01.12.2012 Views

Mercury Removal from Cement Plants by Sorbent Injection ... - CHEC

Mercury Removal from Cement Plants by Sorbent Injection ... - CHEC

Mercury Removal from Cement Plants by Sorbent Injection ... - CHEC

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

TECHNICAL UNIVERSITY OF DENMARK<br />

DEPARTMENT OF CHEMICAL AND BIOCHEMICAL ENGINEERING<br />

Ph.D. Thesis, June 2011<br />

<strong>Mercury</strong> <strong>Removal</strong> <strong>from</strong> <strong>Cement</strong> <strong>Plants</strong> <strong>by</strong> <strong>Sorbent</strong> <strong>Injection</strong><br />

upstream of a Pulse Jet Fabric Filter<br />

Yuanjing Zheng


I<br />

Preface<br />

This thesis is written for partial fulfillment of the requirements to obtain the<br />

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

at the <strong>CHEC</strong> (Combustion and Harmful Emission Control) Research Centre at the<br />

Department of Chemical and Biochemical Engineering under the supervision of Prof.<br />

Anker Degn Jensen <strong>from</strong> <strong>CHEC</strong>, Department Manager Christian Windelin and<br />

Flemming Jensen <strong>from</strong> FLSmidth A/S. The project is financially supported <strong>by</strong> the<br />

Industrial PhD programme of the Danish Ministry of Science, Technology and<br />

Innovation, Danish Advanced Technology Foundation as part of the Research<br />

Platform on New <strong>Cement</strong> Production Technology.<br />

I would like to thank my supervisors, particularly Anker Degn Jensen, for<br />

their support, fruitful discussion and comments. Technician Thomas Wolfe and<br />

department workshop are gratefully acknowledged for help building the fixed-bed<br />

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

part of the manuscript. Student Jacob Clement Nielsen is acknowledged for<br />

performing some of the screening tests. Thanks to all the other people at <strong>CHEC</strong> and<br />

FLSmidth, not mentioned here, for their great help received during my study.<br />

Finally, I would like to thank my family and friends for their support and<br />

encouragement.<br />

Yuanjing Zheng<br />

Kgs. Lyng<strong>by</strong>, June 2011


II<br />

Abstract<br />

There are growing concerns over mercury emissions due to their toxicity,<br />

volatility, persistence, and bioaccumulation in the environment. <strong>Mercury</strong> emissions<br />

<strong>from</strong> cement plants are being regulated <strong>by</strong> environmental agencies in most countries.<br />

Among the available technologies for mercury removal <strong>from</strong> flue gas, sorbent<br />

injection upstream of a polishing fabric filter is considered as the most promising and<br />

suitable technology for cement plant application. <strong>Cement</strong> plants are quite different<br />

<strong>from</strong> power plants and waste incinerators regarding the flue gas composition,<br />

temperature, gas and solid residence time, and inherent material circulation. Thus<br />

knowledge obtained <strong>from</strong> mercury removal in power plants and incinerators might<br />

not be applied to cement plants directly and fundamental investigation under well<br />

controlled cement kiln condition is imperative.<br />

Tests in simulated cement kiln flue gas show that the red brass converter<br />

developed for waste incinerator application does not work properly for either<br />

elemental or total mercury measurement. Sodium sulfite converter is developed and<br />

optimized for oxidized mercury reduction and total mercury measurement. The<br />

response time of the sulfite converter is short, which makes it appropriate for<br />

dynamic measurement of mercury adsorption and oxidation <strong>by</strong> sorbents.<br />

Screening tests of sorbents for mercury removal <strong>from</strong> cement plants have<br />

been conducted in the fixed-bed reactor system using simulated cement kiln flue gas<br />

with elemental mercury and mercury chloride sources. The tested sorbents include<br />

commercial activated carbons, commercial non-carbon sorbents, and cement<br />

materials. With elemental mercury present in the flue gas, no mercury adsorption or<br />

oxidation <strong>by</strong> non-carbon based sorbents and cement materials is observed. Generally<br />

larger amount of adsorbed mercury is obtained with sorbents that have larger mercury<br />

oxidation capacity. While all the non-carbon based sorbents and cement materials<br />

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


activated shows the best performance of adsorption of both elemental and oxidized<br />

mercury and is recommended as the reference sorbent for fundamental investigation.<br />

Parametric studies of mercury adsorption <strong>by</strong> activated carbon have been<br />

conducted in the fixed-bed reactor regarding the effects of adsorption temperature,<br />

flue gas rate, mercury level, carbon particle size, carbon load, and flue gas<br />

composition. The mercury adsorption isotherm follows Henry’s law for the applied<br />

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

derived for model input. The mercury adsorption capacity does not change with O2,<br />

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

concentrations increase. Slight promoting effects of HCl on mercury adsorption are<br />

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

is obtained when HCl is removed <strong>from</strong> the gas. Similar adsorption behaviors of<br />

mercury chloride and elemental mercury <strong>by</strong> Darco Hg activated carbon are observed<br />

using simulated cement kiln flue gas, due to the effective catalytic oxidation of<br />

elemental mercury <strong>by</strong> the activated carbon.<br />

Mathematical models are developed to simulate mercury adsorption <strong>by</strong> a<br />

single carbon particle, fixed carbon bed, in the duct and fabric filter. The developed<br />

fixed bed model can reasonably simulate the mercury breakthrough curve of the fixed<br />

carbon bed. Comparison with fabric filter model simulations and experimental data<br />

<strong>from</strong> slipstream tests at a cement plant shows that the developed two-stage model is a<br />

valuable tool and can reasonably predict the mercury removal <strong>from</strong> cement plants <strong>by</strong><br />

carbon injection upstream of a fabric filter.<br />

III


Resumé (summary in Danish)<br />

Der er voksende bekymringer over kviksølvemissioner grundet disses<br />

giftighed, flygtighed, bestandighed og biologisk akkumulation i miljøet.<br />

Kviksølvemissioner fra cementfabrikker reguleres i de fleste lande af miljøorganer.<br />

Blandt de tilgængelige teknologier til fjernelse af kviksølv fra røggas anses<br />

sorbentinjektion opstrøms for et posefilter for den mest lovende og velegnede<br />

teknologi til anvendelse på cementfabrikker. <strong>Cement</strong>fabrikker er temmelig forskellige<br />

fra kraftværker og affaldsforbrændingsanlæg med hensyn til røggassammensætningen,<br />

temperatur, opholdstid af gas og faststof samt iboende materialecirkulation. Derfor<br />

kan viden opnået fra kviksølvfjernelse i kraftværker og affaldsforbrændingsanlæg<br />

ikke anvendes direkte på cementfabrikker og fundamental undersøgelse under<br />

velkontrollerede forhold svarende til cementfremstilling brændingsovn er essentielt.<br />

Test i simuleret røggas fra cementbrændingsovn viser, at en kommerciel<br />

konverter udviklet til anvendelse på affaldsforbrændingsanlægs ikke virker godt for<br />

hverken elementær kviksølvmåling eller total kviksølvmåling. Som en del af<br />

projektet er der udviklet en natriumsulfitkonverter til reduktion af oxyderet kviksølv<br />

samt total kviksølvmåling. Sulfit konverterens responstid er kort hvilket gør den<br />

velegnet til dynamisk måling af kviksølv adsorption og oxidation med sorbenter.<br />

Screeningsforsøg af sorbenter til fjernelse af kviksølv fra cementfabrikker er<br />

udført i et fixed bed reaktorsystem ved brug af simuleret røggas fra cementsovne med<br />

både elementært kviksølv samt kviksølvklorid. De testede sorbenter inkluderer<br />

kommercielle aktivt kul- og kommercielle ikke-kulstofsorbenter samt<br />

cementmaterialer. Med elementært kviksølv tilstede i røggassen blev hverken<br />

kviksølvadsorption eller -oxidation observeret med de ikke kulstofbaserede sorbenter<br />

og cementmaterialer. Generelt opnås større adsorberet mængde kviksølv med<br />

sorbenter der har større kviksølvoxidationskapacitet. Alle de ikke-kulstofbaserede<br />

sorbenter og cementmaterialer viser nogen adsorption af kviksølvklorid. Blandt de<br />

testede sorbenter udviser Darco Hg aktivt kul den bedste evne til adsorption af både<br />

IV


elementært og oxideret kviksølv og anbefales som referencesorbent i den<br />

fundamentale undersøgelse.<br />

Parameterstudier af kviksølvadsorption med aktivt kul er blevet udført i en<br />

fixed bed reaktor med hensyn til effekter af adsorptionstemperatur, røggasmængde,<br />

kviksølvniveau, kulstofpartikelstørrelse, kulstofbelastning og røggassammensætning.<br />

I dette projekt følger kviksølvadsorptionsisotermen Henrys lov for den anvendte<br />

koncentration af kviksølv. Henrys konstant og adsorptionsvarmen er fundet til<br />

indsættelse i model. Kviksølvadsorptionskapaciteten ændres ikke som følge af O2,<br />

CO og NO niveauer i røggassen, men falder når CO2, H2O, SO2, og NO2<br />

koncentrationerne stiger. En mindre positiv effekt af HCl på kviksølvadsorption er<br />

observeret med HCl i røggassen op til 20 ppmv. Større kviksølv adsorptionskapacitet<br />

opnås når HCl fjernes fra gassen. Lignende adsorptionsmønster for kviksølvklorid og<br />

elementært kviksølv med Darco Hg aktivt kul er observeret ved brug af simuleret<br />

røggas fra cementsovne, på grund af den effektive katalytiske oxidation af elementært<br />

kviksølv med det aktive kul.<br />

Matematiske modeller er udviklet til at simulere kviksølvadsorption på en<br />

enkel kulpartikel, i en fixed bed af aktivt kul, i kanalen og i posefilteret. Den<br />

udviklede fixed bed model med god nøjagtighed simulere kviksølv<br />

gennembrydningskurven for fixed bed forsøgen. Sammenligning af posefiltermodel<br />

simuleringer med eksperimentelle data fra slipstrømstests på en cementfabrik viser at<br />

den udviklede to-trins model er et værdifuldt værktøj der på fornuftigvis kan<br />

forudsige kviksølvfjernelsen fra cementfabrikker med kulstofinjektion opstrøms for et<br />

posefilter.<br />

V


VI<br />

Table of contents<br />

Preface ...........................................................................................................................I<br />

Abstract.........................................................................................................................II<br />

Resumé (summary in Danish).....................................................................................IV<br />

Table of contents.........................................................................................................VI<br />

1. Introduction............................................................................................................... 1<br />

1.1 Project background ............................................................................................. 1<br />

1.2 Project objectives................................................................................................ 3<br />

1.3 Outline of the thesis ............................................................................................ 3<br />

1.4 References........................................................................................................... 4<br />

2. <strong>Mercury</strong> emissions and transformations in cement plants........................................ 6<br />

2.1 <strong>Cement</strong> production processes ............................................................................. 6<br />

2.2 <strong>Mercury</strong> contents in fuels and cement raw materials ....................................... 12<br />

2.3 <strong>Mercury</strong> emissions............................................................................................ 14<br />

2.3 <strong>Mercury</strong> transformation during combustion..................................................... 15<br />

2.3.1 <strong>Mercury</strong> transformation in coal combustion flue gas ................................ 17<br />

2.3.2 <strong>Mercury</strong> transformation within cement kiln system.................................. 23<br />

2.4 Conclusions....................................................................................................... 27<br />

2.5 Further work ..................................................................................................... 28<br />

2.6 References......................................................................................................... 28<br />

3. Review of technologies for mercury removal <strong>from</strong> flue gas .................................. 32<br />

3.1 Introduction....................................................................................................... 32<br />

3.2 <strong>Mercury</strong> avoidance technology......................................................................... 33<br />

3.2.1 Coal cleaning ............................................................................................. 33<br />

3.2.2 <strong>Cement</strong> raw material cleaning ................................................................... 33<br />

3.2.3 Fuel switching............................................................................................ 34<br />

3.3 <strong>Mercury</strong> removal <strong>by</strong> powdered activated carbon injection .............................. 35<br />

3.3.1 Parameters affecting mercury removal <strong>by</strong> activated carbon injection....... 35<br />

3.3.2 Tests of mercury sorbents in lab-scale fixed-bed reactors......................... 38<br />

3.3.3 <strong>Sorbent</strong> injection in power plants .............................................................. 49<br />

3.3.5 Carbon surface chemistry and mechanisms of mercury capture on carbons<br />

............................................................................................................................ 58<br />

3.3.6 Processing and reuse of mercury laden activated carbon .......................... 63<br />

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


3.4 <strong>Mercury</strong> removal <strong>by</strong> activated carbon bed ....................................................... 65<br />

3.5 <strong>Mercury</strong> control <strong>by</strong> flue gas desulphurization systems .................................... 67<br />

3.6 <strong>Mercury</strong> removal <strong>by</strong> sodium tetrasulfide injection........................................... 68<br />

3.7 Enhanced mercury removal <strong>by</strong> oxidation ......................................................... 69<br />

3.8 <strong>Mercury</strong> removal <strong>by</strong> roaster process................................................................. 72<br />

3.9 Conclusions....................................................................................................... 73<br />

3.10 Further research requirement.......................................................................... 75<br />

3.11 Abbreviations.................................................................................................. 75<br />

3.12 References....................................................................................................... 76<br />

4. Experimental methods and materials...................................................................... 86<br />

4.1 Description of the fixed-bed reactor system..................................................... 86<br />

4.1.1 Gas mixing system..................................................................................... 88<br />

4.1.2 <strong>Mercury</strong> vapor addition system ................................................................. 88<br />

4.1.3 Humidifier for water vapor addition.......................................................... 90<br />

4.1.4 Low temperature furnace and fixed-bed reactor........................................ 92<br />

4.1.5 <strong>Mercury</strong> analysis system............................................................................ 93<br />

4.2 Converter and sorbent materials ..................................................................... 100<br />

4.3 Flue gas composition ...................................................................................... 103<br />

4.4 <strong>Sorbent</strong> load in fixed-bed test ......................................................................... 103<br />

4.5 Experimental procedure.................................................................................. 105<br />

4.6 <strong>Sorbent</strong> characterization ................................................................................. 106<br />

4.6.1 Scanning electron microscopy ................................................................. 106<br />

4.6.2 Particle size distribution........................................................................... 107<br />

4.6.3 Analysis of mercury in sorbent................................................................ 108<br />

4.7 References........................................................................................................... 108<br />

Appendix............................................................................................................... 110<br />

4A Check of mercury analyzer ............................................................................. 110<br />

4B Water addition verification ............................................................................. 112<br />

5. Dynamic measurement of mercury adsorption and oxidation on activated carbon in<br />

simulated cement kiln flue gas.................................................................................. 117<br />

5.1 Review of gaseous mercury measurement technology................................... 117<br />

5.2 Performance test of the mercury analyzer ..................................................... 119<br />

5.3 Performance test of the red brass converter.................................................... 121<br />

5.4 Performance of the sulfite converter............................................................... 125<br />

5.5 Examples of dynamic measurement of mercury adsorption and oxidation on<br />

activated carbon .................................................................................................... 131<br />

5.6 Suggestions for practical application of the converter.................................... 132<br />

5.7 Conclusions..................................................................................................... 133<br />

VII


5.8 References....................................................................................................... 134<br />

6. Effects of bed dilution and carbon load on mercury adsorption capacity of activated<br />

carbon........................................................................................................................ 137<br />

6.1 Introduction..................................................................................................... 137<br />

6.2 Effects of carbon load..................................................................................... 137<br />

6.3 Effects of bed dilution..................................................................................... 141<br />

6.4 Effects of sand load......................................................................................... 143<br />

6.5 Effects of carbon loading location.................................................................. 144<br />

6.6 Effects of bed materials .................................................................................. 145<br />

6.7 Effects of carbon type and particle size .......................................................... 146<br />

6.8 Tests with only Portland cement..................................................................... 147<br />

6.9 Conclusions..................................................................................................... 148<br />

6.10 References..................................................................................................... 149<br />

7. Screening tests of mercury sorbents ..................................................................... 151<br />

7.1 Introduction..................................................................................................... 151<br />

7.2 <strong>Sorbent</strong> properties and compositions.............................................................. 153<br />

7.3 SEM-EDX analysis of fresh sorbents ............................................................. 157<br />

7.4 Baseline test .................................................................................................... 160<br />

7.5 Screening tests in nitrogen.............................................................................. 160<br />

7.6 Screening tests in simulated cement kiln flue gas with elemental mercury<br />

source.................................................................................................................... 162<br />

7.7 Screening tests in simulated cement kiln flue gas with HgCl2 source............ 166<br />

7.8 Conclusions..................................................................................................... 170<br />

7.9 References....................................................................................................... 172<br />

8. Fundamental investigation of elemental mercury adsorption <strong>by</strong> activated carbon in<br />

simulated cement kiln flue gas.................................................................................. 176<br />

8.1 Introduction..................................................................................................... 176<br />

8.2 Effect of adsorption temperature .................................................................... 177<br />

8.3 Isotherm tests .................................................................................................. 179<br />

8.4 Effect of carbon particle size .......................................................................... 183<br />

8.5 Effect of flue gas flow rate ............................................................................. 185<br />

8.6 Effects of flue gas compositions..................................................................... 186<br />

8.6.1 Effect of CO2 ........................................................................................... 186<br />

8.6.2 Effect of O2 .............................................................................................. 188<br />

8.6.3 Effect of H2O ........................................................................................... 189<br />

8.6.4 Effect of CO............................................................................................. 192<br />

8.6.5 Effect of SO2............................................................................................ 193<br />

8.6.6 Effect of HCl............................................................................................ 195<br />

VIII


8.6.7 Effect of NO............................................................................................. 197<br />

8.6.8 Effect of NO2 ........................................................................................... 198<br />

8.7 Conclusions..................................................................................................... 201<br />

8.8 References....................................................................................................... 202<br />

9. Fundamental investigation of mercury chloride adsorption <strong>by</strong> activated carbon in<br />

simulated cement kiln flue gas.................................................................................. 206<br />

9.1 Introduction..................................................................................................... 206<br />

9.2 Effect of temperature ...................................................................................... 207<br />

9.3 Effect of flue gas composition ........................................................................ 210<br />

9.4 Conclusions..................................................................................................... 212<br />

9.5 References....................................................................................................... 213<br />

10. Simulation of mercury adsorption <strong>by</strong> fixed carbon bed ..................................... 215<br />

10.1 Adsorption equilibrium................................................................................. 215<br />

10.2 Transport consideration in adsorption process ............................................. 216<br />

10.2.1 External transport................................................................................... 216<br />

10.2.2 Internal transport.................................................................................... 218<br />

10.3 Modeling of adsorption in a single particle .................................................. 220<br />

10.4 Fixed bed adsorption model.......................................................................... 226<br />

10.5 Conclusions................................................................................................... 239<br />

10.6 List of symbols.............................................................................................. 239<br />

10.7 References..................................................................................................... 241<br />

11. Simulation of mercury removal <strong>by</strong> activated carbon injection upstream of a fabric<br />

filter........................................................................................................................... 243<br />

11.1 Common assumptions for mercury removal in the duct and fabric filter..... 243<br />

11.2 Duct model.................................................................................................... 246<br />

11.3 Model for the filter cake ............................................................................... 253<br />

11.4 Fabric filter model ........................................................................................ 257<br />

11.5 Two-stage model........................................................................................... 263<br />

11.6 Conclusions................................................................................................... 269<br />

11.7 List of symbols.............................................................................................. 270<br />

11.8 References..................................................................................................... 271<br />

12. Concluding remarks............................................................................................ 273<br />

13. Suggestions for further work .............................................................................. 277<br />

IX


1.1 Project background<br />

1<br />

1<br />

Introduction<br />

There are growing concerns over mercury emissions due to its toxicity, volatility,<br />

persistence, and bioaccumulation in the environment. According to an inventory of<br />

global mercury emissions to the atmosphere <strong>from</strong> anthropogenic sources <strong>by</strong> Pacyna et al.<br />

[1], the largest emissions of mercury are <strong>from</strong> combustion of fossil fuels. <strong>Mercury</strong><br />

emissions <strong>from</strong> cement and mineral production are the second largest anthropogenic<br />

sources.<br />

While mercury emissions <strong>from</strong> waste incinerators and power plants have been<br />

and continue to be regulated <strong>by</strong> the authorities in many countries, strict mercury emission<br />

limits for cement plants are also established <strong>by</strong> different countries [2-6]. U.S.<br />

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

emissions <strong>from</strong> existing cement kilns and strengthened the limits for new kilns [7-9]. The<br />

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

tons of clinker, respectively. These emission limits correspond to 10 and 4 µg/Nm 3 .<br />

When fully implemented in 2013, EPA estimates the annual mercury emissions will be<br />

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

mercury emission limit without some changes to the system, either through operational<br />

adjustment or use of add-on technology.<br />

<strong>Mercury</strong> is present in both cement raw materials used for kiln feed and fuels used<br />

in the cement production process. Due to rising energy costs and ever stricter energy and<br />

environmental regulations, alternative fuel technology is becoming an important factor in<br />

controlling costs. To gain a competitive edge, many cement and mineral producers<br />

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


- both waste-derived fuel and biomass. High mercury containing alternative fuels such as<br />

chemical waste, domestic waste and sewage sludge are also incinerated in cement plants<br />

and high mercury emission problems have been encountered. To ensure that the mercury<br />

emission limit is met, FLSmidth has initiated research on mercury removal <strong>from</strong> cement<br />

plants.<br />

Due to the extremely low concentration range of mercury in the flue gas, mercury<br />

emission control techniques are technically challenging and expensive. Currently,<br />

activated carbon injection upstream of a particulate control device such as fabric filter<br />

has been shown to have the best potential to remove both elemental and oxidized<br />

mercury <strong>from</strong> the flue gas for combustion facilities not equipped with a wet flue gas<br />

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

flue gas desulphurization unit is installed. In cement plant application sorbent will be<br />

injected upstream of a polishing filter instead of an existing filter in order to separate<br />

carbon <strong>from</strong> the cement materials and save the disposal cost of sorbent and cement<br />

materials mixture.<br />

Although activated carbon is the most studied sorbent for capturing mercury <strong>from</strong><br />

power plant flue gas, mercury adsorption <strong>by</strong> activated carbon is not clearly understood<br />

yet, and research and development efforts are still needed before carbon injection may be<br />

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

developed, the sorbent costs need to be reduced and the amount of carbon injected needs<br />

to be kept to a certain level to minimize the cost. Furthermore, mercury adsorption<br />

stability <strong>by</strong> sorbents needs to be proved.<br />

Extensive research has been carried out to reduce mercury emissions <strong>from</strong> coal<br />

combustion and waste incineration, but very little efforts have been concentrated on<br />

mercury removal in cement plants. The mercury removal not only depends on the sorbent<br />

but also on the speciation of mercury, flue gas composition and temperature, and the<br />

system configuration. The mercury emissions and gas stream characteristics <strong>from</strong> coal<br />

combustion and waste incineration are quite different <strong>from</strong> those <strong>from</strong> cement kilns [4].<br />

Thus knowledge obtained <strong>from</strong> mercury removal in power plants and incinerators might<br />

2


not be applied to cement plant directly. Non-carbon based cement-friendly sorbent is<br />

desired so that the mercury containing sorbent can be used in cement production instead<br />

of costly disposal.<br />

Despite the considerable experimental research that has been carried out to date,<br />

few models for mercury adsorption <strong>by</strong> activated carbon injection in power plant or<br />

incinerator flue gas have been proposed. A comprehensive model is desired to estimate<br />

appropriate design and operating strategies that would lead to efficient and economic<br />

control of mercury.<br />

1.2 Project objectives<br />

The overall goal of this project is to develop and advance improved mercury control<br />

technologies using sorbent injection upstream of a pulse jet fabric filter for cement plant.<br />

Specific objectives are as follows:<br />

1. To obtain updated knowledge of mercury control technologies relevant to cement<br />

plant <strong>by</strong> comprehensive literature review.<br />

2. To develop an experimental lab setup and screen sorbents for capturing mercury<br />

<strong>from</strong> cement kiln flue gas.<br />

3. To test and develop thermal catalytic converters for oxidized mercury reduction and<br />

total mercury measurement.<br />

4. To develop an understanding of sorbent chemistry and provide mechanistic<br />

understanding and kinetic rates for sorbents of interest.<br />

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

predict mercury removal efficiency in cement plant <strong>by</strong> injecting sorbent upstream of a<br />

fabric filter.<br />

1.3 Outline of the thesis<br />

The thesis starts with a chapter (Chapter 2) on introduction of cement production<br />

process and mercury emission and transformation in cement kiln systems. Then in<br />

Chapter 3 available knowledge on mercury removal technologies <strong>from</strong> flue gas is<br />

3


eviewed and the applicability of the reviewed technologies in cement kilns is analyzed.<br />

Properties and performance of typical sorbents are also presented.<br />

Experimental methods and materials are presented in Chapter 4. Chapter 5<br />

particularly deals with the test of a red-brass based converter and development of a<br />

sulfite-based oxidized mercury reduction unit for total gaseous mercury measurement.<br />

Effects of bed dilution and carbon load on equilibrium mercury adsorption capacity of<br />

the activated carbon are investigated in chapter 6. Screening tests of different sorbent<br />

materials in the fixed-bed reactor under simulated cement kiln flue gas are reported in<br />

Chapter 7. Chapter 8 deals with fundamental investigation of mercury adsorption <strong>by</strong><br />

activated carbon in simulated cement kiln flue gas using elemental mercury source.<br />

<strong>Mercury</strong> adsorption mechanism and kinetics <strong>by</strong> the activated carbon will be reported.<br />

The fundamental investigation of mercury chloride adsorption <strong>by</strong> the activated carbon in<br />

simulated cement kiln flue gas will be reported in Chapter 9.<br />

Chapters 10 and 11 will deal with simulations of mercury adsorption <strong>by</strong> the<br />

activated carbon. Chapter 10 focuses on simulation of mercury adsorption <strong>by</strong> a single<br />

carbon particle and a fixed carbon bed. Simulation of mercury adsorption <strong>by</strong> activated<br />

carbon injection upstream of a fabric filter is the topic of Chapter 11. Validation of the<br />

developed duct-fabric filter two-stage model <strong>by</strong> available pilot-scale data is reported.<br />

Finally, conclusions <strong>from</strong> the project are presented in Chapter 12. Suggestions for<br />

further work are given in Chapter 13.<br />

1.4 References<br />

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

emission inventory for 2000, Atmospheric Environment. 40 (2006) 4048-4063.<br />

[2] The European Parliament and the Council of the European Union, Union directive<br />

2000/76/EC on the incineration of waste, 2000.<br />

[3] J. Werther, Gaseous emissions <strong>from</strong> waste combustion, Journal of Hazardous Materials. 144<br />

(2007) 604-613.<br />

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

best available techniques in the cement and lime manufacturing industries, Part I: Lime<br />

manufacturing industries, 2006.<br />

4


[5] German <strong>Cement</strong> Works Association, Environmental protection in cement manufacture, VDZ<br />

activity report 2003-2005.<br />

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

emissions, 2000.<br />

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

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

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

source performance standards for Portland cement manufacturing, 2010.<br />

[9] U.S. EPA, National emission standards for hazardous air pollutants <strong>from</strong> the Portland cement<br />

manufacturing industry and standards of performance for Portland cement plant, 40 CFR Parts 60<br />

and 63, EPA-HQ-OAR-2007-0877, FRLRIN 2060-AO42; EPA-HQ-OAR-2002-0051, FRLRIN<br />

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

accessed January/17, 2011.<br />

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

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

Technology. 82 (2003) 89-165.<br />

5


<strong>Mercury</strong> emissions and transformations in cement<br />

6<br />

2<br />

plants<br />

Knowledge of mercury emissions, speciation, and transformation in cement<br />

plants is important for understanding the transport and fate of mercury released to air<br />

pollution control systems. In this chapter cement production processes are first<br />

introduced and compared with power plants and waste incinerators regarding the flue gas<br />

composition, temperature, residence time, and inherent material circulation. Then<br />

mercury contents in fuels and raw materials applied in cement production and mercury<br />

emission <strong>from</strong> Portland cement plants are presented. Finally mercury transformations in<br />

combustion flue gas and cement kiln system are reviewed.<br />

2.1 <strong>Cement</strong> production processes<br />

Although cement production also involves combustion, the flue gas temperature<br />

and residence time in cement kilns are quite different <strong>from</strong> power plants and waste<br />

incinerators. To help understand the mercury chemistry in the cement kiln systems, a<br />

brief description of the cement production process is necessary. Differences regarding the<br />

gas temperature, residence time, flue gas composition, and material cycles among cement<br />

kilns, power plants and waste incinerators are discussed below.<br />

Depending on how the raw material is handled before being fed to the rotary kiln,<br />

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

Presently, about 78% of Europe's cement production is <strong>from</strong> dry process kilns [1], about<br />

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

production is <strong>from</strong> wet process kilns. Today, all new plants are based on the dry process<br />

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


In the dry process the feed material enters the kiln in a dry, powdered form.<br />

Production of cement can be subdivided into the areas of supply of raw materials,<br />

burning of cement clinker in the rotary kiln, and final cement production <strong>by</strong> adding<br />

interground additives [2].<br />

Raw materials for the manufacture of Portland cement clinker consist basically of<br />

limestone and aluminosilicates. At times, certain corrective materials such as bauxite,<br />

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

composition. Apart <strong>from</strong> natural raw materials, waste materials containing lime,<br />

aluminate, silicate, and iron are also used as raw materials substitutes.<br />

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

raw materials is milled in a raw mill and dried <strong>by</strong> the hot kiln flue gas. In a downstream<br />

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

subsequently transported to raw meal silos. The raw meal is fed into the kiln system,<br />

which is comprised of a tower of cyclone preheaters. The calcination process can almost<br />

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

precalciner, which is located between the kiln and the preheater.<br />

7


Figure 2.1. Sketch of a dry cement production process [2].<br />

In the burning of cement clinker it is necessary to maintain material temperatures<br />

of up to 1450°C to ensure the sintering reactions required [1]. This is achieved <strong>by</strong><br />

applying peak combustion temperatures of about 2000°C with the main burner flame.<br />

Figure 2.2 shows the gas temperature in the kiln system as a function of residence time<br />

and comparison with the gas temperature profiles in a pulverized coal-fired boiler and<br />

waste incinerator [1,3,4]. The combustion gases <strong>from</strong> the main kiln burner remain at<br />

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

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

be burned under oxidizing conditions. The residence time of the solid materials in the<br />

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

flue gas flows through the rotary kiln and preheater in opposite direction to the solids.<br />

The burning conditions in kilns with precalciner firing depend on the precalciner design.<br />

8


Gas temperatures <strong>from</strong> a precalciner burner are typically around 1100°C, and the gas<br />

residence time in the precalciner is approximately 3 seconds. In the cyclone preheater<br />

zone, the gas temperatures typically range <strong>from</strong> approximately 880-890°C at the inlet of<br />

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

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

and the air pollution control device, with gas temperature typically in the range <strong>from</strong><br />

approximately 350-90°C <strong>from</strong> the top of the preheater to the exit stack outlet.<br />

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

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

Generally the gas temperature and residence time in a kiln system is much higher<br />

and longer than those in a pulverized coal-fired boiler and waste incinerator. The<br />

temperature profile in the waste incinerator shown in Figure 2.2 is in the region <strong>from</strong> the<br />

furnace exit to the boiler exit [4] and the gas temperature is much lower than those <strong>from</strong><br />

cement kiln and pulverized coal-fired boiler.<br />

The clinker leaving the rotary kiln is cooled down <strong>by</strong> grate or planetary coolers.<br />

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

cement, which is the most common type of cement. In addition, blended cements are<br />

9


produced <strong>by</strong> intergrinding cement clinker with materials like fly ash, granulated blast<br />

furnace slag, limestone, natural or artificial pozzolanas [5].<br />

Table 2.1 compares the flue gas compositions among coal-fired power plant,<br />

waste incinerator and cement kiln. The major difference between cement kiln flue gas<br />

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

content in the kiln gas is lower than in coal combustion and waste incineration flue gas.<br />

The emission of HCl <strong>from</strong> cement kilns is normally much lower than those <strong>from</strong> waste<br />

incinerators. This could be due to the fact that the environment in cement plants is<br />

effective for absorbing acid gasses [6], such as a range of gas temperatures <strong>from</strong> 100 to<br />

1650°C, gas residence time of about 30s, high levels of turbulence, high concentrations<br />

of alkaline solids including sodium and potassium oxides, and freshly created CaO in<br />

high concentrations. Therefore, gaseous species such as HCl or HF are nearly completely<br />

captured <strong>by</strong> the inherent and efficient alkaline sorption effect of the cement kiln system<br />

[1].<br />

Table 2.1 Typical flue gas compositions in coal-fired boiler, waste incinerator, and<br />

cement kiln before air pollution control device (APCD).<br />

Pulverized coalfired<br />

boiler [1,7-<br />

10,10,11]<br />

10<br />

Waste<br />

incinerator [7,8]<br />

<strong>Cement</strong> kiln<br />

[7-9,12]<br />

O2 (vol.%) 4-6 6-15 2-4<br />

CO2 (vol.%) 10-16 5-14 14-33<br />

H2O (vol.%) 5-12 10-18 5-35<br />

CO (ppmv) 10-100 10-100 600-2600<br />

NO (ppmv) 100-1000 100-1000 475-1900<br />

NO2 (ppmv) 5-50 5-50 25-100<br />

N2O (ppmv)


<strong>Cement</strong> kilns also differ <strong>from</strong> conventional boilers and incinerators in having the<br />

dust recycles in the kiln systems. There are two material cycles in the cement kiln system,<br />

i.e., the internal and external cycle. Because of the countercurrent flow of combustion<br />

products and solids in cement kilns, volatile elements such as mercury, alkalis, sulphur<br />

and chlorine evaporated <strong>from</strong> the solids at the hot end of the kiln near the combustion<br />

zone are carried to the cold end <strong>by</strong> the combustion gases. Some of the volatile<br />

compounds pass through the entire system and exit in vapor phase through the stack.<br />

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

particles and surrounding walls in the cooler regions of the kiln system. With the raw<br />

meal, they are reintroduced to the hot zone thus establishing the internal cycle of volatile<br />

elements.<br />

The external cycle comprises the mass flows that include the raw mill and dust<br />

collectors downstream of the preheater. A small part of the circulating elements leaves<br />

the kiln with the exhaust gas dust and is precipitated in the dedusting device of the<br />

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

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

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

the solid flow in cement plant with a precalciner [13].<br />

With excessive input of volatile elements, the installation of a kiln gas <strong>by</strong>pass<br />

system may become necessary in order to extract part of the circulating elements <strong>from</strong><br />

the kiln system. This <strong>by</strong>pass dust, which is usually highly enriched in alkalis, sulphur or<br />

chloride, is cooled down and then passed through a dust collector before being<br />

discharged.<br />

The operation modes of the cement plants are important for understanding<br />

mercury transformations in the kiln systems as presented in section 2.3. There are two<br />

operation modes [2], i.e., compound operation (raw-mill-on) and direct operation (rawmill-off),<br />

as shown in figure 2.3. Usually these modes are run alternately. The raw mill<br />

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

11


the dust-containing off-gas <strong>from</strong> the cyclone preheater is used for drying and transporting<br />

the raw meal <strong>from</strong> the raw mill. Water injection in the cooler is not applied to cool down<br />

the gas. The raw meal and fly dust <strong>from</strong> the kiln system are collected <strong>by</strong> the ESP or FF<br />

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

dust-containing off-gas <strong>from</strong> the kiln is cooled down in the off-gas cooler <strong>by</strong> the injection<br />

of water and subjected to subsequent dedusting in the ESP or FF.<br />

Figure 2.3. Operation models in cement production [2].<br />

These different modes of operation considerably influence the temperatures and<br />

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

the trace element mass flows in the plant. Increased off-gas temperature during direct<br />

operation causes higher mercury emission level than in the compound mode [2].<br />

Moreover, regular alternation of the operation modes results in weekly cycles of mercury<br />

flows in the cement plant, as discussed in section 2.2.<br />

2.2 <strong>Mercury</strong> contents in fuels and cement raw materials<br />

A comprehensive analysis of mercury content in 291 raw material samples <strong>from</strong><br />

57 cement plants in Canada and U.S. was conducted <strong>by</strong> Hills and Stevenson [15]. Table<br />

2.2 shows the mercury contents in the fuels and raw materials applied in cement<br />

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

materials. The reported average mercury content in the raw materials except for fly ash<br />

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

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

12


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

mercury content and application of fly ash in cement production results in increased<br />

mercury input to the cement kiln and potentially higher mercury emissions. Process<br />

changes in cement plants such as substitution with alternative fuels may result in more<br />

plants needing solutions for mercury emission control.<br />

Table 2.2. <strong>Mercury</strong> contents in raw materials and fuels for cement production. All on dry<br />

weight basis.<br />

Material/fuel Category Sample Average Minimum Maximum<br />

number (ppm) (ppm) (ppm)<br />

Limestone [15] Primary 90 0.017


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

mercury is <strong>from</strong> the kiln feed rather than the fuels when considering the amount of fuels<br />

and raw materials used [22].<br />

2.3 <strong>Mercury</strong> emissions<br />

The U.S. Portland cement association summarized 50 mercury emission tests in<br />

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

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

emission data are only for plants not burning hazardous waste. The information on<br />

mercury speciation is not available. The mercury emission concentrations varied <strong>from</strong><br />

0.02 μg/Nm 3 to 385.6 μg/Nm 3 with a mean value of 28.0 μg/Nm 3 @dry, 7% O2 and a<br />

standard deviation of 62.7 μg/Nm 3 . The maximum mercury concentration was three<br />

times higher than the second highest value.<br />

The U.S. Portland cement association has later gathered and analyzed mercury<br />

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

mercury emissions include particle-bound mercury (Hgp), elemental mercury (Hg 0 ), and<br />

oxidized mercury (Hg 2+ ). The mercury emissions and speciation <strong>from</strong> cement kilns can<br />

vary over time and depend on raw materials and fuels used, and process operation. The<br />

average mercury speciation percentages for cement plants with preheater or precalciner<br />

not firing waste are 5% Hgp, 56% Hg 2+ , 39% Hg 0 for raw-mill-on mode [24], and 4%<br />

Hgp, 62% Hg 2+ , 34% Hg 0 during raw-mill-off mode.<br />

Large variations of mercury speciation during raw-mill-on and -off modes have<br />

been observed in some plants with higher mercury emission during the raw-mill-off<br />

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

mercury concentration during raw-mill-on and raw-mill-off period was 410 and 2250<br />

μg/Nm 3 , respectively [25]. The larger mercury emission during raw-mill-off period is<br />

probably due to high flue gas temperature and lack of mercury adsorption <strong>by</strong> cement raw<br />

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

14


volunteered to install a sorbent injection process for removing at least 75% of the<br />

mercury [26].<br />

The complex mercury mitigation cycles within the cement kiln system make it<br />

difficult to obtain an equilibrium state due to the periodical shut down of raw mills for<br />

maintenance. It typically takes weeks to reach long term equilibrium of the mercury<br />

emission [27].<br />

The German cement manufacturing association has reported mercury emission<br />

results <strong>from</strong> 216 measurements on 44 kilns [28]. Twenty of the results were below the<br />

detection limit. Most of the measurements were below 40 μg/Nm 3 . Only six of the results<br />

were 60 μg/Nm 3 or higher.<br />

The emitted elemental mercury <strong>from</strong> Powder River Basin (PRB) coal-fired power<br />

plants ranges <strong>from</strong> approximately 10 to100 μg/Nm 3 [29]. <strong>Mercury</strong> concentrations in the<br />

flue gas <strong>from</strong> municipal solid waste combustion (200 to 1000 μg/Nm 3 ) are one to two<br />

orders of magnitude higher than for coal combustion sources (5 to 20 μg/Nm 3 ) [30,31].<br />

<strong>Mercury</strong> levels in cement kiln flue gas are generally closer to those found in coal-fired<br />

boilers and lower than those found in waste incinerators.<br />

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

atmosphere <strong>from</strong> anthropogenic sources for the year 2000. The largest emissions of<br />

mercury to the global atmosphere are <strong>from</strong> combustion of fossil fuels, mainly coal in<br />

utility, industrial, and residential boilers. Emissions of mercury <strong>from</strong> coal combustion are<br />

between one and two orders of magnitude higher than emissions <strong>from</strong> oil combustion.<br />

Various industrial processes account for additional 30% of mercury emissions <strong>from</strong><br />

anthropogenic sources worldwide in 2000. <strong>Mercury</strong> emissions <strong>from</strong> cement and mineral<br />

production are the second largest anthropogenic sources.<br />

2.3 <strong>Mercury</strong> transformation during combustion<br />

Knowledge of mercury transformations in combustion flue gas is important for<br />

selection of the mercury control technology and understanding the fate and behavior of<br />

mercury <strong>from</strong> combustion processes. Major chemical forms of mercury <strong>from</strong> combustion<br />

15


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

mercury, which is the portion of mercury deposited on fine particles. Oxidized mercury<br />

species, such as HgCl2 and HgO, are easily removed <strong>by</strong> existing wet type air pollution<br />

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

Also particulate mercury is readily removed <strong>by</strong> the main dust removal control devices<br />

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

because of its high vapor pressure and insolubility in water.<br />

Table 2.3 presents properties of selected mercury compounds. Metallic mercury is<br />

a heavy, silvery-white liquid metal at typical ambient temperatures and pressures, and it<br />

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

numerous inorganic and organic chemical compounds, but the mercurous mercury is<br />

rarely stable under ordinary environmental conditions [23]. The solubility of the mercury<br />

compounds varies greatly <strong>from</strong> negligible (Hg2Cl2, HgS) to very soluble (HgCl2).<br />

Mercuric sulfate reacts with water to produce yellow insoluble basic mercuric subsulfate<br />

and sulfuric acid.<br />

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

Hg 0 Elemental<br />

mercury<br />

Hg2Cl2<br />

HgCl2<br />

Hg2SO4<br />

HgSO4<br />

Name Molar<br />

weight<br />

(g/mol)<br />

Mercurous<br />

chloride<br />

Mercuric<br />

chloride<br />

Mercurous<br />

sulphate<br />

Mercuric<br />

sulphate<br />

Melting<br />

point<br />

(�C)<br />

Boiling<br />

point<br />

(�C)<br />

16<br />

Decomposition<br />

/sublimate<br />

temperature<br />

(�C)<br />

Density<br />

(g/cm 3 )<br />

Aqueous<br />

solubility<br />

(g/l at 25�C)<br />

200.59 -38.8 356.7 n.a. 13.53 5.6�10 -7<br />

472.09 525 n.a. 383 7.15 0.002<br />

271.50 277 302 n.a. 5.43 28.6<br />

497.24 n.a. n.a. n.a. 7.56 0.51<br />

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


HgS <strong>Mercury</strong><br />

sulfide<br />

HgO Mercuric<br />

oxide<br />

Hg2Br2<br />

HgBr2<br />

Hg2I2<br />

HgI2<br />

Hg2F2<br />

Name Molar<br />

weight<br />

(g/mol)<br />

Mercurous<br />

bromide<br />

Mercuric<br />

bromide<br />

Mercurous<br />

iodide<br />

Mercuric<br />

iodide<br />

Mercurous<br />

fluoride<br />

HgF2 Mercuric<br />

fluoride<br />

Hg2(NO3)2 Mercurous<br />

nitrate<br />

Hg(NO3)2 Mercuric<br />

nitrate<br />

Melting<br />

point<br />

(�C)<br />

Boiling<br />

point<br />

(�C)<br />

232.66 n.a. 446-<br />

583<br />

17<br />

Decomposition<br />

/sublimate<br />

temperature<br />

(�C)<br />

Density<br />

(g/cm 3 )<br />

Aqueous<br />

solubility<br />

(g/l at 25�C)<br />

580 8.10 insoluble<br />

216.59 n.a. 356 500 11.14 insoluble<br />

560.99 405 n.a. 340-350 7.31 3.9�10 -4<br />

360.44 237 322 n.a. 6.03 slightly<br />

soluble<br />

654.98 n.a. n.a. 140 7.70 Slightly<br />

soluble<br />

454.40 259 350 n.a. 6.36 0.06<br />

439.18 n.a. n.a. 570 8.73 decomposes<br />

238.59 645 650 645 8.95 soluble,<br />

reacts<br />

525.19 n.a. n.a. 70 (dihydrate) 4.80<br />

(dihydrate)<br />

slightly<br />

soluble, reacts<br />

324.7 79 n.a. n.a. 4.3 0 soluble<br />

2.3.1 <strong>Mercury</strong> transformation in coal combustion flue gas<br />

Figure 2.4 illustrates the potential mercury transformation paths during coal<br />

combustion [33]. All forms of mercury in the coal decompose in the combustion flame to<br />

form Hg 0 (g) [30,33]. In the post combustion section where the gas temperature decreases,<br />

