6.3 Analysis of mercury in sorbent................................................................ 108

4.7 References........................................................................................................... 108

Appendix............................................................................................................... 110

4A Check of mercury analyzer ............................................................................. 110

4B Water addition verification ............................................................................. 112

5. Dynamic measurement of mercury adsorption and oxidation on activated carbon in

simulated cement kiln flue gas.................................................................................. 117

5.1 Review of gaseous mercury measurement technology................................... 117

5.2 Performance test of the mercury analyzer ..................................................... 119

5.3 Performance test of the red brass converter.................................................... 121

5.4 Performance of the sulfite converter............................................................... 125

5.5 Examples of dynamic measurement of mercury adsorption and oxidation on

activated carbon .................................................................................................... 131

5.6 Suggestions for practical application of the converter.................................... 132

5.7 Conclusions..................................................................................................... 133

VIII

5.8 References....................................................................................................... 134

6. Effects of bed dilution and carbon load on mercury adsorption capacity of activated

carbon........................................................................................................................ 137

6.1 Introduction..................................................................................................... 137

6.2 Effects of carbon load ..................................................................................... 137

6.3 Effects of bed dilution..................................................................................... 141

6.4 Effects of sand load......................................................................................... 143

6.5 Effects of carbon loading location .................................................................. 144

6.6 Effects of bed materials .................................................................................. 145

6.7 Effects of carbon type and particle size .......................................................... 146

6.8 Tests with only Portland cement..................................................................... 147

6.9 Conclusions..................................................................................................... 148

6.10 References..................................................................................................... 149

7. Screening tests of mercury sorbents ..................................................................... 151

7.1 Introduction..................................................................................................... 151

7.2 Sorbent properties and compositions.............................................................. 153

7.3 SEM-EDX analysis of fresh sorbents ............................................................. 157

7.4 Baseline test .................................................................................................... 160

7.5 Screening tests in nitrogen.............................................................................. 160

7.6 Screening tests in simulated cement kiln flue gas with elemental mercury

source .................................................................................................................... 162

7.7 Screening tests in simulated cement kiln flue gas with HgCl2 source............ 166

7.8 Conclusions..................................................................................................... 170

7.9 References....................................................................................................... 172

8. Fundamental investigation of elemental mercury adsorption by activated carbon in

simulated cement kiln flue gas.................................................................................. 176

8.1 Introduction..................................................................................................... 176

8.2 Effect of adsorption temperature .................................................................... 177

8.3 Isotherm tests .................................................................................................. 179

8.4 Effect of carbon particle size .......................................................................... 183

8.5 Effect of flue gas flow rate ............................................................................. 185

8.6 Effects of flue gas compositions..................................................................... 186

8.6.1 Effect of CO2 ........................................................................................... 186

8.6.2 Effect of O2 .............................................................................................. 188

8.6.3 Effect of H2O ........................................................................................... 189

8.6.4 Effect of CO............................................................................................. 192

8.6.5 Effect of SO2............................................................................................ 193

8.6.6 Effect of HCl............................................................................................ 195

IX

8.6.7 Effect of NO............................................................................................. 197

8.6.8 Effect of NO2 ........................................................................................... 198

8.7 Conclusions..................................................................................................... 201

8.8 References....................................................................................................... 202

9. Fundamental investigation of mercury chloride adsorption by activated carbon in

simulated cement kiln flue gas.................................................................................. 206

9.1 Introduction..................................................................................................... 206

9.2 Effect of temperature ...................................................................................... 207

9.3 Effect of flue gas composition ........................................................................ 210

9.4 Conclusions..................................................................................................... 212

9.5 References....................................................................................................... 213

10. Simulation of mercury adsorption by fixed carbon bed ..................................... 215

10.1 Adsorption equilibrium................................................................................. 215

10.2 Transport consideration in adsorption process ............................................. 216

10.2.1 External transport................................................................................... 216

10.2.2 Internal transport.................................................................................... 218

10.3 Modeling of adsorption in a single particle .................................................. 220

10.4 Fixed bed adsorption model.......................................................................... 226

10.5 Conclusions................................................................................................... 239

10.6 List of symbols.............................................................................................. 239

10.7 References..................................................................................................... 241

11. Simulation of mercury removal by activated carbon injection upstream of a fabric

filter........................................................................................................................... 243

11.1 Common assumptions for mercury removal in the duct and fabric filter ..... 243

11.2 Duct model.................................................................................................... 246

11.3 Model for the filter cake ............................................................................... 253

11.4 Fabric filter model ........................................................................................ 257

11.5 Two-stage model........................................................................................... 263

11.6 Conclusions................................................................................................... 269

11.7 List of symbols.............................................................................................. 270

11.8 References..................................................................................................... 271

12. Concluding remarks............................................................................................ 273

13. Suggestions for further work .............................................................................. 277

1

1

Introduction

1.1 Project background 

There are growing concerns over mercury emissions due to its toxicity, volatility,

persistence, and bioaccumulation in the environment. According to an inventory of

global mercury emissions to the atmosphere from anthropogenic sources by Pacyna et al.

[1], the largest emissions of mercury are from combustion of fossil fuels. Mercury

emissions from cement and mineral production are the second largest anthropogenic

sources.

While mercury emissions from waste incinerators and power plants have been

and continue to be regulated by the authorities in many countries, strict mercury emission

limits for cement plants are also established by different countries [2-6]. U.S.

Environmental Protection Agency (EPA) recently set the nation’s first limits on mercury

emissions from existing cement kilns and strengthened the limits for new kilns [7-9]. The

mercury emission limit for existing and new cement plants is 55 and 21 pound/million

tons of clinker, respectively. These emission limits correspond to 10 and 4 µg/Nm3.

When fully implemented in 2013, EPA estimates the annual mercury emissions will be

reduced about 92% [8]. It is estimated that few cement kilns in U.S. can achieve this new

mercury emission limit without some changes to the system, either through operational

adjustment or use of add-on technology.

Mercury is present in both cement raw materials used for kiln feed and fuels used

in the cement production process. Due to rising energy costs and ever stricter energy and

environmental regulations, alternative fuel technology is becoming an important factor in

controlling costs. To gain a competitive edge, many cement and mineral producers

worldwide have set ambitious targets for increasing their future usage of alternative fuels

2

- both waste-derived fuel and biomass. High mercury containing alternative fuels such as

chemical waste, domestic waste and sewage sludge are also incinerated in cement plants

and high mercury emission problems have been encountered. To ensure that the mercury

emission limit is met, FLSmidth has initiated research on mercury removal from cement

plants.

Due to the extremely low concentration range of mercury in the flue gas, mercury

emission control techniques are technically challenging and expensive. Currently,

activated carbon injection upstream of a particulate control device such as fabric filter

has been shown to have the best potential to remove both elemental and oxidized

mercury from the flue gas for combustion facilities not equipped with a wet flue gas

desulphurization plant [10]. This also applies to cement plants where typically no wet

flue gas desulphurization unit is installed. In cement plant application sorbent will be

injected upstream of a polishing filter instead of an existing filter in order to separate

carbon from the cement materials and save the disposal cost of sorbent and cement

materials mixture.

Although activated carbon is the most studied sorbent for capturing mercury from

power plant flue gas, mercury adsorption by activated carbon is not clearly understood

yet, and research and development efforts are still needed before carbon injection may be

considered as a commercial technology for wide use [2]. New sorbents need to be

developed, the sorbent costs need to be reduced and the amount of carbon injected needs

to be kept to a certain level to minimize the cost. Furthermore, mercury adsorption

stability by sorbents needs to be proved.

Extensive research has been carried out to reduce mercury emissions from coal

combustion and waste incineration, but very little efforts have been concentrated on

mercury removal in cement plants. The mercury removal not only depends on the sorbent

but also on the speciation of mercury, flue gas composition and temperature, and the

system configuration. The mercury emissions and gas stream characteristics from coal

combustion and waste incineration are quite different from those from cement kilns [4].

Thus knowledge obtained from mercury removal in power plants and incinerators might

3

not be applied to cement plant directly. Non-carbon based cement-friendly sorbent is

desired so that the mercury containing sorbent can be used in cement production instead

of costly disposal.

Despite the considerable experimental research that has been carried out to date,

few models for mercury adsorption by activated carbon injection in power plant or

incinerator flue gas have been proposed. A comprehensive model is desired to estimate

appropriate design and operating strategies that would lead to efficient and economic

control of mercury.

1.2 Project objectives 

The overall goal of this project is to develop and advance improved mercury control

technologies using sorbent injection upstream of a pulse jet fabric filter for cement plant.

Specific objectives are as follows:

1. To obtain updated knowledge of mercury control technologies relevant to cement

plant by comprehensive literature review.

2. To develop an experimental lab setup and screen sorbents for capturing mercury

from cement kiln flue gas.

3. To test and develop thermal catalytic converters for oxidized mercury reduction and

total mercury measurement.

4. To develop an understanding of sorbent chemistry and provide mechanistic

understanding and kinetic rates for sorbents of interest.

5. To develop mathematic models that can describe mercury removal in fixed-bed and

predict mercury removal efficiency in cement plant by injecting sorbent upstream of a

fabric filter.

1.3 Outline of the thesis 

The thesis starts with a chapter (Chapter 2) on introduction of cement production

process and mercury emission and transformation in cement kiln systems. Then in

Chapter 3 available knowledge on mercury removal technologies from flue gas is

4

reviewed and the applicability of the reviewed technologies in cement kilns is analyzed.

Properties and performance of typical sorbents are also presented.

Experimental methods and materials are presented in Chapter 4. Chapter 5

particularly deals with the test of a red-brass based converter and development of a

sulfite-based oxidized mercury reduction unit for total gaseous mercury measurement.

Effects of bed dilution and carbon load on equilibrium mercury adsorption capacity of

the activated carbon are investigated in chapter 6. Screening tests of different sorbent

materials in the fixed-bed reactor under simulated cement kiln flue gas are reported in

Chapter 7. Chapter 8 deals with fundamental investigation of mercury adsorption by

activated carbon in simulated cement kiln flue gas using elemental mercury source.

Mercury adsorption mechanism and kinetics by the activated carbon will be reported.

The fundamental investigation of mercury chloride adsorption by the activated carbon in

simulated cement kiln flue gas will be reported in Chapter 9.

Chapters 10 and 11 will deal with simulations of mercury adsorption by the

activated carbon. Chapter 10 focuses on simulation of mercury adsorption by a single

carbon particle and a fixed carbon bed. Simulation of mercury adsorption by activated

carbon injection upstream of a fabric filter is the topic of Chapter 11. Validation of the

developed duct-fabric filter two-stage model by available pilot-scale data is reported.

Finally, conclusions from the project are presented in Chapter 12. Suggestions for

further work are given in Chapter 13.

1.4 References 

[1] E.G. Pacyna, J.M. Pacyna, F. Steenhuisen, S. Wilson, Global anthropogenic mercury

emission inventory for 2000, Atmospheric Environment. 40 (2006) 4048-4063.

[2] The European Parliament and the Council of the European Union, Union directive

2000/76/EC on the incineration of waste, 2000.

[3] J. Werther, Gaseous emissions from waste combustion, Journal of Hazardous Materials. 144

(2007) 604-613.

[4] G. Ebertsch and S. Plickert, German contribution to the review of the reference document on

best available techniques in the cement and lime manufacturing industries, Part I: Lime

manufacturing industries, 2006.

5

[5] German Cement Works Association, Environmental protection in cement manufacture, VDZ

activity report 2003-2005.

[6] Canadian Council of Ministers of the Environment, Canada-wide standards for mercury

emissions, 2000.

[7] U.S. EPA, EPA sets first national limits to reduce mercury and other toxic emissions from

cement plants, http://yosemite.epa.gov/opa/admpress.nsf, accessed September 6, 2010.

[8] U.S. EPA, Fact sheet, Final amendments to national air toxics emission standards and new

source performance standards for Portland cement manufacturing, 2010.

[9] U.S. EPA, National emission standards for hazardous air pollutants from the Portland cement

manufacturing industry and standards of performance for Portland cement plant, 40 CFR Parts 60

and 63, EPA-HQ-OAR-2007-0877, FRLRIN 2060-AO42; EPA-HQ-OAR-2002-0051, FRLRIN

2060-AO15, http://www.epa.gov /ttn/oarpg/t1/fr_notices/portland _cement_fr_080910.pdf,

accessed January/17, 2011.

[10] J.H. Pavlish, E.A. Sondreal, M.D. Mann, E.S. Olson, K.C. Galbreath, D.L. Laudal, S.A.

Benson, Status review of mercury control options for coal-fired power plants, Fuel Processing

Technology. 82 (2003) 89-165.

6

2

Mercury emissions and transformations in cement

plants

Knowledge of mercury emissions, speciation, and transformation in cement

plants is important for understanding the transport and fate of mercury released to air

pollution control systems. In this chapter cement production processes are first

introduced and compared with power plants and waste incinerators regarding the flue gas

composition, temperature, residence time, and inherent material circulation. Then

mercury contents in fuels and raw materials applied in cement production and mercury

emission from Portland cement plants are presented. Finally mercury transformations in

combustion flue gas and cement kiln system are reviewed.

2.1 Cement production processes 

Although cement production also involves combustion, the flue gas temperature

and residence time in cement kilns are quite different from power plants and waste

incinerators. To help understand the mercury chemistry in the cement kiln systems, a

brief description of the cement production process is necessary. Differences regarding the

gas temperature, residence time, flue gas composition, and material cycles among cement

kilns, power plants and waste incinerators are discussed below.

Depending on how the raw material is handled before being fed to the rotary kiln,

the processes can be categorized as dry, semi-dry, semi-wet and wet processes [1].

Presently, about 78% of Europe's cement production is from dry process kilns [1], about

16% of production is by semi-dry/semi-wet process kilns, and approximate 6% of cement

production is from wet process kilns. Today, all new plants are based on the dry process

and many old wet plants are either replaced or converted to the dry or semi-dry process.

7

In the dry process the feed material enters the kiln in a dry, powdered form.

Production of cement can be subdivided into the areas of supply of raw materials,

burning of cement clinker in the rotary kiln, and final cement production by adding

interground additives [2].

Raw materials for the manufacture of Portland cement clinker consist basically of

limestone and aluminosilicates. At times, certain corrective materials such as bauxite,

iron ore, and sand are used to compensate the specific chemical shortfalls in the raw mix

composition. Apart from natural raw materials, waste materials containing lime,

aluminate, silicate, and iron are also used as raw materials substitutes.

Figure 2.1 illustrates a typical dry cement production process [2].The mixture of

raw materials is milled in a raw mill and dried by the hot kiln flue gas. In a downstream

electrostatic precipitator (ESP) or fabric filter (FF), the raw meal is separated and

subsequently transported to raw meal silos. The raw meal is fed into the kiln system,

which is comprised of a tower of cyclone preheaters. The calcination process can almost

be completed before the raw material enters the kiln if part of the fuel is added in a

precalciner, which is located between the kiln and the preheater.

8

Figure 2.1. Sketch of a dry cement production process [2].

In the burning of cement clinker it is necessary to maintain material temperatures

of up to 1450°C to ensure the sintering reactions required [1]. This is achieved by

applying peak combustion temperatures of about 2000°C with the main burner flame.

Figure 2.2 shows the gas temperature in the kiln system as a function of residence time

and comparison with the gas temperature profiles in a pulverized coal-fired boiler and

waste incinerator [1,3,4]. The combustion gases from the main kiln burner remain at

temperatures above 1200°C for at least 5-10 seconds. An excess of oxygen, typically 2-4

vol.%, is also required in the combustion gases of the rotary kiln as the clinker needs to

be burned under oxidizing conditions. The residence time of the solid materials in the

rotary kiln is 20-30 min and up to 60 min depending on the length of the kiln. The hot

flue gas flows through the rotary kiln and preheater in opposite direction to the solids.

The burning conditions in kilns with precalciner firing depend on the precalciner design.

9

Gas temperatures from a precalciner burner are typically around 1100°C, and the gas

residence time in the precalciner is approximately 3 seconds. In the cyclone preheater

zone, the gas temperatures typically range from approximately 880-890°C at the inlet of

the bottom preheater cyclone to 350°C at the exit of the top preheater and can have a

residence time of 10 to 25 s. The post-preheater zone consists of the cooler, the mill dryer

and the air pollution control device, with gas temperature typically in the range from

approximately 350-90°C from the top of the preheater to the exit stack outlet.

 

Fig. 2.2. Gas temperature and retention time profiles in a cyclone preheater/precalciner

kiln system, pulverized coal-fired boiler, and waste incinerator. Data are from [1,3,4].

Generally the gas temperature and residence time in a kiln system is much higher

and longer than those in a pulverized coal-fired boiler and waste incinerator. The

temperature profile in the waste incinerator shown in Figure 2.2 is in the region from the

furnace exit to the boiler exit [4] and the gas temperature is much lower than those from

cement kiln and pulverized coal-fired boiler.

The clinker leaving the rotary kiln is cooled down by grate or planetary coolers.

After cooling, the clinker is ground with a small amount of gypsum to produce Portland

cement, which is the most common type of cement. In addition, blended cements are

10

produced by intergrinding cement clinker with materials like fly ash, granulated blast

furnace slag, limestone, natural or artificial pozzolanas [5].

Table 2.1 compares the flue gas compositions among coal-fired power plant,

waste incinerator and cement kiln. The major difference between cement kiln flue gas

and other flue gases is the larger water and CO2 content in the kiln flue gas. The oxygen

content in the kiln gas is lower than in coal combustion and waste incineration flue gas.

The emission of HCl from cement kilns is normally much lower than those from waste

incinerators. This could be due to the fact that the environment in cement plants is

effective for absorbing acid gasses [6], such as a range of gas temperatures from 100 to

1650°C, gas residence time of about 30s, high levels of turbulence, high concentrations

of alkaline solids including sodium and potassium oxides, and freshly created CaO in

high concentrations. Therefore, gaseous species such as HCl or HF are nearly completely

captured by the inherent and efficient alkaline sorption effect of the cement kiln system

[1].

Table 2.1 Typical flue gas compositions in coal-fired boiler, waste incinerator, and

cement kiln before air pollution control device (APCD).

Pulverized coal-fired boiler [1,7-10,10,11]

Waste incinerator [7,8]

Cement kiln [7-9,12]

O2 (vol.%) 4-6 6-15 2-4 CO2 (vol.%) 10-16 5-14 14-33 H2O (vol.%) 5-12 10-18 5-35 CO (ppmv) 10-100 10-100 600-2600 NO (ppmv) 100-1000 100-1000 475-1900 NO2 (ppmv) 5-50 5-50 25-100 N2O (ppmv) <1-5 <1 <1 SO2 (ppmv) 100-2000 100-300 10-2500 SO3 (ppmv) 10-40 0-30 HCl (ppmv) 1-100 400-1000 1-25 Flue gas temperature at APCD inlet (°C)

135-180 180-230 85-230

11

Cement kilns also differ from conventional boilers and incinerators in having the

dust recycles in the kiln systems. There are two material cycles in the cement kiln system,

i.e., the internal and external cycle. Because of the countercurrent flow of combustion

products and solids in cement kilns, volatile elements such as mercury, alkalis, sulphur

and chlorine evaporated from the solids at the hot end of the kiln near the combustion

zone are carried to the cold end by the combustion gases. Some of the volatile

compounds pass through the entire system and exit in vapor phase through the stack.

However, as the flue gas cools, some volatile compounds may adsorb/condense onto dust

particles and surrounding walls in the cooler regions of the kiln system. With the raw

meal, they are reintroduced to the hot zone thus establishing the internal cycle of volatile

elements.

The external cycle comprises the mass flows that include the raw mill and dust

collectors downstream of the preheater. A small part of the circulating elements leaves

the kiln with the exhaust gas dust and is precipitated in the dedusting device of the

system. The collected cement kiln dust (CKD) often is blended into the raw meal for

reintroduction, or part of it is fed directly to the cement mill to lower the alkali content of

the clinker and meet product specifications. The CKD typically accounts for about 7% of

the solid flow in cement plant with a precalciner [13].

With excessive input of volatile elements, the installation of a kiln gas bypass

system may become necessary in order to extract part of the circulating elements from

the kiln system. This bypass dust, which is usually highly enriched in alkalis, sulphur or

chloride, is cooled down and then passed through a dust collector before being

discharged.

The operation modes of the cement plants are important for understanding

mercury transformations in the kiln systems as presented in section 2.3. There are two

operation modes [2], i.e., compound operation (raw-mill-on) and direct operation (raw-

mill-off), as shown in figure 2.3. Usually these modes are run alternately. The raw mill

operates typically 80-90% of the time the kiln operates [14]. During compound operation

12

the dust-containing off-gas from the cyclone preheater is used for drying and transporting

the raw meal from the raw mill. Water injection in the cooler is not applied to cool down

the gas. The raw meal and fly dust from the kiln system are collected by the ESP or FF

and passed on to the raw meal silo. During direct operation, the raw mill is not used. The

dust-containing off-gas from the kiln is cooled down in the off-gas cooler by the injection

of water and subjected to subsequent dedusting in the ESP or FF.

Figure 2.3. Operation models in cement production [2].

These different modes of operation considerably influence the temperatures and

material flows between the mill, kiln system, and dust filter. These changes also affect

the trace element mass flows in the plant. Increased off-gas temperature during direct

operation causes higher mercury emission level than in the compound mode [2].

Moreover, regular alternation of the operation modes results in weekly cycles of mercury

flows in the cement plant, as discussed in section 2.2.

2.2 Mercury contents in fuels and cement raw materials 

A comprehensive analysis of mercury content in 291 raw material samples from

57 cement plants in Canada and U.S. was conducted by Hills and Stevenson [15]. Table

2.2 shows the mercury contents in the fuels and raw materials applied in cement

production. There is a wide range of mercury level in both fuels and cement raw

materials. The reported average mercury content in the raw materials except for fly ash

and recycled cement kiln dust is less than 80 ppb. In terms of fuel sources, the majority

of studies reported that the average and maximum levels of mercury in coal, tire-derived

13

fuel, and petroleum coke are under 0.2 and 1 ppm, respectively. Fly ash has a high

mercury content and application of fly ash in cement production results in increased

mercury input to the cement kiln and potentially higher mercury emissions. Process

changes in cement plants such as substitution with alternative fuels may result in more

plants needing solutions for mercury emission control.

Table 2.2. Mercury contents in raw materials and fuels for cement production. All on dry

weight basis.

Material/fuel Category Sample number

Average (ppm)

Minimum (ppm)

Maximum (ppm)

Limestone [15] 90 0.017 <0.001 0.391 Sand [15] 34 0.029 <0.001 0.556 Clay [15] 28 0.052 0.001 0.270 Shale [15]

Primary raw materials

17 0.057 0.002 0.436 Slag [15] 10 0.012 0.002 0.054 Bottom ash [15] 12 0.048 0.003 0.382 Iron ore [15] 12 0.078 0.002 0.672 Fly ash [15]

Secondary raw materials

16 0.205 0.002 0.685 Recycled cement kiln dust (CKD) [15]

19 1.530 0.005 24.56

Petroleum Coke [16] 290 0.050 0.010 0.200 Sub-bituminous coal [16] 2137 0.070 0.010 0.900 Lignite coal [16] 320 0.110 0.020 0.430 Bituminous [16] 6198 0.120 0.000 1.120 Anthracite coal [16]

Regular fuels

13 0.160 0.120 0.210 Tire-derived fuel [16] 30 0.097 0.050 0.400 Tire samples from German cement plants [17]

- 0.170 0.100 0.430

Sewage sludge [18-20]

Secondary fuels

- 1.880 0.600 56.00

In bituminous coals, mercury is generally associated with pyrite (FeS2) and

cinnabar (HgS), while in sub-bituminous coals mercury is largely associated with the

organic fraction [21]. There is no correlation between the mercury content and the pyrite

content in the limestone, which suggests that the mercury in the limestone is not

14

associated primarily with the sulphide phase [21]. In cement production, most of the

mercury is from the kiln feed rather than the fuels when considering the amount of fuels

and raw materials used [22].

2.3 Mercury emissions 

The U.S. Portland cement association summarized 50 mercury emission tests in

the U.S. during 1989-1996 [23]. All the mercury emission data for long dry, preheater,

and precalciner kilns were essentially obtained with the raw-mill-on operating mode. The

emission data are only for plants not burning hazardous waste. The information on

mercury speciation is not available. The mercury emission concentrations varied from

0.02 μg/Nm3 to 385.6 μg/Nm3 with a mean value of 28.0 μg/Nm3 @dry, 7% O2 and a

standard deviation of 62.7 μg/Nm3. The maximum mercury concentration was three

times higher than the second highest value.

The U.S. Portland cement association has later gathered and analyzed mercury

emissions and process data from 645 stack tests in 42 cement plants up to 2007 [24]. The

mercury emissions include particle-bound mercury (Hgp), elemental mercury (Hg0), and

oxidized mercury (Hg2+). The mercury emissions and speciation from cement kilns can

vary over time and depend on raw materials and fuels used, and process operation. The

average mercury speciation percentages for cement plants with preheater or precalciner

not firing waste are 5% Hgp, 56% Hg2+, 39% Hg0 for raw-mill-on mode [24], and 4%

Hgp, 62% Hg2+, 34% Hg0 during raw-mill-off mode.

Large variations of mercury speciation during raw-mill-on and -off modes have

been observed in some plants with higher mercury emission during the raw-mill-off

period [25]. Measurements at Ash Grove’s Durkee plant showed that the average

mercury concentration during raw-mill-on and raw-mill-off period was 410 and 2250

μg/Nm3, respectively [25]. The larger mercury emission during raw-mill-off period is

probably due to high flue gas temperature and lack of mercury adsorption by cement raw

materials. Due to the high mercury emission, the Ash Grove’s Durkee plant has

15

volunteered to install a sorbent injection process for removing at least 75% of the

mercury [26].

The complex mercury mitigation cycles within the cement kiln system make it

difficult to obtain an equilibrium state due to the periodical shut down of raw mills for

maintenance. It typically takes weeks to reach long term equilibrium of the mercury

emission [27].

The German cement manufacturing association has reported mercury emission

results from 216 measurements on 44 kilns [28]. Twenty of the results were below the

detection limit. Most of the measurements were below 40 μg/Nm3. Only six of the results

were 60 μg/Nm3 or higher.

The emitted elemental mercury from Powder River Basin (PRB) coal-fired power

plants ranges from approximately 10 to100 μg/Nm3 [29]. Mercury concentrations in the

flue gas from municipal solid waste combustion (200 to 1000 μg/Nm3) are one to two

orders of magnitude higher than for coal combustion sources (5 to 20 μg/Nm3) [30,31].

Mercury levels in cement kiln flue gas are generally closer to those found in coal-fired

boilers and lower than those found in waste incinerators.

Pacyna et al. [32] presented an inventory of global mercury emissions to the

atmosphere from anthropogenic sources for the year 2000. The largest emissions of

mercury to the global atmosphere are from combustion of fossil fuels, mainly coal in

utility, industrial, and residential boilers. Emissions of mercury from coal combustion are

between one and two orders of magnitude higher than emissions from oil combustion.

Various industrial processes account for additional 30% of mercury emissions from

anthropogenic sources worldwide in 2000. Mercury emissions from cement and mineral

production are the second largest anthropogenic sources.

2.3 Mercury transformation during combustion 

Knowledge of mercury transformations in combustion flue gas is important for

selection of the mercury control technology and understanding the fate and behavior of

mercury from combustion processes. Major chemical forms of mercury from combustion

16

sources are oxidized mercury and elemental mercury [33,34]. Another form is particulate

mercury, which is the portion of mercury deposited on fine particles. Oxidized mercury

species, such as HgCl2 and HgO, are easily removed by existing wet type air pollution

control devices like flue gas desulphurization (FGD), due to its water-soluble property.

Also particulate mercury is readily removed by the main dust removal control devices

such as ESPs and FFs. On the other hand, elemental mercury is difficult to control

because of its high vapor pressure and insolubility in water.

Table 2.3 presents properties of selected mercury compounds. Metallic mercury is

a heavy, silvery-white liquid metal at typical ambient temperatures and pressures, and it

vaporizes under those conditions. Mercurous (Hg+1) and mercuric (Hg+2) mercury form

numerous inorganic and organic chemical compounds, but the mercurous mercury is

rarely stable under ordinary environmental conditions [23]. The solubility of the mercury

compounds varies greatly from negligible (Hg2Cl2, HgS) to very soluble (HgCl2).

Mercuric sulfate reacts with water to produce yellow insoluble basic mercuric subsulfate

and sulfuric acid.

Table 2.3. Properties of selected mercury compounds [23,35,36]. n.a.: not available

Name Molar

weight

(g/mol)

Melting

point

(C)

Boiling

point

(C)

Decomposition

/sublimate

temperature

(C)

Density

(g/cm3)

Aqueous

solubility

(g/l at 25C)

Hg0 Elemental

mercury

200.59 -38.8 356.7 n.a. 13.53 5.610-7

Hg2Cl2 Mercurous

chloride

472.09 525 n.a. 383 7.15 0.002

HgCl2 Mercuric

chloride

271.50 277 302 n.a. 5.43 28.6

Hg2SO4 Mercurous

sulphate

497.24 n.a. n.a. n.a. 7.56 0.51

HgSO4 Mercuric

sulphate

296.66 n.a. n.a. 450 6.47 decomposes

17

Name Molar

weight

(g/mol)

Melting

point

(C)

Boiling

point

(C)

Decomposition

/sublimate

temperature

(C)

Density

(g/cm3)

Aqueous

solubility

(g/l at 25C)

HgS Mercury

sulfide

232.66 n.a. 446-

583

580 8.10 insoluble

HgO Mercuric

oxide

216.59 n.a. 356 500 11.14 insoluble

Hg2Br2 Mercurous

bromide

560.99 405 n.a. 340-350 7.31 3.910-4

HgBr2 Mercuric

bromide

360.44 237 322 n.a. 6.03 slightly

soluble

Hg2I2 Mercurous

iodide

654.98 n.a. n.a. 140 7.70 Slightly

soluble

HgI2 Mercuric

iodide

454.40 259 350 n.a. 6.36 0.06

Hg2F2 Mercurous

fluoride

439.18 n.a. n.a. 570 8.73 decomposes

HgF2 Mercuric

fluoride

238.59 645 650 645 8.95 soluble,

reacts

Hg2(NO3)2 Mercurous

nitrate

525.19 n.a. n.a. 70 (dihydrate) 4.80

(dihydrate)

slightly

soluble, reacts

Hg(NO3)2 Mercuric

nitrate

324.7 79 n.a. n.a. 4.3 0 soluble

 

2.3.1 Mercury transformation in coal combustion flue gas 

Figure 2.4 illustrates the potential mercury transformation paths during coal

combustion [33]. All forms of mercury in the coal decompose in the combustion flame to

form Hg0(g) [30,33]. In the post combustion section where the gas temperature decreases,

Hg0(g) may remain as a monatomic species or react to form inorganic mercurous and

mercuric compounds. The principal oxidized forms of mercury in coal combustion flue

gas are assumed to be Hg2+ compounds. Oxidation of mercury via halogenation does not

reach equilibrium under conditions of rapid quenching [4,7]. The degree of oxidation of

mercury via gas-phase reactions therefore depends on the cooling rate of the flue gas.

18

After mercury chlorination, the resulting HgCl2(g) may remain in the flue gas or adsorb

onto inorganic and carbonaceous ash particles entrained in the flue gas. In addition to

HCl(g) and Cl2(g), O2(g) and NO2(g) are potential mercury oxidants in the flue gas

[30,33].

Figure 2.4. Potential mercury transformation during coal combustion and subsequently in

the resulting flue gas, modified after [33].

Many parameters can potentially affect the formation of various mercury species

throughout a combustion system [30], including fuel type and composition, combustion

environment, heat transfer/cooling rate, residence time at lower temperatures during

convective cooling, configuration of APCD, and operating practices.

As a starting point, the distribution of mercury species in coal combustion flue

gas can be calculated using thermodynamic equilibrium calculations. Senior et al. [3]

calculated the equilibrium mercury speciation in the flue gas from Pittsburgh bituminous

coal combustion. Typical results from 227 to 827C are shown in figure 2.5. At

temperatures below 150C condensed HgSO4 is the only preferred specie (not shown in

figure 2.5). Similar observations were also observed by Frandsen et al. [37]. As

illustrated in figure 2.5, below 450C all of the mercury is predicted to exist as HgCl2.

Above about 700C 99% of mercury is predicted to exist as gaseous elemental mercury.

19

The remaining 1% is predicted to be gaseous HgO. Between 450 and 700C the split

between HgCl2 and elemental mercury is determined by the chlorine content of the coal.

 

200 300 400 500 600 700 800 900

Temperature (oC)

0

20

40

60

80

100

%H

gHgCl2(g) Hg(g)

HgO(g)

Figure 2.5. Equilibrium distribution of mercury species in flue gas from combustion of

Pittsburgh bituminous coal. Modified after [3]. Coal composition: 4.98 wt% H, 1.48 wt%

N, 1.64 wt% S, 8.19 wt% O, 7.01 wt% ash, 980 ppmm Cl, 0.11 ppmm Hg. Gas

composition at a stoichiometric ratio of 1.2: 14.44 vol.% CO2, 5.69 vol.% H2O, 3.86

vol.% O2, 76.59 vol.% N2, 1166 ppmv SO2, 62 ppmv HCl, 1.24 ppbv Hg, 15.5 ppmv SO3.

The effect of HCl concentration on equilibrium partitioning between elemental

mercury and HgCl2 is illustrated in figure 2.6 [38]. The crossover temperature between

the elemental and oxidized forms increases from 530 to 740C as the HCl concentration

increases from 50 to 3000 ppm. The studied HCl level is much higher than real level in

the flue gas and study using low HCl concentration will be more relevant. The crossover

point is not influenced by the mercury concentration as long as hydrochloric acid is

present in excess. At low temperatures, approximately 10% of the mercury is predicted to

be present as HgO (not shown in figure 2.6). This is probably due to the fact that the

calculations do not use simulated flue gas or include gases such as SO2.

20

 

Figure 2.6. Equilibrium distribution of elemental mercury and mercury chloride for

different HCl concentrations [38]. Other gas concentrations include 7.4% O2, 6.2% CO2,

12.3% H2O and N2 as balance.

The high levels of mercury oxidation are most strongly correlated with high

chlorine concentrations in the coal [33]. Iron is thought to catalyze the oxidation and

subsequent capture of mercury [30]. Calcium likely reacts with chlorine and sulphur

during the combustion process and thereby reduces its ability to promote the oxidation of

mercury [33]. The high percentages of elemental mercury typically found in emissions

from lignite and subbituminous coal combustion can likely be attributed to their high

calcium and low chlorine contents.

Full-scale measurements showed that elemental mercury was dominant in the

stack of coal-fired power plants, while oxidized mercury was dominant in the stack of

incinerators [34,39]. This could be due to the formation of mercury compounds in

furnaces and APCDs configuration differences between them. For the study of mercury

removal by sorbent injection upstream of dust collectors, it is important to know the

mercury speciation at the APCD inlet rather than at the stack. The data of mercury

speciation in the flue gas at the inlets of APCDs are very scattered [25,40-43]. This is

again due to different parameters that potentially affect the mercury speciation. Therefore,

21

to develop a mercury control system for a specific plant, measurement of the mercury

speciation at the APCDs’ inlet is necessary.

There is disagreement in the publications on the relative importance of mercury

halogenation in the flue gas by chlorine and bromine. Most literatures suggest that

chlorine plays the most important role in oxidation of mercury [30,33]. However,

research by Vosteen et al. [44] shows that the critical species for the halogenation of

mercury in the flue gases is not chlorine, but rather bromine. The stable form of the

halogens at high combustion temperatures are HCl and HBr. On cooling of the gases, the

diatomic and molecular form of the halogens become stable according to the Deacon

type of reactions [33,44]:

2 2 24 2 2HCl O H O Cl (Chlorine-Deacon-reaction) (R2.1)

2 2 24 2 2HBr O H O Br (Bromine-Deacon-reaction) (R2.2)

The kinetics of the bromine-Deacon-reaction is more favorable [33,44]. Moreover,

molecular chlorine is consumed during boiler passage by SO2 through the chlorine

Griffin reaction:

2 2 2 3 2SO Cl H O SO HCl (Chlorine-Griffin-reaction) (R2.3)

In contrast to chlorine, the bromine-Griffin-reaction is not thermodynamically

favored at temperatures above 100C, because the Gibbs free reaction enthalpy of the

bromine-Griffin-reaction is strongly positive within the whole boiler temperature range.

Therefore, SO2 is not consuming Br2 during boiler passage. To summarize, the primary

reason that bromine is a much more effective mercury oxidizer than chlorine is that HBr

dissociates much more extensively into reactive atomic species than HCl at typical post-

flame conditions [45].

The world average Cl contents in coals for bituminous and lignite coals are,

respectively, 340±40 and 120±20 ppm [46]. The typical bromine content in the coal is

about 1-10 ppm [33,44,47]. Although the chlorine content in the coal is far higher than

the bromine content in the coal, the amount of molecular bromine Br2 in the flue gas may

be many times higher than the amount of Cl2 in the flue gas downstream the combustion

zone [44]. Recently, Niksa [45,48] also stated that homogeneous chemistry with bromine

22

species is much faster than with chlorine species because the bromine atom

concentrations at the furnace exit are three to four orders of magnitude greater. There

might be ample supply of Br2 to oxidize the typical amounts of mercury in the coal flue

gases through direct mercury bromination:

22 HgBrBrHg (Direct Hg bromination) (R2.4)

Based on this knowledge, direct bromine injection into the flue gas has been proposed

and patented to enhance mercury capture by fly ash or sorbents, or mercury oxidation

followed by removal in wet flue gas desulphurization (FGD) unit [44,49]. However, the

higher concentration of Br2 in the post-combustion zone is not verified by full-scale

investigation due to the lack of Br2 and Cl2 measurements.

The arguments on the relative importance of mercury adsorption by bromine are

supported by simulation and full-scale demonstration in power plants [45,50].

Simulations with only homogeneous reaction mechanism by Niksa et al. [45] show that

50% mercury oxidation is obtained for a typical thermal history along a power plant gas

cleaning system with 10 ppmv Br in the flue gas. In contrast, no mercury oxidation is

achieved by 20 ppmv HCl in the flue gas. Homogeneous mercury oxidation by bromine

begins as the flue gas cools below 600C and accelerates sharply when the temperature

drops to below 300C. At the furnace exit, bromine atoms are present in concentrations

that are comparable to HBr levels, in contrast to the much lower concentrations of

chlorine atoms at these conditions.

Liu et al. [50] estimated that a 50% mercury oxidation could be obtained by

injecting 52 ppm Br2 in the flue gases without fly ash for a reaction time of 15 s at 137°C.

Laboratory study of Br2 in the simulated flue gas showed that fly ash in the flue gas

significantly promoted the oxidation of Hg0 by Br2 and the unburned carbon in the fly

ash played a major role in the promotion primarily through the rapid adsorption of Br2

[50]. Hg0 oxidation in the gas phase was found to be less important than fly ash-induced

oxidation by Br2. However, there is an increasing concern on the stability of bromine

impregnated in the AC, added to fuels, or injected directly to the flue gas, which could

23

lead to downstream pollution and pipeline corrosion due to the strong acidic nature of

bromine.

2.3.2 Mercury transformation within cement kiln system 

Larsen et al. [51] made a thermodynamic calculation of potential mercury species

distribution in a cement kiln preheater. In order to get closer to a preheater environment,

chloride as well as sulphide and sulphate compounds were included in the oxygen-

containing system. Detailed compositions of the solid and flue gas can be found in the

figure caption. The alkaline dust was represented by CaO in the calculations, which is in

excess compared to the acidic components such as HCl and SO2. Figure 2.7 illustrates the

equilibrium distribution of mercury species as a function of temperature when the

mercury input in the solid is in ppmm level. The dominant species below 180°C is

oxidized mercury in forms of HgO and HgCl2, while all mercury compounds

thermodynamically preferred above 200°C are gas-phase species and the main species is

Hg0(g).

24

 

Figure 2.7. Equilibrium distribution of mercury species as a function of temperature in

the preheater environment with mercury input in the range of ppmm [51].

Thermodynamic calculation input: solid: 5.00 kmol CaO, 0.000025 kmol HgCl2,

0.000025 kmol HgSO4, 0.000025 kmol HgS, 0.000025 kmol Hg, gas: 0.03 kmol HCl, 1

kmol H2O,1 kmol O2, 30.00 kmol CO2, 0.05 kmol SO2, 67.95 kmol N2.

Figure 2.8 illustrates the equilibrium distribution of mercury species as a function

of temperature when the mercury input is in ‰ level. Presence of CaO and HCl are not

included in the calculation assuming that HCl can be captured by large amount CaO in

the cement raw materials. The results are completely different from the calculation with

ppmm level of mercury input in the solid. The dominant species below 200°C is HgSO4,

while a certain amount of HgCl2(g) is formed above 200°C. The HgSO4(g) decomposes

at around 450°C, thus the dominant species above 450°C are Hg0(g) and HgCl2(g).

25

 

Fig. 2.8. Equilibrium distribution of mercury species as a function of temperature in the

preheater environment with mercury input in the range of ‰ [51]. Thermodynamic

calculation input: solid: 0.025 kmol HgCl2, 0.025 kmol HgSO4, 0.025 kmol HgS, 0.025

kmol Hg, gas: 1 kmol H2O,1 kmol O2,30.00 kmol CO2, 0.05 kmol SO2, 97.95 kmol N2.

General conclusions from thermodynamic calculations for a preheater

representative environment are [51]: HgS will most probably be converted to other

mercury species when entering the preheater, provided the reaction rates are sufficiently

high compared to residence time. Mercury species are preferentially gas-phase

compounds at temperatures above about 400°C. In a CaO rich environment, the

thermodynamically preferred mercury species above 300°C is Hg0(g). This may be

primarily because CaO acts as an HCl drain. Calculation indicates that the

thermodynamically favored mercury species present at the extraction point for a typical

kiln by-pass is Hg0(g).

Detailed experimental information of mercury transformation in cement kiln

system has not been reported. Although cement production also involves combustion, the

26

flue gas composition, temperature and residence time in cement kiln are quite different

from power plants and waste incinerators as explained earlier. When looking at mercury

chemistry in cement kilns, these factors should be taken into consideration.

Schreiber et al. [22] investigated the fate and inherent control of mercury in

cement kiln systems using material balance studies and comprehensive stack tests that

were conducted over the past two decades. They concluded that mercury does not simply

volatilize out from combusted fuels and heated kiln feed materials and leave directly out

of the stack. The cement kiln systems have some inherent ability to control mercury stack

emissions.

Besides adsorption of mercury on the raw material, as shown earlier, new

mercury compounds such as mercury silicates might be formed through reaction of

mercury with silicate in the raw material and exit the system with the clinker product.

The formation of complex silicates in a kiln system is possible due to the high silica

content in the raw feed (typically 13-15 wt.%) and sufficient residence time for reactions

to take place as vaporized mercury cycles through a kiln system. Edgarbaileyite is the

first reported structure to contain both Hg and Si [52,53]. It has the stoichiometry

Hg6Si2O7 with all of the Hg occurring within the structure as (Hg2)2+ dimers. Although

the mineral data of Edgarbaileyite is available, it has not been possible to identify the

thermodynamic properties of the mineral. A chemical equilibrium study was conducted

to estimate probable conditions for the formation of mercury silicates in high temperature

systems [54]. Results from the study suggest that HgSiO3 may form over a temperature

range of 225 to 325°C. However, the equilibrium calculations also indicate that mercury

silicate formation may be inhibited by the presence of chlorine and sulfur. It is reported

by the European cement association that volatile metals are retained in the clinker to a

very small extent only [1]. Unfortunately, there are no laboratory studies to date that

confirm that mercury silicates are stable above temperatures of 325°C. Fundamental

research is required to identify formation of mercury silicates in the cement kiln systems.

27

2.4 Conclusions 

Cement plants are quite different from power plants and waste incinerators

regarding the flue gas composition, temperature, residence time, and inherent material

circulation. The flue gas temperature and residence time in a kiln system are much higher

and longer than those in a pulverized coal-fired boiler and waste incinerator. There are

larger water and CO2 contents in the cement kiln flue gas.

In cement production the raw materials contain mercury – often at much higher

levels than in the fuels. The flue gas mercury level is highly dependent on the type of fuel

and raw materials. The mercury concentrations in the flue gas from cement kilns are

typically in the range of 1-50 μg/m3. Instead of fuel, cement raw materials are the

dominant sources of mercury in the cement kiln flue gas. Higher mercury emissions,

however, are observed for cement plants firing waste.

The mercury emissions and speciation from cement kilns can vary over time and

depend on raw materials and fuels used, and process operation. The average mercury

speciation percentages for cement plants with preheater or precalciner not firing waste

are 5% Hgp, 56% Hg2+, 39% Hg0 for raw-mill-on mode, and 4% Hgp, 62% Hg2+, 34%

Hg0 during raw-mill-off mode.

Mercury transformations in combustion flue gas have been investigated

intensively to get an understanding of the transport and fate of mercury into to air

pollution control systems. All forms of mercury in the fuel decompose in the combustion

flame to form Hg0(g), which is oxidized to Hg2+ in the post combustion section. Mercury

halogenation by chlorine and bromine is the dominant mercury transformation

mechanism in coal combustion flue gas. The resulting HgCl2(g) may remain in the flue

gas or adsorb onto inorganic and carbonaceous ash particles entrained in the flue gas

stream. Equilibrium calculations and experiments show that bromine is a much more

effective mercury oxidizer than chlorine.

The cement kiln systems have some inherent ability to retain mercury in the solid

materials. The mercury evaporated from the solids at the hot end of the kiln is carried to

the cold end by the combustion gases. As the flue gas cools, some mercury may

28

adsorb/condense onto dust particles in the cooler regions of the kiln system. When the

plant is running in raw-mill-on mode, the kiln gas containing volatilized mercury is used

to sweep the mill of the finely ground raw feed particles and some mercury is adsorbed

by the fine particulates. However, the adsorbed mercury is either carried back to the kiln

hot zone or added to the kiln system together with the raw meal, thus forming mercury

cycles in the kiln system.

2.5 Further work 

There is limited literature regarding mercury characteristics, emissions, and

removal from cement kilns. Essentially all of the published data and information apply to

waste incinerators and coal-fired boilers, all of which have mercury emissions and gas

stream characteristics that are quite different from those from cement kilns. Therefore,

comprehensive studies on mercury chemistry in the cement kiln and mercury removal

from cement plants are imperative.

The inherent recycle of mercury in the kiln system should be further investigated.

The interactions between mercury and cement raw materials play an important role in

understanding of mercury chemistry in the cement kiln system. Research is required to

break the mercury cycle in the kiln system, regenerate and implement beneficial

utilization of removed mercury-contained CKD. These treatment systems minimize net

CKD generation by removing mercury, alkalies and other contaminants and returning

treated dust to the system without compromising product quality.

2.6 References 

[1] CEMBUREAU, the European Cement association, Best available technologies for the cement

industry, 1999.

[2] M. Achternbosch, K.R. Bräutigam, M. Gleis, N. Hartlieb, C. Kupsch, U. Richers, P.

Stemmermann, Heavy metals in cement and concrete resulting from the co-incineration of wastes

in cement kilns with regard to the legitimacy of waste utilisation, Wissenschaftliche Berichte,

FZKA 6923, 2003.

[3] C.L. Senior, A.F. Sarofim, T. Zeng, J.J. Helble, R. Mamani-Paco, Gas-phase transformations

of mercury in coal-fired power plants, Fuel Process Technol. 63 (2000) 197-213.

29

[4] D. Shin, S. Choi, J. Oh, Y Chang, Evaluation of polychlorinated dibenzo-p-

dioxin/dibenzofuran (PCDD/F) emission in municipal solid waste incinerators, Environ. Sci.

Technol. 33 (1999) 2657-2666.

[5] K.H. Karstensen, Formation, release and control of dioxins in cement kilns, Chemosphere. 70

(2008) 543-560.

[6] M. V. Seebach and D. Gossman, Cement kilns sources of chlorides not HCl emissions,

http://www.gcisolutions.com/CK&HCL.htm, accessed June 1, 2008.

[7] C. Senior, A. Sarofim and E. Eddings, Behaviour and measurement of mercury in cement

kilns, presented at the IEE-IAS/PCA 45th Cement Industry Technical Conference, Dallas, Texas,

May 4-9 2003.

[8] B. Hall, P. Schager, O. Lindqvist, Chemical-reactions of mercury in combustion flue-gases,

Water Air and Soil Pollution. 56 (1991) 3-14.

[9] Donaldson Membranes, Reducing emissions: Filtering chloride emissions with a bypass cycle,

Filtr. Sep. 45 (2008) 36-37.

[10] H.F. Johnstone, Reactions of sulfur compounds in boiler furnaces, Industrial and

Engineering Chemistry. 23 (1931) 620-624.

[11] A.A. Presto, E.J. Granite, Impact of sulfur oxides on mercury capture by activated carbon,

Environ. Sci. Technol. 41 (2007) 6579-6584.

[12] E. Worrell, L. Price, N. Martin, C. Hendriks, L.O. Meida, Carbon dioxide emissions from

the global cement industry, Annu. Rev. Energy Environ. 26 (2001) 303-329.

[13] C. Senior, C.J. Montgomery, A. Sarofim, Transient model for behaviour of mercury in

Portland cement kilns, Ind Eng Chem Res. 49 (2010) 1436-1443.

[14] Department of Environmental Quality State of Oregon, Ash Grove mercury reduction,

advisory committee’s report, 2007.

[15] L.M. Hills and R.W. Stevenson, Mercury and lead content in raw materials, PCA R&D

Serial No. 2888, 2006.

[16] L.M. Hills, Mercury and lead content in fuels: A literature review, PCA R&D Serial No.

2887, 2006.

[17] S. Sprung, W. Rechenberg, Levels of heavy metals in clinker and cement, Zement-Kalk-

Gips. 47 (1998) 183.

[18] J. Jensen, S. Jepsen, The production, use and quality of sewage sludge in Denmark, Waste

Management. 25 (2005) 239-247.

[19] D. Fytili, A. Zabaniotou, Utilization of sewage sludge in EU application of old and new

methods-A review, Renewable and Sustainable Energy Reviews. 12 (2008) 116-140.

[20] L.E. Åmand, B. Leckner, Metal emissions from co-combustion of sewage sludge and

coal/wood in fluidized bed, Fuel. 83 (2004) 1803-1821.

[21] C. Senior and E. Eddings, Evolution of mercury from limestone, PCA R&D Serial No. 2949,

2006.

30

[22] R.J. Schreiber, C.D. Kellett and N. Joshi, Inherent mercury controls within the Portland

cement kiln system, PCA R&D Serial No. 2841, 2005.

[23] V.C. Johansen and G.J. Hawkins, Mercury speciation in cement kilns: A literature review,

PCA R&D Serial No. 2567, 2003.

[24] R.J. Schreiber and C.D. Kellett, Compilation of mercury emissions data, PCA R&D Serial

No. SN3091, 2009.

[25] Schreiber & Yonley Associates, Mercury emissions test report, Ash Grove Cement

Company Durkee, Oregon, Project No. 060204, 2007.

[26] Scott Learn, Cement plant cuts deal on mercury, http://legacy.lclark.edu/

org/nedc/objects/Ash_Grove.pdf, accessed May 10, 2010.

[27] Ravi Narayan, Mercury monitoring challenges facing the cement industry,

http://www.cemtrex.com/component/content/article/5-monitoring/125-mercury-monitoring-

challenges-facing-the-cement-industry.html, accessed July/22, 2010.

[28] German Cement Works Association, Activity report 1999-2001, 2001.

[29] C. Mones, Removal of elemental mercury from a gas stream facilitated by a non-thermal

plasma device, Final report on jointly sponsored research, task 34 under DE-FC26-98FT40323,

2006.

[30] J.H. Pavlish, E.A. Sondreal, M.D. Mann, E.S. Olson, K.C. Galbreath, D.L. Laudal, S.A.

Benson, Status review of mercury control options for coal-fired power plants, Fuel Processing

Technology. 82 (2003) 89-165.

[31] T.R. Carey, C.F. Richardson, R. Chang, F.B. Meserole, M. Rostam-Abadi, S. Chen,

Assessing sorbent injection mercury control effectiveness in flue gas streams, Environ. Prog. 19

(2000) 167-174.

[32] E.G. Pacyna, J.M. Pacyna, F. Steenhuisen, S. Wilson, Global anthropogenic mercury

emission inventory for 2000, Atmospheric Environment. 40 (2006) 4048-4063.

[33] K.C. Galbreath, C.J. Zygarlicke, Mercury transformations in coal combustion flue gas, Fuel

Processing Technology. 65-66 (2000) 289-310.

[34] K.S. Park, Y.C. Seo, S.J. Lee, J.H. Lee, Emission and speciation of mercury from various

combustion sources, Powder Technology. 180 (2008) 151-156.

[35] R.H. Perry, D.W. Green, J.O. Maloney, (Eds.), Perry’s chemical engineers’ handbook, 7th

Ed. The McGraw-Hill Companies, Inc., 1997.

[36] Wikipedia, Category: Mercury compounds, http://en.wikipedia.org/wiki/Category:

Mercury_compounds, accessed September 8, 2010.

[37] F. Frandsen, K. Dam-Johansen, P. Rasmussen, Trace elements from combustion and

gasification of coal—An equilibrium approach, Progress in Energy and Combustion Science. 20

(1994) 115-138.

[38] R.N. Sliger, J.C. Kramlich, N.M. Marinov, Towards the development of a chemical kinetic

model for the homogeneous oxidation of mercury by chlorine species, Fuel Process Technol. 65-

66 (2000) 423-438.

31

[39] A. Licata, R. Beittel and T. Ake, Multi-pollutant emissions control & strategies: Coal-fired

power plant mercury control by injecting sodium tetrasulfide, Institutes of Clean Air Companies

(ICAC) Forum 2003, Nashville, TN, October 14-15, 2003.

[40] S.J. Lee, Y. Seo, H. Jang, K. Park, J. Baek, H. An, K. Song, Speciation and mass distribution

of mercury in a bituminous coal-fired power plant, Atmospheric Environment. 40 (2006) 2215-

2224.

[41] X. Yang, Y. Zhuo, Y. Duan, L. Chen, L. Yang, L. Zhang, Y. Jiang, X, Xu, Mercury

speciation and its emissions from a 220 MW pulverized coal-fired boiler power plant in flue gas,

Korean Journal of Chemical Engineering. 24 (2007) 711-715.

[42] S. Tang, X. Feng, J. Qiu, G. Yin, Z. Yang, Mercury speciation and emissions from coal

combustion in Guiyang, southwest China, Environmental Research. 105 (2007) 175-182.

[43] M.B. Chang, H.T. Wu, C.K. Huang, Evaluation on speciation and removal efficiencies of

mercury from municipal solid waste incinerators in Taiwan, The Science of The Total

Environment. 246 (2000) 165-173.

[44] B.W. Vosteen, R. Kanefke, H. Köser, Bromine-enhanced mercury abatement from

combustion flue gases-Recent industrial applications and laboratory research, VGB PowerTech.

86 (2006) 70.

[45] S. Niksa, C.V. Naik, M.S. Berry, L. Monroe, Interpreting enhanced Hg oxidation with Br

addition at Plant Miller, Fuel Process Technol. 90 (2009) 1372-1377.

[46] Y.E. Yudovich, M.P. Ketris, Chlorine in coal: A review, International Journal of Coal

Geology. 67 (2006) 127-144.

[47] S.V. Vassilev, G.M. Eskenazy, C.G. Vassileva, Contents, modes of occurrence and origin of

chlorine and bromine in coal, Fuel. 79 (2000) 903-921.

[48] S. Niksa, B. Padak, B. Krishnakumar, C.V. Naik, Process Chemistry of Br Addition to

Utility Flue Gas for Hg Emissions Control, Energy Fuels. 24 (2010) 1020-1029.

[49] M. Holmes and J. Pavlish, Mercury information clearinghouse, Quarter 3- Advanced and

developmental mercury control technologies, July 2004.

[50] S. Liu, N. Yan, Z. Liu, Z. Qu, H.P. Wang, S. Chang, M. Charles, Using bromine gas to

enhance mercury removal from flue gas of coal-fired power plants, Environ. Sci. Technol. 41

(2007) 1405-1412.

[51] M.B. Larsen, I. Schmidt, P. Paone, J. Salmento, A. Petersen and A.W. Jørgensen, Mercury

in cement production-A literature review, FLSmidth internal report, 2007.

[52] R.J. Angel, G. Cressey, A. Criddle, Edgarbaileyite, Hg6Si2O7: The crystal structure of the

first mercury silicate, American Mineralogist. 75 (1990) 1192.

[53] A.C. Robert, M. Bonardi, Erd, Richard C, Edgarbaileyite: the first known silicate of mercury,

from California and Texas, The Mineralogical Record. 21 (1990) 215.

[54] T.M. Owens, C.Y. Wu, P. Biswas, An equilibrium analysis for reaction of metal compounds

with sorbents in high temperature systems, Chem. Eng. Commun. 133 (1995) 31-52.

32

3

Review of technologies for mercury removal from

flue gas

This chapter reviews the available technologies for mercury removal from flue

gas, and the applicability of the technologies in cement plant is discussed. Focus is put on

mercury removal by sorbent injection. Tests of sorbents in lab-scale fixed-bed reactors,

slipstream pilot-scale reactors and full-scale plants are reported.

3.1 Introduction 

The options for mercury control include mercury avoidance by coal and raw

material cleaning, mercury removal by sorbent injection upstream of existing air

pollution control devices (APCDs), and enhanced mercury removal by oxidation.

Differences in fuel type and composition and pollution control devices make it necessary

to develop customized solutions for each plant. The suitable mercury control method for

a specific plant depends on the plant’s configuration, fuel types, and existing flue gas

controls used for other pollutants. In addition, the complicated chemistry and multiple

mechanisms governing mercury speciation in combustion facilities makes it necessary to

investigate mercury emission control technologies at conditions relevant to each specific

plant [1]. Mercury control technologies applied in power plants and waste incinerators

are reviewed in this section and the applicability of these technologies in cement plants is

discussed.

33

3.2 Mercury avoidance technology 

3.2.1 Coal cleaning 

Physical coal cleaning is used primarily to reduce the ash and pyritic sulfur

content of coal [2-4]. Approximately 75% of the U.S. Eastern and Midwestern

bituminous coals undergo physical coal cleaning prior to shipment to power plants. Less

than 20% of the western coals, such as Powder River basin (PRB) coal and Colorado

bituminous coals, are cleaned [5].

Ash and pyritic sulfur are removed due to the difference in the densities of these

materials compared to the organic constituents in the coal. Mercury present in a sulfide

form also has a high density and can be removed during physical coal cleaning.

Reduction in mercury levels in coals ranging from 10% up to 78% have been reported

[3,6,7]. The average mercury reduction resulting from physical coal cleaning is estimated

to be in the range of 20% to 37% [6,8]. The cost of waste water treatment is very high.

As most of the mercury is from the raw material in cement production, the extent of

mercury removal through coal cleaning is expected to be very limited.

3.2.2 Cement raw material cleaning 

Two methods are proposed in the patents for removing mercury from cement raw

materials, i.e., washing and gasification prior to feeding to the kiln [9,10]. By water

washing the water-soluble mercury in the raw materials is removed. In the gasification

process, raw materials are introduced into a heating furnace and mercury and its

compounds contained in the raw materials are gasified. The resulting gas is introduced

into an activated carbon adsorption tower, mercury and its compounds are adsorbed and

separated. However, no results on the mercury removal efficiency by raw material

cleaning are reported in the publications. Due to large amount of raw materials applied in

the cement production, the cost of raw material cleaning is expected to be extremely high

and this technology appears not suitable for mercury removal from cement plants.

34

3.2.3 Fuel switching 

The use of tire-derived fuels (TDF) and substitution of coal and petroleum coke

with natural gas could potentially result in a modest reduction in the mercury emissions

due to the replacement of mercury-containing fossil fuels with low mercury fuels. Co-

firing of TDF with a subbituminous coal in a 55 kw pilot-scale pulverized coal

combustor had a significant effect on mercury speciation in the flue gas [11]. With 100%

coal firing, there was only 16.8% oxidized mercury in the flue gas compared to 47.7%

when 5 wt.% TDF was co-fired and 84.8% when 10% TDF was co-fired. The

significantly enhanced mercury oxidation may be the result of additional homogeneous

gas phase reactions between elemental mercury and the additional chlorine from TDF

combustion. The chlorine content in TDF is about 600 ppmm. However, co-firing of

TDF in the pilot-scale combustor with a hybrid filter for mercury removal demonstrated

only limited improvement on mercury-emission control by the hybrid filter without

sorbent injection. The enhanced mercury oxidation from co-firing TDF has potential in

mercury emission control for power plants equipped with a wet flue gas desulphurization

unit, since oxidized mercury is easily captured in the scrubber. Typically, kilns using

TDF have a replacement rate no greater than 30% of the total fuel requirement. Richards

et al. [12] summarized the available air emissions data for cement plants firing TDF and

literature applicable to cement kilns and concluded that the variability in mercury

concentrations and speciation overshadowed any beneficial impact on emissions due to

the firing of TDF.

Natural gas firing in boilers that are presently being fired with coal results in

direct and significant reductions in mercury. However, as mentioned previously the fuel

is usually not the dominant source of mercury in cement kiln flue gas. The limestone and

possibly other raw materials in the kiln feed provide most of the mercury that is

evaporated and emitted. Accordingly, the substitution of solid fuels would have only

limited impact.

35

3.3 Mercury removal by powdered activated carbon injection 

Powdered activated carbon (PAC) injection systems are well established as

commercial air pollution control processes for a variety of volatile organic compounds,

dioxin-furan, and heavy metals control applications [5]. The following three versions of

PAC processes are being considered for widespread use in coal-fired power plants [6]: (1)

PAC injection upstream of the existing dust collector system; (2) Gas cooling followed

by PAC injection upstream of the existing dust collector system; (3) Gas cooling of the

effluent gas stream of the existing dust collector system followed by PAC injection

upstream of a second dust collector for removal of the adsorbent.

The activated carbon particles remain suspended in the moving gas stream for

periods of one to three seconds. They then deposit onto the dust cake formed on the filter

bags. Additional mercury capture takes place when the mercury-containing gas stream

passes through the sorbent-containing dust cake. Electrostatic precipitators (ESPs) are

rarely used as the downstream polishing dust collector because the precipitated activated

carbon is partially isolated from the gas stream once it reaches the collection plate

surface.

3.3.1 Parameters affecting mercury removal by activated carbon 

injection 

There are a large number of variables that affect the adsorption of mercury on

powdered activated carbon. These include [5]: mercury speciation and concentration,

sorbent physical and chemical properties such as particle size distribution, pore structure

and distribution, and surface characteristics, gas temperature, flue gas composition,

sorbent concentration, mercury-sorbent contact time, and adequacy of sorbent dispersion

into the mercury containing gas stream.

Due to the differences of these variables among plants, there are large variations

in the reported PAC injection rates and mercury removal efficiencies in various studies

and commercial systems [13]. Therefore, it does not make sense to compare the mercury

36

removal efficiencies and sorbent injection rates without considering the actual conditions

in the specific plants.

Mercury speciation determines the mercury capture capacity of sorbents at a

given temperature. Pavlish and others [13] concluded that virgin activated carbon has a

higher rate of capture for mercuric chloride than for elemental mercury. Ho and others

[14] reported that sulphur-impregnated activated carbons have enhanced rates of

elemental mercury capture.

Pavlish et al. [13] conducted a detailed review on possible rate-controlling

mechanisms for mercury removal by sorbent injection. The overall reaction rates may be

limited by mass transfer from the bulk gas to the sorbent surface, the equilibrium

adsorption capacity, and the rates of reactions occurring on the sorbent surface.

All adsorption processes, especially those dependent on physisorption operate

more effectively at low temperatures due to the large adsorption capacity at low

temperatures. Adsorption processes for flue gas cleaning usually are operated in the

temperature range of 150°C to 200°C. The pilot plant studies of PAC injection indicate

that the mercury removal efficiency is strongly dependent on the gas temperatures [5].

Efficiencies of 10% to 70% have been measured at 170°C, and removals of 90% to 99%

have been measured at 100°C. Mercury capture takes place by both physisorption and

chemisorption [13]. With increasing temperature, physical adsorption decreases due to

the nature of exothermal adsorption process whilst chemisorption might be enhanced on

kinetics [15].

Mercury competes with a variety of gases for the adsorption sites on the activated

carbon. Water vapor is important because it is present at concentrations many orders of

magnitude above mercury. At moisture levels above 5% to 10%, moisture competition

can be significant. There are indications that high moisture levels in the flue gas will

suppress the capture of mercury by activated carbon [5,16]. It was postulated that water

molecules are able to fill micropores, thereby blocking adsorption sites for mercury.

Although it is agreed that water plays an important role in the mechanism of

mercury capture, there is disagreement in the literature about the effect of water on

37

mercury removal. Pavlish et al. [13] reported that reintroducing water into flue gas in a

lab-scale reactor at 135C after a period of sorption testing on dry flue gas resulted in an

immediate release of mercury from the activated carbon. However, in another lab-scale

study [17] the presence of moisture on the carbon surface was reported to promote

mercury bonding. About 75–85% reduction in Hg0 adsorption capacity was observed

when the carbon samples’ moisture at a level of 2 wt.% was removed by heating at

110C prior to the Hg0 adsorption experiments at room temperature. These observations

suggest that the moisture adsorbed on activated carbons plays a critical role in retaining

Hg0. It was postulated that adsorbed H2O is closely associated with surface oxygen

complexes and the removal of the H2O from the carbon surface by low-temperature heat

treatment reduces the number of active sites that can chemically bond Hg0 or eliminates

the reactive surface conditions that favor Hg0 adsorption [17]. Liu et al. [18] found that

the mercury adsorption capacity of sulfur impregnated activated carbon did not change

significantly when 5 vol.% water was added to the dry gas at 140C, however, the

adsorption capacity decreased 25% when the water content in the gas increased to 10

vol.%. These observations indicate that it is important to investigate the sorbent using the

same water vapor content as in the full-scale plant flue gas or in a wide range of moisture

content. Further investigations of the effect of water on mercury adsorption are desired to

reveal the dominating effects.

Miller et al. [19] and Ochiai et al. [20] conducted full factorial design

experiments in fixed-bed reactors to determine the relative effects of SO2, HCl, NO, and

NO2 on the elemental mercury capture ability of commercial activated carbons. Without

acid gases present, upon exposure to a baseline gas mixture of 6% O2, 12% CO2, 8%

H2O, and N2, the lignite activated carbon sorbent provided only about 10–20% initial

mercury capture of Hg0 for about 30 min and then fast breakthrough at 107°C [19,20].

Adding 50 ppmv HCl alone with the baseline gas improves the mercury adsorption

significantly [19,20]. It was also found out that adding NO or NO2 alone with the

baseline gas also improves the mercury adsorption capacity significantly. The mercury

capture increases from 10-20% for about 30 min using the baseline gas to 90-100% for

38

more than 2-6 h when 300 ppmv NO or 20 ppmv NO2 was added one at a time to the

baseline gases at 107°C.

When 1600 ppmv SO2 is added to the baseline gas the mercury adsorption

capacity will not be changed or only improved slightly [19,21]. Addition of 20 ppm SO3

to the gas reduced mercury capture by nearly 80%, and higher SO3 concentrations led to

further reductions in the mercury capture [21-23]. The competition between SO3 and

mercury for binding sites on the surface of activated carbon decreases the mercury

adsorption capacity.

The combination of 1600 ppmv SO2 and 20 ppmv NO2 additions resulted in

significant different mercury breakthrough profile compared to adding NO2 alone [19,20].

A highly significant interaction between SO2 and NO2 caused a rapid breakthrough of

mercury and is the controlling mechanism responsible for poor sorbent performance. The

detailed mechanism of SO2 and NO2 interaction is presented in section 3.3.5.2.

3.3.2 Tests of mercury sorbents in lab­scale fixed­bed reactors 

A good sorbent is expected to have high mercury adsorption capacity and fast

kinetics. A sorbent with good capacity but slow kinetics is not a good choice as it takes

mercury compound molecules too long time to reach the particle interior [24]. On the

other hand, a sorbent with fast kinetics but low capacity is not good either as a large

amount of sorbent is required for a given mercury removal. To satisfy these two

requirements, the sorbent must have a reasonably high surface area or micropore volume

and a pore network of relatively large pores for the transport of molecules to the interior.

3.3.2.1 Carbon­based sorbents 

As mentioned previously mercury capture is very sensitive to the flue gas

composition and temperature, and for this reason only the mercury capture capacities of

sorbents tested under simulated flue gas conditions are reported and compared here.

Extensive research has been conducted to study the sorbent mercury capture capacity

mainly using lab-scale fixed-bed reactors [25-36]. The mercury sorbents can be divided

39

into three groups, i.e., virgin carbon sorbents, chemically treated carbons, and non-

carbon based sorbents. The majority of the publications focused on elemental mercury

capture and only few studies investigated capture of HgCl2.

Typical properties of selected sorbents are reported in table 3.1. Coal source and

ash content relate to the composition of the coal and characterize the state of the carbon

such as fixed carbon/volatile ratio [37,38]. High ash content reduces the overall activity

of the activated carbon. Particle surface area is a measure of adsorption capacity and

describes the available surface for mercury adsorption. Pore size and distribution is an

indicator of sorbent quality, with smaller pores preferred. Particle size is used to describe

the degree of sorbent physical preparation. The mean particle size and distribution of

particle size are important parameters for evaluating mercury removal rate and pressure

drop. Smaller size provides faster rate of adsorption and results in larger pressure drop.

Content of bromine/chlorine/sulfur is considered as an indicator of the chemical

characteristics of sorbent responsible for mercury adsorption. Bulk density reflects a

gross approximation of the processing or surface area of a given carbon. The most

investigated sorbents in the literature is Darco FGD, which is a commercial lignite based

powdered activated carbon and is developed for heavy metal removal from incinerators

and power plants.

Table 3.1. Properties of selected sorbents.

Sorbents Sources Ash (wt%)

Sulfur (wt%)

Surface area (m2/g)

Pore volume (cm3/g)

Average pore size (nm)

Mass mean particle size(µm)

Bulk density (g/cm3)

Porosity Reference

Norit Darco FGD

Lignite coal 28.2-32.1

0.86-1.1

481-600

0.535-0.610

3.2 6.8-15 0.51 0.49-0.58

[25-27,29-32,35,39]

Norit Darco Insul

Lignite coal, based on Darco FGD, fine, chemically treated

700 6 0.32 [33,34]

Norit Darco Hg

Lignite coal 1.2 600 16-19 0.51 [40,41]

Norit Darco Hg-LH

Lignite coal, bromine treated

1.2 550 16-19 0.60 [41,42]

40

Sorbents Sources Ash (wt%)

Sulfur (wt%)

Surface area (m2/g)

Pore volume (cm3/g)

Average pore size (nm)

Mass mean particle size(µm)

Bulk density (g/cm3)

Porosity Reference

Calgon FluePac AC

Bituminous coal

5.8 0.7 606 0.285 32 0.585 [33,34,43]

Calgon HGR

Bituminous, sulfur treated

10.9-15

413-486

0.130 2.0 9.8 0.590 [44-46]

HOK standard

Lignite coal 10.0 0.60 300 0.620 63 0.55 [47]

HOK super

Lignite coal 10.0 0.60 300 24 0.44 [47]

Masda GAC

6.0 735 0.300 2.0 [48]

MnO2-AC MnO2 solution impregnated activated carbon

0.43 865 0.290 90 0.45 [28]

FeCl3-AC FeCl3 solution impregnated activated carbon

0.44 1470 0.920 90 0.58 [28]

Shanghai activated carbon

produced from wood by zinc chloride method

0.55 1850 1.050 90 0.67 [28]

Damao activated carbon

Bituminous coal

770 0.330 1.7 280 [49]

1% ZnCl2 Damao

Zncl2 treated Damao carbon

608 0.27 1.8 280 [49]

5% ZnCl2 Damao

Zncl2 treated Damao carbon

277 0.19 2.7 280 [49]

There is disagreement in the publications on the effect of mercury species on

activated carbon and char adsorption capacity. Yang et al. [13,50] reported that the

capture capacities of HgCl2 by bituminous char are larger by a factor of two than those of

elemental mercury using only CO2, O2, H2O and N2. However, other studies, as shown

in figure 3.1, show the opposite trend [25-27]. The elemental mercury adsorption

capacity for the studied carbons is about 0.5-4 times larger than the HgCl2 adsorption

capacity in the temperature of 110-160 C using simulated flue gas with 6% O2, 12%

CO2, 7% H2O, 50 ppmv HCl and 1600 ppmv SO2.

41

0

500

1000

1500

2000

2500

3000

Ad

so

rpti

on

cap

aci

ty,

g H

g/H

gC

l 2/g

-car

bo

n

45

g/m

3 H

g0 ,

Dar

co F

GD

[26

, 27]

54 g

/m3

Hg

0 ,la

b A

C f

rom

hig

h S

co

al [

25]

59

g/m

3 H

g0 ,

Dar

co F

GD

[25

]

60 g

/m3

Hg

Cl 2

,Dar

co F

GD

[2

5]

61

g/m

3 H

gC

l 2,la

b A

C f

rom

hig

h S

co

al [

25]

Figure 3.1. Effect of mercury speciation on mercury adsorption capacity on activated

carbon. Adsorption temperature is 135C and data are from [25-27]. Simulated flue gas

with 6% O2, 12% CO2, 7% H2O, 50 ppmv HCl and 1600 ppmv SO2.

Generally, the mercury adsorption capacities of carbon sorbents decrease when

the gas temperature is increased [13,50]. However, tests of Darco FGD at 100-135C

(see figure 3.2) from different studies are not in agreement with the trend. This is

probably due to the short exposure time for the test at 100C and the presence of SO3 in

the simulated gas for test at 120C. The adsorption capacity at 100C was measured after

2 h and before the complete breakthrough, while other tests were run until the complete

breakthrough of the sorbent bed was obtained. As discussed previously, the presence of

SO3 in the flue gas will decreases the sorbents’ mercury adsorption capacity. This

observation again illustrates the difficulty of analyzing the results from different studies

that are not conducted under the same conditions.

42

0

500

1000

1500

2000

2500

3000

Hg

ad

sorp

tio

n c

apac

ity,

g

Hg

/g_c

arb

on

100

oC

, 40

0 g

Hg

0 /m

3 [2

9-31

]

120

oC

, 29

0 g

Hg

0 /m

3 , S

O3 [

13,3

2]

135

oC

, 45

g

Hg

0/m

3 [2

6,27

]

135

oC

, 59

g

Hg

0/m

3 [2

5]

Figure 3.2. Effect of adsorption temperature on mercury adsorption capacity on Darco

FGD activated carbon. Data are from [13,25-27,29-32]. Gas composition: 100C: 1000

ppmv SO2, 50 ppmv HCl in N2; 120C: 8.3% O2, 14.8% CO2, 7.2% H2O, 278 ppmv NO,

650 ppmv SO2, 107 ppmv CO, 46 ppmv HCl, 27 ppmv SO3 in N2; 135C: 6% O2, 12%

CO2, 7% H2O, 1600 ppmv SO2, 50 ppmv HCl in N2;

Previous studies showed that the low chlorine concentration in the flue gas from

combustion of low-rank coals is a major limiting factor in the mercury control

performance using the virgin activated carbons [13,50]. Various chemically treated

carbons were developed to compensate for the lack of halogens in the combustion flue

gas. These include chloride-impregnated carbons [49,51-55], sulfur-impregnated carbons

[25,51,53,56-61], brominated carbons [50,55,62,63], iodine-impregnated carbons [52,53],

ozone-treated carbon [64], and carbon impregnated with metal compounds such as MnO2,

FeCl3 and CuCl2 [28,65-67]. The price of the chemically treated carbon is typically

higher than the non-treated one. The price of the non-treated Norit Darco Hg was about

1.1 US$/kg in 2007, while the bromine-treated carbons cost about 1.9-2.6 US$/kg [41].

43

Chlorine impregnation of a virgin activated carbon using dilute solutions of

hydrogen chloride leads to increases in fixed-bed capture of both elemental mercury and

mercuric chloride by a factor of 2-3 for a simulated flue gas without HCl, but with 7.1%

O2, 6.9% H2O, 3.4% CO2, 4.5 ppmv CO, 200 ppmv NOx and 500 ppmv SO2 [51]. It is

not reported how the chlorine impregnated carbon behaviors if HCl is present in the gas.

Coal-derived activated carbon from high-organic-sulfur coals was reported to

have a greater equilibrium Hg0 adsorption capacity than that prepared from low-organic-

sulfur coal when tested using a simulated flue gas with 6% O2, 7% H2O, 12% CO2, 50

ppmv HCl, and 1600 ppmv SO2 [25]. At 135C the equilibrium Hg0 adsorption capacity

of carbon derived from high-organic-sulfur coal, which contained 3.7 wt % total sulfur

and 2.9 wt% organic sulfur is 2718 µg Hg0/g_carbon, on the other hand the equilibrium

Hg0 adsorption capacity of carbon prepared from low-organic-sulfur coal with 1.2 wt%

total sulfur and 0.7 wt% organic sulfur was only 1304 µg Hg0/g_carbon. When the low-

organic-sulfur coal-derived activated carbon is impregnated with elemental sulphur at

600°C, its equilibrium Hg0 adsorption capacity is comparable to the adsorption capacity

of the activated carbon from the high-organic-surfur coal. Elemental sulphur-

impregnated carbons enhance elemental mercury removal due to the formation of

mercury sulphide on the carbon surface [68]. A portion of the inherent organic sulphur

in the starting coal, which remained in the activated carbon, plays an important role in

adsorption of elemental mercury. Besides organic sulphur, the surface area and

micropore area of the activated carbon also influence Hg0 adsorption capacity [25]. The

HgCl2 adsorption capacity is not as dependent on the surface area and concentration of

sulphur in the activated carbon as for adsorption of Hg0.

Another method for modifying carbon surfaces is oxidation, using reagents that

include oxygen, ozone, hydrogen peroxide, nitric acid, and permanganate [64]. Ozone

treatment of carbon surfaces leads to large increases in the elemental mercury capture

capacity by more than a factor of 100 when tested in Argon gas, but the activity is easily

destroyed by exposure to air, to water vapor, or by mild heating at 120C [64]. Freshly

ozone-treated carbon surfaces are shown to form labile C–O containing oxidizing groups,

44

which are likely to be epoxides or secondary ozonides. However, this ability fades with

aging. The finding opens the possibility of in-situ carbon ozonolysis to create fresh,

super-active sorbents with the additional benefit of sorbent hydrophilicity useful in

certain applications. Ozone treatment of fly ash carbon has been reported to inhibit the

adsorption of commercial surfactants in concrete paste, thus mitigating the known

negative effects of carbon on ash utilization [69-73]. Therefore, the enhanced mercury

removal could be a co-benefit of the ozone treatment.

3.3.2.2 Non­carbon sorbents 

For cement plant application a non-carbon sorbent is more attractive if the used

sorbent can be added to the final cement product or can be separated and regenerated.

Carbon can deteriorate the cement quality if the used carbon is not separated from the

cement materials by installing an expensive polishing filter. As inspiration to cement

plant application, research on development of non-carbon sorbents that do not adversely

impact sales of fly ash as a coal combustion byproduct for Portland cement and concrete

production are reported in this section.

Chemically synthesized manganese oxides powder has been demonstrated in

power plants to remove mercury, NOx and SO2 from flue gas [74-77]. The reacted

sorbent can be regenerated by a wet chemical process if the sorbent is injected just before

the added polishing fabric filter [74-77]. The simplified capture reactions for these

pollutants are suggested as following:

02Hg + MnO Mn*Hg complex (R3.1)

x 2 3 2NO + MnO Mn(NO ) (R3.2)

x 2 4SO + MnO MnSO (R3.3)

Non-carbonaceous materials or mineral oxides including silica gel, alumina,

molecular sieves, zeolites, and montmorillonite have been modified with various

functional groups such as amine, amide, thiol, urea, and additives such as elemental

sulfur, sodium sulfide, and sodium polysulfide to examine their potential as sorbents for

the removal of mercury vapor at coal-fired utility power plants [78]. A number of sorbent

45

candidates such as amine-silica gel, urea-silica gel, thiol-silica gel, amide-silica gel,

sulfur-alumina, sulfur-molecular sieve, sulfur-montmorillonite, sodium sulfide-

montmorillonite, and sodium polysulfide-montmorillonite, were synthesized and tested in

a lab-scale fixed-bed system under an argon flow for screening purposes at 70°C and

140°C. Several functionalized silica materials used for effective control of heavy metals

in the aqueous phase showed insignificant adsorption capacities for mercury control in

the gas phase, suggesting that mercury removal mechanisms are different in these two

phases. Among the synthesized samples, sodium polysulfide-impregnated

montmorillonite showed a moderate adsorption capacity at 70°C.

The commercial Amended Silicates sorbent uses silicate minerals as substrate

particles on which a chemical reagent with a strong affinity for mercury and mercury

compounds is impregnated [79,80]. A phyllosilicate substrate, for example, vermiculite

or montmorillonite, is used as an inexpensive support to a thin layer for a polyvalent

metal sulfide, ensuring that more of the metal sulfide is engaged in the sorption process.

The sorbent is prepared by ion exchange between the silicate substrate material and a

solution containing one or more of a group of polyvalent metals including tin, iron,

titanium, manganese, zirconium, and molybdenum. Controlled addition of sulfide ions to

the exchanged silicate substrate produces the sorbent. The silicates provide a low-cost

substrate material with average particle size of a few microns and extended surface area

for the amendment process. Due to their high silicate content, they have been proven

compatible with the continued sale of fly ash as a pozzolan material for concrete and

cement production. The price of the Amended Silicates sorbent is about 2.2-4.4 US$/kg

[81], which is comparable to the price of the chemically treated carbons [41]. However,

the performance data of the Amended Silicates sorbent are rarely reported due to the

concern of intelligent property.

A comparison between Darco FGD activated carbon and Ca(OH)2 indicated that

non-carbon-based sorbents with relatively high Ca contents can be fairly effective HgCl2

sorbents [29,30]. The Ca-based sorbents exhibited HgCl2 removal as high as half of the

removal shown by the Darco FGD activated carbon when 100 mg sorbent was tested in a

46

bench-scale fixed-bed reactor using simulated flue gas containing 10% CO2, 7% O2, 5%

H2O, and 173 ppmv SO2 [29,30]. However, the carbon-based sorbent showed superior

efficiency of elemental removal compared to Ca-based sorbent.

Full-scale investigations in coal-fired power plants have observed mercury

capture by unburned carbon in the fly ash [82]. Mercury removal by fly ash has also been

extensively studied to find a solution to the expensive mercury sorbents [36,83-87]. As

shown in figure 3.3, the amount of carbon in the fly ash has a strong effect on mercury

adsorption capacity of the fly ash. The mercury adsorption capacity increases with

carbon content in the fly ash, however, it is not directly proportional to the carbon

content. The mercury adsorption capacity of Nixon fly ash with 2% residual carbon is

about 30% of the commercial activated carbon Darco G60 [86].

0 10 20 30 40

Carbon content in the fly ash (%)

0

100

200

300

400

500

600

700

800

900

Ad

sorb

ed H

g (

g H

g/1

06 g

ash

) Nixon, 2% CCherokee, 8.7% CClark, 32.7% CHuntington, 35.9% C

Figure 3.3. Mercury adsorption capacity on fly ashes with different carbon content. Data

are from [86]. The applied adsorption temperature is 121C and elemental mercury

concentration is 4 mg/m3 with nitrogen as balance gas.

Dunham et al. [85] investigated 16 fly ash samples from a variety of sources and

coal types in a fixed-bed reactor at 121-177C using elemental mercury or HgCl2 in

simulated flue gas mixtures of O2, SO2, NO, NO2, H2O and HCl. While many of the ash

47

samples oxidized elemental mercury to HgCl2 in a range of 15-85%, not all of the

samples that oxidized mercury also captured elemental mercury. However, no capture of

elemental mercury was observed without accompanying oxidation. In general, oxidation

of elemental mercury increased with increasing amount of magnetite (Fe3O4) in the ash.

However, one high-carbon subbituminous ash with no magnetite showed considerable

mercury oxidation that may have been due to the carbon. Dunham et al. [85] suggested

that an iron oxide with a spinel-type structure is active in fly ash with respect to mercury

oxidation. Surface area as well as the nature of the surface, such as the oxygen

functionality and presence of halogen species appeared to be important for oxidation and

adsorption of elemental mercury. For the applied gas composition in Dunham’s study

[85], the capacity of the ash samples for HgCl2 was similar to that for elemental mercury.

There was a good correlation between the capacity for HgCl2 and the surface area. The

correlation between HgCl2 and loss on ignition was not as strong, suggesting that it is

not the carbon content alone but also properties of the ash, such as surface area, that

influence capture of HgCl2.

Based on the research of interactions between mercury and fly ash, carbon that

remains in pulverized coal fly ash could be used as an inexpensive adsorbent for mercury

removal. The fly ash would be injected into the flue gas prior to the particulate control

device [86,88] similarly to the way in which activated carbon is used, thus eliminating

large capital and sorbent costs. Due to the low carbon content and small mercury

adsorption capacity of fly ash, however, a large amount of fly ash may be required.

Another alternative to activated carbon might be the use of noble metal-based

sorbent. Noble metals such as gold and silver form reversible amalgams with mercury

[89]. A class of magnetic zeolite composites with supported silver nanoparticles has been

tested for elemental mercury removal from power plant flue gas [89]. Gaseous mercury is

captured by the sorbent and the mercury-laden sorbent particles are collected by an

existing dust collector and separated from the fly ash by magnetic separation. After mild

heat treatment to release captured mercury the sorbent is regenerated for the next cycle of

mercury capture. The technology is still in early stage and research is required regarding

48

the stability of the sorbent and possible regeneration cycles. Since noble metal is used in

synthesis of the sorbent it is expected that the sorbent is expensive. It is not clear how

this process could be cost effective compared to the activated carbon injection system

and the released mercury also must be captured by some sort of process.

3.2.2.3 In­situ produced sorbents 

To reduce the cost of sorbents, methods for in-situ production of activated carbon

from coal-fired power plants have been invented [33,34,90]. In the so-called Thief

process, partially combusted coal from the furnace of a pulverized coal power generation

plant is extracted by a lance and then re-injected into the ductwork downstream of the air

preheater [33,34,90]. Tests show that the Thief sorbents exhibit capacities for mercury

from flue gas streams that are comparable to those exhibited by commercially available

activated carbons. The process extracts 0.1-0.5% of the furnace gas in the boiler

depending on the desired sorbent injection rate and mercury removal level. The mass of

solids extracted from the furnace is very small in comparison to the mass of coal being

burned. The estimated heat loss is less than 0.3% for a 500 MWe power plant burning

PRB subbituminous coal.

Another process uses an oxy-fuel burner to devolatilize and activate the coal to

produce activated carbon [33,34,90]. In the burner natural gas is combusted together with

an oxygen stream, producing a high temperature oxygen-rich stream which passes

through a nozzle. Downstream of the hot oxygen nozzle the parent coal mixes with the

hot oxygen and begins to burn. Devolatilization and activation take place in a reactor

which leads to a particle separation step where the product is separated from the syngas

stream. The syngas can then be ducted to the boiler to provide added fuel value. At

several points in the process additives can be introduced to dope the product, or to

control the product morphology.

49

3.3.3 Sorbent injection in power plants 

Many activated carbons have been tested in U.S. power plants. Table 3.2 presents

the tested sorbents and applied APCDs and coals. The mercury removal efficiencies are

not included in the table due to various mercury removal efficiencies obtained at

complicated test conditions. Instead the mercury removal efficiencies as a function of

sorbent injection rate are shown in figures.

The most studied sorbents are Darco FGD, Darco Hg, and Darco Hg-LH. The

Darco Hg is formerly known as Darco FGD manufactured specifically for the removal of

mercury in coal fired utility flue gas emission streams [80], while Darco Hg-LH is

bromine impregnated. Although the mercury levels at the inlet of ACPDs are generally

similar, the extents of mercury removal by the existing APCDs without sorbent injection

are quite different. This is due to fact that different ranks of coal and APCD

configurations are applied by different power plants.

Without looking at the detailed data of the specific plants, it is difficult to

evaluate the sorbent performance by comparing the mercury removal efficiency.

However, some trends can be observed by comparing the results obtained under similar

conditions. Figure 3.4 compares the mercury removal at Holcomb and Stanton power

station by injection of Darco Hg-LH upstream of SDA and baghouse. At Holcomb and

Stanton power station the byproduct from SDA is disposed and therefore activated

carbon is injected before the SDA and baghouse. Figure 3.5 illustrates the mercury

removal by Darco FGD injection upstream of a new added so-called COHPAC compact

hybrid particle collector. COHPAC is an EPRI-patented design that places a high air-to-

cloth ratio fabric filter downstream of an existing ESP to improve overall particulate

collection efficiency. The results of mercury removal by Darco Hg injection upstream of

cold-side ESP are presented in figure 3.6. Up to 80 mg/m3 activated carbon is applied for

systems with FF, while up to 320 mg/m3 carbon is injected upstream of cold-side ESP.

50

Table 3.2. Summary of full-scale tests conducted in U.S. power plants. LNB: low NOx

burner, COHPAC: compact hybrid particulate collector

Location Test load MW

Coal APCD Inlet mercury g/Nm3,dry

Mercury removal without

sorbent, %

Sorbents

Holcomb, unit 1, 360 MW [91-93]

180, 360

PRB SDA@143C +Baghouse, LNB

10-12 0-13 Darco Hg, Darco Hg-LH, Calgon 208CP

Stanton, unit 10, 60 MW [93]

60 Lignite SDA, baghouse - - Darco Hg-LH

Stanton unit 1, 150 MW [94]

- PRB Cold-side ESP - 15 Brominated PAC (B-PAC)

Gaston, unit 3, 270 MW [95,96]

135 Low sulfur bituminous coal

COHPAC@143C, hot side ESP, LNB

7-10 6 Darco FGD, fine FGD, Insul, ESP ash

Big Brown, unit 2, 600 MW[97]

150 30%PRB/70% lignite,PRB

ESP, COHPAC @ 177C, LNB

- - Darco FGD, FGD/NaCl/CaCl2

Presque Isle, unit 7-9, 90 MW[98,99]

90 PRB Polishing baghouse

- - Darco FGD

Meramec, unit 2, 140 MW [91,93,100]

70 PRB Cold-side ESP @160C, LNB

10-12 15-30 Darco Hg-LH, Darco Hg

Pleasant Prairie, unit 2, 600 MW [96,101]

150 PRB Cold-side ESP@138C, SO3 conditioning

16-17 5 Darco FGD, Darco Hg, Insul, lime, Sorbalit

Brayton Point, unit 1, 250 MW [102]

125 Low sulfur bituminous coal

Cold-side ESP @138C, SO3 conditioning

17 - Darco FGD, Darco Hg, HOK, LAC

Leland Olds, unit 1, 220 MW [93]

220 Lignite Cold-side ESP, LNB

6-7 - Darco FGD Hg /CaCl2

St. Clair, unit 1, 145 MW [93]

145 85%PRB/15% bituminous coal

Cold-side ESP - - Brominated PAC (B-PAC)

Laramie River unit 3, 550 MW[91,103]

140 PRB SDA+cold-side ESP

10-12 4 Darco Hg-LH Darco Hg

Monroe, unit 4, 775 MW [91]

196 PRB/bituminous coal

Cold-side ESP@125C, SCR

5-10 10-30 Darco Hg-LH, Darco Hg, Darco XTR

Conesville, unit 6, 400 MW [91]

400 Bituminous coal

Cold-side ESP, wet FGD

15-30 50 Darco Hg-LH, Darco Hg

Plant Yates, unit 1, 100 MW[93]

100 Bituminous coal

Cold-side ESP, wet FGD

- - Super HOK

Salem Harbor unit 1, 85 MW [104]

85 Low sulfur bituminous coal

Cold-side ESP@125C, LNB

- - Darco FGD

Ameren Labadie unit 2, 630 MW [105]

630 PRB Cold-side ESP@150-180C. LNB

5-12 <15 Darco Hg, Darco Hg-LH, Darco Hg-E25c, Darco Hg-E26, Basf MS 2000, Calgon FLUEPACTM-MC PLUS

51

As shown in figure 3.4 and 3.5, mercury can be efficiently removed by activated

carbon injection upstream of SDA/baghouse or a polishing baghouse. When 32 mg/m3

Darco Hg-LH in SDA/baghouse system and Darco Hg in polishing baghouse system are

applied, about 80% of mercury can be removed. Further increase of the carbon injection

rate above 32 mg/m3 results in a slow increase of the mercury removal efficiency. The

mercury removal efficiency in SDA/baghouse system by Darco Hg-LH is larger than that

by Darco Hg. This is due to the applied Darco Hg-LH sorbent, which is bromine

impregnated and has larger mercury adsorption capacity than Darco FGD. The waste

disposal cost of sorbent injection upstream of SDA/baghouse is expected to be higher

since used activated carbon cannot be separated from the desulphurization product and

regenerated. Tests at Gaston power station showed that carbon injection significantly

increased the cleaning frequency of the COHPAC baghouse [95,96]. At an injection

concentration of 32 mg/m3 the cleaning frequency increased from 0.5 to 2

pulses/bag/hour, most likely due to the small particle size of the PAC causes a high

pressure drop.

0 20 40 60 80 100

PAC injection rate, mg/m3

0

10

20

30

40

50

60

70

80

90

100

Mer

cury

rem

ova

l eff

icie

ncy

, %

Stanton, lignite, Darco Hg-LHHolcomb, PRB, Darco Hg-LHHolcomb, PRB, Darco Hg

Figure 3.4. Mercury removal as a function of injection rate of Darco Hg-LH sorbent in

power plants using SDA and baghouse as APCDs. Data are from [91-93].

52

0 20 40 60 80

PAC injection rate, mg/m3

0

10

20

30

40

50

60

70

80

90

100

Mer

cury

rem

ova

l eff

icie

ncy

, %

Gaston, bituminousBig Brown, 30%PRB/70%lignitePresque Isle, PRB

Figure 3.5. Mercury removal as a function of injection rate of Darco FGD sorbent in

power plants by sorbent injection upstream of a polishing baghouse. Data are from [95-

99].

As shown in figure 3.6, much more than 32 mg/m3 Darco Hg activated carbon are

required to obtain 80% mercury removal by carbon injection upstream of cold-side ESP,

where the flue gas temperature is about 125-160C. This is due to the short contact time

between mercury vapor and injected carbon in the ESP and mercury is mainly captured

during the carbon particle in-flight period. When bromine treated carbons B-PAC and

Darco Hg-LH is used, the mercury removal efficiency across the cold side ESP increases

significantly.

53

0 50 100 150 200 250 300 350

PAC injection rate, mg/m3

0

10

20

30

40

50

60

70

80

90

100

Mer

cury

rem

ova

l eff

icie

ncy

, %

Pleasant, PRB,Darco Hg

Brayton, bituminous,Darco HgMeramec, PRB,Darco Hg

Leland Olds, lignite,Darco Hg

Monroe,PRB,SCR bypass, Darco Hg

Stanton unit 1, PRB, B-PAC

Meramec, PRB, Darco Hg-LH

Figure 3.6. Mercury removal as a function of injection rate of Darco Hg sorbent in power

plant by sorbent injection upstream of a cold side ESP. Data are from [91,93,96,100,101].

Tests at Pleasant Prairie showed that there was no significant effect on mercury

removal with PAC injection when SO3 was used as flue gas conditioning agent to obtain

optimal dust resistivity and improve ESP performance [96,101]. The level of applied SO3

at Pleasant Prairie was not reported. However, tests at Labadie unit 2 showed that the

presence of SO3 in the flue gas can decrease mercury capture by activated carbon [105].

The applied SO3 concentration in the flue gas at Labadie unit 2 was about 5-10 ppmv.

This is probably due to the competitive adsorption between Hg and SO3 since both

mercury and SO3 bind to the Lewis acid base sites on the activated carbon surface

[21,22].

In some plants burning PRB coals, it was observed that when the carbon injection

rate was increased above 160 mg/m3 the mercury removal efficiency by the cold side

ESP leveled off at about 60% [91,93,96,100,101]. At Brayton Point plant bituminous

coal was fired and the mercury removal increased with carbon injection rates in all the

tested ranges up to 320 mg/m3 reaching 90% mercury removal [102]. This is probably

due to the fact that at the Brayton Point the predominant species of mercury is in the

54

oxidized form since there is a significant amount of HCl present in the flue gas from

Brayton [102], in contrast to Pleasant Prairie where the majority of vapor phase mercury

was in the elemental form.

3.3.4 Sorbent injection tests at cement plant 

There are very limited studies on mercury removal by sorbent injection in cement

plants. In 2007 a six-week test was conducted at Ash Grove Cement Company’s Durkee

plant using a slipstream fabric filter after the main bag filter [106,107].

The overall goal of the tests at Durkee was to perform a parametric test on a

slipstream of actual flue gas to obtain an understanding of how various operating and

design parameters are likely to impact mercury control in the Durkee plant. The

evaluated parameters included activated carbon type, filter bag type, powdered activated

carbon injection rate, and filter air-to-cloth ratio.

The slipstream filter had twelve 152 mm3658 mm bags, corresponding to a

filtration area of 21 m2. The filter chamber and inlet duct were insulated and heated to

maintain a temperature of about 138°C. Mercury concentrations at the filter inlet and

outlet were measured by a Horiba/Nippon Instruments Corporation DM-6B, as well as

the Ontario hydro method. The tested carbons include Darco Hg, Darco Hg e-11, Darco

Hg LH, and Envergex e-sorb e11. The last two carbons are chemically treated. The

Darco Hg is prepared from lignite coal and the Darco Hg e-11 is a coarser version of the

Darco Hg. The particle size of the Darco Hg e-11 carbon is not reported. The test

duration for each parametric study was only about one hour. The flue gas compositions

are not publically reported and baghouse cleaning cycle is unknown.

Figure 3.7 shows the mercury removal efficiency as a function of PAC injection

rate for different carbons. For the Darco Hg and Darco Hg e-11, the mercury removal

efficiency increases only slightly from 80-90% to 90-95% when the PAC injection rate is

further increased above 48 mg/m3. The mercury removal efficiencies by the untreated

carbons are generally larger than those by the treated carbons when low injection rates

are applied. Treated carbons (Darco Hg LH and Envergex) have been shown to perform

55

better than untreated carbons in coal-fired boilers, especially in systems with ESP where

reaction times are short [108]. The halogens in the carbon act to oxidize the Hg in the

system and allow faster adsorption onto the carbon. This is critical in systems with higher

SOx concentrations, because SOx species have been shown to compete for active sites on

the carbon surface [21,22], as discussed earlier. The halogens on the treated carbon allow

the oxidized Hg to bind to the carbon surface before the SOx species consume the active

sites [108]. At the Durkee plant, the SOx concentrations in the slipstream baghouse are

very low compared to a coal-fired utility system. Thus the promoting effects of halogen

treated carbon are less pronounced.

0 20 40 60 80 100

PAC injection rate, mg/m3

0

10

20

30

40

50

60

70

80

90

100

Mer

cury

rem

ova

l eff

icie

ncy

, %

DARCO HgDARCO Hg e-11EnvergexDARCO Hg LH

Figure 3.7. Mercury removal efficiency as a function of PAC injection rate for different

sorbents at 138°C. The applied bag material is polyphenylene sulphide (PPS) and the air-

to cloth ratio is 1.22 m/min. Data are from [106].

At rates higher than 80 mg/m3, the untreated carbons appear to perform similarly

as the treated carbons with a mercury removal efficiency of about 90-95%. However,

injection of the treated carbons at 80 mg/m3 does not result in a significant increase in the

mercury control efficiency as compared to untreated carbon injected at 48 mg/m3. This

shows that there is no reason to choose halogenated carbon over untreated carbon,

56

particularly in light of the higher price and potential concerns associated with the use and

disposal of halogen-treated materials.

The trend of mercury removal efficiency of Darco Hg e-11 is similar to that of

finer Darco Hg, but the Hg removal results were lower by 10%–15%. This was most

likely caused by the larger particle sizes leading to more severe diffusion limitation.

Three bag types were tested, namely, polyphenylene sulphide (PPS), membrane

and fiberglass with membrane, while other conditions of the baghouse are the same. The

primary aim of testing different bag types is to investigate whether retention of carbon

particles on the bag surface can enhance the mercury removal efficiency. The

comparison of the performance of the tree bag types is presented in figure 3.8. At PAC

injection rate above 48 mg/m3 of Darco Hg all three bag types perform quite similarly at

138°C. Considering the uncertainty caused by the variability between the inlet and the

outlet mercury measurements, it is likely that the bags are performing essentially the

same at these conditions [108]. Then the only controlling factor for choosing the bag

type is the working temperature. The flue gas temperature in the bag filter area of the

cement process varies a lot and can exceed 200°C. Among the tested bag types, only the

membrane/fiberglass bag can withstand continuous operating temperatures at 260°C and

is therefore recommended.

57

0 20 40 60 80 100

PAC injection rate, mg/m3

0

10

20

30

40

50

60

70

80

90

100

Mer

cury

rem

ova

l eff

icie

ncy

, %

PPSmembrane/fiberglassMembrane

Figure 3.8. Effects of bag material on mercury removal efficiency at 138°C. The applied

air-to-cloth ratio is 1.22 m/min and the sorbent is Darco Hg. Data are from [106].

The effect of air-to-cloth ratio on mercury removal was tested using the

membrane/fiberglass bag. It is reported that at a PAC injection rate of 16 mg/m3 the

mercury removal efficiency increased with increasing the air-to-cloth ratio in the range of

1.2-3.0 m/min and when the injection rates were higher than 48 mg/m3 the mercury

removal efficiency increases only slightly with further increasing the injection rate since

the mercury removal efficiency is higher than 90% [106]. The increase of mercury

removal efficiency with filtration velocity might be due to the fast accumulation of

carbon on the bag surface. However, care must be taken when discussing the observation.

The cleaning control of the bags was not specified. It is unknown whether the bags were

cleaned at a fixed time interval or defined pressure drop over the filter. Most of the tests

were conducted for a period of only about 20 min, while only several tests were run for

up to 1-2 h.

Injection of Darco Hg before the fabric filter with membrane/fiberglass was also

tested in the raw-mill-off operating period. The mercury concentration at the filter inlet

during the raw-mill-on period was about 485 µg/Nm3, but increased to about 2600

58

µg/Nm3 during raw-mill-off operating period. Using an air-to-cloth ratio of 2.4 m/min,

moderate mercury removal efficiencies of 52% and 58% were obtained at PAC injection

rates of 48-80 mg/m3, respectively. A mercury removal efficiency of 88% was achieved

when the PAC injection rate was increased to 160 mg/m3.

Based on the parametric study at 138C, design parameters for the full-scale

sorbent injection upstream of a polishing filter at Durkee cement plant were

recommended. The untreated carbon, fiberglass with membrane bag type, and air-to-

cloth ratio of 1.8-2.4 m/min were suggested. The proposed sorbent injection rate is 48

mg/m3 and 80 mg/m3 for the raw-mill-on and raw-mill-off operation period, respectively.

The estimated mercury removal efficiency is 90% during raw-mill-on conditions and

60% during raw-mill-off conditions. The weighted mercury removal efficiency expected

is about 77% on annual average.

3.3.5 Carbon surface chemistry and mechanisms of mercury capture on 

carbons 

3.3.5.1 Carbon surface chemistry 

The surface chemistry of carbons determines their moisture content, catalytic

properties, acid-base character, and adsorption of polar species. It is related to the

presence of heteroatoms other than carbon within the carbon matrix. The most common

heteroatoms are oxygen, nitrogen, phosphor, hydrogen, chlorine, and sulphur [109].

During preparation of carbon and particularly during cooling and storage, carbon

materials are in contact with the ambient air so that elements such as H and O are fixed

on the surface, leading to oxygenated chemical functional groups [110]. Several

structures of oxygen functional groups have been proposed as shown in table 3.3.

Functional groups can be acidic, basic, or neutral in character. Surface oxygen groups on

carbon materials decompose upon heating by releasing CO and CO2 at different

temperatures. A CO2 peak results from carboxylic acids at low temperatures, or lactones

at higher temperatures; carboxylic anhydrides originate both a CO and a CO2 peak;

phenols, ethers, carbonyls, and quinones originate a CO peak.

59

Table 8. Surface oxygen groups on carbon and their decomposition by TPD, after

[110,111].

Group name Decomposition product

Decomposition temperature (C)

Carboxyl CO2 100-400

Lactone CO2 190-650 Carboxylic anhydrides CO+CO2 350-627 Phenolic CO 600-700 Ether CO 700 Carbonyl CO 700-980 Quinone CO 700-980

Besides oxygenated functions, nitrogenated functions can be introduced on

carbon surface by reaction of a carbon with a nitrogen-containing reactant or preparation

of a carbon from a nitrogen-containing precursor [110].

3.3.5.2 Mechanisms of mercury capture on carbons 

In order to understand the mercury capture mechanisms, it is important to

understand the chemical and physical nature of the mercury-sorbent interaction. X-ray

absorption fine structure (XAFS) spectroscopy and X-ray photoelectron spectroscopy

(XPS) are techniques that have been previously used to determine information about the

speciation and binding of mercury on a variety of materials [112,113]. XAFS spectra can

be defined by two regions which include X-ray absorption near-edge spectroscopy

(XANES) and extended X-ray absorption fine structure (EXAFS) spectroscopy. XANES

spectra provide information on the oxidation state and characteristics of the first neighbor

coordination environment. EXAFS spectroscopy provides more robust information on

the identity of nearest-neighboring elements, coordination values, and interatomic bond

distances.

XAFS spectroscopy was used to distinguish between elemental and oxidized

mercury in the sorbents by comparing the XAFS spectrum. Elemental mercury exhibits a

60

single peak only in the first-derivative of the mercury XANES spectrum, whereas most

mercuric compounds exhibit a two-peak spectrum [112]. The sorbents were tested for

mercury capture at temperatures lower than 200°C. The studied sorbents included

carbonaceous materials and inorganic-based material, such as lime-derived sorbents and

zeolites.

The XANES data imply that the capture of elemental mercury must involve an

oxidation process, either in the gas phase before interacting with the sorbent, or

simultaneously as the Hg0 atom interacts with the sorbent [112]. This is consistent with

the fact that all Hg-sorbed materials examined exhibit the characteristic dual inflection

point structure in their XANES spectra that is indicative of the formation of Hg–anion

chemical bonds. The anion could be virtually any available electronegative species, as

evidence has been seen for the formation of Hg–I, Hg–Cl, Hg–S, Hg–O, and Hg-Br

[112,113]. Modeling of mercury capture by activated carbon using density functional

theory shows that the mercury binding energies increase with the addition of the

following halogen atoms, F>Cl>Br>I [114]. Data from S and Cl XANES spectra, as well

as from the Hg XAFS data, strongly support the hypothesis that interaction of acidic

species (HCl, HNO3, H2SO4, HBr, etc.) in the flue gas with the sorbent surface is an

important mechanistic process that is responsible for creation of active sites for mercury

capture by chemisorption. The mechanisms of elemental mercury capture on the carbon

sorbents likely consist of surface-enhanced oxidation of the elemental mercury via

interaction with surface-bound halide species with subsequent binding by surface halide

or sulphate species [113].

The catalytic effects of carbon sorbents for mercury capture were investigated by

Olson et al. [115]. The studied carbons were lignite- and bituminous-derived carbon and

catalytic carbon, which were available commercially with enhanced catalytic

functionality for aqueous reactions such as decomposition of peroxides. Catalytic

carbons are produced by recarbonization of urea or ammonia-treated oxidized activated

carbons or by impregnation of nitrogen-containing polymers and pitches [115]. Without

acid gases in the gas stream at 150C, 50% mercury breakthrough was observed after 8

61

min for the catalytic carbon, while less than 1 min for the lignite- and bituminous-derived

carbons. Thus, a catalytic chemisorption mechanism predominates for the sorption of

mercury at these conditions.

The mercury adsorption capacity of the sorbent is inversely proportional to the

temperatures in a studied range of 50-150°C, indicating that a preliminary physisorption

step with mercury associating with a surface site takes place [115]. The chemisorption of

Hg0 is likely a multistep reaction. When the temperature is increased, the rate of each

chemical reaction step increases and the exothermic physisorption of Hg0 at non-

oxidizing binding sites will decrease. If the sorption process includes a preliminary

physisorption equilibration where Hg0 binds and desorbs at the active site, the

equilibration will show a negative temperature effect on the overall reaction rate, since

desorption is favored at higher temperatures. Although chemisorption may account for

the main sorption of mercury, the extent to which increasing the temperature may affect

the sorption rate cannot be predicted.

A detailed mechanism has been proposed to explain the effects of SO2 and NO2

as shown in figure 3.9 [116]. In the presence of NO2, Hg0 is catalytically oxidized on the

carbon surface to form the nonvolatile nitrate Hg(NO3)2, which is bound to basic sites on

the carbon. The Lewis base site refers to the zigzag carbon atom positioned between

aromatic rings [117]. Capture continues until the binding sites are used up and

breakthrough occurs. In the presence of SO2, some of the catalytic sites are converted to

a sulfate form where Hg(NO3)2 is no longer formed. Mercury is still oxidized on the

surface with NO2 acting as the oxidizing agent, but the product formed is a labile sulfur

compound, mercury bisulfate [Hg(SO4H)2]. The bisulfate in turn reacts with NO3- to

form a stable but volatile acidic form of the mercuric nitrate. The emission of Hg(NO3)2

or the hydrate Hg(NO3)2H2O has been confirmed by solvent trapping and gas

chromatography analysis. Sulfurous acid that accumulates from the hydration of SO2

converts the previously formed nonvolatile basic mercuric nitrate into the volatile form,

which explains the slow release of previously captured mercury over time in the presence

of NO2 and SO2.

62

Figure 3.9. Proposed heterogeneous model for mercury capture on carbon showing

potential impact of acid gases [116].