Hg 0 (g) may remain as a monatomic species or react to form inorganic mercurous and<br />

mercuric compounds. The principal oxidized forms of mercury in coal combustion flue<br />

gas are assumed to be Hg 2+ compounds. Oxidation of mercury via halogenation does not<br />

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

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


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

onto inorganic and carbonaceous ash particles entrained in the flue gas. In addition to<br />

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

[30,33].<br />

Figure 2.4. Potential mercury transformation during coal combustion and subsequently in<br />

the resulting flue gas, modified after [33].<br />

Many parameters can potentially affect the formation of various mercury species<br />

throughout a combustion system [30], including fuel type and composition, combustion<br />

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

convective cooling, configuration of APCD, and operating practices.<br />

As a starting point, the distribution of mercury species in coal combustion flue<br />

gas can be calculated using thermodynamic equilibrium calculations. Senior et al. [3]<br />

calculated the equilibrium mercury speciation in the flue gas <strong>from</strong> Pittsburgh bituminous<br />

coal combustion. Typical results <strong>from</strong> 227 to 827�C are shown in figure 2.5. At<br />

temperatures below 150�C condensed HgSO4 is the only preferred specie (not shown in<br />

figure 2.5). Similar observations were also observed <strong>by</strong> Frandsen et al. [37]. As<br />

illustrated in figure 2.5, below 450�C all of the mercury is predicted to exist as HgCl2.<br />

Above about 700�C 99% of mercury is predicted to exist as gaseous elemental mercury.<br />

18


The remaining 1% is predicted to be gaseous HgO. Between 450 and 700�C the split<br />

between HgCl2 and elemental mercury is determined <strong>by</strong> the chlorine content of the coal.<br />

%Hg<br />

100<br />

80<br />

60<br />

40<br />

20<br />

HgCl 2(g)<br />

HgO(g)<br />

0<br />

200 300 400 500 600 700 800 900<br />

Temperature (oC)<br />

19<br />

Hg(g)<br />

Figure 2.5. Equilibrium distribution of mercury species in flue gas <strong>from</strong> combustion of<br />

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

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

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

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

The effect of HCl concentration on equilibrium partitioning between elemental<br />

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

the elemental and oxidized forms increases <strong>from</strong> 530 to 740�C as the HCl concentration<br />

increases <strong>from</strong> 50 to 3000 ppm. The studied HCl level is much higher than real level in<br />

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

point is not influenced <strong>by</strong> the mercury concentration as long as hydrochloric acid is<br />

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

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

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


Figure 2.6. Equilibrium distribution of elemental mercury and mercury chloride for<br />

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

12.3% H2O and N2 as balance.<br />

The high levels of mercury oxidation are most strongly correlated with high<br />

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

subsequent capture of mercury [30]. Calcium likely reacts with chlorine and sulphur<br />

during the combustion process and there<strong>by</strong> reduces its ability to promote the oxidation of<br />

mercury [33]. The high percentages of elemental mercury typically found in emissions<br />

<strong>from</strong> lignite and subbituminous coal combustion can likely be attributed to their high<br />

calcium and low chlorine contents.<br />

Full-scale measurements showed that elemental mercury was dominant in the<br />

stack of coal-fired power plants, while oxidized mercury was dominant in the stack of<br />

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

furnaces and APCDs configuration differences between them. For the study of mercury<br />

removal <strong>by</strong> sorbent injection upstream of dust collectors, it is important to know the<br />

mercury speciation at the APCD inlet rather than at the stack. The data of mercury<br />

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

again due to different parameters that potentially affect the mercury speciation. Therefore,<br />

20


to develop a mercury control system for a specific plant, measurement of the mercury<br />

speciation at the APCDs’ inlet is necessary.<br />

There is disagreement in the publications on the relative importance of mercury<br />

halogenation in the flue gas <strong>by</strong> chlorine and bromine. Most literatures suggest that<br />

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

research <strong>by</strong> Vosteen et al. [44] shows that the critical species for the halogenation of<br />

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

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

diatomic and molecular form of the halogens become stable according to the Deacon<br />

type of reactions [33,44]:<br />

4HCl �O�2HO� 2Cl(Chlorine-Deacon-reaction)<br />

(R2.1)<br />

2 2 2<br />

4HBr �O�2HO� 2Br<br />

(Bromine-Deacon-reaction) (R2.2)<br />

2 2 2<br />

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

molecular chlorine is consumed during boiler passage <strong>by</strong> SO2 through the chlorine<br />

Griffin reaction:<br />

SO �Cl �HO�SO� HCl (Chlorine-Griffin-reaction) (R2.3)<br />

2 2 2 3 2<br />

In contrast to chlorine, the bromine-Griffin-reaction is not thermodynamically<br />

favored at temperatures above 100�C, because the Gibbs free reaction enthalpy of the<br />

bromine-Griffin-reaction is strongly positive within the whole boiler temperature range.<br />

Therefore, SO2 is not consuming Br2 during boiler passage. To summarize, the primary<br />

reason that bromine is a much more effective mercury oxidizer than chlorine is that HBr<br />

dissociates much more extensively into reactive atomic species than HCl at typical postflame<br />

conditions [45].<br />

The world average Cl contents in coals for bituminous and lignite coals are,<br />

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

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

the bromine content in the coal, the amount of molecular bromine Br2 in the flue gas may<br />

be many times higher than the amount of Cl2 in the flue gas downstream the combustion<br />

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

21


species is much faster than with chlorine species because the bromine atom<br />

concentrations at the furnace exit are three to four orders of magnitude greater. There<br />

might be ample supply of Br2 to oxidize the typical amounts of mercury in the coal flue<br />

gases through direct mercury bromination:<br />

Hg � Br � HgBr (Direct Hg bromination) (R2.4)<br />

2<br />

2<br />

Based on this knowledge, direct bromine injection into the flue gas has been proposed<br />

and patented to enhance mercury capture <strong>by</strong> fly ash or sorbents, or mercury oxidation<br />

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

higher concentration of Br2 in the post-combustion zone is not verified <strong>by</strong> full-scale<br />

investigation due to the lack of Br2 and Cl2 measurements.<br />

The arguments on the relative importance of mercury adsorption <strong>by</strong> bromine are<br />

supported <strong>by</strong> simulation and full-scale demonstration in power plants [45,50].<br />

Simulations with only homogeneous reaction mechanism <strong>by</strong> Niksa et al. [45] show that<br />

50% mercury oxidation is obtained for a typical thermal history along a power plant gas<br />

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

achieved <strong>by</strong> 20 ppmv HCl in the flue gas. Homogeneous mercury oxidation <strong>by</strong> bromine<br />

begins as the flue gas cools below 600�C and accelerates sharply when the temperature<br />

drops to below 300�C. At the furnace exit, bromine atoms are present in concentrations<br />

that are comparable to HBr levels, in contrast to the much lower concentrations of<br />

chlorine atoms at these conditions.<br />

Liu et al. [50] estimated that a 50% mercury oxidation could be obtained <strong>by</strong><br />

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

Laboratory study of Br2 in the simulated flue gas showed that fly ash in the flue gas<br />

significantly promoted the oxidation of Hg 0 <strong>by</strong> Br2 and the unburned carbon in the fly<br />

ash played a major role in the promotion primarily through the rapid adsorption of Br2<br />

[50]. Hg 0 oxidation in the gas phase was found to be less important than fly ash-induced<br />

oxidation <strong>by</strong> Br2. However, there is an increasing concern on the stability of bromine<br />

impregnated in the AC, added to fuels, or injected directly to the flue gas, which could<br />

22


lead to downstream pollution and pipeline corrosion due to the strong acidic nature of<br />

bromine.<br />

2.3.2 <strong>Mercury</strong> transformation within cement kiln system<br />

Larsen et al. [51] made a thermodynamic calculation of potential mercury species<br />

distribution in a cement kiln preheater. In order to get closer to a preheater environment,<br />

chloride as well as sulphide and sulphate compounds were included in the oxygencontaining<br />

system. Detailed compositions of the solid and flue gas can be found in the<br />

figure caption. The alkaline dust was represented <strong>by</strong> CaO in the calculations, which is in<br />

excess compared to the acidic components such as HCl and SO2. Figure 2.7 illustrates the<br />

equilibrium distribution of mercury species as a function of temperature when the<br />

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

oxidized mercury in forms of HgO and HgCl2, while all mercury compounds<br />

thermodynamically preferred above 200°C are gas-phase species and the main species is<br />

Hg 0 (g).<br />

23


Figure 2.7. Equilibrium distribution of mercury species as a function of temperature in<br />

the preheater environment with mercury input in the range of ppmm [51].<br />

Thermodynamic calculation input: solid: 5.00 kmol CaO, 0.000025 kmol HgCl2,<br />

0.000025 kmol HgSO4, 0.000025 kmol HgS, 0.000025 kmol Hg, gas: 0.03 kmol HCl, 1<br />

kmol H2O,1 kmol O2, 30.00 kmol CO2, 0.05 kmol SO2, 67.95 kmol N2.<br />

Figure 2.8 illustrates the equilibrium distribution of mercury species as a function<br />

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

included in the calculation assuming that HCl can be captured <strong>by</strong> large amount CaO in<br />

the cement raw materials. The results are completely different <strong>from</strong> the calculation with<br />

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

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

at around 450°C, thus the dominant species above 450°C are Hg 0 (g) and HgCl2(g).<br />

24


Fig. 2.8. Equilibrium distribution of mercury species as a function of temperature in the<br />

preheater environment with mercury input in the range of ‰ [51]. Thermodynamic<br />

calculation input: solid: 0.025 kmol HgCl2, 0.025 kmol HgSO4, 0.025 kmol HgS, 0.025<br />

kmol Hg, gas: 1 kmol H2O,1 kmol O2,30.00 kmol CO2, 0.05 kmol SO2, 97.95 kmol N2.<br />

General conclusions <strong>from</strong> thermodynamic calculations for a preheater<br />

representative environment are [51]: HgS will most probably be converted to other<br />

mercury species when entering the preheater, provided the reaction rates are sufficiently<br />

high compared to residence time. <strong>Mercury</strong> species are preferentially gas-phase<br />

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

thermodynamically preferred mercury species above 300°C is Hg 0 (g). This may be<br />

primarily because CaO acts as an HCl drain. Calculation indicates that the<br />

thermodynamically favored mercury species present at the extraction point for a typical<br />

kiln <strong>by</strong>-pass is Hg 0 (g).<br />

Detailed experimental information of mercury transformation in cement kiln<br />

system has not been reported. Although cement production also involves combustion, the<br />

25


flue gas composition, temperature and residence time in cement kiln are quite different<br />

<strong>from</strong> power plants and waste incinerators as explained earlier. When looking at mercury<br />

chemistry in cement kilns, these factors should be taken into consideration.<br />

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

cement kiln systems using material balance studies and comprehensive stack tests that<br />

were conducted over the past two decades. They concluded that mercury does not simply<br />

volatilize out <strong>from</strong> combusted fuels and heated kiln feed materials and leave directly out<br />

of the stack. The cement kiln systems have some inherent ability to control mercury stack<br />

emissions.<br />

Besides adsorption of mercury on the raw material, as shown earlier, new<br />

mercury compounds such as mercury silicates might be formed through reaction of<br />

mercury with silicate in the raw material and exit the system with the clinker product.<br />

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

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

to take place as vaporized mercury cycles through a kiln system. Edgarbaileyite is the<br />

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

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

the mineral data of Edgarbaileyite is available, it has not been possible to identify the<br />

thermodynamic properties of the mineral. A chemical equilibrium study was conducted<br />

to estimate probable conditions for the formation of mercury silicates in high temperature<br />

systems [54]. Results <strong>from</strong> the study suggest that HgSiO3 may form over a temperature<br />

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

silicate formation may be inhibited <strong>by</strong> the presence of chlorine and sulfur. It is reported<br />

<strong>by</strong> the European cement association that volatile metals are retained in the clinker to a<br />

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

confirm that mercury silicates are stable above temperatures of 325°C. Fundamental<br />

research is required to identify formation of mercury silicates in the cement kiln systems.<br />

26


2.4 Conclusions<br />

<strong>Cement</strong> plants are quite different <strong>from</strong> power plants and waste incinerators<br />

regarding the flue gas composition, temperature, residence time, and inherent material<br />

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

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

larger water and CO2 contents in the cement kiln flue gas.<br />

In cement production the raw materials contain mercury – often at much higher<br />

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

and raw materials. The mercury concentrations in the flue gas <strong>from</strong> cement kilns are<br />

typically in the range of 1-50 μg/m 3 . Instead of fuel, cement raw materials are the<br />

dominant sources of mercury in the cement kiln flue gas. Higher mercury emissions,<br />

however, are observed for cement plants firing waste.<br />

The mercury emissions and speciation <strong>from</strong> cement kilns can vary over time and<br />

depend on raw materials and fuels used, and process operation. The average mercury<br />

speciation percentages for cement plants with preheater or precalciner not firing waste<br />

are 5% Hgp, 56% Hg 2+ , 39% Hg 0 for raw-mill-on mode, and 4% Hgp, 62% Hg 2+ , 34%<br />

Hg 0 during raw-mill-off mode.<br />

<strong>Mercury</strong> transformations in combustion flue gas have been investigated<br />

intensively to get an understanding of the transport and fate of mercury into to air<br />

pollution control systems. All forms of mercury in the fuel decompose in the combustion<br />

flame to form Hg 0 (g), which is oxidized to Hg 2+ in the post combustion section. <strong>Mercury</strong><br />

halogenation <strong>by</strong> chlorine and bromine is the dominant mercury transformation<br />

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

gas or adsorb onto inorganic and carbonaceous ash particles entrained in the flue gas<br />

stream. Equilibrium calculations and experiments show that bromine is a much more<br />

effective mercury oxidizer than chlorine.<br />

The cement kiln systems have some inherent ability to retain mercury in the solid<br />

materials. The mercury evaporated <strong>from</strong> the solids at the hot end of the kiln is carried to<br />

the cold end <strong>by</strong> the combustion gases. As the flue gas cools, some mercury may<br />

27


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

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

to sweep the mill of the finely ground raw feed particles and some mercury is adsorbed<br />

<strong>by</strong> the fine particulates. However, the adsorbed mercury is either carried back to the kiln<br />

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

cycles in the kiln system.<br />

2.5 Further work<br />

There is limited literature regarding mercury characteristics, emissions, and<br />

removal <strong>from</strong> cement kilns. Essentially all of the published data and information apply to<br />

waste incinerators and coal-fired boilers, all of which have mercury emissions and gas<br />

stream characteristics that are quite different <strong>from</strong> those <strong>from</strong> cement kilns. Therefore,<br />

comprehensive studies on mercury chemistry in the cement kiln and mercury removal<br />

<strong>from</strong> cement plants are imperative.<br />

The inherent recycle of mercury in the kiln system should be further investigated.<br />

The interactions between mercury and cement raw materials play an important role in<br />

understanding of mercury chemistry in the cement kiln system. Research is required to<br />

break the mercury cycle in the kiln system, regenerate and implement beneficial<br />

utilization of removed mercury-contained CKD. These treatment systems minimize net<br />

CKD generation <strong>by</strong> removing mercury, alkalies and other contaminants and returning<br />

treated dust to the system without compromising product quality.<br />

2.6 References<br />

[1] CEMBUREAU, the European <strong>Cement</strong> association, Best available technologies for the cement<br />

industry, 1999.<br />

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

Stemmermann, Heavy metals in cement and concrete resulting <strong>from</strong> the co-incineration of wastes<br />

in cement kilns with regard to the legitimacy of waste utilisation, Wissenschaftliche Berichte,<br />

FZKA 6923, 2003.<br />

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

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

28


[4] D. Shin, S. Choi, J. Oh, Y Chang, Evaluation of polychlorinated dibenzo-pdioxin/dibenzofuran<br />

(PCDD/F) emission in municipal solid waste incinerators, Environ. Sci.<br />

Technol. 33 (1999) 2657-2666.<br />

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

(2008) 543-560.<br />

[6] M. V. Seebach and D. Gossman, <strong>Cement</strong> kilns sources of chlorides not HCl emissions,<br />

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

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

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

May 4-9 2003.<br />

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

Water Air and Soil Pollution. 56 (1991) 3-14.<br />

[9] Donaldson Membranes, Reducing emissions: Filtering chloride emissions with a <strong>by</strong>pass cycle,<br />

Filtr. Sep. 45 (2008) 36-37.<br />

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

Engineering Chemistry. 23 (1931) 620-624.<br />

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

Environ. Sci. Technol. 41 (2007) 6579-6584.<br />

[12] E. Worrell, L. Price, N. Martin, C. Hendriks, L.O. Meida, Carbon dioxide emissions <strong>from</strong><br />

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

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

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

[14] Department of Environmental Quality State of Oregon, Ash Grove mercury reduction,<br />

advisory committee’s report, 2007.<br />

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

Serial No. 2888, 2006.<br />

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

2887, 2006.<br />

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

Gips. 47 (1998) 183.<br />

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

Management. 25 (2005) 239-247.<br />

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

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

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

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

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

2006.<br />

29


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

cement kiln system, PCA R&D Serial No. 2841, 2005.<br />

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

PCA R&D Serial No. 2567, 2003.<br />

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

No. SN3091, 2009.<br />

[25] Schreiber & Yonley Associates, <strong>Mercury</strong> emissions test report, Ash Grove <strong>Cement</strong><br />

Company Durkee, Oregon, Project No. 060204, 2007.<br />

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

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

[27] Ravi Narayan, <strong>Mercury</strong> monitoring challenges facing the cement industry,<br />

http://www.cemtrex.com/component/content/article/5-monitoring/125-mercury-monitoringchallenges-facing-the-cement-industry.html,<br />

accessed July/22, 2010.<br />

[28] German <strong>Cement</strong> Works Association, Activity report 1999-2001, 2001.<br />

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

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

2006.<br />

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

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

Technology. 82 (2003) 89-165.<br />

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

Assessing sorbent injection mercury control effectiveness in flue gas streams, Environ. Prog. 19<br />

(2000) 167-174.<br />

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

emission inventory for 2000, Atmospheric Environment. 40 (2006) 4048-4063.<br />

[33] K.C. Galbreath, C.J. Zygarlicke, <strong>Mercury</strong> transformations in coal combustion flue gas, Fuel<br />

Processing Technology. 65-66 (2000) 289-310.<br />

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

combustion sources, Powder Technology. 180 (2008) 151-156.<br />

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

Ed. The McGraw-Hill Companies, Inc., 1997.<br />

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

<strong>Mercury</strong>_compounds, accessed September 8, 2010.<br />

[37] F. Frandsen, K. Dam-Johansen, P. Rasmussen, Trace elements <strong>from</strong> combustion and<br />

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

(1994) 115-138.<br />

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

model for the homogeneous oxidation of mercury <strong>by</strong> chlorine species, Fuel Process Technol. 65-<br />

66 (2000) 423-438.<br />

30


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

power plant mercury control <strong>by</strong> injecting sodium tetrasulfide, Institutes of Clean Air Companies<br />

(ICAC) Forum 2003, Nashville, TN, October 14-15, 2003.<br />

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

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

2224.<br />

[41] X. Yang, Y. Zhuo, Y. Duan, L. Chen, L. Yang, L. Zhang, Y. Jiang, X, Xu, <strong>Mercury</strong><br />

speciation and its emissions <strong>from</strong> a 220 MW pulverized coal-fired boiler power plant in flue gas,<br />

Korean Journal of Chemical Engineering. 24 (2007) 711-715.<br />

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

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

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

mercury <strong>from</strong> municipal solid waste incinerators in Taiwan, The Science of The Total<br />

Environment. 246 (2000) 165-173.<br />

[44] B.W. Vosteen, R. Kanefke, H. Köser, Bromine-enhanced mercury abatement <strong>from</strong><br />

combustion flue gases-Recent industrial applications and laboratory research, VGB PowerTech.<br />

86 (2006) 70.<br />

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

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

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

Geology. 67 (2006) 127-144.<br />

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

chlorine and bromine in coal, Fuel. 79 (2000) 903-921.<br />

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

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

[49] M. Holmes and J. Pavlish, <strong>Mercury</strong> information clearinghouse, Quarter 3- Advanced and<br />

developmental mercury control technologies, July 2004.<br />

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

enhance mercury removal <strong>from</strong> flue gas of coal-fired power plants, Environ. Sci. Technol. 41<br />

(2007) 1405-1412.<br />

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

in cement production-A literature review, FLSmidth internal report, 2007.<br />

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

first mercury silicate, American Mineralogist. 75 (1990) 1192.<br />

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

<strong>from</strong> California and Texas, The Mineralogical Record. 21 (1990) 215.<br />

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

with sorbents in high temperature systems, Chem. Eng. Commun. 133 (1995) 31-52.<br />

31


Review of technologies for mercury removal <strong>from</strong><br />

32<br />

3<br />

flue gas<br />

This chapter reviews the available technologies for mercury removal <strong>from</strong> flue<br />

gas, and the applicability of the technologies in cement plant is discussed. Focus is put on<br />

mercury removal <strong>by</strong> sorbent injection. Tests of sorbents in lab-scale fixed-bed reactors,<br />

slipstream pilot-scale reactors and full-scale plants are reported.<br />

3.1 Introduction<br />

The options for mercury control include mercury avoidance <strong>by</strong> coal and raw<br />

material cleaning, mercury removal <strong>by</strong> sorbent injection upstream of existing air<br />

pollution control devices (APCDs), and enhanced mercury removal <strong>by</strong> oxidation.<br />

Differences in fuel type and composition and pollution control devices make it necessary<br />

to develop customized solutions for each plant. The suitable mercury control method for<br />

a specific plant depends on the plant’s configuration, fuel types, and existing flue gas<br />

controls used for other pollutants. In addition, the complicated chemistry and multiple<br />

mechanisms governing mercury speciation in combustion facilities makes it necessary to<br />

investigate mercury emission control technologies at conditions relevant to each specific<br />

plant [1]. <strong>Mercury</strong> control technologies applied in power plants and waste incinerators<br />

are reviewed in this section and the applicability of these technologies in cement plants is<br />

discussed.


3.2 <strong>Mercury</strong> avoidance technology<br />

3.2.1 Coal cleaning<br />

Physical coal cleaning is used primarily to reduce the ash and pyritic sulfur<br />

content of coal [2-4]. Approximately 75% of the U.S. Eastern and Midwestern<br />

bituminous coals undergo physical coal cleaning prior to shipment to power plants. Less<br />

than 20% of the western coals, such as Powder River basin (PRB) coal and Colorado<br />

bituminous coals, are cleaned [5].<br />

Ash and pyritic sulfur are removed due to the difference in the densities of these<br />

materials compared to the organic constituents in the coal. <strong>Mercury</strong> present in a sulfide<br />

form also has a high density and can be removed during physical coal cleaning.<br />

Reduction in mercury levels in coals ranging <strong>from</strong> 10% up to 78% have been reported<br />

[3,6,7]. The average mercury reduction resulting <strong>from</strong> physical coal cleaning is estimated<br />

to be in the range of 20% to 37% [6,8]. The cost of waste water treatment is very high.<br />

As most of the mercury is <strong>from</strong> the raw material in cement production, the extent of<br />

mercury removal through coal cleaning is expected to be very limited.<br />

3.2.2 <strong>Cement</strong> raw material cleaning<br />

Two methods are proposed in the patents for removing mercury <strong>from</strong> cement raw<br />

materials, i.e., washing and gasification prior to feeding to the kiln [9,10]. By water<br />

washing the water-soluble mercury in the raw materials is removed. In the gasification<br />

process, raw materials are introduced into a heating furnace and mercury and its<br />

compounds contained in the raw materials are gasified. The resulting gas is introduced<br />

into an activated carbon adsorption tower, mercury and its compounds are adsorbed and<br />

separated. However, no results on the mercury removal efficiency <strong>by</strong> raw material<br />

cleaning are reported in the publications. Due to large amount of raw materials applied in<br />

the cement production, the cost of raw material cleaning is expected to be extremely high<br />

and this technology appears not suitable for mercury removal <strong>from</strong> cement plants.<br />

33


3.2.3 Fuel switching<br />

The use of tire-derived fuels (TDF) and substitution of coal and petroleum coke<br />

with natural gas could potentially result in a modest reduction in the mercury emissions<br />

due to the replacement of mercury-containing fossil fuels with low mercury fuels. Cofiring<br />

of TDF with a subbituminous coal in a 55 kw pilot-scale pulverized coal<br />

combustor had a significant effect on mercury speciation in the flue gas [11]. With 100%<br />

coal firing, there was only 16.8% oxidized mercury in the flue gas compared to 47.7%<br />

when 5 wt.% TDF was co-fired and 84.8% when 10% TDF was co-fired. The<br />

significantly enhanced mercury oxidation may be the result of additional homogeneous<br />

gas phase reactions between elemental mercury and the additional chlorine <strong>from</strong> TDF<br />

combustion. The chlorine content in TDF is about 600 ppmm. However, co-firing of<br />

TDF in the pilot-scale combustor with a hybrid filter for mercury removal demonstrated<br />

only limited improvement on mercury-emission control <strong>by</strong> the hybrid filter without<br />

sorbent injection. The enhanced mercury oxidation <strong>from</strong> co-firing TDF has potential in<br />

mercury emission control for power plants equipped with a wet flue gas desulphurization<br />

unit, since oxidized mercury is easily captured in the scrubber. Typically, kilns using<br />

TDF have a replacement rate no greater than 30% of the total fuel requirement. Richards<br />

et al. [12] summarized the available air emissions data for cement plants firing TDF and<br />

literature applicable to cement kilns and concluded that the variability in mercury<br />

concentrations and speciation overshadowed any beneficial impact on emissions due to<br />

the firing of TDF.<br />

Natural gas firing in boilers that are presently being fired with coal results in<br />

direct and significant reductions in mercury. However, as mentioned previously the fuel<br />

is usually not the dominant source of mercury in cement kiln flue gas. The limestone and<br />

possibly other raw materials in the kiln feed provide most of the mercury that is<br />

evaporated and emitted. Accordingly, the substitution of solid fuels would have only<br />

limited impact.<br />

34


3.3 <strong>Mercury</strong> removal <strong>by</strong> powdered activated carbon injection<br />

Powdered activated carbon (PAC) injection systems are well established as<br />

commercial air pollution control processes for a variety of volatile organic compounds,<br />

dioxin-furan, and heavy metals control applications [5]. The following three versions of<br />

PAC processes are being considered for widespread use in coal-fired power plants [6]: (1)<br />

PAC injection upstream of the existing dust collector system; (2) Gas cooling followed<br />

<strong>by</strong> PAC injection upstream of the existing dust collector system; (3) Gas cooling of the<br />

effluent gas stream of the existing dust collector system followed <strong>by</strong> PAC injection<br />

upstream of a second dust collector for removal of the adsorbent.<br />

The activated carbon particles remain suspended in the moving gas stream for<br />

periods of one to three seconds. They then deposit onto the dust cake formed on the filter<br />

bags. Additional mercury capture takes place when the mercury-containing gas stream<br />

passes through the sorbent-containing dust cake. Electrostatic precipitators (ESPs) are<br />

rarely used as the downstream polishing dust collector because the precipitated activated<br />

carbon is partially isolated <strong>from</strong> the gas stream once it reaches the collection plate<br />

surface.<br />

3.3.1 Parameters affecting mercury removal <strong>by</strong> activated carbon<br />

injection<br />

There are a large number of variables that affect the adsorption of mercury on<br />

powdered activated carbon. These include [5]: mercury speciation and concentration,<br />

sorbent physical and chemical properties such as particle size distribution, pore structure<br />

and distribution, and surface characteristics, gas temperature, flue gas composition,<br />

sorbent concentration, mercury-sorbent contact time, and adequacy of sorbent dispersion<br />

into the mercury containing gas stream.<br />

Due to the differences of these variables among plants, there are large variations<br />

in the reported PAC injection rates and mercury removal efficiencies in various studies<br />

and commercial systems [13]. Therefore, it does not make sense to compare the mercury<br />

35


emoval efficiencies and sorbent injection rates without considering the actual conditions<br />

in the specific plants.<br />

<strong>Mercury</strong> speciation determines the mercury capture capacity of sorbents at a<br />

given temperature. Pavlish and others [13] concluded that virgin activated carbon has a<br />

higher rate of capture for mercuric chloride than for elemental mercury. Ho and others<br />

[14] reported that sulphur-impregnated activated carbons have enhanced rates of<br />

elemental mercury capture.<br />

Pavlish et al. [13] conducted a detailed review on possible rate-controlling<br />

mechanisms for mercury removal <strong>by</strong> sorbent injection. The overall reaction rates may be<br />

limited <strong>by</strong> mass transfer <strong>from</strong> the bulk gas to the sorbent surface, the equilibrium<br />

adsorption capacity, and the rates of reactions occurring on the sorbent surface.<br />

All adsorption processes, especially those dependent on physisorption operate<br />

more effectively at low temperatures due to the large adsorption capacity at low<br />

temperatures. Adsorption processes for flue gas cleaning usually are operated in the<br />

temperature range of 150°C to 200°C. The pilot plant studies of PAC injection indicate<br />

that the mercury removal efficiency is strongly dependent on the gas temperatures [5].<br />

Efficiencies of 10% to 70% have been measured at 170°C, and removals of 90% to 99%<br />

have been measured at 100°C. <strong>Mercury</strong> capture takes place <strong>by</strong> both physisorption and<br />

chemisorption [13]. With increasing temperature, physical adsorption decreases due to<br />

the nature of exothermal adsorption process whilst chemisorption might be enhanced on<br />

kinetics [15].<br />

<strong>Mercury</strong> competes with a variety of gases for the adsorption sites on the activated<br />

carbon. Water vapor is important because it is present at concentrations many orders of<br />

magnitude above mercury. At moisture levels above 5% to 10%, moisture competition<br />

can be significant. There are indications that high moisture levels in the flue gas will<br />

suppress the capture of mercury <strong>by</strong> activated carbon [5,16]. It was postulated that water<br />

molecules are able to fill micropores, there<strong>by</strong> blocking adsorption sites for mercury.<br />

Although it is agreed that water plays an important role in the mechanism of<br />

mercury capture, there is disagreement in the literature about the effect of water on<br />

36


mercury removal. Pavlish et al. [13] reported that reintroducing water into flue gas in a<br />

lab-scale reactor at 135�C after a period of sorption testing on dry flue gas resulted in an<br />

immediate release of mercury <strong>from</strong> the activated carbon. However, in another lab-scale<br />

study [17] the presence of moisture on the carbon surface was reported to promote<br />

mercury bonding. About 75–85% reduction in Hg 0 adsorption capacity was observed<br />

when the carbon samples’ moisture at a level of 2 wt.% was removed <strong>by</strong> heating at<br />

110�C prior to the Hg 0 adsorption experiments at room temperature. These observations<br />

suggest that the moisture adsorbed on activated carbons plays a critical role in retaining<br />

Hg 0 . It was postulated that adsorbed H2O is closely associated with surface oxygen<br />

complexes and the removal of the H2O <strong>from</strong> the carbon surface <strong>by</strong> low-temperature heat<br />

treatment reduces the number of active sites that can chemically bond Hg 0 or eliminates<br />

the reactive surface conditions that favor Hg 0 adsorption [17]. Liu et al. [18] found that<br />

the mercury adsorption capacity of sulfur impregnated activated carbon did not change<br />

significantly when 5 vol.% water was added to the dry gas at 140�C, however, the<br />

adsorption capacity decreased 25% when the water content in the gas increased to 10<br />

vol.%. These observations indicate that it is important to investigate the sorbent using the<br />

same water vapor content as in the full-scale plant flue gas or in a wide range of moisture<br />

content. Further investigations of the effect of water on mercury adsorption are desired to<br />

reveal the dominating effects.<br />

Miller et al. [19] and Ochiai et al. [20] conducted full factorial design<br />

experiments in fixed-bed reactors to determine the relative effects of SO2, HCl, NO, and<br />

NO2 on the elemental mercury capture ability of commercial activated carbons. Without<br />

acid gases present, upon exposure to a baseline gas mixture of 6% O2, 12% CO2, 8%<br />

H2O, and N2, the lignite activated carbon sorbent provided only about 10–20% initial<br />

mercury capture of Hg 0 for about 30 min and then fast breakthrough at 107°C [19,20].<br />

Adding 50 ppmv HCl alone with the baseline gas improves the mercury adsorption<br />

significantly [19,20]. It was also found out that adding NO or NO2 alone with the<br />

baseline gas also improves the mercury adsorption capacity significantly. The mercury<br />

capture increases <strong>from</strong> 10-20% for about 30 min using the baseline gas to 90-100% for<br />

37


more than 2-6 h when 300 ppmv NO or 20 ppmv NO2 was added one at a time to the<br />

baseline gases at 107°C.<br />

When 1600 ppmv SO2 is added to the baseline gas the mercury adsorption<br />

capacity will not be changed or only improved slightly [19,21]. Addition of 20 ppm SO3<br />

to the gas reduced mercury capture <strong>by</strong> nearly 80%, and higher SO3 concentrations led to<br />

further reductions in the mercury capture [21-23]. The competition between SO3 and<br />

mercury for binding sites on the surface of activated carbon decreases the mercury<br />

adsorption capacity.<br />

The combination of 1600 ppmv SO2 and 20 ppmv NO2 additions resulted in<br />

significant different mercury breakthrough profile compared to adding NO2 alone [19,20].<br />

A highly significant interaction between SO2 and NO2 caused a rapid breakthrough of<br />

mercury and is the controlling mechanism responsible for poor sorbent performance. The<br />

detailed mechanism of SO2 and NO2 interaction is presented in section 3.3.5.2.<br />

3.3.2 Tests of mercury sorbents in lab­scale fixed­bed reactors<br />

A good sorbent is expected to have high mercury adsorption capacity and fast<br />

kinetics. A sorbent with good capacity but slow kinetics is not a good choice as it takes<br />

mercury compound molecules too long time to reach the particle interior [24]. On the<br />

other hand, a sorbent with fast kinetics but low capacity is not good either as a large<br />

amount of sorbent is required for a given mercury removal. To satisfy these two<br />

requirements, the sorbent must have a reasonably high surface area or micropore volume<br />

and a pore network of relatively large pores for the transport of molecules to the interior.<br />

3.3.2.1 Carbon­based sorbents<br />

As mentioned previously mercury capture is very sensitive to the flue gas<br />

composition and temperature, and for this reason only the mercury capture capacities of<br />

sorbents tested under simulated flue gas conditions are reported and compared here.<br />

Extensive research has been conducted to study the sorbent mercury capture capacity<br />

mainly using lab-scale fixed-bed reactors [25-36]. The mercury sorbents can be divided<br />

38


into three groups, i.e., virgin carbon sorbents, chemically treated carbons, and noncarbon<br />

based sorbents. The majority of the publications focused on elemental mercury<br />

capture and only few studies investigated capture of HgCl2.<br />

Typical properties of selected sorbents are reported in table 3.1. Coal source and<br />

ash content relate to the composition of the coal and characterize the state of the carbon<br />

such as fixed carbon/volatile ratio [37,38]. High ash content reduces the overall activity<br />

of the activated carbon. Particle surface area is a measure of adsorption capacity and<br />

describes the available surface for mercury adsorption. Pore size and distribution is an<br />

indicator of sorbent quality, with smaller pores preferred. Particle size is used to describe<br />

the degree of sorbent physical preparation. The mean particle size and distribution of<br />

particle size are important parameters for evaluating mercury removal rate and pressure<br />

drop. Smaller size provides faster rate of adsorption and results in larger pressure drop.<br />

Content of bromine/chlorine/sulfur is considered as an indicator of the chemical<br />

characteristics of sorbent responsible for mercury adsorption. Bulk density reflects a<br />

gross approximation of the processing or surface area of a given carbon. The most<br />

investigated sorbents in the literature is Darco FGD, which is a commercial lignite based<br />

powdered activated carbon and is developed for heavy metal removal <strong>from</strong> incinerators<br />

and power plants.<br />

Table 3.1. Properties of selected sorbents.<br />

<strong>Sorbent</strong>s Sources Ash<br />

(wt%)<br />

Norit<br />

Darco<br />

FGD<br />

Norit<br />

Darco<br />

Insul<br />

Norit<br />

Darco<br />

Hg<br />

Norit<br />

Darco<br />

Hg-LH<br />

Lignite coal 28.2-<br />

32.1<br />

Sulfur<br />

(wt%)<br />

0.86-<br />

1.1<br />

Surface<br />

area<br />

(m 2 /g)<br />

481-<br />

600<br />

Pore<br />

volume<br />

(cm 3 /g)<br />

0.535-<br />

0.610<br />

39<br />

Average<br />

pore<br />

size<br />

(nm)<br />

Mass<br />

mean<br />

particle<br />

size(µm)<br />

Bulk<br />

density<br />

(g/cm 3 )<br />

3.2 6.8-15 0.51 0.49-<br />

0.58<br />

Porosity Reference<br />

[25-27,29-<br />

32,35,39]<br />

Lignite coal,<br />

based on<br />

Darco FGD,<br />

fine,<br />

chemically<br />

treated<br />

700 6 0.32 [33,34]<br />

Lignite coal 1.2 600 16-19 0.51 [40,41]<br />

Lignite coal,<br />

bromine<br />

treated<br />

1.2 550 16-19 0.60 [41,42]


<strong>Sorbent</strong>s Sources Ash<br />

(wt%)<br />

Calgon<br />

FluePac<br />

AC<br />

Calgon<br />

HGR<br />

HOK<br />

standard<br />

HOK<br />

super<br />

Masda<br />

GAC<br />

Bituminous<br />

coal<br />

Sulfur<br />

(wt%)<br />

Surface<br />

area<br />

(m 2 /g)<br />

Pore<br />

volume<br />

(cm 3 /g)<br />

40<br />

Average<br />

pore<br />

size<br />

(nm)<br />

Mass<br />

mean<br />

particle<br />

size(µm)<br />

Bulk<br />

density<br />

(g/cm 3 )<br />

Porosity Reference<br />

5.8 0.7 606 0.285 32 0.585 [33,34,43]<br />

Bituminous,<br />

10.9- 413- 0.130 2.0 9.8 0.590 [44-46]<br />

sulfur treated<br />

15 486<br />

Lignite coal 10.0 0.60 300 0.620 63 0.55 [47]<br />

MnO2-AC MnO2<br />

solution<br />

impregnated<br />

activated<br />

carbon<br />

FeCl3-AC FeCl3<br />

solution<br />

impregnated<br />

activated<br />

Shanghai<br />

activated<br />

carbon<br />

Damao<br />

activated<br />

carbon<br />

1%<br />

ZnCl2<br />

Damao<br />

5%<br />

ZnCl2<br />

Damao<br />

Lignite coal 10.0 0.60 300 24 0.44 [47]<br />

carbon<br />

produced<br />

<strong>from</strong> wood<br />

<strong>by</strong> zinc<br />

chloride<br />

method<br />

Bituminous<br />

coal<br />

Zncl2<br />

treated<br />

Damao<br />

carbon<br />

Zncl 2<br />

treated<br />

Damao<br />

carbon<br />

6.0 735 0.300 2.0 [48]<br />

0.43 865 0.290 90 0.45 [28]<br />

0.44 1470 0.920 90 0.58 [28]<br />

0.55 1850 1.050 90 0.67 [28]<br />

770 0.330 1.7 280 [49]<br />

608 0.27 1.8 280 [49]<br />

277 0.19 2.7 280 [49]<br />

There is disagreement in the publications on the effect of mercury species on<br />

activated carbon and char adsorption capacity. Yang et al. [13,50] reported that the<br />

capture capacities of HgCl2 <strong>by</strong> bituminous char are larger <strong>by</strong> a factor of two than those of<br />

elemental mercury using only CO2, O2, H2O and N2. However, other studies, as shown<br />

in figure 3.1, show the opposite trend [25-27]. The elemental mercury adsorption<br />

capacity for the studied carbons is about 0.5-4 times larger than the HgCl2 adsorption<br />

capacity in the temperature of 110-160 �C using simulated flue gas with 6% O2, 12%<br />

CO2, 7% H2O, 50 ppmv HCl and 1600 ppmv SO2.