Sulphur trioxide can be present in power plant flue gas through one of the

following paths [21]: (1) During combustion, coal-S is converted to SO2 and a small

fraction of the sulfur is further oxidized to SO3. During combustion of high-sulfur coals,

a minor part of the sulfur is converted to SO3, leading to flue gas concentrations in the

range of 1-40 ppm. (2) SO3 is sometimes added to a level above 10 ppm to flue gas

upstream of an ESP as a conditioning agent and to improve ESP performance. SO3 and

H2SO4 have a low vapor pressure and can condense on fly ash and this reduces the

resistivity of the ash and allows it to be removed more efficiently by the ESP. (3) SO2

can be oxidized to SO3 by SCR catalysts installed for NOx reduction [118-120]. SCR

catalysts typically contain vanadium oxides, which are known catalysts for the oxidation

of SO2 to SO3 and Hg to HgCl2.

The inhibiting effect of SO3 on mercury capture by activated carbon injection has

been observed in full-scale power plant tests [21,121]. Possible mechanisms for the SO3

effect on mercury capture by activated carbon are postulated by Presto and Granite

[21,22]. In addition to removing mercury, activated carbon is also used as catalyst for

oxidation of SO2 to sulphuric acid [122,123]and as SO2 sorbent. There is competitive

adsorption between Hg and SO3 since both mercury and SO3 bind to the Lewis acid base

63

sites on the activated carbon surface. The adsorption of SO3 could be favored both

kinetically and thermodynamically. The concentration of SO3 in flue gas is typically in

the range of 1-40 ppm and this is orders of magnitude larger than typical mercury

concentrations. The bond formed between the S6+ species, such as sulfuric acid and

sulfates, and the carbon surface is stronger than the bond between mercury and the

surface. SO2 can oxidize to sulphate and form a chemical bond with the carbon surface

with a heat of adsorption of >80 kJ/mol. Some activated carbon catalysts for converting

SO2 to H2SO4 are self-poisoned by SO3 or sulfate buildup on the surface. A similar

phenomenon might explain the inhibiting effect of SO3 on mercury capture.

3.3.6 Processing and reuse of mercury laden activated carbon 

The existing production capacity for powdered activated carbon is only 10% of

the capacity required for full implementation of the activated carbon injection technology

to control mercury emissions [124]. The mercury sorption capacity of the activated

carbon is very low, about 1-4 mg of mercury per gram of sorbent, depending on the

mercury concentration in the flue gas [125]. This implies that 250 to 1000 g of activated

carbon are needed to remove 1 g of mercury in the flue gas. Therefore, a large quantity of

spent sorbents contaminated with various forms of mercury is produced.

Presently the PAC with adsorbed mercury must be disposed after use. In addition

to the purchase expense, the disposal of this material is also quite costly. There are strict

regulations for disposal of mercury-containing wastes [126]. Hazardous wastes

containing less than 260 mg/kg of total mercury are required to be treated to 0.20 mg/L,

measured using the toxicity characteristic leaching procedure (TCLP) for mercury

residues from retorting, and 0.025 mg/L TCLP for all other low mercury wastes. Wastes

that contain greater than 260 mg/kg total mercury are required to undergo roasting or

retorting in a thermal processing unit capable of volatilizing mercury and subsequently

condensing the volatilized mercury for recovery.

To reduce the PAC purchase expense and disposal cost of mercury-containing

PAC, a process has been developed to regenerate the used PAC and recover mercury

[124]. To separate PAC from the fly ash, PAC is injected between the main filter and the

64

polishing FF. The collected PAC is periodically removed from the filter and regenerated

in nitrogen process gas and is directed to a multiple activated carbon column gas

treatment system to remove the gaseous mercury from the cooled process gas stream.

After passing through the sulphur impregnated carbon columns, the carrier gas is injected

into the flue gas stream ahead of the carbon injection site. In this way only a small

amount of carbon with high mercury content requires disposal.

An inert atmosphere is required for the tray desorption furnace to avoid

significant losses of the PAC material during mercury desorption. Using a desorption

temperature of 550°C and a duration of 30 minutes, the PAC can be recycled at least 10

times without significant degradation of the adsorption characteristics in nitrogen [124].

It is unknown whether the cycled sorbent works satisfactorily in the real flue gas.

There are only few studies on the mercury desorption from exposed sorbents. It is

worth noting that mercury desorption is relevant both to recover the mercury and to

detoxify the adsorbing material in order to avoid its stabilization before land-filling or to

allow its reuse.

A study of mercury desorption in nitrogen from sulphur impregnated activated

carbon showed that the adsorption rate was faster than the desorption rate [59]. Mercury

desorption from sorbents is strongly affected by desorption temperature, with faster

desorption at high temperature and the mercury-sorbent pair. The desorption rate is

relatively fast initially and then levels off close to zero at a certain concentration of

mercury in sorbents.

Desorption of mercury from activated carbon and fly ash mixture was also carried

out in a fluidized bed reactor at temperatures up to 500°C [127]. All the mixtures had

constant mercury content, i.e., no mercury desorption was observed, until a critical

temperature was reached and then with rapidly decreasing mercury content as the

temperature was increased to higher levels. The critical temperature was found to be a

linear function of carbon contents in the mixtures, increasing from 330°C at 17% carbon

to 370°C at 33% carbon. The temperature at which all of the mercury was removed was

in the 450 to 500°C range.

65

3.3.7 Applicability of sorbent injection in cement plants 

As PAC systems are adapted for control of boilers, it will be possible to evaluate

the feasibility of these control techniques for cement kiln applications having

approximately the same mercury concentrations. Considering the differences between

boiler and kiln applications the possible application of PAC systems to cement kilns

appears to be considerably more challenging than to coal-fired boilers.

Powdered activated carbon injection systems do not appear to be appropriate

upstream of a cement kiln fabric filter system. Cement kilns must recycle a major portion

of the collected dust. Some kilns use the fabric filter system as an integral part of the raw

material processing system. Recycling the mercury laden activated carbon would result

in the revolatilization of the large majority of the mercury. Disposal of the activated

carbon containing cement kiln dust (CKD) also would be complicated because it might

be classified as a hazardous waste due to the presence of mercury.

Due to these issues, a powdered activated carbon injection system would have to

be installed downstream of the main kiln fabric filter to avoid the CKD recycling and

disposal issues. A second fabric filter would have to be installed after the main fabric

filter. The activated carbon injection system would have to be positioned to provide one

to two seconds residence time prior to entering the second fabric filter. The temperature

of this system would have to be controlled to less than 200C to ensure proper mercury

adsorption and reduce the risk of activated carbon fires in the fabric filter or solids

handling system.

3.4 Mercury removal by activated carbon bed 

Fixed and moving bed systems for mercury and dioxin-furan control are also used

in Europe [5]. In both types of systems, contaminant-laden gas is forced through a bed of

granular activated carbon.

One of the fixed bed systems used Sorbalit sorbent instead of activated carbon.

The Sorbalit sorbent consists of Portland cement, lime, carbon, and sulfur compounds

such as sublimed sulfur, Na2S, NaHS, and Na2S4 [16].

66

The quantity of mercury that can be retained on the sorbent at equilibrium is

important. The sorbent can be used at levels that approach the saturation capacity of the

sorbent at the operating gas temperature and gas stream conditions. However, the control

system must have the capability to remove the sorbent on at least a semi-continuous basis.

In fixed bed systems, the activated carbon must be replaced with fresh carbon at a

rate that is dependent primarily on the rate of approach to the mercury saturation level,

and the rate of static pressure increase. Spent carbon can be disposed of by combustion if

the unit is equipped with a wet scrubbing system. The combustion process destroys the

organic compounds captured in the carbon, and the wet scrubber collects the heavy

metals and acid gases. In this case, however, the elemental mercury might not be

removed due the insolubility of elemental mercury in the water. Another disposal option

is to dispose the carbon in a landfill. Because of the adsorbed pollutants, this waste may

require disposal as a hazardous waste. Another option is to heat the carbon and desorb

the pollutants from the carbon.

Slipstream tests of the activated carbon bed have been recently conducted in

several U.S. power plants [128]. Direct adaptation of existing carbon bed technology to

mercury removal from utility power plant flue gas is very costly because of the large flue

gas volumes and low mercury concentrations involved [129]. A thorough engineering

and economic analysis would be necessary to determine the feasibility of modifications

that reduce bed size and the amount of carbon in the bed. The effectiveness of the

modified beds for mercury removal under various flue gas conditions needs to be

determined. Furthermore, the tradeoff between gas velocity to the bed, bed sorbent size

and bed thickness, pressure drop, mercury and ash collection effectiveness, and bed

lifetime should be examined.

For cement plant application, the fixed bed activated carbon systems could not be

installed upstream of the main kiln bag filter or ESP. The high dust loadings in these

locations would quickly blind both types of beds and result in very high activated carbon

usage rates and disposal requirements. Accordingly, it would be necessary to install these

systems downstream of the main particulate matter control system. Similar to the power

67

plant application, installing a fixed bed carbon system in a cement plant will also be very

costly.

3.5 Mercury control by flue gas desulphurization systems 

Dry and wet scrubbers, commonly used in large scale combustion systems for

SO2 and HCl control can be simultaneously used for mercury retention, taking advantage

of the same sorbents used for sulphur or adding a new material for mercury [130-133].

Wet scrubbing systems predominately collect oxidized mercury [5,131]. In the purge

stream, mercuric chloride is collected as a precipitated solid along with the calcium

sulfate. When used as stand alone systems, they have the capability to achieve moderate-

to-high removal efficiencies for oxidized mercury. They are entirely ineffective in the

removal of the highly insoluble elemental mercury.

There are limited number of lime-based scrubbing systems used primarily for

particulate and SO2 control at lime kilns. There are presently only few cement kilns in the

United States equipped with wet SO2 scrubbing systems [5].

The stability of oxidized mercury captured in the flue gas desulphurization (FGD)

systems has been investigated and it was found that the captured oxidized mercury can be

reduced by aqueous phase reactions to form elemental mercury [6]. The insoluble

elemental mercury is rapidly released to the gas stream. Occurrence of mercury in FGD-

gypsum may threaten its re-use for wallboards since mercury can be released during the

heating steps in wallboard manufacturing [131].

Spray dryer absorbers (SDA) with a Ca(OH)2 slurry have been used for sulfur

dioxide and hydrogen chloride control at waste incinerators and coal-fired boilers. SDA

has been applied recently to cement kilns to control HCl that contributes to secondary

plume formation [5].

SDA systems would have to be installed after the main particulate matter control

system to ensure that captured mercury remains with a solid waste product and is not

recycled to the feed end of the kiln. With respect to fossil fuel fired boilers, the reported

mercury removal efficiencies by SDA systems are in the range of 50% to 60% for eastern

68

bituminous coals and 0% to 20% for western lignite and subbituminous coals [5]. The

difference is caused by the lower fraction of oxidized mercury for the lignite coals. With

respect to cement kilns, it appears unlikely that SDA systems will be more effective than

inherent adsorption in cement kiln systems.

3.6 Mercury removal by sodium tetrasulfide injection 

Sodium tetrasulfide (Na2S4) has been used as a sorbent to remove mercury from

flue gas in a number of waste-to-energy plants [1]. This technology should not be

confused with sodium sulfide Na2S that was tried in both Europe and U.S. without

success [134]. The shortcomings of Na2S are that it can leave a strong odor of hydrogen

sulfide (H2S) in the ash and it does not control all species of Hg. The major advantages of

the Na2S4 technology are that it controls elemental as well as ionic forms of Hg.

An aqueous Na2S4 solution is injected into the flue gas duct and such a system

can easily be retrofitted to an existing flue gas cleaning plant [134]. The sodium

tetrasulfide reacts with vapor phase mercury to form solid mercuric sulfide (HgS), which

is a solid at temperatures below about 580°C, and is insoluble [134]. By converting

vapor-phase mercury to an insoluble solid, it may be removed in a FF or ESP. Sodium

tetrasulfide can react with both oxidized and elemental mercury in accordance with the

following simplified reactions [134]:

2 4 2 2 3Na S HgCl HgS NaCl S (R3.4)

Hg S HgS (R3.5)

Decomposition of Na2S4 by an acid such as HCl can provide excess elemental

sulfur. It can also generate an alternate form of ionic sulfur, H2S, for reaction with

oxidized mercury as shown in the following reactions:

2 4 22 3 2Na S HCl H S S NaCl (R3.6)

2 2 2HgCl H S HgS HCl (R3.7)

In the absence of HCl, carbon dioxide may act as an acid for decomposition:

2 4 2 2 2 32 2 3 2Na S CO H O H S S NaHCO (R3.8)

69

Therefore, it is possible to eliminate both the elemental and ionic forms of mercury in the

flue gas.

However, H2S will still be produced in the process as shown in R3.6 and R3.8.

The problem of H2S odor in the ash cannot be avoided. This process is not suitable to

mercury removal from cement plant due to the presence of sodium, which could

deteriorate the cement quality.

3.7 Enhanced mercury removal by oxidation 

Oxidation pretreatment systems may convert elemental mercury to oxidized

mercury upstream of wet scrubber systems and even upstream of conventional particulate

matter control systems. Once in the oxidized form, mercury is captured in these air

pollution control systems at efficiencies approaching 85% [5]. The oxidation

pretreatment systems must be able to withstand the gas stream conditions upstream of the

air pollution control system used for capture of the oxidized mercury. Oxidation

pretreatment systems are only effective for the vapor phase mercury that is not adsorbed

on particle surfaces.

Selective non-catalytic reduction (SNCR) and selective catalytic reduction (SCR)

systems are used in coal-fired boilers and waste incinerators for the control of nitrogen

oxides. The impact of SNCR systems on the chemical form of mercury in a gas stream

appears to be minimal [5].

SCR systems use a catalyst to react ammonia and nitrogen oxides to provide

nitrogen and water. SCR systems are used extensively for coal-fired boilers and waste

incinerators. Full-scale tests have been performed in four U.S. power plants and the

results are presented in table 3.4 [135,136]. Significant oxidation of elemental mercury

across the SCR was observed in plant 2 and 4. While slight mercury oxidation over SCR

was experienced in pant 1 and 3. General conclusions from these tests are the oxidation

effect was quite variable and appears to be coal-specific and possibly catalyst-specific. In

particular, the catalyst type, space velocity, and catalyst age may all be important

variables. Plant 1 burns PRB coal with lower chlorine content and the catalyst is older.

70

More than 90% of the mercury in the flue gas at plant 1 is elemental mercury at the SCR

inlet. One possible explanation for the relatively low oxidation rate of the SCR at plant 3

is the relatively high space velocity, which is nearly double the space velocity compared

to other plants. In addition, the total inlet mercury concentration was more than twice the

levels seen at the other test sites.

Table 3.4. Test conditions and results of mercury oxidation over SCR catalyst in four U.S.

power plants. Data are from [135,136].

Plant 1 2 3 4 Coal PRB

subbituminous

Ohio bituminous high-sulfur

Pennsylvania bituminous low-to-medium sulfur

Kentucky bituminous medium-sulfur

Hg in coal (ppmm) 87 168 400 131 Cl in coal (ppmm) <60 573-1910 721-1420 357-1160 Catalyst vendor Cormetech Siemens KWH Cormetech Catalyst type Honeycomb Plate Honeycomb Honeycomb Catalyst age (h) 8000 2500 3600 3600 SCR space velocity (h-1)

1800 2125 3930 2275

Oxidized Hg increase over SCR (%)

From 8 to 18 From 48 to 91

From 55 to 65 From 9 to 80

Tests were also carried out at two Danish power plants equipped with Topsøe

DNX SCR catalysts and firing bituminous coal with different Cl-levels (0.01 to 0.13%)

[137]. The tests show a high degree of Hg0 oxidation over the SCR catalyst, ranging from

53 to more than 90% depending on operating conditions (load, coal type) and the

sampling method. Furthermore, it is shown that high coal chlorine results in higher Hg0

conversion over the SCR reactor than low chlorine coal.

The effects of different gases on mercury oxidation over SCR catalysts have been

studied [138-143]. Hydrogen halogens (HF, HCl, HBr, and HI) promote mercury

oxidation over the SCR catalyst [141]. It is HCl and not Cl2 that is the major source of

chlorine that dominates the Hg0 oxidation process within the typical SCR temperature

71

range (300-350°C) in a real flue-gas atmosphere [139], while NH3 shows a small

detrimental effect. Adding 2000 ppmv SO2 to baseline gases that contain 6% O2, 12%

CO2, 8% H2O, 550 ppmv NH3, 600 ppmv NO, 18.5 ppmv NO2 without HCl only

increases the mercury oxidation over SCR from 3% using baseline gases to 7% with

2000 ppmv SO2 [139,142].

Adding 50 ppmv SO3 to the baseline gases improves the mercury oxidation to

20% [139,142]. Adding 50 ppmv HCl to the baseline gases without SO2/SO3 results in

71% oxidized mercury in flue gas across the SCR compared to 45% mercury oxidation

when 50 ppmv SO3 was further added to the flue gas. With 50 ppmv HCl and 2000 ppmv

SO2 were added to the baseline gases, mercury oxidation recovered to 64%. The

combination of 2000 ppmv SO2/50 ppmv SO3 and 50 ppmv HCl showed a 63% mercury

oxidation. These observations indicate that both SO2 and SO3 had a negative effect on

mercury–chlorine oxidation over the SCR as a result of slower mercury oxidation by the

sulfated site compared to that of the chlorinated site. The extent of the mitigating effect

by the 2000 ppmv SO2 was not as severe as the 50 ppmv SO3 since the concentration of

SO3 derived through SO2 oxidation over SCR was much lower than the 50 ppmv SO3.

There is possible competition between HCl, SO2, and SO3 over the SCR catalyst.

Conventional SCR catalyst with higher mercury oxidation capacity has been

closely related to higher oxidation of SO2 to SO3 [144]. Higher SO2 oxidation in coal-

fired applications can cause negative downstream impacts such as air heater fouling, flue

duct corrosion and visible stack plumes. The SO2 to SO3

conversion is designed to be less

than 1.5% at SCR operating conditions. Research has been focused on developing new

SCR catalyst that has higher oxidation rate of elemental mercury and very low SO2 to

SO3 conversion [144].

Two studies have demonstrated that ultraviolet radiation in the presence of solid

titanium dioxide (TiO2) results in the photocatalytic conversion of elemental mercury to

HgO when HCl is not present in the gas [6,145,146]. The TiO2 is readily available as a

major component of conventional SCR catalysts. Accordingly, the combination of

ultraviolet light and SCR catalysts could in principle be used to oxidize elemental

72

mercury. Compared to lab-scale study where ultraviolet light can be readily applied,

application of ultraviolet light in monolith catalyst could be a technical challenge.

The application of SCR systems to cement kilns continues to be precluded by

problems associated with alkali metal and arsenic related catalyst poisoning, SO2

oxidation to SO3, particulate matter loadings that can be 5 to 20 times higher than coal-

fired boiler high dust systems, and non-ideal temperature ranges for SCR catalysts

[147,148]. For these reasons, it is unlikely that an SCR system will be used for the

control of nitrogen oxides and oxidation of elemental mercury in cement plants in the

near future.

3.8 Mercury removal by roaster process 

Recently a new process for mercury removal from the cement kiln flue gas by a

roaster was invented and patented [149-151]. As mentioned earlier recycled cement kiln

dust from the main bag filter has high mercury content. The mercury flow in the

collected cement kiln dust is about 60% of the mercury inlet to the cement kiln [152].

This indicates that dust captured in the main baghouse acts as a natural sorbent for

mercury. This mercury enriched dust is taken to the new mercury roaster process for

cleaning before the dust is returned to the system. Figure 3.10 illustrates an example of a

mercury roaster installation. The baghouse dust is fed to a roasting system which uses a

heat source (for example kiln bypass gas, cooler vent gas, or hot gas generator) to heat

the dust above the boiling point of mercury compounds. While the mercury is still in the

gas phase, the gas stream enters a hot ESP which removes most of the cleaned dust. This

dust is taken back to the blending silo to be part of the kiln feed. After the ESP, the gas

stream is cooled below the mercury boiling point so that the mercury can condense on

the dust particles that were not captured in the ESP and additional sorbent is added to the

gas stream here to capture the mercury. The cleaned gas after the baghouse is vented to

the atmosphere. Depending on the type of applied sorbents the mercury enriched

dust/sorbent collected in the baghouse can be transported to the finish mill area to be

73

added to the cement or disposed as waste. The air and sorbent flow rates are expected to

be smaller than what would be seen with a full carbon injection system.

Figure 3.10. Sketch of the roaster process [149].

It should be pointed out that the process is still under development. Information is

lacking on the achievable mercury removal efficiency and operating cost. Since most

dust is removed by the ESP it appears that sorbent is still required for capture the

mercury evaporated from the roaster. Calcium chloride may be required to oxidize the

elemental mercury and enhance mercury capture by the sorbent. It is unclear how

effectively the elemental mercury can be oxidized and how much calcium chloride is

required.

3.9 Conclusions 

Mercury can be removed from the flue gas by fuel cleaning and switching, raw

material cleaning, sorbent injection, sorbent bed, oxidation by catalyst and subsequent

removal by wet scrubber, spray drier absorber, and roaster process with smaller sorbent

injection system. Presently sorbent injection is considered as the most promising and

developed mercury removal technology. Mercury removal by sorbent injection can be

affected by many factors such as mercury speciation and concentration, flue gas

74

composition and temperature, mercury vapor-sorbent contacting time, sorbents and

sorbent dispersion, etc. Due to the high moisture level and lack of carbonaceous particles

in the cement kiln flue gas, and release of the captured mercury during recirculation to

the kiln, the application of sorbent injection to cement kilns will be more challenging and

the obtained knowledge from coal-fired power plants and waste incinerators cannot be

applied to cement kiln directly. The PAC injection system should be installed

downstream of the main kiln fabric filter and upstream of a new added polishing fabric

filter to avoid the cement kiln dust recycling and increased disposal issues.

Powdered activated carbon is the most widely used sorbent for mercury removal

from flue gas. However, there is a lack of fundamental investigation of mercury

adsorption by activated carbon in simulated cement kiln flue gas. Even for power plant

application, mercury capture kinetics is not available in most of the publications and

many of the studies were carried out in air or nitrogen without acid gases presence. The

majority of the publications focused on elemental mercury capture and only few studies

investigated capture of HgCl2 which is a major mercury species. To reduce the cost of

sorbent and possible disposal expense, non-carbon based and concrete/cement friendly

sorbents such as Amended SilicateTM have been developed. Other developments include

regeneration and recirculation of sorbents and in-situ generation of activated carbon. The

performance of these sorbents needs to be proved in the full-scale application.

The carbon-oxygen surface complexes and flue gas composition play an

important role in mercury removal by activated carbon injection. Both physisorption and

chemisorption are involved in mercury capture by carbons. The mechanisms of elemental

mercury capture on the carbons consist of surface-catalyzed oxidation of the elemental

mercury via interaction with surface-bound halide species with subsequent binding by

surface halide or sulphate species. Co-presence of SO2 and NO2 in the flue gas results in

a poor performance of carbons. There is competitive adsorption between Hg and SO3

since both mercury and SO3 bind to the Lewis acid base sites on the activated carbon

surface.

75

3.10 Further research requirement 

Activated carbon injection is a promising technology, but further research is

needed to provide the best sorbent with effective mercury capture at a low cost.

Investigation of mercury capture by the activated carbon using simulated cement kiln

flue gas is imperative to evaluate whether the activated carbon is also a promising

sorbent for cement plant application. Lab-scale tests are desired to obtain kinetics and

study the effects of different operating parameters.

More focus is needed on developing alternative sorbents. Fly ash and cement raw

materials such as clay and silica might be used as cement-friendly sorbents and

alternatives for activated carbon. A better understanding of mercury removal by fly ash

and other cement-friendly sorbents is therefore needed. Fundamental investigation on the

regeneration of the sorbents and enhancement of the sorbents by adding chemical agents

during regeneration is required. Focus should be put on desorption temperature,

separation and purification of collected mercury compounds and possible regeneration

cycle of the sorbents. Possibility of regenerating the sorbent by hot flue gas from the kiln

system and in-situ enhancement of the sorbent should be investigated.

3.11 Abbreviations 

APCD: Air pollution control device

CKD: Cement kiln dust

COHPAC: Compact hybrid particulate collector

ESP: Electrostatic precipitator

EXAFS: Extended X-ray absorption fine structure

FF: Fabric filter

FGD: Flue gas desulphurization

LNB: Low NOx burner

PAC: Powdered activated carbon

PPS: Polyphenylene sulphide

PRB: Powder River Basin

76

SCR: Selective catalytic reduction

SDA: Spray dryer absorber

SNCR: Selective non-catalytic reduction

TCLP: Toxicity characteristic leaching procedure

TDF: Tire-derived fuel

XAFS: X-ray absorption fine structure

XANES: X-ray absorption near-edge spectroscopy

XPS: X-ray photoelectron spectroscopy

3.12 References  

[1] A. Licata, R. Beittel and T. Ake, Multi-pollutant emissions control & strategies: Coal-fired

power plant mercury control by injecting sodium tetrasulfide, Institutes of Clean Air Companies

(ICAC) Forum 2003, Nashville, TN, October 14-15, 2003.

[2] F. Rubiera, S.T. Hall, C.L. Shah, Sulfur removal by fine coal cleaning processes, Fuel. 76

(1997) 1187-1194.

[3] D. Akers, R. Dospoy, Role of coal cleaning in control of air toxics, Fuel Processing

Technology. 39 (1994) 73-86.

[4] R.J. Conzemius, C.D. Chriswell, G.A. Junk, The partitioning of elements during physical

cleaning of coals, Fuel Processing Technology. 19 (1988) 95-106.

[5] J. Richards, Capabilities and limitations of available control technologies for mercury

emissions from cement kilns, PCA R&D Serial No. 2748a, 2005.

[6] T.D. Brown, D.N. Smith, R.A. Hargis Jr., W.J. O'Dowd, Mercury measurement and its

control: What we know, have learned, and need to further investigate, EM: Air and Waste

Management Association's Magazine for Environmental Managers. (1999) 47-50.

[7] D.J. Akers, Coal cleaning controls HAP emissions, Power Eng. 100 (1996) 33.

[8] U.S. EPA, Mercury study report to congress, EPA-452/R-97-003, 1997.

[9] N. Yoshiyuki and T. Haruhiko, Method of producing cement, Patent, JP2007326750, 2007.

[10] T. Junichi, U. Yasuhiro and Y. Osamu, Method of treating exhaust gas from cement kiln,

Patent, JP2003192407, 2003.

[11] Y. Zhuang, S.J. Miller, Impact of supplemental firing of tire-derived fuel (TDF) on mercury

species and mercury capture with the advanced hybrid filter in a western subbituminous coal flue

gas, Energy Fuels. 20 (2006) 1039-1043.

[12] J. Richards, D. Goshaw, D. Speer and T. Holder, Air emissions data summary for Portland

cement pyroprocessing operations firing tire-derived fuels, PCA R&D Serial No. 3050, 2008.

77

[13] J.H. Pavlish, E.A. Sondreal, M.D. Mann, E.S. Olson, K.C. Galbreath, D.L. Laudal, S.A.

Benson, Status review of mercury control options for coal-fired power plants, Fuel Processing

Technology. 82 (2003) 89-165.

[14] T.C. Ho, P. Yang, T.H. Kuo, J.R. Hopper, Characteristics of mercury desorption from

sorbents at elevated temperatures, Waste Management. 18 (1998) 445-452.

[15] R. Yan, D.T. Liang, L. Tsen, Y.P. Wong, Y.K. Lee, Bench-scale experimental evaluation of

carbon performance on mercury vapour adsorption, Fuel. 83 (2004) 2401-2409.

[16] J.J. Cudahy, R.W. Helsel, Removal of products of incomplete combustion with carbon,

Waste Manage. 20 (2000) 339-345.

[17] Y.H. Li, C.W. Lee, B.K. Gullett, The effect of activated carbon surface moisture on low

temperature mercury adsorption, Carbon. 40 (2002) 65-72.

[18] W. Liu, R.D. Vidic, T.D. Brown, Impact of flue gas conditions on mercury uptake by sulfur-

impregnated activated carbon, Environ. Sci. Technol. 34 (2000) 154-159.

[19] S.J. Miller, G.E. Dunham, E.S. Olson, T.D. Brown, Flue gas effects on a carbon-based

mercury sorbent, Fuel Processing Technology. 65-66 (2000) 343-363.

[20] R. Ochiai, M.A. Uddin, E. Sasaoka, S. Wu, Effects of HCl and SO2 concentration on

mercury removal by activated carbon sorbents in coal-derived flue gas, Energy & Fuels. 23 (2009)

4734.

[21] A.A. Presto, E.J. Granite, Impact of sulfur oxides on mercury capture by activated carbon,

Environ. Sci. Technol. 41 (2007) 6579-6584.

[22] A.A. Presto, E.J. Granite, A. Karash, Further investigation of the impact of sulfur oxides on

mercury capture by activated carbon, Ind Eng Chem Res. 46 (2007) 8273-8276.

[23] J. Jarvis, F. Meserole, SO3 effect on mercury control, Power Eng. 112 (2008) 54-60.

[24] D.D. Do, Adsorption analysis: equilibria and kinetics, Imperial College Press, 1998.

[25] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R. Chang,

Preparation and evaluation of coal-derived activated carbons for removal of mercury vapor from

simulated coal combustion flue gases, Energy Fuels. 12 (1998) 1061-1070.

[26] M. Rostam-Abadi, S.G. Chen, H.C. Hsi, M. Rood, R. Chang, T.R. Carey, B., Hargrove, C.

Richardson, W. Rosenhoover, F. Meserole, Novel vapor phase mercury sorbents, Proceedings of

the EPRI-DOE-EPA Combined Utility Air Pollutant Control, Washington, DC, Aug 25–29, 1997.

[27] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang and F.B. Meserole, Factors

affecting mercury control in utility flue gas using sorbent injection, Proceedings of the Air &

Waste Management Association's 90th Annual Meeting, Toronto, Ontario, Canada, June 8-13,

1997.

[28] C.X. Hu, J.S. Zhou, Z.Y. Luo, S. He, G.K. Wang, K.F. Cen, Effect of oxidation treatment on

the adsorption and the stability of mercury on activated carbon, Journal of Environmental

Sciences-China. 18 (2006) 1161-1166.

[29] B. Ghorishi and B.K. Gullett, Fixed-bed control of mercury: Role of acid gases and a

comparison between carbon-based, calcium-based, and coal fly ash sorbents, Proceedings of the

78

EPRI/DOE/EPA Combined Utility Air Pollutant Control Symposium, Washington, DC, August

25–29,1997.

[30] S.B. Ghorishi and C.B. Sedman, Combined mercury and sulfur oxides control using

calcium-based sorbents, Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant

Control Symposium, Washington, DC, August 25–29, 1997.

[31] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption by activated

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy Conference,

Research Triangle Park, NC, 22-25 April, 1997.

[32] N. Hutson, C. Singer, C. Richardson, J. Karwowski and C. Sedman, Practical applications

from observations of mercury oxidation and binding mechanisms, EPRI-DOE-EPA-AWMA

Combined Power Plant Air Pollutant Control Mega Symposium, Washington DC, August 30 -

September 2, 2004.

[33] E.J. Granite, M.C. Freeman, R.A. Hargis, W.J. O'Dowd, H.W. Pennline, The thief process

for mercury removal from flue gas, J. Environ. Manage. 84 (2007) 628-634.

[34] W.J. O’Dowd, H.W. Pennline, M.C. Freeman, E.J. Granite, R.A. Hargis, C.J. Lacher, A.

Karash, A technique to control mercury from flue gas: The thief process, Fuel Processing

Technology. 87 (2006) 1071-1084.

[35] R. Bhardwaj. Impact of temperature and flue gas components on mercury speciation and

uptake by activated carbon sorbents. Master thesis. Master, University of Pittsburgh, 2007.

[36] M.M. Maroto-Valer, Y. Zhang, E.J. Granite, Z. Tang, H.W. Pennline, Effect of porous

structure and surface functionality on the mercury capacity of a fly ash carbon and its activated

sample, Fuel. 84 (2005) 105-108.

[37] J.E. Cichanowicz, R. Stewart, K. Baldrey, Powdered activated carbon (PAC)

characterization tests for coal-fired utility application, Power Plant Air Pollutant Control

"MEGA" Symposium 2008, Baltimore, MD, August 25-28, 2008.

[38] Carbochem, Activated carbon 101, http://www.carbochem.com/activatedcarbon101.html,

accessed September 21, 2010.

[39] Norit Americas Inc., Datasheet of Darco FGD powdered activated carbon, 2008.

[40] Norit Americas Inc., Datasheet of Norit Darco Hg activated carbon, 2007.

[41] K. Dombrowski, Evaluation of low-ash impact sorbent injection technologies at a Texas

lignite/PRB fired power plant. Presented at the DOE/NETL Mercury Control Technology

Conference, Pittsburgh, PA, USA, December 11-13, 2007.

[42] Norit Americas Inc., Datasheet of Norit Darco Hg-LH activated carbon, 2007.

[43] Calgon, Data sheet of FluePac-MC plus powdered activated carbon, 2010.

[44] S.H. Lee, Y.J. Rhim, S.P. Cho, J.I. Baek, Carbon-based novel sorbent for removing gas-

phase mercury, Fuel. 85 (2006) 219-226.

[45] Y. Shin, G. Fryxell, W. Um, K. Parker, S. Mattigod, R. Skaggs, Sulfur-Functionalized

Mesoporous Carbon, Advanced Functional Materials. 17 (2007) 2897-2901.

79

[46] J.W. Graydon, X. Zhang, D.W. Kirk, C.Q. Jia, Sorption and stability of mercury on

activated carbon for emission control, J. Hazard. Mater. 168 (2009) 978-982.

[47] J. Wirling, J. Jablonski, Safety aspects in the use of carbonaceous sorbents during waste gas

treatment, Metallurgical Plant and Technology. 3 (2007) 144.

[48] R. Yan, Y.L. Ng, D.T. Liang, C.S. Lim, J.H. Tay, Bench-scale experimental study on the

effect of flue gas composition on mercury removal by activated carbon adsorption, Energy Fuels.

17 (2003) 1528-1535.

[49] H. Zeng, F. Jin, J. Guo, Removal of elemental mercury from coal combustion flue gas by

chloride-impregnated activated carbon, Fuel. 83 (2004) 143-146.

[50] H. Yang, Z. Xu, M. Fan, A.E. Bland, R.R. Judkins, Adsorbents for capturing mercury in

coal-fired boiler flue gas, Journal of Hazardous Materials. 146 (2007) 1-11.

[51] S.B. Ghorishi, R.M. Keeney, S.D. Serre, B.K. Gullett, W.S. Jozewicz, Development of a Cl-

impregnated activated carbon for entrained-flow capture of elemental mercury, Environ. Sci.

Technol. 36 (2002) 4454-4459.

[52] S.J. Lee, Y. Seo, J. Jurng, T.G. Lee, Removal of gas-phase elemental mercury by iodine-

and chlorine-impregnated activated carbons, Atmospheric Environment. 38 (2004) 4887-4893.

[53] E.J. Granite, H.W. Pennline, R.A. Hargis, Novel sorbents for mercury removal from flue gas,

Ind. Eng. Chem. Res. 39 (2000) 1020-1029.

[54] R.D. Vidic, D.P. Siler, Vapor-phase elemental mercury adsorption by activated carbon

impregnated with chloride and chelating agents, Carbon. 39 (2001) 3-14.

[55] Z. Mei, Z. Shen, T. Yuan, W. Wang, H. Han, Removal of vapor-phase elemental mercury by

N-doped CuCoO4 loaded on activated carbon, Fuel Processing Technology. 88 (2007) 623-629.

[56] H.C. Hsi, M.J. Rood, M. Rostam-Abadi, S. Chen, R. Chang, Effects of sulfur impregnation

temperature on the properties and mercury adsorption capacities of activated carbon fibers

(ACFs), Environ. Sci. Technol. 35 (2001) 2785-2791.

[57] S. Vitolo, M. Seggiani, Mercury removal from geothermal exhaust gas by sulfur-

impregnated and virgin activated carbons, Geothermics. 31 (2002) 431-442.

[58] K.D. Henning, K. Keldenicht, K. Knoblauch, J. Degel, Impregnated activated carbon for

mercury removal, Gas Separation & Purification. 2 (1988) 20-22.

[59] D. Karatza, A. Lancia, D. Musmarra, C. Zucchini, Study of mercury absorption and

desorption on sulfur impregnated carbon, Experimental Thermal and Fluid Science. 21 (2000)

150-155.

[60] W. Liu, R.D. Vidic, T.D. Brown, Optimization of high temperature sulfur impregnation on

activated carbon for permanent sequestration of elemental mercury vapors, Environ. Sci. Technol.

34 (2000) 483-488.

[61] J.A. Korpiel, R.D. Vidic, Effect of sulfur impregnation method on activated carbon uptake

of gas-phase mercury, Environ. Sci. Technol. 31 (1997) 2319-2325.

[62] S. Lee, J. Lee, T.C. Keener, Novel sorbents for mercury emissions control from coal-fired

power plants, Journal of the Chinese Institute of Chemical Engineers. 39 (2008) 137-142.

80

[63] C.A. Lockert, Q. Zhou, Y. Zhang and S. Nelson Jr., Further progress toward concrete-

friendlyTM mercury sorbents, Air Quality V Symposium, Washington, DC, September 19-21,

2005.

[64] S. Manchester, X. Wang, I. Kulaots, Y. Gao, R.H. Hurt, High capacity mercury adsorption

on freshly ozone-treated carbon surfaces, Carbon. 46 (2008) 518-524.

[65] Z. Mei, Z. Shen, Q. Zhao, W. Wang, Y. Zhang, Removal and recovery of gas-phase element

mercury by metal oxide-loaded activated carbon, Journal of Hazardous Materials. 152 (2008)

721-729.

[66] S. Lee, J. Lee, T.C. Keener, Bench-scale studies of in-duct mercury capture using cupric

chloride-impregnated carbons, Environ. Sci. Technol. 43 (2009) 2957-2962.

[67] S. Lee, J. Lee, T.C. Keener, The effect of methods of preparation on the performance of

cupric chloride-impregnated sorbents for the removal of mercury from flue gases, Fuel. 88 (2009)

2053-2056.

[68] J.A. Korpiel, R.D. Vidic, Effect of sulfur impregnation method on activated carbon uptake

of gas-phase mercury, Environ. Sci. Technol. 31 (1997) 2319-2325.

[69] X. Chen, M. Farber, Y. Gao, I. Kulaots, E.M. Suuberg, R.H. Hurt, Mechanisms of surfactant

adsorption on non-polar, air-oxidized and ozone-treated carbon surfaces, Carbon. 41 (2003)

1489-1500.

[70] Y. Gao, I. Külaots, X. Chen, R. Aggarwal, A. Mehta, E.M. Suuberg, et al., Ozonation for the

chemical modification of carbon surfaces in fly ash, Fuel. 80 (2001) 765-768.

[71] K.H. Pedersen, A.D. Jensen, M.S. Skjøth-Rasmussen, K. Dam-Johansen, A review of the

interference of carbon containing fly ash with air entrainment in concrete, Progress in Energy and

Combustion Science. 34 (2008) 135-154.

[72] K.H. Pedersen, M.C. Melia, A.D. Jensen, K. Dam-Johansen, Post-treatment of fly ash by

ozone in a fixed bed reactor, Energy and Fuels. 23 (2009) 280-285.

[73] K.H. Pedersen, A.D. Jensen, M. Berg, L.H. Olsen, K. Dam-Johansen, The effect of

combustion conditions in a full-scale low-NOx coal fired unit on fly ash properties for its

application in concrete mixtures, Fuel Process Technol. 90 (2009) 180-185.

[74] C.A. Wocken, Evaluation of Enviroscrub's multipollutant PahlmanTM process for mercury

removal at facility burning subbituminous coal, final plant report, 05-EERC-04-04, 2004.

[75] R.M. Boren, NETL/DOE mercury control technology R&D program review- mercury

removal results from two coal-fired utility boilers, 2004.

[76] E.S. Tavoulareas and W. Jozewicz, Multipollutant emission control technology options for

coal-fired power plants, EPA-600/R-05/034, 2005.

[77] R. Keeth, S. Peterson and M. Reed, Status of multi-pollutant process development:

LoTOxTM/BOC, GSATM/FLS, PahlmaniteTM, Cost projections, 1004245, 2003.

[78] J.Y. Lee, Y. Ju, T.C. Keener, R.S. Varma, Development of cost-effective noncarbon

sorbents for Hg0 removal from coal-fired power plants, Environ. Sci. Technol. 40 (2006) 2714-

2720.

81

[79] M. Holmes and J. Pavlish, Mercury information clearinghouse, Quarter 3- Advanced and

developmental mercury control technologies, July 2004.

[80] J.R. Butz, T.E. Broderick and C.S. Turchi, Amended Silicates™ for Mercury Control,

project final report, DOE Award Number: DE-FC26-04NT41988, 2006.

[81] ADA Technologies Inc, Arsenic removal technologies, http://www.adatech.com/default.asp,

accessed April 20, 2010.

[82] J.C. Hower, C.L. Senior, E.M. Suuberg, R.H. Hurt, J.L. Wilcox, E.S. Olson, Mercury

capture by native fly ash carbons in coal-fired power plants, Progress in Energy and Combustion

Science. 36 (2010) 510.

[83] D. Karatza, A. Lancia, D. Musmarra, Fly ash capture of mercuric chloride vapors from

exhaust combustion gas, Environ. Sci. Technol. 32 (1998) 3999-4004.

[84] Z. Li, X. Sun, J. Luo, J.Y. Hwang, J.C. Crittenden, Unburned carbon from fly ash for

mercury adsorption: II. Adsorption isotherms and mechanisms, Journal of Minerals & Materials

Characterization & Engineering. 1 (2002) 76.

[85] G.E. Dunham, R.A. DeWall, C.L. Senior, Fixed-bed studies of the interactions between

mercury and coal combustion fly ash, Fuel Processing Technology,. 82 (2003) 197-213.

[86] S.D. Serre, G.D. Silcox, Adsorption of elemental mercury on the residual carbon in coal fly

ash, Ind. Eng. Chem. Res. 39 (2000) 1723-1730.

[87] D.J. Hassett, K.E. Eylands, Mercury capture on coal combustion fly ash, Fuel. 78 (1999)

243-248.

[88] W.D. Owens, A.F. Sarofim, D.W. Pershing, The use of recycle for enhanced volatile metal

capture, Fuel Processing Technology. 39 (1994) 337-356.

[89] J. Dong, Z. Xu, S.M. Kuznicki, Mercury removal from flue gases by novel regenerable

magnetic nanocomposite sorbents, Environ. Sci. Technol. 43 (2009) 3266-3271.

[90] Lawrence E. Bool III, Jurron Bradley and David R. Thompson, A novel process for onsite

production of mercury sorbents, http://www.icac.com/files/public/Clearwater_2006_Bool_.pdf,

accessed June 30, 2008.

[91] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for DTE

Energy’s Monroe Station, DOE Award Number DE-FC26-03NT41986, Report Number

41986R16, 2006.

[92] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for

Sunflower Electric’s Holcomb Station, DE-FC26-03NT41986, Topical Report No. 41986R07,

2005.

[93] A.P. Jones, J.W. Hoffmann, D.N. Smith, T.J. Feeley, J.T. Murphy, DOE/NETL's Phase II

Mercury Control Technology Field Testing Program: Preliminary Economic Analysis of

Activated Carbon Injection, Environ. Sci. Technol. 41 (2007) 1365-1371.

[94] A.P. Jones, J.W. Hoffmann, D.N. Smith, T.J. Feeley and J.T. Murphy, DOE/NETL's Phase

II mercury control technology field test program: Updated economic analysis of activated carbon

injection, 2007.

82

[95] ADA-ES, Field test program to develop comprehensive design, operating, and cost data for

mercury control systems: Final site report for E.C. Gaston Unit 3 sorbent injection into COHPAC

for mercury control, DE-FC26-00NT41005, Topical Report No. 41005R11, 2003.

[96] J. Bustard, M. Durham, T. Starns, C. Lindsey, C. Martin, R. Schlager, K. Baldrey, Full-scale

evaluation of sorbent injection for mercury control on coal-fired power plants, Fuel Processing

Technology. 85 (2004) 549-562.

[97] J. Pavlish, Baghouse balance-of-plant effects at TXU’s Big Brown Station during field

testing of sorbent injection for Hg control, 10th Annual Electric Utilities Environmental

Conference, Tucson, Arizona, January 23, 2007.

[98] S. Derenne, P. Sartorelli, D. Johnson, R. Stewart, C.J. Bustad, S. Sjostrom, D. Muggli, T.

Starns, T. McMahon, F. Sudhoff, R. Chang, R. Utter, J. Cummings, TOXECON™ retrofit for

mercury and multi-pollutant control, 2006 Symposium on Western Fuels, Denver, 24-26 October,

2006.

[99] S.T. Derenne, TOXECON™ retrofit for mercury and multi-pollutant control on three 90-MW

coal fired boilers. Quarterly technical progress report, 1st January-31st March 2008, Report No.

41766R08, 2008.

[100] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for

AmerenUE’s Meramec Station Unit 2, DE-FC26-03NT41986, Topical Report No. 41986R09,

2005.

[101] ADA-ES, Field test program to develop comprehensive design, operating, and cost data for

mercury control systems: Final site report for Pleasant Prairie Power Plant Unit 2 sorbent

injection into a cold-side ESP for mercury control, DE-FC26-00NT41005, Topical Report No.

41005R12, 2003.

[102] ADA-ES, Field test program to develop comprehensive design, operating, and cost data for

mercury control systems: Final site report for Brayton Point Generating Station Unit 1 sorbent

injection into a cold-side ESP for mercury control, DE-FC26-00NT41005, Topical Report No.

41005R21, 2005.

[103] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for Basin

Electric Power Cooperative’s Laramie River Station, DE-FC26-03NT41986, Topical Report No.

41986R11, 2006.

[104] ADA-ES, Field test program to develop comprehensive design, operating, and cost data for

mercury control systems: Final site report for PG&E NEG Salem Harbor Station Unit 1 sorbent

injection into a cold-side ESP for mercury control, U.S. DOE Cooperative Agreement No. DE-

FC26-00NT41005, 2004.

[105] S. Sjostrom, M. Dillon, B. Donnelly, J. Bustard, G. Filippelli, R. Glesmann, T. Orscheln,

S.,Wahlert, R. Chang, A. O’Palko, Influence of SO3 on mercury removal with activated carbon:

Full-scale results, Fuel Process Technol. 90 (2009) 1419-1423.

[106] L. Hayes-Gorman, Regulating mercury emissions: Ash Grove Cement in Durkee, Air

toxics summit 2008, Boise, Idaho, 4-7 August, 2008.

83

[107] Schreiber & Yonley Associates, Mercury emissions test report, Ash Grove Cement

Company Durkee, Oregon, Project No. 060204, 2007.

[108] M.L. Jones, J.P. Kay, B.M. Pavlish and K.C. Galbreath, Mercury control slipstream

baghouse at Ash Grove’s Durkee cement facility, 2007.

[109] I.I. Salame, T.J. Bandosz, Surface chemistry of activated carbons: Combining the results of

temperature-programmed desorption, Boehm, and potentiometric titrations, Journal of Colloid

and Interface Science. 240 (2001) 252-258.

[110] J. Lahaye, The chemistry of carbon surfaces, Fuel. 77 (1998) 543-547.

[111] J.L. Figueiredo, M.F.R. Pereira, M.M.A. Freitas, J.J.M. Órfão, Modification of the surface

chemistry of activated carbons, Carbon. 37 (1999) 1379-1389.

[112] F.E. Huggins, N. Yap, G.P. Huffman, C.L. Senior, XAFS characterization of mercury

captured from combustion gases on sorbents at low temperatures, Fuel Processing Technology.

82 (2003) 167-196.

[113] N.D. Hutson, B.C. Attwood, K.G. Scheckel, XAS and XPS characterization of mercury

binding on brominated activated carbon, Environ. Sci. Technol. 41 (2007) 1747-1752.

[114] B. Podak, M. Brunetti, A. Lewis, J. Wilcox, Mercury binding on activated carbon, Environ.

Prog. 25 (2006) 319-326.

[115] E.S. Olson, S.J. Miller, R.K. Sharma, G.E. Dunham, S.A. Benson, Catalytic effects of

carbon sorbents for mercury capture, Journal of Hazardous Materials. 74 (2000) 61-79.

[116] G.E. Dunham, E.S. Olson and S.J. Miller, Impact of flue gas constituents on carbon

sorbents, McLean, VA, September 19-21, 2000.

[117] E.S. Olson, B.A. Mibeck, S.A. Benson, J.D. Laumb, C.R. Crocker, G.E. Dunham, et al.,

The mechanistic model for flue gas-mercury interactions on activated carbons: The oxidation site,

Prepr. Pap. -Am. Chem. Soc. , Div. Fuel Chem. 49 (2004) 279-280.

[118] Y. Zheng, A.D. Jensen, J.E. Johnsson, Deactivation of V2O5-WO3-TiO2 SCR catalyst at a

biomass-fired combined heat and power plant, Applied Catalysis B: Environmental. 60 (2005)

253-264.

[119] Y. Zheng, A.D. Jensen and J.E. Johnsson, DTU final report of EU project (NNE5-2001-

00164), Influences from Biofuel (Co-)Combustion on Catalytic Converters in Coal-fired Power

Plants, 2004.

[120] A. Yilmaz, L. Hindiyarti, A.D. Jensen, P. Glarborg, P. Marshall, Thermal dissociation of

SO3 at 1000-1400 K, Journal of Physical Chemistry a. 110 (2006) 6654-6659.

[121] J. Jarvis, F. Meserole, SO3 Effect on Mercury Control, Power Eng. 112 (2008) 54-60.

[122] A.A. Lizzio, J.A. DeBarr, Mechanism of SO2 removal by carbon, Energy & Fuels. 11

(1997) 284-291.

[123] F.H. Yang, R.T. Yang, Ab initio molecular orbital study of the mechanism of SO2

oxidation catalyzed by carbon, Carbon. 41 (2003) 2149-2158.

[124] D.L. Nickels and T. Weyand, Processing and reuse of activated carbon used to adsorb

mercury from power plant flue gases. Final report, Contract#: 2447-PMET-DOE-0350, 2004.

84

[125] T.C. Ho, P. Yang, T.H. Kuo, J.R. Hopper, Characteristics of mercury desorption from

sorbents at elevated temperatures, Waste Management. 18 (1998) 445-452.

[126] U.S. EPA, Land disposal restrictions for mercury-containing wastes,

http://www.epa.gov/epaoswer/hazwaste/mercury/regs.htm, accessed April 16, 2008.

[127] E.K. Levy, C. Kiely and Z. Yao, Investigation of fly ash and activated carbon obtained

from pulverized coal boilers, DE–FG26-03NT41796, Final report, 2006.

[128] T. Machalek, C. Richardson, J. Noblett, Field investigations of fixed-bed sorbents for

mercury capture from coal-fired flue gas, Power Plant Air Pollutant Control "MEGA"

Symposium 2008, Baltimore, MD, August 25-28, 2008.

[129] R. Change, G.R. Offen, Mercury emission control technologies: An EPRI synopsis, Power

Engineering. 99 (1995).

[130] M. Dίaz-Somoano, S. Unterberger, K.R.G. Hein, Mercury emission control in coal-fired

plants: The role of wet scrubbers, Fuel Processing Technology. 88 (2007) 259-263.

[131] C.L. Kairies, K.T. Schroeder, C.R. Cardone, Mercury in gypsum produced from flue gas

desulfurization, Fuel. 85 (2006) 2530-2536.

[132] S. Niksa, N. Fujiwara, The impact of wet flue gas desulfurization scrubbing on mercury

emissions from coal-fired power stations, J. Air Waste Manage. Assoc. 55 (2005) 970-977.

[133] P.S. Nolan, K.E. Redinger, G.T. Amrhein, G.A. Kudlac, Demonstration of additive use for

enhanced mercury emissions control in wet FGD systems, Fuel Processing Technology. 85 (2004)

587-600.

[134] A. Licata, W. Schuettenhelm and M. Klein, Mercury controls for MWCs using the sodium

tetrasulfide process, 8th Annual North American Waste-to-Energy Conference, Nashville, TN,

May 22, 2000.

[135] D.L. Laudal, J.S. Thompson, J.H. Pavlish, L. Brickett, P. Chu, R.K. Srivastava, J. Kilgroe,

C.W. Lee, Mercury speciation at power plants using SCR and SNCR control technologies, EM:

Air and Waste Management Association's Magazine for Environmental Managers. (2003) 16-22.

[136] Thomas J. Feeley, Lynn A. Brickett and James T. Murphy, Evaluation of the effect of SCR

NOx control technology on mercury speciation, http://www.netl.doe.gov/technologies/coalpower

/ewr/pubs/SCRHgPaperFinal030503.pdf, accessed September 9, 2010.

[137] H.G. Pedersen, L.S. Pedersen, H. Rostgaard and K. Pedersen, Oxidation of mercury on

DNX catalysts. Proceedings of the Air Quality V: Mercury, Trace Elements, SO3, and Particulate

Matter Conference, Arlington, VA, Sept 19–21, 2005.

[138] S. Eswaran, H.G. Stenger, Understanding mercury conversion in selective catalytic

reduction (SCR) catalysts, Energy & Fuels. 19 (2005) 2328-2334.

[139] Y. Cao, B. Chen, J. Wu, H. Cui, J. Smith, C.K. Chen, P. Chu, W. Pan, Study of mercury

oxidation by a selective catalytic reduction catalyst in a pilot-scale slipstream reactor at a utility

boiler burning bituminous coal, Energy & Fuels. 21 (2007) 145-156.

[140] H. Kamata, S. Ueno, T. Naito, A. Yukimura, Mercury oxidation over the V2O5(WO3)/TiO2

commercial SCR catalyst, Ind Eng Chem Res. 47 (2008) 8136-8141.

85

[141] Y. Cao, Z. Gao, J. Zhu, Q. Wang, Y. Huang, C. Chiu, B. Parker, P. Chu, W. Pan, Impacts

of halogen additions on mercury oxidation in a slipstream selective catalyst reduction (SCR)

reactor when burning sub-bituminous coal, Environ. Sci. Technol. 42 (2008) 256-261.

[142] Y. Zhuang, J. Laumb, R. Liggett, M. Holmes, J. Pavlish, Impacts of acid gases on mercury

oxidation across SCR catalyst, Fuel Processing Technology. 88 (2007) 929-934.

[143] S. Straube, T. Hahn, H. Koeser, Adsorption and oxidation of mercury in tail-end SCR-

DeNOx plants—Bench scale investigations and speciation experiments, Applied Catalysis B:

Environmental. 79 (2008) 286-295.

[144] K. Kai, Y. Nagai, Y. Kato, W. Gretta and H. Kikkawa, SCR catalyst with high mercury

oxidation and low SO2 to SO3 conversion, Baltimore, MD, August 28-31, 2006, .

[145] S.H. Jeon, Y. Eom, T.G. Lee, Photocatalytic oxidation of gas-phase elemental mercury by

nanotitanosilicate fibers, Chemosphere. 71 (2008) 969-974.