Adsorption capacity, �g Hg/HgCl2/g-carbon<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

45 �g/m3 Hg0,Darco FGD [26, 27]<br />

54 �g/m3 Hg0,lab AC <strong>from</strong> high S coal [25]<br />

Figure 3.1. Effect of mercury speciation on mercury adsorption capacity on activated<br />

carbon. Adsorption temperature is 135�C and data are <strong>from</strong> [25-27]. Simulated flue gas<br />

with 6% O2, 12% CO2, 7% H2O, 50 ppmv HCl and 1600 ppmv SO2.<br />

Generally, the mercury adsorption capacities of carbon sorbents decrease when<br />

the gas temperature is increased [13,50]. However, tests of Darco FGD at 100-135�C<br />

(see figure 3.2) <strong>from</strong> different studies are not in agreement with the trend. This is<br />

probably due to the short exposure time for the test at 100�C and the presence of SO3 in<br />

the simulated gas for test at 120�C. The adsorption capacity at 100�C was measured after<br />

2 h and before the complete breakthrough, while other tests were run until the complete<br />

breakthrough of the sorbent bed was obtained. As discussed previously, the presence of<br />

SO3 in the flue gas will decreases the sorbents’ mercury adsorption capacity. This<br />

observation again illustrates the difficulty of analyzing the results <strong>from</strong> different studies<br />

that are not conducted under the same conditions.<br />

41<br />

59 �g/m3 Hg0,Darco FGD [25]<br />

60 �g/m3 HgCl 2 ,Darco FGD [25]<br />

61 �g/m3 HgCl 2 ,lab AC <strong>from</strong> high S coal [25]


Hg adsorption capacity, �g Hg/g_carbon<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

100oC, 400 �g Hg 0 /m3 [29-31]<br />

120oC, 290 �g Hg 0 /m3, SO 3 [13,32]<br />

Figure 3.2. Effect of adsorption temperature on mercury adsorption capacity on Darco<br />

FGD activated carbon. Data are <strong>from</strong> [13,25-27,29-32]. Gas composition: 100�C: 1000<br />

ppmv SO2, 50 ppmv HCl in N2; 120�C: 8.3% O2, 14.8% CO2, 7.2% H2O, 278 ppmv NO,<br />

650 ppmv SO2, 107 ppmv CO, 46 ppmv HCl, 27 ppmv SO3 in N2; 135�C: 6% O2, 12%<br />

CO2, 7% H2O, 1600 ppmv SO2, 50 ppmv HCl in N2;<br />

Previous studies showed that the low chlorine concentration in the flue gas <strong>from</strong><br />

combustion of low-rank coals is a major limiting factor in the mercury control<br />

performance using the virgin activated carbons [13,50]. Various chemically treated<br />

carbons were developed to compensate for the lack of halogens in the combustion flue<br />

gas. These include chloride-impregnated carbons [49,51-55], sulfur-impregnated carbons<br />

[25,51,53,56-61], brominated carbons [50,55,62,63], iodine-impregnated carbons [52,53],<br />

ozone-treated carbon [64], and carbon impregnated with metal compounds such as MnO2,<br />

FeCl3 and CuCl2 [28,65-67]. The price of the chemically treated carbon is typically<br />

higher than the non-treated one. The price of the non-treated Norit Darco Hg was about<br />

1.1 US$/kg in 2007, while the bromine-treated carbons cost about 1.9-2.6 US$/kg [41].<br />

42<br />

135oC, 45 �g Hg 0 /m3 [26,27]<br />

135oC, 59 �g Hg 0 /m3 [25]


Chlorine impregnation of a virgin activated carbon using dilute solutions of<br />

hydrogen chloride leads to increases in fixed-bed capture of both elemental mercury and<br />

mercuric chloride <strong>by</strong> a factor of 2-3 for a simulated flue gas without HCl, but with 7.1%<br />

O2, 6.9% H2O, 3.4% CO2, 4.5 ppmv CO, 200 ppmv NOx and 500 ppmv SO2 [51]. It is<br />

not reported how the chlorine impregnated carbon behaviors if HCl is present in the gas.<br />

Coal-derived activated carbon <strong>from</strong> high-organic-sulfur coals was reported to<br />

have a greater equilibrium Hg 0 adsorption capacity than that prepared <strong>from</strong> low-organicsulfur<br />

coal when tested using a simulated flue gas with 6% O2, 7% H2O, 12% CO2, 50<br />

ppmv HCl, and 1600 ppmv SO2 [25]. At 135�C the equilibrium Hg 0 adsorption capacity<br />

of carbon derived <strong>from</strong> high-organic-sulfur coal, which contained 3.7 wt % total sulfur<br />

and 2.9 wt% organic sulfur is 2718 µg Hg 0 /g_carbon, on the other hand the equilibrium<br />

Hg 0 adsorption capacity of carbon prepared <strong>from</strong> low-organic-sulfur coal with 1.2 wt%<br />

total sulfur and 0.7 wt% organic sulfur was only 1304 µg Hg 0 /g_carbon. When the loworganic-sulfur<br />

coal-derived activated carbon is impregnated with elemental sulphur at<br />

600°C, its equilibrium Hg 0 adsorption capacity is comparable to the adsorption capacity<br />

of the activated carbon <strong>from</strong> the high-organic-surfur coal. Elemental sulphurimpregnated<br />

carbons enhance elemental mercury removal due to the formation of<br />

mercury sulphide on the carbon surface [68]. A portion of the inherent organic sulphur<br />

in the starting coal, which remained in the activated carbon, plays an important role in<br />

adsorption of elemental mercury. Besides organic sulphur, the surface area and<br />

micropore area of the activated carbon also influence Hg 0 adsorption capacity [25]. The<br />

HgCl2 adsorption capacity is not as dependent on the surface area and concentration of<br />

sulphur in the activated carbon as for adsorption of Hg 0 .<br />

Another method for modifying carbon surfaces is oxidation, using reagents that<br />

include oxygen, ozone, hydrogen peroxide, nitric acid, and permanganate [64]. Ozone<br />

treatment of carbon surfaces leads to large increases in the elemental mercury capture<br />

capacity <strong>by</strong> more than a factor of 100 when tested in Argon gas, but the activity is easily<br />

destroyed <strong>by</strong> exposure to air, to water vapor, or <strong>by</strong> mild heating at 120�C [64]. Freshly<br />

ozone-treated carbon surfaces are shown to form labile C–O containing oxidizing groups,<br />

43


which are likely to be epoxides or secondary ozonides. However, this ability fades with<br />

aging. The finding opens the possibility of in-situ carbon ozonolysis to create fresh,<br />

super-active sorbents with the additional benefit of sorbent hydrophilicity useful in<br />

certain applications. Ozone treatment of fly ash carbon has been reported to inhibit the<br />

adsorption of commercial surfactants in concrete paste, thus mitigating the known<br />

negative effects of carbon on ash utilization [69-73]. Therefore, the enhanced mercury<br />

removal could be a co-benefit of the ozone treatment.<br />

3.3.2.2 Non­carbon sorbents<br />

For cement plant application a non-carbon sorbent is more attractive if the used<br />

sorbent can be added to the final cement product or can be separated and regenerated.<br />

Carbon can deteriorate the cement quality if the used carbon is not separated <strong>from</strong> the<br />

cement materials <strong>by</strong> installing an expensive polishing filter. As inspiration to cement<br />

plant application, research on development of non-carbon sorbents that do not adversely<br />

impact sales of fly ash as a coal combustion <strong>by</strong>product for Portland cement and concrete<br />

production are reported in this section.<br />

Chemically synthesized manganese oxides powder has been demonstrated in<br />

power plants to remove mercury, NOx and SO2 <strong>from</strong> flue gas [74-77]. The reacted<br />

sorbent can be regenerated <strong>by</strong> a wet chemical process if the sorbent is injected just before<br />

the added polishing fabric filter [74-77]. The simplified capture reactions for these<br />

pollutants are suggested as following:<br />

0<br />

Hg + MnO2 Mn*Hg complex<br />

� (R3.1)<br />

NO x + MnO 2 � Mn(NO 3) 2<br />

(R3.2)<br />

SO x + MnO2 � MnSO4<br />

(R3.3)<br />

Non-carbonaceous materials or mineral oxides including silica gel, alumina,<br />

molecular sieves, zeolites, and montmorillonite have been modified with various<br />

functional groups such as amine, amide, thiol, urea, and additives such as elemental<br />

sulfur, sodium sulfide, and sodium polysulfide to examine their potential as sorbents for<br />

the removal of mercury vapor at coal-fired utility power plants [78]. A number of sorbent<br />

44


candidates such as amine-silica gel, urea-silica gel, thiol-silica gel, amide-silica gel,<br />

sulfur-alumina, sulfur-molecular sieve, sulfur-montmorillonite, sodium sulfidemontmorillonite,<br />

and sodium polysulfide-montmorillonite, were synthesized and tested in<br />

a lab-scale fixed-bed system under an argon flow for screening purposes at 70°C and<br />

140°C. Several functionalized silica materials used for effective control of heavy metals<br />

in the aqueous phase showed insignificant adsorption capacities for mercury control in<br />

the gas phase, suggesting that mercury removal mechanisms are different in these two<br />

phases. Among the synthesized samples, sodium polysulfide-impregnated<br />

montmorillonite showed a moderate adsorption capacity at 70°C.<br />

The commercial Amended Silicates sorbent uses silicate minerals as substrate<br />

particles on which a chemical reagent with a strong affinity for mercury and mercury<br />

compounds is impregnated [79,80]. A phyllosilicate substrate, for example, vermiculite<br />

or montmorillonite, is used as an inexpensive support to a thin layer for a polyvalent<br />

metal sulfide, ensuring that more of the metal sulfide is engaged in the sorption process.<br />

The sorbent is prepared <strong>by</strong> ion exchange between the silicate substrate material and a<br />

solution containing one or more of a group of polyvalent metals including tin, iron,<br />

titanium, manganese, zirconium, and molybdenum. Controlled addition of sulfide ions to<br />

the exchanged silicate substrate produces the sorbent. The silicates provide a low-cost<br />

substrate material with average particle size of a few microns and extended surface area<br />

for the amendment process. Due to their high silicate content, they have been proven<br />

compatible with the continued sale of fly ash as a pozzolan material for concrete and<br />

cement production. The price of the Amended Silicates sorbent is about 2.2-4.4 US$/kg<br />

[81], which is comparable to the price of the chemically treated carbons [41]. However,<br />

the performance data of the Amended Silicates sorbent are rarely reported due to the<br />

concern of intelligent property.<br />

A comparison between Darco FGD activated carbon and Ca(OH)2 indicated that<br />

non-carbon-based sorbents with relatively high Ca contents can be fairly effective HgCl2<br />

sorbents [29,30]. The Ca-based sorbents exhibited HgCl2 removal as high as half of the<br />

removal shown <strong>by</strong> the Darco FGD activated carbon when 100 mg sorbent was tested in a<br />

45


ench-scale fixed-bed reactor using simulated flue gas containing 10% CO2, 7% O2, 5%<br />

H2O, and 173 ppmv SO2 [29,30]. However, the carbon-based sorbent showed superior<br />

efficiency of elemental removal compared to Ca-based sorbent.<br />

Full-scale investigations in coal-fired power plants have observed mercury<br />

capture <strong>by</strong> unburned carbon in the fly ash [82]. <strong>Mercury</strong> removal <strong>by</strong> fly ash has also been<br />

extensively studied to find a solution to the expensive mercury sorbents [36,83-87]. As<br />

shown in figure 3.3, the amount of carbon in the fly ash has a strong effect on mercury<br />

adsorption capacity of the fly ash. The mercury adsorption capacity increases with<br />

carbon content in the fly ash, however, it is not directly proportional to the carbon<br />

content. The mercury adsorption capacity of Nixon fly ash with 2% residual carbon is<br />

about 30% of the commercial activated carbon Darco G60 [86].<br />

Adsorbed Hg (g Hg/106 g ash)<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Nixon, 2% C<br />

Cherokee, 8.7% C<br />

Clark, 32.7% C<br />

Huntington, 35.9% C<br />

0 10 20 30 40<br />

Carbon content in the fly ash (%)<br />

Figure 3.3. <strong>Mercury</strong> adsorption capacity on fly ashes with different carbon content. Data<br />

are <strong>from</strong> [86]. The applied adsorption temperature is 121�C and elemental mercury<br />

concentration is 4 mg/m 3 with nitrogen as balance gas.<br />

Dunham et al. [85] investigated 16 fly ash samples <strong>from</strong> a variety of sources and<br />

coal types in a fixed-bed reactor at 121-177�C using elemental mercury or HgCl2 in<br />

simulated flue gas mixtures of O2, SO2, NO, NO2, H2O and HCl. While many of the ash<br />

46


samples oxidized elemental mercury to HgCl2 in a range of 15-85%, not all of the<br />

samples that oxidized mercury also captured elemental mercury. However, no capture of<br />

elemental mercury was observed without accompanying oxidation. In general, oxidation<br />

of elemental mercury increased with increasing amount of magnetite (Fe3O4) in the ash.<br />

However, one high-carbon subbituminous ash with no magnetite showed considerable<br />

mercury oxidation that may have been due to the carbon. Dunham et al. [85] suggested<br />

that an iron oxide with a spinel-type structure is active in fly ash with respect to mercury<br />

oxidation. Surface area as well as the nature of the surface, such as the oxygen<br />

functionality and presence of halogen species appeared to be important for oxidation and<br />

adsorption of elemental mercury. For the applied gas composition in Dunham’s study<br />

[85], the capacity of the ash samples for HgCl2 was similar to that for elemental mercury.<br />

There was a good correlation between the capacity for HgCl2 and the surface area. The<br />

correlation between HgCl2 and loss on ignition was not as strong, suggesting that it is<br />

not the carbon content alone but also properties of the ash, such as surface area, that<br />

influence capture of HgCl2.<br />

Based on the research of interactions between mercury and fly ash, carbon that<br />

remains in pulverized coal fly ash could be used as an inexpensive adsorbent for mercury<br />

removal. The fly ash would be injected into the flue gas prior to the particulate control<br />

device [86,88] similarly to the way in which activated carbon is used, thus eliminating<br />

large capital and sorbent costs. Due to the low carbon content and small mercury<br />

adsorption capacity of fly ash, however, a large amount of fly ash may be required.<br />

Another alternative to activated carbon might be the use of noble metal-based<br />

sorbent. Noble metals such as gold and silver form reversible amalgams with mercury<br />

[89]. A class of magnetic zeolite composites with supported silver nanoparticles has been<br />

tested for elemental mercury removal <strong>from</strong> power plant flue gas [89]. Gaseous mercury is<br />

captured <strong>by</strong> the sorbent and the mercury-laden sorbent particles are collected <strong>by</strong> an<br />

existing dust collector and separated <strong>from</strong> the fly ash <strong>by</strong> magnetic separation. After mild<br />

heat treatment to release captured mercury the sorbent is regenerated for the next cycle of<br />

mercury capture. The technology is still in early stage and research is required regarding<br />

47


the stability of the sorbent and possible regeneration cycles. Since noble metal is used in<br />

synthesis of the sorbent it is expected that the sorbent is expensive. It is not clear how<br />

this process could be cost effective compared to the activated carbon injection system<br />

and the released mercury also must be captured <strong>by</strong> some sort of process.<br />

3.2.2.3 In­situ produced sorbents<br />

To reduce the cost of sorbents, methods for in-situ production of activated carbon<br />

<strong>from</strong> coal-fired power plants have been invented [33,34,90]. In the so-called Thief<br />

process, partially combusted coal <strong>from</strong> the furnace of a pulverized coal power generation<br />

plant is extracted <strong>by</strong> a lance and then re-injected into the ductwork downstream of the air<br />

preheater [33,34,90]. Tests show that the Thief sorbents exhibit capacities for mercury<br />

<strong>from</strong> flue gas streams that are comparable to those exhibited <strong>by</strong> commercially available<br />

activated carbons. The process extracts 0.1-0.5% of the furnace gas in the boiler<br />

depending on the desired sorbent injection rate and mercury removal level. The mass of<br />

solids extracted <strong>from</strong> the furnace is very small in comparison to the mass of coal being<br />

burned. The estimated heat loss is less than 0.3% for a 500 MWe power plant burning<br />

PRB subbituminous coal.<br />

Another process uses an oxy-fuel burner to devolatilize and activate the coal to<br />

produce activated carbon [33,34,90]. In the burner natural gas is combusted together with<br />

an oxygen stream, producing a high temperature oxygen-rich stream which passes<br />

through a nozzle. Downstream of the hot oxygen nozzle the parent coal mixes with the<br />

hot oxygen and begins to burn. Devolatilization and activation take place in a reactor<br />

which leads to a particle separation step where the product is separated <strong>from</strong> the syngas<br />

stream. The syngas can then be ducted to the boiler to provide added fuel value. At<br />

several points in the process additives can be introduced to dope the product, or to<br />

control the product morphology.<br />

48


3.3.3 <strong>Sorbent</strong> injection in power plants<br />

Many activated carbons have been tested in U.S. power plants. Table 3.2 presents<br />

the tested sorbents and applied APCDs and coals. The mercury removal efficiencies are<br />

not included in the table due to various mercury removal efficiencies obtained at<br />

complicated test conditions. Instead the mercury removal efficiencies as a function of<br />

sorbent injection rate are shown in figures.<br />

The most studied sorbents are Darco FGD, Darco Hg, and Darco Hg-LH. The<br />

Darco Hg is formerly known as Darco FGD manufactured specifically for the removal of<br />

mercury in coal fired utility flue gas emission streams [80], while Darco Hg-LH is<br />

bromine impregnated. Although the mercury levels at the inlet of ACPDs are generally<br />

similar, the extents of mercury removal <strong>by</strong> the existing APCDs without sorbent injection<br />

are quite different. This is due to fact that different ranks of coal and APCD<br />

configurations are applied <strong>by</strong> different power plants.<br />

Without looking at the detailed data of the specific plants, it is difficult to<br />

evaluate the sorbent performance <strong>by</strong> comparing the mercury removal efficiency.<br />

However, some trends can be observed <strong>by</strong> comparing the results obtained under similar<br />

conditions. Figure 3.4 compares the mercury removal at Holcomb and Stanton power<br />

station <strong>by</strong> injection of Darco Hg-LH upstream of SDA and baghouse. At Holcomb and<br />

Stanton power station the <strong>by</strong>product <strong>from</strong> SDA is disposed and therefore activated<br />

carbon is injected before the SDA and baghouse. Figure 3.5 illustrates the mercury<br />

removal <strong>by</strong> Darco FGD injection upstream of a new added so-called COHPAC compact<br />

hybrid particle collector. COHPAC is an EPRI-patented design that places a high air-tocloth<br />

ratio fabric filter downstream of an existing ESP to improve overall particulate<br />

collection efficiency. The results of mercury removal <strong>by</strong> Darco Hg injection upstream of<br />

cold-side ESP are presented in figure 3.6. Up to 80 mg/m 3 activated carbon is applied for<br />

systems with FF, while up to 320 mg/m 3 carbon is injected upstream of cold-side ESP.<br />

49


Table 3.2. Summary of full-scale tests conducted in U.S. power plants. LNB: low NOx<br />

burner, COHPAC: compact hybrid particulate collector<br />

Location Test<br />

load<br />

MW<br />

Holcomb, unit 1, 360<br />

MW [91-93]<br />

180,<br />

360<br />

Coal APCD Inlet mercury<br />

�g/Nm 3 ,dry<br />

PRB SDA@143�C<br />

+Baghouse,<br />

LNB<br />

50<br />

<strong>Mercury</strong><br />

removal<br />

without<br />

sorbent, %<br />

<strong>Sorbent</strong>s<br />

10-12 0-13 Darco Hg, Darco<br />

Hg-LH, Calgon<br />

208CP<br />

Stanton, unit 10, 60<br />

MW [93]<br />

60 Lignite SDA, baghouse - - Darco Hg-LH<br />

Stanton unit 1, 150 - PRB Cold-side ESP - 15 Brominated PAC<br />

MW [94]<br />

(B-PAC)<br />

Gaston, unit 3, 270 135 Low sulfur COHPAC@143� 7-10 6 Darco FGD, fine<br />

MW [95,96]<br />

bituminous C, hot side ESP,<br />

FGD, Insul, ESP<br />

coal<br />

LNB<br />

ash<br />

Big Brown, unit 2, 150 30%PRB/70 ESP, COHPAC - - Darco FGD,<br />

600 MW[97]<br />

% lignite,PRB @ 177�C, LNB<br />

FGD/NaCl/CaCl2 Presque Isle, unit 7-9, 90 PRB Polishing - - Darco FGD<br />

90 MW[98,99]<br />

baghouse<br />

Meramec, unit 2, 140 70 PRB Cold-side ESP 10-12 15-30 Darco Hg-LH,<br />

MW [91,93,100]<br />

@160�C, LNB<br />

Darco Hg<br />

Pleasant Prairie, unit 150 PRB Cold-side 16-17 5 Darco FGD, Darco<br />

2, 600 MW [96,101]<br />

ESP@138�C,<br />

Hg, Insul, lime,<br />

SO3 conditioning<br />

Sorbalit<br />

Brayton Point, unit 1, 125 Low sulfur Cold-side ESP 17 - Darco FGD, Darco<br />

250 MW [102]<br />

bituminous @138�C, SO3<br />

Hg, HOK, LAC<br />

coal<br />

conditioning<br />

Leland Olds, unit 1, 220 Lignite Cold-side ESP, 6-7 - Darco FGD Hg<br />

220 MW [93]<br />

LNB<br />

/CaCl2<br />

St. Clair, unit 1, 145 145 85%PRB/15 Cold-side ESP - - Brominated PAC<br />

MW [93]<br />

% bituminous<br />

coal<br />

(B-PAC)<br />

Laramie River unit 3, 140 PRB SDA+cold-side 10-12 4 Darco Hg-LH<br />

550 MW[91,103]<br />

ESP<br />

Darco Hg<br />

Monroe, unit 4, 775 196 PRB/bitumino Cold-side 5-10 10-30 Darco Hg-LH,<br />

MW [91]<br />

us coal ESP@125�C,<br />

Darco Hg, Darco<br />

SCR<br />

XTR<br />

Conesville, unit 6, 400 Bituminous Cold-side ESP, 15-30 50 Darco Hg-LH,<br />

400 MW [91]<br />

coal<br />

wet FGD<br />

Darco Hg<br />

Plant Yates, unit 1, 100 Bituminous Cold-side ESP, - - Super HOK<br />

100 MW[93]<br />

coal<br />

wet FGD<br />

Salem Harbor unit 1, 85 Low sulfur Cold-side - - Darco FGD<br />

85 MW [104]<br />

bituminous ESP@125�C,<br />

coal<br />

LNB<br />

Ameren Labadie unit 630 PRB Cold-side 5-12


As shown in figure 3.4 and 3.5, mercury can be efficiently removed <strong>by</strong> activated<br />

carbon injection upstream of SDA/baghouse or a polishing baghouse. When 32 mg/m 3<br />

Darco Hg-LH in SDA/baghouse system and Darco Hg in polishing baghouse system are<br />

applied, about 80% of mercury can be removed. Further increase of the carbon injection<br />

rate above 32 mg/m 3 results in a slow increase of the mercury removal efficiency. The<br />

mercury removal efficiency in SDA/baghouse system <strong>by</strong> Darco Hg-LH is larger than that<br />

<strong>by</strong> Darco Hg. This is due to the applied Darco Hg-LH sorbent, which is bromine<br />

impregnated and has larger mercury adsorption capacity than Darco FGD. The waste<br />

disposal cost of sorbent injection upstream of SDA/baghouse is expected to be higher<br />

since used activated carbon cannot be separated <strong>from</strong> the desulphurization product and<br />

regenerated. Tests at Gaston power station showed that carbon injection significantly<br />

increased the cleaning frequency of the COHPAC baghouse [95,96]. At an injection<br />

concentration of 32 mg/m 3 the cleaning frequency increased <strong>from</strong> 0.5 to 2<br />

pulses/bag/hour, most likely due to the small particle size of the PAC causes a high<br />

pressure drop.<br />

<strong>Mercury</strong> removal efficiency, %<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Stanton, lignite, Darco Hg-LH<br />

Holcomb, PRB, Darco Hg-LH<br />

Holcomb, PRB, Darco Hg<br />

0 20 40 60 80 100<br />

PAC injection rate, mg/m 3<br />

Figure 3.4. <strong>Mercury</strong> removal as a function of injection rate of Darco Hg-LH sorbent in<br />

power plants using SDA and baghouse as APCDs. Data are <strong>from</strong> [91-93].<br />

51


<strong>Mercury</strong> removal efficiency, %<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Gaston, bituminous<br />

Big Brown, 30%PRB/70%lignite<br />

Presque Isle, PRB<br />

0 20 40 60 80<br />

PAC injection rate, mg/m 3<br />

Figure 3.5. <strong>Mercury</strong> removal as a function of injection rate of Darco FGD sorbent in<br />

power plants <strong>by</strong> sorbent injection upstream of a polishing baghouse. Data are <strong>from</strong> [95-<br />

99].<br />

As shown in figure 3.6, much more than 32 mg/m 3 Darco Hg activated carbon are<br />

required to obtain 80% mercury removal <strong>by</strong> carbon injection upstream of cold-side ESP,<br />

where the flue gas temperature is about 125-160�C. This is due to the short contact time<br />

between mercury vapor and injected carbon in the ESP and mercury is mainly captured<br />

during the carbon particle in-flight period. When bromine treated carbons B-PAC and<br />

Darco Hg-LH is used, the mercury removal efficiency across the cold side ESP increases<br />

significantly.<br />

52


<strong>Mercury</strong> removal efficiency, %<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 50 100 150 200 250 300 350<br />

PAC injection rate, mg/m 3<br />

53<br />

Pleasant, PRB,Darco Hg<br />

Brayton, bituminous,Darco Hg<br />

Meramec, PRB,Darco Hg<br />

Leland Olds, lignite,Darco Hg<br />

Monroe,PRB,SCR <strong>by</strong>pass, Darco Hg<br />

Stanton unit 1, PRB, B-PAC<br />

Meramec, PRB, Darco Hg-LH<br />

Figure 3.6. <strong>Mercury</strong> removal as a function of injection rate of Darco Hg sorbent in power<br />

plant <strong>by</strong> sorbent injection upstream of a cold side ESP. Data are <strong>from</strong> [91,93,96,100,101].<br />

Tests at Pleasant Prairie showed that there was no significant effect on mercury<br />

removal with PAC injection when SO3 was used as flue gas conditioning agent to obtain<br />

optimal dust resistivity and improve ESP performance [96,101]. The level of applied SO3<br />

at Pleasant Prairie was not reported. However, tests at Labadie unit 2 showed that the<br />

presence of SO3 in the flue gas can decrease mercury capture <strong>by</strong> activated carbon [105].<br />

The applied SO3 concentration in the flue gas at Labadie unit 2 was about 5-10 ppmv.<br />

This is probably due to the competitive adsorption between Hg and SO3 since both<br />

mercury and SO3 bind to the Lewis acid base sites on the activated carbon surface<br />

[21,22].<br />

In some plants burning PRB coals, it was observed that when the carbon injection<br />

rate was increased above 160 mg/m 3 the mercury removal efficiency <strong>by</strong> the cold side<br />

ESP leveled off at about 60% [91,93,96,100,101]. At Brayton Point plant bituminous<br />

coal was fired and the mercury removal increased with carbon injection rates in all the<br />

tested ranges up to 320 mg/m 3 reaching 90% mercury removal [102]. This is probably<br />

due to the fact that at the Brayton Point the predominant species of mercury is in the


oxidized form since there is a significant amount of HCl present in the flue gas <strong>from</strong><br />

Brayton [102], in contrast to Pleasant Prairie where the majority of vapor phase mercury<br />

was in the elemental form.<br />

3.3.4 <strong>Sorbent</strong> injection tests at cement plant<br />

There are very limited studies on mercury removal <strong>by</strong> sorbent injection in cement<br />

plants. In 2007 a six-week test was conducted at Ash Grove <strong>Cement</strong> Company’s Durkee<br />

plant using a slipstream fabric filter after the main bag filter [106,107].<br />

The overall goal of the tests at Durkee was to perform a parametric test on a<br />

slipstream of actual flue gas to obtain an understanding of how various operating and<br />

design parameters are likely to impact mercury control in the Durkee plant. The<br />

evaluated parameters included activated carbon type, filter bag type, powdered activated<br />

carbon injection rate, and filter air-to-cloth ratio.<br />

The slipstream filter had twelve �152 mm�3658 mm bags, corresponding to a<br />

filtration area of 21 m 2 . The filter chamber and inlet duct were insulated and heated to<br />

maintain a temperature of about 138°C. <strong>Mercury</strong> concentrations at the filter inlet and<br />

outlet were measured <strong>by</strong> a Horiba/Nippon Instruments Corporation DM-6B, as well as<br />

the Ontario hydro method. The tested carbons include Darco Hg, Darco Hg e-11, Darco<br />

Hg LH, and Envergex e-sorb e11. The last two carbons are chemically treated. The<br />

Darco Hg is prepared <strong>from</strong> lignite coal and the Darco Hg e-11 is a coarser version of the<br />

Darco Hg. The particle size of the Darco Hg e-11 carbon is not reported. The test<br />

duration for each parametric study was only about one hour. The flue gas compositions<br />

are not publically reported and baghouse cleaning cycle is unknown.<br />

Figure 3.7 shows the mercury removal efficiency as a function of PAC injection<br />

rate for different carbons. For the Darco Hg and Darco Hg e-11, the mercury removal<br />

efficiency increases only slightly <strong>from</strong> 80-90% to 90-95% when the PAC injection rate is<br />

further increased above 48 mg/m 3 . The mercury removal efficiencies <strong>by</strong> the untreated<br />

carbons are generally larger than those <strong>by</strong> the treated carbons when low injection rates<br />

are applied. Treated carbons (Darco Hg LH and Envergex) have been shown to perform<br />

54


etter than untreated carbons in coal-fired boilers, especially in systems with ESP where<br />

reaction times are short [108]. The halogens in the carbon act to oxidize the Hg in the<br />

system and allow faster adsorption onto the carbon. This is critical in systems with higher<br />

SOx concentrations, because SOx species have been shown to compete for active sites on<br />

the carbon surface [21,22], as discussed earlier. The halogens on the treated carbon allow<br />

the oxidized Hg to bind to the carbon surface before the SOx species consume the active<br />

sites [108]. At the Durkee plant, the SOx concentrations in the slipstream baghouse are<br />

very low compared to a coal-fired utility system. Thus the promoting effects of halogen<br />

treated carbon are less pronounced.<br />

<strong>Mercury</strong> removal efficiency, %<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 20 40 60 80 100<br />

PAC injection rate, mg/m 3<br />

55<br />

DARCO Hg<br />

DARCO Hg e-11<br />

Envergex<br />

DARCO Hg LH<br />

Figure 3.7. <strong>Mercury</strong> removal efficiency as a function of PAC injection rate for different<br />

sorbents at 138°C. The applied bag material is polyphenylene sulphide (PPS) and the airto<br />

cloth ratio is 1.22 m/min. Data are <strong>from</strong> [106].<br />

At rates higher than 80 mg/m 3 , the untreated carbons appear to perform similarly<br />

as the treated carbons with a mercury removal efficiency of about 90-95%. However,<br />

injection of the treated carbons at 80 mg/m 3 does not result in a significant increase in the<br />

mercury control efficiency as compared to untreated carbon injected at 48 mg/m 3 . This<br />

shows that there is no reason to choose halogenated carbon over untreated carbon,


particularly in light of the higher price and potential concerns associated with the use and<br />

disposal of halogen-treated materials.<br />

The trend of mercury removal efficiency of Darco Hg e-11 is similar to that of<br />

finer Darco Hg, but the Hg removal results were lower <strong>by</strong> 10%–15%. This was most<br />

likely caused <strong>by</strong> the larger particle sizes leading to more severe diffusion limitation.<br />

Three bag types were tested, namely, polyphenylene sulphide (PPS), membrane<br />

and fiberglass with membrane, while other conditions of the baghouse are the same. The<br />

primary aim of testing different bag types is to investigate whether retention of carbon<br />

particles on the bag surface can enhance the mercury removal efficiency. The<br />

comparison of the performance of the tree bag types is presented in figure 3.8. At PAC<br />

injection rate above 48 mg/m 3 of Darco Hg all three bag types perform quite similarly at<br />

138°C. Considering the uncertainty caused <strong>by</strong> the variability between the inlet and the<br />

outlet mercury measurements, it is likely that the bags are performing essentially the<br />

same at these conditions [108]. Then the only controlling factor for choosing the bag<br />

type is the working temperature. The flue gas temperature in the bag filter area of the<br />

cement process varies a lot and can exceed 200°C. Among the tested bag types, only the<br />

membrane/fiberglass bag can withstand continuous operating temperatures at 260°C and<br />

is therefore recommended.<br />

56


<strong>Mercury</strong> removal efficiency, %<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 20 40 60 80 100<br />

PAC injection rate, mg/m 3<br />

57<br />

PPS<br />

membrane/fiberglass<br />

Membrane<br />

Figure 3.8. Effects of bag material on mercury removal efficiency at 138°C. The applied<br />

air-to-cloth ratio is 1.22 m/min and the sorbent is Darco Hg. Data are <strong>from</strong> [106].<br />

The effect of air-to-cloth ratio on mercury removal was tested using the<br />

membrane/fiberglass bag. It is reported that at a PAC injection rate of 16 mg/m 3 the<br />

mercury removal efficiency increased with increasing the air-to-cloth ratio in the range of<br />

1.2-3.0 m/min and when the injection rates were higher than 48 mg/m 3 the mercury<br />

removal efficiency increases only slightly with further increasing the injection rate since<br />

the mercury removal efficiency is higher than 90% [106]. The increase of mercury<br />

removal efficiency with filtration velocity might be due to the fast accumulation of<br />

carbon on the bag surface. However, care must be taken when discussing the observation.<br />

The cleaning control of the bags was not specified. It is unknown whether the bags were<br />

cleaned at a fixed time interval or defined pressure drop over the filter. Most of the tests<br />

were conducted for a period of only about 20 min, while only several tests were run for<br />

up to 1-2 h.<br />

<strong>Injection</strong> of Darco Hg before the fabric filter with membrane/fiberglass was also<br />

tested in the raw-mill-off operating period. The mercury concentration at the filter inlet<br />

during the raw-mill-on period was about 485 µg/Nm 3 , but increased to about 2600


µg/Nm 3 during raw-mill-off operating period. Using an air-to-cloth ratio of 2.4 m/min,<br />

moderate mercury removal efficiencies of 52% and 58% were obtained at PAC injection<br />

rates of 48-80 mg/m 3 , respectively. A mercury removal efficiency of 88% was achieved<br />

when the PAC injection rate was increased to 160 mg/m 3 .<br />

Based on the parametric study at 138�C, design parameters for the full-scale<br />

sorbent injection upstream of a polishing filter at Durkee cement plant were<br />

recommended. The untreated carbon, fiberglass with membrane bag type, and air-tocloth<br />

ratio of 1.8-2.4 m/min were suggested. The proposed sorbent injection rate is 48<br />

mg/m 3 and 80 mg/m 3 for the raw-mill-on and raw-mill-off operation period, respectively.<br />

The estimated mercury removal efficiency is 90% during raw-mill-on conditions and<br />

60% during raw-mill-off conditions. The weighted mercury removal efficiency expected<br />

is about 77% on annual average.<br />

3.3.5 Carbon surface chemistry and mechanisms of mercury capture on<br />

carbons<br />

3.3.5.1 Carbon surface chemistry<br />

The surface chemistry of carbons determines their moisture content, catalytic<br />

properties, acid-base character, and adsorption of polar species. It is related to the<br />

presence of heteroatoms other than carbon within the carbon matrix. The most common<br />

heteroatoms are oxygen, nitrogen, phosphor, hydrogen, chlorine, and sulphur [109].<br />

During preparation of carbon and particularly during cooling and storage, carbon<br />

materials are in contact with the ambient air so that elements such as H and O are fixed<br />

on the surface, leading to oxygenated chemical functional groups [110]. Several<br />

structures of oxygen functional groups have been proposed as shown in table 3.3.<br />

Functional groups can be acidic, basic, or neutral in character. Surface oxygen groups on<br />

carbon materials decompose upon heating <strong>by</strong> releasing CO and CO2 at different<br />

temperatures. A CO2 peak results <strong>from</strong> carboxylic acids at low temperatures, or lactones<br />

at higher temperatures; carboxylic anhydrides originate both a CO and a CO2 peak;<br />

phenols, ethers, carbonyls, and quinones originate a CO peak.<br />

58


Table 8. Surface oxygen groups on carbon and their decomposition <strong>by</strong> TPD, after<br />

[110,111].<br />

Group name Decomposition Decomposition<br />

product temperature (�C)<br />

Carboxyl CO2 100-400<br />

Lactone CO2 190-650<br />

Carboxylic anhydrides CO+CO2 350-627<br />

Phenolic CO 600-700<br />

Ether CO 700<br />

Carbonyl CO 700-980<br />

Quinone CO 700-980<br />

Besides oxygenated functions, nitrogenated functions can be introduced on<br />

carbon surface <strong>by</strong> reaction of a carbon with a nitrogen-containing reactant or preparation<br />

of a carbon <strong>from</strong> a nitrogen-containing precursor [110].<br />

3.3.5.2 Mechanisms of mercury capture on carbons<br />

In order to understand the mercury capture mechanisms, it is important to<br />

understand the chemical and physical nature of the mercury-sorbent interaction. X-ray<br />

absorption fine structure (XAFS) spectroscopy and X-ray photoelectron spectroscopy<br />

(XPS) are techniques that have been previously used to determine information about the<br />

speciation and binding of mercury on a variety of materials [112,113]. XAFS spectra can<br />

be defined <strong>by</strong> two regions which include X-ray absorption near-edge spectroscopy<br />

(XANES) and extended X-ray absorption fine structure (EXAFS) spectroscopy. XANES<br />

spectra provide information on the oxidation state and characteristics of the first neighbor<br />

coordination environment. EXAFS spectroscopy provides more robust information on<br />

the identity of nearest-neighboring elements, coordination values, and interatomic bond<br />

distances.<br />

XAFS spectroscopy was used to distinguish between elemental and oxidized<br />

mercury in the sorbents <strong>by</strong> comparing the XAFS spectrum. Elemental mercury exhibits a<br />

59


single peak only in the first-derivative of the mercury XANES spectrum, whereas most<br />

mercuric compounds exhibit a two-peak spectrum [112]. The sorbents were tested for<br />

mercury capture at temperatures lower than 200°C. The studied sorbents included<br />

carbonaceous materials and inorganic-based material, such as lime-derived sorbents and<br />

zeolites.<br />

The XANES data imply that the capture of elemental mercury must involve an<br />

oxidation process, either in the gas phase before interacting with the sorbent, or<br />

simultaneously as the Hg 0 atom interacts with the sorbent [112]. This is consistent with<br />

the fact that all Hg-sorbed materials examined exhibit the characteristic dual inflection<br />

point structure in their XANES spectra that is indicative of the formation of Hg–anion<br />

chemical bonds. The anion could be virtually any available electronegative species, as<br />

evidence has been seen for the formation of Hg–I, Hg–Cl, Hg–S, Hg–O, and Hg-Br<br />

[112,113]. Modeling of mercury capture <strong>by</strong> activated carbon using density functional<br />

theory shows that the mercury binding energies increase with the addition of the<br />

following halogen atoms, F>Cl>Br>I [114]. Data <strong>from</strong> S and Cl XANES spectra, as well<br />

as <strong>from</strong> the Hg XAFS data, strongly support the hypothesis that interaction of acidic<br />

species (HCl, HNO3, H2SO4, HBr, etc.) in the flue gas with the sorbent surface is an<br />

important mechanistic process that is responsible for creation of active sites for mercury<br />

capture <strong>by</strong> chemisorption. The mechanisms of elemental mercury capture on the carbon<br />

sorbents likely consist of surface-enhanced oxidation of the elemental mercury via<br />

interaction with surface-bound halide species with subsequent binding <strong>by</strong> surface halide<br />

or sulphate species [113].<br />

The catalytic effects of carbon sorbents for mercury capture were investigated <strong>by</strong><br />

Olson et al. [115]. The studied carbons were lignite- and bituminous-derived carbon and<br />

catalytic carbon, which were available commercially with enhanced catalytic<br />

functionality for aqueous reactions such as decomposition of peroxides. Catalytic<br />

carbons are produced <strong>by</strong> recarbonization of urea or ammonia-treated oxidized activated<br />

carbons or <strong>by</strong> impregnation of nitrogen-containing polymers and pitches [115]. Without<br />

acid gases in the gas stream at 150�C, 50% mercury breakthrough was observed after 8<br />

60


min for the catalytic carbon, while less than 1 min for the lignite- and bituminous-derived<br />

carbons. Thus, a catalytic chemisorption mechanism predominates for the sorption of<br />

mercury at these conditions.<br />

The mercury adsorption capacity of the sorbent is inversely proportional to the<br />

temperatures in a studied range of 50-150°C, indicating that a preliminary physisorption<br />

step with mercury associating with a surface site takes place [115]. The chemisorption of<br />

Hg 0 is likely a multistep reaction. When the temperature is increased, the rate of each<br />

chemical reaction step increases and the exothermic physisorption of Hg 0 at nonoxidizing<br />

binding sites will decrease. If the sorption process includes a preliminary<br />

physisorption equilibration where Hg 0 binds and desorbs at the active site, the<br />

equilibration will show a negative temperature effect on the overall reaction rate, since<br />

desorption is favored at higher temperatures. Although chemisorption may account for<br />

the main sorption of mercury, the extent to which increasing the temperature may affect<br />

the sorption rate cannot be predicted.<br />

A detailed mechanism has been proposed to explain the effects of SO2 and NO2<br />

as shown in figure 3.9 [116]. In the presence of NO2, Hg 0 is catalytically oxidized on the<br />

carbon surface to form the nonvolatile nitrate Hg(NO3)2, which is bound to basic sites on<br />

the carbon. The Lewis base site refers to the zigzag carbon atom positioned between<br />

aromatic rings [117]. Capture continues until the binding sites are used up and<br />

breakthrough occurs. In the presence of SO2, some of the catalytic sites are converted to<br />

a sulfate form where Hg(NO3)2 is no longer formed. <strong>Mercury</strong> is still oxidized on the<br />

surface with NO2 acting as the oxidizing agent, but the product formed is a labile sulfur<br />

compound, mercury bisulfate [Hg(SO4H)2]. The bisulfate in turn reacts with NO3 - to<br />

form a stable but volatile acidic form of the mercuric nitrate. The emission of Hg(NO3)2<br />

or the hydrate Hg(NO3)2�H2O has been confirmed <strong>by</strong> solvent trapping and gas<br />

chromatography analysis. Sulfurous acid that accumulates <strong>from</strong> the hydration of SO2<br />

converts the previously formed nonvolatile basic mercuric nitrate into the volatile form,<br />

which explains the slow release of previously captured mercury over time in the presence<br />

of NO2 and SO2.<br />

61


Figure 3.9. Proposed heterogeneous model for mercury capture on carbon showing<br />

potential impact of acid gases [116].<br />

Sulphur trioxide can be present in power plant flue gas through one of the<br />

following paths [21]: (1) During combustion, coal-S is converted to SO2 and a small<br />

fraction of the sulfur is further oxidized to SO3. During combustion of high-sulfur coals,<br />

a minor part of the sulfur is converted to SO3, leading to flue gas concentrations in the<br />

range of 1-40 ppm. (2) SO3 is sometimes added to a level above 10 ppm to flue gas<br />

upstream of an ESP as a conditioning agent and to improve ESP performance. SO3 and<br />

H2SO4 have a low vapor pressure and can condense on fly ash and this reduces the<br />

resistivity of the ash and allows it to be removed more efficiently <strong>by</strong> the ESP. (3) SO2<br />

can be oxidized to SO3 <strong>by</strong> SCR catalysts installed for NOx reduction [118-120]. SCR<br />

catalysts typically contain vanadium oxides, which are known catalysts for the oxidation<br />

of SO2 to SO3 and Hg to HgCl2.<br />

The inhibiting effect of SO3 on mercury capture <strong>by</strong> activated carbon injection has<br />

been observed in full-scale power plant tests [21,121]. Possible mechanisms for the SO3<br />

effect on mercury capture <strong>by</strong> activated carbon are postulated <strong>by</strong> Presto and Granite<br />

[21,22]. In addition to removing mercury, activated carbon is also used as catalyst for<br />

oxidation of SO2 to sulphuric acid [122,123]and as SO2 sorbent. There is competitive<br />

adsorption between Hg and SO3 since both mercury and SO3 bind to the Lewis acid base<br />

62


sites on the activated carbon surface. The adsorption of SO3 could be favored both<br />

kinetically and thermodynamically. The concentration of SO3 in flue gas is typically in<br />

the range of 1-40 ppm and this is orders of magnitude larger than typical mercury<br />

concentrations. The bond formed between the S 6+ species, such as sulfuric acid and<br />

sulfates, and the carbon surface is stronger than the bond between mercury and the<br />

surface. SO2 can oxidize to sulphate and form a chemical bond with the carbon surface<br />

with a heat of adsorption of >80 kJ/mol. Some activated carbon catalysts for converting<br />