[146] Y. Li, C. Wu, Kinetic study for photocatalytic oxidation of elemental mercury on a SiO2-

TiO2 nanocomposite, Environ. Eng. Sci. 24 (2007) 3-12.

[147] Robert J. Schreiber, Christa O. Russell and Jeff Evers, Evaluation of suitability of selective

catalytic and selective non-catalytic reduction for use in Portland cement industry,

http://www.dep.state.pa.us/dep/DEPUTATE/airwaste/aq/transport/comments/CEMEX_Attachme

nt.pdf, accessed June 3, 2008.

[148] U. Leibacher, NOx elimination by the SCR process for cement kiln systems, ZKG

International. 49 (1996) 43.

[149] FLSmidth A/S, New patent-pending system for mercury control, http://ehighlights.

flsmidth.com/node/257, accessed June 30, 2009.

[150] P. Paone, Mercury reduction technologies in cement production, presented at 7th Colloquia

of managers and technicians of cement plants, Palacio de Congresos, Malaga, Spain, 17-19,

September 2009, .

[151] P. Paone, Mercury removal systems for the cement industry, presented at 2008 profit-

production technology seminar, Lehigh Country Club, Allentown, Pennsylvania, USA, 2 October,

2008.

[152] T.L. Mlakar, M. Horvat, T. Vuk, A. Stergaršek, J. Kotnik, J. Tratnik, F. Vesna, Mercury

species, mass flows and processes in a cement plant, Fuel. 89 (2010) 1936.

86

4

Experimental methods and materials

To provide a simple means for screening the performance of candidate sorbents

and derive mercury capture kinetics for promising sorbents in a mercury-laden simulated

cement kiln flue gas, a fixed-bed reactor system was designed and built in this project. In

this chapter the entire reactor setup will be described. Furthermore, the choice of some of

the core parts, i.e., the mercury vapor generator, humidifier for water vapor addition and

mercury analysis system are described in more details. Materials and methods applied in

this project are also presented.

4.1 Description of the fixed­bed reactor system 

Tests of the oxidized mercury converter and sorbents were conducted in a fixed-

bed reactor system as illustrated in figure 4.1. A photo of the system is shown in figure

4.2. Main equipments in the reactor system include a gas mixing system with water vapor

addition by a humidifier and mercury source in a calibration gas generator to simulate the

cement kiln flue gas, a low temperature oven with a glass reactor, mercury analysis

system, and mercury traps for exhaust gas treatment. To avoid mercury condensation and

accumulation in the system, all the gas lines before the analyzer are heated to 150C. All

temperatures including temperatures of heated lines, ovens, reactors, converter, and

analytical cell in mercury analyzer are sampled. The mercury source, reactor and hot

panel are located in a dedicated ventilation hood.

87

MFC

MFC

MFC

MFC

MFC

MFC

MFC

MFC

Flow meter

Vent Vent

Vent

N2

N2

CO2

O2

HCl

SO2

NOx

N2

Reactor

Analyzer

Evaporator

Hg source

Filter

Filter

Rotameter

Heat trace

Converter

Distributionbox

Air

Figure 4.1. Sketch of the fixed-bed reactor system for converter and sorbent tests.

Figure 4.2. Photo of the fixed-bed reactor system.

88

4.1.1 Gas mixing system 

The gas mixing system consists of valves and mass flow controllers for adding

different gases to simulate cement kiln flue gas. Gas addition includes carrier nitrogen to

the mercury source, carrier nitrogen to humidifier, CO2, O2, HCl, SO2, NO, NO2 and

balance nitrogen. The addition of the gases is controlled by the mass flow controllers and

the actual flow rate of each gas is measured by a bubble flow meter.

4.1.2 Mercury vapor addition system 

Mercury sources are added using a commercial calibration gas generator from

VICI Metronics, Dynacalibrator Model 150-06e-C with a gas flow capacity up to 750

ml/min. The Model 150 calibration gas generator is a constant temperature system

designed to generate precise ppm or ppb concentrations of chemical compounds in a gas

stream, using a permeation tube as the trace gas source. Figure 4.3 shows a picture and

sketch of the calibration gas generator. A passivated glass-coated permeation chamber

houses the permeation device, with measured inert carrier gas nitrogen sweeping the

calibration gas/vapor from the chamber. A digital temperature controller maintains the

chamber temperature at a set point with an accuracy of ±0.01°C and a wide range of

temperature settings (5°C above ambient to 110°C).

Figure 4.3. Left: picture of the calibration gas generator. Right: sketch of the calibration

gas generator [1].

89

The permeation tubes are small, inert capsules containing pure liquid elemental

mercury or solid mercury chloride in a two- phase equilibrium between its gas phase and

its liquid or solid phase, respectively. At a constant temperature, the device emits the

compound through its permeable portion at a constant rate. Figure 4.4 illustrates the

working principle of elemental mercury permeation tube.

Figure 4.4. Sketch of the permeation tube with elemental mercury [2].

The amount of mercury released from the tube is governed by the permeability of

the material used for the tube, the length of the tube, and the temperature at which the

tube is maintained. When the permeation rate at that temperature and the carrier flow rate

are known, the concentration of the calibration stream can be estimated. Table 4.1 shows

the specifications of the permeation tubes used in this project.

Table 4.1. Specifications of the elemental mercury and mercury chloride permeation

tubes used in this project.

Elemental mercury

Mercury chloride

Mercury chloride

Working temperature (C) 70 70 50 Tube diameter (mm) 9.8 9.8 9.8 Tube length (mm) 100 13 70 Release rate (ng/min) 378 2445 823

90

It is difficult to control the release rate of the mercury chloride tube to a similar

value as the elemental mercury tube. The originally supplied mercury chloride tube has a

release rate that is five times larger than the quoted tube at 70C. After several trials

VICI can provide a tube with a release rate of 823 ng/min at 50C.

4.1.3 Humidifier for water vapor addition 

The water vapor is not removed before mercury analyzer and mercury

concentration is therefore measured on a wet basis. Thus it is important to get a precise

control of the water addition.

Survey and quotation of water vapor addition methods and equipments were

carried out to get a reliable water vapor addition device. Current water addition methods

include direct liquid injection, bubblers, porous membrane contactor and non-porous

membrane contactor. The first three methods are already applied in CHEC. However,

recent studies show that fluctuation is a main problem [3,4]. Due to the small amount of

water injected in the direct injection method, it is difficult to control such small flow rate.

Water condensation test of the bubbler evaporator shows that the water saturation

fluctuates and cannot reach the calculated level [3,4]. This method is low cost, but has

inaccuracies due to the temperature of the gas and liquid, operating pressure, and liquid

level.

Porous membrane contactor uses Nafion selective permeable membrane tube and

water to continuously humidify gas streams. The producer suggests recirculation of the

water at 4% of the gas flow [5]. It is difficult to find such a small pump that works at

temperatures above 50C. Flow with greater pressure needs to be flowing inside tubes to

prevent tubing collapse. CHEC has made an evaporator using the membrane tube.

However, a lot of humidity fluctuation has been observed and the reasons have not been

identified [3,4]. Swedish company Cellkraft produces commercial evaporator using the

membrane tube [6]. The system uses similar design of CHEC’s membrane evaporator,

but has a water trap to remove water droplet. The water tank is heated from outside and

there is an integrated heating tape for the gas lines. The water level in the tank is

91

controlled and automatically filled. The producer can provide calibration curve and

guarantee for water droplet free and working properly at small carrier gas flow rate of

300 ml/min. The evaporator system without a dew point sensor costs about 64,000 DKK.

American company Rasirc produces a water vapor addition unit using the non-

porous membrane tube and integrated water temperature, level and dew point control

[7,8]. The unit purifies and controls water vapor addition for a wide range of flow rates

and process pressures. The membrane excludes particles, micro-droplet, volatile gases

and other opposite charged species and ensures only water vapor is added. Figure 4.5

illustrates the configuration of the Rasirc humidifier. Carrier gas to be humidified flows

into the humidification unit. The water is heated to match the desired dew point

temperature or humidification level. Water diffuses across the membrane to saturate the

gas to be humidified. Temperature of the humidified gas is measured and fed back to a

temperature controller to adjust the humidification level. Internal pressure control

maintains independence from variations in downstream process pressures which allows

operation into atmospheric and vacuum pressure environments. The unit with integrated

humidity sensor costs about 50,000 DKK. The Rasirc humidifier has been widely used in

the fabrication of semiconductors, nanotechnology, photovoltaics, fuel cells and other

applications [8]. After comparison, a Rasirc RHS-IP-3-HT humidifier with an internal

dew point sensor to regulate the dew point of the saturated gas was purchased.

92

Figure 4.5. Sketch of the Rasirc humidifier [8]. Internal dew point sensor is not shown in

the sketch.

4.1.4 Low temperature furnace and fixed­bed reactor 

The low temperature oven is a three-zone electrically heated furnace and can heat

up to 300C. The oven has an internal diameter of 50 mm and a length of 450 mm. The

heating tape for the top, middle and bottom zone is 170W/1m, 700W/4m, and 170W/1m,

respectively. Three thermocouples are installed to measure the temperature at each zone

and the heating is controlled by a center control box of the whole setup. The bottom of

the oven is closed and the glass reactor has a u-shape. To avoid losses of sorbent powder

in the gas stream, a downward flow is applied in the reactor. Quartz wool plugs are used

at both ends of the sorbent bed. The top of the reactor and oven is heated by a heating

tape.

The temperature profiles at different setpoints are measured, as shown in figure

4.6. The location of the sorbent bed is also illustrated in the figure. The measurements

confirm that an isothermal reactor zone of about 300 mm is obtained with an estimated

temperature uncertainty of ±2C to the setpoints.

93

0 50 100 150 200 250 300 350 400

Distance to oven bottom (mm)

0

50

100

150

200

250

300

Ov

en t

emp

erat

ure

(oC

)

Setpoint: 250 oCSetpoint: 200 oCSetpoint: 150 oCSetpoint: 120 oC Sorbent bed

position

Figure 4.6. Temperature profile of the low temperature reactor oven.

The glass reactor applied in this project is shown in figure 4.7. The reactor has an

outer diameter of 20 mm (internal diameter of 18 mm) and a glass fiber porous plate to

hold sorbent sample. With the dimension of the reactor shown in figure 4.7, the sorbent

bed is located in the middle height of the low temperature furnace.

Figure 4.7. Pictures with dimensions for the glass reactor.

4.1.5 Mercury analysis system 

The mercury analysis system consists of a Lumex RA-915 AMFG elemental

mercury analyzer, a gas distribution box and a oxidized mercury converter. Figure 4.8

illustrates the sketch of the analysis system. A photo of the analysis system is presented

in Figure 4.9.

94

Figure 4.8. Sketch of the mercury analysis system.

Figure 4.9. Picture of the mercury analysis system with box open. The oxidized mercury

converter is behind the mercury analyzer and gas distribution box.

95

4.1.5.1 The Lumex analyzer

The Lumex mercury analyzer has a measuring range of 0-500 g/Nm3 and

automatic zero and span calibration functions. The lower detection limit is about 2

g/Nm3. All the gas lines and analytical cell inside the analyzer are heated. The analyzer

can analyze gas of up to 30% water, and therefore no drying of the gas is needed. The

analyzer determines the mercury concentration by Zeeman atomic absorption

spectrometry using high frequency modulated polarized light. It is possible to measure

only elemental mercury bypassing the converter and only total mercury passing the gases

through the converter, and to change the frequency of elemental and total mercury

measurement switching through the sampling software.

A block diagram of the analyzer is shown in figure 4.10 and a photo the analyzer

internal parts are shown in figure 4.11. A membrane pump P draws flue gas from a

sampling point via heated lines through a gas distribution box. There are four valves in

the gas distribution box. The flue gas stream is either directed through the converter

which reduces oxidized mercury to elemental mercury (valve V3 opened, valve V4

closed) or is passed directly to the analytical cell AC, which is kept at a temperature of

about 150°C. In the cell AC, which has an optical path length of about 0.4 m, a

spectrometer determines the mercury concentration by Zeeman atomic absorption

spectrometry using high frequency modulated polarized light (ZAAS-HFM). After

leaving the cell, the gas is passing through a heated gas line and is then vented to a

carbon trap before the ventilation. Temperature of the cell is constantly monitored by

temperature sensors T. The whole unit is controlled by an industrial panel PC, and

powered by a power module PM.

96

Figure 4.10. Block diagram of the mercury analyzer. The gas distribution box and

converter are not integrated in the analyzer.

97

Figure 4.11. Photo of the analyzer internal parts.

The measurement principle of the analyzer is illustrated in figure 4.12 [9]. A

mercury electronic discharge lamp is placed in a strong magnetic field H, by which the

mercury resonance line at 254 nm is split into the three polarized Zeeman components -,

, and +. Only the -components of the electromagnetic radiation will be registered by

the photo detector D. - and + are separated by a polarization modulator. As long as

mercury vapor is absent in the multipath cell, the intensities of both -components are

equal. When mercury is admitted to the cell, the difference in intensities between the two

-components increases as a function of the mercury concentration. As the spectral shift

between the -components is significantly smaller than the widths of molecular

absorption bands and scattering spectra, background absorption by interfering

compounds can be neglected.

98

Figure 4.12. Illustration principle of the Zeeman atomic absorption spectrometry using

high frequency modulated polarized light (ZAAS-HFM) [9].

Sample gas connection to the analyzer is maintained at ambient pressure, with

any excess flow vented to the atmosphere. Heated inlet and outlet lines are connected to

the analytical cell inside the analyzer by means of 6 mm Swagelok-type fittings. The

analyzer requires between 1 and 12 l/min of sample gas at all times and the flow can be

controlled by the needle valve before the pump.

4.1.5.2 Gas distribution box

The gas distribution box contains valves for switching the gas and air to the

analyzer and converter. The valves are controlled by the sampling software. The box is

heat traced and isolated. There is a switch valve before the gas distribution box. Addition

of air or sample gas to the analysis system can be selected.

99

4.1.5.3 The oxidized mercury converters

Two converters are used in this project. Originally Lumex supplied a low

temperature converter for the red brass catalyst. Figure 4.13 shows a picture of the

converter with a glass container loaded with red brass chips. The converter is designed to

work at 180C and the highest temperature is about 250C. The glass container has an

outer diameter of 20 mm and can hold about 20 g red brass chips.

Figure 4.13. Picture of the low temperature converter with a glass container loaded with

20 g red brass chips.

Later a high temperature converter was used for the sulfite based converter

material. The high temperature oven is a three-zone electrically heated furnace with a

quartz reactor, which has an inner diameter of 17 mm and can hold up to 30 g sulfite-

based converter pellets. The converter is a fixed-bed reactor made of quartz as shown in

figure 4.14. The inner and bottom tubes of the reactor were removable. The sulfite-based

pellets are placed on the porous quartz plate. The converter temperature is measured

below the porous quartz plate by a thermocouple shielded in a quartz tube.

100

Figure 4.14. Sketch of the high temperature converter with quartz reactor in the furnace.

4.2 Converter and sorbent materials 

The red brass chips are obtained through Lumex. The idea of using red brass at

low temperature is to bind free halogens in the flue gas and thus prevent back reaction

into mercury halides and corrosion problem caused by SO2 oxidation at high

temperatures [10]. Figure 4.15 shows picture of the red brass chips which have a

thickness of about 0.5 mm and are rolled to a diameter of about 2 mm and a length of

about 10 mm.

101

Figure 4.15. Picture of the red brass chips supplied by Lumex.

The sulfite converter material is prepared according to the work of Akiyama et al.

[11]. Alumina pellets or zeolite pellets are first dried at 600C for 24 h. Then the pellets

are impregnated with water glass by forming a thin layer of water glass on the surfaces of

the pellets. Sodium sulfate powders are added and mixed with the impregnated pellets.

To inhibit crystallization of the salts, CaSO4 is added to the sulfite salts at a ratio of 50

wt.%. About 15 to 45 wt.% of the sulfite salts and CaSO4 mixture are adhered almost

uniformly to the thin layer of water glass. Immediately after mixing the product is placed

in an oven and vacuum-dried at room temperature for 1 h, then it is vacuum-dried at

150C for 12 h. Figure 4.16 shows the picture of the prepared sulfite-based converter

material. White powders of sodium sulfite and calcium sulfate are doped on the zeolite

pellets with a diameter of 3 mm.

102

Figure 4.16. Picture of the prepared sulfite-based converter material.

To be able to quantify the oxidized mercury reduction efficiency, the oxidized

mercury is produced by passing the flue gas with known concentration of elemental

mercury to the reactor with 4 g catalyst for selective catalytic reduction of NOx. The

catalyst piece was cut from a corrugated-type monolith obtained from Haldor Topsøe

A/S. The catalyst is based on a fiber reinforced titania (TiO2) carrier, which is

impregnated by vanadium (V2O5) and tungsten (WO3). The vanadium loading (3 wt.%

V2O5) was uniformly distributed across the wall thickness of the monolith [12,13]. The

efficiency of the converter is evaluated by the recovery extent of measured total mercury

through the SCR catalyst and converter compared to the elemental mercury level at the

inlet of the SCR catalyst.

The most investigated sorbents in this project is Darco Hg activated carbon,

which is a commercial lignite based powdered activated carbon and is developed for

heavy metal removal from incinerators and power plants. The Darco Hg carbon has a

bulk density of 0.51 g/cm3 and a surface area of about 600 m2/g. The average particle

103

size is 16 m and the porosity is about 58% [14-22]. Properties of other sorbents are

presented in the chapter of sorbent screening.

4.3 Flue gas composition 

The total flow rate through the reactor is 2.75 Nl/min of which about 2 Nl/min is

passed through the analyzer. The typical composition of the simulated cement kiln flue

gas applied in this work includes 21% CO2, 6% O2, 1% H2O, 10 ppmv HCl, 1000 ppmv

NO, 23 ppmv NO2, and 1000 ppmv SO2. The applied mercury concentration is about

160-180 µg/Nm3 by keeping the elemental mercury and mercury chloride source at 70ºC

and 50ºC, respectively, and using 0.275 Nl/min nitrogen as carrier gas. The water level in

the simulated flue gas is lower than real level in the cement kiln flue gas. This is due to

the limitation of the humidifier. Although the humidifier can add water vapor relatively

precisely, it is not robust. The membrane can be easily broken and the unit cannot stand

high over pressure. After short period of operation, the unit was repaired twice by

changing the membrane and installing of a pressure release valve. It seems that the unit

can run properly only for short period. Another reason is the fluctuation measurement of

the mercury analyzer with more than 5% water in the simulated flue gas. Therefore it is

decided to use 1% water in most of the tests by adding water through a bubbling bottle.

In few cases high water contents are used to cover a wide range of water level in the

simulated flue gas.

4.4 Sorbent load in fixed­bed test 

For the applied reactor in this project, at least 500 mg activated carbon is need to

form a fixed-bed covering the cross area of the reactor. The amount of the sorbent sample

is determined by the sample saturation time. Literature reported that approximately 600

mg of sorbent initially was placed into the reactor; however, the samples were reduced

from 600 mg to between 100 and 150 mg after it was observed that extremely long

durations (up to weeks) would be required to saturate the larger quantity of sorbent [23].

To avoid channeling the sorbent sample is usually mixed with some inert materials such

104

as sand and glass beads. Application of dilution by sand powder can also accelerate the

tests. The reaction gas flows downward through the bed to minimize the chance of

selective flow or channeling through the bed. Reactor sizes and sample dilutions applied

in the literature are reviewed and summarized in table 4.2.

Table 4.2. Reported reactor sizes, flow rates and sorbent sample loads in the literature.

Sorbent loading Reactor

size

ID (mm)

Flow rate

(Nl/min)

Superficial

velocity (cm/s)

@150C

References

50 mg fly ash mixed with 3 g

glass beads, 2.5 mm bed

thickness, additional 57.5 mm

glass beads upstream to the bed

for better flow distribution

35 3.2 8.7 [24]

0.2-0.6 g sample (copper

compound based sorbents and

commercial proprietary sorbents)

held by a glass wool plug, 16 mm

bed thickness

4 0.15 30.8 [25]

20-30 mg sorbent (Norit FGD

carbon and functionalized silica)

in 6 g silica, glass fiber filter at

two ends

12.7 0.91 18.4 [26]

0.61 g Darco G60 carbon with 3 g

glass beads, 4 mm bed length

35 2.8 7.5 [27]

20 mg activated carbon mixed

with 1 g sand

6.35 0.17 13.8 [28]

5 mg carbon on 3 g glass beads, 4

mm bed thickness

35 4.2 7.52 [29]

20 mg carbon in 10 g sand,

supported by quartz wool

12.7 1 20.3 [30]

10-100 mg sorbent mixed with 2

g sand, bed thickness of about 5

mm, quartz wool at two ends

18 2.75 22.1 This work

105

4.5 Experimental procedure 

An experimental procedure is developed for sorbent tests and measures are taken to

avoid mercury accumulation in the system. The detailed procedure is as following:

The mercury source is maintained at the operating temperature with carrier gas through all the time.

Increase the sulfite converter temperature setpoints from 100C to 500C.

Check the temperatures of mercury source, heated lines, low temperature oven,

gas distribution box and analyzer, set the low temperature reactor oven to desired

temperature.

Weight desired amount of sorbent and mix with 2g sand powder, load the sample

to the glass reactor.

Check and measure the flow rates of different gases.

When the converter temperature reaches 500C for about 30 min, change the

mercury analyzer measurement mode from elemental to total mercury

measurement.

Switch the valve before the glass reactor to bypassing the reactor position; add

gases except mercury to the system.

Add the gases to the mercury analyzer to check whether there is some mercury

accumulated in the hot system; if there is some mercury detected, wait the

mercury reading decreases to zero value and then switch mercury source to the

system.

Start the test and data sampling, make note in the sampling program, measure the

mercury inlet concentration for at least 30 min to ensure that the glass reactor is

heated for about 1 h and stable reactor temperature is obtained.

Switch the gases to fixed sorbent bed, make note in the sampling program.

After full mercury breakthrough is observed for 30 min, switch air to the analyzer

system while keeping simulated flue gas to the reactor, and switch the mercury

analyzer to measure elemental mercury.

106

When the elemental mercury measurement mode is ready, switch the simulated

flue gas to the analysis system.

After 20 min, switch the simulated flue gas to bypass the reactor and measure the

inlet mercury concentration for another 20 min.

Stop the test, switch air to the analysis system, switch mercury source to the

carbon trap and ventilation, and stop other gases.

Remove the reactor from the oven and remove the sample after the reactor is

cooled, store the sample in a closed plastic bottle with label.

At the end of the day, turn all gases off except nitrogen and H2O when elemental

mercury source is applied and add also HCl when mercury chloride is applied to

flush the system overnight. Decrease the converter temperature to 100C, to

ensure the mercury analyzer is in elemental mercury measurement mode and air

is added to the analysis system.

Always keep the whole reactor system hot.

4.6 Sorbent characterization 

4.6.1 Scanning electron microscopy 

Scanning electron microscopy with energy-dispersive X-ray spectroscopy (SEM-

EDX) analysis is used to understand mercury capture mechanisms by different powder

sorbent. The main goals of the SEM-EDX analysis is to study the sorbents’ topography

(surface features), morphology (shape and size), and composition. Morphology study

will be used to identify particle agglomeration and compare with particle size

measurement.

The SEM-EDX analysis is conducted at Center for Electron Nanoscopy, DTU.

Micrographs and EDX analysis of carbon samples are carried out using Quanta FEGSEM

200F. The carbon samples are not coated, while the non-carbon samples are coated with

14 nm carbon and analyzed on Inspect ‘S’ SEM. The FEGSEM is a high resolution

flexible microscope with field emission gun (FEG). The Inspect ‘S’ is a scanning

electron microscope with a tungsten filament electron source. To support the surface

107

information obtained by the imaging detectors both SEMs are equipped with Oxford

Instruments INCA EDX analyzer which gives possibilities to analyze chemical elements

position on the sample surface in single spots or over a selected area. The microscope can

operate in as well high- and low vacuum as in environmental mode. A typical working

distance of 10 mm is applied.

A thin layer of sample powders is spread on a double sided conductive carbon

table. If particles are piled on each other charge-up easily takes place, causing them to

move during observation. A low accelerating voltage of 5 kV is applied during imaging

on the FEGSEM to obtain detailed information on the particle surface and minimize

specimen charging problem.

4.6.2 Particle size distribution 

The particle size distributions of the sorbent powders are analyzed by a Malvern

Mastersizer S analyzer using laser diffraction. The technique of laser diffraction is based

around the principle that particles passing through a laser beam will scatter light at an

angle that is directly related to their size. Large particles scatter light at narrow angles

with high intensity [31], whereas small particles scatter at wider angles but with low

intensity. The analyzer consists of a laser to provide a source of coherent, intense light of

fixed wavelength, a sample presentation system to ensure that the material under test

passes through the laser beam as a homogeneous stream of particles in a known,

reproducible state of dispersion, and a series of detectors which are used to measure the

light pattern produced over a wide range of angles.

Based on previous analysis experience at CHEC, the samples are dispersed either

in ethanol or distilled water for one minute before measurement to avoid agglomeration

and the result is the average of 5 measurements.

108

4.6.3 Analysis of mercury in sorbent

Mercury content in the exposed sorbent is analyzed by a DMA-80 analyzer from

Milestone at FLSmidth Dania lab. 100 mg of powder sample is first weighted and loaded

in a sample boat. The boats are transported automatically into the furnace. The sample is

initially dried and then thermally decomposed in a continuous flow of oxygen.

Combustion products are carried off and further decomposed in a hot catalyst bed [32].

Mercury vapors are trapped on a gold amalgamator and subsequently desorbed for

quantization. The mercury content is determined using an atomic absorption

spectrophotometer at 254 nm. The instrument determines the absolute amount of Hg and

then the software calculates its concentration in the sample.

4.7 References  

[1] VICI Metronics Inc., Dynacalibrator® Model 150 calibration gas generator brochure, 2008.

[2] VICI Metronics Inc., Dynacal permeation tubes, 2011.

[3] B. Maribo-Mogensen and J. Christensen, Internal steam reforming in solid oxide fuel cells,

Bachelor, Department of Chemical and Biochemical Engineering, Technical University of

Denmark, 2008.

[4] A.F. Castells. Steam reforming kinetics over Ni-YSZ used as anode material for solid fuel

cells, Master, Department of Chemical and Biochemical Engineering, Technical University of

Denmark, 2009.

[5] Perma Pure, MHTM-series humidifier user manual, http://www.permapure.com /PDF%20Files

/MH%20Manual.pdf, accessed March/20, 2009.

[6] Cellkraft AB, P-series humidifier manual, http://www.cellkraft.se/humidity_and_steam/P-

Series.html, accessed March 20, 2009.

[7] RASIRC, RASIRC RainmakerTM humidification system manual, http://www.rasirc.com

/resources/datasheets/datasheet_RASIRC_RainMaker_HS.pdf, accessed March 20, 2009.

[8] Jeffrey Spiegelman, RainMaker humidification system for precise delivery of water vapor

into atmospheric and vacuum applications, http://www.rasirc.com/resources/whitepapers

/whitepaper_RHS.pdf, accessed March 20, 2009.

[9] Lumex Ltd, RA-915 AMFG automatic mercury monitor for flue gas operational manual,

2009.

[10] R. Kanefke, H. Köser, B. Vosteen, F. Kristina, B. Frank, S. Raik, Method for the production

of elemental mercury from mercury compounds, patent WO 2008/064667 A2, 2008.

109

[11] S. Akiyama, J. Kato, F. Koga, K. Ishikawa, Catalysts for reducing mercury, a mercury

conversion unit, and an apparatus for measuring total mercury in combustion exhaust gas by

using the same, patent US2007/0232488 A1, 2007.

[12] Y. Zheng, A.D. Jensen, J.E. Johnsson, Deactivation of V2O5-WO3-TiO2 SCR catalyst at a

biomass-fired combined heat and power plant, Applied Catalysis B: Environmental. 60 (2005)

253-264.

[13] Y. Zheng, A.D. Jensen, J.E. Johnsson, J.R. Thøgersen, Deactivation of V2O5-WO3-TiO2

SCR catalyst at biomass fired power plants: Elucidation of mechanisms by lab- and pilot-scale

experiments, Applied Catalysis B: Environmental. 83 (2008) 186-194.

[14] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R. Chang,

Preparation and evaluation of coal-derived activated carbons for removal of mercury vapor from

simulated coal combustion flue gases, Energy & Fuels. 12 (1998) 1061-1070.

[15] M. Rostam-Abadi, S.G. Chen, H.C. Hsi, M. Rood, R. Chang, T.R. Carey, B. Hargrove, C.

Richardson, W. Rosenhoover, F. Meserole, Novel vapor phase mercury sorbents, Proceedings of

the EPRI-DOE-EPA Combined Utility Air Pollutant Control, Washington, DC, Aug 25–29,1997.

[16] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang and F.B. Meserole, Factors

affecting mercury control in utility flue gas using sorbent injection, Proceedings of the Air &

Waste Management Association's 90th Annual Meeting, Toronto, Ontario, Canada, June 8-13,

1997.

[17] B. Ghorishi and B.K. Gullett, Fixed-bed control of mercury: Role of acid gases and a

comparison between carbon-based, calcium-based, and coal fly ash sorbents, Proceedings of the

EPRI/DOE/EPA Combined Utility Air Pollutant Control Symposium, Washington, DC, August

25–29,1997.

[18] S.B. Ghorishi and C.B. Sedman, Combined mercury and sulfur oxides control using

calcium-based sorbents, Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant

Control Symposium, Washington, DC, August 25–29, 1997.

[19] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption by activated

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy Conference,

Research Triangle Park, NC, 22-25 April, 1997.

[20] N. Hutson, C. Singer, C. Richardson, J. Karwowski and C. Sedman, Practical applications

from observations of mercury oxidation and binding mechanisms, EPRI-DOE-EPA-AWMA

Combined Power Plant Air Pollutant Control Mega Symposium, Washington DC, August 30 -

September 2, 2004.

[21] Norit Americas Inc., Datasheet of Darco FGD powdered activated carbon, 2008.

[22] R. Bhardwaj. Impact of temperature and flue gas components on mercury speciation and

uptake by activated carbon sorbents. Master thesis. University of Pittsburgh, 2007.

[23] D.J. Hassett, K.E. Eylands, Mercury capture on coal combustion fly ash, Fuel. 78 (1999)

243-248.

110

[24] D. Karatza, A. Lancia, D. Musmarra, Fly ash capture of mercuric chloride vapors from

exhaust combustion gas, Environ. Sci. Technol. 32 (1998) 3999-4004.

[25] J.W. Portzer, J.R. Albritton, C.C. Allen, R.P. Gupta, Development of novel sorbents for

mercury control at elevated temperatures in coal-derived syngas: results of initial screening of

candidate materials, Fuel Processing Technology. 85 (2004) 621-630.

[26] J.Y. Lee, Y. Ju, T.C. Keener, R.S. Varma, Development of cost-effective noncarbon

sorbents for Hg0 removal from coal-fired power plants, Environ. Sci. Technol. 40 (2006) 2714-

2720.

[27] D. Karata, A. Lancia, D. Musmarra, F. Pepe, Adsorption of metallic mercury on activated

carbon, Symposium (International) on Combustion. 26 (1996) 2439-2445.

[28] G. Skodras, I. Diamantopoulou, G. Pantoleontos, G.P. Sakellaropoulos, Kinetic studies of

elemental mercury adsorption in activated carbon fixed bed reactor, Journal of Hazardous

Materials. 158 (2008) 1-13.

[29] D. Karatza, A. Lancia, D. Musmarra, C. Zucchini, Study of mercury absorption and

desorption on sulfur impregnated carbon, Experimental Thermal and Fluid Science. 21 (2000)

150-155.

[30] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors affecting

mercury control in utility flue gas using activated carbon, Journal of the Air & Waste

Management Association. 48 (1998) 1166.

[31] Malvern Instruments Ltd, Understanding how laser diffraction works,

http://www.malvern.com/LabEng/technology/laser_diffraction/laser_diffraction.htm, accessed

January/3, 2011.

[32] Milestone Srl, DMA-80: Principle of operation, http://www.milestonesrl.com

/analytical/products-mercury-determination-dma-80-and-dma-803-principle-of-operation.html,

accessed January 6, 2011.

[33] A. Ocklind. Calculation from gas dewpoint to water content, 2009.

[34] R.H. Perry, D.W. Green, J.O. Maloney, (Eds.), Perry’s chemical engineers’ handbook, 7th

ed. The McGraw-Hill Companies, Inc., 1997.

Appendix 

4A Check of mercury analyzer 

Since many valves and fittings are used in the mercury analysis system, it is

important to make a leakage test of the system. Lumex service technicians performed the

leakage test. The Lumex analyzer cannot stand high pressure and the system therefore

cannot be tested by plugging the system and checking the pressure change. Instead, flow

rates at the gas distribution box, converter, and analyzer are measured simultaneously

111

using rotating flow meters and a separate pump from Lumex. Flow rates measured by the

portable rotating flow meters at the analysis system inlet and outlet and flow rate

measured by the integrated flow meter in the analyzer are the same at both cold and hot

condition, indicating that there is no leakage in the Lumex analysis system. The function

of heating the gas panel and lines to avoid mercury accumulation in the system was also

checked. The test was conducted by first passing mercury contained gas through the hot

panel and then stopping mercury addition and the hot panel was flushed by nitrogen. No

mercury was detected in the flushing nitrogen, confirming that no mercury was

accumulated in the lines.

The required flow rate for the mercury analyzer was determined. The needle

valve for controlling the flow rate was changed from bypass line between the analytical

cell and the pump to on the line between the analytical cell and the pump. The flow rate

is better controlled; however, it cannot be reduced to as low as l l/min. The flow rate

through the analyzer is set to 2 l/min. The analyzer requires overflow. Therefore a 3

l/min gas through the reactor is used and the excess flow will bypass the analyzer and

exhaust through the ventilation.

Lumex service technician brought a portable mercury analyzer RA-915+, which

is the same as the Haldor Topsøe analyzer except that it has only a single analytical cell.

These two analyzers were compared by making tests at Haldor Topsøe’s mercury

research facility. Using a single analytical cell at analyzers, the measured mercury

concentration by Lumex and Haldor Topsøe analyzer was 9936 and 10080 ng/m3,

respectively. Mercury concentration measured by Haldor Topsøe analyzer using multiple

cells was 10200 ng/m3. This indicates that the Lumex portable analyzer works properly

and seems reliable.

Comparison of Lumex AMFG monitor, which is used in present project, with

Lumex portable analyzer was conducted by running these analyzers simultaneously. The

measured mercury concentrations by the portable analyzer are about 10% higher than the

AMFG, as shown in table 4.3.

112

Table 4.3. Comparison of mercury measurement by the portable Lumex RA-915+ and

the Lumex AMFG analyzer applied in this project.

Concentration measured by portable analyzer, g/Nm3

Concentration measured by AMFG analyzer, g/Nm3

Ratio of portable/AMFG

249 220 1.13 217 201 1.08 153 137 1.11

Lumex said that the difference is caused by different conditions used during

calibration at the factory and test at CHEC lab. The gas lines before the analyzer at the

factory was not heated and the analytical cell was tested at a temperature which was

50C lower than at CHEC lab. Since comparison with Haldor Topsøe analyzer shows

that the portable analyzer works properly, the AMFG is calibrated using linearity test. A

factor of 1.10 was used as the calibration coefficient.

4B Water addition verification 

To verify the water addition stability and accuracy condensation tests are

conducted. The configuration of the test system is shown in figure 4.17. A water column

of a height of about 500 mm is hanged 1 m above the humidifier as water source. The top

of the water column is open. A stop valve is installed below the water column to allow

disconnection of the water column for weight measurement. Nitrogen is used as the

carrier gas. The water level in the humidifier is always kept at full level by a liquid level

switch and a micro pump. The gas line between the humidifier and the condensation

water bottle is heated at 110C to avoid water condensation in the gas line. The water

bath temperature is kept at 4C. The condensation bottle is filled with some water to

enhance the heat exchange and capture of water. It takes about 70 min for the humidifier

to reach the desired dew point. During this period the gas is bypassed to the water bath

and passes through a water bottle before ventilated. When the desired dew point is

reached and stabilized, the water tank is disconnected for weight measurement. The gas

is still running through the humidifier and water is continuously added. During the

113

measuring of the water tank weight about 80 mm water inside the ¼’’ Teflon tube is

added and the corresponding weight is about 1 g. When the water tank is connected back

there will be an 80 mm air plug inside the Teflon tube. This means that 1 g water is not

added and should be deducted from the theoretical calculation for comparison with the

measured water addition. The amount of water added can be evaluated either by

measuring weight change of the water tank or weight of water collected in the

condensation bottles.

Figure 4.17. Sketch of the condensation test system.

When the dew point is reached and stabilized, the dew point reading of the

saturated gas is changing about 0.1C from the set point. This indicates that the relative

humidity of the gas is quite stable and close to 100%.

To calculate the amount of water added the saturated water vapor pressure at

given dew point should be calculated first. Table and figure of saturated water vapor

pressure can be readily found in text book. The saturated water vapor pressure can be

calculated by using the empirical expression from Cellkraft, which is a Swedish

membrane humidifier produce [6,33]:

114

3 5 2 7 3 9 4 13 5

16 6 18 7 22 8 25 9

10.4592 4.05 10 4.18 10 3.69 10 1.02 10 8.65 10

9.04 10 2.00 10 7.79 10 1.91 10 3968.06 /( 39.57)

slnP T T T T T

T T T T T

(4.1) where the water saturated water vapor pressure Ps is in MPa, T is in K.

Partial pressures of the gases are proportional to the gas flow rates:

2 2

2 2

H O H O

N N

P F

P F (4.2)

2 2tot H O NP P P (4.3)

where T is the desired dew point of the gas, 82 C, 355.15 K 3 5 2 7 3

9 4 13 5 16 6 18 7

22 8 25 9

10.4592 4.05 10 355.15 4.18 10 355.15 3.69 10 355.15

1.02 10 355.15 8.65 10 355.15 9.04 10 355.15 2.00 10 355.15

7.79 10 355.15 1.91 10 355.15 3968.06 /( 355.15 39.5

slnP

7) 2.9686 (4.4) and 513.76sP mbar

With room temperature of 25C, carrier nitrogen flow, 0.3l/min, equals 0.275 Nl/min,

one can calculate:

Water addition rate (g/min):

(18 / 22.4) 0.275 (513.76 /1013.25)0.227

1 513.76 /1013.25

Water flow rate (Nl/min):

0.275 (513.76 /1013.25)0.283

1 513.76 /1013.25

The water vapor pressure can also be calculated using Antoine equation [34]:

10 8.07131 1730.63/(233.426 )sPlog T (4.5)

where Ps is in Torr (1 mmHg), T is in C.

10 8.07131 1730.63 /(233.426 82) 2.5847sPlog

then 512.35sP mbar

Calculated water addition rate (g/min):

(18 / 22.4) 0.275 (512.35 /1013.25)0.226

1 513.76 /1013.25

Water flow rate (Nl/min):

115

0.275 (512.35 /1013.25)0.281

1 513.76 /1013.25

Table 4.4 presents the calculated water addition rates and flow rates at different

dew points. Calculations using Cellkraft equation and Antoine equation give almost the

same results.

Table 4.4. Calculated water addition rate and flow rate at different dew points. Cellkraft equation Antoine equation Dew point (C)

H2O addition rate (g/min)

H2O flow rate (Nl/min)

H2O addition rate (g/min)

H2O flow rate (Nl/min)

82 0.227 0.283 0.226 0.281 77 0.156 0.194 0.155 0.193 72 0.112 0.139 0.111 0.138 63 0.064 0.080 0.064 0.080 4 0.002 0.002 0.002 0.002

The comparison between the measured water addition and calculated water

addition is presented in table 4.5. For the first two tests only one condensation bottle is

used and about 1/3 of the bottle is filled with water. The water collected in the

condensation bottle is only 50-60% of the calculated value. This is probably due to the

short gas residence time in the condensation bottle and small amount of water filled in

the condensation bottle. The amount of water added by water tank weight measurement

is reasonably in agreement with the calculation. Later more water is filled in the bottle

and two or four bottles are connected in series. Then the amounts of water added by

measuring the water tank weight change and water collected in the bottle are similar and

about 90-95% of the calculated values. For the four bottle in-series tests the weight

change of each bottle is measured. Almost all the water is collected in the first bottle and

the water collected in other bottles is negligible. The test results show that the humidifier

works well.

116

Table 4.5. Comparison between the measured water addition and calculated water

addition.

Dew point (C) 82 82 82 82 82 82 N2 flow (ml/min) 300 300 300 300 300 400 Room temp. (C) 24.9 24.7 26.3 25.7 24.9 24.9 Duration (min) 450 318 368.5 240 143 120 Calculated water addition (g) 100.25 70.55 82.17 52.52 28.89 35.12 Measured water addition through tank weight (g)

102.4 63.5 74.10 47.7 27.6 31.8

Measured/calculated (%) 102 90 90.5 90.8 95.5 90.6 Measured water addition through collection in water bottle (g)

61.48 36.27 73.99 47.08 - 32.32

Water bottle number 1 1 2 2 4 2 Measured/calculated (%) 61.4 51.4 90.3 89.6 - 92

117

5

Dynamic measurement of mercury adsorption and

oxidation on activated carbon in simulated cement

kiln flue gas

This chapter starts with a review of available gaseous mercury measurement

technologies. Pros and cons of the technologies will be discussed. Then tests of the

commercial red brass converter in simulated cement kiln flue gas are presented. Finally

development of sulfite-based converter for oxidized mercury reduction in simulated

cement kiln flue gas is reported. Suggestions for practical applications of the sulfite

converter in both lab and cement plants are presented.

5.1 Review of gaseous mercury measurement technology  

Presently, the accepted methods for mercury measurement are wet-chemistry

procedures such as EPA methods 29 and 101 A for total mercury measurement and the

Ontario Hydro method for total mercury and speciation measurement [1,2]. These

methods often have 2-week or more turn-around time for results. The sorbent trap

method was developed to shorten the analysis time of the collected samples. These

methods can only provide an average mercury concentration over a 1-2 hour period, and

cannot characterize the variability in mercury emissions due to process and operating

changes with time.

To obtain an understanding of the process of mercury removal by sorbent

injection upstream of a fabric filter, it is necessary to study them under more controlled

conditions such as in a laboratory scale setup, for example using a fixed bed reactor. In

such experiments it is also necessary to use a continuous emission monitor (CEM) to

118

obtain knowledge of uptake of total and speciated mercury in simulated flue gas to fully

evaluate the control technologies under development. Fixed-bed experiments have been

used by many laboratories to test the relative effectiveness of different mercury sorbents.

The critical assumption of this experimental method is that the performance of a sorbent

over a long exposure time (hours) reflects the filtration/reaction on bags, where the

sorbent contacts the flue gas for about 25 minutes [1]. Therefore, it is preferable to verify

this assumption by supplementing the final mercury content data with breakthrough data

obtained using a CEM.

Real- or near-real-time mercury emission measurement can in principle be

obtained depending on the applied detection method. Generally, real-time measurements

can be achieved by analyzers using cold vapor atomic absorption spectroscopy. The cold

vapor atomic fluorescence spectrophotometer collects mercury in flue gas on alternating

gold traps and thermally desorbs the mercury in about five minute intervals allowing for

semi-continuous measurements [3].

Available commercial mercury analyzers can only measure elemental mercury.

The measurement of total mercury as well as mercury speciation can only be achieved

indirectly. For this purpose, all oxidized mercury is reduced to its elemental form by a

converter system. It should be noted that the technology for the analytical part of the

detection system is somehow matured and provides accurate and sensitive detection of

elemental mercury [3]. The conversion unit, on the other hand, is a subject of continuing

research and improvement efforts [2].

The converter can be based either on wet chemistry or dry conversion. In a wet-

chemistry conversion unit the Hg2+ is converted to Hg0 via a liquid phase reducing agent,

often stannous chloride (SnCl2), prior to entering the analysis unit. There is interference

with SO2, which can affect the reduction of Hg2+ when using SnCl2 [2,4,5]. Furthermore,

the wet chemicals themselves are very corrosive and need frequent replenishment.

For on-line measurements, a dry converter is usually preferred over a wet

chemical converter for the reasons mentioned above [6]. Several dry converter types

exist. In a pure thermal conversion unit, the flue gas is heated to reduce all Hg2+ to Hg0.

119

However, reoxidation of the reduced mercury before reaching the analysis unit is a

concern. Furthermore, the required temperature depends on the HCl concentration in the

gas. In case of a thermocatalytic conversion, the potential short lifetime of the catalyst is

an issue due to the possible poisoning by acidic gases in the sample gas [2,4,5].

Compared to mercury measurements in power plants and waste incinerators, there

is a lack of experience related to continuous measurement of mercury emissions from

cement kilns. Furthermore, the experience gained from power plant and waste incinerator

may not be applied directly to cement plant due to the different process conditions and

flue gas compositions [7]. In this work a commercial red brass converter, which is

developed for application in waste incinerators, is tested in simulated cement kiln flue

gas and an improved sodium sulfite-based converter is developed and tested.

5.2 Performance test of the mercury analyzer 

The analyzer has an internal mercury source for span calibration. However, the

span calibration is conducted at an elemental mercury concentration of about 16 g/Nm3,

which is much lower than typical mercury concentration of about 180 g/Nm3 applied in

this project. If the linearity of the analyzer is poor then the measured mercury

concentration at typical mercury levels in this project could be wrong. To check the

analyzer linearity some tests were conducted. The carrier nitrogen flow rate through the

mercury source was kept at 275 Nml/min. Firstly, the mercury concentration in the outlet

gas from the mercury source was calculated using the measured mercury concentration in

the mixed gas and applied flow rates. Then part of the gas from the outlet of mercury

source was bypassed to ventilation and more nitrogen was added to the empty reactor to

dilute the mercury-contained gas and keep the total flow through the reactor at 2.75

Nl/min. Figure 5.1 shows the comparison between the measured and calculated mercury

concentration in the mixed gas. Very good agreement is obtained between the measured

and calculated mercury concentration, confirming that the linearity of the analyzer is

good.

120

0 50 100 150 200 250

Calculated Hg concentration (g/Nm3)

0

50

100

150

200

250

Me

asu

reed

Hg

co

nce

ntr

atio

n (g

/Nm

3 )

Figure 5.1. Linearity of the Lumex mercury analyzer. Measured elemental mercury

concentration is compared with the calculated values in the range of 0-250 g/Nm3.

The effects of different gases on elemental mercury measurement were

investigated by adding gases separately. Figure 5.2 shows the measured mercury

concentration under different conditions. Nitrogen, water, and CO2 were used as baseline

gas. Further addition of O2, SO2, NOx and HCl step by step to the baseline gases gives

the same mercury level. This indicates that these gases at the applied level do not have

influence on elemental mercury measurement. The consistent mercury concentration also

implies no mercury oxidation in the lines.

121

0 10 20 30 40 50 60 70

Time (min)

-40

0

40

80

120

160

200

240H

g c

on

cen

trat

ion

(g

/Nm

3 )Baselinegas

air to analyzer

Baselinegas,O2

Baselinegas,SO2

Baselinegas,NOX

Baselinegas,HCl

Baseline gasO2, SO2, NOX,HCl

Figure 5.2. Effects of different gases on elemental mercury measurement bypassing the

converter.

5.3 Performance test of the red brass converter 

The red brass chips are obtained through the analyzer supplier Lumex. The

typical composition of red brass includes 85% Cu, 5% Sn, 5% Zn and 5% Pb [8]. The

idea of using red brass at low temperature is to bind free halogens in the flue gas and thus

prevent back reaction into mercury halides as illustrated in following reaction [9]:

2 2Cl Cu CuCl (5R1)

The red brass converter is designed to convert Hg2+ to Hg0 at low temperatures of

120-250C to minimize the corrosion problem caused by SO2 oxidation at high

temperatures [9]. The principle of the converter is to convert oxidized mercury according

to following reaction:

2 2HgCl Me Hg MeCl (5R2)

where Me could be Cu, Sn, Zn, and Pb that are contained in the red brass.

122

The performance of the red brass converter on elemental mercury measurement

was first investigated. Test of the converter at 180C in nitrogen atmosphere with only

elemental mercury shows that the converter works well, since measurements through and

bypass of the converter give the same mercury concentrations and the response time is

short. However, tests of the converter using simulated flue gas and elemental mercury

show that the performance of the converter degrades as a function of time. After short

term exposure to the simulated flue gas the measured mercury level through the

converter starts to decrease and is lower than that measured bypassing the converter.

Detailed investigations were then conducted to study the possible effects of gases

on Hg0 measurement through the converter. The applied gas concentrations are: 15 ppmv

HCl, 1000 ppmv NO, 30 ppmv NO2, 1000 ppmv SO2, 1% H2O. Figure 5.3 shows the

measured Hg0 through the converter after adding different gases. When HCl, SO2 and

NOx is added alone with water, the measured Hg0 after the converter are the same as the

inlet. However, when HCl is added either with SO2 or NOx the measured Hg0 through the

converter decreases with time, indicating that the catalyst surface is modified and starts

to adsorb mercury or oxidize it to HgCl2.

123

0 60 120 180 240 300 360 420 480

Time (min)

0

50

100

150

200

250

300

Hg

co

nce

ntr

atio

n (g

/Nm

3 ) N2+Hg+H2O

N2+Hg+H2O+NOx

N2+Hg+H2O+HCl N2+Hg+H2O+SO2

N2+Hg+H2O+HCl+SO2

N2+Hg+H2O+HCl+SO2

+NOx

N2+Hg+H2O+HCl+NOx

N2+Hg+H2O+SO2+NOx

Figure 5.3. Measured elemental mercury concentration through the converter with 20 g

red brass chips at 180C after adding different gases.

Besides reaction with mercury chloride, copper in the red brass can also react

with other gases and form oxidized copper compounds. Possible reactions include:

22 2Cu O CuO (5R3)

2 22CuO HCl CuCl H O (5R4)

2 2 42 2 2CuO SO O CuSO (5R5)

Similar reactions could also take place for metals such as Sn, Zn and Pb contained in the

red brass. It has been reported that NO2 is a very good oxidizing agent for preparing ZnO

from metallic zinc through the reaction [10]

2NO Zn NO ZnO (5R6)

A similar reaction might take place between NO2 and copper. Copper chloride and

copper sulfate have been used as promoters to improve mercury oxidation and adsorption

by different sorbents [11-15]. These possible reactions might explain why elemental

124

mercury adsorption and oxidation takes place on the red brass chips in the simulated

cement kiln flue gas.

The oxidized mercury is added and produced by passing gases to the reactor with

4 g SCR catalyst plate at 150C. Oxidation of Hg0 by the SCR system has been reported

in both power plants[16,17] and in bench-scale tests [18-21]. The oxidation of mercury

by the SCR catalyst is fast and about 70% mercury oxidation is obtained when 4g SCR

catalyst is exposed to the simulated flue gas with15 ppmv HCl, 1% H2O, 1000 ppmv NO,

30 ppmv NO2 and 1000 ppmv SO2 at 150C. Figure 5.4 shows the result for using only

15 ppmv HCl, 1% H2O and with N2 as balance. It takes about 5 h for the converter to

obtain full oxidized mercury reduction, indicating that red brass converter cannot be used

for dynamic measurement.

0 2 4 6 8 10 12

Time (hour)

0

40

80

120

160

200

240

Hg

co

nce

ntr

atio

n (g

/Nm

3 )

Hg0, bypass reactor

Hg0 through reactor with SCR

Hgtotal through converter

Figure 5.4. Measured total mercury concentration using 4g SCR catalyst at 150C and

2.75 Nl/min flue gas containing only 15 ppmv HCl and 1% H2O, 20 g red brass in the

converter at 180C.

Based on these tests it was suggested by the supplier that the converter material

reaches stability only after a period of several hours of operation under the gas mixture

investigated in this project [22]. Tests were then conducted by conditioning the converter

125

with gases excluding Hg0 addition. Detailed results are shown in figure 5.5. After

conditioning the red brass catalyst with 15 ppmv HCl, 1000 ppmv SO2, 500 ppmv NO,

15 ppmv NO2 and 1% H2O for 40 h, the measured mercury level through the converter is

about 90 g/Nm3 compared to Hg0 inlet level of 210 g/Nm3. Furthermore, the measured

mercury concentration through the converter keeps decreasing with time. The results

show that the red brass converter does not work for the present conditions.

0 1 2 3 4 5

Time (hour)

0

40

80

120

160

200

240

Hg

co

nce

ntr

atio

n (g

/Nm

3 )

Hgtotal through converter

Hg0 bypassconverter

Figure 5.5. Measured total mercury concentration after passing SCR reactor with 4g SCR

catalyst at 150C. 20 g red brass in the converter at 180C was preconditioned by 2.75

Nl/min flue gas containing 15 ppmv HCl, 1000 ppmv SO2, 1% H2O, 500 ppmv NO, 15

ppmv NO2 for 40 h.

5.4 Performance of the sulfite converter 

The principle of the sulfite converter is that oxidized mercury such as HgCl2 can

be reduced to Hg0 through following reaction [23]:

122 2 3 2 22HgCl Na SO Hg NaCl SO O (5R7)

126

It is reported that 95% or more HgCl2 reduction efficiency can be obtained at

300-500C [23]. It is not stated in the patent for which gas composition this conversion

was obtained and for how long. Thus a fundamental investigation of the sulfite converter

under simulated cement kiln flue gas is necessary. The effect of converter temperature on

Hg0 recovery was tested and the results are illustrated in table 5.1. With 5 g sulfite

compounds at 350C, full mercury recovery can only be obtained for 1 h. Then the Hg0

recovery decreases with time and drops to 43% after another 3.5 h. For 10 g sulfite

compounds at 450C, full mercury recovery was obtained for 2 h. Then the mercury

recovery decreased with time and dropped to 87% after another 4 h.

Fast deactivation of sodium sulfite can take place at high temperatures [23].

Sodium sulfite is water soluble and can recrystallize when water is present in the gas.

When recrystallization occurs, the resistance of a layer of the sodium sulfite to gas

transport is increased and the oxidized mercury reduction efficiency may be reduced. The

higher the temperature the more recrystallization of sodium sulfite may take place. To

minimize the deactivation, the converter was first tested at 250C. However, only about

50% mercury recovery was obtained immediately after switching gas to the converter

and the Hg0 recovery kept decreasing to 36% after another 1.5 h. Then the converter

temperature was increased to 500C. The mercury concentration after the converter

increased sharply to 180% of the inlet elemental mercury level right after increasing the

converter temperature. This is probably due to the fact that mercury is first adsorbed on

the converter material at 250C and then desorbs at high temperatures. Full mercury

recovery was obtained for 1 h and then the mercury recovery decreased slowly to 88%

after another 13 h.

127

Table 5.1. Test results of elemental mercury recovery by the sulfite converter in 2.75

Nl/min simulated cement kiln flue gas containing 21% CO2, 6% O2, 1% H2O, 1000 ppmv

NO, 30 ppmv NO2, and 1000 ppmv SO2.

Sulfite compound load

(g)

Converter temperature

(C)

HCl level (ppmv)

Short time performance

Long time performance

5 350 15 Full Hg0 recovery for 1 h

Hg0 recovery decreases to 43% after 3.5 h

10 250 15 50% Hg0 recovery for 0.5 h

Hg0 recovery decreases to 36% after 1.5 h

10 450 15 Full Hg0 recovery for 2 h

Hg0 recovery decreases to 87% after 4 h

20 500 (after 250C test)

15 Full Hg0 recovery for 1 h

Hg0 recovery decreases to 88% after 13 h

20 500 2 Full Hg0 recovery for 72 h

Not tested

20 500 6 Full Hg0 recovery for 24 h

Not tested

20 500 10 Full Hg0 recovery for 15 h

Hg0 recovery decreases to 95% after 35 h

The level of HCl in the flue gas is a key factor that determines the efficiency and

lifetime of the converter. Since only short time of full oxidized mercury reduction was

observed with 15 ppmv HCl in the simulated flue gas, the HCl level was decreased to

study the effects. With 2, 6, and 10 ppmv HCl in the simulated gas, full oxidized

reduction can be obtained for at least 72, 24, and 15 h, respectively, for short-term test.

With 10 ppmv HCl in the simulated cement kiln flue gas continuous operation of the

converter with 20g sulfite material up to 2-3 months has been achieved. The presence of

HCl in the gas can result in mercury oxidation both in the flue gas and on the sorbent.

The recombination of elemental mercury and HCl after the converter might also be

enhanced with high levels of HCl in the gas.

It should be noted that the sulfite can be oxidized to sulfate by oxygen in the flue

gas:

122 3 2 2 4Na SO O Na SO (5R8)

128

To study the effects of sodium sulfite oxidation to sulfate on oxidized mercury reduction

efficiency, the sulfite pellets were first exposed to 12% O2 at 350C for 18 h. Figure 5.6

shows that the maximum Hg0 recovery is about 90% after preconditioning by oxygen and

slowly decreases with time. This indicates that the sodium sulfite was partly oxidized to

sodium sulfate during preconditioning by oxygen. The formed sulfate is not active for

reduction of oxidized mercury to elemental mercury. Normally the analyzer is running all

the time to avoid damage of the lamp by restarting of the analyzer. When the experiment

is not run, air is added to the analyzer. The test of oxygen precondition indicates that the

converter should be closed to avoid oxidation of the sulfite compound when the

experiment is not running. Instead the analyzer is running in Hg0 measurement mode

with air to the analyzer, bypassing the converter.

0 30 60 90 120

Time (min)

0

40

80

120

160

200

Hg

co

nce

ntr

atio

n (g

/Nm

3 ) Hgtotal bypassreactor Hgtotal through reactor

Hg0 inlet

Figure 5.6. Test of 20 g sodium sulfite converter materials at 350C using 2.75 Nl/min

simulated flue gas with 15 ppmv HCl. The sulfite pellets were pre-conditioned by 12%

O2 for 18 h. Oxidized mercury is produced by passing gases through 4 g SCR catalyst at

150C.

129

The dynamics of the converter were investigated by studying the response time of

mercury measurement to the change of mercury addition and switching between the

reactor and bypass. This was carried out by step up and step down tests [7]. The

dynamics of Hg0 measurement bypassing the converter were first investigated. The steps

of the dynamics test are illustrated in figure 5.7 Air was used as zero gas and added to the

analyzer directly until a stable reading was achieved for about 5 min. Then air addition

was stopped and Hg0 in simulated flue gas was added bypassing the reactor with SCR

catalyst and the sulfite converter to measure the Hg0 inlet level. After a stable reading of

the inlet Hg0 level was obtained for about 10 min, zero air was added to the analyzer and

step down test of Hg0 measurement was finished when stable reading was obtained. The

95% response time of both step up and down is less than 0.5 min for elemental mercury

measurement bypassing the sulfite converter. The step change is very similar to that seen

in the elemental measurement shown in figure 5.2.

0 30 60 90 120

Time (min)

0

20

40

60

80

100

120

140

Hg

co

nce

ntr

atio

n (g

/Nm

3 )

BypassSCR

Air to converter, gas to SCR

ThroughSCR

Air to converter

BypassSCR

BypassSCRstop Hgaddition

Figure 5.7. Response of total mercury measurement with elemental Hg inlet level of 112

g/Nm3. 20 g sodium sulfite converter materials is used in the converter at 500C using

2.75 Nl/min simulated flue gas with 10 ppmv HCl. Oxidized mercury is produced by

passing gases through 4 g SCR catalyst at 150C.

130

Then the measurement was switched to total mercury measurement through the

converter. The Hg0 in the simulated flue gas was added to the reactor with SCR catalyst

and the converter. The step up test of total mercury measurement was finished when

stable mercury measurement through the converter was achieved for 50 min. Then air

was added to the converter for a step down test. The dynamic tests were conducted at

different Hg0 inlet levels of 41, 112, and 150 µg/Nm3. Figure 5.7 shows the response of

both Hg0 measurement bypassing the converter and SCR catalyst and total mercury

measurement through the SCR catalyst and converter with a Hg0 inlet level of 112

µg/Nm3. Close look at the response for step change in figure 5.8 shows that the response

of mercury measurement is very fast for both step up and down tests. The 95% response

time for step up and down change is 1.5 and 0.6 min, respectively.

34 36 38 40 42 44 46 48 50

Time (min)

0

20

40

60

80

100

120

Hg

co

nce

ntr

atio

n (g

/Nm

3)

78 80 82 84 86

Time (min)

95% step down change

95% step up change

0.60 min1.50 min

Figure 5.8. Close look of the response time test shown in figure 5.7. Left: step up test by

switching gas addition to the sulfite converter from air to simulated flue gas with

oxidized mercury produced by SCR catalyst. Right: step down test by switching gas

addition to the sulfite converter from simulated flue gas with oxidized mercury produced

by SCR catalyst to air.

131

5.5 Examples of dynamic measurement of mercury adsorption and 

oxidation on activated carbon 

Commercial activated carbons, Darco Hg and HOK standard were investigated in

simulated cement kiln flue gas at 150C. Figure 5.9 shows the mercury profiles for the

Darco Hg activated carbon. The experiments are conducted twice using separate total and

elemental mercury measurement. Comparison of the elemental and total mercury

measurement shows that both adsorption and oxidation of mercury by the carbon occur.

After mercury breakthrough is achieved, the mercury oxidation is stable at about 92%.

0 0.5 1 1.5 2 2.5 3 3.5

Time (hour)

0

40

80

120

160

200

Gas

eou

s H

g (g

/Nm

3 )

Gaseous Hgtotal

Gaseous Hg0

bypassreactor

throughreactor

Figure 5.9. Total and elemental mercury profile of 30 mg Darco Hg activated carbon

mixed with 2 g sand at 150C using 2.75 Nl/min simulated cement kiln flue gas with 10

ppmv HCl, two separate tests and measurements.

Rather than running the test of the same carbon twice as shown in figure 5.9, it is

possible to run the test once and evaluate both the mercury adsorption and oxidation by

the carbon. The mercury breakthrough was first obtained by measuring total mercury

through the converter, then zero air is added to the converter and the analyzer was

changed to Hg0 measurement mode. During this period mercury in simulated gas was

still added to the sorbent to avoid possible desorption of mercury from the carbon. When

132

the analyzer was running for Hg0 measurement, gases after the reactor were switched to

the analyzer to measure Hg0. Figure 5.10 illustrates the mercury adsorption and oxidation

by HOK standard carbon at 150C. In this case 57% mercury oxidation was observed.

Both the tests of Darco Hg and HOK in simulated cement kiln flue gas show that the

sulfite converter and analysis system are capable of following the transient mercury

outlet concentration in a satisfactory way.

0 1 2 3 4 5

Time (min)

0

40

80

120

160

200

Gas

eou

s H

g (g

/Nm

3 )

bypassreactorHgtotal

through reactor Hgtotal

Gas to reactor, air to analyzerchange from Hgtotal to Hg0 measurement

through reactor Hg0

Figure 5.10. Total and elemental mercury profile of 30 mg HOK standard activated

carbon mixed with 2 g sand at 150C using 2.75 Nl/min simulated cement kiln flue gas

with 10 ppmv HCl.

5.6 Suggestions for practical application of the converter 

The conditions in full-scale application are much more demanding than in the lab-

scale investigation. In this work no particles in the gas stream were applied. On the other

hand, the dust load in the flue gas between the raw mill and filter could be up to 800-

1000 g/Nm3. The sampling probe needs to be able to separate the particles from the flue

gas efficiently to avoid plugging of probe. Adsorption of mercury by the dust and probe

should be minimized by high sampling flow rate and high filter temperature.

133

The HCl content in the cement kiln flue gas can be up to 20-25 ppmv [24]. As

found in this work, the full Hg0 recovery can only be maintained for short period when

more than 10 ppmv HCl is present in the flue gas. It is therefore necessary to remove HCl

before or in the converter. Lime pellets can be used together with the sulfite compounds

and the converter temperature should be high enough to avoid mercury adsorption on the

lime pellets. Alternatively, large amount of converter material might be used.

Compared to power plants and incinerators, the emission levels of CO and

volatile organic compounds such as hydrocarbons are higher in cement plants. The

emission level of volatile organic compound in the stack gas of cement kilns is usually

between 10 and 100 mg/Nm3, with a few excessive cases up to 500 mg/Nm3[25]. The CO

concentration in the stack gas can be as high as 1000 mg/Nm3, even exceeding 2000

mg/Nm3 in some cases. High levels of CO and hydrocarbons in the flue gas will cause

fast contamination of the windows in the analytical cells and interruption of the mercury

measurement [22]. Measures such as dilution should be applied to minimize the problem.

The sulfite material should be kept in a closed box to avoid oxidation by air and

moisture. It is important that the sulfite powders are adhered uniformly to the surface of

the thin layer of water glass on the zeolite pellets. Thoroughly mixing the water glass

with zeolite pellets in a plastic container can improve the sulfite converter performance.

In this way, the sulfite converter can work well up to months, as observed in this work.

For both lab-scale and full-scale application of the analysis system, it is important

to avoid cold parts in the system. All the connections, Teflon lines and gas contacting

parts in the analyzer before the spectrometry should be heated above 150C. When the

converter is not used, the converter temperature should be decreased to 100C. The

converter should be closed to avoid deactivation of the converter material due to

oxidation of sulfite to sulphate by air.

5.7 Conclusions  

To be able to perform dynamic measurement of mercury adsorption by sorbents,

red brass chips and sulfite converter were investigated in simulated cement kiln flue gas

134

in a fixed-bed reactor system. The converter with red brass chips works only when

measuring elemental mercury in nitrogen (i.e., without carrying out actual conversion)

and does not work properly even when only elemental mercury was added to the

simulated flue gas. The red brass is poisoned or oxidized within a short time and adsorbs

elemental mercury. When oxidized mercury was produced by passing gases through a

separate reactor with an SCR catalyst, the red brass converter cannot fully reduce HgCl2

to elemental mercury under any relevant condition.

Sodium sulfite converter material was prepared by dry impregnation of sodium

sulfite and calcium sulfate powders on zeolite pellets using water glass as binder. The

optimal operating temperature of the sulfite converter is 500C. The level of HCl in the

flue gas is a key factor that determines the efficiency and lifetime of the converter. Full

elemental mercury recovery can only be obtained for short period with 15 ppmv HCl in

the simulated gas, but the sulfite converter works well at 500C with up to 10 ppmv HCl

in the simulated cement kiln flue gas. When the converter is not used, the converter

temperature was decreased to 100C without air passing through to avoid deactivation of

the converter material by oxidation of the sodium sulfite to sodium sulfate. The response

time of the sulfite converter is short and typically within at most two minutes, which

makes it appropriate for not too fast dynamic measurements, as verified by dynamic

mercury adsorption tests on commercial activated carbons Darco Hg and HOK standard

in a fixed-bed reactor. Suggestions for practical application of the sulfite converter in

cement plant with high dust load are provided.

5.8 References 

[1] R.J. Schreiber and C.D. Kellett, Compilation of mercury emissions data, PCA R&D Serial No.

SN3091, 2009.

[2] D.L. Laudal, J.S. Thompson, J.H. Pavlish, L.A. Brickett, P. Chu, Use of continuous mercury

monitors at coal-fired utilities, Fuel Processing Technology. 85 (2004) 501-511.

[3] J. Wu, Y. Du and W. Pan, J. Ren, P. He, W. Wang, M. Shen, X. Leng, Y. Jin, Z. Dai, L. Zhao,

X. Ming, Y. Cao, W. Pan, Study on different measurement methods of mercury emission in the

135

coal-fired power station, 3rd International Conference on Bioinformatics and Biomedical

Engineering, Beijing, China, June 11-13, 2009.

[4] V. Schmid, Continuous monitoring of mercury emissions from stationary sources, 2002.

[5] M. Holmes and J. Pavlish, Mercury information of clearinghouse, Quarterly 2–mercury

measurement, 2004.

[6] J. Wang, Z. Xiao, O. Lindqvist, On-line measurement of mercury in simulated flue gas, Water,

Air, & Soil Pollution. 80 (1995) 1217-1226.

[7] M.L. Jones, D.L. Laudal and J.H. Pavlish, Mercury emission monitoring for the cement

industry, Cement Industry Technical Conference Record, 2008 IEEE, Miami, Florida, May 18-22,

2008.

[8] Wikipedia, Brass, http://en.wikipedia.org/wiki/Brass, accessed December 7, 2010.

[9] R. Kanefke, H. Köser, B. Vosteen, F. Kristina, B. Frank, S. Raik, Method for the production

of elemental mercury from mercury compounds, patent WO 2008/064667 A2, 2008.

[10] J.A. Rodriguez, T. Jirsak, J. Dvorak, S. Sambasivan, D. Fischer, Reaction of NO2 with Zn

and ZnO: Photoemission, XANES, and density functional studies on the formation of NO3, The

Journal of Physical Chemistry B. 104 (2000) 319-328.

[11] S. Lee, J. Lee, T.C. Keener, Bench-scale studies of in-duct mercury capture using cupric

chloride-impregnated carbons, Environ. Sci. Technol. 43 (2009) 2957-2962.

[12] S. Lee, J. Lee, T.C. Keener, The effect of methods of preparation on the performance of

cupric chloride-impregnated sorbents for the removal of mercury from flue gases, Fuel. 88 (2009)

2053-2056.

[13] A. Makkuni, R.S. Varma, S.K. Sikdar, D. Bhattacharyya, Vapor phase mercury sorption by

organic sulfide modified bimetallic iron-copper nanoparticle aggregates, Ind Eng Chem Res. 46

(2007) 1305-1315.

[14] D.E. Meyer, S.K. Sikdar, N.D. Hutson, D. Bhattacharyya, Examination of sulfur-

functionalized, copper-doped iron nanoparticles for vapor-phase mercury capture in entrained-

flow and fixed-bed systems, Energy & Fuels. 21 (2007) 2688-2697.

[15] D.E. Meyer, N. Meeks, S. Sikdar, N.D. Hutson, D. Hua, D. Bhattacharyya, Copper-doped

silica materials silanized with bis-(triethoxy silyl propyl)-tetra sulfide for mercury vapor capture,

Energy Fuels. 22 (2008) 2290-2298.

[16] D.L. Laudal, J.S. Thompson, J.H. Pavlish, L. Brickett, P. Chu, R.K. Srivastava, J. Kilgroe,

C.W. Lee, Mercury speciation at power plants using SCR and SNCR control technologies, EM:

Air and Waste Management Association's Magazine for Environmental Managers. (2003) 16-22.

[17] H.G. Pedersen, L.S. Pedersen, H. Rostgaard and K. Pedersen, Oxidation of mercury on DNX

catalysts. Proceedings of the Air Quality V: Mercury, Trace Elements, SO3, and Particulate

Matter Conference, Arlington, VA, Sept 19–21, 2005.

136

[18] S. Straube, T. Hahn, H. Koeser, Adsorption and oxidation of mercury in tail-end SCR-

DeNOx plants—Bench scale investigations and speciation experiments, Applied Catalysis B:

Environmental. 79 (2008) 286-295.

[19] Y. Zhuang, J. Laumb, R. Liggett, M. Holmes, J. Pavlish, Impacts of acid gases on mercury

oxidation across SCR catalyst, Fuel Processing Technology. 88 (2007) 929-934.

[20] Y. Cao, Z. Gao, J. Zhu, Q. Wang, Y. Huang, C. Chiu, B. Parker, P. Chu, W. Pan, Impacts of

halogen additions on mercury oxidation in a slipstream selective catalyst reduction (SCR) reactor

when burning sub-bituminous coal, Environ. Sci. Technol. 42 (2008) 256-261.

[21] H. Kamata, S. Ueno, T. Naito, A. Yukimura, Mercury oxidation over the V2O5(WO3)/TiO2

commercial SCR catalyst, Ind Eng Chem Res. 47 (2008) 8136-8141.

[22] R. Moeseler. Issues about red brass converter, personal communication, Lumex Analytical

GmbH, 2010.

[23] S. Akiyama, J. Kato, F. Koga, K. Ishikawa, Catalysts for reducing mercury, a mercury

conversion unit, and an apparatus for measuring total mercury in combustion exhaust gas by

using the same, patent US2007/0232488 A1, 2007.

[24] C. Senior, A. Sarofim and E. Eddings, Behavor and measurement of mercury in cement

kilns, presented at the IEE-IAS/PCA 45th Cement Industry Technical Conference, Dallas, Texas,

May 4-9 2003.

[25] CEMBUREAU, the European Cement association, Best available technologies for the

cement industry, 1999.

137

6

Effects of bed dilution and carbon load on

mercury adsorption capacity of activated

carbon

This chapter reports the effects of bed dilution and carbon load on the equilibrium

mercury adsorption capacity of the activated carbon. The mercury adsorption capacity

per unit mass of the activated carbon decreases when the carbon load is increased.

Detailed investigations are conducted to reveal the cause.

6.1 Introduction 

Most of the studies on mercury adsorption use bed dilution [1-9], while only few

investigations apply pure sorbent bed [10-14] when the mercury sorbents are evaluated in

fixed-bed reactors. The sorbent beds are often diluted with inert particles to suppress

other potential disturbing effects such as axial dispersion and bypassing [15,16]. Low-

surface-area materials such as glass beads and sand/quartz powder are preferred as

diluting solids because of their relative inertness and good heat transfer properties. The

effects of sorbent load on the mercury adsorption capacity of the sorbent are rarely

reported in the literature.

6.2 Effects of carbon load 

The direct result of the fixed-bed test is the mercury adsorption breakthrough

curve. The percentage breakthrough is determined as a function of time by normalizing

the measured total mercury concentration at the outlet of the sorbent bed to the inlet

mercury concentration.

138

From the mercury breakthrough curve, the amount of mercury adsorbed on unit

mass of the sorbent as a function of time can be calculated from the expression:

t

toutint dtCCW

Fq

0

, )( (6.1)

where F is the flow rate through the sorbent bed, W is the mass of the sorbent, Cin is the

inlet mercury concentration, Cout,t is the mercury concentration at the reactor outlet at

time t. The mercury adsorption capacity of a known weight of a sorbent is calculated in

terms of µg Hg adsorbed/mg_sorbent from the breakthrough curve for the sorbent. The

equilibrium adsorption capacity is defined by the time when the outlet Hg concentration

is first equal to the inlet concentration.

Figure 6.1 presents the mercury breakthrough curves for different loads of Darco

Hg activated carbon mixed with 2 g sand powder at 150C using simulated cement kiln

flue gas with elemental mercury. Faster mercury breakthrough is observed for smaller

carbon load as expected. The calculated amount of adsorbed mercury and equilibrium

mercury adsorption capacity per unit mass of the activated carbon are illustrated in figure

6.2. The calculated amount of mercury adsorbed in the carbon does not increase

proportionally to the mass of carbon, i.e., the mercury adsorption capacities of the carbon

apparently decreases when the carbon load is increased. It seems that there is promotion

of mercury adsorption by the sand when it is mixed with activated carbon. The trend line

indicates that about 8.39 µg mercury is adsorbed by 2 g sand powder. Stuart [17] also

reported that activated carbon mixed with sand had larger mercury uptake capacity than

the carbon tightly packed in the reactor. He postulated that the incoming gas might be

short circuiting and allowing the gas flow through the reactor without encountering all

the tightly packed carbon. This argument is doubtful since even if the contact is poor

uptake of mercury would just be lower and eventually the same uptake will be reached.

139

0 1 2 3 4 5 6

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Co

ut/C

in

A B C D

A: 10 mgB: 30 mgC: 60 mgD: 100 mg

Figure 6.1. Mercury breakthrough curves of different Darco Hg activated carbon loads

mixed with 2 g sand powder and tested at 150C using 2.75 Nl/min simulated flue gas

with 170 µg/Nm3 elemental mercury, 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10

ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

0 20 40 60 80 100 120

Carbon load in 2 g sand (mg)

0

10

20

30

40

50

Ad

sorb

ed H

g (g

)

0

0.2

0.4

0.6

0.8

1

1.2H

g A

dso

rpti

on

cap

acit

y (

g H

g/m

g_c

arb

on

)

Adsorbed Hg

Adsorbed HgY=0.3797X+8.3921,R2=0.99Adsorption capacity

Figure 6.2. Calculated amount of adsorbed mercury and equilibrium mercury adsorption

capacity of Darco Hg activated carbon mixed with 2 g sand powder and tested at 150C

with different carbon loads using 2.75 Nl/min simulated flue gas with 170 µg/Nm3

elemental mercury, 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv HCl, 1

vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

140

To investigate whether the oxidation of elemental mercury could influence the

mercury adsorption capacity, mercury adsorption by different carbon loads using HgCl2

source are conducted. As shown in figure 6.3, similar trends as tests with elemental

mercury source are observed. This implies that the decrease of mercury adsorption

capacity with increased carbon load in the sand is not caused by the oxidation of

elemental mercury.

0 10 20 30 40 50 60 70 80

Carbon load (mg)

0

5

10

15

20

25

30

35

Ad

sorb

ed m

ercu

ry (g

)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

Hg

ad

sorp

tio

n c

apac

ity

(g

Hg

/mg

_ca

rbo

n)

Adsorbed HgAdsorbed Hg Y=0.426X+6.739,R2=0.96Adsorption capacity

Figure 6.3. Calculated amount of adsorbed mercury and equilibrium mercury adsorption

capacity of Darco Hg activated carbon mixed with 2 g sand powder and tested at 150C

with different carbon loads using 2.75 Nl/min simulated flue gas with 170 µg/Nm3

mercury from HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv

HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

One possible cause of the decrease of mercury adsorption capacity with increased

carbon load could be the wall effect. Carbon particles separate from sand powders during

loading the sample to the reactor. A layer of carbon particles deposits on the sample

holder. Later these carbon particles are loaded to the reactor by knocking the sample

holder. During loading of the sample to the reactor, carbon particles stick on the reactor

wall. A quartz wool plug is used to move the carbon particles that adhere on the reactor

wall to the top of the carbon bed. As a result some carbon particles are loaded to the area

141

close to the reactor wall. More carbon particles adhere on the reactor wall when larger

carbon load is applied.

Another reason might be the leakage of the system and mercury adsorption by the

quartz wool and sand powder. However, leakage tests of the system at different stages of

the project show that the system is tight. The pressure drop over the carbon bed is about 5

mbar. The deviation of the flow rates at the reactor inlet and outlet to the flow rate after

the gas mixing panel is within 2.2%. Tests with empty reactor, reactor with quartz wool,

and sand powder do not show any adsorption of either elemental mercury or mercury

chloride.

6.3 Effects of bed dilution 

Negative deviation of conversion caused by dilution of the catalyst bed with inert

particles in gas-solid systems has been reported [15,16]. Dilution of activated carbon by

inert sand powder is applied in this work; it is therefore relevant to evaluate the possible

effects caused by the dilution. The extent of negative effect depends on the amount of

dilution, the reaction/adsorption kinetics, the particles and reactor geometry, and the

degree of segregation of carbon and sand. Since the mercury removal fraction by the

carbon bed changes with time, the dilution effect as a relative measure of the deviation

in the conversion can be calculated for different time:

( ) ( )( ) undiluted diluted

undiluted

x t x tt

x

(6.2)

where xdiulted(t) and xundiluted(t) is the mercury removal fraction at time t for diluted and

undiluted bed, respectively.

For practical application the relative deviation in conversion can be estimated

from observable parameters [15,16] :

( )( ) ( )

1 2p diluted

bed

d x tbt

b h

(6.3)

where b is the volume of inert sand as fraction of total volume of solids, dp is carbon

particle diameter, and hbed is the bed height.

For 10 mg Darco Hg carbon mixed with 2 g sand, the b is calculated as:

142

3

6 3

/ 2 10 /16020.98

/ / 10 10 / 510 2 10 /1602sand sand

carbon carbon sand sand

mb

m m

Figure 6.4 presents the calculated relative deviation in mercury adsorption as a

function of time for different loads of Darco Hg carbon tested at 180C in simulated

cement kiln flue gas. Larger relative deviation in short period is observed for smaller

carbon loads, i.e., larger dilution ratio. However, the area under the relative deviation

curve and above the zero deviation appears to be similar for different carbon loads. This

indicates that the influence of bed dilution on the equilibrium mercury adsorption

capacity of the carbon is similar. Therefore the decrease of mercury adsorption capacity

with increase of carbon loads is probably not caused by the bed dilution.

0 0.5 1 1.5 2 2.5

Time (hour)

-0.01

0

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

Rel

ativ

e d

evia

tio

n,

A

B

C

A: 10 mg carbon, b=0.98B: 30 mg carbon, b=0.96C: 60 mg carbon, b=0.91

Figure 6.4. The calculated relative deviation as a function of time for tests at 180C

with different carbon loads using 2.75 Nl/min simulated flue gas with 170 µg/Nm3

mercury from HgCl2, 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv HCl, 1

vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

Tests using same dilution ratio, i.e., 10 mg carbon is mixed with 2 g sand, 30 mg

carbon with 6 g sand, 60 mg carbon with 12 g sand, are also performed at 180C using

simulated cement kiln flue gas with HgCl2. Figure 6.5 shows the calculated amount of

adsorbed mercury and equilibrium mercury adsorption capacity of the activated carbon.

143

Similar mercury adsorption capacity is still not obtained for different carbon loads, which

behaves as with 2 g sand.

0 10 20 30 40 50 60 70 80

Carbon load (mg)

0

5

10

15

20

25

30

Ad

sorb

ed m

ercu

ry (g

)

0

0.2

0.4

0.6

0.8

1

Hg

ad

sorp

tio

n c

ap

acit

y (

g H

g/m

g_c

arb

on

)

Adsorbed HgAdsorbed HgY=0.3777X+7.0386,R2=0.97Adsorption capacity

Figure 6.5. Calculated amount of adsorbed mercury and equilibrium mercury adsorption

capacity of Darco Hg activated carbon mixed with sand powder using same dilution rate

and tested at 180C with different carbon loads using 2.75 Nl/min simulated flue gas

with 170 µg/Nm3 mercury from HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000

ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

6.4 Effects of sand load 

If the mercury adsorption by activated carbon is promoted by the sand mixing, it

would be interesting to investigate the effects of sand load on the promotion of mercury

adsorption capacity of the carbon by running tests with different masses of sand. Table

6.2 presents the calculated amount of adsorbed mercury and equilibrium mercury

adsorption capacity of 10 mg Darco Hg activated carbon mixed with different amounts of

sand powder at 150C in simulated cement kiln flue gas with HgCl2. When the sand load

is above 20 mg the mercury adsorption capacities of the Darco Hg do not increase further

and level off at a value of about 1.135 µg Hg/mg_carbon. This also indicates that the

repeatability of the experiment is reasonable.

144

Table 6.2. Calculated amount of adsorbed mercury and equilibrium mercury adsorption

capacity of 10 mg Darco Hg activated carbon mixed with different amounts of sand

powder and tested at 150C using 2.75 Nl/min simulated flue gas with HgCl2 source,

1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2,

and 21 vol.% CO2.

Sand load (g) Hg inlet (µg/Nm3) Adsorbed Hg (µg) Adsorption capacity (µg Hg/mg_carbon)

0 164 5.06 0.506 0.01 161 9.81 0.981 0.02 160 11.36 1.136 0.05 166 11.48 1.148 0.10 164 11.07 1.107 0.25 166 12.17 1.217 0.5 174 10.50 1.050 1 217 10.24 1.024 2 183 12.24 1.224 2 163 11.80 1.180 4 206 11.29 1.129

6.5 Effects of carbon loading location 

The carbon sample was separated from the sand powder with quartz wool plug to

test possible effect of carbon loading location. Different locations of the carbon sample

are applied to investigate whether the promotion of mercury adsorption is caused by the

preconditioning of the gas by the sand. When the carbon is on top of the sand, the flue

gas first contacts the carbon powder. However, as shown in figure 6.6, the equilibrium

mercury adsorption capacity is almost the same when the carbon sample is loaded on top

of and under the sand powder. This implies that the promotion of mercury adsorption

only occurs when the carbon is mixed with sand powder. The slightly larger mercury

adsorption capacity with sand powder in the reactor compared to only carbon in the

reactor might be due to the improved contact of carbon with gas flow.

145

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Hg

ad

sorp

tio

n c

apac

ity

(g

Hg

/mg

-ca

rbo

n)

No

san

d

Car

bo

n o

n t

op

San

d o

n t

op

Figure 6.6. Equilibrium mercury adsorption capacity of 10 mg Darco Hg activated

carbon on top of and under 1 g sand powder at 150C using 2.75 Nl/min simulated flue

gas with 170 µg/Nm3 mercury from HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000

ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

6.6 Effects of bed materials 

Different bed materials are applied to investigate the possible effects of bed

materials on the mercury adsorption capacity of activated carbon. The investigated bed

materials include sand powder, fine quartz powder, and glass beads and the mean

diameters of these materials are 215, 2, and 180 µm, respectively. Baseline tests with

only fine quartz powder and glass beads show that fine quartz powder is inert for

mercury adsorption and the mercury adsorption by the glass beads is negligible. Figure

6.7 compares the mercury adsorption capacity of Darco Hg activated carbon tested with

different bed materials. The mercury adsorption capacity of Darco Hg carbon tested with

fine quartz powder is much smaller than those with sand powder and glass beads. The

fine quartz powder behaves like paste and might hinder the contact of gas with the

carbon particles. Inconsistent mercury adsorption capacity is still obtained when different

amounts of carbon are mixed with 2 g sand powder, fine quartz powder, and glass beads.

146

0

0.2

0.4

0.6

0.8

1

1.2

1.4

Hg

ad

sorp

tio

n c

ap

aity

(g

Hg

/mg

_c

arb

on

)

10 mg carbon in 2 g bed material

30 mg carbon in 2 g bed material

San

d p

ow

der

Fin

e q

uar

tz p

ow

der

Gla

ss b

ead

s

San

d p

ow

der

Fin

e q

uar

tz p

ow

de

r

Gla

ss b

ead

s

Figure 6.7. Equilibrium mercury adsorption capacity of Darco Hg activated carbon

mixed with different bed materials and tested at 150C using 2.75 Nl/min simulated flue

gas with 170 µg/Nm3 mercury from HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000

ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

6.7 Effects of carbon type and particle size 

To study the effects of carbon type and particle size on the mercury adsorption

capacity, commercial activated carbon pellets of Norit RB4 are crushed and sieved to

size of 165 and less than 32 µm in diameter. Figure 6.8 shows the mercury adsorption

capacity obtained with different carbon loads, carbon types and particle size. Inconsistent

mercury adsorption capacity at different carbon loads is observed for both Darco Hg and

Norit RB4 carbons with different sizes. The effects of carbon load are much smaller for

Norit RB4 carbon.

147

0 10 20 30 40 50 60 70 80

Carbon load (mg)

Darco Hg, 16 m

Norit RB 4, 165 m

Norit RB 4, <32 m

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2

Hg

ad

sorp

tio

n c

apac

ity

(g

Hg

/mg

_car

bo

n)

Figure 6.8. Equilibrium mercury adsorption capacity as a function of carbon loads for

Darco Hg activated carbon and Norit RB4 with different sizes mixed with 2 g sand

powder and tested at 150C using 2.75 Nl/min simulated flue gas with 170 µg/Nm3

mercury from HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv

HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. 6.8 Tests with only Portland cement 

Instead of using sorbent and sand mixture, 1-4 g Portland cement is tested at

150C with HgCl2 in the simulated cement kiln flue gas. Figure 6.9 shows that the

adsorbed mercury is proportional to the cement load and the equilibrium mercury

adsorption capacity of the cement is similar for different cement loads.

148

1 2 3 4 5

Portland cement load (g)

2

3

4

5

6

7

8

9

10

Ad

sorb

ed H

g (g

)

0

0.0005

0.001

0.0015

0.002

0.0025

Hg

ad

sorp

tio

n c

apac

ity

(g

Hg

/mg

_cem

ent)

Adsorbed HgAdsorption capacity

Figure 6.9. Calculated amount of adsorbed mercury and equilibrium mercury adsorption

capacity as a function of Portland cement load at 150C using 2.75 Nl/min simulated flue

gas with 170 µg/Nm3 mercury from HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000

ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. 6.9 Conclusions 

Inconsistent mercury adsorption capacity of activated carbon is observed at

different carbon loads when mixed with sand. Smaller mercury adsorption capacity is

obtained with larger carbon load. Tests with elemental mercury and mercury chloride,

different carbon type and particle sizes show the same trend. Effects of bed dilution at

fixed carbon load on the equilibrium mercury adsorption capacity appear to be limited.

The mercury adsorption capacity of activated carbon obtained using sand and

carbon mixture is larger than that obtained with only activated carbon. The mercury

adsorption capacity with 10 mg carbon increases with sand load up to 20 mg and then

levels off when the sand load is further increased.

Similar mercury adsorption capacities are obtained with different Portland cement

loads in the reactor. This implies that the inconsistent mercury adsorption capacity of

carbon obtained using different carbon loads might be due to possible adsorption of

149

mercury by sand when it is mixed with carbon, rather than the failure of the experimental

setup. The sand powder alone is inert for mercury adsorption, while after modification

with chemical reagent it can be used for mercury adsorption [18,19]. In-situ analysis

technology is required to reveal whether mercury is adsorbed by the sand when it is

mixed with activated carbon.

The problem of inconsistent mercury adsorption capacity was encountered in the

late stage of the project when performing a fundamental parametric study. Although

detailed tests are conducted to reveal the cause, the problem is not solved due to the lack

of analysis techniques and time. It is impossible to repeat and run all the tests with only

large carbon load within the time schedule of the project. For a full-scale application in

the cement plant it is impossible to exclude all the cement materials in the flue gas even

with a polishing filter. Instead of providing actual kinetics data relevant to full-scale

application conditions, this work aims at evaluation the effects of different operating

parameters and mathematical model development. Therefore, mercury adsorption

kinetics obtained using 10 mg activated carbon mixed with 2 g sand powder is used in

the following chapters dealing with parametric study and model development. It may be

argued that it is difficult to state what the real capacity is if it depends on the mixing

condition. In reality there will always be least 20 mg diluter with the sorbent and this is

where it has stabilized.

6.10 References 

[1] S. Sjostrom, T. Ebner, T. Ley, R. Slye, C. Richardson, T. Machalek, M. Richardson, R.

Chang, Assessing sorbents for mercury control in coal-combustion flue gas, J. Air & Waste

Manage. Assoc. 52 (2002) 902.

[2] S.J. Lee, Y. Seo, J. Jurng, T.G. Lee, Removal of gas-phase elemental mercury by iodine- and

chlorine-impregnated activated carbons, Atmospheric Environment. 38 (2004) 4887-4893.

[3] D. Karatza, A. Lancia, D. Musmarra, Fly ash capture of mercuric chloride vapors from

exhaust combustion gas, Environ. Sci. Technol. 32 (1998) 3999-4004.

[4] J.W. Portzer, J.R. Albritton, C.C. Allen, R.P. Gupta, Development of novel sorbents for

mercury control at elevated temperatures in coal-derived syngas: results of initial screening of

candidate materials, Fuel Processing Technology. 85 (2004) 621-630.

150

[5] J.Y. Lee, Y. Ju, T.C. Keener, R.S. Varma, Development of cost-effective noncarbon sorbents

for Hg0 removal from coal-fired power plants, Environ. Sci. Technol. 40 (2006) 2714-2720.

[6] D. Karata, A. Lancia, D. Musmarra, F. Pepe, Adsorption of metallic mercury on activated

carbon, Symposium (International) on Combustion,. 26 (1996) 2439-2445.

[7] G. Skodras, I. Diamantopoulou, G. Pantoleontos, G.P. Sakellaropoulos, Kinetic studies of

elemental mercury adsorption in activated carbon fixed bed reactor, Journal of Hazardous

Materials. 158 (2008) 1-13.

[8] D. Karatza, A. Lancia, D. Musmarra, C. Zucchini, Study of mercury absorption and

desorption on sulfur impregnated carbon, Experimental Thermal and Fluid Science. 21 (2000)

150-155.

[9] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors affecting

mercury control in utility flue gas using activated carbon, Journal of the Air & Waste

Management Association. 48 (1998) 1166.

[10] R. Yan, Y.L. Ng, D.T. Liang, C.S. Lim, J.H. Tay, Bench-scale experimental study on the

effect of flue gas composition on mercury removal by activated carbon adsorption, Energy &

Fuels. 17 (2003) 1528-1535.

[11] R. Yan, D.T. Liang, L. Tsen, Y.P. Wong, Y.K. Lee, Bench-scale experimental evaluation of

carbon performance on mercury vapour adsorption, Fuel. 83 (2004) 2401-2409.

[12] G.E. Dunham, R.A. DeWall, C.L. Senior, Fixed-bed studies of the interactions between

mercury and coal combustion fly ash, Fuel Processing Technology. 82 (2003) 197-213.

[13] B.A.F. Mibeck, E.S. Olson, S.J. Miller, HgCl2 sorption on lignite activated carbon: Analysis

of fixed-bed results, Fuel Process Technol. 90 (2009) 1364-1371.

[14] G.E. Dunham, S.J. Miller, Mercury capture by an activated carbon in a fixed-bed bench-

scale system, Environmental Progress. 17 (1998) 203.

[15] R.J. Berger, J. Pérez-Ramírez, F. Kapteijn, J.A. Moulijn, Catalyst performance testing: the

influence of catalyst bed dilution on the conversion observed, Chem. Eng. J. 90 (2002) 173-183.

[16] R.J. Berger, J. Pérez-Ramírez, F. Kapteijn, J.A. Moulijn, Catalyst performance testing: bed

dilution revisited, Chemical Engineering Science. 57 (2002) 4921-4932.

[17] J.L. Stuart. Development of an experimental system to study mercury uptake by activated

carbon under simulated flue gas conditions, Master thesis, University of Pittsburgh, 2002.

[18] M. Holmes and J. Pavlish, Mercury information clearinghouse, Quarter 3- Advanced and

developmental mercury control technologies, July 2004.

[19] J.R. Butz, T.E. Broderick and C.S. Turchi, Amended Silicates™ for Mercury Control,

project final report, DOE Award Number: DE-FC26-04NT41988, 2006.

151

7

Screening tests of mercury sorbents

This chapter first deals with screening of mercury sorbents in simulated cement

kiln flue gas using elemental mercury source. Then screening tests using mercury

chloride source are reported. The results are used to suggest promising sorbents for

application in cement plant and provide explanation of mercury adsorption in cement

production by cement materials.

7.1 Introduction 

Sorbents are screened in the laboratory using simulated flue gas before field-

testing in actual flue gas. The purpose of these laboratory tests is to evaluate a number of

sorbents at conditions similar to those expected at typical cement plants. These test

results then are used to determine the most appropriate samples for large scale tests.

Basing on the screen tests, promising sorbents will be further investigated in the lab in

detail to obtain adsorption kinetics and study the influence of different operational

parameters. The fixed bed tests are not intended to simulate the conditions where a

sorbent is injected continuously upstream of a fabric filter but they provide a good

indication of sorbent effectiveness, providing the exposure conditions are similar.

Screening measurements are used to evaluate mercury capture effectiveness, oxidation

potential, and capacity for the selected sorbents.

16 sorbent materials are collected and compared. The selected sorbents are tested

in the fixed-bed reactor with continuous mercury measurement, following closely the

experimental procedure as described in chapter 4.

The empty reactor, quartz wool plug, and sand for sorbent dilution are tested first

to investigate whether there is some mercury adsorption by the empty reactor, quartz

152

wool and sand powder. Then the sorbents are tested in the simulated cement kiln flue gas

using either elemental mercury or mercury chloride source to study whether the sorbents

behavior differently using different mercury sources.

During mercury adsorption tests, the elemental Hg can be fully or partially

oxidized because of reactions between the elemental Hg, sorbent, and flue gas

components. Then the extent of mercury oxidation is calculated by comparing the

measured elemental mercury after breakthrough and the inlet level of added elemental

mercury.

The direct result of the fixed-bed test is the mercury adsorption breakthrough

curve. The percentage breakthrough is determined as a function of time by normalizing

the measured total mercury concentration at the outlet of the sorbent bed to the inlet

mercury concentration.

From the mercury breakthrough curve, the amount of mercury adsorbed on unit

mass of the sorbent as a function of time can be calculated from the expression:

,0( )

t

t in out t

Fq C C dt

W (7.1)

where F is the flow rate through the sorbent bed, W is the mass of the sorbent, Cin is the

inlet mercury concentration, Cout,t is the mercury concentration at the reactor outlet at

time t.

The mercury adsorption capacity of a known weight of a sorbent is calculated in

terms of µg Hg adsorbed/g_sorbent material from the breakthrough curve for the sorbent.

The area under the inlet mercury concentration line and a breakthrough curve is used to

determine how much mercury is adsorbed by the sorbent. By material balance, the area

between the curve and the line provides the information on the total mercury adsorbed

onto a sorbent if the entire bed reaches equilibrium with mercury vapor. The equilibrium

adsorption capacity is defined by the time when the outlet Hg concentration is first equal

to the inlet concentration.

Previous bench-scale studies have reported performance of the sorbents in terms

of the adsorption capacity and/or the time taken for complete breakthrough of Hg from a

sorbent [1-13]. However, the final application of the sorbent injection is in a full-scale

153

plant where the sorbent is injected in the duct and captured either in a fabric filter, where

the contact time of mercury with carbon particles is very short. Therefore, the adsorption

rate within these time scales is the most important parameter when evaluating sorbent

performance.

The amount of mercury adsorbed at time t (mt) can be calculated using

12,( ). .t t t in out t tm m C C F t (7.2)

where 2)( ,,, 21 ttouttoutttout CCC

The adsorption rate at each time step is calculated using

t t tt

m mrate

dt

(7.3)

The initial rate is evaluated as the slope of the cumulative adsorption curve in the

first 25 min.

7.2 Sorbent properties and compositions 

The collected sorbent candidates include both commercial sorbents and cement

materials. Virgin activated carbon Dacro Hg, formerly known as Darco FGD [14] is

prepared from lignite coal and has been widely studied in the literature [1,2,5-7,15-24]

and is therefore tested here as a reference sorbent. Darco Hg-LH is Darco Hg treated with

bromine and developed for application with low chlorine concentration in the flue gas

from combustion of low-rank coals. Activated Lignite HOK is produced according to the

so-called rotary-hearth furnace process [25-27]. Unlike activated carbon, activated

Lignite HOK is produced as mass product with an annual output of 200,000 tons at a

much lower price than that of activated carbon. HOK is the most widely used sorbent for

waste incinerator flue gas cleaning in Europe. Sorbalit is a mixture of reagents, surface-

active substances and chemical additives [28]. Reagents are calcium based compounds

such as CaCO3, CaO and Ca(OH)2. Examples of surface-active substances are activated

carbon, aluminum oxide and zeolite. Chemical additives are sulfur and sulfur compounds

such as Na2S, NaHS, Na2S4. Sorbalit can be produced with carbon contents ranging from

4% to 65%. Minsorb DM and ME are non-carbon based sorbent and for removal of

154

dioxin/furan and mercury, respectively [29,30]. The hydrated lime is standard Sorbacal

product used for SO2 and SO3 removal [31].

Cement materials are obtained from FLSmidth Dania lab. Clay contains

essentially hydrous aluminum silicates, with minor amount of magnesium, iron, alkalies

or alkaline earths [32]. Cement kiln dust is a fine-grained solid material with high

alkaline content removed from the cement kiln exhaust gas by filters. The cement kiln

dust contains mainly incompletely reacted raw material, including a raw mix at various

stages of burning, and particles of clinker. The primary constituents are silicates, calcium

oxide, carbonates, potassium oxide, sulfates, chlorides, various metal oxides, and sodium

oxide [33].

The kaolin sample is from Prolabo Merck. Hydroxyapatite has a formula of

Ca5(PO4)3(OH) and has been used as sorbent to removal of heavy metals from waste

incinerators [34]. Initial tests show that hydroxyapatite is a new promising sorbent for

heavy metal removal from waste incineration flue gas [34]. Hydroxyapatite is chemically

similar to the mineral component of bones and suitable for biomedical application.

Properties of the sorbents are presented in table 7.1. Carbon-based sorbents have

much larger surface area than the non-carbon based sorbents and cement raw materials.

The volume median diameter D(v,0.5) is the diameter where 50% of the distribution is

above and 50% is below. D(v,0.9) diameter means that 90% of the volume distribution is

below this value. Similarly D(v,0.1) diameter means that 10% of the volume distribution

is below this value. Generally the cement materials have a smaller particle size than the

commercial sorbents. The cement materials are the cheapest due to the availability of

large quantity in the cement plant and saving of transport cost. The bromine treated

Darco Hg-LH carbon is much more expensive than the virgin activated carbons and non-

carbon sorbents. The high price of hydroxyapatite is because it is pharmaceutical grade.

Table 7.1. Properties of sorbents studied in this work.

Sorbent D(v,0.1) m

D(v,0.5) m

D(v,0.9) m

BET area

Bulk density

Price USD/kg

155

m2/g g/cm3 Darco Hg 1.27 15.99 43.07 600 0.51 1-2 Darco Hg-LH 1.12 15.36 44.70 550 0.60 2-4 HOK standard 63 300 0.55 1-2 HOK super 24 300 0.44 1-2 Sorbalit 0.85 12.60 52.24 58.5 0.42 1-2 Minsorb DM 6.76 52.02 168.95 120 0.60 1-2 Minsorb ME 3.20 39.17 177.07 70 1.10 1-2 Hydrated lime 0.30 3.35 13.15 21.5 0.35 0.2 Saklei fly ash 3.77 36.15 115.07 0.7 - 0.1 Gypsum 1.63 18.78 62.20 18.5 - 0.1-0.15 Raw meal 0.30 9.47 77.56 1.8 - 0.1 Portland cement 0.32 16.16 46.10 1.8 - 0.1-0.2 Cement kiln dust 0.33 3.36 63.93 6.5 - 0.05 Clay 0.36 9.73 58.40 15.2 - 0.1 Kaolin 1.24 5.60 20.65 13.0 - 0.1-0.2 Hydroxyapatite 0.20 3.80 47.99 70.2 - 160

Table 7.2 presents chemical composition of some selected sorbents. The

compositions of HOK carbons and Minsorb sorbents are from the literature published by

the manufacture and the product datasheet [26,27,29,30]. The compositions of Darco

carbons and Sorbalit are obtained by averaging 10-20 spot analyses of the samples by

SEM-EDX. Compositions of other materials are obtained by inductively coupled plasma

(ICP) spectrometry. The Darco carbons have larger ash content than the HOK carbons.

SEM-EDX analyses show that the Darco Hg-LH has a bromine content of about 7.8 wt%.

The main elements of Sorbalit are C and Ca, in agreement with the statement by the

producer [28]. Minsorb ME has larger Al and Fe contents than Minsorb DM. Ca is the

main element in the raw meal. Kaolin and Saklei fly ash from bituminous coal

combustion have similar composition with large Al and Si contents.

156

Table 7.2 Chemical composition of selected sorbents. All in wt%

Sorbent Moisture Ash C Cl S K Na Mg Ca Fe Al Si Reference Darco Hg <8 32 65 0.1 1.5 - 0.2 0.8 5.2 0.9 0.5 6.6 [1] Darco Hg-LH <12 - 56 0.2 2.9 - 3.2 0.7 4.6 0.9 0.7 1.0 HOK standard 0.5 10 89 - 0.6 - 0.8 2.6 - - - [26,27] HOK super 0.5 10 89 - 0.6 - 0.8 2.6 - - - [26,27] Sorbalit - - 27 0.2 - 0.2 0.2 28 0.1 - - Minsorb DM <8 - - 3.0 0.2 0.8 0.3 12 5.7 2.1 6.4 23.3 [29] Minsorb ME <8 - - 2-5 0.1 0.5 0.1 1.8 2.1 14.0 18.6 18.7 [30] Saklei fly ash - - 2.5 - 0.1 0.5 0.1 0.6 3.5 2.1 18.1 23.4 Raw meal - - - - 0.4 0.6 0.2 1.2 31.4 1.6 1.7 6.3 Kaolin - - - - - 1.3 - 0.2 - 0.6 20.1 22.8

157

7.3 SEM­EDX analysis of fresh sorbents 

The main goals of the SEM-EDX analysis is to study the sorbents’ topography

(surface features), morphology (shape and size), and composition. Morphology study

will be used to identify particle agglomeration and compare with particle size

measurement.

Figure 7.1 shows a typical micrograph of the fresh Darco Hg carbon which

has various single carbon particles of irregular surface with different shape and

brightness. Close observation of the big particles at higher magnification shows that

there are many small floc-like particles agglomerated on the big particle. Images at

lower magnification (not shown in figure 7.1) show that most of the particles are

within the range of 5-30 m and this is in reasonable agreement with the particle size

measurements by laser diffraction. However, it should be noted that these SEM

pictures provide only semi-quantitative results of particle sizing since the technique

uses two-dimension information to infer a three-dimensional quantity

Figure 7.1. SEM micrographs of the fresh Darco Hg activated carbon at different

magnifications. Scale bar from left to right is 10 and 5 m, respectively.

Figure 7.2 illustrates the difference in information provided by secondary

electron (SE) image and backscattered electron (BSE) image. The SE image is

superior for displaying surface detail and particle morphology but does not generally

show chemical heterogeneity. EDX analysis shows that in the BSE image the small

bright spots in the left (area 7) have high iron content, the bright spot in the center on

158

the big particle (area 4) and big bright particle on the up-right corner (area 1) have

high silica content. Area 5 has high content of calcium and the particle is crystal-like.

The carbon particles have similar brightness level as the carbon substrate on the

carbon table and are not clearly seen in the BSE image.

Figure 7.2. SE (left) and BSE (right) images of the fresh Darco Hg sorbent at the

same location. Positions for SEM-EDX analysis are marked on the BSE image. Scale

bar is 30 m. As shown in figure 7.3 the morphology of the fresh Darco Hg-LH is very

similar to that of the fresh Darco Hg and this is not surprised since the Darco Hg-LH

is prepared from Darco Hg by a brominating process.

Figure 7.3. SE images of fresh Darco Hg-LH activated carbon. Scale bar is 5 m.

159

Compared to the fresh Darco Hg activated carbon there are high contents of

Na, S, and Br in the Darco Hg-LH sample. The average molar ratio of Na/Br is about

1.74, while the molar ratio of Na/(0.5S+Br) is about 1.18, suggesting the sample is

brominated by exposing to NaBr and Na2SO4/Na2SO3/Na2S compounds instead of to

HBr or Br2.

The SE images of fresh Sorbalit sorbent are presented in figure 7.4. The

particles are much less porous than the carbon particles. A thin layer of small crystal-

like flakes agglomerate on the big particles. The small dots on the background are

from the carbon table for sample holding.

Figure 7.4. SE images of fresh Sorbalit sorbent. Scale bar is 5 m.

The SE images of the fresh Minsorb ME sorbent are shown in figure 7.5. The

particle size is generally lager than carbon particle size and in agreement with the

particle size measurement by the laser diffraction.

160

Figure 7.5. SE images of fresh Minsorb ME sorbent. Scale bar is 50 m.

7.4 Baseline test 

As a starting point, baseline tests of the empty glass reactor, quartz wool plug

and sand powder are conducted first to investigate whether mercury can be adsorbed

by these parts and materials. Tests are conducted in both nitrogen and simulated

cement kiln flue gas with either elemental mercury or mercury chloride sources. In all

cases, simultaneous mercury breakthroughs are observed indicating no mercury

adsorption is adsorbed by these materials. The mercury exposed sand powder is

analyzed for mercury content in the sample. No mercury is detected in the exposed

sand, which again verifies that no mercury adsorption by the sand takes place.

7.5 Screening tests in nitrogen 

Preliminary tests of some sorbents were conducted by mixing 5-10 mg

sorbent with 2 g sand in nitrogen using elemental mercury source. Only elemental

mercury was measured due to the fact that the problem of converter for total mercury

measurement was not solved at that time. Tests at 150C show that instantaneous

mercury breakthrough was observed for all the sorbents except the bromine treated

Darco Hg-LH carbon. As shown in figure 7.6, even in nitrogen the bromine treated

161

Darco Hg-LH carbon can both oxidize and adsorb some mercury. Compared to

instantaneous mercury breakthrough observed by the non-treated Darco Hg carbon,

mercury adsorption by the Darco Hg-LH carbon is due to the promoting effects of

bromine in the Darco Hg-LH carbon. Part of the mercury is probably oxidized on the

Darco Hg-LH carbon by the bromine compounds. However, most of the mercury is

still in the form of elemental mercury. Figure 7.6 also shows the breakthrough curve

of 10 mg Darco Hg tested in nitrogen at 150C with HgCl2 source and total mercury

measurement. In contrast to test using elemental mercury source, it takes about 15 h

to reach the breakthrough. These tests indicate that mercury oxidation is an important

step during mercury adsorption by the sorbent. Elemental mercury needs to be

oxidized first either in the gas phase or on the sorbent before being adsorbed by the

sorbent.

0 2 4 6 8 10 12 14 16 18

Time (hour)

0

0.4

0.8

1.2

Gas

eou

s H

g, C

ou

t/Cin

12

1, Hg0 source, 230 g/Nm3, 5 mgDarco Hg-LH, Hg0 measurement2, HgCl2 source, 209 g/Nm3, 10 mgDarco Hg, Hgtotal measurement

Figure 7.6. Mercury breakthrough curves at 150 C for 5 mg Darco Hg-LH carbon

tested in N2 with elemental source and elemental mercury measurement and 10 mg

Darco Hg carbon tested in N2 with HgCl2 source and total mercury measurement. 2 g

sand as bed mixing material.

162

7.6  Screening  tests  in  simulated  cement  kiln  flue  gas  with 

elemental mercury source 

Total mercury measurement was conducted using the sulfite-based converter

to obtain mercury breakthrough curves using elemental mercury source in the

simulated cement kiln flue gas. Figure 7.7 illustrates the screening results of 30 mg

different sorbents in 2 g sand at 150C. From the mercury breakthrough curves, the

amount of adsorbed mercury by the sorbent and the average initial adsorption rate for

the first 25 min are calculated and presented in table 7.3. The extents of mercury

oxidation by different sorbents are illustrated in figure 7.8.