SO2 to H2SO4 are self-poisoned <strong>by</strong> SO3 or sulfate buildup on the surface. A similar<br />

phenomenon might explain the inhibiting effect of SO3 on mercury capture.<br />

3.3.6 Processing and reuse of mercury laden activated carbon<br />

The existing production capacity for powdered activated carbon is only 10% of<br />

the capacity required for full implementation of the activated carbon injection technology<br />

to control mercury emissions [124]. The mercury sorption capacity of the activated<br />

carbon is very low, about 1-4 mg of mercury per gram of sorbent, depending on the<br />

mercury concentration in the flue gas [125]. This implies that 250 to 1000 g of activated<br />

carbon are needed to remove 1 g of mercury in the flue gas. Therefore, a large quantity of<br />

spent sorbents contaminated with various forms of mercury is produced.<br />

Presently the PAC with adsorbed mercury must be disposed after use. In addition<br />

to the purchase expense, the disposal of this material is also quite costly. There are strict<br />

regulations for disposal of mercury-containing wastes [126]. Hazardous wastes<br />

containing less than 260 mg/kg of total mercury are required to be treated to 0.20 mg/L,<br />

measured using the toxicity characteristic leaching procedure (TCLP) for mercury<br />

residues <strong>from</strong> retorting, and 0.025 mg/L TCLP for all other low mercury wastes. Wastes<br />

that contain greater than 260 mg/kg total mercury are required to undergo roasting or<br />

retorting in a thermal processing unit capable of volatilizing mercury and subsequently<br />

condensing the volatilized mercury for recovery.<br />

To reduce the PAC purchase expense and disposal cost of mercury-containing<br />

PAC, a process has been developed to regenerate the used PAC and recover mercury<br />

[124]. To separate PAC <strong>from</strong> the fly ash, PAC is injected between the main filter and the<br />

63


polishing FF. The collected PAC is periodically removed <strong>from</strong> the filter and regenerated<br />

in nitrogen process gas and is directed to a multiple activated carbon column gas<br />

treatment system to remove the gaseous mercury <strong>from</strong> the cooled process gas stream.<br />

After passing through the sulphur impregnated carbon columns, the carrier gas is injected<br />

into the flue gas stream ahead of the carbon injection site. In this way only a small<br />

amount of carbon with high mercury content requires disposal.<br />

An inert atmosphere is required for the tray desorption furnace to avoid<br />

significant losses of the PAC material during mercury desorption. Using a desorption<br />

temperature of 550°C and a duration of 30 minutes, the PAC can be recycled at least 10<br />

times without significant degradation of the adsorption characteristics in nitrogen [124].<br />

It is unknown whether the cycled sorbent works satisfactorily in the real flue gas.<br />

There are only few studies on the mercury desorption <strong>from</strong> exposed sorbents. It is<br />

worth noting that mercury desorption is relevant both to recover the mercury and to<br />

detoxify the adsorbing material in order to avoid its stabilization before land-filling or to<br />

allow its reuse.<br />

A study of mercury desorption in nitrogen <strong>from</strong> sulphur impregnated activated<br />

carbon showed that the adsorption rate was faster than the desorption rate [59]. <strong>Mercury</strong><br />

desorption <strong>from</strong> sorbents is strongly affected <strong>by</strong> desorption temperature, with faster<br />

desorption at high temperature and the mercury-sorbent pair. The desorption rate is<br />

relatively fast initially and then levels off close to zero at a certain concentration of<br />

mercury in sorbents.<br />

Desorption of mercury <strong>from</strong> activated carbon and fly ash mixture was also carried<br />

out in a fluidized bed reactor at temperatures up to 500°C [127]. All the mixtures had<br />

constant mercury content, i.e., no mercury desorption was observed, until a critical<br />

temperature was reached and then with rapidly decreasing mercury content as the<br />

temperature was increased to higher levels. The critical temperature was found to be a<br />

linear function of carbon contents in the mixtures, increasing <strong>from</strong> 330°C at 17% carbon<br />

to 370°C at 33% carbon. The temperature at which all of the mercury was removed was<br />

in the 450 to 500°C range.<br />

64


3.3.7 Applicability of sorbent injection in cement plants<br />

As PAC systems are adapted for control of boilers, it will be possible to evaluate<br />

the feasibility of these control techniques for cement kiln applications having<br />

approximately the same mercury concentrations. Considering the differences between<br />

boiler and kiln applications the possible application of PAC systems to cement kilns<br />

appears to be considerably more challenging than to coal-fired boilers.<br />

Powdered activated carbon injection systems do not appear to be appropriate<br />

upstream of a cement kiln fabric filter system. <strong>Cement</strong> kilns must recycle a major portion<br />

of the collected dust. Some kilns use the fabric filter system as an integral part of the raw<br />

material processing system. Recycling the mercury laden activated carbon would result<br />

in the revolatilization of the large majority of the mercury. Disposal of the activated<br />

carbon containing cement kiln dust (CKD) also would be complicated because it might<br />

be classified as a hazardous waste due to the presence of mercury.<br />

Due to these issues, a powdered activated carbon injection system would have to<br />

be installed downstream of the main kiln fabric filter to avoid the CKD recycling and<br />

disposal issues. A second fabric filter would have to be installed after the main fabric<br />

filter. The activated carbon injection system would have to be positioned to provide one<br />

to two seconds residence time prior to entering the second fabric filter. The temperature<br />

of this system would have to be controlled to less than 200�C to ensure proper mercury<br />

adsorption and reduce the risk of activated carbon fires in the fabric filter or solids<br />

handling system.<br />

3.4 <strong>Mercury</strong> removal <strong>by</strong> activated carbon bed<br />

Fixed and moving bed systems for mercury and dioxin-furan control are also used<br />

in Europe [5]. In both types of systems, contaminant-laden gas is forced through a bed of<br />

granular activated carbon.<br />

One of the fixed bed systems used Sorbalit sorbent instead of activated carbon.<br />

The Sorbalit sorbent consists of Portland cement, lime, carbon, and sulfur compounds<br />

such as sublimed sulfur, Na2S, NaHS, and Na2S4 [16].<br />

65


The quantity of mercury that can be retained on the sorbent at equilibrium is<br />

important. The sorbent can be used at levels that approach the saturation capacity of the<br />

sorbent at the operating gas temperature and gas stream conditions. However, the control<br />

system must have the capability to remove the sorbent on at least a semi-continuous basis.<br />

In fixed bed systems, the activated carbon must be replaced with fresh carbon at a<br />

rate that is dependent primarily on the rate of approach to the mercury saturation level,<br />

and the rate of static pressure increase. Spent carbon can be disposed of <strong>by</strong> combustion if<br />

the unit is equipped with a wet scrubbing system. The combustion process destroys the<br />

organic compounds captured in the carbon, and the wet scrubber collects the heavy<br />

metals and acid gases. In this case, however, the elemental mercury might not be<br />

removed due the insolubility of elemental mercury in the water. Another disposal option<br />

is to dispose the carbon in a landfill. Because of the adsorbed pollutants, this waste may<br />

require disposal as a hazardous waste. Another option is to heat the carbon and desorb<br />

the pollutants <strong>from</strong> the carbon.<br />

Slipstream tests of the activated carbon bed have been recently conducted in<br />

several U.S. power plants [128]. Direct adaptation of existing carbon bed technology to<br />

mercury removal <strong>from</strong> utility power plant flue gas is very costly because of the large flue<br />

gas volumes and low mercury concentrations involved [129]. A thorough engineering<br />

and economic analysis would be necessary to determine the feasibility of modifications<br />

that reduce bed size and the amount of carbon in the bed. The effectiveness of the<br />

modified beds for mercury removal under various flue gas conditions needs to be<br />

determined. Furthermore, the tradeoff between gas velocity to the bed, bed sorbent size<br />

and bed thickness, pressure drop, mercury and ash collection effectiveness, and bed<br />

lifetime should be examined.<br />

For cement plant application, the fixed bed activated carbon systems could not be<br />

installed upstream of the main kiln bag filter or ESP. The high dust loadings in these<br />

locations would quickly blind both types of beds and result in very high activated carbon<br />

usage rates and disposal requirements. Accordingly, it would be necessary to install these<br />

systems downstream of the main particulate matter control system. Similar to the power<br />

66


plant application, installing a fixed bed carbon system in a cement plant will also be very<br />

costly.<br />

3.5 <strong>Mercury</strong> control <strong>by</strong> flue gas desulphurization systems<br />

Dry and wet scrubbers, commonly used in large scale combustion systems for<br />

SO2 and HCl control can be simultaneously used for mercury retention, taking advantage<br />

of the same sorbents used for sulphur or adding a new material for mercury [130-133].<br />

Wet scrubbing systems predominately collect oxidized mercury [5,131]. In the purge<br />

stream, mercuric chloride is collected as a precipitated solid along with the calcium<br />

sulfate. When used as stand alone systems, they have the capability to achieve moderateto-high<br />

removal efficiencies for oxidized mercury. They are entirely ineffective in the<br />

removal of the highly insoluble elemental mercury.<br />

There are limited number of lime-based scrubbing systems used primarily for<br />

particulate and SO2 control at lime kilns. There are presently only few cement kilns in the<br />

United States equipped with wet SO2 scrubbing systems [5].<br />

The stability of oxidized mercury captured in the flue gas desulphurization (FGD)<br />

systems has been investigated and it was found that the captured oxidized mercury can be<br />

reduced <strong>by</strong> aqueous phase reactions to form elemental mercury [6]. The insoluble<br />

elemental mercury is rapidly released to the gas stream. Occurrence of mercury in FGDgypsum<br />

may threaten its re-use for wallboards since mercury can be released during the<br />

heating steps in wallboard manufacturing [131].<br />

Spray dryer absorbers (SDA) with a Ca(OH)2 slurry have been used for sulfur<br />

dioxide and hydrogen chloride control at waste incinerators and coal-fired boilers. SDA<br />

has been applied recently to cement kilns to control HCl that contributes to secondary<br />

plume formation [5].<br />

SDA systems would have to be installed after the main particulate matter control<br />

system to ensure that captured mercury remains with a solid waste product and is not<br />

recycled to the feed end of the kiln. With respect to fossil fuel fired boilers, the reported<br />

mercury removal efficiencies <strong>by</strong> SDA systems are in the range of 50% to 60% for eastern<br />

67


ituminous coals and 0% to 20% for western lignite and subbituminous coals [5]. The<br />

difference is caused <strong>by</strong> the lower fraction of oxidized mercury for the lignite coals. With<br />

respect to cement kilns, it appears unlikely that SDA systems will be more effective than<br />

inherent adsorption in cement kiln systems.<br />

3.6 <strong>Mercury</strong> removal <strong>by</strong> sodium tetrasulfide injection<br />

Sodium tetrasulfide (Na2S4) has been used as a sorbent to remove mercury <strong>from</strong><br />

flue gas in a number of waste-to-energy plants [1]. This technology should not be<br />

confused with sodium sulfide Na2S that was tried in both Europe and U.S. without<br />

success [134]. The shortcomings of Na2S are that it can leave a strong odor of hydrogen<br />

sulfide (H2S) in the ash and it does not control all species of Hg. The major advantages of<br />

the Na2S4 technology are that it controls elemental as well as ionic forms of Hg.<br />

An aqueous Na2S4 solution is injected into the flue gas duct and such a system<br />

can easily be retrofitted to an existing flue gas cleaning plant [134]. The sodium<br />

tetrasulfide reacts with vapor phase mercury to form solid mercuric sulfide (HgS), which<br />

is a solid at temperatures below about 580°C, and is insoluble [134]. By converting<br />

vapor-phase mercury to an insoluble solid, it may be removed in a FF or ESP. Sodium<br />

tetrasulfide can react with both oxidized and elemental mercury in accordance with the<br />

following simplified reactions [134]:<br />

Na2S4 �HgCl2 � HgS �2NaCl � 3S<br />

(R3.4)<br />

Hg�S � HgS<br />

(R3.5)<br />

Decomposition of Na2S4 <strong>by</strong> an acid such as HCl can provide excess elemental<br />

sulfur. It can also generate an alternate form of ionic sulfur, H2S, for reaction with<br />

oxidized mercury as shown in the following reactions:<br />

Na S �2HCl �HS�3S� 2NaCl<br />

(R3.6)<br />

2 4 2<br />

HgCl2 �H2S�HgS � 2HCl<br />

(R3.7)<br />

In the absence of HCl, carbon dioxide may act as an acid for decomposition:<br />

Na S �2CO�2HO�HS�3S� 2NaHCO<br />

(R3.8)<br />

2 4 2 2 2 3<br />

68


Therefore, it is possible to eliminate both the elemental and ionic forms of mercury in the<br />

flue gas.<br />

However, H2S will still be produced in the process as shown in R3.6 and R3.8.<br />

The problem of H2S odor in the ash cannot be avoided. This process is not suitable to<br />

mercury removal <strong>from</strong> cement plant due to the presence of sodium, which could<br />

deteriorate the cement quality.<br />

3.7 Enhanced mercury removal <strong>by</strong> oxidation<br />

Oxidation pretreatment systems may convert elemental mercury to oxidized<br />

mercury upstream of wet scrubber systems and even upstream of conventional particulate<br />

matter control systems. Once in the oxidized form, mercury is captured in these air<br />

pollution control systems at efficiencies approaching 85% [5]. The oxidation<br />

pretreatment systems must be able to withstand the gas stream conditions upstream of the<br />

air pollution control system used for capture of the oxidized mercury. Oxidation<br />

pretreatment systems are only effective for the vapor phase mercury that is not adsorbed<br />

on particle surfaces.<br />

Selective non-catalytic reduction (SNCR) and selective catalytic reduction (SCR)<br />

systems are used in coal-fired boilers and waste incinerators for the control of nitrogen<br />

oxides. The impact of SNCR systems on the chemical form of mercury in a gas stream<br />

appears to be minimal [5].<br />

SCR systems use a catalyst to react ammonia and nitrogen oxides to provide<br />

nitrogen and water. SCR systems are used extensively for coal-fired boilers and waste<br />

incinerators. Full-scale tests have been performed in four U.S. power plants and the<br />

results are presented in table 3.4 [135,136]. Significant oxidation of elemental mercury<br />

across the SCR was observed in plant 2 and 4. While slight mercury oxidation over SCR<br />

was experienced in pant 1 and 3. General conclusions <strong>from</strong> these tests are the oxidation<br />

effect was quite variable and appears to be coal-specific and possibly catalyst-specific. In<br />

particular, the catalyst type, space velocity, and catalyst age may all be important<br />

variables. Plant 1 burns PRB coal with lower chlorine content and the catalyst is older.<br />

69


More than 90% of the mercury in the flue gas at plant 1 is elemental mercury at the SCR<br />

inlet. One possible explanation for the relatively low oxidation rate of the SCR at plant 3<br />

is the relatively high space velocity, which is nearly double the space velocity compared<br />

to other plants. In addition, the total inlet mercury concentration was more than twice the<br />

levels seen at the other test sites.<br />

Table 3.4. Test conditions and results of mercury oxidation over SCR catalyst in four U.S.<br />

power plants. Data are <strong>from</strong> [135,136].<br />

Plant 1 2 3 4<br />

Coal PRB<br />

subbituminous<br />

Ohio<br />

bituminous<br />

high-sulfur<br />

70<br />

Pennsylvania<br />

bituminous<br />

low-to-medium<br />

sulfur<br />

Kentucky<br />

bituminous<br />

medium-sulfur<br />

Hg in coal (ppmm) 87 168 400 131<br />

Cl in coal (ppmm)


ange (300-350°C) in a real flue-gas atmosphere [139], while NH3 shows a small<br />

detrimental effect. Adding 2000 ppmv SO2 to baseline gases that contain 6% O2, 12%<br />

CO2, 8% H2O, 550 ppmv NH3, 600 ppmv NO, 18.5 ppmv NO2 without HCl only<br />

increases the mercury oxidation over SCR <strong>from</strong> 3% using baseline gases to 7% with<br />

2000 ppmv SO2 [139,142].<br />

Adding 50 ppmv SO3 to the baseline gases improves the mercury oxidation to<br />

20% [139,142]. Adding 50 ppmv HCl to the baseline gases without SO2/SO3 results in<br />

71% oxidized mercury in flue gas across the SCR compared to 45% mercury oxidation<br />

when 50 ppmv SO3 was further added to the flue gas. With 50 ppmv HCl and 2000 ppmv<br />

SO2 were added to the baseline gases, mercury oxidation recovered to 64%. The<br />

combination of 2000 ppmv SO2/50 ppmv SO3 and 50 ppmv HCl showed a 63% mercury<br />

oxidation. These observations indicate that both SO2 and SO3 had a negative effect on<br />

mercury–chlorine oxidation over the SCR as a result of slower mercury oxidation <strong>by</strong> the<br />

sulfated site compared to that of the chlorinated site. The extent of the mitigating effect<br />

<strong>by</strong> the 2000 ppmv SO2 was not as severe as the 50 ppmv SO3 since the concentration of<br />

SO3 derived through SO2 oxidation over SCR was much lower than the 50 ppmv SO3.<br />

There is possible competition between HCl, SO2, and SO3 over the SCR catalyst.<br />

Conventional SCR catalyst with higher mercury oxidation capacity has been<br />

closely related to higher oxidation of SO2 to SO3 [144]. Higher SO2 oxidation in coalfired<br />

applications can cause negative downstream impacts such as air heater fouling, flue<br />

duct corrosion and visible stack plumes. The SO2 to SO3 conversion is designed to be less<br />

than 1.5% at SCR operating conditions. Research has been focused on developing new<br />

SCR catalyst that has higher oxidation rate of elemental mercury and very low SO2 to<br />

SO3 conversion [144].<br />

Two studies have demonstrated that ultraviolet radiation in the presence of solid<br />

titanium dioxide (TiO2) results in the photocatalytic conversion of elemental mercury to<br />

HgO when HCl is not present in the gas [6,145,146]. The TiO2 is readily available as a<br />

major component of conventional SCR catalysts. Accordingly, the combination of<br />

ultraviolet light and SCR catalysts could in principle be used to oxidize elemental<br />

71


mercury. Compared to lab-scale study where ultraviolet light can be readily applied,<br />

application of ultraviolet light in monolith catalyst could be a technical challenge.<br />

The application of SCR systems to cement kilns continues to be precluded <strong>by</strong><br />

problems associated with alkali metal and arsenic related catalyst poisoning, SO2<br />

oxidation to SO3, particulate matter loadings that can be 5 to 20 times higher than coalfired<br />

boiler high dust systems, and non-ideal temperature ranges for SCR catalysts<br />

[147,148]. For these reasons, it is unlikely that an SCR system will be used for the<br />

control of nitrogen oxides and oxidation of elemental mercury in cement plants in the<br />

near future.<br />

3.8 <strong>Mercury</strong> removal <strong>by</strong> roaster process<br />

Recently a new process for mercury removal <strong>from</strong> the cement kiln flue gas <strong>by</strong> a<br />

roaster was invented and patented [149-151]. As mentioned earlier recycled cement kiln<br />

dust <strong>from</strong> the main bag filter has high mercury content. The mercury flow in the<br />

collected cement kiln dust is about 60% of the mercury inlet to the cement kiln [152].<br />

This indicates that dust captured in the main baghouse acts as a natural sorbent for<br />

mercury. This mercury enriched dust is taken to the new mercury roaster process for<br />

cleaning before the dust is returned to the system. Figure 3.10 illustrates an example of a<br />

mercury roaster installation. The baghouse dust is fed to a roasting system which uses a<br />

heat source (for example kiln <strong>by</strong>pass gas, cooler vent gas, or hot gas generator) to heat<br />

the dust above the boiling point of mercury compounds. While the mercury is still in the<br />

gas phase, the gas stream enters a hot ESP which removes most of the cleaned dust. This<br />

dust is taken back to the blending silo to be part of the kiln feed. After the ESP, the gas<br />

stream is cooled below the mercury boiling point so that the mercury can condense on<br />

the dust particles that were not captured in the ESP and additional sorbent is added to the<br />

gas stream here to capture the mercury. The cleaned gas after the baghouse is vented to<br />

the atmosphere. Depending on the type of applied sorbents the mercury enriched<br />

dust/sorbent collected in the baghouse can be transported to the finish mill area to be<br />

72


added to the cement or disposed as waste. The air and sorbent flow rates are expected to<br />

be smaller than what would be seen with a full carbon injection system.<br />

Figure 3.10. Sketch of the roaster process [149].<br />

It should be pointed out that the process is still under development. Information is<br />

lacking on the achievable mercury removal efficiency and operating cost. Since most<br />

dust is removed <strong>by</strong> the ESP it appears that sorbent is still required for capture the<br />

mercury evaporated <strong>from</strong> the roaster. Calcium chloride may be required to oxidize the<br />

elemental mercury and enhance mercury capture <strong>by</strong> the sorbent. It is unclear how<br />

effectively the elemental mercury can be oxidized and how much calcium chloride is<br />

required.<br />

3.9 Conclusions<br />

<strong>Mercury</strong> can be removed <strong>from</strong> the flue gas <strong>by</strong> fuel cleaning and switching, raw<br />

material cleaning, sorbent injection, sorbent bed, oxidation <strong>by</strong> catalyst and subsequent<br />

removal <strong>by</strong> wet scrubber, spray drier absorber, and roaster process with smaller sorbent<br />

injection system. Presently sorbent injection is considered as the most promising and<br />

developed mercury removal technology. <strong>Mercury</strong> removal <strong>by</strong> sorbent injection can be<br />

affected <strong>by</strong> many factors such as mercury speciation and concentration, flue gas<br />

73


composition and temperature, mercury vapor-sorbent contacting time, sorbents and<br />

sorbent dispersion, etc. Due to the high moisture level and lack of carbonaceous particles<br />

in the cement kiln flue gas, and release of the captured mercury during recirculation to<br />

the kiln, the application of sorbent injection to cement kilns will be more challenging and<br />

the obtained knowledge <strong>from</strong> coal-fired power plants and waste incinerators cannot be<br />

applied to cement kiln directly. The PAC injection system should be installed<br />

downstream of the main kiln fabric filter and upstream of a new added polishing fabric<br />

filter to avoid the cement kiln dust recycling and increased disposal issues.<br />

Powdered activated carbon is the most widely used sorbent for mercury removal<br />

<strong>from</strong> flue gas. However, there is a lack of fundamental investigation of mercury<br />

adsorption <strong>by</strong> activated carbon in simulated cement kiln flue gas. Even for power plant<br />

application, mercury capture kinetics is not available in most of the publications and<br />

many of the studies were carried out in air or nitrogen without acid gases presence. The<br />

majority of the publications focused on elemental mercury capture and only few studies<br />

investigated capture of HgCl2 which is a major mercury species. To reduce the cost of<br />

sorbent and possible disposal expense, non-carbon based and concrete/cement friendly<br />

sorbents such as Amended Silicate TM have been developed. Other developments include<br />

regeneration and recirculation of sorbents and in-situ generation of activated carbon. The<br />

performance of these sorbents needs to be proved in the full-scale application.<br />

The carbon-oxygen surface complexes and flue gas composition play an<br />

important role in mercury removal <strong>by</strong> activated carbon injection. Both physisorption and<br />

chemisorption are involved in mercury capture <strong>by</strong> carbons. The mechanisms of elemental<br />

mercury capture on the carbons consist of surface-catalyzed oxidation of the elemental<br />

mercury via interaction with surface-bound halide species with subsequent binding <strong>by</strong><br />

surface halide or sulphate species. Co-presence of SO2 and NO2 in the flue gas results in<br />

a poor performance of carbons. There is competitive adsorption between Hg and SO3<br />

since both mercury and SO3 bind to the Lewis acid base sites on the activated carbon<br />

surface.<br />

74


3.10 Further research requirement<br />

Activated carbon injection is a promising technology, but further research is<br />

needed to provide the best sorbent with effective mercury capture at a low cost.<br />

Investigation of mercury capture <strong>by</strong> the activated carbon using simulated cement kiln<br />

flue gas is imperative to evaluate whether the activated carbon is also a promising<br />

sorbent for cement plant application. Lab-scale tests are desired to obtain kinetics and<br />

study the effects of different operating parameters.<br />

More focus is needed on developing alternative sorbents. Fly ash and cement raw<br />

materials such as clay and silica might be used as cement-friendly sorbents and<br />

alternatives for activated carbon. A better understanding of mercury removal <strong>by</strong> fly ash<br />

and other cement-friendly sorbents is therefore needed. Fundamental investigation on the<br />

regeneration of the sorbents and enhancement of the sorbents <strong>by</strong> adding chemical agents<br />

during regeneration is required. Focus should be put on desorption temperature,<br />

separation and purification of collected mercury compounds and possible regeneration<br />

cycle of the sorbents. Possibility of regenerating the sorbent <strong>by</strong> hot flue gas <strong>from</strong> the kiln<br />

system and in-situ enhancement of the sorbent should be investigated.<br />

3.11 Abbreviations<br />

APCD: Air pollution control device<br />

CKD: <strong>Cement</strong> kiln dust<br />

COHPAC: Compact hybrid particulate collector<br />

ESP: Electrostatic precipitator<br />

EXAFS: Extended X-ray absorption fine structure<br />

FF: Fabric filter<br />

FGD: Flue gas desulphurization<br />

LNB: Low NOx burner<br />

PAC: Powdered activated carbon<br />

PPS: Polyphenylene sulphide<br />

PRB: Powder River Basin<br />

75


SCR: Selective catalytic reduction<br />

SDA: Spray dryer absorber<br />

SNCR: Selective non-catalytic reduction<br />

TCLP: Toxicity characteristic leaching procedure<br />

TDF: Tire-derived fuel<br />

XAFS: X-ray absorption fine structure<br />

XANES: X-ray absorption near-edge spectroscopy<br />

XPS: X-ray photoelectron spectroscopy<br />

3.12 References<br />

[1] A. Licata, R. Beittel and T. Ake, Multi-pollutant emissions control & strategies: Coal-fired<br />

power plant mercury control <strong>by</strong> injecting sodium tetrasulfide, Institutes of Clean Air Companies<br />

(ICAC) Forum 2003, Nashville, TN, October 14-15, 2003.<br />

[2] F. Rubiera, S.T. Hall, C.L. Shah, Sulfur removal <strong>by</strong> fine coal cleaning processes, Fuel. 76<br />

(1997) 1187-1194.<br />

[3] D. Akers, R. Dospoy, Role of coal cleaning in control of air toxics, Fuel Processing<br />

Technology. 39 (1994) 73-86.<br />

[4] R.J. Conzemius, C.D. Chriswell, G.A. Junk, The partitioning of elements during physical<br />

cleaning of coals, Fuel Processing Technology. 19 (1988) 95-106.<br />

[5] J. Richards, Capabilities and limitations of available control technologies for mercury<br />

emissions <strong>from</strong> cement kilns, PCA R&D Serial No. 2748a, 2005.<br />

[6] T.D. Brown, D.N. Smith, R.A. Hargis Jr., W.J. O'Dowd, <strong>Mercury</strong> measurement and its<br />

control: What we know, have learned, and need to further investigate, EM: Air and Waste<br />

Management Association's Magazine for Environmental Managers. (1999) 47-50.<br />

[7] D.J. Akers, Coal cleaning controls HAP emissions, Power Eng. 100 (1996) 33.<br />

[8] U.S. EPA, <strong>Mercury</strong> study report to congress, EPA-452/R-97-003, 1997.<br />

[9] N. Yoshiyuki and T. Haruhiko, Method of producing cement, Patent, JP2007326750, 2007.<br />

[10] T. Junichi, U. Yasuhiro and Y. Osamu, Method of treating exhaust gas <strong>from</strong> cement kiln,<br />

Patent, JP2003192407, 2003.<br />

[11] Y. Zhuang, S.J. Miller, Impact of supplemental firing of tire-derived fuel (TDF) on mercury<br />

species and mercury capture with the advanced hybrid filter in a western subbituminous coal flue<br />

gas, Energy Fuels. 20 (2006) 1039-1043.<br />

[12] J. Richards, D. Goshaw, D. Speer and T. Holder, Air emissions data summary for Portland<br />

cement pyroprocessing operations firing tire-derived fuels, PCA R&D Serial No. 3050, 2008.<br />

76


[13] J.H. Pavlish, E.A. Sondreal, M.D. Mann, E.S. Olson, K.C. Galbreath, D.L. Laudal, S.A.<br />

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

Technology. 82 (2003) 89-165.<br />

[14] T.C. Ho, P. Yang, T.H. Kuo, J.R. Hopper, Characteristics of mercury desorption <strong>from</strong><br />

sorbents at elevated temperatures, Waste Management. 18 (1998) 445-452.<br />

[15] R. Yan, D.T. Liang, L. Tsen, Y.P. Wong, Y.K. Lee, Bench-scale experimental evaluation of<br />

carbon performance on mercury vapour adsorption, Fuel. 83 (2004) 2401-2409.<br />

[16] J.J. Cudahy, R.W. Helsel, <strong>Removal</strong> of products of incomplete combustion with carbon,<br />

Waste Manage. 20 (2000) 339-345.<br />

[17] Y.H. Li, C.W. Lee, B.K. Gullett, The effect of activated carbon surface moisture on low<br />

temperature mercury adsorption, Carbon. 40 (2002) 65-72.<br />

[18] W. Liu, R.D. Vidic, T.D. Brown, Impact of flue gas conditions on mercury uptake <strong>by</strong> sulfurimpregnated<br />

activated carbon, Environ. Sci. Technol. 34 (2000) 154-159.<br />

[19] S.J. Miller, G.E. Dunham, E.S. Olson, T.D. Brown, Flue gas effects on a carbon-based<br />

mercury sorbent, Fuel Processing Technology. 65-66 (2000) 343-363.<br />

[20] R. Ochiai, M.A. Uddin, E. Sasaoka, S. Wu, Effects of HCl and SO2 concentration on<br />

mercury removal <strong>by</strong> activated carbon sorbents in coal-derived flue gas, Energy & Fuels. 23 (2009)<br />

4734.<br />

[21] A.A. Presto, E.J. Granite, Impact of sulfur oxides on mercury capture <strong>by</strong> activated carbon,<br />

Environ. Sci. Technol. 41 (2007) 6579-6584.<br />

[22] A.A. Presto, E.J. Granite, A. Karash, Further investigation of the impact of sulfur oxides on<br />

mercury capture <strong>by</strong> activated carbon, Ind Eng Chem Res. 46 (2007) 8273-8276.<br />

[23] J. Jarvis, F. Meserole, SO3 effect on mercury control, Power Eng. 112 (2008) 54-60.<br />

[24] D.D. Do, Adsorption analysis: equilibria and kinetics, Imperial College Press, 1998.<br />

[25] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R. Chang,<br />

Preparation and evaluation of coal-derived activated carbons for removal of mercury vapor <strong>from</strong><br />

simulated coal combustion flue gases, Energy Fuels. 12 (1998) 1061-1070.<br />

[26] M. Rostam-Abadi, S.G. Chen, H.C. Hsi, M. Rood, R. Chang, T.R. Carey, B., Hargrove, C.<br />

Richardson, W. Rosenhoover, F. Meserole, Novel vapor phase mercury sorbents, Proceedings of<br />

the EPRI-DOE-EPA Combined Utility Air Pollutant Control, Washington, DC, Aug 25–29, 1997.<br />

[27] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang and F.B. Meserole, Factors<br />

affecting mercury control in utility flue gas using sorbent injection, Proceedings of the Air &<br />

Waste Management Association's 90th Annual Meeting, Toronto, Ontario, Canada, June 8-13,<br />

1997.<br />

[28] C.X. Hu, J.S. Zhou, Z.Y. Luo, S. He, G.K. Wang, K.F. Cen, Effect of oxidation treatment on<br />

the adsorption and the stability of mercury on activated carbon, Journal of Environmental<br />

Sciences-China. 18 (2006) 1161-1166.<br />

[29] B. Ghorishi and B.K. Gullett, Fixed-bed control of mercury: Role of acid gases and a<br />

comparison between carbon-based, calcium-based, and coal fly ash sorbents, Proceedings of the<br />

77


EPRI/DOE/EPA Combined Utility Air Pollutant Control Symposium, Washington, DC, August<br />

25–29,1997.<br />

[30] S.B. Ghorishi and C.B. Sedman, Combined mercury and sulfur oxides control using<br />

calcium-based sorbents, Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant<br />

Control Symposium, Washington, DC, August 25–29, 1997.<br />

[31] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption <strong>by</strong> activated<br />

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy Conference,<br />

Research Triangle Park, NC, 22-25 April, 1997.<br />

[32] N. Hutson, C. Singer, C. Richardson, J. Karwowski and C. Sedman, Practical applications<br />

<strong>from</strong> observations of mercury oxidation and binding mechanisms, EPRI-DOE-EPA-AWMA<br />

Combined Power Plant Air Pollutant Control Mega Symposium, Washington DC, August 30 -<br />

September 2, 2004.<br />

[33] E.J. Granite, M.C. Freeman, R.A. Hargis, W.J. O'Dowd, H.W. Pennline, The thief process<br />

for mercury removal <strong>from</strong> flue gas, J. Environ. Manage. 84 (2007) 628-634.<br />

[34] W.J. O’Dowd, H.W. Pennline, M.C. Freeman, E.J. Granite, R.A. Hargis, C.J. Lacher, A.<br />

Karash, A technique to control mercury <strong>from</strong> flue gas: The thief process, Fuel Processing<br />

Technology. 87 (2006) 1071-1084.<br />

[35] R. Bhardwaj. Impact of temperature and flue gas components on mercury speciation and<br />

uptake <strong>by</strong> activated carbon sorbents. Master thesis. Master, University of Pittsburgh, 2007.<br />

[36] M.M. Maroto-Valer, Y. Zhang, E.J. Granite, Z. Tang, H.W. Pennline, Effect of porous<br />

structure and surface functionality on the mercury capacity of a fly ash carbon and its activated<br />

sample, Fuel. 84 (2005) 105-108.<br />

[37] J.E. Cichanowicz, R. Stewart, K. Baldrey, Powdered activated carbon (PAC)<br />

characterization tests for coal-fired utility application, Power Plant Air Pollutant Control<br />

"MEGA" Symposium 2008, Baltimore, MD, August 25-28, 2008.<br />

[38] Carbochem, Activated carbon 101, http://www.carbochem.com/activatedcarbon101.html,<br />

accessed September 21, 2010.<br />

[39] Norit Americas Inc., Datasheet of Darco FGD powdered activated carbon, 2008.<br />

[40] Norit Americas Inc., Datasheet of Norit Darco Hg activated carbon, 2007.<br />

[41] K. Dombrowski, Evaluation of low-ash impact sorbent injection technologies at a Texas<br />

lignite/PRB fired power plant. Presented at the DOE/NETL <strong>Mercury</strong> Control Technology<br />

Conference, Pittsburgh, PA, USA, December 11-13, 2007.<br />

[42] Norit Americas Inc., Datasheet of Norit Darco Hg-LH activated carbon, 2007.<br />

[43] Calgon, Data sheet of FluePac-MC plus powdered activated carbon, 2010.<br />

[44] S.H. Lee, Y.J. Rhim, S.P. Cho, J.I. Baek, Carbon-based novel sorbent for removing gasphase<br />

mercury, Fuel. 85 (2006) 219-226.<br />

[45] Y. Shin, G. Fryxell, W. Um, K. Parker, S. Mattigod, R. Skaggs, Sulfur-Functionalized<br />

Mesoporous Carbon, Advanced Functional Materials. 17 (2007) 2897-2901.<br />

78


[46] J.W. Graydon, X. Zhang, D.W. Kirk, C.Q. Jia, Sorption and stability of mercury on<br />

activated carbon for emission control, J. Hazard. Mater. 168 (2009) 978-982.<br />

[47] J. Wirling, J. Jablonski, Safety aspects in the use of carbonaceous sorbents during waste gas<br />

treatment, Metallurgical Plant and Technology. 3 (2007) 144.<br />

[48] R. Yan, Y.L. Ng, D.T. Liang, C.S. Lim, J.H. Tay, Bench-scale experimental study on the<br />

effect of flue gas composition on mercury removal <strong>by</strong> activated carbon adsorption, Energy Fuels.<br />

17 (2003) 1528-1535.<br />

[49] H. Zeng, F. Jin, J. Guo, <strong>Removal</strong> of elemental mercury <strong>from</strong> coal combustion flue gas <strong>by</strong><br />

chloride-impregnated activated carbon, Fuel. 83 (2004) 143-146.<br />

[50] H. Yang, Z. Xu, M. Fan, A.E. Bland, R.R. Judkins, Adsorbents for capturing mercury in<br />

coal-fired boiler flue gas, Journal of Hazardous Materials. 146 (2007) 1-11.<br />

[51] S.B. Ghorishi, R.M. Keeney, S.D. Serre, B.K. Gullett, W.S. Jozewicz, Development of a Climpregnated<br />

activated carbon for entrained-flow capture of elemental mercury, Environ. Sci.<br />

Technol. 36 (2002) 4454-4459.<br />

[52] S.J. Lee, Y. Seo, J. Jurng, T.G. Lee, <strong>Removal</strong> of gas-phase elemental mercury <strong>by</strong> iodine-<br />

and chlorine-impregnated activated carbons, Atmospheric Environment. 38 (2004) 4887-4893.<br />

[53] E.J. Granite, H.W. Pennline, R.A. Hargis, Novel sorbents for mercury removal <strong>from</strong> flue gas,<br />

Ind. Eng. Chem. Res. 39 (2000) 1020-1029.<br />

[54] R.D. Vidic, D.P. Siler, Vapor-phase elemental mercury adsorption <strong>by</strong> activated carbon<br />

impregnated with chloride and chelating agents, Carbon. 39 (2001) 3-14.<br />

[55] Z. Mei, Z. Shen, T. Yuan, W. Wang, H. Han, <strong>Removal</strong> of vapor-phase elemental mercury <strong>by</strong><br />

N-doped CuCoO4 loaded on activated carbon, Fuel Processing Technology. 88 (2007) 623-629.<br />

[56] H.C. Hsi, M.J. Rood, M. Rostam-Abadi, S. Chen, R. Chang, Effects of sulfur impregnation<br />

temperature on the properties and mercury adsorption capacities of activated carbon fibers<br />

(ACFs), Environ. Sci. Technol. 35 (2001) 2785-2791.<br />

[57] S. Vitolo, M. Seggiani, <strong>Mercury</strong> removal <strong>from</strong> geothermal exhaust gas <strong>by</strong> sulfurimpregnated<br />

and virgin activated carbons, Geothermics. 31 (2002) 431-442.<br />

[58] K.D. Henning, K. Keldenicht, K. Knoblauch, J. Degel, Impregnated activated carbon for<br />

mercury removal, Gas Separation & Purification. 2 (1988) 20-22.<br />

[59] D. Karatza, A. Lancia, D. Musmarra, C. Zucchini, Study of mercury absorption and<br />

desorption on sulfur impregnated carbon, Experimental Thermal and Fluid Science. 21 (2000)<br />

150-155.<br />

[60] W. Liu, R.D. Vidic, T.D. Brown, Optimization of high temperature sulfur impregnation on<br />

activated carbon for permanent sequestration of elemental mercury vapors, Environ. Sci. Technol.<br />

34 (2000) 483-488.<br />

[61] J.A. Korpiel, R.D. Vidic, Effect of sulfur impregnation method on activated carbon uptake<br />

of gas-phase mercury, Environ. Sci. Technol. 31 (1997) 2319-2325.<br />

[62] S. Lee, J. Lee, T.C. Keener, Novel sorbents for mercury emissions control <strong>from</strong> coal-fired<br />

power plants, Journal of the Chinese Institute of Chemical Engineers. 39 (2008) 137-142.<br />

79


[63] C.A. Lockert, Q. Zhou, Y. Zhang and S. Nelson Jr., Further progress toward concretefriendly<br />

TM mercury sorbents, Air Quality V Symposium, Washington, DC, September 19-21,<br />

2005.<br />

[64] S. Manchester, X. Wang, I. Kulaots, Y. Gao, R.H. Hurt, High capacity mercury adsorption<br />

on freshly ozone-treated carbon surfaces, Carbon. 46 (2008) 518-524.<br />

[65] Z. Mei, Z. Shen, Q. Zhao, W. Wang, Y. Zhang, <strong>Removal</strong> and recovery of gas-phase element<br />

mercury <strong>by</strong> metal oxide-loaded activated carbon, Journal of Hazardous Materials. 152 (2008)<br />

721-729.<br />

[66] S. Lee, J. Lee, T.C. Keener, Bench-scale studies of in-duct mercury capture using cupric<br />

chloride-impregnated carbons, Environ. Sci. Technol. 43 (2009) 2957-2962.<br />

[67] S. Lee, J. Lee, T.C. Keener, The effect of methods of preparation on the performance of<br />

cupric chloride-impregnated sorbents for the removal of mercury <strong>from</strong> flue gases, Fuel. 88 (2009)<br />

2053-2056.<br />

[68] J.A. Korpiel, R.D. Vidic, Effect of sulfur impregnation method on activated carbon uptake<br />

of gas-phase mercury, Environ. Sci. Technol. 31 (1997) 2319-2325.<br />

[69] X. Chen, M. Farber, Y. Gao, I. Kulaots, E.M. Suuberg, R.H. Hurt, Mechanisms of surfactant<br />

adsorption on non-polar, air-oxidized and ozone-treated carbon surfaces, Carbon. 41 (2003)<br />

1489-1500.<br />

[70] Y. Gao, I. Külaots, X. Chen, R. Aggarwal, A. Mehta, E.M. Suuberg, et al., Ozonation for the<br />

chemical modification of carbon surfaces in fly ash, Fuel. 80 (2001) 765-768.<br />

[71] K.H. Pedersen, A.D. Jensen, M.S. Skjøth-Rasmussen, K. Dam-Johansen, A review of the<br />

interference of carbon containing fly ash with air entrainment in concrete, Progress in Energy and<br />

Combustion Science. 34 (2008) 135-154.<br />

[72] K.H. Pedersen, M.C. Melia, A.D. Jensen, K. Dam-Johansen, Post-treatment of fly ash <strong>by</strong><br />

ozone in a fixed bed reactor, Energy and Fuels. 23 (2009) 280-285.<br />

[73] K.H. Pedersen, A.D. Jensen, M. Berg, L.H. Olsen, K. Dam-Johansen, The effect of<br />

combustion conditions in a full-scale low-NOx coal fired unit on fly ash properties for its<br />

application in concrete mixtures, Fuel Process Technol. 90 (2009) 180-185.<br />

[74] C.A. Wocken, Evaluation of Enviroscrub's multipollutant Pahlman TM process for mercury<br />

removal at facility burning subbituminous coal, final plant report, 05-EERC-04-04, 2004.<br />

[75] R.M. Boren, NETL/DOE mercury control technology R&D program review- mercury<br />

removal results <strong>from</strong> two coal-fired utility boilers, 2004.<br />

[76] E.S. Tavoulareas and W. Jozewicz, Multipollutant emission control technology options for<br />

coal-fired power plants, EPA-600/R-05/034, 2005.<br />

[77] R. Keeth, S. Peterson and M. Reed, Status of multi-pollutant process development:<br />

LoTOx TM /BOC, GSA TM /FLS, Pahlmanite TM , Cost projections, 1004245, 2003.<br />

[78] J.Y. Lee, Y. Ju, T.C. Keener, R.S. Varma, Development of cost-effective noncarbon<br />

sorbents for Hg 0 removal <strong>from</strong> coal-fired power plants, Environ. Sci. Technol. 40 (2006) 2714-<br />

2720.<br />

80


[79] M. Holmes and J. Pavlish, <strong>Mercury</strong> information clearinghouse, Quarter 3- Advanced and<br />

developmental mercury control technologies, July 2004.<br />

[80] J.R. Butz, T.E. Broderick and C.S. Turchi, Amended Silicates for <strong>Mercury</strong> Control,<br />

project final report, DOE Award Number: DE-FC26-04NT41988, 2006.<br />

[81] ADA Technologies Inc, Arsenic removal technologies, http://www.adatech.com/default.asp,<br />

accessed April 20, 2010.<br />

[82] J.C. Hower, C.L. Senior, E.M. Suuberg, R.H. Hurt, J.L. Wilcox, E.S. Olson, <strong>Mercury</strong><br />

capture <strong>by</strong> native fly ash carbons in coal-fired power plants, Progress in Energy and Combustion<br />