0 1 2 3 4 5 6

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Co

ut/C

in

A B C DE

F

A:Minsorb ME, 169g/Nm3 Hg0

B: Darco Hg, 180g/Nm3 Hg0

C: HOK super, 167g/Nm3 Hg0

D: Sorbalit, 164g/Nm3 Hg0

E: HOK standard, 171g/Nm3 Hg0

F: Darco Hg-LH, 167g/Nm3 Hg0

Figure 7.7. Mercury breakthrough profiles for 30 mg sorbets in 2 g sand tested at

150C in simulated cement kiln flue gas with 164-180 µg Hg0/Nm3, 1000 ppmv NO,

23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6

vol.% O2.

163

0

20

40

60

80

100

Me

rcu

ry o

xid

ati

on

(%

)

Dar

co H

g 1

6 m

Dar

co H

g-L

H 1

5 m

HO

K S

up

er, 2

4 m

HO

K S

tan

dar

d 6

3

m

So

bal

it s

up

er 1

3 m

Min

sorb

M

E 4

0 m

Figure 7.8. Percentages of mercury oxidation by 30 mg sorbets in 2 g sand tested at

150C in simulated cement kiln flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO,

23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6

vol.% O2.

Table 7.3. Mercury 99% breakthrough time, adsorbed mercury and initial adsorption

rates for 30 mg sorbets in 2 g sand tested at 150C in simulated cement kiln flue gas

with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv

SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

Sorbent 99% breakthrough

time (min)

Adsorbed Hg (µg Hg/g_sorbent)

Initial adsorption rate (µg Hg/g_sorbent/h)

Hydroxyapatite 0 0 0 Minsorb DM 0 0 0 Minsorb ME 28 31 70

Sorbalit 110 270 310 HOK super 111 632 690

HOK standard 164 699 560 Darco Hg 90 726 890

Darco Hg-LH 320 1305 690

164

Neither mercury adsorption nor oxidation is observed by 30 mg non-carbon

based sorbents Minsorb DM, hydroxyapatite, and cement materials at 150C. The

bromine treated carbon Darco Hg-LH has larger adsorption capacity but smaller

adsorption rate compared to the non-treaded Darco Hg carbon. As shown in figure

7.8 and table 7.3, there is a clear trend between the extent of mercury oxidation and

amount of adsorbed mercury. Generally larger amount of adsorbed mercury is

obtained with sorbents that have larger mercury oxidation capacity. The initial

adsorption rate of coarse HOK standard carbon is slightly smaller than the fine HOK

super due to the larger diffusion resistance within the larger carbon particles. Sorbalit,

which is a mixture of lime and carbon, shows poorer performance than the carbons.

Minsorb ME, which is aluminumsilicates based sorbent shows the poorest

performance among the tested commercial sorbents despite that it has much larger

surface area than the Sorbalit sorbent. This is probably due to its capacity for mercury

oxidation is much smaller than the Sorbalit sorbent.

The adsorption of mercury in the Darco Hg carbon is attempted by analyzing

the mercury content in the exposed carbon sample. Table 7.4 compares the measured

and calculated mercury contents in the carbons from the breakthrough curve. The

calculated mercury contents are much larger than the measured values for the carbon

and sand mixtures. The analysis of mercury content in the sample uses only 100 mg

of the sample for analysis and one reason for the disagreement could be that the

sample analyzed might not be representative. Only 30 mg carbon is mixed with 2 g

sand and the carbon may separate from the sand. This is often observed during

loading the sample to the reactor. To ensure most of the carbon is loaded to the

reactor, the sample holder is shaken to remove the carbon deposited on the sample

holder and a big quartz wool plug is used to clean carbon deposited on the reactor

wall and move the carbon to the fixed-bed bed. The carbon particle might deposit on

the container wall and therefore the analyzed sample could contain relatively more

sand powder. The mercury content in the carbon is calculated from the measured

mercury level in the carbon-sand mixture and the carbon-sand mixing ratio. Since no

mercury adsorbed by the sand powder, analysis using non-representative carbon-sand

mixture could result in small mercury content in the carbon.

165

Table 7.4. Comparison of measured and calculated mercury contents in the carbons

from the breakthrough curve. Flue gas composition: 141-183 µg Hg0/Nm3, 1000

ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.%

CO2, and 6 vol.% O2.

Sorbent description Measured Hg in carbon (ppmm)

Calculated results from breakthrough (ppmm)

30 mg Darco Hg in 2g sand, 150°C, 141 µg/Nm3 Hg in flue gas

2.22 5.96

30 mg Darco Hg in 2g sand, 150°C, 183 µg/Nm3 Hg in flue gas

2.27 13.14

30 mg Darco Hg in 2g sand, 150°C. 143 µg/Nm3 Hg in flue gas

1.49 7.06

500 mg Darco Hg, 200°C, 160 µg/Nm3 Hg in flue gas

46.79 50.87

To check whether the method for calculating the mercury content in the

carbon is reasonable, a new test was performed by using only carbon sample to avoid

the problem of non-representative sample caused by carbon-sand mixing. As shown

in table 7.4, the measured mercury content in the carbon is about 92% of the

calculated value from the breakthrough curve. This reasonable agreement between the

measured and calculated value confirms that the method of calculating mercury

content in the carbon from the breakthrough curve works to a satisfactory extent.

To be able to observe some mercury adsorption by the cement materials, the

adsorption temperature was decreased to 75C. However, still no mercury adsorption

was observed by 30 mg cement materials at 75C. Then the sorbent load is increased

to 2 g. Among the tested cement materials only raw meal shows some mercury

adsorption as shown in figure 7.9. This can to some extent explain the low mercury

emission from cement plants during raw mill-on period. The dust load in the flue gas

after the raw mill could be up to 800-1000 g/m3 and therefore noticeable amount of

mercury could be adsorbed by the raw meal both in and after the raw mill.

166

0 30 60 90 120 150

Time (min)

0

20

40

60

80

100

120

140

160

180

Gas

eou

s H

g (g

/Nm

3 )

0

0.0004

0.0008

0.0012

0.0016

Cal

cula

ted

Hg

in s

orb

ent

(mg

Hg

/g_s

orb

ent)

Gaseous Hgtotal

Calculated Hg in sorbent

bypassreactor through reactor

Figure 7.9. Mercury breakthrough profile and calculated mercury adsorption in 2 g

cement raw meal tested at 75C in simulated cement kiln flue gas with 160-170 µg

Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.%

H2O, 21 vol.% CO2, and 6 vol.% O2.

7.7 Screening  tests  in simulated cement kiln  flue gas with HgCl2 

source 

The collected sorbents are also tested in simulated cement kiln flue gas with

HgCl2 source. Figure 7.10 shows the mercury breakthrough curves for 10 mg

sorbents in 2 g sand at 150C using simulated cement kiln flue gas with170±10

µg/Nm3 mercury from HgCl2 source. The 99% breakthrough time, calculated amount

of adsorbed mercury by the sorbent from the breakthrough curve, and the average

initial adsorption rate for the first 25 min are presented in table 7.5.

167

Figure 7.10. Mercury breakthrough profiles of 10 mg sorbents tested in 2g sand at

150C in simulated cement kiln flue gas with HgCl2 source. The inlet mercury level

is 170±10 µg/Nm3, other gases include 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl,

1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

Table 7.5. Mercury 99% breakthrough time, adsorbed mercury and initial adsorption

rates for 10 mg sorbets in 2 g sand tested at 150C in simulated cement kiln flue gas

using HgCl2 source. The inlet mercury level is 170±10 µg/Nm3, other gases include

1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21

vol.% CO2, and 6 vol.% O2.

Sorbent 99%

breakthrough time (min)

Adsorbed Hg (µg Hg/g_sorbent)

Initial adsorption rate (µg Hg/g_sorbent/h)

Hydroxyapatite 0 0 - Minsorb DM 5 12 - Minsorb ME 49 363 780

Sorbalit 86 429 730 HOK standard 74 1021 1390

HOK super 55 1153 2170 Darco Hg 43 1224 2510

Darco Hg-LH 52 1290 2510

168

The hydroxyapatite sorbent still does not adsorb any mercury even using the

mercury chloride source. This means that the mercury adsorption by hydroxyapatite

is not only limited by its ability of oxidizing mercury, but also other properties. The

Minsorb DM sorbent shows low adsorption of HgCl2 compared to no adsorption of

elemental mercury. Minsorb ME and Sorbalit show similar mercury adsorption in

terms of mercury adsorption capacity and initial adsorption rate. All the carbons show

similar mercury adsorption capacity; while the HOK standard has the smallest initial

adsorption rate. Compared to similar initial adsorption rate of elemental mercury, the

initial adsorption rate of HgCl2 for HOK super is about 50% larger than the HOK

standard. The elemental mercury adsorption capacity of Darco Hg-LH is about 79%

larger and initial Hg0 adsorption rate is about 23% smaller in comparison with the

virgin Darco Hg carbon. Similar HgCl2 adsorption capacity and initial adsorption rate

of Darco Hg and Darco Hg-LH indicate that Darco Hg is a better choice at least for

removing HgCl2 from cement kiln flue gas.

Cement materials were also tested for HgCl2 capture from simulated cement

kiln flue gas using a sorbent load of 2 g without mixing with sand powder. Figure

7.11 presents the mercury breakthrough curves of 2 g cement materials tested at

150C using simulated cement kiln flue gas with170±10 µg/Nm3 mercury from

HgCl2 source. The 99% breakthrough time and calculated amount of adsorbed

mercury by the sorbent from the breakthrough curve are given in table 7.7. The

fluctuation of some breakthrough curves is due to the aging of the sulfite-based

converter material. After changing the converter material used for about 3 months

smooth mercury breakthrough is obtained again.

169

Figure 7.11. Mercury breakthrough profiles of 2 g cement materials tested at 150C

in simulated cement kiln flue gas with HgCl2 source. The inlet mercury level is

170±10 µg/Nm3, other gases include 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl,

1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. Data of 10 mg Darco

Hg carbon in 2 g sand are shown for comparison.

Table 7.7. Mercury breakthrough time and adsorbed mercury for 2 g cement materials

tested at 150C in simulated cement kiln flue gas using HgCl2 source. The inlet

mercury level is 170±10 µg/Nm3, other gases include 1000 ppmv NO, 23 ppmv NO2,

10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

Sorbents Breakthrough time (min) Adsorbed Hg (µg Hg/g_sorbent)

Hydrated lime 11 0.60 Clay 14 0.90

Kaolin 23 1.77 Cement kiln dust 60 1.81

Gypsum 49 1.73 Raw meal 54 0.90

Saklei fly ash 38 1.01 Portland cement 60 2.28

0 0.2 0.4 0.6 0.8 1

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Gas

eou

s H

g (

Co

ut/C

in)

A BC

FE

D

A: 2 g hydrated limeB: 2 g clayC: 2 g kaolinD: 10 mg Darco Hg in 2 g sandE: 2 g cement kiln dustF: 2 g gypsum

0 0.2 0.4 0.6 0.8 1

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Gas

eou

s H

g (

Co

ut/C

in)

G

H

I

D

D: 10 mg Darco Hg in 2 g sandG: 2 g raw mealH: 2 g Portland cementI: 2 g Saklei fly ash

170

Compared to null adsorption of elemental mercury at temperature as low as

75C, all the tested cement materials show some adsorption of mercury chloride

at150C. The lack of ability for mercury oxidation is probably the main limitation for

these materials to be used for mercury removal from flue gas, which contains both

elemental and oxidized mercury. 2 g of kaolin, cement kiln dust, gypsum, and

Portland cement adsorb similar amount of mercury chloride as 10 mg Minsorb ME

and Sorbalit commercial sorbents. Raw meal, hydrated lime, clay, Saklei fly ash have

similar mercury adsorption capacities which are about half of the capacities of kaolin,

cement kiln dust, gypsum, and Portland cement. These results can further explain low

mercury emission from the cement plant during raw mill-on period, as about 55-65 %

of the mercury in the cement kiln flue gas is oxidized mercury [35]. Considering the

low cost and abundance of the cement materials, injection of cement materials for

mercury control in cement plant is feasible provided that the elemental mercury in the

flue gas can be oxidized by adding an oxidant. However, raw meal, hydrated lime,

clay, cement kiln dust, and kaolin have to be recycled to the kiln in the cement

production and the adsorbed mercury will be released again in the hot zone. If these

materials are not recycled the disposal cost will be high since larger amount of these

materials have to be used compared to activated carbon for same amount of mercury

removal. Gypsum and Saklei fly ash can be added to the final cement product and the

release of captured mercury and high disposal cost are avoided. This also applied for

the Portland cement. However, the stability of mercury in the final cement product

requires further investigation.

7.8 Conclusions 

Screening tests of sorbents for mercury removal from cement plants have

been conducted in the fixed-bed reactor system. The tested sorbents include

commercial activated carbons, commercial non-carbon sorbents, and cement

materials. Screening measurements are used to evaluate initial mercury capture rate,

oxidation potential, and capacity for the selected sorbents. The amount of mercury

adsorbed is calculated from the mercury breakthrough curve and the initial mercury

171

adsorption rate is further evaluated for application regarding sorbent injection

upstream of a fabric filter.

Baseline tests of empty reactor, quartz wool plug, and sand powder show that

no mercury adsorption is observed either in nitrogen or simulated cement kiln flue

gas with elemental mercury or mercury chloride source.

Initial tests of sorbent in nitrogen with elemental mercury at 150C find that

only the bromine treated Darco Hg-LH activated carbon shows some mercury

adsorption among the collected sorbents. However, the virgin Darco Hg carbon

adsorbs mercury chloride in nitrogen. This indicates that mercury oxidation is an

important factor for mercury adsorption by the sorbents. Elemental mercury needs to

be oxidized either in the flue gas or on the sorbent.

Tests a collection of sorbents (30 mg in 2 g sand) at 150C in simulated

cement kiln flue gas with elemental mercury show that no mercury adsorption or

oxidation takes place on the non-carbon based sorbents Minsorb DM, hydroxyapatite,

and cement materials. The mercury adsorption capacity of bromine treated carbon

Darco Hg-LH is 79% larger than the non-treated Darco Hg carbon, but the initial

adsorption rate is 23% smaller. Generally a larger amount of adsorbed mercury is

obtained with sorbents that have larger mercury oxidation capacity. A lower amount

of mercury is adsorbed by the HOK carbon compared to Darco Hg carbon, probably

be due to both the smaller surface area and mercury oxidation capacity of the HOK

carbon. The initial adsorption rate of coarse HOK standard carbon is slightly lower

than the fine HOK super due to the larger diffusion resistance within the larger HOK

standard carbon particles. Sorbalit shows poorer performance than the carbons, while

Minsorb ME shows the poorest performance among the tested commercial sorbents

despite that it has much larger surface area than the Sorbalit sorbent. Among the

tested sorbents Darco Hg has the largest initial adsorption rate of elemental mercury.

The collected sorbents are also tested in simulated cement kiln flue gas with

mercury chloride using 10 mg sorbents in 2 g sand at 150C. The hydroxyapatite

sorbent still does not adsorb any mercury. The Minsorb DM sorbent shows negligible

adsorption of HgCl2 compared to no adsorption of elemental mercury. Minsorb ME

and Sorbalit show similar mercury adsorption of about 400 µg Hg/g_sorbent. All the

172

carbons show similar mercury adsorption capacity; while the HOK standard has the

smallest initial adsorption rate. Similar HgCl2 adsorption capacity and initial

adsorption rate of Darco Hg and Darco Hg-LH indicate that Darco Hg is a better

choice at least for removing HgCl2 from cement kiln flue gas.

Compared to non-observable adsorption of elemental mercury on 30 mg

sample at temperature as low as 75C, all the tested cement materials show some

adsorption of mercury chloride at 150C using a sorbent load of 2 g. Similar amount

of mercury chloride adsorption is observed by 2 g of kaolin, cement kiln dust,

gypsum, Portland cement, and 10 mg Minsorb ME, Sorbalit commercial sorbents.

Among the tested sorbents the Darco Hg activated shows the best

performance of adsorption of both elemental and oxidized mercury, with the largest

initial adsorption rate and second largest mercury adsorption capacity and a lower

price than the treated carbon. Therefore, the Darco Hg carbon is recommended as the

reference sorbent for a fundamental investigation of mercury adsorption in simulated

cement kiln flue gas and large-scale tests. Adsorption by cement materials at larger

load can explain the phenomena of low mercury emission from the cement plant

during raw mill-on period, when larger amount of cement materials are present in the

flue gas at relative low temperature. Considering the low cost and abundance of the

cement materials, injection of cement materials for mercury control in cement plant is

feasible provided that the elemental mercury in the flue gas can be oxidized by

adding of oxidant. Compared to raw meal, clay, kaolin, and cement kiln dust, gypsum,

Saklei fly ash, and Portland cement are more preferred due to avoidance of captured

mercury release in the kiln and high disposal cost since these materials will be added

to the finished cement product. However, the stability of mercury in the exposed

cement materials requires further investigation to study whether it will be released

from the final cement product.

7.9 References 

[1] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R. Chang,

Preparation and evaluation of coal-derived activated carbons for removal of mercury vapor

from simulated coal combustion flue gases, Energy & Fuels. 12 (1998) 1061-1070.

173

[2] M. Rostam-Abadi, S.G. Chen, H.C. Hsi, M. Rood, R. Chang, T.R. Carey, B. Hargrove, C.

Richardson, W. Rosenhoover, F. Meserole, Novel vapor phase mercury sorbents,

Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant Control, Washington,

DC, Aug 25–29, 1997.

[3] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang and F.B. Meserole, Factors

affecting mercury control in utility flue gas using sorbent injection, Proceedings of the Air &

Waste Management Association's 90th Annual Meeting, Toronto, Ontario, Canada, June 8-13,

1997.

[4] C.X. Hu, J.S. Zhou, Z.Y. Luo, S. He, G.K. Wang, K.F. Cen, Effect of oxidation treatment

on the adsorption and the stability of mercury on activated carbon, Journal of Environmental

Sciences-China. 18 (2006) 1161-1166.

[5] B. Ghorishi and B.K. Gullett, Fixed-bed control of mercury: Role of acid gases and a

comparison between carbon-based, calcium-based, and coal fly ash sorbents, Proceedings of

the EPRI/DOE/EPA Combined Utility Air Pollutant Control Symposium, Washington, DC,

August 25–29, 1997.

[6] S.B. Ghorishi and C.B. Sedman, Combined mercury and sulfur oxides control using

calcium-based sorbents, Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant

Control Symposium, Washington, DC, August 25–29, 1997.

[7] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption by activated

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy

Conference, Research Triangle Park, NC, 22-25 April, 1997.

[8] N. Hutson, C. Singer, C. Richardson, J. Karwowski and C. Sedman, Practical applications

from observations of mercury oxidation and binding mechanisms, EPRI-DOE-EPA-AWMA

Combined Power Plant Air Pollutant Control Mega Symposium, Washington DC, August 30

- September 2, 2004.

[9] E.J. Granite, M.C. Freeman, R.A. Hargis, W.J. O'Dowd, H.W. Pennline, The thief process

for mercury removal from flue gas, J. Environ. Manage. 84 (2007) 628-634.

[10] W.J. O’Dowd, H.W. Pennline, M.C. Freeman, E.J. Granite, R.A. Hargis, C.J. Lacher, K.

Andrew, A technique to control mercury from flue gas: The thief process, Fuel Processing

Technology. 87 (2006) 1071-1084.

[11] R. Bhardwaj. Impact of temperature and flue gas components on mercury speciation and

uptake by activated carbon sorbents . Master thesis. University of Pittsburgh, 2007.

[12] M.M. Maroto-Valer, Y. Zhang, E.J. Granite, Z. Tang, H.W. Pennline, Effect of porous

structure and surface functionality on the mercury capacity of a fly ash carbon and its

activated sample, Fuel. 84 (2005) 105-108.

[13] S. Eswaran, H.G. Stenger, Z. Fan, Gas-phase mercury adsorption rate studies, Energy &

Fuels. 21 (2007) 852-857.

[14] J.R. Butz, T.E. Broderick and C.S. Turchi, Amended Silicates™ for Mercury Control,

project final report, DOE Award Number: DE-FC26-04NT41988, 2006.

174

[15] J.H. Pavlish, E.A. Sondreal, M.D. Mann, E.S. Olson, K.C. Galbreath, D.L. Laudal, S.A.

Benson, Status review of mercury control options for coal-fired power plants, Fuel

Processing Technology. 82 (2003) 89-165.

[16] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors

affecting mercury control in utility flue gas using activated carbon, Journal of the Air &

Waste Management Association. 48 (1998) 1166.

[17] N. Hutson, C. Singer, C. Richardson and J. Karwowski, Practical applications from

observations of mercury oxidation and binding mechanisms, The fifth mega symposium on

air pollutant controls for power plants, Washington, DC, August, 2004.

[18] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for DTE

Energy’s Monroe Station, DOE Award Number DE-FC26-03NT41986, Report Number

41986R16, 2006.

[19] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for

Sunflower Electric’s Holcomb Station, DE-FC26-03NT41986, Topical Report No. 41986R07,

2005.

[20] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for

AmerenUE’s Meramec Station Unit 2, DE-FC26-03NT41986, Topical Report No. 41986R09,

2005.

[21] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for

Basin Electric Power Cooperative’s Laramie River Station, DE-FC26-03NT41986, Topical

Report No. 41986R11, 2006.

[22] ADA-ES, Field test program to develop comprehensive design, operating, and cost data

for mercury control systems: Final site report for E.C. Gaston Unit 3 sorbent injection into

COHPAC for mercury control, DE-FC26-00NT41005, Topical Report No. 41005R11, 2003.

[23] ADA-ES, Field test program to develop comprehensive design, operating, and cost data

for mercury control systems: Final site report for Brayton Point Generating Station Unit 1

sorbent injection into a cold-side ESP for mercury control, DE-FC26-00NT41005, Topical

Report No. 41005R21, 2005.

[24] ADA-ES, Field test program to develop comprehensive design, operating, and cost data

for mercury control systems: Final site report for PG&E NEG Salem Harbor Station Unit 1

sorbent injection into a cold-side ESP for mercury control, U.S. DOE Cooperative Agreement

No. DE-FC26-00NT41005, 2004.

[25] J. Wirling, Implementation of process-integrated waste gas cleaning using activated

lignite, A&WMA specialty conference on hazardous waste combustors, Kansas City, Kansas,

March 28-30, 2001, .

[26] J. Wirling, Adsorptive waste gas cleaning in an industrial-scale coal-fired power plant,

A&WMA specialty conference on mercury emissions, Chicago, Illinos, August 21-23, 2001, .

[27] J. Wirling, J. Jablonski, Safety aspects in the use of carbonaceous sorbents during waste

gas treatment, Metallurgical Plant and Technology. 3 (2007) 144.

175

[28] A. Licata, M. Babu and L. Nethe, Acid gases, mercury, and dioxin from MWCs,

National Waste Processing Conference Proceedings ASME, Boston, Massachusetts, June 5-8,

1994.

[29] Lhoist Group, Data sheet of Minsorb® DM sorbent, 2010.

[30] Lhoist Group, Data sheet of Minsorb® ME sorbent, 2010.

[31] Lhoist Group, Sorbacal flue gas treatment, http://www.sorbacal.com/, accessed January

25, 2011.

[32] J.I. Bhatty, F.M. Miller, S.H. Kosmatka, (Eds.), Innovations in Portland cement

manufacturing, Portland Cement Association, Skokie, Illinois, U.S.A, 2004.

[33] U.S. EPA, Materials characterization paper in support of the proposed rulemaking:

Identification of nonhazardous secondary materials that are solid waste-cement kiln dust

(CKD), 2010.

[34] C. Verwilghen, S. Rio, J. Ramaroson, A. Nzihou and P. Sharrock, The use of

hydroxyapatite for the removal of heavy metals from industrial flue gas PART B:

Investigation in pilot scale, DustConf 2007, Maastricht, The Netherlands, April 23-24, 2007.

[35] R.J. Schreiber and C.D. Kellett, Compilation of mercury emission data, PCA R&D

Serial No. SN3091, 2009.

176

8

Fundamental investigation of elemental

mercury adsorption by activated carbon in

simulated cement kiln flue gas

This chapter reports a fundamental investigation of elemental mercury

adsorption by Darco Hg activated carbon in simulated cement kiln flue gas. The

investigation includes the effects of temperature and gas composition on mercury

adsorption kinetics and equilibrium uptake.

8.1 Introduction 

The investigation is mainly conducted using 10 mg Darco Hg mixed with 2 g

sand in simulated cement flue gas of a baseline composition of 1000 ppmv NO, 23

ppmv NO2, 1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 21 vol.% CO2, 6 vol.% O2.

Table 8.1 shows the investigated parameters and the tested ranges. The idea of using

wide range of parameters is to simulate possible cement kiln flue gas composition

and derive kinetics correlations that can be used to predict mercury adsorption by

activated carbon under different conditions. The isotherms and kinetics are obtained

using different adsorption temperatures, mercury inlet levels, and gas composition.

The percentages of mercury oxidation are also investigated by measuring the

elemental mercury level after the complete mercury breakthrough is obtained and

comparison with inlet elemental mercury concentration.

Another commercial carbon, Norit RB4, is also investigated to the study the

effects of carbon particle size. The granular Norit RB4 has a diameter of 4 mm and

length of 10 mm, respectively. The Norit RB4 pellet has a surface area of 1060-1320

m2/g, a microporous volume of 0.41-0.54 cm3/g, 0.8% of water, and 5.6% of ash [1,2].

The pellets are crushed and sieved for studying the effects of particle size.

177

Table. 8.1. Parameters for lab-scale fundamental investigation of elemental mercury

adsorption by the activated carbon.

Parameters Baseline values Range tested Flue gas rate (Nl/min) 2.75 1.1-2.75 Adsorption temperature (C) 150 75-250 Gas composition Hg0 (µg/Nm3) 160-170 0-170 NO (ppmv) 1000 100-1000 NO2 (ppmv) 23 0-100 SO2 (ppmv) 1000 100-1000 HCl (ppmv) 10 0-20 CO (ppmv) 0 0-1000 H2O (vol.%) 1 0-15 CO2 (vol.%) 21 1-31 O2 (vol.%) 6 1-16

8.2 Effect of adsorption temperature 

Figure 8.1 shows the effect of adsorption temperature on mercury

breakthrough profiles after the carbon bed. As expected, faster mercury breakthrough

is obtained at higher adsorption temperature. This pronounced effect of temperature

on the mercury adsorption capacity of the activated carbon evidences a physical

adsorption mechanism between the mercury and Darco Hg carbon. Physical

adsorption from the gas phase is accompanied by a decrease in free energy of the

system [3,4]. The gaseous molecules in the adsorbed state have fewer degrees of

freedom than in the gaseous state. This results in a decrease in entropy during

adsorption. Using the thermodynamic relationship:

G H T S (8.1)

It follows that the term ΔH, which is the heat of adsorption, must be negative

indicating that adsorption is always an exothermic process, respective of the nature of

the forces involved in the adsorption process.

178

0 0.5 1 1.5 2 2.5 3 3.5 4

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Gas

eou

s H

g, C

ou

t/Cin

1 2 3

1: 150oC2:120oC3:75oC

Figure 8.1. Effect of adsorption temperature on mercury breakthrough of 10 mg

Darco Hg mixed with 2 g sand using 2.75 Nl/min simulated flue gas with 160-170 µg

Hg0/Nm3,1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.%

H2O, 21 vol.% CO2, and 6 vol.% O2.

The effect of temperature on the extent of mercury oxidation by the Darco Hg

carbon is present in figure 8.2. The adsorption temperature does not affect the

oxidation of mercury by the Darco Hg carbon. The mercury oxidation is always

larger than 92% and the average mercury oxidation percentage is about 97% in the

studied temperature range of 75-250C.

179

50 100 150 200 250 300

Temperature (0C)

0

10

20

30

40

50

60

70

80

90

100

110

Mer

cury

oxi

dat

ion

(%

)

Figure 8.2. Effect of adsorption temperature on mercury oxidation by 10 mg Darco

Hg mixed with 2 g sand using 2.75 Nl/min simulated flue gas with 160-170 µg

Hg0/Nm3,1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.%

H2O, 21 vol.% CO2, and 6 vol.% O2.

8.3 Isotherm tests  

In order to simulate the performance of a given sorbent, the adsorption

equilibrium information such as the isotherm and characteristics of the sorbent must

be known. The adsorption isotherm is the most extensively employed method for

representing the equilibrium states of an adsorption system [3,4]. It can give useful

information regarding the adsorbate, the adsorbent, and the adsorption process. It

helps in the determination of the heat of adsorption, and the relative absorbability of a

gas on a given adsorbent.

Sorbent equilibrium data can be generated by conducting adsorption

breakthrough tests in the fixed-bed reactor. Figure 8.3 illustrates the mercury

breakthrough curves of 10 mg Norit Hg activated carbon tested at 120C with

different elemental mercury inlet levels. The time necessary for saturation of 10 mg

carbon is in the order of 0.6-1.2 h for the elemental mercury inlet level of 27-95

µg/Nm3. It takes longer time to reach the complete breakthrough when the mercury

inlet level is lower. This is in agreement with the observation by Karatza et al. [5] that

180

the saturation time decreased when the inlet mercury level was increased from 1 to

5.5 mg/m3. The driving force of mercury adsorption is the difference between the

amount of adsorbed mercury by unit carbon at a particular mercury inlet

concentration and the theoretical amount of mercury that could be adsorbed by unit

carbon at that concentration and this driving force disappears when the adsorption

gradually approaches its equilibrium state. Initially the rate of adsorption is large as

the whole carbon surface is bare but as more and more of the surface becomes

covered by the mercury molecules, the available bare surface decreases and so does

the rate of adsorption. The driving force theory can therefore explain the sigmoidal

shape of the breakthrough curve. The driving force for higher mercury inlet level is

larger at the initial stage of the adsorption (the first 12 min as shown in figure 8.3)

and becomes smaller and similar for all the applied mercury inlet levels due to the

accumulation of mercury in the carbon. As a result, faster mercury breakthrough is

obtained for larger mercury inlet concentration.

0 0.2 0.4 0.6 0.8 1 1.2

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Ga

seo

us

Hg

, C

ou

t/Cin

1 2 3

1: 95 g/Nm3 Hg0

2: 57 g/Nm3 Hg0

3: 27 g/Nm3 Hg0

Figure 8.3. Effect of elemental mercury inlet level on mercury breakthrough of 10 mg

Darco Hg mixed with 2 g sand tested at 120C using 2.75 Nl/min simulated flue gas

with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21

vol.% CO2, and 6 vol.% O2. The isotherm studies are conducted at 75, 100, 120, and 150C. As shown in

figure 8.4 there is a linear correlation between the amounts of mercury adsorbed on

the unit mass of carbon and the inlet mercury concentrations. For the mercury levels

181

applied in this work (<200 µg/Nm3), the adsorption isotherm follows the Henry’s law,

which corresponds physically to the situation where the adsorbed phase is so dilute

that there is neither competition for adsorption sites nor interaction between adsorbed

molecules. At higher loadings both these effects become significant, leading to

curvature of the equilibrium isotherm and variation of the heat of adsorption with

loading [6]. Ho’s work [7] shows that when the mercury inlet level is under 500

µg/m3 the adsorption isotherm follows Henry’s law and at higher mercury inlet levels

the isotherm follows Langmuir equation. Similarly, Karatza et al. [5] showed that the

isotherm follows Langmuir equation and of the favorable kind when 1-5.5 mg/m3

mercury inlet levels were applied.

0 20 40 60 80 100 120 140

Inlet gaseous Hg0 (g/m3)

0

0.4

0.8

1.2

1.6

2

2.4

Ad

sorb

ed H

g (g

Hg

/mg

_car

bo

n)

75oC,Y=0.01755X,R2=0.99100oC,Y=0.01259X,R2=0.97120oC,Y=0.01170X,R2=0.99150oC,Y=0.01024X,R2=0.97

Figure 8.4. Isotherms at 75, 100, 120, and 150C. 10 mg Darco Hg mixed with 2 g

sand is tested using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv

NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

The Henry constant is simply the thermodynamic equilibrium constant for

adsorption, and the temperature dependence follows a van’t Hoff expression [6]:

0 exp( )adsHk k

RT

(8.2)

where k0 is the preexponential factor, adsH is the heat of adsorption, R is the

universal constant, and T is temperature in K. The heat of adsorption provides a direct

measure of the strength of the binding mercury and the carbon surface [8]. Figure 8.5

182

plots lnk as a function of the reciprocal of temperature. Since adsorption is

exothermic, the Henry constant decreases with temperature. There is a linear relation

between lnk and 1/T and from the slope and intercept the calculated value for k0 and

adsH is 0.869 m3/g and -8543 J/mol, respectively. In the work of Karatza et al. [5] a

heat of adsorption of -22000 J/mol was found for Darco G60 activated carbon tested

in nitrogen. The Darco G60 carbon has surface area of 600 m2/g and is typically used

for treating fine chemicals and pharmaceutical intermediates [9]. Calculated binding

energy of elemental mercury on activated carbon at room temperature using density

functional theory and fused-benzene ring cluster approach is -18100 J/mol [10].

Effects of other flue gas constituents have not been considered in the simulations. The

derived heat of adsorption for Darco Hg carbon in simulated cement kiln flue gas is

about half of both the experimental data for Darco G60 in nitrogen and theoretical

calculation of binding energy of elemental mercury on activated carbon in nitrogen at

room temperature.

0.0022 0.0024 0.0026 0.0028 0.003

1/T (1/K)

-4.8

-4.6

-4.4

-4.2

-4

ln(k

)

DataY=1027.4881X-7.0484R2=0.93

Figure 8.5. Plot of lnk as a function of 1/T. The Henry’s constants k are derived from

isotherms at 75, 100, 120, and 150C. 10 mg Darco Hg mixed with 2 g sand is tested

using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv

HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

The oxidation of mercury by the Darco Hg carbon seems not to be affected by

the elemental mercury inlet level, as shown in figure 8.6. The mercury oxidation is

183

always larger than 90% and the average mercury oxidation percentage is about 96%

in the studied elemental mercury inlet level of 18-180 µg/Nm3. While almost no

mercury oxidation takes place on the activated carbon tested in nitrogen. It is

expected that the heat of adsorption is different for elemental mercury and oxidized

mercury adsorption by the carbon. This might explain the difference between the

derived heat of adsorption from this work and both the experimental data for Darco

G60 in nitrogen and theoretical calculation of binding energy of elemental mercury

on activated carbon in nitrogen.

0 40 80 120 160 200

Hg inlet concentration (g/Nm3)

0

10

20

30

40

50

60

70

80

90

100

110

Mer

cury

oxi

dat

ion

(%

)

Figure 8.6. Effect of elemental mercury inlet level on mercury oxidation by 10 mg

Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min simulated flue gas

with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21

vol.% CO2, and 6 vol.% O2. 8.4 Effect of carbon particle size 

The Darco Hg carbon has a mean diameter of only 16 µm. To be able to

observe the influence of carbon particle size on mercury adsorption, the Norit RB4

pellets are crushed and sieved to fractions having mean diameter of 38, 98, 165, and

325 µm. Figure 8.7 presents the mercury breakthrough curves for crushed Norit RB4

carbon with different particle sizes tested at 150C using simulated cement kiln flue

gas. Figure 8.8 further illustrates the effect of carbon particle size on the percentage

of mercury oxidation and initial adsorption rate. Faster mercury breakthrough is

184

observed for smaller carbon particle. The final mercury adsorption capacity is the

same at a value of 0.725 µg Hg/mg_carbon, which is about 65% of the Darco Hg

adsorption capacity at 150C. Higher mercury oxidation and initial adsorption rate are

also observed for smaller carbon particles.

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2G

aseo

us

Hg

(C

ou

t/Cin)

1 2 3

1: 38 m2: 98 m3: 325 m

Figure 8.7. Effect of particle size on mercury breakthrough for 10 mg crushed Norit

pellets in 2 g sand tested at 150C using 2.75 Nl/min simulated flue gas with 160-170

µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.%

H2O, 21 vol.% CO2, and 6 vol.% O2.

0 50 100 150 200 250 300 350

Carbon particle size (m)

0

10

20

30

40

50

60

70

80

90

100

Mer

cury

oxi

dat

ion

(%

)

0

0.5

1

1.5

2

Mer

cury

ad

sorp

tio

n r

ate

(g

Hg

/mg

_ca

rbo

n/h

)

Hg oxidationAdsorption rate for first 25 min

0 2 4 6 8

Thiele modulus

185

Figure 8.8. Effects of particle size on mercury oxidation and initial adsorption rate for

10 mg crushed Norit RB4 pellets in 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10

ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

The observation on effect of carbon particle size can be quantified using the

well-known Thiele modulus, expressed for the first order kinetics as [11]:

e

paoxp D

SkR

(8.3)

where Rp is the carbon particle radius, kox is the oxidation rate constant for the first

order reaction, Sa is the surface area per unit mass of carbon, p is the carbon particle

density, and De is the effective diffusivity. The Thiele modulus is defined as the ratio

of an intrinsic reaction rate in the absence of mass transfer limitations to the rate of

diffusion into the particle under specified conditions. When the carbon particle size

increases, Thiele modulus becomes larger. For the smallest particle size of 38 µm the

calculated Thiele modulus is about 1. For the larger particles the Thiele modulus are

much larger than 1, indicating that the mercury oxidation by the large particles might

be limited by the internal diffusion resistance. Similar trends are observed for the

mercury oxidation percentage and initial adsorption rate, which again indicates that

mercury oxidation is an important step in elemental adsorption by the activated

carbon.

8.5 Effect of flue gas flow rate 

The effect of flue gas flow rate on mercury breakthrough profile is presented

in figure 8.9. With larger flue gas flow rate the active carbon is saturated and reach

the equilibrium capacity in shorter time. The initial mercury breakthrough time,

which is defined as time when the mercury concentration after the carbon bed starts

to increase, decreases when the flow rate is increased. The final adsorption capacities

for all tested flow rates are almost the same.

A higher superficial velocity is associated with a higher total mercury input,

resulting in a faster consumption of the sorption capacity of the activated carbon and

186

corresponding higher mercury outlet concentration. While a higher superficial

velocity enhances the mass transfer rate and the corresponding mercury sorption rate.

0 0.4 0.8 1.2 1.6 2

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Gas

eou

s H

g C

ou

t/Cin

2750 Nml/min1830 Nml/min1100 Nml/min

Figure 8.9. Effect of flue gas flow rate on mercury breakthrough of 10 mg Darco Hg

mixed with 2 g sand tested at 150C using 2.75 Nl/min simulated flue gas with 160-

170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1

vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

8.6 Effects of flue gas compositions 

Most of the previous work conducted the tests either in nitrogen or added

single gas to the baseline gas of either only N2 or a mixture of CO2, O2, H2O and N2

[5,11-19]. Some studies use simulated flue gas that does not contain all the relevant

gases, especially the acid gases [20-26]. It is more reasonable to change ranges of the

relevant gases instead of completely removing the gases from the baseline to simulate

the real flue gas. In this work, the concentration of relevant gas component is varied

while the concentrations of the other flue gas components remain at the baseline

values.

8.6.1 Effect of CO2 

The effects of CO2 concentration in the flue gas on mercury adsorption

capacity of Darco Hg carbon at 150C are illustrated in figure 8.10. The mercury

187

adsorption capacity slightly decreases when the CO2 level in the gas is increased from

1 to 31 vol.%. The negative effects of CO2 in the flue gas on mercury adsorption by

the virgin activated carbon were also observed by Yan et al. [13]. The decrease of the

adsorption capacity in the presence of CO2 is probably due to the reduction in the

active sites for mercury adsorption due to the competitive adsorption of CO2 and

mercury on the carbon. The weak effect of CO2 on mercury adsorption might due to

the fact that significant CO2 adsorption on the activated carbon only occur at low

temperature (<50C) and high pressure [27-32].

The observation on the effects of CO2 is different from tests using a gas

mixture of Hg, N2, and CO2, which shows that the mercury adsorption capacity on

activated carbon does not change when the CO2 level is increased from 0 to, 5 and 15

vol.% [24]. Under such gas conditions CO2 in the gas behaves as an inert gas, i.e., it

neither impacts adsorption capacity of activated carbon nor does it affects the

oxidation of mercury. Tests in this work show that the mercury oxidation percentage

is about 94% for CO2 level in the gas less than 21 vol.%, while only 82% mercury

oxidation is obtained at a CO2 level of 31 vol.%.

0 5 10 15 20 25 30 35

CO2 level in gas (%)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Data

Y=1.3779X-0.066

R2=0.72

Figure 8.10. Effect of CO2 concentration in the flue gas on mercury adsorption

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

188

simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10

ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, and 6 vol.% O2.

8.6.2 Effect of O2 

Figure 8.11 shows the effects of oxygen concentration in the flue gas on the

mercury adsorption capacity of Darco Hg at 150C. The mercury adsorption capacity

hardly changes with increasing the oxygen level from 1 vol.% to 16 vol.% when

taking the experimental uncertainty into account. In all the cases the mercury

oxidation by the carbon is about 97%.

0 4 8 12 16 20

O2 level in gas (%)

0

0.2

0.4

0.6

0.8

1

1.2

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Figure 8.11. Effect of O2 concentration in the flue gas on mercury adsorption capacity

of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min simulated

flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl,

1000 ppmv SO2, 1 vol.% H2O, and 21 vol.% CO2.

The effects of oxygen level in the gas on mercury adsorption by activated

carbon were investigated using a gas mixture of Hg, N2, and O2 at 140C in the

literature [24]. When the oxygen concentration was increased from 0 to 3 vol.%, the

mercury adsorption capacity on the studied activated carbon remained almost

unchanged. The mercury adsorption capacity increased by 16 and 33% when the

oxygen level was further increased to 6 and 9 vol.%, respectively.

189

The possibility of carbon-oxygen complexes formation during the fixed-bed

test and their impact on mercury adsorption was investigated [24]. Pretreatment of the

unoxidized carbon by air for 7 days had no impact on the mercury adsorption

performance of the carbon. The air can oxidize carbon surface and increases its acidic

surface functional group content. However, these changes have no impact on the

performance of active carbon for mercury adsorption.

Thermodynamic calculations of mercury-oxygen reactions suggest that about

30% of the mercury could be present as HgO(g) at 200C, while at lower

temperatures HgO(s) is the dominant form, when acid gases such HCl and NOx are

not present in the gas [33,34]. This could lead to very high mercury adsorption

capacity as it is not adsorption of mercury but precipation of HgO(s) on the carbon

that takes place. However, the exact temperature range at which HgO(s) is the

dominant form was not reported. Homogenous gas phase reaction of mercury with

oxygen in an atmosphere of N2, O2, and Hg was investigated by Hall et al [34].

Results suggest that a homogeneous gas phase reaction between oxygen and

elemental mercury is not an important factor in flue gas reaction processes. The

enhanced mercury adsorption in the presence of oxygen can be explained by the

conversion of mercury to mercuric oxides as there is no reaction between oxygen and

mercury in the absence of activated carbon surface.

When HCl is present in the gas even at ppmv level, compared to vol.% level

of oxygen, elemental mercury will be mainly oxidized to HgCl2 instead of HgO as

shown by the thermodynamic calculations presented in chapter 2 [35,36]. Oxygen

was found to be a weak oxidant of mercury [37]. This could explain the weak effect

of oxygen on the mercury adsorption by Darco Hg carbon tested in simulated flue gas

in this work.

8.6.3 Effect of H2O 

The effect of water in the flue gas on the mercury breakthrough profile and

mercury adsorption capacity of Darco Hg at 150C is shown in figure 8.12 and 8.13,

respectively. The presence of water in the flue gas generally accelerates the mercury

breakthrough and therefore decreases the amount of mercury adsorbed on the carbon.

190

The mercury adsorption capacity is increased significantly when water is removed

from the simulated flue gas. The mercury adsorption capacity of Darco Hg carbon

tested without water in the flue gas is about 5.5 times of that with 1 vol.% water in

the gas. However, this result is not practically important since full-scale flue gas

always contains water in percentage level [38]. Compared to CO2, the effects of H2O

in the flue gas on mercury adsorption are more pronounced. The mercury oxidation

percentage is about 97% for water level in the range of 0-15 vol.%, however, the

mercury oxidation decreases to 68% when 25 vol.% of water is added to the flue gas.

As a result the mercury adsorption capacity of Darco Hg with 25 vol.% water is only

about 44% of that with 1 vol.% water in the flue gas.

0 1 2 3 4 5 6

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Gas

eou

s H

g, C

ou

t/Cin

1 2 3

1: 8% H2O2:1% H2O3:0% H2O

Figure 8.12. Effects of water concentration in the flue gas on mercury breakthrough

profile of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10

ppmv HCl, 1000 ppmv SO2, 6 vol.% O2, and 21 vol.% CO2.

191

0 5 10 15 20 25 30

H2O level in gas (%)

0

0.2

0.4

0.6

0.8

1

1.2

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Data

Y=1.1277X-0.261

R2=0.98

Figure 8.13. Effects of water concentration in the flue gas on mercury adsorption

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10

ppmv HCl, 1000 ppmv SO2, 6 vol.% O2, and 21 vol.% CO2. The regressed equation

of mercury adsorption capacity is not valid for H2O level of zero.

Tests in N2, Hg, and H2O at 140C show that the mercury adsorption capacity

of sulfur impregnated activated carbon does not change significantly when the water

content in the gas is increased from 0 to 5 vol.% [24]. While the adsorption capacity

decreases about 25% when the water content is further increased to 10 vol.%. This is

in contrast to tests with simulated flue gas in this work, as the mercury adsorption

capacity always decreases with water addition in the flue gas.

The negative effects of water presence in the gas on mercury adsorption by

the carbon could be due to the competitive adsorption of water on the carbon. Water

adsorption on carbon has been studied by several researchers [39-42]. Due to the

strong chemisorption of water molecules with the acidic oxygen functional group on

the carbon, the initial water adsorption occurs at the functional groups, and further

water adsorption will occur on top of the chemisorbed water molecules via hydrogen

bonding [42].

192

8.6.4 Effect of CO 

Unlike most other combustion processes, organic constituents in the raw

material for clinker production result in CO emissions, even under optimized

combustion conditions. In the preheater, these organic components in the raw

material are liberated, part of them being emitted with the exhaust gas. The carbon

monoxide concentration in the exhaust gas from cement rotary kiln systems ranges

between 0.1 and 5 g/Nm³ (80-4000 ppmv) [43]. There is no reported investigation on

possible effect of CO on mercury adsorption by the activated carbon.

The effect of CO concentration in the flue gas on the mercury adsorption

capacity of Darco Hg carbon at 150C is illustrated in figure 8.14. Similar to the

effects of oxygen, the mercury adsorption capacity is not affected by the presence of

CO in the range of 0-1000 ppmv.

0 200 400 600 800 1000 1200

CO level in the gas (ppmv)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

Hg

/mg

_car

bo

n)

Figure 8.14. Effects of CO concentration in the flue gas on mercury adsorption

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10

ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

As shown in figure 8.15, there is a slight decrease of mercury oxidation

percentage when the CO level in the flue gas is increased. The mercury oxidation

decreases from 98% when less than 100 ppmv CO is present in the flue gas to 85%

193

with 1000 ppmv CO in the flue gas. This is probably because oxidized mercury is

reduced at higher CO levels.

0 200 400 600 800 1000

CO level in flue gas (ppmv)

0

10

20

30

40

50

60

70

80

90

100

110

Mer

cury

oxi

dat

ion

(%

)

Figure 8.15. Effects of CO concentration in the flue gas on mercury oxidation by 10

mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min simulated flue

gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

8.6.5 Effect of SO2 

The effect of SO2 in the flue gas on mercury breakthrough profile and

adsorption capacity of Darco Hg at 150C is presented in figure 8.16 and 8.17,

respectively. There are strong effects of SO2 in the flue gas on the mercury adsorption

by the activated carbon. The mercury adsorption capacity decreases when the SO2

level in the flue gas is increased. The mercury adsorption capacity of Darco Hg

carbon tested with 100 ppmv SO2 in the flue gas is about 4 times of that tested with

1000 ppmv SO2 in the flue gas. Dunham et al. [44] also found that the mercury

adsorption capacity of activated carbon is inversely affected by SO2 in the flue gas

with reductions in adsorption capacity noted at concentrations as low as 100 ppmv

SO2. The mercury oxidation is not affected by changing the SO2 level in the flue gas

and is about 97% for SO2 added in the range of 100-1000 ppmv.

194

0 0.5 1 1.5 2 2.5 3 3.5

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Gas

eou

s H

g, C

ou

t/Cin

1 2

3 4

1: 1000 ppmv SO2

2: 500 ppmv SO2

3: 300 ppmv SO2

4: 100 ppmv SO2

Figure 8.16. Effects of SO2 concentration in the flue gas on mercury breakthrough

profile of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10

ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

0 200 400 600 800 1000 1200

SO2 level in gas (ppmv)

0

1

2

3

4

5

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Data,Hg

Y=69.176X-0.592

R2=0.99

Figure 8.17. Effects of SO2 concentration in the flue gas on mercury adsorption

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10

ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. The regressed equation of

mercury adsorption capacity is valid for SO2 level higher than 100 ppmv.

195

Without the presence of NOx in the gas, the mercury adsorption capacity of

activated carbon is also reported to decrease when the SO2 concentration increases

[38]. This might be explained by the oxidation of SO2 to SO3 on the activated carbon

and the inhibiting effect of SO3 on mercury capture by activated carbon injection has

been observed in full-scale power plant tests [45,46]. In addition to removing

mercury, activated carbon is also used as catalyst for oxidation SO2 to sulphuric acid

and as SO2 sorbent [47,48]. There is competitive adsorption between Hg and SO3

since both mercury and SO3 bind to the Lewis base sites on the activated carbon

surface [45,49]. Some activated carbon catalysts for converting SO2 to H2SO4 are

self-poisoned by SO3 or sulfate buildup on the surface. Therefore, a similar

phenomenon might explain the inhibiting effect of SO3 on mercury capture.

Previous results demonstrated that the oxidation of SO2 on carbon particles

was greatly enhanced by the presence of trace quantities of gaseous NO2 [50-53].

NO2 is an efficient oxidant for SO2 sorbed on carbon. According to the mechanisms

of flue gas and mercury interactions on activated carbon proposed by Dunham et al.

[44] and Olson et al. [54], sulfurous acid that accumulates from the hydration of SO2

converts the previously formed nonvolatile basic mercuric nitrate into the volatile

form. This results in the slow release of previously captured mercury over time in the

presence of NO2 and SO2.

8.6.6 Effect of HCl 

Figure 8.18 illustrates the effects of HCl in the flue gas on the mercury

adsorption capacity of Darco Hg tested at 150C. There are weak effects of HCl

concentration in the flue gas on mercury adsorption capacity when 0.5-20 ppmv HCl

is added to the flue gas. The mercury adsorption capacity increases gradually when

the HCl level is increased from 0.5 to 5 ppmv and then it levels off when the HCl

level is further increased. The mercury oxidation percentage is about 97% for all the

tested HCl levels except that only 87% mercury oxidation is obtained when 0.5 ppmv

HCl is added to the gas.

196

0 4 8 12 16 20 24

HCl level in gas (ppmv)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Data

Y=0.9728X0.071

R2=0.82

Figure 8.18. Effects of HCl concentration in the flue gas on mercury adsorption

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 1000

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. The regressed equation of

mercury adsorption capacity is valid for HCl level higher than 0.5 ppmv.

Strong promoting effects of HCl on mercury adsorption by activated carbon

in the simulated flue gas without NOx at 135C have been reported by Carey et al.

[38]. Addition of HCl to the simulated gases of 1600 ppmv SO2, 6% O2, 12% CO2

and 7% H2O results in an increase of equilibrium adsorption capacity for elemental

mercury from 0 at 0 ppmv HCl to a value approaching 3 µg Hg/mg_carbon in the

range of 50–100 ppmv HCl. The adsorption capacity does not change significantly

above 50 ppm HCl.

When NOx is included the simulated flue gas, the promoting effects of HCl on

adsorption capacity of activated carbon becomes less pronounced. Mercury can be

adsorbed by the carbon without HCl presence in the gas, provided that NOx is present.

To study whether mercury can be adsorbed by the activated carbon in the absence of

HCl, one test was conducted by removing HCl from the baseline flue gas applied in

this work. Figure 8.19 shows the mercury breakthrough curve for 10 mg Darco Hg in

2 g sand at150C.

197

0 0.4 0.8 1.2 1.6

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Gas

eou

s H

g, C

ou

t/Cin

10 ppmv HCl

0 ppmv HCl

Figure 8.19. Comparison of mercury breakthrough curves with 0 and 10 ppmv HCl in

the simulated flue gas, 10 mg Darco Hg mixed with 2 g sand and tested at 150C

using 2.75 Nl/min simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv NO, 23

ppmv NO2, 1000 ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

Figure 8.19 compares the mercury breakthrough curves of 0 and 10 ppmv

HCl in the simulated cement kiln flue gas. Test without HCl in the gas shows that

large amount of mercury is adsorbed on the carbon and mercury breakthrough is not

obtained after 10 h (not shown in figure 8.19). This might be due to the fact that

HgCl2 will form when HCl is present in the gas and on the other hand HgO or HgSO4

will form when HCl is not present through following reactions:

HgOOHg 22 2 (8R1)

NOHgONOHg 2 (8R2)

422 HgSOSOOHg (8R3)

HgO(s) is easily captured by the carbon since it might condense on the carbon at the

applied adsorption temperature of 150C.

8.6.7 Effect of NO 

As shown in figure 8.20, changing of NO concentration in the simulated flue

gas does not affect the adsorption capacity of Darco Hg tested at 150C. The mercury

198

oxidation extent by the Darco Hg carbon is about 98% for tested NO in the range of

100-1000 ppmv.

0 200 400 600 800 1000 1200

NO level in gas ( ppmv)

0

0.2

0.4

0.6

0.8

1

1.2

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Figure 8.20. Effects of NO concentration in the flue gas on mercury adsorption

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 10 ppmv HCl, 23 ppmv NO2, 1000

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

Strong promoting effects of NO on mercury adsorption by the activated

carbon have been reported when 300 ppmv NO was added to baseline gas with 8

vol.% H2O, 6 vol.% O2, and 12 vol.% CO2 at 107C [5,11-19]. Liu et al. [24] reported

that adding 500 ppmv NO to nitrogen did not change the mercury adsorption by the

activated carbon at 140C. Fan et al. [55] proposed that the promoting effects of NO

on mercury adsorption by the activated carbon is due to the reaction of NO with O2 to

form NO2 and active O atoms that could further react with elemental mercury. Thus

the effects of NO on mercury adsorption depend on the presence of other acid gases

in the flue gas. With HCl and NO2 presence in the gas the effects of NO are less

significant.

8.6.8 Effect of NO2 

The effect of NO2 in the flue gas on mercury breakthrough profile and

adsorption capacity of Darco Hg at 150C is presented in figure 8.21 and 8.22,

199

respectively. The mercury oxidation by the Darco Hg carbon is about 98% for tested

NO2 in the range of 0-100 ppmv. Dunham et al. [44] also found that the mercury

adsorption capacity of activated carbon is inversely proportional to the concentrations

of NO2 in the simulated flue gas. A decrease in the mercury adsorption capacity of

activated carbon was observed at concentrations as low as 2.5 ppmv NO2. The

negative effects of NO2 are again due to the interaction between NO2 and SO2, as

discussed in section 8.6.5.

0 0.2 0.4 0.6 0.8 1 1.2

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Gas

eou

s H

g, C

ou

t/Cin

1 2 3

1: 100 ppmv NO2

2: 23 ppmv NO2

3: 5 ppmv NO2

Figure 8.21. Effects of NO2 concentration in the flue gas on mercury breakthrough

curves of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 10 ppmv HCl, 1000 ppmv NO, 1000

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

200

0 20 40 60 80 100 120

NO2 level in gas (ppmv)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Data

Y=2.0685X-0.199

R2=0.89

Figure 8.22. Effects of NO2 concentration in the flue gas on mercury adsorption

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 160-170 µg Hg0/Nm3, 10 ppmv HCl, 1000 ppmv NO, 1000

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. The regressed equation of

mercury adsorption capacity is not valid when NO2 is not presented in the gas.

Table 8.2 summarizes the regressed equations of Darco Hg carbon mercury

adsorption capacity as a function of flue gas composition at 150C. From these

equations it is possible estimate mercury adsorption capacity in a wide range of glue

gas composition.

Table 8.2 Summary of regressed equations of Darco Hg carbon mercury adsorption

capacity as a function of flue gas composition at 150C.

Gases Concentration unit Concentration range Equations CO2 % >0 Y=1.3779X-0.066 H2O % >0 Y=1.1277X-0.261 SO2 ppmv ≥100 Y=69.176X-0.592 HCl ppmv ≥0.5 Y=0.9728X0.071 NO2 ppmv >0 Y=2.0685X-0.199

201

8.7 Conclusions  

A parametric study of elemental mercury adsorption by activated carbon has

been conducted in the fixed-bed reactor by mixing 10 mg Darco Hg carbon with 2 g

sand and using simulated cement kiln flue gas. Equilibrium mercury adsorption

capacity, initial adsorption rate and mercury oxidation percentage are evaluated.

Increasing adsorption temperature results in decreased equilibrium mercury

adsorption capacity of the activated carbon. The mercury adsorption isotherm follows

Henry’s law for the applied mercury inlet levels in this project at all tested

temperatures. All these are consistent with a physical adsorption mechanism. The

derived heat of adsorption is -8543 J/mol for elemental mercury adsorption by Darco

Hg activated carbon in simulated cement kiln flue gas.

The effects of carbon particle size were investigated using the crushed Norit

RB4 pellets. Higher mercury oxidation and initial adsorption rate are observed for

smaller carbon particles, while the equilibrium mercury adsorption capacity is the

same.

The effects of flue gas composition are investigated by varying the

concentrations of relevant gases instead of complete removal of the single gas from

the baseline to simulate the real flue gas. The mercury adsorption capacity does not

change with changes in the O2, CO, and NO levels in the flue gas. The mercury

adsorption capacity decreases when CO2, H2O, SO2, and NO2 concentrations in the

flue gas increase. The following correlation between mercury adsorption capacity and

these gas concentrations are obtained at 150C: mercury adsorption capacity is

proportional to CCO2-0.066, CH2O

-0.261, CSO2-0.592, CNO2

-0.199. The decrease of mercury

adsorption capacity is due to the competition for active site with mercury by CO2 and

H2O, and conversion of the previously formed nonvolatile basic mercuric nitrate into

the volatile form by interactions between SO2 and NO2.

Slight promoting effects of HCl on mercury adsorption are observed when

HCl concentration is varied in the range of 0.5-20 ppmv. A larger mercury adsorption

capacity is obtained when HCl is removed from baseline gas. This might due to the

fact that HgCl2 will form when HCl is present in the gas while HgO(s) will form

202

when HCl is not present. HgO is more easily captured by the carbon since it

condenses on the carbon at the applied adsorption temperature of 150C.

Significant mercury oxidation is observed by the activated carbon. Even for

10 mg carbon typically an oxidation level of 94-97% is found. Increasing CO and

water level in the flue gas causes a slight decrease of mercury oxidation. Larger

mercury oxidation percentage is obtained with smaller carbon particle size. All these

observations indicate that mercury oxidation by HCl, when this is present in the gas,

is an important step in elemental adsorption by the activated carbon.

8.8 References  

[1] A.J. Cruz, J. Pires, A.P. Carvalho, D.C. Brotas, Adsorption of acetic acid by activated

carbons, zeolites, and other adsorbent materials related with the preventive conservation of

lead objects in museum showcases, Journal of Chemical & Engineering Data. 49 (2004) 725-

731.

[2] A. Bagreev, T.J. Bandosz, H2S adsorption/oxidation on unmodified activated carbons:

importance of prehumidification, Carbon. 39 (2001) 2303-2311.

[3] D.D. Do, Adsorption analysis: equilibria and kinetics, Imperial College Press, 1998.

[4] R.C. Bansal, M. Goyal, Activated carbon adsorption, CRC Press, 2005.

[5] D. Karatza, A. Lancia, D. Musmarra, F. Pepe, Adsorption of metallic mercury on

activated carbon, Symp. Int. Combust. 26 (1996) 2439-2445.

[6] H.G. Karge, J. Weitkamp, (Eds.), Molecular sieves-science and technology, volume 7,

Adsorption and diffusion,Springer-Verlag Berlin Heidelberg, 2008.

[7] T.C. Ho, N. Kobayashi, Y. Lee, J. Lin, J.R. Hopper, Experimental and kinetic study of

mercury adsorption on various activated carbons in a fixed-bed adsorber, Environ. Eng. Sci.

21 (2004) 21-27.

[8] D.M. Ruthven, Principles of adsorption and adsorption processes, John Wiley & Sons,

Inc., 1984.

[9] D.O. Cooney, A.L. Hines, Extractive purification of activated carbon. 1. Effect of acid

normality, volume, and repeated contact, Industrial & Engineering Chemistry Process Design

and Development. 22 (1983) 208-211.

[10] B. Padak, M. Brunetti, A. Lewis, J. Wilcox, Mercury binding on activated carbon,

Environmental Progress. 25(2006) 319-326.

[11] G. Skodras, I. Diamantopoulou, G. Pantoleontos, G.P. Sakellaropoulos, Kinetic studies

of elemental mercury adsorption in activated carbon fixed bed reactor, Journal of Hazardous

Materials. 158 (2008) 1-13.

[12] S.J. Miller, G.E. Dunham, E.S. Olson, T.D. Brown, Flue gas effects on a carbon-based

mercury sorbent, Fuel Processing Technology. 65-66 (2000) 343-363.

203

[13] R. Yan, Y.L. Ng, D.T. Liang, C.S. Lim, J.H. Tay, Bench-scale experimental study on the

effect of flue gas composition on mercury removal by activated carbon adsorption, Energy

Fuels. 17 (2003) 1528-1535.

[14] I. Diamantopoulou, G. Skodras, G.P. Sakellaropoulos, Sorption of mercury by activated

carbon in the presence of flue gas components, Fuel Process Technol. 91 (2010) 158-163.

[15] R. Yan, D.T. Liang, L. Tsen, Y.P. Wong, Y.K. Lee, Bench-scale experimental

evaluation of carbon performance on mercury vapour adsorption, Fuel. 83 (2004) 2401-2409.

[16] J.A. Korpiel, R.D. Vidic, Effect of sulfur impregnation method on activated carbon

uptake of gas-phase mercury, Environ. Sci. Technol. 31 (1997) 2319-2325.

[17] S. Lee, Y. Park, Gas-phase mercury removal by carbon-based sorbents, Fuel Process

Technol. 84 (2003) 197-206.

[18] S.J. Lee, Y. Seo, J. Jurng, T.G. Lee, Removal of gas-phase elemental mercury by iodine-

and chlorine-impregnated activated carbons, Atmospheric Environment. 38 (2004) 4887-

4893.

[19] D. Karatza, A. Lancia, D. Musmarra, C. Zucchini, Study of mercury absorption and

desorption on sulfur impregnated carbon, Experimental Thermal and Fluid Science. 21 (2000)

150-155.

[20] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R.

Chang, Preparation and evaluation of coal-derived activated carbons for removal of mercury

vapor from simulated coal combustion flue gases, Energy Fuels. 12 (1998) 1061-1070.

[21] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption by activated

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy

Conference, Research Triangle Park, NC, 22-25 April, 1997.

[22] S. Eswaran, H.G. Stenger, Z. Fan, Gas-phase mercury adsorption rate studies, Energy

Fuels. 21 (2007) 852-857.

[23] R. Ochiai, M.A. Uddin, E. Sasaoka, S. Wu, Effects of HCl and SO2 concentration on

mercury removal by activated carbon sorbents in coal-derived flue gas, Energy & Fuels. 23

(2009) 4734.

[24] W. Liu, R.D. Vidic, T.D. Brown, Impact of flue gas conditions on mercury uptake by

sulfur-impregnated activated carbon, Environ. Sci. Technol. 34 (2000) 154-159.

[25] G.E. Dunham, S.J. Miller, Mercury capture by an activated carbon in a fixed-bed bench-

scale system, Environmental Progress. 17 (1998) 203.

[26] S. Lee, J. Lee, T.C. Keener, Novel sorbents for mercury emissions control from coal-

fired power plants, Journal of the Chinese Institute of Chemical Engineers. 39 (2008) 137-

142.

[27] M. Molina-Sabio, A.M.A. Muñecas, F. Rodríguez-Reinoso, B. McEnaney, Adsorption

of CO2 and SO2 on activated carbons with a wide range of micropore size distribution,

Carbon. 33 (1995) 1777-1782.

204

[28] T.C. Drage, O. Kozynchenko, C. Pevida, M.G. Plaza, F. Rubiera, J.J. Pis, C.E. Snape, S.

Tennison, Developing activated carbon adsorbents for pre-combustion CO2 capture, Energy

Procedia. 1 (2009) 599-605.

[29] M.G. Plaza, S. García, F. Rubiera, J.J. Pis, C. Pevida, Post-combustion CO2 capture with

a commercial activated carbon: Comparison of different regeneration strategies, Chem. Eng.

J. 163 (2010) 41-47.

[30] C. Shen, C.A. Grande, P. Li, J. Yu, A.E. Rodrigues, Adsorption equilibria and kinetics

of CO2 and N2 on activated carbon beads, Chem. Eng. J. 160 (2010) 398-407.

[31] Y. Sun, Y. Wang, Y. Zhang, Y. Zhou, L. Zhou, CO2 sorption in activated carbon in the

presence of water, Chemical Physics Letters. 437 (2007) 14-16.

[32] B. Guo, L. Chang, K. Xie, Adsorption of carbon dioxide on activated carbon, Journal of

Natural Gas Chemistry. 15 (2006) 223-229.

[33] B. Hall, O. Lindqvist, E. Ljungstroem, Mercury chemistry in simulated flue gases related

to waste incineration conditions, Environ. Sci. Technol. 24 (1990) 108-111.

[34] B. Hall, P. Schager, J. Weesmaa, The homogeneous gas phase reaction of mercury with

oxygen, and the corresponding heterogeneous reactions in the presence of activated carbon

and fly ash, Chemosphere. 30 (1995) 611-627.

[35] C.L. Senior, A.F. Sarofim, T. Zeng, J.J. Helble, R. Mamani-Paco, Gas-phase

transformations of mercury in coal-fired power plants, Fuel Process Technol. 63 (2000) 197-

213.

[36] R.N. Sliger, J.C. Kramlich, N.M. Marinov, Towards the development of a chemical

kinetic model for the homogeneous oxidation of mercury by chlorine species, Fuel Process

Technol. 65-66 (2000) 423-438.

[37] Niksa, S. and J. J. Helble, Interpreting laboratory test data on homogeneous mercury

oxidation in coal-derived exhausts. EPA-DOE-EPRI Combined Power Plant Air Pollution

Control Symposium: The Mega Symposium and the A&WMA Specialty Conference on

Mercury Emissions: Fate, Effects, and Control, Chicago, Illinois, August 2001.

[38] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors

affecting mercury control in utility flue gas using activated carbon, Journal of the Air &

Waste Management Association. 48 (1998) 1166.

[39] M.M. Dubinin, V.V. Serpinsky, Isotherm equation for water vapor adsorption by

microporous carbonaceous adsorbents, Carbon. 19 (1981) 402-403.

[40] M.M. Dubinin, Water vapor adsorption and the microporous structures of carbonaceous

adsorbents, Carbon. 18 (1980) 355-364.

[41] F. Stoeckli, T. Jakubov, A. Lavanchy, Water adsorption in active carbons described by

the Dubinin-Astakhov equation, Journal of the Chemical Society, Faraday Transactions. 90

(1994) 783-786.

[42] D.D. Do, H.D. Do, A model for water adsorption in activated carbon, Carbon. 38 (2000)

767-773.

205

[43] ECRA (European cement research academy), Carbon monoxide formation and burn-out

during the clinker burning process, News letter 2, 2008.

[44] G.E. Dunham, E.S. Olson and S.J. Miller, Impact of flue gas constituents on carbon

sorbents, McLean, VA, September 19-21, 2000.

[45] A.A. Presto, E.J. Granite, Impact of sulfur oxides on mercury capture by activated

carbon, Environ. Sci. Technol. 41 (2007) 6579-6584.

[46] J. Jarvis, F. Meserole, SO3 Effect on Mercury Control, Power Eng. 112 (2008) 54-60.

[47] E. Raymundo-Piñero, D. Cazorla-Amorós, C. Salinas-Martinez de Lecea, A. Linares-

Solano, Factors controlling the SO2 removal by porous carbons: relevance of the SO2

oxidation step, Carbon. 38 (2000) 335-344.

[48] E. Raymundo-Piñero, D. Cazorla-Amorós, A. Linares-Solano, Temperature programmed

desorption study on the mechanism of SO2 oxidation by activated carbon and activated

carbon fibres, Carbon. 39 (2001) 231-242.

[49] A.A. Presto, E.J. Granite, A. Karash, Further investigation of the impact of sulfur oxides

on mercury capture by activated carbon, Ind Eng Chem Res. 46 (2007) 8273-8276.

[50] W.R. Cofer III, D.R. Schryer, R.S. Rogowski, The enhanced oxidation of SO2 by NO2

on carbon particulates, Atmospheric Environment (1967). 14 (1980) 571-575.

[51] W.R. Cofer III, D.R. Schryer, R.S. Rogowski, The oxidation of SO2 on carbon particles

in the presence of O3, NO2 and N2O, Atmospheric Environment (1967). 15 (1981) 1281-1286.

[52] W.R. Cofer III, D.R. Schryer, R.S. Rogowski, Oxidation of SO2 by NO2 and O3 on

carbon: Implications to tropospheric chemistry, Atmospheric Environment (1967). 18 (1984)

243-245.

[53] J.A. Rodriguez, T. Jirsak, J. Dvorak, S. Sambasivan, D. Fischer, Reaction of NO2 with

Zn and ZnO: Photoemission, XANES, and density functional studies on the formation of NO3,

The Journal of Physical Chemistry B. 104 (2000) 319-328.

[54] E.S. Olson, B.A. Mibeck, S.A. Benson, J.D. Laumb, C.R. Crocker, G.E. Dunham, et al.,

The mechanistic model for flue gas-mercury interactions on activated carbons: The oxidation

site, Prepr. Pap. -Am. Chem. Soc. , Div. Fuel Chem. 49 (2004) 279-280.

[55] X. Fan, C. Li, Zeng Guangming, Z. Gao, L. Chen, W. Zhang, et al., Removal of gas-

phase element mercury by activated carbon fiber impregnated with CeO2, Energy & Fuels. 24

(2010) 4250-4254.

206

9

Fundamental investigation of mercury

chloride adsorption by activated carbon in

simulated cement kiln flue gas

This chapter deals with a fundamental investigation of mercury chloride

adsorption by the Darco Hg activated carbon in simulated cement kiln flue gas. The

results are compared with tests using elemental mercury.

9.1 Introduction 

Compared to research on elemental mercury adsorption [1-13], there are few

studies on mercury chloride adsorption by activated carbon [1,2,14-16]. Some of

these studies conducted tests using simulated flue gases containing 1600 ppmv SO2,

50 ppmv HCl, 12 vol.% CO2, 7 vol.% H2O, 6 vol.% O2, but without NOx [1,2,14].

Carey et al. [15] performed tests using simulated flue gas with 1600 ppmv SO2, 1-50

ppmv HCl, 10-12 vol.% CO2, 8 vol.% H2O, 6 vol.% O2, and 200-400 ppmv NOx. The

research [15] focused on a comparison of mercury adsorption capacity obtained in a

fixed-bed reactor using simulated flue gas and real power plant flue gas. Mibeck et al.

[16] investigated the effects of acid gases by adding 1600 ppmv SO2, 50 ppmv HCl,

400 ppmv NO, and 20 ppmv NO2 alone or in combination to baseline gas of 12 vol.%

CO2, 8 vol.% H2O, 6 vol.% O2. Only breakthrough curves are presented, neither

adsorption capacity nor kinetics is reported in their work.

In this work, mercury chloride adsorption capacity and kinetics are

investigated by varying the relevant gas concentrations and operating parameters.

207

9.2 Effect of temperature 

The effect of temperature on mercury chloride adsorption is investigated by

applying the same conditions as for the study with elemental mercury source reported

in chapter 8. The breakthrough curves using HgCl2 source are compared with those

obtained with elemental mercury at 100, 120, and 150C, as illustrated in figure 9.1-3.

In contrast to the breakthrough curve obtained with elemental mercury, the

breakthrough curve of mercury chloride often has an introduction period. It takes

some time to reach the lowest outlet mercury concentration after switching the flue

gas with mercury chloride to the carbon bed. While the outlet mercury decreases to

the lowest value almost simultaneously after switching the flue gas with elemental

mercury to the carbon bed. This phenomenon is also observed by Mibeck et al. [16].

They also reported that about 90-95% of mercury after the carbon bed is oxidized

mercury. Measurement in this work shows that all the mercury after the carbon bed is

oxidized mercury, i.e., no reduction takes place.

0 0.2 0.4 0.6 0.8 1

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Ga

se

ou

s H

g, C

ou

t/Cin

Hg0, 170g/Nm3 Hg, 1.183 g Hg/mg_carbon

HgCl2, 183g/Nm3 Hg, 1.224 g Hg/mg_carbon

Figure 9.1. Comparison of breakthrough curves obtained using mercury chloride and

elemental mercury source for 10 mg Darco Hg carbon mixed with 2 g sand tested at

150C using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2,

1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

208

0 0.2 0.4 0.6 0.8 1 1.2

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Gas

eou

s H

g, C

ou

t/Cin

Hg0, 166g/Nm3 Hg, 1.335 g Hg/mg_carbon

HgCl2, 156g/Nm3 Hg, 1.375 g Hg/mg_carbon

Figure 9.2. Comparison of breakthrough curves obtained using mercury chloride and

elemental mercury source for 10 mg Darco Hg carbon mixed with 2 g sand tested at

120C using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2,

1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

Time (hour)

0

0.2

0.4

0.6

0.8

1

1.2

Ga

seo

us

Hg

, C

ou

t/Cin

Hg0, 167g/Nm3 Hg, 1.506 g Hg/mg_carbon

HgCl2, 159g/Nm3 Hg, 1.429 g Hg/mg_carbon

Figure 9.3. Comparison of breakthrough curves obtained using mercury chloride and

elemental mercury source for 10 mg Darco Hg carbon mixed with 2 g sand tested at

100C using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2,

1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.

The equilibrium mercury adsorption capacities are very similar for tests using

elemental mercury and mercury chloride under same gas conditions at different

temperatures. This is probably due to the catalytic oxidation of elemental mercury by

209

the activated carbon. As shown in chapter 8, almost all elemental mercury is oxidized

by the activated carbon. Thus it is not surprising to observe similar adsorption

behavior of elemental mercury and mercury chloride adsorption by the activated

carbon for the present flue gas containing HCl. For sorbents with poor mercury

oxidation ability, the adsorption behavior of elemental mercury and oxidized mercury

is expected to be different. Increasing adsorption temperature also decreases the

adsorption capacity of mercury chloride.

Similarly to tests with elemental mercury, the adsorption constants of HgCl2

are also derived. Figure 9.4 shows that there is a linear relation between lnk and 1/T

and from the slope and intercept the calculated value for k0 and adsH is 1.595 m3/g

and -6587 J/mol, respectively. The corresponding k0 and adsH for tests with

elemental mercury is 0.869 m3/g and -8543 J/mol, respectively.

0.0022 0.0024 0.0026 0.0028 0.003

1/T (1/K)

-4.8

-4.6

-4.4

-4.2

-4

ln(k

)

Hg0 dataY=1027.4881X-7.0484R2=0.93HgCl2 data

Y=792.2246X-6.4409R2=0.93

Figure 9.4. Plot of lnk as a function of 1/T. The Henry’s constants k are derived from

isotherms at 100, 120, 150, and 180C. 10 mg Darco Hg mixed with 2 g sand is tested

using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv

HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. Data for tests

using elemental mercury are shown for comparison.

210

9.3 Effect of flue gas composition 

Since negligible effects of CO2, O2, CO, and NO on elemental mercury

adsorption by the activated carbon are observed and preliminary tests of mercury

chloride adsorption by the activated carbon shows similar behavior as elemental

mercury, only effects of H2O, SO2, HCl, and NO2 on mercury chloride adsorption by

the Darco Hg activated carbon are investigated. The effects of H2O, SO2, HCl, and

NO2 on equilibrium adsorption capacity of mercury chloride are presented in figure

9.5-8, respectively. Data for tests using elemental mercury are shown for comparison.

0 5 10 15 20 25 30

H2O level in gas (%)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Data,Hg0

Y=1.1277X-0.261

R2=0.98Data, HgCl2

Y=1.2389X-0.240

R2=0.98

Figure 9.5. Effects of water in the flue gas on mercury chloride adsorption capacity of

10 mg Darco Hg carbon mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv

HCl, 6 vol.% O2, and 21 vol.% CO2.

211

0 200 400 600 800 1000 1200

SO2 level in gas (ppmv)

0

1

2

3

4

5

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Data,Hg0

Y=69.176X-0.592

R2=0.99Data,HgCl2

Y=58.23X-0.565

R2=0.99

Figure 9.6. Effects of SO2 in the flue gas on mercury chloride adsorption capacity of

10 mg Darco Hg carbon mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 6 vol.% O2, 1

vol.% H2O, and 21 vol.% CO2.

0 4 8 12 16 20 24

HCl level in gas (ppmv)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Data,Hg0

Y=0.9728X0.071

R2=0.82Data,HgCl2Y=0.9162X0.1158

R2=0.76

Figure 9.7. Effects of HCl in the flue gas on mercury chloride adsorption capacity of

10 mg Darco Hg carbon mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 6 vol.% O2,

1 vol.% H2O, and 21 vol.% CO2.

212

0 20 40 60 80 100 120

NO2 level in gas (ppmv)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

Mer

cury

ad

sorp

tio

n c

apac

ity

(g

_Hg

/mg

_car

bo

n)

Data, Hg0

Y=2.0685X-0.199

R2=0.89Data, HgCl2

Y=2.0732X-0.197

R2=0.88

Figure 9.8. Effects of NO2 in the flue gas on mercury chloride adsorption capacity of

10 mg Darco Hg carbon mixed with 2 g sand tested at 150C using 2.75 Nl/min

simulated flue gas with 1000 ppmv NO, 1000 ppmv SO2, 10 ppmv HCl, 6 vol.% O2,

1 vol.% H2O, and 21 vol.% CO2.

The adsorption capacity of mercury chloride is slightly larger than the

elemental mercury when the water content in the flue gas is above 1 vol.%. Almost

the same tendency of adsorption capacity as a function of SO2, HCl, and NO2

concentration in the flue gas is observed for mercury chloride and elemental mercury.

This is again due to the high oxidation rate of elemental mercury by the Darco Hg

carbon.

9.4 Conclusions 

Similar adsorption behaviors of mercury chloride and elemental mercury by

Darco Hg activated carbon are observed using simulated cement kiln flue gas at

different temperatures. Increasing adsorption temperature also decreases the

adsorption capacity of mercury chloride.

The effects of H2O, SO2, HCl, and NO2 on mercury chloride adsorption by

Darco Hg are investigated by varying their concentrations in the baseline gas.

Compared to elemental mercury adsorption, a slightly larger adsorption capacity of

mercury chloride is obtained when the water content in the flue gas is above 1 vol.%.

213

The dependence of mercury chloride adsorption capacity on SO2, HCl, and NO2

concentrations in the flue gas is the same as elemental mercury adsorption capacity.

The similar behavior of mercury chloride and elemental mercury is due to the

effective catalytic oxidation of elemental mercury by the activated carbon in the

presence of HCl.

9.5 References 

[1] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R. Chang,

Preparation and evaluation of coal-derived activated carbons for removal of mercury vapor

from simulated coal combustion flue gases, Energy & Fuels. 12 (1998) 1061-1070.

[2] M. Rostam-Abadi, S.G. Chen, H.C. Hsi, M. Rood, R. Chang, T.R. Carey, B., Hargrove, C.

Richardson, W. Rosenhoover, F. Meserole, Novel vapor phase mercury sorbents,

Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant Control, Washington,

DC, Aug 25–29, 1997.

[3] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang and F.B. Meserole, Factors

affecting mercury control in utility flue gas using sorbent injection, Proceedings of the Air &

Waste Management Association's 90th Annual Meeting, Toronto, Ontario, Canada, June 8-13,

1997.

[4] C.X. Hu, J.S. Zhou, Z.Y. Luo, S. He, G.K. Wang, K.F. Cen, Effect of oxidation treatment

on the adsorption and the stability of mercury on activated carbon, Journal of Environmental

Sciences-China. 18 (2006) 1161-1166.

[5] B. Ghorishi and B.K. Gullett, Fixed-bed control of mercury: Role of acid gases and a

comparison between carbon-based, calcium-based, and coal fly ash sorbents, Proceedings of

the EPRI/DOE/EPA Combined Utility Air Pollutant Control Symposium, Washington, DC,

August 25–29, 1997.

[6] S.B. Ghorishi and C.B. Sedman, Combined mercury and sulfur oxides control using

calcium-based sorbents, Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant

Control Symposium, Washington, DC, August 25–29, 1997.

[7] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption by activated

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy

Conference, Research Triangle Park, NC, 22-25 April, 1997.

[8] N. Hutson, C. Singer, C. Richardson, J. Karwowski and C. Sedman, Practical applications

from observations of mercury oxidation and binding mechanisms, EPRI-DOE-EPA-AWMA

Combined Power Plant Air Pollutant Control Mega Symposium, Washington DC, August 30

- September 2, 2004.

[9] E.J. Granite, M.C. Freeman, R.A. Hargis, W.J. O'Dowd, H.W. Pennline, The thief process

for mercury removal from flue gas, J. Environ. Manage. 84 (2007) 628-634.

214

[10] W.J. O’Dowd, H.W. Pennline, M.C. Freeman, E.J. Granite, R.A. Hargis, C.J. Lacher, A.

Karash, A technique to control mercury from flue gas: The thief process, Fuel Processing

Technology. 87 (2006) 1071-1084.

[11] R. Bhardwaj. Impact of temperature and flue gas components on mercury speciation and

uptake by activated carbon sorbents . Master thesis. Master, University of Pittsburgh, 2007.

[12] M.M. Maroto-Valer, Y. Zhang, E.J. Granite, Z. Tang, H.W. Pennline, Effect of porous

structure and surface functionality on the mercury capacity of a fly ash carbon and its

activated sample, Fuel. 84 (2005) 105-108.

[13] S. Eswaran, H.G. Stenger, Z. Fan, Gas-phase mercury adsorption rate studies, Energy &

Fuels. 21 (2007) 852-857.

[14] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors

affecting mercury control in utility flue gas using activated carbon, Journal of the Air &

Waste Management Association. 48 (1998) 1166.

[15] T.R. Carey, C.F. Richardson, R. Chang, F.B. Meserole, M. Rostam-Abadi, S. Chen,

Assessing sorbent injection mercury control effectiveness in flue gas streams, Environ. Prog.

19 (2000) 167-174.

[16] B.A.F. Mibeck, E.S. Olson, S.J. Miller, HgCl2 sorption on lignite activated carbon:

Analysis of fixed-bed results, Fuel Process Technol. 90 (2009) 1364-1371.

215

10

Simulation of mercury adsorption by fixed

carbon bed

To properly understand an adsorption process, two basic ingredients, i.e.,

equilibrium and transport processes must be investigated. Understanding of the

adsorptive capacity is within the domain of equilibrium, and understanding of the

diffusion resistance is within the domain of transport process. These two aspects are

first introduced in this chapter. The remaining part of this chapter deals with

mathematical models describing the behavior of isothermal adsorption of mercury in

a carbon particle and a fixed carbon bed including adsorption isotherm, mass balance

for the gas phase and mass balance inside the adsorbent.

10.1 Adsorption equilibrium 

A summary of commonly used isotherm equations for pure gas adsorption is

given in the following table 10.1. Considering the typical mercury level in the flue

gas of ppb level, the simple Henry’s law is able to describe the isotherm well, as

shown in chapter 8.

216

Table 10.1. Summary of commonly used isotherm equations for pure gas adsorption

[1].

Isotherm Equation Remarks Henry law C KC Low pressure range

Langmuir

1s

bCC C

bC

Has Henry law limit and finite saturation limit

Frendlich 1/nFC K C Does not have Henry law

limit and no saturation limit Langmuir-Frendlich 1/

1/

( )

1 ( )

n

s n

bCC C

bC

Does not have Henry law limit, but has finite saturation limit

Toth 1/[1 ( ) ]s t t

bCC C

bC

Has Henry law limit and finite saturation limit

Unilan 1( )

2 1

ss

s

C be CC ln

s be C

Has Henry law limit and finite saturation limit

K: Henry constant; C: gaseous adsorbate concentration; C: adsorbed concentration in

the sorbent; Cs: saturation adsorbed concentration in the sorbent; b: Langmuir

constant; KF: Frendlich constant; s: heterogeneity parameter

10.2 Transport consideration in adsorption process 

Adsorption of an adsorbate molecule on to the porous surface of an adsorbent

include following steps [2]:

1. External (or interphase) mass transfer of the adsorbate from the bulk fluid by

convection through a thin film or boundary layer.

2. Internal (intraphase) mass transfer of the adsorbate by pore diffusion from the

outer surface of the adsorbent to the inner surface of the internal porous structure.

3. Surface diffusion along the porous surface.

4. Adsorption of the adsorbate onto the porous surface.

10.2.1 External transport 

Rates of convection mass and heat transfer between the outer surface of a

particle and the surrounding bulk fluid during an adsorption process are given,

respectively, by [2]:

217

m b s

dNk A c c

dt (10.1)

s b

dQhA T T

dt (10.2)

where km is the external mass transfer coefficient, A is the particle external surface

area, Cb and Cs is the gas concentration in the bulk and at the particle surface,

respectively. h is the heat transfer coefficient, Tb and Ts is the gas temperature in the

bulk and at the particle surface, respectively.

When fluid flows past a single particle, experimental transport data

correlations are usually developed for coefficients averaged over the particle surface.

Some typical correlations published by Ranz and Marshall for Nusselt numbers as

high as 30, and Sherwood numbers to 160 are the following [2]: 11

32Re Pr2 0.60NuN N N (10.3)

1132

Re2 0.60Sh ScN N N (10.4)

where Prandtl number NPr=k

C p ; Schmidt number NSc=iD

; Reynolds number

NRe= Gd p , and G is the fluid mass velocity.

When particles are packed in a bed, the fluid flow patterns are restricted, and

the single particle correlations cannot be used to estimate the average external

transport coefficients for particles in the bed. A correlation of 37 sets of mass-transfer

data including Sherwood number corrections for axial dispersion result in an

expression of the form [2]:

315.0

Re1.12 Sci

pmSh NN

D

dkN (10.5)

This equation covers a Schmidt number range from 0.6 to 70600, a Reynolds number

range from 3 to 10000. Particle shapes applicable include spheres, short cylinders,

flakes and granules. By analogy, the corresponding equation for fluid-particle

convection heat transfer in packed beds is:

31

Pr5.0

Re1.12 NNk

hdN p

Nu (10.6)

218

When these equations are used with beds packed with non-spherical particles, dp, is

the equivalent diameter of a spherical particle.

10.2.2 Internal transport 

Porous particles in most cases have a sufficiently high effective thermal

conductivity so that temperature gradients within the particle are negligible. In

contrast, internal mass transfer within the particle must be considered. In sorption

processes, transport is from the exterior to the interior for adsorption and from the

interior to the exterior for desorption processes. The flux of mercury transported to

the carbon particle can be expressed as:

dr

dCDN eA (10.7)

where De is the effective diffusion coefficient. There are basically three modes of

transport of molecules inside a porous medium: Knudsen diffusion, molecular

diffusion, and surface diffusion [1].

10.2.2.1 Molecular Diffusion

When the adsorbate is in a macropore or in the fluid phase, the frequency of

collision with a surface is minimal and transport of the molecule occurs via

intermolecular collisions only. This mode of transport is due to a partial pressure

gradient of a continuum fluid mixture.

For binary gas mixtures at low pressure (<10 atm), the diffusion coefficient is

inversely proportional to the pressure, increases with increasing temperature and is

almost independent of composition for a given gas pair. For nonpolar gases the

prediction of the Chapman-Enskog kinetic theory is usually preferred for diffusion

coefficient calculation [1]:

(10.8)

where DAB is in cm2/s, T in K, P in atm, MW in g/mol. AB and D,AB are

characteristic molecular properties that are based on the Lennard-Jones potential

3 1 1

2,

( )0.0018583 A BMW MW

ABAB D AB

TD

P

219

parameters of the individual species in the system. MWA and MWB are molecular

weights of species A and B. The molecular collision diameter, AB, is calculated as

the arithmetic average of the two species:

12 ( )AB A B (10.9)

D,AB is a dimensionless function of temperature and the intermolecular potential

field for a molecular of A and B. The interaction is described by the individual

Lennard-Jones 12-6-potentials, A and B, in accordance with following equation:

.AB A Bk k k

(10.10)

D,AB can be calculated according to:

, 0.15610

1.06036 0.19300 1.03587 1.76474

( ) exp(0.47635 ) exp(1.52996 ) exp(3.89411 )D AB

AB AB AB AB

kT kT kT kT

(10.11)

where k is the Boltzmann’s constant.

The Lennard-Jones potential parameter for N2 can be easily found while for

elemental mercury only few data are available. Table 10.2 lists the Lennard-Jones

potential parameter for N2 and elemental mercury.

Table 10.2. Lennard-Jones potential parameter for N2 and elemental mercury [3].

(Å) /k (K) N2 3.681 91.5 Hg 3.23 627

10.2.2.2 Knudsen Diffusion

This diffusion process occurs when the mean free path of the adsorbate is

much larger than the diameter of the channel in which the diffusing molecules reside.

This normally occurs at very low pressure and channels of small size, usually of order

of 10 nm to 100 nm [1]. The flow is induced by collision of gaseous molecules with

the pore wall.

829700

3g

K

R Tr TD r

MW MW (10.12)

where r is the pore radius in cm, T in K, MW in g/mol, Dk, in cm2/s.

220

10.2.2.3 Surface Diffusion

In most cases, the surface diffusion coefficient is unknown as the heat of

adsorption is not available. Furthermore, the surface diffusivity is a strong function of

the amount of mercury adsorbed and the sorbent surface coverage is low due to the

low mercury level in the flue gas, it is therefore reasonable to assume that the surface

diffusion resistance can be neglected.

10.3 Modeling of adsorption in a single particle 

The final aim of this project is to develop a mathematical model that can

simulate mercury adsorption by a carbon cake on the fabric filter bags. A single

particle model is the core and starting points of the filter model. The single particle

model can be used to study how an adsorption process would vary with parameters

such as particle size, bulk concentration, pressure, temperature, pore size, and

adsorption affinity. Analytical solution of the single particle model is available when

linear adsorption isotherm is used. However, the single particle model works only at

constant gas atmosphere. The gas concentration changes in time for both the fixed-

bed and fabric filter adsorption processes. Therefore a numerical solution of the

single particle model is required in order to incorporate it to fixed-bed and fabric

filter models.

Do [1] has made a detailed description of the single particle adsorption model

using linear isotherm. Both analytical solution and numerical solution using

orthogonal collocation method with subroutines in MATLAB are provided in his

book. Fixed-bed and fabric filter models in this work are further developed on the

basis of the single particle model.

Since the mercury level in the flue gas is very low, the adsorption system can

be treated as isothermal. Mass balance around a thin shell element in the particle

gives [1]:

1(1 ) ( )s

p p e s

CC CD r

t t r r r

(10.13)

221

where p is the porosity of the particle, C is gaseous mercury concentration, C is the

mercury concentration in the adsorbed phase, De is the pore diffusivity, and s is the

particle shape factor (s=0, 1, and 2 for slab, cylinder, and sphere, respectively).

The free molecules of mercury in the pore space and the adsorbed mercury

molecules at any point within a particle are assumed in equilibrium with each other.

The local linear isotherm takes the form:

C KC (10.14)

where K is the Henry constant.

Substituting the local equilibrium into the mass balance equation, we can

obtain [1]:

2 1( )

(1 )se

app sp p

DC CD C r

t K r r r

(10.15)

with initial condition: t=0, C=Ci, (10.16)

and typical boundary conditions:

0, 0C

rr

(10.17)

, ( )p

p e m bRr R

Cr R D k C C

r

(10.18)

For slab object R is the half thickness, while for cylindrical and spherical objects, R is

their respective radius. Cb is the concentration of the adsorbate in the bulk

surrounding the particle, km is the external mass transfer coefficient.

An analytical solution of the concentration distribution within the particle is

given in the form of an infinite series [1]. The solution is only valid for a particle

surrounded by a gas atmosphere not changing in time. To get a numerical solution,

equation 10.15 is written in a dimensionless form by defining following non-

dimensional variables and parameters:

20 0 0

; ; ; ;app b ib i

p p

D t C CC ry x y y

C R R C C

1( )s

s

y yx

x x x

(10.19)

Initial condition: =0, y=yi (10.20)

222

Boundary conditions:

0, 0y

xx

(10.21)

)(;1 yyBix

yx b

(10.22)

where Bi is the Biot number epm DRk .

The problem has symmetry at x=0, and it is useful to utilize this by making

the transformation of u=x2, and the differential equation becomes [1]:

2

24 2( 1)

y y yu s

u u

(10.23)

The equation is solved by the orthogonal collocation method [4]. The domain u(0,1)

is represented by n interior collocation points. Taking the boundary point (u=1) as the

(n+1)-th point, we have a total of n+1 interpolation points. The first and second

derivatives at these interpolation points are related to the functional values at all

points as given below:

1n

ij jji

yA y

u

(10.24)

2 1

2

n

ij jji

yB y

u

(10.25)

The matrices A and B are constant matrices once n+1 interpolation points have been

chosen. The mass balance equation is valid at any point within the u domain.

Evaluating the equation at the ith interior collocation point we get:

1

1

ni

ij jj

yC y

(10.26)

For i=1, 2,…n, where

4 2(1 )ij i ij ijC u B s A (10.27)

Numerical calculation of the average gaseous mercury concentration inside the

particle is obtained by:

1

0

1( ) ( , ) ( 1) ( , ) s

V

C t C t x dV s C t x x dsV

(10.28)

223

1 1

2

0

( 1)( ) ( , )

2

ssC t C t x u du

(10.29)

The integration is evaluated by Radau quadrature [1,4]:

1 11 1

2

10 0

1( , ) (1 ) ( , ) ; 0;

2

s n

k kk

sC t x u du u u C t x du w C

(10.30)

where the weight factors wk are the Radau quadrature weights.

The calculated Radau quadratue weights from the program are normalized [4]:

( , )k

k

wW

I (10.31)

( , ) ( 1) ( 1) 2 10

( 2) 1 2

sI for and

s

(10.32)

2

1k kw Ws

(10.33)

1 1 12

10

( 1)( ) ( , )

2

s n

k kk

sC t C t x u du W C

(10.34)

The boundary condition at the particle surface becomes:

1; ( )2 b

y Biu y y

u

(10.35)

1

1, 11

( )2

N

n j j b nj

BiA y y y

(10.36)

From which we can solve for the concentration at the boundary in terms of other

dependent variables [1]:

1,1

1

1, 1

2

21

n

b n j jj

n

n n

y A yBi

yA

Bi

(10.37)

The mass balance equation together with initial and boundary conditions are solved

numerically by combination of collocation and Runge-Kutta methods using

MATLAB [1]. Both the analytical solution and numerical solution give the same

results for the single particle adsorption model using local linear isotherm [1].

The developed model is used to simulate mercury adsorption by a single

Darco Hg activated carbon particle exposed to elemental mercury at 150C in

224

simulated cement kiln flue gas with 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2,

10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. The inputs to the model

are presented in table 10.3. The Henry’s constant is derived from fixed-bed

experiments using 10 mg Darco Hg carbon in 2 g sand. Table 10.4 presents the

calculated diffusion coefficients, external mass transfer coefficient and Biot number

by the program.

Table 10.3. Inputs to the single particle adsorption model.

Parameters Unit Value Temperature C 150

Carbon particle diameter µm 16 Carbon true density kg/m3 2200 [1]

Carbon particle porosity - 0.73 [1] Carbon pore radius nm 10 [1]

Bed porosity - 0.5 Reactor diameter mm 18

Flow rate Nl/min 2.75 Hg inlet concentration µg/Nm3 170

Henry’s constant m3/g 10.24 Collocation point number - 10

Table 10.4. Calculated diffusion coefficients, external mass transfer coefficient and

Biot number by the single particle adsorption model.

Parameters Unit Value Diffusion coefficient of Hg0 in N2, DHg,N2 m2/s 2.44e-5 Knudsen diffusion coefficient of Hg0, DK m2/s 1.41e-6

Pore diffusion coefficient, Dp m2/s 1.33e-6 Effective pore diffusion coefficient, De m2/s 7.10e-7

Apparent diffusion coefficient, Dapp m2/s 1.17e-13 External mass transfer coefficient, km m/s 3.37

Biot number - 37.95

The concentration profiles of elemental mercury inside the particle at different

times are illustrated in figure 10.1. At the beginning there is a sharp concentration

profile inside the particle, indicating larger diffusion resistance inside the particle

compared to the boundary layer. At later stage, the concentration profile becomes flat

225

until the mercury concentration reaches the bulk level at all positions inside the

particle when the adsorption equilibrium is obtained.

Figure 10.1. Simulated mercury concentration profile inside the particle at different

time. The corresponding time for each curve from bottom to top is 0.06, 0.29, 0.77,

2.20, 5.40, 14.04, 33.84, 94.32, 323.64, 1425.24, 3600 s, respectively. Inputs to the

model are given in table 10.3.

The amount of elemental mercury adsorbed as a function of time for different

particle sizes is illustrated in figure 10.2. The calculated external mass transfer

coefficient, Biot number, and time for equilibrium adsorption are given in the table

10.5. The larger the particle is, the larger the Biot numbers are. This indicates that the

larger particle has larger internal mass transfer resistance. As a result, it takes the

larger particle longer time to reach the equilibrium. For a 16 µm particle it takes 7.2

min to reach the equilibrium, while for a 100 µm particle it takes more than four

hours.

226

Figure 10.2. Simulated amount of mercury adsorbed in the Darco Hg carbon particle

as a function of time for particles with different diameters.

Table 10.5. Calculated external mass transfer coefficient, Biot number, and simulated

equilibrium approach time for Darco Hg carbon with different particle sizes.

d=5 m d=16 m d=50 m d=100 m d=200 m d=300 m External

mass transfer coefficient,

km (m/s)

10.26 3.37 1.18 0.64 0.36 0.26

Biot number 36.14 37.95 41.48 45.15 50.73 55.24 99%

equilibrium approach time (h)

0.01 0.12 0.89 4.28 17.04 36.41

 

10.4 Fixed bed adsorption model 

In this project a plug flow model with linear equilibrium isotherm, external

and intraparticle mass transfer resistances is developed. Due to the low level of

227

mercury applied in this project the system can be treated as isothermal. The plug-flow

model means that the fluid velocity profile is uniform at all radial positions, a fact

which generally involves turbulent flow conditions. In addition, it is assumed that the

fixed-bed adsorption reactor is packed randomly with adsorbent particles. The

adsorption process is supposed to be very fast relative to the convection and diffusion

effects; subsequently, local equilibrium will exist inside the adsorbent particles [5].

If the solid particles are small, the axial diffusion effects can be ignored and

the main mode of transport in the mobile fluid phase is by convection [6]. Consider a

section of the fixed bed column with a length of z, cross section area of A, and bed

porosity of b, as shown in figure 10.3, a mass balance of the mercury contained in

both phase, we get [6]:

t

qzA

t

CzAtzzACvtzACv bb

)1(),(),( 00 (10.38)

where v0 is the superficial fluid velocity. Dividing through by Az and taking limit,

we get the overall balance of the mercury [6]:

0 (1 ) 0b b

C C qv

z t t

(10.39)

q is the volume-average mercury loading per unit volume of porous pellet,

Figure 10.3. Sketch of a fixed-bed absorber.

Using the void velocity u, we get:

10b

b

C C qu

z t t

(10.40)

228

q can be expressed as [2]:

23

0

3 PR

P

q r qdrR

(10.41)

where Rp is the radius of the carbon particle.

Equation 10.40 gives the concentration of the mercury in the bulk gas as a

function of time and location in the bed. The concentration of mercury in the gas

within the pores of a carbon particle is obtained by solving equation 10.13.

The simultaneous solution of equation 10.13 and10.41 is a hard task, which

can be avoided by using the tank-in-series method. The fixed-bed is divided into N

equal size well-mixed tanks and the mercury mass balance in the bulk gas phase for

each tank can be written as:

(10.42)

where Vi is the volume of each tank, F is the flow rate through the bed, Cbin,i and Cb,i

is the inlet and outlet mercury concentration in tank i, respectively, Np is the particle

number in tank i, As is the outer surface area of one particle, Cs,i is the gaseous

mercury concentration at the particle surface in tank i. Giving the bed cross area A,

bed height h, total mass of sorbent in the bed M, void velocity u, particle radius Rp,

and density p, the above equation can be expressed as:

(10.43)

Further arranging equation 10.43 into:

,, , , ,

3( ) ( )

(1 )b i m

bin i b i b i s ib p p p

dC MkuNC C C C

dt h R Ah

(10.44)

Initial condition:

t=0, C=0, Cbin,i=Cb0, all tanks (10.45)

t>0 Cbin= Cb0, tank 1 (10.46)

Boundary conditions:

0, 0C

rr

(10.47)

,, ( )e m b iRr R

Cr R D k C C

r

(10.48)

,, , , ,( )b i

b i bin i b i m p s b i s i

dCV FC FC k N A C C

dt

,, , , ,

3( ) ( )

(1 )b i

b b bin i b i m b i s ip p p

dCAh Mu A C C k C C

N dt NR

229

Dimensionless equation can be written as:

(10.49)

Initial condition:

=0, y=0, ybin,i=1, all tanks (10.50)

>0 ybin= 1, first tank (10.51)

Boundary conditions become:

0, 0y

uu

(10.52)

1; ( )2 b

y Biu y y

u

(10.53)

The boundary-value partial differential equation along the particle radius

(equation 10.19) is solved by the orthogonal collocation method [4]. The particle

radius is represented by n interior collocation points. The boundary point is the

(n+1)th point. For each tank the initial-value ordinary equations contain n+1 equation

for the particle collocation points and another equation for the bulk phase mercury in

the bed:

(10.54)

For the whole fixed-bed the resulting system of N(n+2) ordinary differential

equations are solved by the MATLAB routine ode15s.

Besides the inputs to the single particle adsorption model, the inputs to the

developed fixed bed adsorption model also include tank number of 20 after parameter

study the effect of tank number, a bed thickness of 5 mm, which corresponds to a

mixture of 10 mg Darco Hg carbon with 2 g sand powder in the reactor, actual and

baseline concentrations of SO2, NO2, and H2O in the simulated cement kiln flue gas,

preexponential factor of Henry’s constant of 0.869 m3/g and heat of adsorption of -

8543 J/mol as presented in chapter 8.

2,

, , , 1,

3( ) ( )

(1 )pb i m

bin i b i b i iapp b p p p

Rdy MkuNy y y y

d D h R Ah

22

, 2 2 1

3( ) ( )

(1 )n m

bin i n n napp b p

dy MkR uNy y y y

d D h R Ah

230

A parameter study of the model was first conducted to evaluate the effect of

collocation point number inside the carbon particle and tank number on the mercury

breakthrough curve of the carbon bed. Figure 10.4 illustrates the effects of collocation

point number inside the carbon particle on the mercury breakthrough curve of 10 mg

Darco Hg carbon tested at 75C with 90 µg/Nm3 mercury in the simulated cement

kiln flue gas. Reasonable agreement between the simulation and experimental data is

already obtained using one collocation point inside the carbon particle. Simulations

using 2, 5, and 10 collocation points generate the same mercury breakthrough curve.

Generally more accurate solutions can be obtained using more collocation points.

Since the simulation can be done within 30 s, a collocation point of 10 is used as

default input to the program.

Figure 10.4. Effects of collocation point number inside the carbon particle on the

simulated mercury breakthrough curves of 10 mg Darco Hg mixed with 2 g sand

tested at 75C using 2.75 Nl/min simulated flue gas with 90 µg Hg0/Nm3, 10 ppmv

HCl, 1000 ppmv NO, 1000 ppmv SO2, 23 ppmv NO2, 1 vol.% H2O, 6 vol.% O2, and

21 vol.% CO2. A tank number of 20 is applied in the simulation.

231

Figure 10.5 presents the effects of applied tank number in the simulation on

the predicted mercury breakthrough curve of 10 mg Darco Hg carbon tested at 75C

with 90 µg/Nm3 mercury in the simulated cement kiln flue gas. Better agreement

between the simulation and experimental data is obtained when larger tank number is

applied. When the tank number is above 20, the produced breakthrough profile is

almost the same.

Figure 10.5. Effects of applied tank number on the simulated mercury breakthrough

curves of 10 mg Darco Hg mixed with 2 g sand tested at 75C using 2.75 Nl/min

simulated flue gas with 90 µg Hg0/Nm3, 10 ppmv HCl, 1000 ppmv NO, 1000 ppmv

SO2, 23 ppmv NO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. A collocation point

number of 10 inside the carbon particle is applied in the simulation.

Validation and parametric study of mercury adsorption by the activated

carbon is conducted by simulation and comparison with the experimental data as

shown in figure 10.6-10.12. The developed fixed bed model can reasonably simulate

the effects of temperature, mercury inlet concentration, flow gas rate, carbon particle

size, and SO2, H2O, NO2 level in the flue gas on the mercury breakthrough curve of

fixed bed with 10 mg carbon in 2g sand powder. Isotherm study of crushed Norit

232

RB4 pellets is not performed and the Henry’s constant for Norit RB4 carbon is

calculated by comparing the equilibrium adsorption capacity with Darco Hg carbon at

the same conditions.

Figure 10.6. Comparison of simulation and experimental data for effect of adsorption

temperature on mercury breakthrough of 10 mg Darco Hg mixed with 2 g sand using

2.75 Nl/min simulated flue gas with 90 µg Hg0/Nm3,1000 ppmv NO, 23 ppmv NO2,

10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

The model can clearly simulate the effect of adsorption temperature on

mercury breakthrough curve of the carbon bed, i.e., faster mercury breakthrough is

obtained at higher adsorption temperature. The Henry constants at each temperature

are calculated from the derived preexponential factor and heat of adsorption. The best

agreement between the simulation and experimental data is obtained for adsorption

test at 75C as shown in figure 10.6.

233

Figure 10.7. Comparison of simulation and experimental data for effect of elemental

mercury inlet level on mercury breakthrough of 10 mg Darco Hg mixed with 2 g sand

tested at 120C using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv

NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

Figure 10.7 shows that the best agreement between the simulation and

experimental data regarding the effects of mercury inlet level is the mercury

breakthrough curve of 57 µg/Nm3 mercury level. The simulation slightly overpredicts

the mercury adsorption with 95 µg/Nm3 and underpredicts the mercury adsorption for

tests with 27 µg/Nm3 in the flue gas. When taking the experimental uncertainty into

account, the simulation is acceptable as the average uncertainty of the equilibrium

mercury adsorption capacity is about ±10%.

234

Figure 10.8. Comparison of simulation and experimental data for effect of flue gas

flow rate on mercury breakthrough of 10 mg Darco Hg mixed with 2 g sand tested at

150C using 2.75 Nl/min simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv

NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and

6 vol.% O2.

Figure 10.8 illustrates that there are good agreements between the simulations

and experimental data regarding the initial mercury breakthrough time when the

mercury concentration after the carbon bed starts to increase and the model correctly

predicts that. The model overpredicts the effect of changing gas flow especially for

high flow rates.

235

Figure 10.9. Comparison of simulation and experimental data for effect of particle

size on mercury breakthrough for 10 mg crushed Norit RB4 pellets in 2 g sand tested

at 150C using 2.75 Nl/min simulated flue gas with 160-170 µg Hg0/Nm3, 1000 ppmv

NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and

6 vol.% O2.

There are good agreements between the simulations and experimental data for

the effect of particle size for Norit RB4 over the size range of 38-325 µm, as

illustrated in figure 10.9.

236

Figure 10.10. Comparison of simulation and experimental data for effect of SO2

concentration in the flue gas on mercury breakthrough profile of 10 mg Darco Hg

mixed with 2 g sand tested at 150C using 2.75 Nl/min simulated flue gas with 160-

170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1 vol.% H2O, 6

vol.% O2, and 21 vol.% CO2.

237

Figure 10.11. Comparison of simulation and experimental data for effect of water

concentration in the flue gas on mercury breakthrough profile of 10 mg Darco Hg

mixed with 2 g sand tested at 150C using 2.75 Nl/min simulated flue gas with 160-

170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 6

vol.% O2, and 21 vol.% CO2.

238

Figure 10.12. Comparison of simulation and experimental data for effect of NO2

concentration in the flue gas on mercury breakthrough curves of 10 mg Darco Hg

mixed with 2 g sand tested at 150C using 2.75 Nl/min simulated flue gas with 160-

170 µg Hg0/Nm3, 10 ppmv HCl, 1000 ppmv NO, 1000 ppmv SO2, 1 vol.% H2O, 6

vol.% O2, and 21 vol.% CO2.

Figure 10.10 and 10.11 show that there are good agreements between the

simulations and experimental data on the effects of SO2 and H2O levels in the

simulated cement kiln flue gas. The model slightly overpredicts the mercury

adsorption rate with 100 and 5 ppmv NO2 in the flue gas and slightly underpredicts

the mercury adsorption with 23 ppmv NO2 as shown in figure 10.12. The effects of

these gases on the mercury adsorption capacity are evaluated by the derived

correlations between mercury adsorption capacity and gas concentrations, as

presented in chapter 8.

239

10.5 Conclusions 

Mathematical models for mercury adsorption by a single carbon particle and a

fixed carbon bed are developed. Local equilibrium within the carbon particle is

assumed and the adsorption system is assumed to be isothermal due to the low

mercury concentration presented in the flue gas. The models account for both the

external and internal mass transfer resistances. The orthogonal collocation method is

used to solve mercury diffusion and adsorption inside a sorbent particle. The fixed-

bed model is solved by a tank-in-series method.