Science. 36 (2010) 510.<br />

[83] D. Karatza, A. Lancia, D. Musmarra, Fly ash capture of mercuric chloride vapors <strong>from</strong><br />

exhaust combustion gas, Environ. Sci. Technol. 32 (1998) 3999-4004.<br />

[84] Z. Li, X. Sun, J. Luo, J.Y. Hwang, J.C. Crittenden, Unburned carbon <strong>from</strong> fly ash for<br />

mercury adsorption: II. Adsorption isotherms and mechanisms, Journal of Minerals & Materials<br />

Characterization & Engineering. 1 (2002) 76.<br />

[85] G.E. Dunham, R.A. DeWall, C.L. Senior, Fixed-bed studies of the interactions between<br />

mercury and coal combustion fly ash, Fuel Processing Technology,. 82 (2003) 197-213.<br />

[86] S.D. Serre, G.D. Silcox, Adsorption of elemental mercury on the residual carbon in coal fly<br />

ash, Ind. Eng. Chem. Res. 39 (2000) 1723-1730.<br />

[87] D.J. Hassett, K.E. Eylands, <strong>Mercury</strong> capture on coal combustion fly ash, Fuel. 78 (1999)<br />

243-248.<br />

[88] W.D. Owens, A.F. Sarofim, D.W. Pershing, The use of recycle for enhanced volatile metal<br />

capture, Fuel Processing Technology. 39 (1994) 337-356.<br />

[89] J. Dong, Z. Xu, S.M. Kuznicki, <strong>Mercury</strong> removal <strong>from</strong> flue gases <strong>by</strong> novel regenerable<br />

magnetic nanocomposite sorbents, Environ. Sci. Technol. 43 (2009) 3266-3271.<br />

[90] Lawrence E. Bool III, Jurron Bradley and David R. Thompson, A novel process for onsite<br />

production of mercury sorbents, http://www.icac.com/files/public/Clearwater_2006_Bool_.pdf,<br />

accessed June 30, 2008.<br />

[91] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for DTE<br />

Energy’s Monroe Station, DOE Award Number DE-FC26-03NT41986, Report Number<br />

41986R16, 2006.<br />

[92] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for<br />

Sunflower Electric’s Holcomb Station, DE-FC26-03NT41986, Topical Report No. 41986R07,<br />

2005.<br />

[93] A.P. Jones, J.W. Hoffmann, D.N. Smith, T.J. Feeley, J.T. Murphy, DOE/NETL's Phase II<br />

<strong>Mercury</strong> Control Technology Field Testing Program: Preliminary Economic Analysis of<br />

Activated Carbon <strong>Injection</strong>, Environ. Sci. Technol. 41 (2007) 1365-1371.<br />

[94] A.P. Jones, J.W. Hoffmann, D.N. Smith, T.J. Feeley and J.T. Murphy, DOE/NETL's Phase<br />

II mercury control technology field test program: Updated economic analysis of activated carbon<br />

injection, 2007.<br />

81


[95] ADA-ES, Field test program to develop comprehensive design, operating, and cost data for<br />

mercury control systems: Final site report for E.C. Gaston Unit 3 sorbent injection into COHPAC<br />

for mercury control, DE-FC26-00NT41005, Topical Report No. 41005R11, 2003.<br />

[96] J. Bustard, M. Durham, T. Starns, C. Lindsey, C. Martin, R. Schlager, K. Baldrey, Full-scale<br />

evaluation of sorbent injection for mercury control on coal-fired power plants, Fuel Processing<br />

Technology. 85 (2004) 549-562.<br />

[97] J. Pavlish, Baghouse balance-of-plant effects at TXU’s Big Brown Station during field<br />

testing of sorbent injection for Hg control, 10 th Annual Electric Utilities Environmental<br />

Conference, Tucson, Arizona, January 23, 2007.<br />

[98] S. Derenne, P. Sartorelli, D. Johnson, R. Stewart, C.J. Bustad, S. Sjostrom, D. Muggli, T.<br />

Starns, T. McMahon, F. Sudhoff, R. Chang, R. Utter, J. Cummings, TOXECON retrofit for<br />

mercury and multi-pollutant control, 2006 Symposium on Western Fuels, Denver, 24-26 October,<br />

2006.<br />

[99] S.T. Derenne, TOXECON retrofit for mercury and multi-pollutant control on three 90-MW<br />

coal fired boilers. Quarterly technical progress report, 1 st January-31 st March 2008, Report No.<br />

41766R08, 2008.<br />

[100] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for<br />

AmerenUE’s Meramec Station Unit 2, DE-FC26-03NT41986, Topical Report No. 41986R09,<br />

2005.<br />

[101] ADA-ES, Field test program to develop comprehensive design, operating, and cost data for<br />

mercury control systems: Final site report for Pleasant Prairie Power Plant Unit 2 sorbent<br />

injection into a cold-side ESP for mercury control, DE-FC26-00NT41005, Topical Report No.<br />

41005R12, 2003.<br />

[102] ADA-ES, Field test program to develop comprehensive design, operating, and cost data for<br />

mercury control systems: Final site report for Brayton Point Generating Station Unit 1 sorbent<br />

injection into a cold-side ESP for mercury control, DE-FC26-00NT41005, Topical Report No.<br />

41005R21, 2005.<br />

[103] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for Basin<br />

Electric Power Cooperative’s Laramie River Station, DE-FC26-03NT41986, Topical Report No.<br />

41986R11, 2006.<br />

[104] ADA-ES, Field test program to develop comprehensive design, operating, and cost data for<br />

mercury control systems: Final site report for PG&E NEG Salem Harbor Station Unit 1 sorbent<br />

injection into a cold-side ESP for mercury control, U.S. DOE Cooperative Agreement No. DE-<br />

FC26-00NT41005, 2004.<br />

[105] S. Sjostrom, M. Dillon, B. Donnelly, J. Bustard, G. Filippelli, R. Glesmann, T. Orscheln,<br />

S.,Wahlert, R. Chang, A. O’Palko, Influence of SO3 on mercury removal with activated carbon:<br />

Full-scale results, Fuel Process Technol. 90 (2009) 1419-1423.<br />

[106] L. Hayes-Gorman, Regulating mercury emissions: Ash Grove <strong>Cement</strong> in Durkee, Air<br />

toxics summit 2008, Boise, Idaho, 4-7 August, 2008.<br />

82


[107] Schreiber & Yonley Associates, <strong>Mercury</strong> emissions test report, Ash Grove <strong>Cement</strong><br />

Company Durkee, Oregon, Project No. 060204, 2007.<br />

[108] M.L. Jones, J.P. Kay, B.M. Pavlish and K.C. Galbreath, <strong>Mercury</strong> control slipstream<br />

baghouse at Ash Grove’s Durkee cement facility, 2007.<br />

[109] I.I. Salame, T.J. Bandosz, Surface chemistry of activated carbons: Combining the results of<br />

temperature-programmed desorption, Boehm, and potentiometric titrations, Journal of Colloid<br />

and Interface Science. 240 (2001) 252-258.<br />

[110] J. Lahaye, The chemistry of carbon surfaces, Fuel. 77 (1998) 543-547.<br />

[111] J.L. Figueiredo, M.F.R. Pereira, M.M.A. Freitas, J.J.M. Órfão, Modification of the surface<br />

chemistry of activated carbons, Carbon. 37 (1999) 1379-1389.<br />

[112] F.E. Huggins, N. Yap, G.P. Huffman, C.L. Senior, XAFS characterization of mercury<br />

captured <strong>from</strong> combustion gases on sorbents at low temperatures, Fuel Processing Technology.<br />

82 (2003) 167-196.<br />

[113] N.D. Hutson, B.C. Attwood, K.G. Scheckel, XAS and XPS characterization of mercury<br />

binding on brominated activated carbon, Environ. Sci. Technol. 41 (2007) 1747-1752.<br />

[114] B. Podak, M. Brunetti, A. Lewis, J. Wilcox, <strong>Mercury</strong> binding on activated carbon, Environ.<br />

Prog. 25 (2006) 319-326.<br />

[115] E.S. Olson, S.J. Miller, R.K. Sharma, G.E. Dunham, S.A. Benson, Catalytic effects of<br />

carbon sorbents for mercury capture, Journal of Hazardous Materials. 74 (2000) 61-79.<br />

[116] G.E. Dunham, E.S. Olson and S.J. Miller, Impact of flue gas constituents on carbon<br />

sorbents, McLean, VA, September 19-21, 2000.<br />

[117] E.S. Olson, B.A. Mibeck, S.A. Benson, J.D. Laumb, C.R. Crocker, G.E. Dunham, et al.,<br />

The mechanistic model for flue gas-mercury interactions on activated carbons: The oxidation site,<br />

Prepr. Pap. -Am. Chem. Soc. , Div. Fuel Chem. 49 (2004) 279-280.<br />

[118] Y. Zheng, A.D. Jensen, J.E. Johnsson, Deactivation of V2O5-WO3-TiO2 SCR catalyst at a<br />

biomass-fired combined heat and power plant, Applied Catalysis B: Environmental. 60 (2005)<br />

253-264.<br />

[119] Y. Zheng, A.D. Jensen and J.E. Johnsson, DTU final report of EU project (NNE5-2001-<br />

00164), Influences <strong>from</strong> Biofuel (Co-)Combustion on Catalytic Converters in Coal-fired Power<br />

<strong>Plants</strong>, 2004.<br />

[120] A. Yilmaz, L. Hindiyarti, A.D. Jensen, P. Glarborg, P. Marshall, Thermal dissociation of<br />

SO3 at 1000-1400 K, Journal of Physical Chemistry a. 110 (2006) 6654-6659.<br />

[121] J. Jarvis, F. Meserole, SO3 Effect on <strong>Mercury</strong> Control, Power Eng. 112 (2008) 54-60.<br />

[122] A.A. Lizzio, J.A. DeBarr, Mechanism of SO2 removal <strong>by</strong> carbon, Energy & Fuels. 11<br />

(1997) 284-291.<br />

[123] F.H. Yang, R.T. Yang, Ab initio molecular orbital study of the mechanism of SO2<br />

oxidation catalyzed <strong>by</strong> carbon, Carbon. 41 (2003) 2149-2158.<br />

[124] D.L. Nickels and T. Weyand, Processing and reuse of activated carbon used to adsorb<br />

mercury <strong>from</strong> power plant flue gases. Final report, Contract#: 2447-PMET-DOE-0350, 2004.<br />

83


[125] T.C. Ho, P. Yang, T.H. Kuo, J.R. Hopper, Characteristics of mercury desorption <strong>from</strong><br />

sorbents at elevated temperatures, Waste Management. 18 (1998) 445-452.<br />

[126] U.S. EPA, Land disposal restrictions for mercury-containing wastes,<br />

http://www.epa.gov/epaoswer/hazwaste/mercury/regs.htm, accessed April 16, 2008.<br />

[127] E.K. Levy, C. Kiely and Z. Yao, Investigation of fly ash and activated carbon obtained<br />

<strong>from</strong> pulverized coal boilers, DE–FG26-03NT41796, Final report, 2006.<br />

[128] T. Machalek, C. Richardson, J. Noblett, Field investigations of fixed-bed sorbents for<br />

mercury capture <strong>from</strong> coal-fired flue gas, Power Plant Air Pollutant Control "MEGA"<br />

Symposium 2008, Baltimore, MD, August 25-28, 2008.<br />

[129] R. Change, G.R. Offen, <strong>Mercury</strong> emission control technologies: An EPRI synopsis, Power<br />

Engineering. 99 (1995).<br />

[130] M. Dίaz-Somoano, S. Unterberger, K.R.G. Hein, <strong>Mercury</strong> emission control in coal-fired<br />

plants: The role of wet scrubbers, Fuel Processing Technology. 88 (2007) 259-263.<br />

[131] C.L. Kairies, K.T. Schroeder, C.R. Cardone, <strong>Mercury</strong> in gypsum produced <strong>from</strong> flue gas<br />

desulfurization, Fuel. 85 (2006) 2530-2536.<br />

[132] S. Niksa, N. Fujiwara, The impact of wet flue gas desulfurization scrubbing on mercury<br />

emissions <strong>from</strong> coal-fired power stations, J. Air Waste Manage. Assoc. 55 (2005) 970-977.<br />

[133] P.S. Nolan, K.E. Redinger, G.T. Amrhein, G.A. Kudlac, Demonstration of additive use for<br />

enhanced mercury emissions control in wet FGD systems, Fuel Processing Technology. 85 (2004)<br />

587-600.<br />

[134] A. Licata, W. Schuettenhelm and M. Klein, <strong>Mercury</strong> controls for MWCs using the sodium<br />

tetrasulfide process, 8 th Annual North American Waste-to-Energy Conference, Nashville, TN,<br />

May 22, 2000.<br />

[135] D.L. Laudal, J.S. Thompson, J.H. Pavlish, L. Brickett, P. Chu, R.K. Srivastava, J. Kilgroe,<br />

C.W. Lee, <strong>Mercury</strong> speciation at power plants using SCR and SNCR control technologies, EM:<br />

Air and Waste Management Association's Magazine for Environmental Managers. (2003) 16-22.<br />

[136] Thomas J. Feeley, Lynn A. Brickett and James T. Murphy, Evaluation of the effect of SCR<br />

NOx control technology on mercury speciation, http://www.netl.doe.gov/technologies/coalpower<br />

/ewr/pubs/SCRHgPaperFinal030503.pdf, accessed September 9, 2010.<br />

[137] H.G. Pedersen, L.S. Pedersen, H. Rostgaard and K. Pedersen, Oxidation of mercury on<br />

DNX catalysts. Proceedings of the Air Quality V: <strong>Mercury</strong>, Trace Elements, SO3, and Particulate<br />

Matter Conference, Arlington, VA, Sept 19–21, 2005.<br />

[138] S. Eswaran, H.G. Stenger, Understanding mercury conversion in selective catalytic<br />

reduction (SCR) catalysts, Energy & Fuels. 19 (2005) 2328-2334.<br />

[139] Y. Cao, B. Chen, J. Wu, H. Cui, J. Smith, C.K. Chen, P. Chu, W. Pan, Study of mercury<br />

oxidation <strong>by</strong> a selective catalytic reduction catalyst in a pilot-scale slipstream reactor at a utility<br />

boiler burning bituminous coal, Energy & Fuels. 21 (2007) 145-156.<br />

[140] H. Kamata, S. Ueno, T. Naito, A. Yukimura, <strong>Mercury</strong> oxidation over the V2O5(WO3)/TiO2<br />

commercial SCR catalyst, Ind Eng Chem Res. 47 (2008) 8136-8141.<br />

84


[141] Y. Cao, Z. Gao, J. Zhu, Q. Wang, Y. Huang, C. Chiu, B. Parker, P. Chu, W. Pan, Impacts<br />

of halogen additions on mercury oxidation in a slipstream selective catalyst reduction (SCR)<br />

reactor when burning sub-bituminous coal, Environ. Sci. Technol. 42 (2008) 256-261.<br />

[142] Y. Zhuang, J. Laumb, R. Liggett, M. Holmes, J. Pavlish, Impacts of acid gases on mercury<br />

oxidation across SCR catalyst, Fuel Processing Technology. 88 (2007) 929-934.<br />

[143] S. Straube, T. Hahn, H. Koeser, Adsorption and oxidation of mercury in tail-end SCR-<br />

DeNOx plants—Bench scale investigations and speciation experiments, Applied Catalysis B:<br />

Environmental. 79 (2008) 286-295.<br />

[144] K. Kai, Y. Nagai, Y. Kato, W. Gretta and H. Kikkawa, SCR catalyst with high mercury<br />

oxidation and low SO2 to SO3 conversion, Baltimore, MD, August 28-31, 2006, .<br />

[145] S.H. Jeon, Y. Eom, T.G. Lee, Photocatalytic oxidation of gas-phase elemental mercury <strong>by</strong><br />

nanotitanosilicate fibers, Chemosphere. 71 (2008) 969-974.<br />

[146] Y. Li, C. Wu, Kinetic study for photocatalytic oxidation of elemental mercury on a SiO2-<br />

TiO2 nanocomposite, Environ. Eng. Sci. 24 (2007) 3-12.<br />

[147] Robert J. Schreiber, Christa O. Russell and Jeff Evers, Evaluation of suitability of selective<br />

catalytic and selective non-catalytic reduction for use in Portland cement industry,<br />

http://www.dep.state.pa.us/dep/DEPUTATE/airwaste/aq/transport/comments/CEMEX_Attachme<br />

nt.pdf, accessed June 3, 2008.<br />

[148] U. Leibacher, NOx elimination <strong>by</strong> the SCR process for cement kiln systems, ZKG<br />

International. 49 (1996) 43.<br />

[149] FLSmidth A/S, New patent-pending system for mercury control, http://ehighlights.<br />

flsmidth.com/node/257, accessed June 30, 2009.<br />

[150] P. Paone, <strong>Mercury</strong> reduction technologies in cement production, presented at 7 th Colloquia<br />

of managers and technicians of cement plants, Palacio de Congresos, Malaga, Spain, 17-19,<br />

September 2009, .<br />

[151] P. Paone, <strong>Mercury</strong> removal systems for the cement industry, presented at 2008 profitproduction<br />

technology seminar, Lehigh Country Club, Allentown, Pennsylvania, USA, 2 October,<br />

2008.<br />

[152] T.L. Mlakar, M. Horvat, T. Vuk, A. Stergaršek, J. Kotnik, J. Tratnik, F. Vesna, <strong>Mercury</strong><br />

species, mass flows and processes in a cement plant, Fuel. 89 (2010) 1936.<br />

85


Experimental methods and materials<br />

To provide a simple means for screening the performance of candidate sorbents<br />

and derive mercury capture kinetics for promising sorbents in a mercury-laden simulated<br />

cement kiln flue gas, a fixed-bed reactor system was designed and built in this project. In<br />

this chapter the entire reactor setup will be described. Furthermore, the choice of some of<br />

the core parts, i.e., the mercury vapor generator, humidifier for water vapor addition and<br />

mercury analysis system are described in more details. Materials and methods applied in<br />

this project are also presented.<br />

4.1 Description of the fixed­bed reactor system<br />

Tests of the oxidized mercury converter and sorbents were conducted in a fixedbed<br />

reactor system as illustrated in figure 4.1. A photo of the system is shown in figure<br />

4.2. Main equipments in the reactor system include a gas mixing system with water vapor<br />

addition <strong>by</strong> a humidifier and mercury source in a calibration gas generator to simulate the<br />

cement kiln flue gas, a low temperature oven with a glass reactor, mercury analysis<br />

system, and mercury traps for exhaust gas treatment. To avoid mercury condensation and<br />

accumulation in the system, all the gas lines before the analyzer are heated to 150�C. All<br />

temperatures including temperatures of heated lines, ovens, reactors, converter, and<br />

analytical cell in mercury analyzer are sampled. The mercury source, reactor and hot<br />

panel are located in a dedicated ventilation hood.<br />

86<br />

4


N 2<br />

N 2<br />

CO 2<br />

O 2<br />

HCl<br />

SO 2<br />

NO x<br />

N 2<br />

Flow meter<br />

MFC<br />

MFC<br />

MFC<br />

MFC<br />

MFC<br />

MFC<br />

MFC<br />

MFC<br />

Vent Vent<br />

Hg source<br />

Evaporator<br />

Rotameter<br />

Filter<br />

87<br />

Heat trace<br />

Reactor<br />

Distribution<br />

box<br />

Analyzer<br />

Converter<br />

Figure 4.1. Sketch of the fixed-bed reactor system for converter and sorbent tests.<br />

Figure 4.2. Photo of the fixed-bed reactor system.<br />

Air<br />

Filter<br />

Vent


4.1.1 Gas mixing system<br />

The gas mixing system consists of valves and mass flow controllers for adding<br />

different gases to simulate cement kiln flue gas. Gas addition includes carrier nitrogen to<br />

the mercury source, carrier nitrogen to humidifier, CO2, O2, HCl, SO2, NO, NO2 and<br />

balance nitrogen. The addition of the gases is controlled <strong>by</strong> the mass flow controllers and<br />

the actual flow rate of each gas is measured <strong>by</strong> a bubble flow meter.<br />

4.1.2 <strong>Mercury</strong> vapor addition system<br />

<strong>Mercury</strong> sources are added using a commercial calibration gas generator <strong>from</strong><br />

VICI Metronics, Dynacalibrator Model 150-06e-C with a gas flow capacity up to 750<br />

ml/min. The Model 150 calibration gas generator is a constant temperature system<br />

designed to generate precise ppm or ppb concentrations of chemical compounds in a gas<br />

stream, using a permeation tube as the trace gas source. Figure 4.3 shows a picture and<br />

sketch of the calibration gas generator. A passivated glass-coated permeation chamber<br />

houses the permeation device, with measured inert carrier gas nitrogen sweeping the<br />

calibration gas/vapor <strong>from</strong> the chamber. A digital temperature controller maintains the<br />

chamber temperature at a set point with an accuracy of ±0.01°C and a wide range of<br />

temperature settings (5°C above ambient to 110°C).<br />

Figure 4.3. Left: picture of the calibration gas generator. Right: sketch of the calibration<br />

gas generator [1].<br />

88


The permeation tubes are small, inert capsules containing pure liquid elemental<br />

mercury or solid mercury chloride in a two- phase equilibrium between its gas phase and<br />

its liquid or solid phase, respectively. At a constant temperature, the device emits the<br />

compound through its permeable portion at a constant rate. Figure 4.4 illustrates the<br />

working principle of elemental mercury permeation tube.<br />

Figure 4.4. Sketch of the permeation tube with elemental mercury [2].<br />

The amount of mercury released <strong>from</strong> the tube is governed <strong>by</strong> the permeability of<br />

the material used for the tube, the length of the tube, and the temperature at which the<br />

tube is maintained. When the permeation rate at that temperature and the carrier flow rate<br />

are known, the concentration of the calibration stream can be estimated. Table 4.1 shows<br />

the specifications of the permeation tubes used in this project.<br />

Table 4.1. Specifications of the elemental mercury and mercury chloride permeation<br />

tubes used in this project.<br />

Elemental <strong>Mercury</strong> <strong>Mercury</strong><br />

mercury<br />

chloride chloride<br />

Working temperature (�C) 70 70 50<br />

Tube diameter (mm) 9.8 9.8 9.8<br />

Tube length (mm) 100 13 70<br />

Release rate (ng/min) 378 2445 823<br />

89


It is difficult to control the release rate of the mercury chloride tube to a similar<br />

value as the elemental mercury tube. The originally supplied mercury chloride tube has a<br />

release rate that is five times larger than the quoted tube at 70�C. After several trials<br />

VICI can provide a tube with a release rate of 823 ng/min at 50�C.<br />

4.1.3 Humidifier for water vapor addition<br />

The water vapor is not removed before mercury analyzer and mercury<br />

concentration is therefore measured on a wet basis. Thus it is important to get a precise<br />

control of the water addition.<br />

Survey and quotation of water vapor addition methods and equipments were<br />

carried out to get a reliable water vapor addition device. Current water addition methods<br />

include direct liquid injection, bubblers, porous membrane contactor and non-porous<br />

membrane contactor. The first three methods are already applied in <strong>CHEC</strong>. However,<br />

recent studies show that fluctuation is a main problem [3,4]. Due to the small amount of<br />

water injected in the direct injection method, it is difficult to control such small flow rate.<br />

Water condensation test of the bubbler evaporator shows that the water saturation<br />

fluctuates and cannot reach the calculated level [3,4]. This method is low cost, but has<br />

inaccuracies due to the temperature of the gas and liquid, operating pressure, and liquid<br />

level.<br />

Porous membrane contactor uses Nafion selective permeable membrane tube and<br />

water to continuously humidify gas streams. The producer suggests recirculation of the<br />

water at 4% of the gas flow [5]. It is difficult to find such a small pump that works at<br />

temperatures above 50�C. Flow with greater pressure needs to be flowing inside tubes to<br />

prevent tubing collapse. <strong>CHEC</strong> has made an evaporator using the membrane tube.<br />

However, a lot of humidity fluctuation has been observed and the reasons have not been<br />

identified [3,4]. Swedish company Cellkraft produces commercial evaporator using the<br />

membrane tube [6]. The system uses similar design of <strong>CHEC</strong>’s membrane evaporator,<br />

but has a water trap to remove water droplet. The water tank is heated <strong>from</strong> outside and<br />

there is an integrated heating tape for the gas lines. The water level in the tank is<br />

90


controlled and automatically filled. The producer can provide calibration curve and<br />

guarantee for water droplet free and working properly at small carrier gas flow rate of<br />

300 ml/min. The evaporator system without a dew point sensor costs about 64,000 DKK.<br />

American company Rasirc produces a water vapor addition unit using the nonporous<br />

membrane tube and integrated water temperature, level and dew point control<br />

[7,8]. The unit purifies and controls water vapor addition for a wide range of flow rates<br />

and process pressures. The membrane excludes particles, micro-droplet, volatile gases<br />

and other opposite charged species and ensures only water vapor is added. Figure 4.5<br />

illustrates the configuration of the Rasirc humidifier. Carrier gas to be humidified flows<br />

into the humidification unit. The water is heated to match the desired dew point<br />

temperature or humidification level. Water diffuses across the membrane to saturate the<br />

gas to be humidified. Temperature of the humidified gas is measured and fed back to a<br />

temperature controller to adjust the humidification level. Internal pressure control<br />

maintains independence <strong>from</strong> variations in downstream process pressures which allows<br />

operation into atmospheric and vacuum pressure environments. The unit with integrated<br />

humidity sensor costs about 50,000 DKK. The Rasirc humidifier has been widely used in<br />

the fabrication of semiconductors, nanotechnology, photovoltaics, fuel cells and other<br />

applications [8]. After comparison, a Rasirc RHS-IP-3-HT humidifier with an internal<br />

dew point sensor to regulate the dew point of the saturated gas was purchased.<br />

91


Figure 4.5. Sketch of the Rasirc humidifier [8]. Internal dew point sensor is not shown in<br />

the sketch.<br />

4.1.4 Low temperature furnace and fixed­bed reactor<br />

The low temperature oven is a three-zone electrically heated furnace and can heat<br />

up to 300�C. The oven has an internal diameter of 50 mm and a length of 450 mm. The<br />

heating tape for the top, middle and bottom zone is 170W/1m, 700W/4m, and 170W/1m,<br />

respectively. Three thermocouples are installed to measure the temperature at each zone<br />

and the heating is controlled <strong>by</strong> a center control box of the whole setup. The bottom of<br />

the oven is closed and the glass reactor has a u-shape. To avoid losses of sorbent powder<br />

in the gas stream, a downward flow is applied in the reactor. Quartz wool plugs are used<br />

at both ends of the sorbent bed. The top of the reactor and oven is heated <strong>by</strong> a heating<br />

tape.<br />

The temperature profiles at different setpoints are measured, as shown in figure<br />

4.6. The location of the sorbent bed is also illustrated in the figure. The measurements<br />

confirm that an isothermal reactor zone of about 300 mm is obtained with an estimated<br />

temperature uncertainty of ±2�C to the setpoints.<br />

92


Oven temperature ( o C)<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

Setpoint: 250 o C<br />

Setpoint: 200 o C<br />

Setpoint: 150 o C<br />

Setpoint: 120 o C<br />

0 50 100 150 200 250 300 350 400<br />

Distance to oven bottom (mm)<br />

93<br />

<strong>Sorbent</strong> bed<br />

position<br />

Figure 4.6. Temperature profile of the low temperature reactor oven.<br />

The glass reactor applied in this project is shown in figure 4.7. The reactor has an<br />

outer diameter of 20 mm (internal diameter of 18 mm) and a glass fiber porous plate to<br />

hold sorbent sample. With the dimension of the reactor shown in figure 4.7, the sorbent<br />

bed is located in the middle height of the low temperature furnace.<br />

Figure 4.7. Pictures with dimensions for the glass reactor.<br />

4.1.5 <strong>Mercury</strong> analysis system<br />

The mercury analysis system consists of a Lumex RA-915 AMFG elemental<br />

mercury analyzer, a gas distribution box and a oxidized mercury converter. Figure 4.8<br />

illustrates the sketch of the analysis system. A photo of the analysis system is presented<br />

in Figure 4.9.


Figure 4.8. Sketch of the mercury analysis system.<br />

Figure 4.9. Picture of the mercury analysis system with box open. The oxidized mercury<br />

converter is behind the mercury analyzer and gas distribution box.<br />

94


4.1.5.1 The Lumex analyzer<br />

The Lumex mercury analyzer has a measuring range of 0-500 �g/Nm 3 and<br />

automatic zero and span calibration functions. The lower detection limit is about 2<br />

�g/Nm 3 . All the gas lines and analytical cell inside the analyzer are heated. The analyzer<br />

can analyze gas of up to 30% water, and therefore no drying of the gas is needed. The<br />

analyzer determines the mercury concentration <strong>by</strong> Zeeman atomic absorption<br />

spectrometry using high frequency modulated polarized light. It is possible to measure<br />

only elemental mercury <strong>by</strong>passing the converter and only total mercury passing the gases<br />

through the converter, and to change the frequency of elemental and total mercury<br />

measurement switching through the sampling software.<br />

A block diagram of the analyzer is shown in figure 4.10 and a photo the analyzer<br />

internal parts are shown in figure 4.11. A membrane pump P draws flue gas <strong>from</strong> a<br />

sampling point via heated lines through a gas distribution box. There are four valves in<br />

the gas distribution box. The flue gas stream is either directed through the converter<br />

which reduces oxidized mercury to elemental mercury (valve V3 opened, valve V4<br />

closed) or is passed directly to the analytical cell AC, which is kept at a temperature of<br />

about 150°C. In the cell AC, which has an optical path length of about 0.4 m, a<br />

spectrometer determines the mercury concentration <strong>by</strong> Zeeman atomic absorption<br />

spectrometry using high frequency modulated polarized light (ZAAS-HFM). After<br />

leaving the cell, the gas is passing through a heated gas line and is then vented to a<br />

carbon trap before the ventilation. Temperature of the cell is constantly monitored <strong>by</strong><br />

temperature sensors T. The whole unit is controlled <strong>by</strong> an industrial panel PC, and<br />

powered <strong>by</strong> a power module PM.<br />

95


Figure 4.10. Block diagram of the mercury analyzer. The gas distribution box and<br />

converter are not integrated in the analyzer.<br />

96


Figure 4.11. Photo of the analyzer internal parts.<br />

The measurement principle of the analyzer is illustrated in figure 4.12 [9]. A<br />

mercury electronic discharge lamp is placed in a strong magnetic field H, <strong>by</strong> which the<br />

mercury resonance line at 254 nm is split into the three polarized Zeeman components �-,<br />

�, and �+. Only the �-components of the electromagnetic radiation will be registered <strong>by</strong><br />

the photo detector D. �- and �+ are separated <strong>by</strong> a polarization modulator. As long as<br />

mercury vapor is absent in the multipath cell, the intensities of both �-components are<br />

equal. When mercury is admitted to the cell, the difference in intensities between the two<br />

�-components increases as a function of the mercury concentration. As the spectral shift<br />

between the �-components is significantly smaller than the widths of molecular<br />

absorption bands and scattering spectra, background absorption <strong>by</strong> interfering<br />

compounds can be neglected.<br />

97


Figure 4.12. Illustration principle of the Zeeman atomic absorption spectrometry using<br />

high frequency modulated polarized light (ZAAS-HFM) [9].<br />

Sample gas connection to the analyzer is maintained at ambient pressure, with<br />

any excess flow vented to the atmosphere. Heated inlet and outlet lines are connected to<br />

the analytical cell inside the analyzer <strong>by</strong> means of 6 mm Swagelok-type fittings. The<br />

analyzer requires between 1 and 12 l/min of sample gas at all times and the flow can be<br />

controlled <strong>by</strong> the needle valve before the pump.<br />

4.1.5.2 Gas distribution box<br />

The gas distribution box contains valves for switching the gas and air to the<br />

analyzer and converter. The valves are controlled <strong>by</strong> the sampling software. The box is<br />

heat traced and isolated. There is a switch valve before the gas distribution box. Addition<br />

of air or sample gas to the analysis system can be selected.<br />

98


4.1.5.3 The oxidized mercury converters<br />

Two converters are used in this project. Originally Lumex supplied a low<br />

temperature converter for the red brass catalyst. Figure 4.13 shows a picture of the<br />

converter with a glass container loaded with red brass chips. The converter is designed to<br />

work at 180�C and the highest temperature is about 250�C. The glass container has an<br />

outer diameter of 20 mm and can hold about 20 g red brass chips.<br />

Figure 4.13. Picture of the low temperature converter with a glass container loaded with<br />

20 g red brass chips.<br />

Later a high temperature converter was used for the sulfite based converter<br />

material. The high temperature oven is a three-zone electrically heated furnace with a<br />

quartz reactor, which has an inner diameter of 17 mm and can hold up to 30 g sulfitebased<br />

converter pellets. The converter is a fixed-bed reactor made of quartz as shown in<br />

figure 4.14. The inner and bottom tubes of the reactor were removable. The sulfite-based<br />

pellets are placed on the porous quartz plate. The converter temperature is measured<br />

below the porous quartz plate <strong>by</strong> a thermocouple shielded in a quartz tube.<br />

99


Figure 4.14. Sketch of the high temperature converter with quartz reactor in the furnace.<br />

4.2 Converter and sorbent materials<br />

The red brass chips are obtained through Lumex. The idea of using red brass at<br />

low temperature is to bind free halogens in the flue gas and thus prevent back reaction<br />

into mercury halides and corrosion problem caused <strong>by</strong> SO2 oxidation at high<br />

temperatures [10]. Figure 4.15 shows picture of the red brass chips which have a<br />

thickness of about 0.5 mm and are rolled to a diameter of about 2 mm and a length of<br />

about 10 mm.<br />

100


Figure 4.15. Picture of the red brass chips supplied <strong>by</strong> Lumex.<br />

The sulfite converter material is prepared according to the work of Akiyama et al.<br />

[11]. Alumina pellets or zeolite pellets are first dried at 600�C for 24 h. Then the pellets<br />

are impregnated with water glass <strong>by</strong> forming a thin layer of water glass on the surfaces of<br />

the pellets. Sodium sulfate powders are added and mixed with the impregnated pellets.<br />

To inhibit crystallization of the salts, CaSO4 is added to the sulfite salts at a ratio of 50<br />

wt.%. About 15 to 45 wt.% of the sulfite salts and CaSO4 mixture are adhered almost<br />

uniformly to the thin layer of water glass. Immediately after mixing the product is placed<br />

in an oven and vacuum-dried at room temperature for 1 h, then it is vacuum-dried at<br />

150�C for 12 h. Figure 4.16 shows the picture of the prepared sulfite-based converter<br />

material. White powders of sodium sulfite and calcium sulfate are doped on the zeolite<br />

pellets with a diameter of 3 mm.<br />

101


Figure 4.16. Picture of the prepared sulfite-based converter material.<br />

To be able to quantify the oxidized mercury reduction efficiency, the oxidized<br />

mercury is produced <strong>by</strong> passing the flue gas with known concentration of elemental<br />

mercury to the reactor with 4 g catalyst for selective catalytic reduction of NOx. The<br />

catalyst piece was cut <strong>from</strong> a corrugated-type monolith obtained <strong>from</strong> Haldor Topsøe<br />

A/S. The catalyst is based on a fiber reinforced titania (TiO2) carrier, which is<br />

impregnated <strong>by</strong> vanadium (V2O5) and tungsten (WO3). The vanadium loading (3 wt.%<br />

V2O5) was uniformly distributed across the wall thickness of the monolith [12,13]. The<br />

efficiency of the converter is evaluated <strong>by</strong> the recovery extent of measured total mercury<br />

through the SCR catalyst and converter compared to the elemental mercury level at the<br />

inlet of the SCR catalyst.<br />

The most investigated sorbents in this project is Darco Hg activated carbon,<br />

which is a commercial lignite based powdered activated carbon and is developed for<br />

heavy metal removal <strong>from</strong> incinerators and power plants. The Darco Hg carbon has a<br />

bulk density of 0.51 g/cm 3 and a surface area of about 600 m 2 /g. The average particle<br />

102


size is 16 �m and the porosity is about 58% [14-22]. Properties of other sorbents are<br />

presented in the chapter of sorbent screening.<br />

4.3 Flue gas composition<br />

The total flow rate through the reactor is 2.75 Nl/min of which about 2 Nl/min is<br />

passed through the analyzer. The typical composition of the simulated cement kiln flue<br />

gas applied in this work includes 21% CO2, 6% O2, 1% H2O, 10 ppmv HCl, 1000 ppmv<br />

NO, 23 ppmv NO2, and 1000 ppmv SO2. The applied mercury concentration is about<br />

160-180 µg/Nm 3 <strong>by</strong> keeping the elemental mercury and mercury chloride source at 70ºC<br />

and 50ºC, respectively, and using 0.275 Nl/min nitrogen as carrier gas. The water level in<br />

the simulated flue gas is lower than real level in the cement kiln flue gas. This is due to<br />

the limitation of the humidifier. Although the humidifier can add water vapor relatively<br />

precisely, it is not robust. The membrane can be easily broken and the unit cannot stand<br />

high over pressure. After short period of operation, the unit was repaired twice <strong>by</strong><br />

changing the membrane and installing of a pressure release valve. It seems that the unit<br />

can run properly only for short period. Another reason is the fluctuation measurement of<br />

the mercury analyzer with more than 5% water in the simulated flue gas. Therefore it is<br />

decided to use 1% water in most of the tests <strong>by</strong> adding water through a bubbling bottle.<br />

In few cases high water contents are used to cover a wide range of water level in the<br />

simulated flue gas.<br />

4.4 <strong>Sorbent</strong> load in fixed­bed test<br />

For the applied reactor in this project, at least 500 mg activated carbon is need to<br />

form a fixed-bed covering the cross area of the reactor. The amount of the sorbent sample<br />

is determined <strong>by</strong> the sample saturation time. Literature reported that approximately 600<br />

mg of sorbent initially was placed into the reactor; however, the samples were reduced<br />

<strong>from</strong> 600 mg to between 100 and 150 mg after it was observed that extremely long<br />

durations (up to weeks) would be required to saturate the larger quantity of sorbent [23].<br />

To avoid channeling the sorbent sample is usually mixed with some inert materials such<br />

103


as sand and glass beads. Application of dilution <strong>by</strong> sand powder can also accelerate the<br />

tests. The reaction gas flows downward through the bed to minimize the chance of<br />

selective flow or channeling through the bed. Reactor sizes and sample dilutions applied<br />

in the literature are reviewed and summarized in table 4.2.<br />

Table 4.2. Reported reactor sizes, flow rates and sorbent sample loads in the literature.<br />

<strong>Sorbent</strong> loading Reactor<br />

size<br />

ID (mm)<br />

50 mg fly ash mixed with 3 g<br />

glass beads, 2.5 mm bed<br />

thickness, additional 57.5 mm<br />

glass beads upstream to the bed<br />

for better flow distribution<br />

0.2-0.6 g sample (copper<br />

compound based sorbents and<br />

commercial proprietary sorbents)<br />

held <strong>by</strong> a glass wool plug, 16 mm<br />

bed thickness<br />

20-30 mg sorbent (Norit FGD<br />

carbon and functionalized silica)<br />

in 6 g silica, glass fiber filter at<br />

two ends<br />

0.61 g Darco G60 carbon with 3 g<br />

glass beads, 4 mm bed length<br />

20 mg activated carbon mixed<br />

with 1 g sand<br />

5 mg carbon on 3 g glass beads, 4<br />

mm bed thickness<br />

20 mg carbon in 10 g sand,<br />

supported <strong>by</strong> quartz wool<br />

10-100 mg sorbent mixed with 2<br />

g sand, bed thickness of about 5<br />

mm, quartz wool at two ends<br />

104<br />

Flow rate<br />

(Nl/min)<br />

Superficial<br />

velocity (cm/s)<br />

@150�C<br />

35 3.2 8.7 [24]<br />

4 0.15 30.8 [25]<br />

12.7 0.91 18.4 [26]<br />

35 2.8 7.5 [27]<br />

6.35 0.17 13.8 [28]<br />

35 4.2 7.52 [29]<br />

12.7 1 20.3 [30]<br />

References<br />

18 2.75 22.1 This work


4.5 Experimental procedure<br />

An experimental procedure is developed for sorbent tests and measures are taken to<br />

avoid mercury accumulation in the system. The detailed procedure is as following:<br />

� The mercury source is maintained at the operating temperature with carrier gas<br />

through all the time.<br />

� Increase the sulfite converter temperature setpoints <strong>from</strong> 100�C to 500�C.<br />

� Check the temperatures of mercury source, heated lines, low temperature oven,<br />

gas distribution box and analyzer, set the low temperature reactor oven to desired<br />

temperature.<br />

� Weight desired amount of sorbent and mix with 2g sand powder, load the sample<br />

to the glass reactor.<br />

� Check and measure the flow rates of different gases.<br />

� When the converter temperature reaches 500�C for about 30 min, change the<br />

mercury analyzer measurement mode <strong>from</strong> elemental to total mercury<br />

measurement.<br />

� Switch the valve before the glass reactor to <strong>by</strong>passing the reactor position; add<br />

gases except mercury to the system.<br />

� Add the gases to the mercury analyzer to check whether there is some mercury<br />

accumulated in the hot system; if there is some mercury detected, wait the<br />

mercury reading decreases to zero value and then switch mercury source to the<br />

system.<br />

� Start the test and data sampling, make note in the sampling program, measure the<br />

mercury inlet concentration for at least 30 min to ensure that the glass reactor is<br />

heated for about 1 h and stable reactor temperature is obtained.<br />

� Switch the gases to fixed sorbent bed, make note in the sampling program.<br />

� After full mercury breakthrough is observed for 30 min, switch air to the analyzer<br />

system while keeping simulated flue gas to the reactor, and switch the mercury<br />

analyzer to measure elemental mercury.<br />

105


� When the elemental mercury measurement mode is ready, switch the simulated<br />

flue gas to the analysis system.<br />

� After 20 min, switch the simulated flue gas to <strong>by</strong>pass the reactor and measure the<br />

inlet mercury concentration for another 20 min.<br />

� Stop the test, switch air to the analysis system, switch mercury source to the<br />

carbon trap and ventilation, and stop other gases.<br />

� Remove the reactor <strong>from</strong> the oven and remove the sample after the reactor is<br />

cooled, store the sample in a closed plastic bottle with label.<br />

� At the end of the day, turn all gases off except nitrogen and H2O when elemental<br />

mercury source is applied and add also HCl when mercury chloride is applied to<br />

flush the system overnight. Decrease the converter temperature to 100�C, to<br />

ensure the mercury analyzer is in elemental mercury measurement mode and air<br />

is added to the analysis system.<br />

� Always keep the whole reactor system hot.<br />

4.6 <strong>Sorbent</strong> characterization<br />

4.6.1 Scanning electron microscopy<br />

Scanning electron microscopy with energy-dispersive X-ray spectroscopy (SEM-<br />

EDX) analysis is used to understand mercury capture mechanisms <strong>by</strong> different powder<br />

sorbent. The main goals of the SEM-EDX analysis is to study the sorbents’ topography<br />

(surface features), morphology (shape and size), and composition. Morphology study<br />

will be used to identify particle agglomeration and compare with particle size<br />

measurement.<br />

The SEM-EDX analysis is conducted at Center for Electron Nanoscopy, DTU.<br />

Micrographs and EDX analysis of carbon samples are carried out using Quanta FEGSEM<br />

200F. The carbon samples are not coated, while the non-carbon samples are coated with<br />

14 nm carbon and analyzed on Inspect ‘S’ SEM. The FEGSEM is a high resolution<br />

flexible microscope with field emission gun (FEG). The Inspect ‘S’ is a scanning<br />

electron microscope with a tungsten filament electron source. To support the surface<br />

106


information obtained <strong>by</strong> the imaging detectors both SEMs are equipped with Oxford<br />

Instruments INCA EDX analyzer which gives possibilities to analyze chemical elements<br />

position on the sample surface in single spots or over a selected area. The microscope can<br />

operate in as well high- and low vacuum as in environmental mode. A typical working<br />

distance of 10 mm is applied.<br />

A thin layer of sample powders is spread on a double sided conductive carbon<br />

table. If particles are piled on each other charge-up easily takes place, causing them to<br />

move during observation. A low accelerating voltage of 5 kV is applied during imaging<br />

on the FEGSEM to obtain detailed information on the particle surface and minimize<br />

specimen charging problem.<br />

4.6.2 Particle size distribution<br />

The particle size distributions of the sorbent powders are analyzed <strong>by</strong> a Malvern<br />

Mastersizer S analyzer using laser diffraction. The technique of laser diffraction is based<br />

around the principle that particles passing through a laser beam will scatter light at an<br />

angle that is directly related to their size. Large particles scatter light at narrow angles<br />

with high intensity [31], whereas small particles scatter at wider angles but with low<br />

intensity. The analyzer consists of a laser to provide a source of coherent, intense light of<br />

fixed wavelength, a sample presentation system to ensure that the material under test<br />

passes through the laser beam as a homogeneous stream of particles in a known,<br />

reproducible state of dispersion, and a series of detectors which are used to measure the<br />

light pattern produced over a wide range of angles.<br />

Based on previous analysis experience at <strong>CHEC</strong>, the samples are dispersed either<br />

in ethanol or distilled water for one minute before measurement to avoid agglomeration<br />

and the result is the average of 5 measurements.<br />

107


4.6.3 Analysis of mercury in sorbent<br />

<strong>Mercury</strong> content in the exposed sorbent is analyzed <strong>by</strong> a DMA-80 analyzer <strong>from</strong><br />

Milestone at FLSmidth Dania lab. 100 mg of powder sample is first weighted and loaded<br />

in a sample boat. The boats are transported automatically into the furnace. The sample is<br />

initially dried and then thermally decomposed in a continuous flow of oxygen.<br />

Combustion products are carried off and further decomposed in a hot catalyst bed [32].<br />

<strong>Mercury</strong> vapors are trapped on a gold amalgamator and subsequently desorbed for<br />

quantization. The mercury content is determined using an atomic absorption<br />

spectrophotometer at 254 nm. The instrument determines the absolute amount of Hg and<br />

then the software calculates its concentration in the sample.<br />

4.7 References<br />

[1] VICI Metronics Inc., Dynacalibrator ® Model 150 calibration gas generator brochure, 2008.<br />

[2] VICI Metronics Inc., Dynacal permeation tubes, 2011.<br />

[3] B. Maribo-Mogensen and J. Christensen, Internal steam reforming in solid oxide fuel cells,<br />

Bachelor, Department of Chemical and Biochemical Engineering, Technical University of<br />

Denmark, 2008.<br />

[4] A.F. Castells. Steam reforming kinetics over Ni-YSZ used as anode material for solid fuel<br />

cells, Master, Department of Chemical and Biochemical Engineering, Technical University of<br />

Denmark, 2009.<br />

[5] Perma Pure, MH TM -series humidifier user manual, http://www.permapure.com /PDF%20Files<br />

/MH%20Manual.pdf, accessed March/20, 2009.<br />

[6] Cellkraft AB, P-series humidifier manual, http://www.cellkraft.se/humidity_and_steam/P-<br />

Series.html, accessed March 20, 2009.<br />

[7] RASIRC, RASIRC Rainmaker TM humidification system manual, http://www.rasirc.com<br />

/resources/datasheets/datasheet_RASIRC_RainMaker_HS.pdf, accessed March 20, 2009.<br />

[8] Jeffrey Spiegelman, RainMaker humidification system for precise delivery of water vapor<br />

into atmospheric and vacuum applications, http://www.rasirc.com/resources/whitepapers<br />

/whitepaper_RHS.pdf, accessed March 20, 2009.<br />

[9] Lumex Ltd, RA-915 AMFG automatic mercury monitor for flue gas operational manual,<br />

2009.<br />

[10] R. Kanefke, H. Köser, B. Vosteen, F. Kristina, B. Frank, S. Raik, Method for the production<br />

of elemental mercury <strong>from</strong> mercury compounds, patent WO 2008/064667 A2, 2008.<br />

108


[11] S. Akiyama, J. Kato, F. Koga, K. Ishikawa, Catalysts for reducing mercury, a mercury<br />

conversion unit, and an apparatus for measuring total mercury in combustion exhaust gas <strong>by</strong><br />

using the same, patent US2007/0232488 A1, 2007.<br />

[12] Y. Zheng, A.D. Jensen, J.E. Johnsson, Deactivation of V2O5-WO3-TiO2 SCR catalyst at a<br />

biomass-fired combined heat and power plant, Applied Catalysis B: Environmental. 60 (2005)<br />

253-264.<br />

[13] Y. Zheng, A.D. Jensen, J.E. Johnsson, J.R. Thøgersen, Deactivation of V2O5-WO3-TiO2<br />

SCR catalyst at biomass fired power plants: Elucidation of mechanisms <strong>by</strong> lab- and pilot-scale<br />

experiments, Applied Catalysis B: Environmental. 83 (2008) 186-194.<br />

[14] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R. Chang,<br />

Preparation and evaluation of coal-derived activated carbons for removal of mercury vapor <strong>from</strong><br />

simulated coal combustion flue gases, Energy & Fuels. 12 (1998) 1061-1070.<br />

[15] M. Rostam-Abadi, S.G. Chen, H.C. Hsi, M. Rood, R. Chang, T.R. Carey, B. Hargrove, C.<br />

Richardson, W. Rosenhoover, F. Meserole, Novel vapor phase mercury sorbents, Proceedings of<br />

the EPRI-DOE-EPA Combined Utility Air Pollutant Control, Washington, DC, Aug 25–29,1997.<br />

[16] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang and F.B. Meserole, Factors<br />

affecting mercury control in utility flue gas using sorbent injection, Proceedings of the Air &<br />

Waste Management Association's 90th Annual Meeting, Toronto, Ontario, Canada, June 8-13,<br />

1997.<br />

[17] B. Ghorishi and B.K. Gullett, Fixed-bed control of mercury: Role of acid gases and a<br />

comparison between carbon-based, calcium-based, and coal fly ash sorbents, Proceedings of the<br />

EPRI/DOE/EPA Combined Utility Air Pollutant Control Symposium, Washington, DC, August<br />

25–29,1997.<br />

[18] S.B. Ghorishi and C.B. Sedman, Combined mercury and sulfur oxides control using<br />

calcium-based sorbents, Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant<br />

Control Symposium, Washington, DC, August 25–29, 1997.<br />

[19] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption <strong>by</strong> activated<br />

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy Conference,<br />

Research Triangle Park, NC, 22-25 April, 1997.<br />

[20] N. Hutson, C. Singer, C. Richardson, J. Karwowski and C. Sedman, Practical applications<br />

<strong>from</strong> observations of mercury oxidation and binding mechanisms, EPRI-DOE-EPA-AWMA<br />

Combined Power Plant Air Pollutant Control Mega Symposium, Washington DC, August 30 -<br />

September 2, 2004.<br />

[21] Norit Americas Inc., Datasheet of Darco FGD powdered activated carbon, 2008.<br />

[22] R. Bhardwaj. Impact of temperature and flue gas components on mercury speciation and<br />

uptake <strong>by</strong> activated carbon sorbents. Master thesis. University of Pittsburgh, 2007.<br />

[23] D.J. Hassett, K.E. Eylands, <strong>Mercury</strong> capture on coal combustion fly ash, Fuel. 78 (1999)<br />

243-248.<br />

109


[24] D. Karatza, A. Lancia, D. Musmarra, Fly ash capture of mercuric chloride vapors <strong>from</strong><br />

exhaust combustion gas, Environ. Sci. Technol. 32 (1998) 3999-4004.<br />

[25] J.W. Portzer, J.R. Albritton, C.C. Allen, R.P. Gupta, Development of novel sorbents for<br />

mercury control at elevated temperatures in coal-derived syngas: results of initial screening of<br />

candidate materials, Fuel Processing Technology. 85 (2004) 621-630.<br />

[26] J.Y. Lee, Y. Ju, T.C. Keener, R.S. Varma, Development of cost-effective noncarbon<br />

sorbents for Hg 0 removal <strong>from</strong> coal-fired power plants, Environ. Sci. Technol. 40 (2006) 2714-<br />

2720.<br />

[27] D. Karata, A. Lancia, D. Musmarra, F. Pepe, Adsorption of metallic mercury on activated<br />

carbon, Symposium (International) on Combustion. 26 (1996) 2439-2445.<br />

[28] G. Skodras, I. Diamantopoulou, G. Pantoleontos, G.P. Sakellaropoulos, Kinetic studies of<br />

elemental mercury adsorption in activated carbon fixed bed reactor, Journal of Hazardous<br />

Materials. 158 (2008) 1-13.<br />

[29] D. Karatza, A. Lancia, D. Musmarra, C. Zucchini, Study of mercury absorption and<br />

desorption on sulfur impregnated carbon, Experimental Thermal and Fluid Science. 21 (2000)<br />

150-155.<br />

[30] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors affecting<br />

mercury control in utility flue gas using activated carbon, Journal of the Air & Waste<br />

Management Association. 48 (1998) 1166.<br />

[31] Malvern Instruments Ltd, Understanding how laser diffraction works,<br />

http://www.malvern.com/LabEng/technology/laser_diffraction/laser_diffraction.htm, accessed<br />

January/3, 2011.<br />

[32] Milestone Srl, DMA-80: Principle of operation, http://www.milestonesrl.com<br />

/analytical/products-mercury-determination-dma-80-and-dma-803-principle-of-operation.html,<br />

accessed January 6, 2011.<br />

[33] A. Ocklind. Calculation <strong>from</strong> gas dewpoint to water content, 2009.<br />

[34] R.H. Perry, D.W. Green, J.O. Maloney, (Eds.), Perry’s chemical engineers’ handbook, 7th<br />

ed. The McGraw-Hill Companies, Inc., 1997.<br />

Appendix<br />

4A Check of mercury analyzer<br />

Since many valves and fittings are used in the mercury analysis system, it is<br />

important to make a leakage test of the system. Lumex service technicians performed the<br />

leakage test. The Lumex analyzer cannot stand high pressure and the system therefore<br />

cannot be tested <strong>by</strong> plugging the system and checking the pressure change. Instead, flow<br />

rates at the gas distribution box, converter, and analyzer are measured simultaneously<br />

110


using rotating flow meters and a separate pump <strong>from</strong> Lumex. Flow rates measured <strong>by</strong> the<br />

portable rotating flow meters at the analysis system inlet and outlet and flow rate<br />

measured <strong>by</strong> the integrated flow meter in the analyzer are the same at both cold and hot<br />

condition, indicating that there is no leakage in the Lumex analysis system. The function<br />

of heating the gas panel and lines to avoid mercury accumulation in the system was also<br />

checked. The test was conducted <strong>by</strong> first passing mercury contained gas through the hot<br />

panel and then stopping mercury addition and the hot panel was flushed <strong>by</strong> nitrogen. No<br />

mercury was detected in the flushing nitrogen, confirming that no mercury was<br />

accumulated in the lines.<br />

The required flow rate for the mercury analyzer was determined. The needle<br />

valve for controlling the flow rate was changed <strong>from</strong> <strong>by</strong>pass line between the analytical<br />

cell and the pump to on the line between the analytical cell and the pump. The flow rate<br />

is better controlled; however, it cannot be reduced to as low as l l/min. The flow rate<br />

through the analyzer is set to 2 l/min. The analyzer requires overflow. Therefore a 3<br />

l/min gas through the reactor is used and the excess flow will <strong>by</strong>pass the analyzer and<br />

exhaust through the ventilation.<br />

Lumex service technician brought a portable mercury analyzer RA-915+, which<br />

is the same as the Haldor Topsøe analyzer except that it has only a single analytical cell.<br />

These two analyzers were compared <strong>by</strong> making tests at Haldor Topsøe’s mercury<br />

research facility. Using a single analytical cell at analyzers, the measured mercury<br />

concentration <strong>by</strong> Lumex and Haldor Topsøe analyzer was 9936 and 10080 ng/m 3 ,<br />

respectively. <strong>Mercury</strong> concentration measured <strong>by</strong> Haldor Topsøe analyzer using multiple<br />

cells was 10200 ng/m 3 . This indicates that the Lumex portable analyzer works properly<br />

and seems reliable.<br />

Comparison of Lumex AMFG monitor, which is used in present project, with<br />

Lumex portable analyzer was conducted <strong>by</strong> running these analyzers simultaneously. The<br />

measured mercury concentrations <strong>by</strong> the portable analyzer are about 10% higher than the<br />

AMFG, as shown in table 4.3.<br />

111


Table 4.3. Comparison of mercury measurement <strong>by</strong> the portable Lumex RA-915+ and<br />

the Lumex AMFG analyzer applied in this project.<br />

Concentration measured <strong>by</strong><br />

portable analyzer, �g/Nm 3<br />

Concentration measured<br />

<strong>by</strong> AMFG analyzer,<br />

�g/Nm 3<br />

249 220 1.13<br />

217 201 1.08<br />

153 137 1.11<br />

112<br />

Ratio of portable/AMFG<br />

Lumex said that the difference is caused <strong>by</strong> different conditions used during<br />

calibration at the factory and test at <strong>CHEC</strong> lab. The gas lines before the analyzer at the<br />

factory was not heated and the analytical cell was tested at a temperature which was<br />

50�C lower than at <strong>CHEC</strong> lab. Since comparison with Haldor Topsøe analyzer shows<br />

that the portable analyzer works properly, the AMFG is calibrated using linearity test. A<br />

factor of 1.10 was used as the calibration coefficient.<br />

4B Water addition verification<br />

To verify the water addition stability and accuracy condensation tests are<br />

conducted. The configuration of the test system is shown in figure 4.17. A water column<br />

of a height of about 500 mm is hanged 1 m above the humidifier as water source. The top<br />

of the water column is open. A stop valve is installed below the water column to allow<br />

disconnection of the water column for weight measurement. Nitrogen is used as the<br />

carrier gas. The water level in the humidifier is always kept at full level <strong>by</strong> a liquid level<br />

switch and a micro pump. The gas line between the humidifier and the condensation<br />

water bottle is heated at 110�C to avoid water condensation in the gas line. The water<br />

bath temperature is kept at 4�C. The condensation bottle is filled with some water to<br />

enhance the heat exchange and capture of water. It takes about 70 min for the humidifier<br />

to reach the desired dew point. During this period the gas is <strong>by</strong>passed to the water bath<br />

and passes through a water bottle before ventilated. When the desired dew point is<br />

reached and stabilized, the water tank is disconnected for weight measurement. The gas<br />

is still running through the humidifier and water is continuously added. During the


measuring of the water tank weight about 80 mm water inside the ¼’’ Teflon tube is<br />

added and the corresponding weight is about 1 g. When the water tank is connected back<br />

there will be an 80 mm air plug inside the Teflon tube. This means that 1 g water is not<br />

added and should be deducted <strong>from</strong> the theoretical calculation for comparison with the<br />

measured water addition. The amount of water added can be evaluated either <strong>by</strong><br />

measuring weight change of the water tank or weight of water collected in the<br />

condensation bottles.<br />

Figure 4.17. Sketch of the condensation test system.<br />

When the dew point is reached and stabilized, the dew point reading of the<br />

saturated gas is changing about � 0.1�C <strong>from</strong> the set point. This indicates that the relative<br />

humidity of the gas is quite stable and close to 100%.<br />

To calculate the amount of water added the saturated water vapor pressure at<br />

given dew point should be calculated first. Table and figure of saturated water vapor<br />

pressure can be readily found in text book. The saturated water vapor pressure can be<br />

calculated <strong>by</strong> using the empirical expression <strong>from</strong> Cellkraft, which is a Swedish<br />

membrane humidifier produce [6,33]:<br />

113


lnP �10.4592 �4.05�10 T �4.18�10T�3.69�10 T �1.02�10 T �8.65�10 T<br />

� � � � � � � � � �<br />

(4.1)<br />

where the water saturated water vapor pressure Ps is in MPa, T is in K.<br />

Partial pressures of the gases are proportional to the gas flow rates:<br />

P HO F<br />

2 HO 2 �<br />

P F<br />

(4.2)<br />

�3 �5 2 �7 3 �9 4 �13<br />

5<br />

s<br />

�16 6<br />

9.04 10 T<br />

�18 7<br />

2.00 10 T<br />

�22 8<br />

7.79 10 T<br />

�25<br />

9<br />

1.91 10 T 3968.06 /( T 39.57)<br />

N2 N2<br />

Ptot � PH2O� PN<br />

(4.3)<br />

2<br />

where T is the desired dew point of the gas, 82 �C, 355.15 K<br />

�3 �5 2 �7<br />

3<br />

lnPs<br />

�10.4592 �4.05�10 �355.15 �4.18�10�355.15 �3.69�10 �355.15<br />

� � � �<br />

�1.02�10 �355.15 �8.65�10 �355.15 �9.04�10 �355.15 �2.00�10�355.15 �22 8 �25<br />

9<br />

�7.79�10 �355.15 �1.91�10 �355.15 �3968.06 /( �355.15 �39.57) ��2.9686<br />

(4.4)<br />

and Ps � 513.76 mbar<br />

With room temperature of 25�C, carrier nitrogen flow, 0.3l/min, equals 0.275 Nl/min,<br />

9 4 13 5 16 6 18 7<br />

one can calculate:<br />

Water addition rate (g/min):<br />

(18 / 22.4) �0.275 �(513.76<br />

/1013.25)<br />

� 0.227<br />

1� 513.76 /1013.25<br />

Water flow rate (Nl/min):<br />

0.275 �(513.76<br />

/1013.25)<br />

� 0.283<br />

1� 513.76 /1013.25<br />

The water vapor pressure can also be calculated using Antoine equation [34]:<br />

Ps<br />

log10 �8.07131�1730.63/(233.426 � T )<br />

(4.5)<br />

where Ps is in Torr (1 mmHg), T is in �C.<br />

Ps<br />

log10 �8.07131�1730.63/(233.426 �82) � 2.5847<br />

then P � 512.35 mbar<br />

s<br />

Calculated water addition rate (g/min):<br />

(18 / 22.4) �0.275 �(512.35<br />

/1013.25)<br />

� 0.226<br />

1� 513.76 /1013.25<br />

Water flow rate (Nl/min):<br />

114


0.275 �(512.35<br />

/1013.25)<br />

� 0.281<br />

1� 513.76 /1013.25<br />

Table 4.4 presents the calculated water addition rates and flow rates at different<br />

dew points. Calculations using Cellkraft equation and Antoine equation give almost the<br />

same results.<br />

Table 4.4. Calculated water addition rate and flow rate at different dew points.<br />

Dew point (�C) Cellkraft equation Antoine equation<br />

H2O addition H2O flow rate H2O addition H2O flow rate<br />

rate (g/min) (Nl/min) rate (g/min) (Nl/min)<br />

82 0.227 0.283 0.226 0.281<br />

77 0.156 0.194 0.155 0.193<br />

72 0.112 0.139 0.111 0.138<br />

63 0.064 0.080 0.064 0.080<br />

4 0.002 0.002 0.002 0.002<br />

The comparison between the measured water addition and calculated water<br />

addition is presented in table 4.5. For the first two tests only one condensation bottle is<br />

used and about 1/3 of the bottle is filled with water. The water collected in the<br />

condensation bottle is only 50-60% of the calculated value. This is probably due to the<br />

short gas residence time in the condensation bottle and small amount of water filled in<br />

the condensation bottle. The amount of water added <strong>by</strong> water tank weight measurement<br />

is reasonably in agreement with the calculation. Later more water is filled in the bottle<br />

and two or four bottles are connected in series. Then the amounts of water added <strong>by</strong><br />

measuring the water tank weight change and water collected in the bottle are similar and<br />

about 90-95% of the calculated values. For the four bottle in-series tests the weight<br />

change of each bottle is measured. Almost all the water is collected in the first bottle and<br />

the water collected in other bottles is negligible. The test results show that the humidifier<br />

works well.<br />

115


Table 4.5. Comparison between the measured water addition and calculated water<br />

addition.<br />

Dew point (�C) 82 82 82 82 82 82<br />

N2 flow (ml/min) 300 300 300 300 300 400<br />

Room temp. (�C) 24.9 24.7 26.3 25.7 24.9 24.9<br />

Duration (min) 450 318 368.5 240 143 120<br />

Calculated water addition (g) 100.25 70.55 82.17 52.52 28.89 35.12<br />

Measured water addition<br />

through tank weight (g)<br />

102.4 63.5 74.10 47.7 27.6 31.8<br />

Measured/calculated (%) 102 90 90.5 90.8 95.5 90.6<br />

Measured water addition<br />

through collection in water<br />

bottle (g)<br />

61.48 36.27 73.99 47.08 - 32.32<br />

Water bottle number 1 1 2 2 4 2<br />

Measured/calculated (%) 61.4 51.4 90.3 89.6 - 92<br />

116


Dynamic measurement of mercury adsorption and<br />

oxidation on activated carbon in simulated cement<br />

117<br />

5<br />

kiln flue gas<br />

This chapter starts with a review of available gaseous mercury measurement<br />

technologies. Pros and cons of the technologies will be discussed. Then tests of the<br />

commercial red brass converter in simulated cement kiln flue gas are presented. Finally<br />

development of sulfite-based converter for oxidized mercury reduction in simulated<br />

cement kiln flue gas is reported. Suggestions for practical applications of the sulfite<br />

converter in both lab and cement plants are presented.<br />

5.1 Review of gaseous mercury measurement technology<br />

Presently, the accepted methods for mercury measurement are wet-chemistry<br />

procedures such as EPA methods 29 and 101 A for total mercury measurement and the<br />

Ontario Hydro method for total mercury and speciation measurement [1,2]. These<br />

methods often have 2-week or more turn-around time for results. The sorbent trap<br />

method was developed to shorten the analysis time of the collected samples. These<br />

methods can only provide an average mercury concentration over a 1-2 hour period, and<br />

cannot characterize the variability in mercury emissions due to process and operating<br />

changes with time.<br />

To obtain an understanding of the process of mercury removal <strong>by</strong> sorbent<br />

injection upstream of a fabric filter, it is necessary to study them under more controlled<br />

conditions such as in a laboratory scale setup, for example using a fixed bed reactor. In<br />

such experiments it is also necessary to use a continuous emission monitor (CEM) to


obtain knowledge of uptake of total and speciated mercury in simulated flue gas to fully<br />

evaluate the control technologies under development. Fixed-bed experiments have been<br />

used <strong>by</strong> many laboratories to test the relative effectiveness of different mercury sorbents.<br />

The critical assumption of this experimental method is that the performance of a sorbent<br />

over a long exposure time (hours) reflects the filtration/reaction on bags, where the<br />

sorbent contacts the flue gas for about 25 minutes [1]. Therefore, it is preferable to verify<br />

this assumption <strong>by</strong> supplementing the final mercury content data with breakthrough data<br />

obtained using a CEM.<br />

Real- or near-real-time mercury emission measurement can in principle be<br />

obtained depending on the applied detection method. Generally, real-time measurements<br />

can be achieved <strong>by</strong> analyzers using cold vapor atomic absorption spectroscopy. The cold<br />

vapor atomic fluorescence spectrophotometer collects mercury in flue gas on alternating<br />

gold traps and thermally desorbs the mercury in about five minute intervals allowing for<br />

semi-continuous measurements [3].<br />

Available commercial mercury analyzers can only measure elemental mercury.<br />

The measurement of total mercury as well as mercury speciation can only be achieved<br />

indirectly. For this purpose, all oxidized mercury is reduced to its elemental form <strong>by</strong> a<br />

converter system. It should be noted that the technology for the analytical part of the<br />

detection system is somehow matured and provides accurate and sensitive detection of<br />

elemental mercury [3]. The conversion unit, on the other hand, is a subject of continuing<br />

research and improvement efforts [2].<br />

The converter can be based either on wet chemistry or dry conversion. In a wetchemistry<br />

conversion unit the Hg 2+ is converted to Hg 0 via a liquid phase reducing agent,<br />

often stannous chloride (SnCl2), prior to entering the analysis unit. There is interference<br />

with SO2, which can affect the reduction of Hg 2+ when using SnCl2 [2,4,5]. Furthermore,<br />

the wet chemicals themselves are very corrosive and need frequent replenishment.<br />

For on-line measurements, a dry converter is usually preferred over a wet<br />

chemical converter for the reasons mentioned above [6]. Several dry converter types<br />

exist. In a pure thermal conversion unit, the flue gas is heated to reduce all Hg 2+ to Hg 0 .<br />

118


However, reoxidation of the reduced mercury before reaching the analysis unit is a<br />

concern. Furthermore, the required temperature depends on the HCl concentration in the<br />

gas. In case of a thermocatalytic conversion, the potential short lifetime of the catalyst is<br />

an issue due to the possible poisoning <strong>by</strong> acidic gases in the sample gas [2,4,5].<br />

Compared to mercury measurements in power plants and waste incinerators, there<br />

is a lack of experience related to continuous measurement of mercury emissions <strong>from</strong><br />

cement kilns. Furthermore, the experience gained <strong>from</strong> power plant and waste incinerator<br />

may not be applied directly to cement plant due to the different process conditions and<br />

flue gas compositions [7]. In this work a commercial red brass converter, which is<br />

developed for application in waste incinerators, is tested in simulated cement kiln flue<br />

gas and an improved sodium sulfite-based converter is developed and tested.<br />

5.2 Performance test of the mercury analyzer<br />

The analyzer has an internal mercury source for span calibration. However, the<br />

span calibration is conducted at an elemental mercury concentration of about 16 �g/Nm 3 ,<br />

which is much lower than typical mercury concentration of about 180 �g/Nm 3 applied in<br />

this project. If the linearity of the analyzer is poor then the measured mercury<br />

concentration at typical mercury levels in this project could be wrong. To check the<br />

analyzer linearity some tests were conducted. The carrier nitrogen flow rate through the<br />

mercury source was kept at 275 Nml/min. Firstly, the mercury concentration in the outlet<br />

gas <strong>from</strong> the mercury source was calculated using the measured mercury concentration in<br />

the mixed gas and applied flow rates. Then part of the gas <strong>from</strong> the outlet of mercury<br />

source was <strong>by</strong>passed to ventilation and more nitrogen was added to the empty reactor to<br />

dilute the mercury-contained gas and keep the total flow through the reactor at 2.75<br />

Nl/min. Figure 5.1 shows the comparison between the measured and calculated mercury<br />

concentration in the mixed gas. Very good agreement is obtained between the measured<br />

and calculated mercury concentration, confirming that the linearity of the analyzer is<br />

good.<br />

119


Measureed Hg concentration (�g/Nm3)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

0 50 100 150 200 250<br />

Calculated Hg concentration (�g/Nm3)<br />

Figure 5.1. Linearity of the Lumex mercury analyzer. Measured elemental mercury<br />

concentration is compared with the calculated values in the range of 0-250 �g/Nm 3 .<br />

The effects of different gases on elemental mercury measurement were<br />

investigated <strong>by</strong> adding gases separately. Figure 5.2 shows the measured mercury<br />

concentration under different conditions. Nitrogen, water, and CO2 were used as baseline<br />

gas. Further addition of O2, SO2, NOx and HCl step <strong>by</strong> step to the baseline gases gives<br />

the same mercury level. This indicates that these gases at the applied level do not have<br />

influence on elemental mercury measurement. The consistent mercury concentration also<br />

implies no mercury oxidation in the lines.<br />

120


Hg concentration (�g/Nm3)<br />

240<br />

200<br />

160<br />

120<br />

80<br />

40<br />

0<br />

-40<br />

Baseline<br />

gas<br />

Baseline<br />

gas,O 2<br />

air to analyzer<br />

Baseline<br />

gas,SO 2<br />

0 10 20 30 40 50 60 70<br />

Time (min)<br />

Figure 5.2. Effects of different gases on elemental mercury measurement <strong>by</strong>passing the<br />

converter.<br />

121<br />

Baseline<br />

gas,NO X<br />

Baseline<br />

gas,HCl<br />

5.3 Performance test of the red brass converter<br />

Baseline gas<br />

O 2, SO 2, NO X,HCl<br />

The red brass chips are obtained through the analyzer supplier Lumex. The<br />

typical composition of red brass includes 85% Cu, 5% Sn, 5% Zn and 5% Pb [8]. The<br />

idea of using red brass at low temperature is to bind free halogens in the flue gas and thus<br />

prevent back reaction into mercury halides as illustrated in following reaction [9]:<br />

Cl2 �Cu � CuCl2<br />

(5R1)<br />

The red brass converter is designed to convert Hg 2+ to Hg 0 at low temperatures of<br />

120-250�C to minimize the corrosion problem caused <strong>by</strong> SO2 oxidation at high<br />

temperatures [9]. The principle of the converter is to convert oxidized mercury according<br />

to following reaction:<br />

HgCl2 �Me�Hg� MeCl2<br />

(5R2)<br />

where Me could be Cu, Sn, Zn, and Pb that are contained in the red brass.


The performance of the red brass converter on elemental mercury measurement<br />

was first investigated. Test of the converter at 180�C in nitrogen atmosphere with only<br />

elemental mercury shows that the converter works well, since measurements through and<br />

<strong>by</strong>pass of the converter give the same mercury concentrations and the response time is<br />

short. However, tests of the converter using simulated flue gas and elemental mercury<br />

show that the performance of the converter degrades as a function of time. After short<br />

term exposure to the simulated flue gas the measured mercury level through the<br />

converter starts to decrease and is lower than that measured <strong>by</strong>passing the converter.<br />

Detailed investigations were then conducted to study the possible effects of gases<br />

on Hg 0 measurement through the converter. The applied gas concentrations are: 15 ppmv<br />

HCl, 1000 ppmv NO, 30 ppmv NO2, 1000 ppmv SO2, 1% H2O. Figure 5.3 shows the<br />

measured Hg 0 through the converter after adding different gases. When HCl, SO2 and<br />

NOx is added alone with water, the measured Hg 0 after the converter are the same as the<br />

inlet. However, when HCl is added either with SO2 or NOx the measured Hg 0 through the<br />

converter decreases with time, indicating that the catalyst surface is modified and starts<br />

to adsorb mercury or oxidize it to HgCl2.<br />

122


Hg concentration (�g/Nm 3 )<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

N 2+Hg+H 2O<br />

N 2+Hg+<br />

H 2O+NO x<br />

N 2+Hg+<br />

H 2 O+HCl N 2+Hg+H 2O+SO 2 N 2+Hg+H 2O<br />

+HCl+SO 2<br />

0 60 120 180 240 300 360 420 480<br />

Time (min)<br />

Figure 5.3. Measured elemental mercury concentration through the converter with 20 g<br />

red brass chips at 180�C after adding different gases.<br />

123<br />

N 2+Hg+H 2O<br />

+HCl+SO 2<br />

+NO x<br />

N 2+Hg+H 2O<br />

+SO 2 +NO x<br />

N 2+Hg+H 2O<br />

+HCl+NO x<br />

Besides reaction with mercury chloride, copper in the red brass can also react<br />

with other gases and form oxidized copper compounds. Possible reactions include:<br />

2Cu �O� 2CuO<br />

(5R3)<br />

2<br />

CuO �2HCl �CuCl � H O<br />

(5R4)<br />

2 2<br />

2CuO �2SO�O� 2CuSO<br />

(5R5)<br />

2 2 4<br />

Similar reactions could also take place for metals such as Sn, Zn and Pb contained in the<br />

red brass. It has been reported that NO2 is a very good oxidizing agent for preparing ZnO<br />

<strong>from</strong> metallic zinc through the reaction [10]<br />

NO2 �Zn � NO � ZnO<br />

(5R6)<br />

A similar reaction might take place between NO2 and copper. Copper chloride and<br />

copper sulfate have been used as promoters to improve mercury oxidation and adsorption<br />

<strong>by</strong> different sorbents [11-15]. These possible reactions might explain why elemental


mercury adsorption and oxidation takes place on the red brass chips in the simulated<br />

cement kiln flue gas.<br />

The oxidized mercury is added and produced <strong>by</strong> passing gases to the reactor with<br />

4 g SCR catalyst plate at 150�C. Oxidation of Hg 0 <strong>by</strong> the SCR system has been reported<br />

in both power plants[16,17] and in bench-scale tests [18-21]. The oxidation of mercury<br />

<strong>by</strong> the SCR catalyst is fast and about 70% mercury oxidation is obtained when 4g SCR<br />

catalyst is exposed to the simulated flue gas with15 ppmv HCl, 1% H2O, 1000 ppmv NO,<br />

30 ppmv NO2 and 1000 ppmv SO2 at 150�C. Figure 5.4 shows the result for using only<br />

15 ppmv HCl, 1% H2O and with N2 as balance. It takes about 5 h for the converter to<br />

obtain full oxidized mercury reduction, indicating that red brass converter cannot be used<br />

for dynamic measurement.<br />

Hg concentration (�g/Nm3)<br />

240<br />

200<br />

160<br />

120<br />

80<br />

40<br />

0<br />

Hg0, <strong>by</strong>pass reactor<br />

Hg0 through reactor with SCR<br />

0 2 4 6 8 10 12<br />

Time (hour)<br />

124<br />

Hgtotal through converter<br />

Figure 5.4. Measured total mercury concentration using 4g SCR catalyst at 150�C and<br />

2.75 Nl/min flue gas containing only 15 ppmv HCl and 1% H2O, 20 g red brass in the<br />

converter at 180�C.<br />

Based on these tests it was suggested <strong>by</strong> the supplier that the converter material<br />

reaches stability only after a period of several hours of operation under the gas mixture<br />

investigated in this project [22]. Tests were then conducted <strong>by</strong> conditioning the converter


with gases excluding Hg 0 addition. Detailed results are shown in figure 5.5. After<br />

conditioning the red brass catalyst with 15 ppmv HCl, 1000 ppmv SO2, 500 ppmv NO,<br />

15 ppmv NO2 and 1% H2O for 40 h, the measured mercury level through the converter is<br />

about 90 �g/Nm 3 compared to Hg 0 inlet level of 210 �g/Nm 3 . Furthermore, the measured<br />

mercury concentration through the converter keeps decreasing with time. The results<br />

show that the red brass converter does not work for the present conditions.<br />

Hg concentration (�g/Nm3)<br />

240<br />

200<br />

160<br />

120<br />

80<br />

40<br />

0<br />

Hgtotal through converter<br />

0 1 2 3 4 5<br />

Time (hour)<br />

125<br />

Hg0 <strong>by</strong>pass<br />

converter<br />

Figure 5.5. Measured total mercury concentration after passing SCR reactor with 4g SCR<br />

catalyst at 150�C. 20 g red brass in the converter at 180�C was preconditioned <strong>by</strong> 2.75<br />

Nl/min flue gas containing 15 ppmv HCl, 1000 ppmv SO2, 1% H2O, 500 ppmv NO, 15<br />

ppmv NO2 for 40 h.<br />

5.4 Performance of the sulfite converter<br />

The principle of the sulfite converter is that oxidized mercury such as HgCl2 can<br />

be reduced to Hg 0 through following reaction [23]:<br />

HgCl �NaSO �Hg�2NaCl �SO � O<br />

(5R7)<br />

1<br />

2 2 3 2 2 2


It is reported that 95% or more HgCl2 reduction efficiency can be obtained at<br />

300-500�C [23]. It is not stated in the patent for which gas composition this conversion<br />

was obtained and for how long. Thus a fundamental investigation of the sulfite converter<br />

under simulated cement kiln flue gas is necessary. The effect of converter temperature on<br />

Hg 0 recovery was tested and the results are illustrated in table 5.1. With 5 g sulfite<br />

compounds at 350�C, full mercury recovery can only be obtained for 1 h. Then the Hg 0<br />

recovery decreases with time and drops to 43% after another 3.5 h. For 10 g sulfite<br />

compounds at 450�C, full mercury recovery was obtained for 2 h. Then the mercury<br />

recovery decreased with time and dropped to 87% after another 4 h.<br />

Fast deactivation of sodium sulfite can take place at high temperatures [23].<br />

Sodium sulfite is water soluble and can recrystallize when water is present in the gas.<br />

When recrystallization occurs, the resistance of a layer of the sodium sulfite to gas<br />

transport is increased and the oxidized mercury reduction efficiency may be reduced. The<br />

higher the temperature the more recrystallization of sodium sulfite may take place. To<br />

minimize the deactivation, the converter was first tested at 250�C. However, only about<br />

50% mercury recovery was obtained immediately after switching gas to the converter<br />

and the Hg 0 recovery kept decreasing to 36% after another 1.5 h. Then the converter<br />

temperature was increased to 500�C. The mercury concentration after the converter<br />

increased sharply to 180% of the inlet elemental mercury level right after increasing the<br />

converter temperature. This is probably due to the fact that mercury is first adsorbed on<br />

the converter material at 250�C and then desorbs at high temperatures. Full mercury<br />

recovery was obtained for 1 h and then the mercury recovery decreased slowly to 88%<br />

after another 13 h.<br />

126


Table 5.1. Test results of elemental mercury recovery <strong>by</strong> the sulfite converter in 2.75<br />

Nl/min simulated cement kiln flue gas containing 21% CO2, 6% O2, 1% H2O, 1000 ppmv<br />

NO, 30 ppmv NO2, and 1000 ppmv SO2.<br />

Sulfite<br />

compound load<br />

(g)<br />

Converter<br />

temperature<br />

(�C)<br />

HCl level<br />

(ppmv)<br />

127<br />

Short time<br />

performance<br />

5 350 15 Full Hg 0 recovery<br />

for 1 h<br />

10 250 15 50% Hg 0 recovery<br />

for 0.5 h<br />

10 450 15 Full Hg 0 recovery<br />

for 2 h<br />

20 500 (after 15 Full Hg<br />

250�C test)<br />

0 recovery<br />

for 1 h<br />

20 500 2 Full Hg 0 recovery<br />

for 72 h<br />

20 500 6 Full Hg 0 recovery<br />

for 24 h<br />

20 500 10 Full Hg 0 recovery<br />

for 15 h<br />

Long time performance<br />

Hg 0 recovery decreases to<br />

43% after 3.5 h<br />

Hg 0 recovery decreases to<br />

36% after 1.5 h<br />

Hg 0 recovery decreases to<br />

87% after 4 h<br />

Hg 0 recovery decreases to<br />

88% after 13 h<br />

Not tested<br />

Not tested<br />

Hg 0 recovery decreases to<br />

95% after 35 h<br />

The level of HCl in the flue gas is a key factor that determines the efficiency and<br />

lifetime of the converter. Since only short time of full oxidized mercury reduction was<br />

observed with 15 ppmv HCl in the simulated flue gas, the HCl level was decreased to<br />

study the effects. With 2, 6, and 10 ppmv HCl in the simulated gas, full oxidized<br />

reduction can be obtained for at least 72, 24, and 15 h, respectively, for short-term test.<br />

With 10 ppmv HCl in the simulated cement kiln flue gas continuous operation of the<br />

converter with 20g sulfite material up to 2-3 months has been achieved. The presence of<br />

HCl in the gas can result in mercury oxidation both in the flue gas and on the sorbent.<br />

The recombination of elemental mercury and HCl after the converter might also be<br />

enhanced with high levels of HCl in the gas.<br />

It should be noted that the sulfite can be oxidized to sulfate <strong>by</strong> oxygen in the flue<br />

gas:<br />

Na SO � O � Na SO<br />

(5R8)<br />

1<br />

2 3 2 2 2 4


To study the effects of sodium sulfite oxidation to sulfate on oxidized mercury reduction<br />

efficiency, the sulfite pellets were first exposed to 12% O2 at 350�C for 18 h. Figure 5.6<br />

shows that the maximum Hg 0 recovery is about 90% after preconditioning <strong>by</strong> oxygen and<br />

slowly decreases with time. This indicates that the sodium sulfite was partly oxidized to<br />

sodium sulfate during preconditioning <strong>by</strong> oxygen. The formed sulfate is not active for<br />

reduction of oxidized mercury to elemental mercury. Normally the analyzer is running all<br />

the time to avoid damage of the lamp <strong>by</strong> restarting of the analyzer. When the experiment<br />

is not run, air is added to the analyzer. The test of oxygen precondition indicates that the<br />

converter should be closed to avoid oxidation of the sulfite compound when the<br />

experiment is not running. Instead the analyzer is running in Hg 0 measurement mode<br />

with air to the analyzer, <strong>by</strong>passing the converter.<br />

Hg concentration (�g/Nm 3 )<br />

200<br />

160<br />

120<br />

80<br />

40<br />

0<br />

Hgtotal<br />

<strong>by</strong>pass<br />

reactor<br />

0 30 60 90 120<br />

Time (min)<br />

128<br />

Hg0 inlet<br />

Hgtotal through reactor<br />

Figure 5.6. Test of 20 g sodium sulfite converter materials at 350�C using 2.75 Nl/min<br />

simulated flue gas with 15 ppmv HCl. The sulfite pellets were pre-conditioned <strong>by</strong> 12%<br />

O2 for 18 h. Oxidized mercury is produced <strong>by</strong> passing gases through 4 g SCR catalyst at<br />

150�C.


The dynamics of the converter were investigated <strong>by</strong> studying the response time of<br />

mercury measurement to the change of mercury addition and switching between the<br />

reactor and <strong>by</strong>pass. This was carried out <strong>by</strong> step up and step down tests [7]. The<br />

dynamics of Hg 0 measurement <strong>by</strong>passing the converter were first investigated. The steps<br />

of the dynamics test are illustrated in figure 5.7 Air was used as zero gas and added to the<br />

analyzer directly until a stable reading was achieved for about 5 min. Then air addition<br />

was stopped and Hg 0 in simulated flue gas was added <strong>by</strong>passing the reactor with SCR<br />

catalyst and the sulfite converter to measure the Hg 0 inlet level. After a stable reading of<br />

the inlet Hg 0 level was obtained for about 10 min, zero air was added to the analyzer and<br />

step down test of Hg 0 measurement was finished when stable reading was obtained. The<br />

95% response time of both step up and down is less than 0.5 min for elemental mercury<br />

measurement <strong>by</strong>passing the sulfite converter. The step change is very similar to that seen<br />

in the elemental measurement shown in figure 5.2.<br />

Hg concentration (�g/Nm3)<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Bypass<br />

SCR<br />

Air to<br />

converter,<br />

gas to SCR<br />

Through<br />

SCR<br />

0 30 60 90 120<br />

Time (min)<br />

129<br />

Air to<br />

converter<br />

Bypass<br />

SCR<br />

Bypass<br />

SCR<br />

stop Hg<br />

addition<br />

Figure 5.7. Response of total mercury measurement with elemental Hg inlet level of 112<br />

�g/Nm 3 . 20 g sodium sulfite converter materials is used in the converter at 500�C using<br />

2.75 Nl/min simulated flue gas with 10 ppmv HCl. Oxidized mercury is produced <strong>by</strong><br />

passing gases through 4 g SCR catalyst at 150�C.