Henry’s constant obtained from fixed-bed investigation of mercury adsorption

by activated carbon in the simulated cement kiln flue gas is used as input to the

models. The single particle model can simulate the mercury concentration profile and

amount of adsorbed mercury inside the carbon particle as a function of adsorption

time.

The developed fixed bed model can reasonably simulate the effects of

adsorption temperature, mercury inlet concentration, flow gas rate, carbon particle

size, and SO2, H2O, NO2 level in the flue gas on the mercury breakthrough curve of

fixed bed with 10 mg carbon in 2 g sand powder. The developed models are useful

tools for understanding the mercury adsorption by the activated carbon and

interpretation of the experimental results.

10.6 List of symbols 

A: carbon particle external surface area (m2) A: cross area of the fixed-bed (m2) A: matrix in equation 10.24 As: outer surface area of one carbon particle (m2) b: Langmuir equilibrium constant (m3/g) B: matrix in equation 10.25 Bi: dimensionless Biot number C: gaseous mercury concentration (µg/m3) C: matrix in equation 10.26 Cb: gas bulk mercury concentration (µg/m3) Cbo: initial gas bulk mercury concentration (µg/m3) Cb,i: outlet mercury concentration in tank i (µg/m3)

240

Cbin,i: inlet mercury concentration in tank i (µg/m3) Ci: initial mercury concentration (µg/m3) Cs,i : gaseous mercury concentration at the particle surface in tank i (µg/m3) cp: specific heat (J/(kg.K) Cs: gaseous mercury concentration at the particle surface (µg/m3) Cµ: adsorbed mercury concentration in the sorbent (µg/m3) Cµs: saturated concentration of adsorbed mercury in the sorbent (µg/m3) DAB: binary molecular diffusion coefficient (m2/s) Dapp: apparent diffusion coefficient (m2/s) De: effective diffusion coefficient (m2/s) Di: molecular diffusion coefficient (m2/s) Dk: Knudsen pore diffusion coefficient (cm2/s) dp: particle diameter (m) F: gas flow rate (m3/s) G: fluid mass velocity (kg/(m2.s)) h: heat transfer coefficient (W/(m2.K)) h: bed height (m) k: thermal conductivity (W/(m.K)) k : the Boltzmann’s constant (J/K) K: Henry’s constant KF: Frendlich constant km: gas film mass transfer coefficient (m/s) M: carbon load in the fixed-bed (mg) MW: mole weight (g/mol) n: exponent in isotherm equations of Frendlich, and Langmuir-Frendlich n: number of interior collocation points N: amount of transported mercury in equation 10.1 (µg) N: tank number Np: carbon particle number in the tank NPr: Prandtl number NSc: Schmidt number NRe: Reynolds number P: pressure (atm) q: mercury concentration in the sorbent (µg/m3) Q: heat (W) r: radial coordinate (m) r: pore radius (cm) Rg: universal gas constant, 8.314 (J/(mol.K)) Rp: sorbent particle radius (m) s: heterogeneity parameter in Unilan isotherm equation

241

s: the particle shape factor (s=0, 1, and 2 for slab, cylinder, and sphere, respectively) t: exponent in isotherm equations of Toth t: time (s) T: temperature (K) Tb : gas temperature in the bulk (K) Ts: gas temperature at the particle surface (K) u: dimensionless parameter, u=x2 u: void velocity (m/s) v0: superficial fluid velocity (m/s) Vi: volume of the tank (m3) Vp: volume of the carbon particle (m3) wk: Radau quadratue weight Wk: normalized Radau quadratue weight x: dimensionless radius y: dimensionless mercury concentration yb: dimensionless gas bulk mercury concentration (µg/m3) yi: dimensionless initial mercury concentration (µg/m3) z: axial coordinate (m) Greek symbols : parameter defined by equation 10.31 : parameter defined by equation 10.31 AB: Lennard-Jones 12-6-potentials for specie A and B b: bed void fraction p: sorbent particle porosity p: sorbent particle density (kg/m3) µ: dynamic viscosity (kg/(m.s)) AB : molecular collision diameter (Å) D,AB: dimensionless parameter in equation 10.8 and 10.11 Hads: heat of adsorption (J/mol) : dimensionless time 10.7 References 

[1] D.D. Do. Adsorption analysis: equilibria and kinetics, Imperial College Press, 1998.

[2] J.D. Seader, E.J. Henley, Separation process principles, John Wiley & Sons, Inc. 1998.

[3] P.J. Gardner, P. Pang, S.R. Preston, Binary gaseous-diffusion coefficients of mercury and

of zinc in hydrogen, helium, argon, nitrogen, and carbon-dioxide, J. Chem. Eng. Data. 36

(1991) 265-268.

242

[4] J. Villadsen, M.L. Michelsen, Solutions of differential equation models by polynomial

approximation, Prentice-Hall, Inc., 1978.

[5] D.M. Ruthven, Principles of adsorption and adsorption processes, John Wiley & Sons,

Inc., 1984.

[6] R.G. Rice, D.D. Do. Applied mathematics and modelling for chemical engineers, John

Wiley & Sons, Inc, 1995.

243

11

Simulation of mercury removal by activated

carbon injection upstream of a fabric filter

This chapter deals with the development of a two-stage model for simulation

of mercury removal by carbon injection upstream of a fabric filter. First the

development of duct-fabric filter models is presented, and then the models are

compared with available experimental data from pilot-scale investigation.

11.1 Common assumptions  for mercury removal  in  the duct and 

fabric filter 

Mercury removal by the sorbent injection upstream of a fabric filter consists

of two stages, i.e., the duct and filter sections as illustrated in figure 11.1. Powdered

sorbent such as activated carbon is metered to the injection point at a rate

proportional to the gas stream flow. Once dispersed, mercury species diffuse to the

particle surface and migrate into pores of the activated carbon particle. The carbon

particles remain suspended in the moving gas stream in the duct for periods of one to

three seconds. It then deposits onto the carbon cake formed on the filter bags.

Additional mercury capture takes place when the mercury-containing gas stream

passes through the carbon cake. The carbon cake grows with filtration time and after

a certain time the pressure drop across the filter reaches its threshold value and the

cleaning process is initiated by pulse injection of compressed air. A fraction of the

filter bags is periodically cleaned to relieve the pressure drop across the fabric filter.

244

Figure 11.1. Sketch of the mercury removal process by carbon injection upstream of a

fabric filter.

A mathematical model is a useful tool to simulate the mercury capture and

evaluate the mercury removal efficiency for various operational conditions. An

advanced model can provide a rational basis for describing and characterizing the

effectiveness of mercury removal by sorbent injection and provide guidelines for

developing new types of sorbents and improve of the process.

To make the mathematics tractable, following assumptions are made:

1. The relevant mercury species in the gas phase is assumed to be either

elemental mercury or mercuric chloride. Elemental mercury is much more difficult to

remove if the sorbent cannot oxidize it. As shown in chapter 8 and 9, similar

adsorption behavior of elemental mercury and mercury chloride by activated carbon

is observed since significant oxidation of mercury by the activated carbon occurs if

HCl is present in the gas above few ppmv. This is the case in most practical systems

and will be assumed here.

2. Activated carbon particles are spherical, uniform in size and uniformly

dispersed in the duct and filter cake.

245

3. The temperature is constant and uniform through the system. Mercury

adsorption heat effects are neglected due to the trace level mercury concentrations.

The adsorption equilibrium is described by Henry’s law as shown in chapter 8.

4. Both the gas and the solid flow rates are constant. In reality, there are changes

of both differential pressures over the filter bag and cake porosity with time. The

cleaned section of the filter by pulse jet would have less hydraulic resistance,

resulting in a larger fraction of the flow diverted to this section. There would be a

dynamic redistribution of the flow as the cake grows on the filter bag surface. Flora et

al. [1] evaluated the effect of the dynamic redistribution of flow on removal of

mercury using activated carbon injection in a fabric filter system. The magnitude of

this impact is small compared with the potential impact caused by uncertainties in the

isotherm and mass transfer parameters. When the differential pressure over the bag

increases the system fan speed will be increased correspondingly to maintain the

same filtration velocity. It is therefore reasonable to assume constant gas flow

through the filter cake.

5. Mercury adsorption on the duct walls and filter fabric is negligible.

Equilibrium conditions are reached between the gas phase and walls/fabric so that no

net exchange of mercury is present.

6. Removal of mercury from the bulk gas phase is caused solely by adsorption

on the activated carbon.

7. A mass transfer boundary layer causes resistance to mass transfer from the

bulk gas phase to the activated carbon particle external surface, and mass transfer

within the carbon particle is controlled by pore diffusion.

8. Since the mercury level in the flue gas is very low and the surface diffusivity

is a strong function of the amount of mercury adsorbed, it is reasonable to assume

that the surface diffusion resistance can be neglected.

9. The free gaseous mercury molecules in the pore and the adsorbed mercury

molecules at any point within a particle are in equilibrium with each other. The local

adsorption kinetics is much faster than the diffusion process into the particle.

246

11.2 Duct model 

Part of the mercury is removed in the duct section. Simulation of mercury

removal in the duct is presented in this section taking into account relevant

mechanisms.

The flue gas is assumed to travel in plug flow along the duct. To verify

whether the slip velocity between the activated carbon particles and the gas is

relevant, Scala evaluated the terminal velocity of the particles for the particle sizes of

interest, e.g., less than 100 m [2-4]. Results indicated that terminal velocities are

always more than one order of magnitude lower than typical flue gas velocity so that

it is reasonable to assume that particles travel at the same velocity as the flue gas. The

particle Reynolds number was always smaller than one, justifying the assumption of

Stokes regime.

Mass balance around a thin shell element in the spherical particle gives:

22

1(1 ) ( )p p p e

CC CD r

t t r r r

(11.1)

where p is the porosity of the particle, C is the gaseous mercury concentration, C is

the adsorbed mercury per unit mass of the particle, p is the particle density and De is

the effective diffusivity.

The local linear isotherm takes the form:

C KC (11.2)

where K is the Henry’s constant.

Substituting the local equilibrium into the mass balance equation, we can get:

2 22

1( )

(1 )e

appp p p

DC CD C r

t K r r r

(11.3)

Assuming plug flow and no slip velocity, mercury adsorption in the duct can be

treated as a batch adsorber. Assuming perfect mixing, the mass balance of mercury in

the bulk phase is:

p

bR

dCV A J

dt (11.4)

where V is the volume of the adsorber, Cb is the concentration of mercury in the

adsorber, A is the total exterior surface area of all carbon particles in the adsorber, and

247

pRJ is the mass transfer into the carbon particle per unit interfacial area. If the

particles are spheres, the total exterior surface area is

3

(1 )p

p p p

mA

R

(11.5)

where mp is the mass of the particles and Rp is the particle radius. Equation 11.4 can

be rearranged into:

33( )

(1 ) (1 ) (1 )p p p

pb mbR R R

p p p p p p p p p

mdC k1 3J J C C

dt V R R R

(11.6)

where is the carbon load in the flue gas (kg/m3), km is the external mass transfer

coefficient, pR

C is the gaseous mercury concentration at the carbon particle surface.

Initial condition: t=0, C=0, Cb=Cb0, (11.7)

Boundary conditions:

0, 0C

rr

(11.8)

,

, ( )p

p

p e m bRR t

Cr R D k C C

r

(11.9)

Equations 11.3 and 11.6 are written in a dimensionless form by defining the

following non-dimensional variables and parameters:

20 0

; ; ; ;app bb

b p p b

D t CC ry x y

C R R C

where y is the interparticle gas concentration.

22

1( )

y yx

x x x

(11.10)

2

1 1

33( ) ( )

(1 ) (1 )p m pb m

b bp p p app p p app

R k Rdy ky y y y

d R D D

(11.11)

Initial condition: =0, y=0, yb=1 (11.12)

Boundary conditions become:

0, 0y

xx

(11.13)

248

1; ( ) ( )m pb b

e

k Ryx y y Bi y y

x D

(11.14)

The problem of diffusion and adsorption in the carbon particle has symmetry at x=0,

and it is useful to utilize this by making the transformation of u=x2, and the

differential equation becomes:

2

24 6

y y yu

u u

(11.15)

Boundary conditions become:

0, 0y

uu

(11.16)

1; ( )2 b

y Biu y y

u

(11.17)

The equation is solved by the orthogonal collocation method [5]. The domain

u(0,1) is represented by n interior collocation points. Taking the boundary point

(u=1) as the (n+1)-th point, we have a total of n+1 interpolation points. The first and

second derivatives at these interpolation points are related to the functional values at

all points as given below:

1n

ij jji

yA y

u

(11.18)

2 1

2

n

ij jji

yB y

u

(11.19)

The matrices A and B are constant matrices once n+1 interpolation points have been

chosen. The mass balance equation is valid at any point within the u domain.

Evaluating the equation at the ith interior collocation point we get:

1

1

ni

ij jj

yC y

(11.20)

For i=1, 2,…n+1, where

4 6ij i ij ijC u B A (11.21)

, 1 11

ni

ij j i n nj

yC y C y

(11.22)

The boundary condition at the carbon particle surface is:

249

1

1, 11

( )2

n

n j j b nj

BiA y y y

(11.23)

From which we can solve for the concentration at the boundary in terms of other

dependent variables [6]:

1,1

1

1, 1

2

21

n

b n j jj

n

n n

y A yBi

yA

Bi

(11.24)

Including the equation for the bulk phase mercury,

22 1

3( )

(1 )m pn

n np p app

k Rdyy y

d D

(11.25)

n+1 initial-value ordinary differential equations are solved simultaneously by

MATLAB routine ode15s.

The developed duct model is very similar to that developed by Scala [2-4].

The main difference between the models is that Scale used Langmuir isotherm and

dynamic adsorption, i.e., local equilibrium is not assumed.

The model input parameters are listed in table 11.1 for simulation of mercury

adsorption by injection of Darco Hg activated carbon into the duct. The Henry’s

constant is derived from fixed-bed investigation as presented in chapter 8. When the

gas composition is different from the baseline test, the effect of individual gas on the

mercury adsorption is evaluated using correlations derived from chapter 8. Since full-

scale data are not available for comparison it is the intention here to test the model

ability instead of simulating the full-scale application. The simulation results are

analyzed by selecting a set of operating variables as a base case for computations and

to assess the influence of the relevant input variables on the process by varying them

one at a time.

250

Table 11.1. Inputs to the duct adsorption model.

Parameters Unit Value Temperature C 75-150

Actual SO2 concentration ppmv 1000 Baseline SO2 concentration ppmv 1000 Actual NO2 concentration ppmv 23

Baseline NO2 concentration ppmv 23 Actual H2O concentration % 1

Baseline H2O concentration % 1 Hg inlet concentration µg/Nm3 170

Carbon particle diameter µm 5-200 Carbon true density kg/m3 2200

Carbon particle porosity - 0.73 Carbon pore radius nm 10

Carbon injection rate g/m3 0.05-10 Residence time in the duct s 0-10

Henry’s constant preexponential factor

m3/g 0.869

Heat of adsorption J/mol -8543 Collocation point number - 10

Figure 11.2 illustrates the simulated bulk mercury concentration in the duct as

a function of flight time for different injection rates of 16 µm Darco Hg carbon at

150C to the baseline gas of 170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2, 10

ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. The gas

bulk mercury concentration is normalized with the inlet mercury concentration. The

mercury concentration decreases with increasing of residence time in the duct and the

carbon load. To obtain 80% mercury removal by injection of 16 µm Darco Hg carbon

at 150C, it needs either a long residence time in the duct, i.e., long duct (> 1 g/m3

load and 10 s) or large carbon injection rate (10 g/m3 load and 0.16 s). After 10 s in

the duct the mercury removal efficiency is 77.9% and 97.6% for a carbon injection

rate of 1 and 10 g/m3, respectively.

251

Figure 11.2. Simulated gaseous mercury concentration as a function of residence time

in the duct and carbon injection rate. Darco Hg carbon with a diameter of 16 µm is

injected at 150C to simulated cement kiln flue gas with 170 µg Hg0/Nm3, 1000

ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.%

CO2, and 6 vol.% O2.

The effects of carbon particle size on mercury concentration in the duct are

illustrated in figure 11.3. The gas bulk mercury concentration decreases faster for

smaller carbon particles during the first 2 s in the duct and a larger mercury removal

is obtained by smaller carbon particles at all residence time, indicating that diffusion

resistance is relevant for mercury adsorption on carbon particles. Decreasing the

particle size from 16 to 5 µm can increase the mercury removal efficiency from 77.9

to 87.6% using an injection rate of 1 g/m3 at 150C. The improvement of mercury

removal efficiency by further lowering the particle size is less pronounced (not shown

in figure 11.3).

252

Figure 11.3. Simulated gaseous mercury concentration as a function of residence time

in the duct and carbon particle size. 1 g/m3 Darco Hg carbon is injected at 150C to

simulated cement kiln flue gas with 170 µg Hg0/Nm3, 1000 ppmv NO, 23 ppmv NO2,

10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

The effects of particle size on mercury removal can be explained by Biot

numbers for different particle sizes as shown in table 10.5. Calculations show that the

larger the particle is, the smaller external mass transfer coefficient, the larger the Biot

numbers are. This indicates that the larger particle has relatively larger internal

transfer resistance. As a result, it takes the larger particle longer time to reach the

equilibrium. In all the cases calculated here, the Biot numbers are much larger than

36, indicating that the internal diffusional resistance is much larger than the external

mass transfer resistance.

The effects of temperature on mercury removal in the duct are presented in

figure 11.4. Similar mercury outlet concentrations are observed for injection 0.5 g/m3

Darco Hg carbon with a size of 16 µm for the first 2 s in the duct, and then lower

mercury outlet concentrations are obtained with lower flue gas temperature and

longer residence time in the duct.

253

Figure 11.4. Simulated gaseous mercury concentration as a function of residence time

in the duct and flue gas temperature. 0.5 g/m3 Darco Hg carbon with a diameter of 16

µm is injected at 150C to simulated cement kiln flue gas with 170 µg Hg0/Nm3,

1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21

vol.% CO2, and 6 vol.% O2.

11.3 Model for the filter cake 

The filter cake formed on the bags is similar to a fixed-bed with the difference

that new adsorbent is continuously fed. The fixed-bed model with constant bed

thickness developed in chapter 10 is extended here to deal with this situation. During

filtration without cleaning of the bag the carbon cake thickness grows with time. At

the beginning no carbon particles are collected on the bag surface. After a short time

t a layer of carbon is collected and the corresponding carbon cake thickness is:

(1 )(1 )b

p p b

uL t

(11.26)

where is the carbon load in the flue gas, u is the face velocity on the filter bags.

This thin layer of carbon particles can be treated as a continuous stirred-tank reactor

254

(CSTR). The bed volume is AL, the mass of carbon in the tank

is (1 ) (1 )(1 )b p b p p bM V AL . The mass balance for mercury in the bulk

phase in the tank becomes:

3(1 )( ) ( )

p

b b mbin b b R

b p

dC kuC C C C

dt L R

(11.27)

The dimensionless equation can be written as:

2

1

3(1 )( ) ( )pb b m

bin b bapp b p

Rdy kuy y y y

d D L R

(11.28)

n+1 collocation points are used for the carbon particle. The bulk phase mercury

balance equation becomes:

2

22 2 1

3(1 )( ) ( )pn b m

bin n n napp b p

Rdy kuy y y y

d D L R

(11.29)

Initial and boundary conditions:

=0, y=0, ybin=1 (11.30)

0, 0y

uu

(11.31)

)(2

;1 12

nn yyBi

u

yu (11.32)

Theses n+1 equations are solved in a time interval of [0 t] by MATLAB

routine ode15s. After another t, another layer of carbon with thickness of L is

formed on the bag surface and on top of the first layer and is termed as tank 2. Now

the system contains 2(n+1) initial-value ordinary differential equations which are

solved simultaneously in a time interval of [0 t] by MATLAB routine ode15s. The

initial conditions for equations in tank 1 are the calculated concentration from last t

interval. The initial conditions for tank 2 are y=0, ybin,2=1. At >0, ybin,1= yb,2. The

cycle is conducted to the desired filtration time. Here it is assumed that the carbon

particles are injected just at the filter inlet. The combination of duct injection and

filter cake model will be presented in section 11.5. In the later case the carbon

particles arriving in the filter will have already adsorbed mercury with some radial

profile, i.e., y0.

255

Simulation of mercury removal by the fixed-bed with moving boundary is

performed using the conditions from the Durkee pilot plant study [7,8]. The inputs to

the model are given in table 11.2. The flue gas temperature is taken from the field test

report [7,8] and the flue gas compositions are supplied by Paone [9]. Other gas

concentrations are the same as the baseline gas. The effects of CO2 and HCl are not

accounted for since the effects of these gases are less pronounced compared to SO2,

NO2 and H2O. Referring to results from chapter 8, when the CO2 level in the flue gas

is above 21 vol.%, which is used in baseline test and deriving of the adsorption

kinetics, the mercury adsorption capacity of the carbon is only slightly decreased.

With HCl in the gas up to 15 ppmv, the mercury adsorption capacity is almost not

affected by changing the HCl level in the gas. Large adsorption capacity is obtained

without HCl in the gas.

Table 11.2. Inputs to the filter cake model.

Parameters Unit Value Temperature C 138

Actual SO2 concentration ppmv 5 Baseline SO2 concentration ppmv 1000 Actual NO2 concentration ppmv 5

Baseline NO2 concentration ppmv 23 Actual H2O concentration % 15

Baseline H2O concentration % 1 Hg inlet concentration µg/Nm3 200

Carbon particle diameter µm 16 Carbon true density kg/m3 2200

Carbon particle porosity - 0.73 Carbon pore radius nm 10

Carbon injection rate mg/m3 8-80 Filtration time s 1500

Air to cloth ratio m/min 1.2 Time for new cake layer min 0.5-5

Henry’s constant preexponential factor

m3/g 0.869

Heat of adsorption J/mol -8543 Collocation point number - 10

256

The effect of the time for new carbon layer addition t on the mercury

removal efficiency of the fabric filter is illustrated in figure 11.5. Generally the more

frequently the new layer is added the larger mercury removal efficiency is obtained

for filtration time less than 1200 s. Smoother mercury removal efficiency curve will

be obtained using a smaller time interval for adding a carbon layer. For short

filtration time the new carbon layer should be added very fast in the simulation,

otherwise the simulated mercury removal efficiency will be smaller due to the delay

of new carbon layer addition. For filtration time larger than 1200 s same mercury

removal efficiency is predicted with 5 min interval for new carbon layer addition as

with smaller time interval. However, the computation time of the program is

considerably reduced using an interval of 5 min compared to 1 min.

0 300 600 900 1200 1500 18000

10

20

30

40

50

60

70

Time (s)

Mer

cury

rem

ova

l, %

30 s60 s120 s300s

Figure 11.5. Simulated effects of new cake layer addition frequency on the mercury

removal efficiency of a fabric filter without cleaning of the bags. 16 mg/m3 Darco Hg

carbon with a diameter of 16 µm is injected at 138C to simulated cement kiln flue

gas with 200 µg Hg0/Nm3, 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2,

15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

257

Figure 11.6 presents the simulated mercury removal efficiency by a fabric

filter without cleaning of the bags as a function of filtration time and injection rate of

carbon. Every minute a layer of carbon is added to the bag surface. As expected,

larger mercury removal efficiency is obtained with higher carbon injection rate. The

mercury removal efficiency increases fast with time after initiating carbon injection

up to 600 s and then it slowly increases with filtration time.

Figure 11.6. Simulated mercury removal efficiency by a fabric filter without cleaning

of the bags as a function of filtration time and injection rate of carbon. Darco Hg

carbon with a diameter of 16 µm is injected at 138C to simulated cement kiln flue

gas with 200 µg Hg0/Nm3, 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2,

15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. t= 1 min.

11.4 Fabric filter model 

In reality the filter bags are periodically cleaned. Therefore the filter cake

model needs to be extended with periodical cleaning of the bags in order to simulate

mercury adsorption in the bag filter. Assuming the cleaning cycle interval is tclean,

fraction of filter cleaned per cycle is fclean. After first tclean, the carbon cakes on the

258

bag surface have different exposure time to mercury and the mercury removal

efficiency is the average of the mercury removal efficiency in different sections of the

filter. Table 11.3 illustrates the carbon cake life time in different sections of the fabric

filter and the calculation of average mercury removal efficiency across the filter as a

function of filtration time. Pulse duration is selected as 0.1 second. Symbol * means a

pulse with 0.1/60 min and indicates that a fraction of the filter bags is cleaned.

Table 11.3 Illustration of carbon cake lifetime for different filter sections due to

periodic cleaning of bags. Here tclean=25 min; fclean=0.1.

Filtration time (min)

Exposure time of different filter sections

Average mercury removal efficiency

0 0 0 0-25 10@[0 25]

[0 25]

25 10@[25] [25]

25* 1@[0], 9@[25] [ 25 ]

9

10

25*-50 1@[0 25], 9@[0 50] [ 0 25 ] [ 0 50 ]

19

1 0

50 1@[25], 9@[50] [ 25 ] [50 ]

19

10

50* 1@[0], 1@[25], 8@[50] [ 2 5 ] [ 5 0 ]

18

1 0

50*-75 1@[0-25], 1@[0 50], 8@[0 75]

[ 0 2 5 ] [ 0 5 0 ] [ 0 7 5 ]

18

1 0

75 1@[25], 1@[50], 8@[75] [ 25 ] [50 ] [ 75 ]

18

10

75* 1@[0], 1@[25], 1@[50], 7@[75]

[ 2 5 ] [ 5 0 ] [ 7 5 ]

17

10

75*-100 1@[0 25], 1@[0 50], 1@[ 0 75], 7@[0 100] [ 0 25 ] [ 0 50 ] [ 0 75 ] [ 0 100 ]

17

10

100 1@[25], 1@[50], 1@[75], 7@[100] [ 25 ] [ 50 ] [ 75 ] [100 ]

17

10

100* 1@[0], 1@[25], 1@[50], 1@[75], 6@[100] [ 25 ] [50 ] [ 75 ] [100 ]

16

10

100*-125 1@[0 25], 1@[0 50], 1@[0 75], 1@[0 100],

6@[0 125]

[0 25] [0 50 ] [0 75] [0 100 ]

[0 125]

1

610

125 1@[25], 1@[50], 1@[75], 1@[100],

6@[125]

[25] [50] [75] [100]

[125]

1

610

125* 1@[0], 1@[25], 1@[50], 1@[75], 1@[100],

5@[125]

[25] [50] [75] [100]

[125]

1

510

259

Filtration time (min)

Exposure time of different filter sections

Average mercury removal efficiency

125*-150 1@[0 25], 1@[0 50], 1@[0 75], 1@[0 100], 1@[0 125], 5@[0 150]

[0 25] [0 50] [0 75] [0 100]

[0 125] [0 150]

1

510

150 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 5@[150]

[25] [50] [75] [100]

[125] [150]

1

510

150* 1@[0], 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 4@[150]

[25] [50] [75] [100]

[125] [150]

1

410

150*-175 1@[0 25], 1@[0 50], 1@[0 75], 1@[0 100], 1@[0 125], 1@[0 150],

4@[0 175]

[0 25] [0 50] [0 75] [0 100]

[0 125] [0 150] [0 175]

1

410

175 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150],

4@[175]

[25] [50] [75] [100]

[125] [150] [175]

1

410

175* 1@[0],1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150],

3@[175]

[25] [50] [75] [100]

[125] [150] [175]

1

310

175*-200 1@[0 25], 1@[0 50], 1@[0 75], 1@[0 100], 1@[0 125], 1@[0 150], 1@[0 175],3@[0 200]

[0 25] [0 50] [0 75] [0 100]

[0 125] [0 150] [0 175] [0 200]

1

310

200 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150], 1@[175], 3@[200]

[25] [50] [75] [100]

[125] [150] [175] [200]

1

310

200* 1@[0], 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150], 1@[175], 2@[200]

[25] [50] [75] [100]

[125] [150] [175] [200]

1

210

200*-225 1@[0 25], 1@[0 50], 1@[0 75], 1@[0 100], 1@[0 125], 1@[0 150], 1@[0 175], 1@[0 200],

2@[0 225]

[0 25] [0 50] [0 75] [0 100]

[0 125] [0 150] [0 175] [0 200]

[0 225]

1

102

225 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150], 1@[175], 1@[200],

2@[225]

[25] [50] [75] [100]

[125] [150] [175] [200]

[225]

1

102

225* 1@[0], 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150], 1@[175], 1@[200],

1@[225]

[25] [50] [75] [100]

[125] [150] [175] [200]

[225]

1

10

260

225*-250 1@[0 25], 1@[0 50], 1@[0 75], 1@[0 100], 1@[0 125], 1@[0 150], 1@[0 175], 1@[0 200], 1@[0 225], 1@[0 250]

[0 25] [0 50] [0 75] [0 100]

[0 125] [0 150] [0 175] [0 200]

[0 225] [0 250]

1

10

250 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150], 1@[175], 1@[200], 1@[225], 1@[250]

[25] [50] [75] [100]

[125] [150] [175] [200]

[225] [250]

1

10

250* 1@[0], 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150], 1@[175], 1@[200],

1@[225]

[25] [50] [75] [100]

[125] [150] [175] [200]

[225]

1

10

250*-275 1@[0 25], 1@[0 50], 1@[0 75], 1@[0 100], 1@[0 125], 1@[0 150], 1@[0 175], 1@[0 200], 1@[0 225], 1@[0 250]

[0 25] [0 50] [0 75] [0 100]

[0 125] [0 150] [0 175] [0 200]

[0 225] [0 250]

1

10

275 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150], 1@[175], 1@[200], 1@[225], 1@[250]

[25] [50] [75] [100]

[125] [150] [175] [200]

[225] [250]

1

10

275* 1@[0], 1@[25], 1@[50], 1@[75], 1@[100], 1@[125], 1@[150], 1@[175], 1@[200],

1@[225]

[25] [50] [75] [100]

[125] [150] [175] [200]

[225]

1

10

… … …

The filter cake model is run for time interval of [0 250] min. The calculated

mercury removal efficiencies at different time are used to calculate the corresponding

average mercury removal efficiency across the whole fabric filter.

The input parameters from Durkee slipstream tests listed in table 11.2 are

again used as model inputs to the fabric filter model. Other inputs include a bag

cleaning interval of 25 min and a cleaning fraction of 0.1. It is assumed that a new

sorbent layer is accumulated on the filter bag every 5 min.

Figure 11.7 shows the simulated mercury removal efficiency if the filter was

running without periodical cleaning up to 4 h. Compared to the short filtration time of

25 min as shown in figure 11.5, the mercury removal efficiency reaches a stable value

after about 1 h for the applied injection rates of powdered activated carbon. This

261

behavior is due to the growing thickness of the carbon cake. Fresh carbon is

continuously injected to the filter, providing increased mercury adsorption. At long

times, the inner layers of the carbon cake, consisting of almost fully spent carbon,

gives negligible contribution to the process so that asymptotic conditions are reached.

Figure 11.7. Simulated mercury removal efficiency by 1/10 of the fabric filter without

cleaning of the bags as a function of filtration time and injection rate of carbon. Darco

Hg carbon with a diameter of 16 µm is injected at 138C to simulated cement kiln

flue gas with 200 µg Hg0/Nm3, 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv

SO2, 15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. t=5 min, Air to cloth ratio

ufilter= 1.2 m/min.

Figure 11.8 presents the simulated mercury removal efficiency across the

fabric filter with periodical cleaning 10% of the bags every 25 min. When 10% of the

bags is cleaned the mercury removal efficiency decreases. At the beginning the

mercury removal efficiency decreases slightly and it decreases more at later stage.

This is due to the fact that more carbon is collected on the filter bag and is removed

by pulse cleaning. The model assumes that all the carbon collected on the bag is

completely removed from the bag surface and the corresponding mercury removal

efficiency for this fraction of bags drops to zero when the pulse cleaning is initiated.

The mercury removal efficiency across the whole filter reaches a stable level after all

262

the bags have been cleaned once. The less smooth curve is due to the applied time

interval of 5 min for a new carbon layer addition. Only five data points are used in a

cleaning interval of 25 min. This can be easily improved by decreasing the time

interval of new carbon cake layer addition at the expense of longer computation time.

Figure 11.8. Simulated mercury removal efficiency of the fabric filter with cleaning

of the bags as a function of filtration time and carbon injection rate. Darco Hg carbon

with a diameter of 16 µm is injected at 138C to simulated cement kiln flue gas with

200 µg Hg0/Nm3, 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15

vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. tclean=25 min, t=5 min, ufilter=1.2 m/min.

Figure 11.9 shows the simulated effects of flue gas temperature on the

mercury removal efficiency of the fabric filter. When the flue gas temperature is

reduced from 138C to 75C an improvement of about 8% mercury removal

efficiency is obtained. However, whether this improvement is economical needs to be

compared with additional costs by cooling down the flue gas.

263

0 1 2 3 4 5 6 70

10

20

30

40

50

60

70

80

Time (hour)

Ove

rall

bag

filte

r m

ercu

ry r

emov

al e

ffici

ency

, %1234

1: 75 degree C2: 100 degree C3: 115 degree C4: 138 degree C

Figure 11.9. Simulated effects of flue gas temperature on mercury removal efficiency

of the fabric filter. 16 mg/m3 Darco Hg carbon with a diameter of 16 µm is injected to

simulated cement kiln flue gas with 200 µg Hg0/Nm3, 1000 ppmv NO, 5 ppmv NO2,

10 ppmv HCl, 5 ppmv SO2, 15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. tclean=25

min, t=5 min, ufilter= 1.2 m/min.

 

11.5 Two­stage model 

Models developed in previous sections simulate mercury removal by separate

parts of the full-scale process. In this section mercury adsorption in the duct section is

coupled with the fabric filter section. First the duct model is run. The mercury

concentrations inside the particle and the bulk mercury concentration at the end of the

duct are used as initial conditions for the fabric filter model. Then the fabric filter

model is run to desired filtration time. The model inputs are the same as the fabric

filter model. A flight time of 1 s in the duct is applied.

Figure 11.10 shows the simulated mercury removal efficiency in the duct

section at Durkee cement plant. When a smaller carbon injection rate is applied the

mercury removal efficiency after 1 s in the duct is negligible. As shown in figure

264

11.10, about 2% mercury removal is obtained when 16 mg/m3 Darco Hg carbon is

injected. However, at larger carbon injection rates the mercury removal efficiency

after 1 s in the duct is noticeable. Therefore, high mercury removal efficiency can be

obtained by increasing the residence time of carbon particles in the duct, i.e., by

applying long duct, provided that there is enough space in the plant and large carbon

injection rate is applied.

Figure 11.10. Simulated mercury removal efficiency in the duct as a function of

residence time and injection rate of carbon. Darco Hg carbon with a diameter of 16

µm is injected at 138C to simulated cement kiln flue gas with 200 µg Hg0/Nm3,

1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15 vol.% H2O, 21 vol.%

CO2, and 6 vol.% O2.

Figure 11.11 compares the simulated and measured mercury removal

efficiency of the fabric filter at Durkee slipstream plant. Generally there is good

agreement between the simulation and pilot-scale data. This indicates that the

adsorption kinetics derived from 10 mg Darco Hg carbon in 2 g sand is reasonable

and the developed model is a useful tool to simulate and optimize the carbon injection

process.

265

0 10 20 30 40 50 60 70 80 90 100

Activated carbon injection rate (mg/m3)

0

10

20

30

40

50

60

70

80

90

100

Mer

cury

rem

ova

l ove

r th

e fi

lter

(%

)

ModelData

Figure 11.11. Comparison of simulated and measured mercury removal efficiency of

the fabric filter at Durkee slipstream plant. Darco Hg carbon with a diameter of 16

µm is injected at 138C to simulated cement kiln flue gas with 200 µg Hg0/Nm3,

1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15 vol.% H2O, 21 vol.%

CO2, and 6 vol.% O2. tclean=25 min, t=5 min, ufilter= 1.2 m/min.

The overall mercury removal efficiency of the sorbent injection system refers

to mercury removal from the carbon injection point to the fabric filter outlet and can

be evaluated as following:

100 (1 (1 %)(1 %))total duct filter (11.33)

Table 11.4 summarizes the calculated mercury removal efficiencies in the duct, fabric

filter and the whole carbon injection system for different carbon injection rates. The

contribution of mercury removal in the duct is much smaller to the mercury removal

in the whole carbon injection system. However, data regarding overall mercury

removal for the Durkee slipstream plant are not available for comparison.

266

Table 11.4 Simulated mercury removal efficiencies in the duct, fabric filter and the

whole carbon injection system for different carbon injection rates. Darco Hg carbon

with a diameter of 16 µm is injected at 138C to simulated cement kiln flue gas with

200 µg Hg0/Nm3, 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15

vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. Carbon particle residence time in the

duct is 1 s. Every five minute a layer of carbon is added to the bag surface. Air to

cloth ration is 1.2 m/min.

Carbon load (mg/m3)

Mercury removal in duct (%)

Mercury removal in fabric filter (%)

Total Mercury removal (%)

8 1.0 54.0 54.5 16 2.0 69.6 70.2 48 5.9 86.0 86.9 80 9.7 90.3 91.2

The applied carbon injection rate for mercury control is much smaller than the

typical dust load in the flue gas for particulate emission control process. Therefore,

the pressure drop over the fabric filter is expected to increase slowly with filtration

time for mercury control process. It is then feasible to extend the bag cleaning

interval, i.e., use less frequent cleaning of the bags. Figure 11.12 illustrates the

simulated effects of bag cleaning frequency on the mercury removal efficiency of the

fabric filter. The mercury removal efficiency slightly increases when the bag cleaning

interval is increased. Extending the bag cleaning interval from 25 min to 100 min

results in a 1.3% improvement of mercury removal efficiency. A longer bag cleaning

cycle results in longer retention time of the carbon particles on the bags, which allows

the carbon particles to adsorb more mercury from the flue gas.

267

25 50 75 100 125

Bag cleaning interval (min)

86

86.4

86.8

87.2

87.6

88

Hg

rem

ova

l eff

icie

nc

y, %

Figure 11.12. Simulated effects of bag cleaning frequency on mercury removal

efficiency of the fabric filter. 48 mg/m3 Darco Hg carbon with a diameter of 16 µm is

injected at 138C to simulated cement kiln flue gas with 200 µg Hg0/Nm3, 1000

ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15 vol.% H2O, 21 vol.% CO2,

and 6 vol.% O2. tclean=25 min, t=5 min, ufilter= 1.2 m/min.

The only carbon injection system for mercury control from cement production

is operated at Ash Grove’s Durkee cement plant [10]. Instead of continuous injection

of powdered activated carbon, the fabric filter works in fixed-bed adsorber mode.

During the first run period, the activated carbon was added to the system through the

first day and removed from the system through the eighth day. The average removal

efficiency during those intervening 6 days was 92.8% [10]. However, the carbon

injection rate and duration in the first day is not reported.

Simulations are performed to simulate the fabric filter fixed-bed operation

mode. Firstly activated carbon is injected at high load for 30 min to form a carbon

cake on the bags. Then the activated carbon injection is stopped and the fabric filter

works as a fixed-bed adsorber. When the initial mercury breakthrough occurs or the

mercury emission limit is reached, the bags are cleaned by pulse-jet compressed air

and later activated carbon is injected again. The injection-adsorption-cleaning cycle is

repeated. The carbon injection without bag cleaning period is simulated by the filter

cake model and the fixed-bed adsorption period is simulated by the fixed-bed model.

268

Figure 11.13 shows the simulated mercury breakthrough curves of the fabric

filter injected with 0.5-2.0 g/m3 Darco Hg carbon for 30 min. The actual carbon

injection rate at full-scale test is not available and high carbon injection rates are

tested here to show the effect. The initial mercury breakthrough time for carbon load

of 0.5, 1.0, 1.5, and 2.0 g/m3 is 17.9, 35.7, 53.6, and 71.5 h, respectively. This means

that about 100% mercury removal efficiency is obtained within the initial mercury

breakthrough periods. With an activated carbon injection of 1-2 g/m3 for 30 min, the

fabric filter can work for 53.6 to 71.5 h before the initial breakthrough occurs. This

means that the activated carbon injection is only required 2-3 times per week. This is

in agreement with the information obtained from Paone [9] on practice of the Durkee

sorbent injection plant. However, the actual carbon injection rate and duration at

Durkee plant are unknown and are required for validation of the simulation.

0 50 100 150 200 250 300 350 4000

0.2

0.4

0.6

0.8

1

1.2

Time (hour)

Gas

eous

mer

cury

out

let

(Cou

t/C

in)

A

A: 0.5 g/m3B: 1.0 g/m3C: 1.5 g/m3D: 2.0 g/m3

CB D

Figure 11.13. Simulated mercury breakthrough curves of fabric filter injected with

large carbon loads for 30 min and then in fixed-bed operation mode. Darco Hg

carbon with a diameter of 16 µm is injected at 138C to simulated cement kiln flue

gas with 200 µg Hg0/Nm3, 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2,

15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.

269

To compare with mercury removal by continuous injection of carbon,

simulations are conducted by injection of the same total amount of carbon within the

initial mercury breakthrough period. The corresponding carbon injection rate is 14

mg/m3 and the simulated mercury removal efficiency is only 66.8%. The fabric filter

works more efficiently for mercury removal in fixed-bed operation model than

continuous injection of carbon for the same total amount of carbon injected. This is

probably due to the larger amount of carbon accumulated on the bags in the fixed-bed

operation mode. The fixed-bed operation model is limited by the pressure drop over

the fabric filter. It is expected that a larger pressure drop over the filter and power

consumption of the system is required when a large carbon injection rate is applied in

a reasonable period.

11.6 Conclusions 

The developed single particle and fixed-bed adsorption models are further

extended to duct and fabric filter models to simulate mercury removal by carbon

injection upstream of a fabric filter. The fabric filter model is accounted for by adding

a new carbon layer on the bag surface after a short time and treating each layer as a

well mixed tank. Finally the duct model and fabric filter model are coupled to a two-

stage model. The mercury concentrations inside the particle and the bulk mercury

concentration at the end of the duct are used as initial conditions for the fabric filter

model. The models are based on materials balances in both gaseous and adsorbed

phase along the duct length/growing filter cake and inside the carbon particles. The

models account for adsorption kinetics, both the external and internal mass transfer

resistances, accumulation of carbon layer on the bags, and periodical cleaning of the

bags.

Henry’s constant obtained from fixed-bed investigation of mercury adsorption

by activated carbon in the simulated cement kiln flue gas is used as input to the

models. The effects of SO2, H2O, NO2 levels in the flue gas on mercury removal are

accounted by using correlations derived from the fixed-bed investigation.

Duct model simulations indicate that large carbon loading in the flue gas are

required to obtain high mercury removal efficiency due to the short residence time.

270

To minimize the carbon feed rate it is advisable to lower the operating temperature.

Improvements in the mercury removal efficiency can be obtained also by increasing

the in-duct particle residence time and decreasing the carbon particle size.

In contrast to the in-duct removal process, simulations of mercury adsorption

in the fabric filter show that higher mercury removal efficiency can be achieved with

moderate carbon consumption due to the effective gas/carbon contact on the filter

bags. The effects of carbon load, temperature, frequency of new carbon layer addition

and bag cleaning on mercury removal efficiency are simulated. The fabric filter

model can predict the mercury removal profile with jagged nature because of the

intermittent partial cleaning of the bags. Comparison with simulation and

experimental data from Durkee cement plant slipstream tests shows that the

developed two-stage model can reasonably predict the mercury removal from cement

plants by carbon injection upstream of a fabric filter.

Minor benefits can be obtained by increasing the cleaning cycle time of the

fabric filter compartments. The fabric filter works more efficiently on mercury

removal when it is operated as fixed-bed adsorbed by injection of high carbon load in

short time and then stopping carbon injection and cleaning of the bags.

11.7 List of symbols 

A: total exterior surface area of all particles in the adsorber (m2) A: matrix in equation 11.18 B: matrix in equation 11.19 Bi: dimensionless Biot number C: gaseous mercury concentration (µg/m3) C: matrix in equation 11.20 and 11.21 Cb: gas bulk mercury concentration (µg/m3) Cbo: initial gas bulk mercury concentration (µg/m3) Cbin: inlet mercury concentration in tank (µg/m3) Cµ: adsorbed mercury concentration in the sorbent (µg/m3) Dapp: apparent diffusion coefficient (m2/s) De: effective diffusion coefficient (m2/s) fclean: fraction of bags cleaned per pulse cleaning J: mercury flux (µg/m2) K: Henry’s constant

271

km: gas film mass transfer coefficient (m/s) L: thickness of carbon cake (m) mp: mass of carbon particle in the adsorber (g) M: carbon load in the tank (mg) n: number of interior collocation points r: radial coordinate (m) Rp: carbon particle radius (m) t: time (s) tclean: time interval for bag cleaning (25) u: dimensionless parameter, u=x2 u: face velocity on the filter bags (m/s) ufilter: air to cloth ratio (m/min) V: volume of the adsorber (m3) Vb: volume of the adsorber (m3) x: dimensionless radius y: dimensionless mercury concentration yb: dimensionless gas bulk mercury concentration (µg/m3) ybin: dimensionless inlet mercury concentration in tank (µg/m3) Greek symbols b: bed void fraction p: carbon particle porosity p: carbon particle density (kg/m3) : dimensionless time : carbon load in the flue gas (kg/m3) : mercury removal efficiency (%)

11.8 References 

[1] J.R.V. Flora, R.A. Hargis, W.J. O'Dowd, A. Karash, H.W. Pennline, R.D. Vidic, The role

of pressure drop and flow redistribution on modeling mercury control using sorbent injection

in baghouse filters, J. Air Waste Manage. Assoc. 56 (2006) 343-349.

[2] F. Scala, Simulation of mercury capture by activated carbon injection in incinerator flue

gas. 1. In-duct removal, Environ. Sci. Technol. 35 (2001) 4367-4372.

[3] F. Scala, Simulation of mercury capture by activated carbon injection in incinerator flue

gas. 2. Fabric filter removal, Environ. Sci. Technol. 35 (2001) 4373-4378.

[4] F. Scala, Modeling mercury capture in coal-fired power plant flue gas, Ind Eng Chem Res.

43 (2004) 2575-2589.

[5] J. Villadsen, M.L. Michelsen, Solutions of differential equation models by polynomial

approximation, Prentice-Hall, Inc., 1978.

272

[6] D.D. Do, Adsorption analysis: equilibria and kinetics, Imperial College Press, 1998.

[7] L. Hayes-Gorman, Regulating mercury emissions: Ash Grove Cement in Durkee, Air

toxics summit 2008, Boise, Idaho, 4-7 August, 2008.

[8] Schreiber & Yonley Associates, Mercury emissions test report, Ash Grove Cement

Company Durkee, Oregon, Project No. 060204, 2007.

[9] P. Paone, Personal communication about flue gas compositions and temperature for pilot-

scale sorbent injection tests at Ash Grove Durkee plant and FLSmidth Mineral Lab, 2010.

[10] Curtis D. Lesslie, Mail to U.S.EPA about initial results of Ash Grove's Durkee sorbent

injection system, http://www.whitehouse.gov/sites/default/files/omb/assets/ oira_2060/2060_

07292010-3.pdf, visited March 21, 2011.

273

12

Concluding remarks

To develop and get a better understanding of mercury removal from cement

plant by sorbent injection upstream of a pulse jet fabric filter, this project has focused

on four areas: comprehensive review of mercury emission from cement plants and

analysis applicability of available technologies for mercury removal from cement

plants, test and development of thermal catalytic converters for oxidized mercury

reduction and dynamic total mercury measurement, screening tests and fundamental

investigation of mercury adsorption by sorbents in simulated cement kiln flue gas,

and development of mathematic models that can describe mercury removal by fixed-

bed and carbon injection upstream of a fabric filter.

Cement plants are quite different from power plants and waste incinerators

regarding the flue gas composition, temperature, residence time, and inherent

material circulation. Instead of fuel, cement raw materials are the dominant sources of

mercury in the cement kiln flue gas. The mercury emissions and speciation from

cement kilns can vary over time and depend on raw materials and fuels used, and

process operation. Among the available technologies for mercury removal from flue

gas, sorbent injection upstream of a polishing fabric filter is considered as the most

promising and suitable technology for cement plant application.

To be able to perform dynamic measurement of mercury adsorption by

sorbents, a red brass chips and sulfite converter were investigated in simulated

cement kiln flue gas. The red brass converter works only when measuring elemental

mercury in nitrogen and does not work properly even when only elemental mercury is

added to the simulated flue gas. The red brass converter cannot fully reduce HgCl2 to

elemental mercury under any relevant condition.

The sodium sulfite converter material was prepared by dry impregnation of

sodium sulfite and calcium sulfate powders on zeolite pellets using water glass as

274

binder. The sulfite converter works well at 500C when less than 10 ppmv HCl is

present in the simulated cement kiln flue gas. The response time of the sulfite

converter is short and typically within at most two minutes, which makes it

appropriate for not too fast dynamic measurements.

Inconsistent mercury adsorption capacity of activated carbon is observed at

different carbon loads in 2 g sand. A smaller mercury adsorption capacity is obtained

with larger carbon load. Tests with elemental mercury and mercury chloride, different

carbon type and particle sizes show the same trend. Effects of bed dilution on the

equilibrium mercury adsorption capacity appear to be limited.

Screening tests of sorbents for mercury removal from cement plants have

been conducted in the fixed-bed reactor system using simulated cement kiln flue gas

with elemental mercury and mercury chloride source. The tested sorbents include

commercial activated carbons, commercial non-carbon sorbents, and cement

materials. Screening measurements are used to evaluate initial mercury capture rate,

oxidation potential, and capacity for the selected sorbents.

The sorbents don’t adsorb any mercury when tested with elemental mercury

in nitrogen. Tests of a range of 30 mg collected non-carbon based sorbents and

cement materials as sorbents in 2 g sand at 150C in simulated cement kiln flue gas

with elemental mercury do not show any mercury adsorption or oxidation. Generally

a larger amount of adsorbed mercury is obtained with sorbents that have larger

mercury oxidation capacity. While all the non-carbon based sorbents and cement

materials show some adsorption of mercury chloride. This indicates that mercury

oxidation is an important factor for mercury adsorption by the sorbents. Elemental

mercury needs to be oxidized either in the flue gas with HCl or on the sorbent.

Among the tested sorbents the Darco Hg activated shows the best performance of

adsorption of both elemental and oxidized mercury and is recommended as the

reference sorbent for fundamental investigation.

A parametric study of elemental mercury adsorption by activated carbon has

been conducted in the fixed-bed reactor by mixing 10 mg Darco Hg carbon with 2 g

sand. Increasing adsorption temperature results in decreased equilibrium mercury

adsorption capacity of the activated carbon. The mercury adsorption isotherm follows

275

Henry’s law for the applied mercury inlet levels in this project. The derived heat of

adsorption is -8540 J/mol for elemental mercury adsorption by Darco Hg activated

carbon in simulated cement kiln flue gas. Higher mercury oxidation and initial

adsorption rate are also observed for smaller carbon particles, while the equilibrium

mercury adsorption capacity is the same.

The mercury adsorption capacity does not change with O2, CO, and NO levels

in the flue gas, but decreases when CO2, H2O, SO2, and NO2 concentrations increase.

The decrease of mercury adsorption capacity is due to the competition for active site

with mercury by CO2 and H2O, and conversion of the previously formed nonvolatile

basic mercuric nitrate into the volatile form by interactions between SO2 and NO2.

Slight promoting effects of HCl on mercury adsorption are observed when HCl

concentration is varied in the range of 0.5-20 ppmv. Larger mercury adsorption

capacity is obtained when HCl is removed from baseline gas because HgO(s) is

formed on the carbon.

Similar adsorption behaviors of mercury chloride and elemental mercury by

Darco Hg activated carbon are observed using simulated cement kiln flue gas. This is

due to the effective catalytic oxidation of elemental mercury by the activated carbon.

Mathematical models are developed to simulate mercury adsorption by a

single carbon particle, fixed carbon bed, in the duct and fabric filter. Orthogonal

collocation method is used to solve mercury diffusion and adsorption inside a carbon

particle. The fixed-bed model is solved by tank-in-series method. The fabric filter

model is accounted for by adding a new carbon layer on the bag surface after a short

time as a well mixed tank. The two-stage duct-fabric filter model accounts for

adsorption kinetics, both the external and internal mass transfer resistances,

accumulation of carbon layer on the bags, and periodical cleaning of the bags.

Henry’s constant obtained from fixed-bed investigation are used as input to

the models. The developed fixed bed model can reasonably simulate the effects of

adsorption temperature, mercury inlet concentration, flow gas rate, carbon particle

size, and SO2, H2O, NO2 level in the flue gas on the mercury breakthrough curve of

the fixed carbon bed.

276

Duct model simulations indicate that a large carbon load is required to obtain

a high mercury removal efficiency due to the short residence time. Simulations of

mercury adsorption in the fabric filter show that higher mercury removal efficiency

can be achieved with moderate carbon consumption due to the effective gas/carbon

contact on the filter bags. The effects of carbon load, temperature, frequency of new

carbon layer addition and bag cleaning on mercury removal efficiency are simulated.

The fabric filter model can predict the mercury removal profile with jagged nature

because of the intermittent partial cleaning of the bags. Comparison with simulation

and experimental data from Durkee cement plant slipstream tests shows that the

developed two-stage model can reasonably predict the mercury removal from cement

plants by carbon injection upstream of a fabric filter.

277

13

Suggestions for further work

This work has investigated the mercury removal by carbon injection upstream

of a fabric filter under more controlled conditions using a fixed bed reactor. Pilot or

full-scale tests are desired to demonstrate the ability of the studied sorbents and

technology to control emissions of mercury from cement plant over a typical range of

operating conditions for an extended period of time and to further validate the

developed models. The condition in full-scale application is much more demanding

than in the lab-scale investigation. Further development and test of the sulfite

converter is required for dynamic measurement of mercury in large scale

investigation. A sampling probe is needed to separate the particles from the flue gas

efficiently without plugging. Adsorption of mercury by the dust and probe should be

minimized by high sampling flow rate and high heating temperature.

The problem of inconsistent mercury adsorption capacity for different carbon

loads could not be solved within the project. More thorough investigation is

necessary to reveal the cause. New analysis technology is required to reveal whether

mercury is adsorbed by the sand when it is mixed with activated carbon.

This project investigates only mercury removal by the activated carbon. In the

future multipollutants control by the activated carbon should be studied by measuring

also other harmful species such as SO2 and NOx. When more than one component is

involved in the adsorption system, adsorption equilibrium involving competition

between molecules of different types is needed for the understanding of the system as

well as for the design purposes.

To reduce the sorbent cost, regeneration of used sorbents should be

investigated. Recycling sorbent collected by the fabric filter to the injection process

also requires more investigation. Modification of cement materials by additives that

278

can oxidize mercury is attractive. However, the influence of the additives on the

cement quality needs to be investigated.

Models developed in this work assume that all the particles have a uniform

size. It is interesting to take the particle size distribution into account in the more

advanced model. The developed fabric filter model does not include pressure drop

over the filter. It is useful to incorporate the pressure development of the fabric filter

and pulse jet cleaning instead of assuming a constant bag cleaning interval. Current

models assume local equilibrium inside the carbon particle, simulation with a full

kinetics description of the adsorption process is necessary to investigate whether this

assumption is reasonable.