Then the measurement was switched to total mercury measurement through the<br />

converter. The Hg 0 in the simulated flue gas was added to the reactor with SCR catalyst<br />

and the converter. The step up test of total mercury measurement was finished when<br />

stable mercury measurement through the converter was achieved for 50 min. Then air<br />

was added to the converter for a step down test. The dynamic tests were conducted at<br />

different Hg 0 inlet levels of 41, 112, and 150 µg/Nm 3 . Figure 5.7 shows the response of<br />

both Hg 0 measurement <strong>by</strong>passing the converter and SCR catalyst and total mercury<br />

measurement through the SCR catalyst and converter with a Hg 0 inlet level of 112<br />

µg/Nm 3 . Close look at the response for step change in figure 5.8 shows that the response<br />

of mercury measurement is very fast for both step up and down tests. The 95% response<br />

time for step up and down change is 1.5 and 0.6 min, respectively.<br />

Hg concentration (�g/Nm 3 )<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1.50 min<br />

95% step<br />

up change<br />

34 36 38 40 42 44 46 48 50<br />

Time (min)<br />

130<br />

0.60 min<br />

95% step<br />

down change<br />

78 80 82 84 86<br />

Time (min)<br />

Figure 5.8. Close look of the response time test shown in figure 5.7. Left: step up test <strong>by</strong><br />

switching gas addition to the sulfite converter <strong>from</strong> air to simulated flue gas with<br />

oxidized mercury produced <strong>by</strong> SCR catalyst. Right: step down test <strong>by</strong> switching gas<br />

addition to the sulfite converter <strong>from</strong> simulated flue gas with oxidized mercury produced<br />

<strong>by</strong> SCR catalyst to air.


5.5 Examples of dynamic measurement of mercury adsorption and<br />

oxidation on activated carbon<br />

Commercial activated carbons, Darco Hg and HOK standard were investigated in<br />

simulated cement kiln flue gas at 150�C. Figure 5.9 shows the mercury profiles for the<br />

Darco Hg activated carbon. The experiments are conducted twice using separate total and<br />

elemental mercury measurement. Comparison of the elemental and total mercury<br />

measurement shows that both adsorption and oxidation of mercury <strong>by</strong> the carbon occur.<br />

After mercury breakthrough is achieved, the mercury oxidation is stable at about 92%.<br />

Gaseous Hg (�g/Nm3)<br />

200<br />

160<br />

120<br />

80<br />

40<br />

0<br />

<strong>by</strong>pass<br />

reactor<br />

0 0.5 1 1.5 2 2.5 3 3.5<br />

Time (hour)<br />

131<br />

through<br />

reactor<br />

Gaseous Hg total<br />

Gaseous Hg 0<br />

Figure 5.9. Total and elemental mercury profile of 30 mg Darco Hg activated carbon<br />

mixed with 2 g sand at 150�C using 2.75 Nl/min simulated cement kiln flue gas with 10<br />

ppmv HCl, two separate tests and measurements.<br />

Rather than running the test of the same carbon twice as shown in figure 5.9, it is<br />

possible to run the test once and evaluate both the mercury adsorption and oxidation <strong>by</strong><br />

the carbon. The mercury breakthrough was first obtained <strong>by</strong> measuring total mercury<br />

through the converter, then zero air is added to the converter and the analyzer was<br />

changed to Hg 0 measurement mode. During this period mercury in simulated gas was<br />

still added to the sorbent to avoid possible desorption of mercury <strong>from</strong> the carbon. When


the analyzer was running for Hg 0 measurement, gases after the reactor were switched to<br />

the analyzer to measure Hg 0 . Figure 5.10 illustrates the mercury adsorption and oxidation<br />

<strong>by</strong> HOK standard carbon at 150�C. In this case 57% mercury oxidation was observed.<br />

Both the tests of Darco Hg and HOK in simulated cement kiln flue gas show that the<br />

sulfite converter and analysis system are capable of following the transient mercury<br />

outlet concentration in a satisfactory way.<br />

Gaseous Hg (�g/Nm 3 )<br />

200<br />

160<br />

120<br />

80<br />

40<br />

0<br />

<strong>by</strong>pass<br />

reactor<br />

Hg total<br />

through<br />

reactor Hg total<br />

Gas to reactor,<br />

air to analyzer<br />

change <strong>from</strong> Hg total<br />

to Hg 0 measurement<br />

0 1 2 3 4 5<br />

Time (min)<br />

132<br />

through<br />

reactor Hg 0<br />

Figure 5.10. Total and elemental mercury profile of 30 mg HOK standard activated<br />

carbon mixed with 2 g sand at 150�C using 2.75 Nl/min simulated cement kiln flue gas<br />

with 10 ppmv HCl.<br />

5.6 Suggestions for practical application of the converter<br />

The conditions in full-scale application are much more demanding than in the labscale<br />

investigation. In this work no particles in the gas stream were applied. On the other<br />

hand, the dust load in the flue gas between the raw mill and filter could be up to 800-<br />

1000 g/Nm 3 . The sampling probe needs to be able to separate the particles <strong>from</strong> the flue<br />

gas efficiently to avoid plugging of probe. Adsorption of mercury <strong>by</strong> the dust and probe<br />

should be minimized <strong>by</strong> high sampling flow rate and high filter temperature.


The HCl content in the cement kiln flue gas can be up to 20-25 ppmv [24]. As<br />

found in this work, the full Hg 0 recovery can only be maintained for short period when<br />

more than 10 ppmv HCl is present in the flue gas. It is therefore necessary to remove HCl<br />

before or in the converter. Lime pellets can be used together with the sulfite compounds<br />

and the converter temperature should be high enough to avoid mercury adsorption on the<br />

lime pellets. Alternatively, large amount of converter material might be used.<br />

Compared to power plants and incinerators, the emission levels of CO and<br />

volatile organic compounds such as hydrocarbons are higher in cement plants. The<br />

emission level of volatile organic compound in the stack gas of cement kilns is usually<br />

between 10 and 100 mg/Nm 3 , with a few excessive cases up to 500 mg/Nm 3 [25]. The CO<br />

concentration in the stack gas can be as high as 1000 mg/Nm 3 , even exceeding 2000<br />

mg/Nm 3 in some cases. High levels of CO and hydrocarbons in the flue gas will cause<br />

fast contamination of the windows in the analytical cells and interruption of the mercury<br />

measurement [22]. Measures such as dilution should be applied to minimize the problem.<br />

The sulfite material should be kept in a closed box to avoid oxidation <strong>by</strong> air and<br />

moisture. It is important that the sulfite powders are adhered uniformly to the surface of<br />

the thin layer of water glass on the zeolite pellets. Thoroughly mixing the water glass<br />

with zeolite pellets in a plastic container can improve the sulfite converter performance.<br />

In this way, the sulfite converter can work well up to months, as observed in this work.<br />

For both lab-scale and full-scale application of the analysis system, it is important<br />

to avoid cold parts in the system. All the connections, Teflon lines and gas contacting<br />

parts in the analyzer before the spectrometry should be heated above 150�C. When the<br />

converter is not used, the converter temperature should be decreased to 100�C. The<br />

converter should be closed to avoid deactivation of the converter material due to<br />

oxidation of sulfite to sulphate <strong>by</strong> air.<br />

5.7 Conclusions<br />

To be able to perform dynamic measurement of mercury adsorption <strong>by</strong> sorbents,<br />

red brass chips and sulfite converter were investigated in simulated cement kiln flue gas<br />

133


in a fixed-bed reactor system. The converter with red brass chips works only when<br />

measuring elemental mercury in nitrogen (i.e., without carrying out actual conversion)<br />

and does not work properly even when only elemental mercury was added to the<br />

simulated flue gas. The red brass is poisoned or oxidized within a short time and adsorbs<br />

elemental mercury. When oxidized mercury was produced <strong>by</strong> passing gases through a<br />

separate reactor with an SCR catalyst, the red brass converter cannot fully reduce HgCl2<br />

to elemental mercury under any relevant condition.<br />

Sodium sulfite converter material was prepared <strong>by</strong> dry impregnation of sodium<br />

sulfite and calcium sulfate powders on zeolite pellets using water glass as binder. The<br />

optimal operating temperature of the sulfite converter is 500�C. The level of HCl in the<br />

flue gas is a key factor that determines the efficiency and lifetime of the converter. Full<br />

elemental mercury recovery can only be obtained for short period with 15 ppmv HCl in<br />

the simulated gas, but the sulfite converter works well at 500�C with up to 10 ppmv HCl<br />

in the simulated cement kiln flue gas. When the converter is not used, the converter<br />

temperature was decreased to 100�C without air passing through to avoid deactivation of<br />

the converter material <strong>by</strong> oxidation of the sodium sulfite to sodium sulfate. The response<br />

time of the sulfite converter is short and typically within at most two minutes, which<br />

makes it appropriate for not too fast dynamic measurements, as verified <strong>by</strong> dynamic<br />

mercury adsorption tests on commercial activated carbons Darco Hg and HOK standard<br />

in a fixed-bed reactor. Suggestions for practical application of the sulfite converter in<br />

cement plant with high dust load are provided.<br />

5.8 References<br />

[1] R.J. Schreiber and C.D. Kellett, Compilation of mercury emissions data, PCA R&D Serial No.<br />

SN3091, 2009.<br />

[2] D.L. Laudal, J.S. Thompson, J.H. Pavlish, L.A. Brickett, P. Chu, Use of continuous mercury<br />

monitors at coal-fired utilities, Fuel Processing Technology. 85 (2004) 501-511.<br />

[3] J. Wu, Y. Du and W. Pan, J. Ren, P. He, W. Wang, M. Shen, X. Leng, Y. Jin, Z. Dai, L. Zhao,<br />

X. Ming, Y. Cao, W. Pan, Study on different measurement methods of mercury emission in the<br />

134


coal-fired power station, 3 rd International Conference on Bioinformatics and Biomedical<br />

Engineering, Beijing, China, June 11-13, 2009.<br />

[4] V. Schmid, Continuous monitoring of mercury emissions <strong>from</strong> stationary sources, 2002.<br />

[5] M. Holmes and J. Pavlish, <strong>Mercury</strong> information of clearinghouse, Quarterly 2–mercury<br />

measurement, 2004.<br />

[6] J. Wang, Z. Xiao, O. Lindqvist, On-line measurement of mercury in simulated flue gas, Water,<br />

Air, & Soil Pollution. 80 (1995) 1217-1226.<br />

[7] M.L. Jones, D.L. Laudal and J.H. Pavlish, <strong>Mercury</strong> emission monitoring for the cement<br />

industry, <strong>Cement</strong> Industry Technical Conference Record, 2008 IEEE, Miami, Florida, May 18-22,<br />

2008.<br />

[8] Wikipedia, Brass, http://en.wikipedia.org/wiki/Brass, accessed December 7, 2010.<br />

[9] R. Kanefke, H. Köser, B. Vosteen, F. Kristina, B. Frank, S. Raik, Method for the production<br />

of elemental mercury <strong>from</strong> mercury compounds, patent WO 2008/064667 A2, 2008.<br />

[10] J.A. Rodriguez, T. Jirsak, J. Dvorak, S. Sambasivan, D. Fischer, Reaction of NO2 with Zn<br />

and ZnO: Photoemission, XANES, and density functional studies on the formation of NO3, The<br />

Journal of Physical Chemistry B. 104 (2000) 319-328.<br />

[11] S. Lee, J. Lee, T.C. Keener, Bench-scale studies of in-duct mercury capture using cupric<br />

chloride-impregnated carbons, Environ. Sci. Technol. 43 (2009) 2957-2962.<br />

[12] S. Lee, J. Lee, T.C. Keener, The effect of methods of preparation on the performance of<br />

cupric chloride-impregnated sorbents for the removal of mercury <strong>from</strong> flue gases, Fuel. 88 (2009)<br />

2053-2056.<br />

[13] A. Makkuni, R.S. Varma, S.K. Sikdar, D. Bhattacharyya, Vapor phase mercury sorption <strong>by</strong><br />

organic sulfide modified bimetallic iron-copper nanoparticle aggregates, Ind Eng Chem Res. 46<br />

(2007) 1305-1315.<br />

[14] D.E. Meyer, S.K. Sikdar, N.D. Hutson, D. Bhattacharyya, Examination of sulfurfunctionalized,<br />

copper-doped iron nanoparticles for vapor-phase mercury capture in entrainedflow<br />

and fixed-bed systems, Energy & Fuels. 21 (2007) 2688-2697.<br />

[15] D.E. Meyer, N. Meeks, S. Sikdar, N.D. Hutson, D. Hua, D. Bhattacharyya, Copper-doped<br />

silica materials silanized with bis-(triethoxy silyl propyl)-tetra sulfide for mercury vapor capture,<br />

Energy Fuels. 22 (2008) 2290-2298.<br />

[16] D.L. Laudal, J.S. Thompson, J.H. Pavlish, L. Brickett, P. Chu, R.K. Srivastava, J. Kilgroe,<br />

C.W. Lee, <strong>Mercury</strong> speciation at power plants using SCR and SNCR control technologies, EM:<br />

Air and Waste Management Association's Magazine for Environmental Managers. (2003) 16-22.<br />

[17] H.G. Pedersen, L.S. Pedersen, H. Rostgaard and K. Pedersen, Oxidation of mercury on DNX<br />

catalysts. Proceedings of the Air Quality V: <strong>Mercury</strong>, Trace Elements, SO3, and Particulate<br />

Matter Conference, Arlington, VA, Sept 19–21, 2005.<br />

135


[18] S. Straube, T. Hahn, H. Koeser, Adsorption and oxidation of mercury in tail-end SCR-<br />

DeNOx plants—Bench scale investigations and speciation experiments, Applied Catalysis B:<br />

Environmental. 79 (2008) 286-295.<br />

[19] Y. Zhuang, J. Laumb, R. Liggett, M. Holmes, J. Pavlish, Impacts of acid gases on mercury<br />

oxidation across SCR catalyst, Fuel Processing Technology. 88 (2007) 929-934.<br />

[20] Y. Cao, Z. Gao, J. Zhu, Q. Wang, Y. Huang, C. Chiu, B. Parker, P. Chu, W. Pan, Impacts of<br />

halogen additions on mercury oxidation in a slipstream selective catalyst reduction (SCR) reactor<br />

when burning sub-bituminous coal, Environ. Sci. Technol. 42 (2008) 256-261.<br />

[21] H. Kamata, S. Ueno, T. Naito, A. Yukimura, <strong>Mercury</strong> oxidation over the V2O5(WO3)/TiO2<br />

commercial SCR catalyst, Ind Eng Chem Res. 47 (2008) 8136-8141.<br />

[22] R. Moeseler. Issues about red brass converter, personal communication, Lumex Analytical<br />

GmbH, 2010.<br />

[23] S. Akiyama, J. Kato, F. Koga, K. Ishikawa, Catalysts for reducing mercury, a mercury<br />

conversion unit, and an apparatus for measuring total mercury in combustion exhaust gas <strong>by</strong><br />

using the same, patent US2007/0232488 A1, 2007.<br />

[24] C. Senior, A. Sarofim and E. Eddings, Behavor and measurement of mercury in cement<br />

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

May 4-9 2003.<br />

[25] CEMBUREAU, the European <strong>Cement</strong> association, Best available technologies for the<br />

cement industry, 1999.<br />

136


Effects of bed dilution and carbon load on<br />

mercury adsorption capacity of activated<br />

137<br />

6<br />

carbon<br />

This chapter reports the effects of bed dilution and carbon load on the equilibrium<br />

mercury adsorption capacity of the activated carbon. The mercury adsorption capacity<br />

per unit mass of the activated carbon decreases when the carbon load is increased.<br />

Detailed investigations are conducted to reveal the cause.<br />

6.1 Introduction<br />

Most of the studies on mercury adsorption use bed dilution [1-9], while only few<br />

investigations apply pure sorbent bed [10-14] when the mercury sorbents are evaluated in<br />

fixed-bed reactors. The sorbent beds are often diluted with inert particles to suppress<br />

other potential disturbing effects such as axial dispersion and <strong>by</strong>passing [15,16]. Lowsurface-area<br />

materials such as glass beads and sand/quartz powder are preferred as<br />

diluting solids because of their relative inertness and good heat transfer properties. The<br />

effects of sorbent load on the mercury adsorption capacity of the sorbent are rarely<br />

reported in the literature.<br />

6.2 Effects of carbon load<br />

The direct result of the fixed-bed test is the mercury adsorption breakthrough<br />

curve. The percentage breakthrough is determined as a function of time <strong>by</strong> normalizing<br />

the measured total mercury concentration at the outlet of the sorbent bed to the inlet<br />

mercury concentration.


From the mercury breakthrough curve, the amount of mercury adsorbed on unit<br />

mass of the sorbent as a function of time can be calculated <strong>from</strong> the expression:<br />

q<br />

t<br />

�<br />

F<br />

W<br />

t<br />

� ( Cin<br />

� Cout,<br />

t ) dt<br />

(6.1)<br />

0<br />

where F is the flow rate through the sorbent bed, W is the mass of the sorbent, Cin is the<br />

inlet mercury concentration, Cout,t is the mercury concentration at the reactor outlet at<br />

time t. The mercury adsorption capacity of a known weight of a sorbent is calculated in<br />

terms of µg Hg adsorbed/mg_sorbent <strong>from</strong> the breakthrough curve for the sorbent. The<br />

equilibrium adsorption capacity is defined <strong>by</strong> the time when the outlet Hg concentration<br />

is first equal to the inlet concentration.<br />

Figure 6.1 presents the mercury breakthrough curves for different loads of Darco<br />

Hg activated carbon mixed with 2 g sand powder at 150�C using simulated cement kiln<br />

flue gas with elemental mercury. Faster mercury breakthrough is observed for smaller<br />

carbon load as expected. The calculated amount of adsorbed mercury and equilibrium<br />

mercury adsorption capacity per unit mass of the activated carbon are illustrated in figure<br />

6.2. The calculated amount of mercury adsorbed in the carbon does not increase<br />

proportionally to the mass of carbon, i.e., the mercury adsorption capacities of the carbon<br />

apparently decreases when the carbon load is increased. It seems that there is promotion<br />

of mercury adsorption <strong>by</strong> the sand when it is mixed with activated carbon. The trend line<br />

indicates that about 8.39 µg mercury is adsorbed <strong>by</strong> 2 g sand powder. Stuart [17] also<br />

reported that activated carbon mixed with sand had larger mercury uptake capacity than<br />

the carbon tightly packed in the reactor. He postulated that the incoming gas might be<br />

short circuiting and allowing the gas flow through the reactor without encountering all<br />

the tightly packed carbon. This argument is doubtful since even if the contact is poor<br />

uptake of mercury would just be lower and eventually the same uptake will be reached.<br />

138


C out/C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

A B C D<br />

0 1 2 3 4 5 6<br />

Time (hour)<br />

139<br />

A: 10 mg<br />

B: 30 mg<br />

C: 60 mg<br />

D: 100 mg<br />

Figure 6.1. <strong>Mercury</strong> breakthrough curves of different Darco Hg activated carbon loads<br />

mixed with 2 g sand powder and tested at 150�C using 2.75 Nl/min simulated flue gas<br />

with 170 µg/Nm 3 elemental mercury, 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10<br />

ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

Adsorbed Hg (�g)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Adsorbed Hg<br />

Adsorbed Hg<br />

Y=0.3797X+8.3921,R<br />

0.2<br />

0<br />

0<br />

0 20 40 60 80 100 120<br />

2 =0.99<br />

Adsorption capacity<br />

Carbon load in 2 g sand (mg)<br />

Figure 6.2. Calculated amount of adsorbed mercury and equilibrium mercury adsorption<br />

capacity of Darco Hg activated carbon mixed with 2 g sand powder and tested at 150�C<br />

with different carbon loads using 2.75 Nl/min simulated flue gas with 170 µg/Nm 3<br />

elemental mercury, 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv HCl, 1<br />

vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

Hg Adsorption capacity (�g Hg/mg_carbon)


To investigate whether the oxidation of elemental mercury could influence the<br />

mercury adsorption capacity, mercury adsorption <strong>by</strong> different carbon loads using HgCl2<br />

source are conducted. As shown in figure 6.3, similar trends as tests with elemental<br />

mercury source are observed. This implies that the decrease of mercury adsorption<br />

capacity with increased carbon load in the sand is not caused <strong>by</strong> the oxidation of<br />

elemental mercury.<br />

Adsorbed mercury (�g)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Adsorbed Hg<br />

0.4<br />

Adsorbed Hg<br />

Y=0.426X+6.739,R 0.2<br />

0<br />

0 10 20 30 40 50 60 70 80<br />

2 =0.96<br />

Adsorption capacity<br />

Carbon load (mg)<br />

Figure 6.3. Calculated amount of adsorbed mercury and equilibrium mercury adsorption<br />

capacity of Darco Hg activated carbon mixed with 2 g sand powder and tested at 150�C<br />

with different carbon loads using 2.75 Nl/min simulated flue gas with 170 µg/Nm 3<br />

mercury <strong>from</strong> HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv<br />

HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

One possible cause of the decrease of mercury adsorption capacity with increased<br />

carbon load could be the wall effect. Carbon particles separate <strong>from</strong> sand powders during<br />

loading the sample to the reactor. A layer of carbon particles deposits on the sample<br />

holder. Later these carbon particles are loaded to the reactor <strong>by</strong> knocking the sample<br />

holder. During loading of the sample to the reactor, carbon particles stick on the reactor<br />

wall. A quartz wool plug is used to move the carbon particles that adhere on the reactor<br />

wall to the top of the carbon bed. As a result some carbon particles are loaded to the area<br />

140<br />

2<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

Hg adsorption capacity (�g Hg/mg_carbon)


close to the reactor wall. More carbon particles adhere on the reactor wall when larger<br />

carbon load is applied.<br />

Another reason might be the leakage of the system and mercury adsorption <strong>by</strong> the<br />

quartz wool and sand powder. However, leakage tests of the system at different stages of<br />

the project show that the system is tight. The pressure drop over the carbon bed is about 5<br />

mbar. The deviation of the flow rates at the reactor inlet and outlet to the flow rate after<br />

the gas mixing panel is within 2.2%. Tests with empty reactor, reactor with quartz wool,<br />

and sand powder do not show any adsorption of either elemental mercury or mercury<br />

chloride.<br />

6.3 Effects of bed dilution<br />

Negative deviation of conversion caused <strong>by</strong> dilution of the catalyst bed with inert<br />

particles in gas-solid systems has been reported [15,16]. Dilution of activated carbon <strong>by</strong><br />

inert sand powder is applied in this work; it is therefore relevant to evaluate the possible<br />

effects caused <strong>by</strong> the dilution. The extent of negative effect depends on the amount of<br />

dilution, the reaction/adsorption kinetics, the particles and reactor geometry, and the<br />

degree of segregation of carbon and sand. Since the mercury removal fraction <strong>by</strong> the<br />

carbon bed changes with time, the dilution effect � as a relative measure of the deviation<br />

in the conversion can be calculated for different time:<br />

xundiluted () t � xdiluted () t<br />

�() t � (6.2)<br />

xundiluted<br />

where xdiulted(t) and xundiluted(t) is the mercury removal fraction at time t for diluted and<br />

undiluted bed, respectively.<br />

For practical application the relative deviation in conversion can be estimated<br />

<strong>from</strong> observable parameters [15,16] :<br />

b d p xdiluted () t<br />

�() t �(<br />

)<br />

(6.3)<br />

1�b hbed<br />

2<br />

where b is the volume of inert sand as fraction of total volume of solids, dp is carbon<br />

particle diameter, and hbed is the bed height.<br />

For 10 mg Darco Hg carbon mixed with 2 g sand, the b is calculated as:<br />

141


msand<br />

/ �sand<br />

�3<br />

2�10 /1602<br />

carbon �carbon � sand �sand<br />

�<br />

�6 � �<br />

�3<br />

b � � �0.98<br />

m / m / 10 10 / 510 2 10 /1602<br />

Figure 6.4 presents the calculated relative deviation in mercury adsorption as a<br />

function of time for different loads of Darco Hg carbon tested at 180�C in simulated<br />

cement kiln flue gas. Larger relative deviation in short period is observed for smaller<br />

carbon loads, i.e., larger dilution ratio. However, the area under the relative deviation<br />

curve and above the zero deviation appears to be similar for different carbon loads. This<br />

indicates that the influence of bed dilution on the equilibrium mercury adsorption<br />

capacity of the carbon is similar. Therefore the decrease of mercury adsorption capacity<br />

with increase of carbon loads is probably not caused <strong>by</strong> the bed dilution.<br />

Relative deviation, �<br />

0.08<br />

0.07<br />

0.06<br />

0.05<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

-0.01<br />

A<br />

B<br />

C<br />

0 0.5 1 1.5 2 2.5<br />

Time (hour)<br />

142<br />

A: 10 mg carbon, b=0.98<br />

B: 30 mg carbon, b=0.96<br />

C: 60 mg carbon, b=0.91<br />

Figure 6.4. The calculated relative deviation � as a function of time for tests at 180�C<br />

with different carbon loads using 2.75 Nl/min simulated flue gas with 170 µg/Nm 3<br />

mercury <strong>from</strong> HgCl2, 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv HCl, 1<br />

vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

Tests using same dilution ratio, i.e., 10 mg carbon is mixed with 2 g sand, 30 mg<br />

carbon with 6 g sand, 60 mg carbon with 12 g sand, are also performed at 180�C using<br />

simulated cement kiln flue gas with HgCl2. Figure 6.5 shows the calculated amount of<br />

adsorbed mercury and equilibrium mercury adsorption capacity of the activated carbon.


Similar mercury adsorption capacity is still not obtained for different carbon loads, which<br />

behaves as with 2 g sand.<br />

Adsorbed mercury (�g)<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Adsorbed Hg<br />

Adsorbed Hg<br />

Y=0.3777X+7.0386,R<br />

0.2<br />

0<br />

0 10 20 30 40 50 60 70 80<br />

2 =0.97<br />

Adsorption capacity<br />

Carbon load (mg)<br />

Figure 6.5. Calculated amount of adsorbed mercury and equilibrium mercury adsorption<br />

capacity of Darco Hg activated carbon mixed with sand powder using same dilution rate<br />

and tested at 180�C with different carbon loads using 2.75 Nl/min simulated flue gas<br />

with 170 µg/Nm 3 mercury <strong>from</strong> HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000<br />

ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

6.4 Effects of sand load<br />

If the mercury adsorption <strong>by</strong> activated carbon is promoted <strong>by</strong> the sand mixing, it<br />

would be interesting to investigate the effects of sand load on the promotion of mercury<br />

adsorption capacity of the carbon <strong>by</strong> running tests with different masses of sand. Table<br />

6.2 presents the calculated amount of adsorbed mercury and equilibrium mercury<br />

adsorption capacity of 10 mg Darco Hg activated carbon mixed with different amounts of<br />

sand powder at 150�C in simulated cement kiln flue gas with HgCl2. When the sand load<br />

is above 20 mg the mercury adsorption capacities of the Darco Hg do not increase further<br />

and level off at a value of about 1.135 µg Hg/mg_carbon. This also indicates that the<br />

repeatability of the experiment is reasonable.<br />

143<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

Hg adsorption capacity (�g Hg/mg_carbon)


Table 6.2. Calculated amount of adsorbed mercury and equilibrium mercury adsorption<br />

capacity of 10 mg Darco Hg activated carbon mixed with different amounts of sand<br />

powder and tested at 150�C using 2.75 Nl/min simulated flue gas with HgCl2 source,<br />

1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2,<br />

and 21 vol.% CO2.<br />

Sand load (g) Hg inlet (µg/Nm 3 ) Adsorbed Hg (µg) Adsorption capacity<br />

(µg Hg/mg_carbon)<br />

0 164 5.06 0.506<br />

0.01 161 9.81 0.981<br />

0.02 160 11.36 1.136<br />

0.05 166 11.48 1.148<br />

0.10 164 11.07 1.107<br />

0.25 166 12.17 1.217<br />

0.5 174 10.50 1.050<br />

1 217 10.24 1.024<br />

2 183 12.24 1.224<br />

2 163 11.80 1.180<br />

4 206 11.29 1.129<br />

6.5 Effects of carbon loading location<br />

The carbon sample was separated <strong>from</strong> the sand powder with quartz wool plug to<br />

test possible effect of carbon loading location. Different locations of the carbon sample<br />

are applied to investigate whether the promotion of mercury adsorption is caused <strong>by</strong> the<br />

preconditioning of the gas <strong>by</strong> the sand. When the carbon is on top of the sand, the flue<br />

gas first contacts the carbon powder. However, as shown in figure 6.6, the equilibrium<br />

mercury adsorption capacity is almost the same when the carbon sample is loaded on top<br />

of and under the sand powder. This implies that the promotion of mercury adsorption<br />

only occurs when the carbon is mixed with sand powder. The slightly larger mercury<br />

adsorption capacity with sand powder in the reactor compared to only carbon in the<br />

reactor might be due to the improved contact of carbon with gas flow.<br />

144


Hg adsorption capacity (�g Hg/mg-carbon)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

No sand<br />

0<br />

Figure 6.6. Equilibrium mercury adsorption capacity of 10 mg Darco Hg activated<br />

carbon on top of and under 1 g sand powder at 150�C using 2.75 Nl/min simulated flue<br />

gas with 170 µg/Nm 3 mercury <strong>from</strong> HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000<br />

ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

6.6 Effects of bed materials<br />

Different bed materials are applied to investigate the possible effects of bed<br />

materials on the mercury adsorption capacity of activated carbon. The investigated bed<br />

materials include sand powder, fine quartz powder, and glass beads and the mean<br />

diameters of these materials are 215, 2, and 180 µm, respectively. Baseline tests with<br />

only fine quartz powder and glass beads show that fine quartz powder is inert for<br />

mercury adsorption and the mercury adsorption <strong>by</strong> the glass beads is negligible. Figure<br />

6.7 compares the mercury adsorption capacity of Darco Hg activated carbon tested with<br />

different bed materials. The mercury adsorption capacity of Darco Hg carbon tested with<br />

fine quartz powder is much smaller than those with sand powder and glass beads. The<br />

fine quartz powder behaves like paste and might hinder the contact of gas with the<br />

carbon particles. Inconsistent mercury adsorption capacity is still obtained when different<br />

amounts of carbon are mixed with 2 g sand powder, fine quartz powder, and glass beads.<br />

145<br />

Carbon on top<br />

Sand on top


Hg adsorption capaity (�g Hg/mg_carbon)<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

Sand powder<br />

Fine quartz powder<br />

Glass beads<br />

10 mg carbon in<br />

2 g bed material<br />

146<br />

Sand powder<br />

Fine quartz powder<br />

Glass beads<br />

30 mg carbon in<br />

2 g bed material<br />

Figure 6.7. Equilibrium mercury adsorption capacity of Darco Hg activated carbon<br />

mixed with different bed materials and tested at 150�C using 2.75 Nl/min simulated flue<br />

gas with 170 µg/Nm 3 mercury <strong>from</strong> HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000<br />

ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

6.7 Effects of carbon type and particle size<br />

To study the effects of carbon type and particle size on the mercury adsorption<br />

capacity, commercial activated carbon pellets of Norit RB4 are crushed and sieved to<br />

size of 165 and less than 32 µm in diameter. Figure 6.8 shows the mercury adsorption<br />

capacity obtained with different carbon loads, carbon types and particle size. Inconsistent<br />

mercury adsorption capacity at different carbon loads is observed for both Darco Hg and<br />

Norit RB4 carbons with different sizes. The effects of carbon load are much smaller for<br />

Norit RB4 carbon.


Hg adsorption capacity (�g Hg/mg_carbon)<br />

2<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 10 20 30 40 50 60 70 80<br />

Carbon load (mg)<br />

147<br />

Darco Hg, 16 �m<br />

Norit RB 4, 165 �m<br />

Norit RB 4,


Adsorbed Hg (�g)<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

Adsorbed Hg<br />

Adsorption capacity<br />

1 2 3 4 5<br />

Portland cement load (g)<br />

148<br />

0.0025<br />

0.002<br />

0.0015<br />

0.001<br />

0.0005<br />

Figure 6.9. Calculated amount of adsorbed mercury and equilibrium mercury adsorption<br />

capacity as a function of Portland cement load at 150�C using 2.75 Nl/min simulated flue<br />

gas with 170 µg/Nm 3 mercury <strong>from</strong> HgCl2 source, 1000 ppmv NO, 23 ppmv NO2, 1000<br />

ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

6.9 Conclusions<br />

Inconsistent mercury adsorption capacity of activated carbon is observed at<br />

different carbon loads when mixed with sand. Smaller mercury adsorption capacity is<br />

obtained with larger carbon load. Tests with elemental mercury and mercury chloride,<br />

different carbon type and particle sizes show the same trend. Effects of bed dilution at<br />

fixed carbon load on the equilibrium mercury adsorption capacity appear to be limited.<br />

The mercury adsorption capacity of activated carbon obtained using sand and<br />

carbon mixture is larger than that obtained with only activated carbon. The mercury<br />

adsorption capacity with 10 mg carbon increases with sand load up to 20 mg and then<br />

levels off when the sand load is further increased.<br />

Similar mercury adsorption capacities are obtained with different Portland cement<br />

loads in the reactor. This implies that the inconsistent mercury adsorption capacity of<br />

carbon obtained using different carbon loads might be due to possible adsorption of<br />

0<br />

Hg adsorption capacity (�g Hg/mg_cement)


mercury <strong>by</strong> sand when it is mixed with carbon, rather than the failure of the experimental<br />

setup. The sand powder alone is inert for mercury adsorption, while after modification<br />

with chemical reagent it can be used for mercury adsorption [18,19]. In-situ analysis<br />

technology is required to reveal whether mercury is adsorbed <strong>by</strong> the sand when it is<br />

mixed with activated carbon.<br />

The problem of inconsistent mercury adsorption capacity was encountered in the<br />

late stage of the project when performing a fundamental parametric study. Although<br />

detailed tests are conducted to reveal the cause, the problem is not solved due to the lack<br />

of analysis techniques and time. It is impossible to repeat and run all the tests with only<br />

large carbon load within the time schedule of the project. For a full-scale application in<br />

the cement plant it is impossible to exclude all the cement materials in the flue gas even<br />

with a polishing filter. Instead of providing actual kinetics data relevant to full-scale<br />

application conditions, this work aims at evaluation the effects of different operating<br />

parameters and mathematical model development. Therefore, mercury adsorption<br />

kinetics obtained using 10 mg activated carbon mixed with 2 g sand powder is used in<br />

the following chapters dealing with parametric study and model development. It may be<br />

argued that it is difficult to state what the real capacity is if it depends on the mixing<br />

condition. In reality there will always be least 20 mg diluter with the sorbent and this is<br />

where it has stabilized.<br />

6.10 References<br />

[1] S. Sjostrom, T. Ebner, T. Ley, R. Slye, C. Richardson, T. Machalek, M. Richardson, R.<br />

Chang, Assessing sorbents for mercury control in coal-combustion flue gas, J. Air & Waste<br />

Manage. Assoc. 52 (2002) 902.<br />

[2] S.J. Lee, Y. Seo, J. Jurng, T.G. Lee, <strong>Removal</strong> of gas-phase elemental mercury <strong>by</strong> iodine- and<br />

chlorine-impregnated activated carbons, Atmospheric Environment. 38 (2004) 4887-4893.<br />

[3] D. Karatza, A. Lancia, D. Musmarra, Fly ash capture of mercuric chloride vapors <strong>from</strong><br />

exhaust combustion gas, Environ. Sci. Technol. 32 (1998) 3999-4004.<br />

[4] J.W. Portzer, J.R. Albritton, C.C. Allen, R.P. Gupta, Development of novel sorbents for<br />

mercury control at elevated temperatures in coal-derived syngas: results of initial screening of<br />

candidate materials, Fuel Processing Technology. 85 (2004) 621-630.<br />

149


[5] J.Y. Lee, Y. Ju, T.C. Keener, R.S. Varma, Development of cost-effective noncarbon sorbents<br />

for Hg 0 removal <strong>from</strong> coal-fired power plants, Environ. Sci. Technol. 40 (2006) 2714-2720.<br />

[6] D. Karata, A. Lancia, D. Musmarra, F. Pepe, Adsorption of metallic mercury on activated<br />

carbon, Symposium (International) on Combustion,. 26 (1996) 2439-2445.<br />

[7] G. Skodras, I. Diamantopoulou, G. Pantoleontos, G.P. Sakellaropoulos, Kinetic studies of<br />

elemental mercury adsorption in activated carbon fixed bed reactor, Journal of Hazardous<br />

Materials. 158 (2008) 1-13.<br />

[8] D. Karatza, A. Lancia, D. Musmarra, C. Zucchini, Study of mercury absorption and<br />

desorption on sulfur impregnated carbon, Experimental Thermal and Fluid Science. 21 (2000)<br />

150-155.<br />

[9] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors affecting<br />

mercury control in utility flue gas using activated carbon, Journal of the Air & Waste<br />

Management Association. 48 (1998) 1166.<br />

[10] R. Yan, Y.L. Ng, D.T. Liang, C.S. Lim, J.H. Tay, Bench-scale experimental study on the<br />

effect of flue gas composition on mercury removal <strong>by</strong> activated carbon adsorption, Energy &<br />

Fuels. 17 (2003) 1528-1535.<br />

[11] R. Yan, D.T. Liang, L. Tsen, Y.P. Wong, Y.K. Lee, Bench-scale experimental evaluation of<br />

carbon performance on mercury vapour adsorption, Fuel. 83 (2004) 2401-2409.<br />

[12] G.E. Dunham, R.A. DeWall, C.L. Senior, Fixed-bed studies of the interactions between<br />

mercury and coal combustion fly ash, Fuel Processing Technology. 82 (2003) 197-213.<br />

[13] B.A.F. Mibeck, E.S. Olson, S.J. Miller, HgCl2 sorption on lignite activated carbon: Analysis<br />

of fixed-bed results, Fuel Process Technol. 90 (2009) 1364-1371.<br />

[14] G.E. Dunham, S.J. Miller, <strong>Mercury</strong> capture <strong>by</strong> an activated carbon in a fixed-bed benchscale<br />

system, Environmental Progress. 17 (1998) 203.<br />

[15] R.J. Berger, J. Pérez-Ramírez, F. Kapteijn, J.A. Moulijn, Catalyst performance testing: the<br />

influence of catalyst bed dilution on the conversion observed, Chem. Eng. J. 90 (2002) 173-183.<br />

[16] R.J. Berger, J. Pérez-Ramírez, F. Kapteijn, J.A. Moulijn, Catalyst performance testing: bed<br />

dilution revisited, Chemical Engineering Science. 57 (2002) 4921-4932.<br />

[17] J.L. Stuart. Development of an experimental system to study mercury uptake <strong>by</strong> activated<br />

carbon under simulated flue gas conditions, Master thesis, University of Pittsburgh, 2002.<br />

[18] M. Holmes and J. Pavlish, <strong>Mercury</strong> information clearinghouse, Quarter 3- Advanced and<br />

developmental mercury control technologies, July 2004.<br />

[19] J.R. Butz, T.E. Broderick and C.S. Turchi, Amended Silicates for <strong>Mercury</strong> Control,<br />

project final report, DOE Award Number: DE-FC26-04NT41988, 2006.<br />

150


Screening tests of mercury sorbents<br />

This chapter first deals with screening of mercury sorbents in simulated cement<br />

kiln flue gas using elemental mercury source. Then screening tests using mercury<br />

chloride source are reported. The results are used to suggest promising sorbents for<br />

application in cement plant and provide explanation of mercury adsorption in cement<br />

production <strong>by</strong> cement materials.<br />

7.1 Introduction<br />

<strong>Sorbent</strong>s are screened in the laboratory using simulated flue gas before fieldtesting<br />

in actual flue gas. The purpose of these laboratory tests is to evaluate a number of<br />

sorbents at conditions similar to those expected at typical cement plants. These test<br />

results then are used to determine the most appropriate samples for large scale tests.<br />

Basing on the screen tests, promising sorbents will be further investigated in the lab in<br />

detail to obtain adsorption kinetics and study the influence of different operational<br />

parameters. The fixed bed tests are not intended to simulate the conditions where a<br />

sorbent is injected continuously upstream of a fabric filter but they provide a good<br />

indication of sorbent effectiveness, providing the exposure conditions are similar.<br />

Screening measurements are used to evaluate mercury capture effectiveness, oxidation<br />

potential, and capacity for the selected sorbents.<br />

16 sorbent materials are collected and compared. The selected sorbents are tested<br />

in the fixed-bed reactor with continuous mercury measurement, following closely the<br />

experimental procedure as described in chapter 4.<br />

The empty reactor, quartz wool plug, and sand for sorbent dilution are tested first<br />

to investigate whether there is some mercury adsorption <strong>by</strong> the empty reactor, quartz<br />

151<br />

7


wool and sand powder. Then the sorbents are tested in the simulated cement kiln flue gas<br />

using either elemental mercury or mercury chloride source to study whether the sorbents<br />

behavior differently using different mercury sources.<br />

During mercury adsorption tests, the elemental Hg can be fully or partially<br />

oxidized because of reactions between the elemental Hg, sorbent, and flue gas<br />

components. Then the extent of mercury oxidation is calculated <strong>by</strong> comparing the<br />

measured elemental mercury after breakthrough and the inlet level of added elemental<br />

mercury.<br />

The direct result of the fixed-bed test is the mercury adsorption breakthrough<br />

curve. The percentage breakthrough is determined as a function of time <strong>by</strong> normalizing<br />

the measured total mercury concentration at the outlet of the sorbent bed to the inlet<br />

mercury concentration.<br />

From the mercury breakthrough curve, the amount of mercury adsorbed on unit<br />

mass of the sorbent as a function of time can be calculated <strong>from</strong> the expression:<br />

F t<br />

qt� ( C<br />

0<br />

in �Cout,<br />

t ) dt<br />

W � (7.1)<br />

where F is the flow rate through the sorbent bed, W is the mass of the sorbent, Cin is the<br />

inlet mercury concentration, Cout,t is the mercury concentration at the reactor outlet at<br />

time t.<br />

The mercury adsorption capacity of a known weight of a sorbent is calculated in<br />

terms of µg Hg adsorbed/g_sorbent material <strong>from</strong> the breakthrough curve for the sorbent.<br />

The area under the inlet mercury concentration line and a breakthrough curve is used to<br />

determine how much mercury is adsorbed <strong>by</strong> the sorbent. By material balance, the area<br />

between the curve and the line provides the information on the total mercury adsorbed<br />

onto a sorbent if the entire bed reaches equilibrium with mercury vapor. The equilibrium<br />

adsorption capacity is defined <strong>by</strong> the time when the outlet Hg concentration is first equal<br />

to the inlet concentration.<br />

Previous bench-scale studies have reported performance of the sorbents in terms<br />

of the adsorption capacity and/or the time taken for complete breakthrough of Hg <strong>from</strong> a<br />

sorbent [1-13]. However, the final application of the sorbent injection is in a full-scale<br />

152


plant where the sorbent is injected in the duct and captured either in a fabric filter, where<br />

the contact time of mercury with carbon particles is very short. Therefore, the adsorption<br />

rate within these time scales is the most important parameter when evaluating sorbent<br />

performance.<br />

The amount of mercury adsorbed at time t (mt) can be calculated using<br />

m �m �( C �C ). F. � t<br />

(7.2)<br />

t��t t in out, t�1 2�t<br />

where � C � C ) 2<br />

Cout, t�<br />

1 (<br />

2�t<br />

out,<br />

t out,<br />

t�<br />

�t<br />

The adsorption rate at each time step is calculated using<br />

mt��t mt<br />

ratet<br />

dt<br />

�<br />

� (7.3)<br />

The initial rate is evaluated as the slope of the cumulative adsorption curve in the<br />

first 25 min.<br />

7.2 <strong>Sorbent</strong> properties and compositions<br />

The collected sorbent candidates include both commercial sorbents and cement<br />

materials. Virgin activated carbon Dacro Hg, formerly known as Darco FGD [14] is<br />

prepared <strong>from</strong> lignite coal and has been widely studied in the literature [1,2,5-7,15-24]<br />

and is therefore tested here as a reference sorbent. Darco Hg-LH is Darco Hg treated with<br />

bromine and developed for application with low chlorine concentration in the flue gas<br />

<strong>from</strong> combustion of low-rank coals. Activated Lignite HOK is produced according to the<br />

so-called rotary-hearth furnace process [25-27]. Unlike activated carbon, activated<br />

Lignite HOK is produced as mass product with an annual output of 200,000 tons at a<br />

much lower price than that of activated carbon. HOK is the most widely used sorbent for<br />

waste incinerator flue gas cleaning in Europe. Sorbalit is a mixture of reagents, surfaceactive<br />

substances and chemical additives [28]. Reagents are calcium based compounds<br />

such as CaCO3, CaO and Ca(OH)2. Examples of surface-active substances are activated<br />

carbon, aluminum oxide and zeolite. Chemical additives are sulfur and sulfur compounds<br />

such as Na2S, NaHS, Na2S4. Sorbalit can be produced with carbon contents ranging <strong>from</strong><br />

4% to 65%. Minsorb DM and ME are non-carbon based sorbent and for removal of<br />

153


dioxin/furan and mercury, respectively [29,30]. The hydrated lime is standard Sorbacal<br />

product used for SO2 and SO3 removal [31].<br />

<strong>Cement</strong> materials are obtained <strong>from</strong> FLSmidth Dania lab. Clay contains<br />

essentially hydrous aluminum silicates, with minor amount of magnesium, iron, alkalies<br />

or alkaline earths [32]. <strong>Cement</strong> kiln dust is a fine-grained solid material with high<br />

alkaline content removed <strong>from</strong> the cement kiln exhaust gas <strong>by</strong> filters. The cement kiln<br />

dust contains mainly incompletely reacted raw material, including a raw mix at various<br />

stages of burning, and particles of clinker. The primary constituents are silicates, calcium<br />

oxide, carbonates, potassium oxide, sulfates, chlorides, various metal oxides, and sodium<br />

oxide [33].<br />

The kaolin sample is <strong>from</strong> Prolabo Merck. Hydroxyapatite has a formula of<br />

Ca5(PO4)3(OH) and has been used as sorbent to removal of heavy metals <strong>from</strong> waste<br />

incinerators [34]. Initial tests show that hydroxyapatite is a new promising sorbent for<br />

heavy metal removal <strong>from</strong> waste incineration flue gas [34]. Hydroxyapatite is chemically<br />

similar to the mineral component of bones and suitable for biomedical application.<br />

Properties of the sorbents are presented in table 7.1. Carbon-based sorbents have<br />

much larger surface area than the non-carbon based sorbents and cement raw materials.<br />

The volume median diameter D(v,0.5) is the diameter where 50% of the distribution is<br />

above and 50% is below. D(v,0.9) diameter means that 90% of the volume distribution is<br />

below this value. Similarly D(v,0.1) diameter means that 10% of the volume distribution<br />

is below this value. Generally the cement materials have a smaller particle size than the<br />

commercial sorbents. The cement materials are the cheapest due to the availability of<br />

large quantity in the cement plant and saving of transport cost. The bromine treated<br />

Darco Hg-LH carbon is much more expensive than the virgin activated carbons and noncarbon<br />

sorbents. The high price of hydroxyapatite is because it is pharmaceutical grade.<br />

Table 7.1. Properties of sorbents studied in this work.<br />

<strong>Sorbent</strong> D(v,0.1)<br />

�m<br />

D(v,0.5)<br />

�m<br />

154<br />

D(v,0.9)<br />

�m<br />

BET<br />

area<br />

Bulk<br />

density<br />

Price<br />

USD/kg


m 2 /g g/cm 3<br />

Darco Hg 1.27 15.99 43.07 600 0.51 1-2<br />

Darco Hg-LH 1.12 15.36 44.70 550 0.60 2-4<br />

HOK standard 63 300 0.55 1-2<br />

HOK super 24 300 0.44 1-2<br />

Sorbalit 0.85 12.60 52.24 58.5 0.42 1-2<br />

Minsorb DM 6.76 52.02 168.95 120 0.60 1-2<br />

Minsorb ME 3.20 39.17 177.07 70 1.10 1-2<br />

Hydrated lime 0.30 3.35 13.15 21.5 0.35 0.2<br />

Saklei fly ash 3.77 36.15 115.07 0.7 - 0.1<br />

Gypsum 1.63 18.78 62.20 18.5 - 0.1-0.15<br />

Raw meal 0.30 9.47 77.56 1.8 - 0.1<br />

Portland cement 0.32 16.16 46.10 1.8 - 0.1-0.2<br />

<strong>Cement</strong> kiln dust 0.33 3.36 63.93 6.5 - 0.05<br />

Clay 0.36 9.73 58.40 15.2 - 0.1<br />

Kaolin 1.24 5.60 20.65 13.0 - 0.1-0.2<br />

Hydroxyapatite 0.20 3.80 47.99 70.2 - 160<br />

Table 7.2 presents chemical composition of some selected sorbents. The<br />

compositions of HOK carbons and Minsorb sorbents are <strong>from</strong> the literature published <strong>by</strong><br />

the manufacture and the product datasheet [26,27,29,30]. The compositions of Darco<br />

carbons and Sorbalit are obtained <strong>by</strong> averaging 10-20 spot analyses of the samples <strong>by</strong><br />

SEM-EDX. Compositions of other materials are obtained <strong>by</strong> inductively coupled plasma<br />

(ICP) spectrometry. The Darco carbons have larger ash content than the HOK carbons.<br />

SEM-EDX analyses show that the Darco Hg-LH has a bromine content of about 7.8 wt%.<br />

The main elements of Sorbalit are C and Ca, in agreement with the statement <strong>by</strong> the<br />

producer [28]. Minsorb ME has larger Al and Fe contents than Minsorb DM. Ca is the<br />

main element in the raw meal. Kaolin and Saklei fly ash <strong>from</strong> bituminous coal<br />

combustion have similar composition with large Al and Si contents.<br />

155


Table 7.2 Chemical composition of selected sorbents. All in wt%<br />

<strong>Sorbent</strong> Moisture Ash C Cl S K Na Mg Ca Fe Al Si Reference<br />

Darco Hg


7.3 SEM­EDX analysis of fresh sorbents<br />

The main goals of the SEM-EDX analysis is to study the sorbents’ topography<br />

(surface features), morphology (shape and size), and composition. Morphology study<br />

will be used to identify particle agglomeration and compare with particle size<br />

measurement.<br />

Figure 7.1 shows a typical micrograph of the fresh Darco Hg carbon which<br />

has various single carbon particles of irregular surface with different shape and<br />

brightness. Close observation of the big particles at higher magnification shows that<br />

there are many small floc-like particles agglomerated on the big particle. Images at<br />

lower magnification (not shown in figure 7.1) show that most of the particles are<br />

within the range of 5-30 �m and this is in reasonable agreement with the particle size<br />

measurements <strong>by</strong> laser diffraction. However, it should be noted that these SEM<br />

pictures provide only semi-quantitative results of particle sizing since the technique<br />

uses two-dimension information to infer a three-dimensional quantity<br />

Figure 7.1. SEM micrographs of the fresh Darco Hg activated carbon at different<br />

magnifications. Scale bar <strong>from</strong> left to right is 10 and 5 �m, respectively.<br />

Figure 7.2 illustrates the difference in information provided <strong>by</strong> secondary<br />

electron (SE) image and backscattered electron (BSE) image. The SE image is<br />

superior for displaying surface detail and particle morphology but does not generally<br />

show chemical heterogeneity. EDX analysis shows that in the BSE image the small<br />

bright spots in the left (area 7) have high iron content, the bright spot in the center on<br />

157


the big particle (area 4) and big bright particle on the up-right corner (area 1) have<br />

high silica content. Area 5 has high content of calcium and the particle is crystal-like.<br />

The carbon particles have similar brightness level as the carbon substrate on the<br />

carbon table and are not clearly seen in the BSE image.<br />

Figure 7.2. SE (left) and BSE (right) images of the fresh Darco Hg sorbent at the<br />

same location. Positions for SEM-EDX analysis are marked on the BSE image. Scale<br />

bar is 30 �m.<br />

As shown in figure 7.3 the morphology of the fresh Darco Hg-LH is very<br />

similar to that of the fresh Darco Hg and this is not surprised since the Darco Hg-LH<br />

is prepared <strong>from</strong> Darco Hg <strong>by</strong> a brominating process.<br />

Figure 7.3. SE images of fresh Darco Hg-LH activated carbon. Scale bar is 5 �m.<br />

158


Compared to the fresh Darco Hg activated carbon there are high contents of<br />

Na, S, and Br in the Darco Hg-LH sample. The average molar ratio of Na/Br is about<br />

1.74, while the molar ratio of Na/(0.5S+Br) is about 1.18, suggesting the sample is<br />

brominated <strong>by</strong> exposing to NaBr and Na2SO4/Na2SO3/Na2S compounds instead of to<br />

HBr or Br2.<br />

The SE images of fresh Sorbalit sorbent are presented in figure 7.4. The<br />

particles are much less porous than the carbon particles. A thin layer of small crystallike<br />

flakes agglomerate on the big particles. The small dots on the background are<br />

<strong>from</strong> the carbon table for sample holding.<br />

Figure 7.4. SE images of fresh Sorbalit sorbent. Scale bar is 5 �m.<br />

The SE images of the fresh Minsorb ME sorbent are shown in figure 7.5. The<br />

particle size is generally lager than carbon particle size and in agreement with the<br />

particle size measurement <strong>by</strong> the laser diffraction.<br />

159


Figure 7.5. SE images of fresh Minsorb ME sorbent. Scale bar is 50 �m.<br />

7.4 Baseline test<br />

As a starting point, baseline tests of the empty glass reactor, quartz wool plug<br />

and sand powder are conducted first to investigate whether mercury can be adsorbed<br />

<strong>by</strong> these parts and materials. Tests are conducted in both nitrogen and simulated<br />

cement kiln flue gas with either elemental mercury or mercury chloride sources. In all<br />

cases, simultaneous mercury breakthroughs are observed indicating no mercury<br />

adsorption is adsorbed <strong>by</strong> these materials. The mercury exposed sand powder is<br />

analyzed for mercury content in the sample. No mercury is detected in the exposed<br />

sand, which again verifies that no mercury adsorption <strong>by</strong> the sand takes place.<br />

7.5 Screening tests in nitrogen<br />

Preliminary tests of some sorbents were conducted <strong>by</strong> mixing 5-10 mg<br />

sorbent with 2 g sand in nitrogen using elemental mercury source. Only elemental<br />

mercury was measured due to the fact that the problem of converter for total mercury<br />

measurement was not solved at that time. Tests at 150�C show that instantaneous<br />

mercury breakthrough was observed for all the sorbents except the bromine treated<br />

Darco Hg-LH carbon. As shown in figure 7.6, even in nitrogen the bromine treated<br />

160


Darco Hg-LH carbon can both oxidize and adsorb some mercury. Compared to<br />

instantaneous mercury breakthrough observed <strong>by</strong> the non-treated Darco Hg carbon,<br />

mercury adsorption <strong>by</strong> the Darco Hg-LH carbon is due to the promoting effects of<br />

bromine in the Darco Hg-LH carbon. Part of the mercury is probably oxidized on the<br />

Darco Hg-LH carbon <strong>by</strong> the bromine compounds. However, most of the mercury is<br />

still in the form of elemental mercury. Figure 7.6 also shows the breakthrough curve<br />

of 10 mg Darco Hg tested in nitrogen at 150�C with HgCl2 source and total mercury<br />

measurement. In contrast to test using elemental mercury source, it takes about 15 h<br />

to reach the breakthrough. These tests indicate that mercury oxidation is an important<br />

step during mercury adsorption <strong>by</strong> the sorbent. Elemental mercury needs to be<br />

oxidized first either in the gas phase or on the sorbent before being adsorbed <strong>by</strong> the<br />

sorbent.<br />

Gaseous Hg, C out /C in<br />

1.2<br />

0.8<br />

0.4<br />

0<br />

1<br />

1, Hg 0 source, 230 �g/Nm 3 , 5 mg<br />

Darco Hg-LH, Hg 0 measurement<br />

2, HgCl 2 source, 209 �g/Nm 3 , 10 mg<br />

Darco Hg, Hg total measurement<br />

0 2 4 6 8 10 12 14 16 18<br />

Time (hour)<br />

Figure 7.6. <strong>Mercury</strong> breakthrough curves at 150 �C for 5 mg Darco Hg-LH carbon<br />

tested in N2 with elemental source and elemental mercury measurement and 10 mg<br />

Darco Hg carbon tested in N2 with HgCl2 source and total mercury measurement. 2 g<br />

sand as bed mixing material.<br />

161<br />

2


7.6 Screening tests in simulated cement kiln flue gas with<br />

elemental mercury source<br />

Total mercury measurement was conducted using the sulfite-based converter<br />

to obtain mercury breakthrough curves using elemental mercury source in the<br />

simulated cement kiln flue gas. Figure 7.7 illustrates the screening results of 30 mg<br />

different sorbents in 2 g sand at 150�C. From the mercury breakthrough curves, the<br />

amount of adsorbed mercury <strong>by</strong> the sorbent and the average initial adsorption rate for<br />

the first 25 min are calculated and presented in table 7.3. The extents of mercury<br />

oxidation <strong>by</strong> different sorbents are illustrated in figure 7.8.<br />

C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

A B C D<br />

0 1 2 3 4 5 6<br />

Time (hour)<br />

162<br />

E<br />

A:Minsorb ME, 169�g/Nm 3 Hg 0<br />

B: Darco Hg, 180�g/Nm 3 Hg 0<br />

C: HOK super, 167�g/Nm 3 Hg 0<br />

D: Sorbalit, 164�g/Nm 3 Hg 0<br />

E: HOK standard, 171�g/Nm 3 Hg 0<br />

F: Darco Hg-LH, 167�g/Nm 3 Hg 0<br />

Figure 7.7. <strong>Mercury</strong> breakthrough profiles for 30 mg sorbets in 2 g sand tested at<br />

150�C in simulated cement kiln flue gas with 164-180 µg Hg 0 /Nm 3 , 1000 ppmv NO,<br />

23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6<br />

vol.% O2.<br />

F


<strong>Mercury</strong> oxidation (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Darco Hg 16 �m<br />

Darco Hg-LH 15 �m<br />

Figure 7.8. Percentages of mercury oxidation <strong>by</strong> 30 mg sorbets in 2 g sand tested at<br />

150�C in simulated cement kiln flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO,<br />

23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6<br />

vol.% O2.<br />

Table 7.3. <strong>Mercury</strong> 99% breakthrough time, adsorbed mercury and initial adsorption<br />

rates for 30 mg sorbets in 2 g sand tested at 150�C in simulated cement kiln flue gas<br />

with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv<br />

SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

<strong>Sorbent</strong> 99%<br />

breakthrough<br />

time (min)<br />

HOK Super, 24 �m<br />

163<br />

HOK Standard 63 �m<br />

Sobalit super 13 �m<br />

Adsorbed Hg (µg<br />

Hg/g_sorbent)<br />

Minsorb ME 40 �m<br />

Initial adsorption rate<br />

(µg Hg/g_sorbent/h)<br />

Hydroxyapatite 0 0 0<br />

Minsorb DM 0 0 0<br />

Minsorb ME 28 31 70<br />

Sorbalit 110 270 310<br />

HOK super 111 632 690<br />

HOK standard 164 699 560<br />

Darco Hg 90 726 890<br />

Darco Hg-LH 320 1305 690


Neither mercury adsorption nor oxidation is observed <strong>by</strong> 30 mg non-carbon<br />

based sorbents Minsorb DM, hydroxyapatite, and cement materials at 150�C. The<br />

bromine treated carbon Darco Hg-LH has larger adsorption capacity but smaller<br />

adsorption rate compared to the non-treaded Darco Hg carbon. As shown in figure<br />

7.8 and table 7.3, there is a clear trend between the extent of mercury oxidation and<br />

amount of adsorbed mercury. Generally larger amount of adsorbed mercury is<br />

obtained with sorbents that have larger mercury oxidation capacity. The initial<br />

adsorption rate of coarse HOK standard carbon is slightly smaller than the fine HOK<br />

super due to the larger diffusion resistance within the larger carbon particles. Sorbalit,<br />

which is a mixture of lime and carbon, shows poorer performance than the carbons.<br />

Minsorb ME, which is aluminumsilicates based sorbent shows the poorest<br />

performance among the tested commercial sorbents despite that it has much larger<br />

surface area than the Sorbalit sorbent. This is probably due to its capacity for mercury<br />

oxidation is much smaller than the Sorbalit sorbent.<br />

The adsorption of mercury in the Darco Hg carbon is attempted <strong>by</strong> analyzing<br />

the mercury content in the exposed carbon sample. Table 7.4 compares the measured<br />

and calculated mercury contents in the carbons <strong>from</strong> the breakthrough curve. The<br />

calculated mercury contents are much larger than the measured values for the carbon<br />

and sand mixtures. The analysis of mercury content in the sample uses only 100 mg<br />

of the sample for analysis and one reason for the disagreement could be that the<br />

sample analyzed might not be representative. Only 30 mg carbon is mixed with 2 g<br />

sand and the carbon may separate <strong>from</strong> the sand. This is often observed during<br />

loading the sample to the reactor. To ensure most of the carbon is loaded to the<br />

reactor, the sample holder is shaken to remove the carbon deposited on the sample<br />

holder and a big quartz wool plug is used to clean carbon deposited on the reactor<br />

wall and move the carbon to the fixed-bed bed. The carbon particle might deposit on<br />

the container wall and therefore the analyzed sample could contain relatively more<br />

sand powder. The mercury content in the carbon is calculated <strong>from</strong> the measured<br />

mercury level in the carbon-sand mixture and the carbon-sand mixing ratio. Since no<br />

mercury adsorbed <strong>by</strong> the sand powder, analysis using non-representative carbon-sand<br />

mixture could result in small mercury content in the carbon.<br />

164


Table 7.4. Comparison of measured and calculated mercury contents in the carbons<br />

<strong>from</strong> the breakthrough curve. Flue gas composition: 141-183 µg Hg 0 /Nm 3 , 1000<br />

ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.%<br />

CO2, and 6 vol.% O2.<br />

<strong>Sorbent</strong> description Measured Hg in Calculated results <strong>from</strong><br />

carbon (ppmm) breakthrough (ppmm)<br />

30 mg Darco Hg in 2g sand, 150°C,<br />

141 µg/Nm 3 Hg in flue gas<br />

2.22 5.96<br />

30 mg Darco Hg in 2g sand, 150°C,<br />

183 µg/Nm 3 Hg in flue gas<br />

2.27 13.14<br />

30 mg Darco Hg in 2g sand, 150°C.<br />

143 µg/Nm 3 Hg in flue gas<br />

1.49 7.06<br />

500 mg Darco Hg, 200°C, 160 µg/Nm 3<br />

Hg in flue gas<br />

46.79 50.87<br />

To check whether the method for calculating the mercury content in the<br />

carbon is reasonable, a new test was performed <strong>by</strong> using only carbon sample to avoid<br />

the problem of non-representative sample caused <strong>by</strong> carbon-sand mixing. As shown<br />

in table 7.4, the measured mercury content in the carbon is about 92% of the<br />

calculated value <strong>from</strong> the breakthrough curve. This reasonable agreement between the<br />

measured and calculated value confirms that the method of calculating mercury<br />

content in the carbon <strong>from</strong> the breakthrough curve works to a satisfactory extent.<br />

To be able to observe some mercury adsorption <strong>by</strong> the cement materials, the<br />

adsorption temperature was decreased to 75�C. However, still no mercury adsorption<br />

was observed <strong>by</strong> 30 mg cement materials at 75�C. Then the sorbent load is increased<br />

to 2 g. Among the tested cement materials only raw meal shows some mercury<br />

adsorption as shown in figure 7.9. This can to some extent explain the low mercury<br />

emission <strong>from</strong> cement plants during raw mill-on period. The dust load in the flue gas<br />

after the raw mill could be up to 800-1000 g/m 3 and therefore noticeable amount of<br />

mercury could be adsorbed <strong>by</strong> the raw meal both in and after the raw mill.<br />

165


Gaseous Hg (�g/Nm 3 )<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

<strong>by</strong>pass<br />

reactor<br />

through reactor<br />

Gaseous Hg total<br />

Calculated Hg in sorbent<br />

0 30 60 90 120 150<br />

Time (min)<br />

166<br />

0.0016<br />

0.0012<br />

0.0008<br />

0.0004<br />

Figure 7.9. <strong>Mercury</strong> breakthrough profile and calculated mercury adsorption in 2 g<br />

cement raw meal tested at 75�C in simulated cement kiln flue gas with 160-170 µg<br />

Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.%<br />

H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

7.7 Screening tests in simulated cement kiln flue gas with HgCl2<br />

source<br />

The collected sorbents are also tested in simulated cement kiln flue gas with<br />

HgCl2 source. Figure 7.10 shows the mercury breakthrough curves for 10 mg<br />

sorbents in 2 g sand at 150�C using simulated cement kiln flue gas with170±10<br />

µg/Nm 3 mercury <strong>from</strong> HgCl2 source. The 99% breakthrough time, calculated amount<br />

of adsorbed mercury <strong>by</strong> the sorbent <strong>from</strong> the breakthrough curve, and the average<br />

initial adsorption rate for the first 25 min are presented in table 7.5.<br />

0<br />

Calculated Hg in sorbent (mg Hg/g_sorbent)


Figure 7.10. <strong>Mercury</strong> breakthrough profiles of 10 mg sorbents tested in 2g sand at<br />

150�C in simulated cement kiln flue gas with HgCl2 source. The inlet mercury level<br />

is 170±10 µg/Nm 3 , other gases include 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl,<br />

1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

Table 7.5. <strong>Mercury</strong> 99% breakthrough time, adsorbed mercury and initial adsorption<br />

rates for 10 mg sorbets in 2 g sand tested at 150�C in simulated cement kiln flue gas<br />

using HgCl2 source. The inlet mercury level is 170±10 µg/Nm 3 , other gases include<br />

1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21<br />

vol.% CO2, and 6 vol.% O2.<br />

<strong>Sorbent</strong> 99%<br />

breakthrough<br />

time (min)<br />

Adsorbed Hg (µg<br />

Hg/g_sorbent)<br />

167<br />

Initial adsorption rate<br />

(µg Hg/g_sorbent/h)<br />

Hydroxyapatite 0 0 -<br />

Minsorb DM 5 12 -<br />

Minsorb ME 49 363 780<br />

Sorbalit 86 429 730<br />

HOK standard 74 1021 1390<br />

HOK super 55 1153 2170<br />

Darco Hg 43 1224 2510<br />

Darco Hg-LH 52 1290 2510


The hydroxyapatite sorbent still does not adsorb any mercury even using the<br />

mercury chloride source. This means that the mercury adsorption <strong>by</strong> hydroxyapatite<br />

is not only limited <strong>by</strong> its ability of oxidizing mercury, but also other properties. The<br />

Minsorb DM sorbent shows low adsorption of HgCl2 compared to no adsorption of<br />

elemental mercury. Minsorb ME and Sorbalit show similar mercury adsorption in<br />

terms of mercury adsorption capacity and initial adsorption rate. All the carbons show<br />

similar mercury adsorption capacity; while the HOK standard has the smallest initial<br />

adsorption rate. Compared to similar initial adsorption rate of elemental mercury, the<br />

initial adsorption rate of HgCl2 for HOK super is about 50% larger than the HOK<br />

standard. The elemental mercury adsorption capacity of Darco Hg-LH is about 79%<br />

larger and initial Hg 0 adsorption rate is about 23% smaller in comparison with the<br />

virgin Darco Hg carbon. Similar HgCl2 adsorption capacity and initial adsorption rate<br />

of Darco Hg and Darco Hg-LH indicate that Darco Hg is a better choice at least for<br />

removing HgCl2 <strong>from</strong> cement kiln flue gas.<br />

<strong>Cement</strong> materials were also tested for HgCl2 capture <strong>from</strong> simulated cement<br />

kiln flue gas using a sorbent load of 2 g without mixing with sand powder. Figure<br />

7.11 presents the mercury breakthrough curves of 2 g cement materials tested at<br />

150�C using simulated cement kiln flue gas with170±10 µg/Nm 3 mercury <strong>from</strong><br />

HgCl2 source. The 99% breakthrough time and calculated amount of adsorbed<br />

mercury <strong>by</strong> the sorbent <strong>from</strong> the breakthrough curve are given in table 7.7. The<br />

fluctuation of some breakthrough curves is due to the aging of the sulfite-based<br />

converter material. After changing the converter material used for about 3 months<br />

smooth mercury breakthrough is obtained again.<br />

168


Gaseous Hg (C out /C in )<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

A B<br />

F<br />

E<br />

C<br />

0 0.2 0.4 0.6 0.8 1<br />

Time (hour)<br />

D<br />

A: 2 g hydrated lime<br />

B: 2 g clay<br />

C: 2 g kaolin<br />

D: 10 mg Darco Hg in 2 g sand<br />

E: 2 g cement kiln dust<br />

F: 2 g gypsum<br />

Figure 7.11. <strong>Mercury</strong> breakthrough profiles of 2 g cement materials tested at 150�C<br />

in simulated cement kiln flue gas with HgCl2 source. The inlet mercury level is<br />

170±10 µg/Nm 3 , other gases include 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl,<br />

1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. Data of 10 mg Darco<br />

Hg carbon in 2 g sand are shown for comparison.<br />

Table 7.7. <strong>Mercury</strong> breakthrough time and adsorbed mercury for 2 g cement materials<br />

tested at 150�C in simulated cement kiln flue gas using HgCl2 source. The inlet<br />

mercury level is 170±10 µg/Nm 3 , other gases include 1000 ppmv NO, 23 ppmv NO2,<br />

10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

<strong>Sorbent</strong>s Breakthrough time (min) Adsorbed Hg (µg<br />

Hg/g_sorbent)<br />

Hydrated lime 11 0.60<br />

Clay 14 0.90<br />

Kaolin 23 1.77<br />

<strong>Cement</strong> kiln dust 60 1.81<br />

Gypsum 49 1.73<br />

Raw meal 54 0.90<br />

Saklei fly ash 38 1.01<br />

Portland cement 60 2.28<br />

169<br />

Gaseous Hg (C out /C in )<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

G<br />

I<br />

H<br />

D: 10 mg Darco Hg in 2 g sand<br />

G: 2 g raw meal<br />

H: 2 g Portland cement<br />

I: 2 g Saklei fly ash<br />

0 0.2 0.4 0.6 0.8 1<br />

Time (hour)<br />

D


Compared to null adsorption of elemental mercury at temperature as low as<br />

75�C, all the tested cement materials show some adsorption of mercury chloride<br />

at150�C. The lack of ability for mercury oxidation is probably the main limitation for<br />

these materials to be used for mercury removal <strong>from</strong> flue gas, which contains both<br />

elemental and oxidized mercury. 2 g of kaolin, cement kiln dust, gypsum, and<br />

Portland cement adsorb similar amount of mercury chloride as 10 mg Minsorb ME<br />

and Sorbalit commercial sorbents. Raw meal, hydrated lime, clay, Saklei fly ash have<br />

similar mercury adsorption capacities which are about half of the capacities of kaolin,<br />

cement kiln dust, gypsum, and Portland cement. These results can further explain low<br />

mercury emission <strong>from</strong> the cement plant during raw mill-on period, as about 55-65 %<br />

of the mercury in the cement kiln flue gas is oxidized mercury [35]. Considering the<br />

low cost and abundance of the cement materials, injection of cement materials for<br />

mercury control in cement plant is feasible provided that the elemental mercury in the<br />

flue gas can be oxidized <strong>by</strong> adding an oxidant. However, raw meal, hydrated lime,<br />

clay, cement kiln dust, and kaolin have to be recycled to the kiln in the cement<br />

production and the adsorbed mercury will be released again in the hot zone. If these<br />

materials are not recycled the disposal cost will be high since larger amount of these<br />

materials have to be used compared to activated carbon for same amount of mercury<br />

removal. Gypsum and Saklei fly ash can be added to the final cement product and the<br />

release of captured mercury and high disposal cost are avoided. This also applied for<br />

the Portland cement. However, the stability of mercury in the final cement product<br />

requires further investigation.<br />

7.8 Conclusions<br />

Screening tests of sorbents for mercury removal <strong>from</strong> cement plants have<br />

been conducted in the fixed-bed reactor system. The tested sorbents include<br />

commercial activated carbons, commercial non-carbon sorbents, and cement<br />

materials. Screening measurements are used to evaluate initial mercury capture rate,<br />

oxidation potential, and capacity for the selected sorbents. The amount of mercury<br />

adsorbed is calculated <strong>from</strong> the mercury breakthrough curve and the initial mercury<br />

170


adsorption rate is further evaluated for application regarding sorbent injection<br />

upstream of a fabric filter.<br />

Baseline tests of empty reactor, quartz wool plug, and sand powder show that<br />

no mercury adsorption is observed either in nitrogen or simulated cement kiln flue<br />

gas with elemental mercury or mercury chloride source.<br />

Initial tests of sorbent in nitrogen with elemental mercury at 150�C find that<br />

only the bromine treated Darco Hg-LH activated carbon shows some mercury<br />

adsorption among the collected sorbents. However, the virgin Darco Hg carbon<br />

adsorbs mercury chloride in nitrogen. This indicates that mercury oxidation is an<br />

important factor for mercury adsorption <strong>by</strong> the sorbents. Elemental mercury needs to<br />

be oxidized either in the flue gas or on the sorbent.<br />

Tests a collection of sorbents (30 mg in 2 g sand) at 150�C in simulated<br />

cement kiln flue gas with elemental mercury show that no mercury adsorption or<br />

oxidation takes place on the non-carbon based sorbents Minsorb DM, hydroxyapatite,<br />

and cement materials. The mercury adsorption capacity of bromine treated carbon<br />

Darco Hg-LH is 79% larger than the non-treated Darco Hg carbon, but the initial<br />

adsorption rate is 23% smaller. Generally a larger amount of adsorbed mercury is<br />

obtained with sorbents that have larger mercury oxidation capacity. A lower amount<br />

of mercury is adsorbed <strong>by</strong> the HOK carbon compared to Darco Hg carbon, probably<br />

be due to both the smaller surface area and mercury oxidation capacity of the HOK<br />

carbon. The initial adsorption rate of coarse HOK standard carbon is slightly lower<br />

than the fine HOK super due to the larger diffusion resistance within the larger HOK<br />

standard carbon particles. Sorbalit shows poorer performance than the carbons, while<br />

Minsorb ME shows the poorest performance among the tested commercial sorbents<br />

despite that it has much larger surface area than the Sorbalit sorbent. Among the<br />

tested sorbents Darco Hg has the largest initial adsorption rate of elemental mercury.<br />

The collected sorbents are also tested in simulated cement kiln flue gas with<br />

mercury chloride using 10 mg sorbents in 2 g sand at 150�C. The hydroxyapatite<br />

sorbent still does not adsorb any mercury. The Minsorb DM sorbent shows negligible<br />

adsorption of HgCl2 compared to no adsorption of elemental mercury. Minsorb ME<br />

and Sorbalit show similar mercury adsorption of about 400 µg Hg/g_sorbent. All the<br />

171


carbons show similar mercury adsorption capacity; while the HOK standard has the<br />

smallest initial adsorption rate. Similar HgCl2 adsorption capacity and initial<br />

adsorption rate of Darco Hg and Darco Hg-LH indicate that Darco Hg is a better<br />

choice at least for removing HgCl2 <strong>from</strong> cement kiln flue gas.<br />

Compared to non-observable adsorption of elemental mercury on 30 mg<br />

sample at temperature as low as 75�C, all the tested cement materials show some<br />

adsorption of mercury chloride at 150�C using a sorbent load of 2 g. Similar amount<br />

of mercury chloride adsorption is observed <strong>by</strong> 2 g of kaolin, cement kiln dust,<br />

gypsum, Portland cement, and 10 mg Minsorb ME, Sorbalit commercial sorbents.<br />

Among the tested sorbents the Darco Hg activated shows the best<br />

performance of adsorption of both elemental and oxidized mercury, with the largest<br />

initial adsorption rate and second largest mercury adsorption capacity and a lower<br />

price than the treated carbon. Therefore, the Darco Hg carbon is recommended as the<br />

reference sorbent for a fundamental investigation of mercury adsorption in simulated<br />

cement kiln flue gas and large-scale tests. Adsorption <strong>by</strong> cement materials at larger<br />

load can explain the phenomena of low mercury emission <strong>from</strong> the cement plant<br />

during raw mill-on period, when larger amount of cement materials are present in the<br />

flue gas at relative low temperature. Considering the low cost and abundance of the<br />

cement materials, injection of cement materials for mercury control in cement plant is<br />

feasible provided that the elemental mercury in the flue gas can be oxidized <strong>by</strong><br />

adding of oxidant. Compared to raw meal, clay, kaolin, and cement kiln dust, gypsum,<br />

Saklei fly ash, and Portland cement are more preferred due to avoidance of captured<br />

mercury release in the kiln and high disposal cost since these materials will be added<br />

to the finished cement product. However, the stability of mercury in the exposed<br />

cement materials requires further investigation to study whether it will be released<br />

<strong>from</strong> the final cement product.<br />

7.9 References<br />

[1] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R. Chang,<br />

Preparation and evaluation of coal-derived activated carbons for removal of mercury vapor<br />

<strong>from</strong> simulated coal combustion flue gases, Energy & Fuels. 12 (1998) 1061-1070.<br />

172


[2] M. Rostam-Abadi, S.G. Chen, H.C. Hsi, M. Rood, R. Chang, T.R. Carey, B. Hargrove, C.<br />

Richardson, W. Rosenhoover, F. Meserole, Novel vapor phase mercury sorbents,<br />

Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant Control, Washington,<br />

DC, Aug 25–29, 1997.<br />

[3] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang and F.B. Meserole, Factors<br />

affecting mercury control in utility flue gas using sorbent injection, Proceedings of the Air &<br />

Waste Management Association's 90th Annual Meeting, Toronto, Ontario, Canada, June 8-13,<br />

1997.<br />

[4] C.X. Hu, J.S. Zhou, Z.Y. Luo, S. He, G.K. Wang, K.F. Cen, Effect of oxidation treatment<br />

on the adsorption and the stability of mercury on activated carbon, Journal of Environmental<br />

Sciences-China. 18 (2006) 1161-1166.<br />

[5] B. Ghorishi and B.K. Gullett, Fixed-bed control of mercury: Role of acid gases and a<br />

comparison between carbon-based, calcium-based, and coal fly ash sorbents, Proceedings of<br />

the EPRI/DOE/EPA Combined Utility Air Pollutant Control Symposium, Washington, DC,<br />

August 25–29, 1997.<br />

[6] S.B. Ghorishi and C.B. Sedman, Combined mercury and sulfur oxides control using<br />

calcium-based sorbents, Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant<br />

Control Symposium, Washington, DC, August 25–29, 1997.<br />

[7] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption <strong>by</strong> activated<br />

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy<br />

Conference, Research Triangle Park, NC, 22-25 April, 1997.<br />

[8] N. Hutson, C. Singer, C. Richardson, J. Karwowski and C. Sedman, Practical applications<br />

<strong>from</strong> observations of mercury oxidation and binding mechanisms, EPRI-DOE-EPA-AWMA<br />

Combined Power Plant Air Pollutant Control Mega Symposium, Washington DC, August 30<br />

- September 2, 2004.<br />

[9] E.J. Granite, M.C. Freeman, R.A. Hargis, W.J. O'Dowd, H.W. Pennline, The thief process<br />

for mercury removal <strong>from</strong> flue gas, J. Environ. Manage. 84 (2007) 628-634.<br />

[10] W.J. O’Dowd, H.W. Pennline, M.C. Freeman, E.J. Granite, R.A. Hargis, C.J. Lacher, K.<br />

Andrew, A technique to control mercury <strong>from</strong> flue gas: The thief process, Fuel Processing<br />

Technology. 87 (2006) 1071-1084.<br />

[11] R. Bhardwaj. Impact of temperature and flue gas components on mercury speciation and<br />

uptake <strong>by</strong> activated carbon sorbents . Master thesis. University of Pittsburgh, 2007.<br />

[12] M.M. Maroto-Valer, Y. Zhang, E.J. Granite, Z. Tang, H.W. Pennline, Effect of porous<br />

structure and surface functionality on the mercury capacity of a fly ash carbon and its<br />

activated sample, Fuel. 84 (2005) 105-108.<br />

[13] S. Eswaran, H.G. Stenger, Z. Fan, Gas-phase mercury adsorption rate studies, Energy &<br />

Fuels. 21 (2007) 852-857.<br />

[14] J.R. Butz, T.E. Broderick and C.S. Turchi, Amended Silicates for <strong>Mercury</strong> Control,<br />

project final report, DOE Award Number: DE-FC26-04NT41988, 2006.<br />

173


[15] J.H. Pavlish, E.A. Sondreal, M.D. Mann, E.S. Olson, K.C. Galbreath, D.L. Laudal, S.A.<br />

Benson, Status review of mercury control options for coal-fired power plants, Fuel<br />

Processing Technology. 82 (2003) 89-165.<br />

[16] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors<br />

affecting mercury control in utility flue gas using activated carbon, Journal of the Air &<br />

Waste Management Association. 48 (1998) 1166.<br />

[17] N. Hutson, C. Singer, C. Richardson and J. Karwowski, Practical applications <strong>from</strong><br />

observations of mercury oxidation and binding mechanisms, The fifth mega symposium on<br />

air pollutant controls for power plants, Washington, DC, August, 2004.<br />

[18] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for DTE<br />

Energy’s Monroe Station, DOE Award Number DE-FC26-03NT41986, Report Number<br />

41986R16, 2006.<br />

[19] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for<br />

Sunflower Electric’s Holcomb Station, DE-FC26-03NT41986, Topical Report No. 41986R07,<br />

2005.<br />

[20] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for<br />

AmerenUE’s Meramec Station Unit 2, DE-FC26-03NT41986, Topical Report No. 41986R09,<br />

2005.<br />

[21] S. Sjostrom, Evaluation of sorbent injection for mercury control: Topical report for<br />

Basin Electric Power Cooperative’s Laramie River Station, DE-FC26-03NT41986, Topical<br />

Report No. 41986R11, 2006.<br />

[22] ADA-ES, Field test program to develop comprehensive design, operating, and cost data<br />

for mercury control systems: Final site report for E.C. Gaston Unit 3 sorbent injection into<br />

COHPAC for mercury control, DE-FC26-00NT41005, Topical Report No. 41005R11, 2003.<br />

[23] ADA-ES, Field test program to develop comprehensive design, operating, and cost data<br />

for mercury control systems: Final site report for Brayton Point Generating Station Unit 1<br />

sorbent injection into a cold-side ESP for mercury control, DE-FC26-00NT41005, Topical<br />

Report No. 41005R21, 2005.<br />

[24] ADA-ES, Field test program to develop comprehensive design, operating, and cost data<br />

for mercury control systems: Final site report for PG&E NEG Salem Harbor Station Unit 1<br />

sorbent injection into a cold-side ESP for mercury control, U.S. DOE Cooperative Agreement<br />

No. DE-FC26-00NT41005, 2004.<br />

[25] J. Wirling, Implementation of process-integrated waste gas cleaning using activated<br />

lignite, A&WMA specialty conference on hazardous waste combustors, Kansas City, Kansas,<br />

March 28-30, 2001, .<br />

[26] J. Wirling, Adsorptive waste gas cleaning in an industrial-scale coal-fired power plant,<br />

A&WMA specialty conference on mercury emissions, Chicago, Illinos, August 21-23, 2001, .<br />

[27] J. Wirling, J. Jablonski, Safety aspects in the use of carbonaceous sorbents during waste<br />

gas treatment, Metallurgical Plant and Technology. 3 (2007) 144.<br />

174


[28] A. Licata, M. Babu and L. Nethe, Acid gases, mercury, and dioxin <strong>from</strong> MWCs,<br />

National Waste Processing Conference Proceedings ASME, Boston, Massachusetts, June 5-8,<br />

1994.<br />

[29] Lhoist Group, Data sheet of Minsorb ® DM sorbent, 2010.<br />

[30] Lhoist Group, Data sheet of Minsorb ® ME sorbent, 2010.<br />

[31] Lhoist Group, Sorbacal flue gas treatment, http://www.sorbacal.com/, accessed January<br />

25, 2011.<br />

[32] J.I. Bhatty, F.M. Miller, S.H. Kosmatka, (Eds.), Innovations in Portland cement<br />

manufacturing, Portland <strong>Cement</strong> Association, Skokie, Illinois, U.S.A, 2004.<br />

[33] U.S. EPA, Materials characterization paper in support of the proposed rulemaking:<br />

Identification of nonhazardous secondary materials that are solid waste-cement kiln dust<br />

(CKD), 2010.<br />

[34] C. Verwilghen, S. Rio, J. Ramaroson, A. Nzihou and P. Sharrock, The use of<br />

hydroxyapatite for the removal of heavy metals <strong>from</strong> industrial flue gas PART B:<br />

Investigation in pilot scale, DustConf 2007, Maastricht, The Netherlands, April 23-24, 2007.<br />

[35] R.J. Schreiber and C.D. Kellett, Compilation of mercury emission data, PCA R&D<br />

Serial No. SN3091, 2009.<br />

175


Fundamental investigation of elemental<br />

mercury adsorption <strong>by</strong> activated carbon in<br />

simulated cement kiln flue gas<br />

This chapter reports a fundamental investigation of elemental mercury<br />

adsorption <strong>by</strong> Darco Hg activated carbon in simulated cement kiln flue gas. The<br />

investigation includes the effects of temperature and gas composition on mercury<br />

adsorption kinetics and equilibrium uptake.<br />

8.1 Introduction<br />

The investigation is mainly conducted using 10 mg Darco Hg mixed with 2 g<br />

sand in simulated cement flue gas of a baseline composition of 1000 ppmv NO, 23<br />

ppmv NO2, 1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 21 vol.% CO2, 6 vol.% O2.<br />

Table 8.1 shows the investigated parameters and the tested ranges. The idea of using<br />

wide range of parameters is to simulate possible cement kiln flue gas composition<br />

and derive kinetics correlations that can be used to predict mercury adsorption <strong>by</strong><br />

activated carbon under different conditions. The isotherms and kinetics are obtained<br />

using different adsorption temperatures, mercury inlet levels, and gas composition.<br />

The percentages of mercury oxidation are also investigated <strong>by</strong> measuring the<br />

elemental mercury level after the complete mercury breakthrough is obtained and<br />

comparison with inlet elemental mercury concentration.<br />

Another commercial carbon, Norit RB4, is also investigated to the study the<br />

effects of carbon particle size. The granular Norit RB4 has a diameter of 4 mm and<br />

length of 10 mm, respectively. The Norit RB4 pellet has a surface area of 1060-1320<br />

m 2 /g, a microporous volume of 0.41-0.54 cm 3 /g, 0.8% of water, and 5.6% of ash [1,2].<br />

The pellets are crushed and sieved for studying the effects of particle size.<br />

176<br />

8


Table. 8.1. Parameters for lab-scale fundamental investigation of elemental mercury<br />

adsorption <strong>by</strong> the activated carbon.<br />

Parameters Baseline values Range tested<br />

Flue gas rate (Nl/min) 2.75 1.1-2.75<br />

Adsorption temperature (�C) 150 75-250<br />

Gas composition<br />

Hg 0 (µg/Nm 3 ) 160-170 0-170<br />

NO (ppmv) 1000 100-1000<br />

NO2 (ppmv) 23 0-100<br />

SO2 (ppmv) 1000 100-1000<br />

HCl (ppmv) 10 0-20<br />

CO (ppmv) 0 0-1000<br />

H2O (vol.%) 1 0-15<br />

CO2 (vol.%) 21 1-31<br />

O2 (vol.%) 6 1-16<br />

8.2 Effect of adsorption temperature<br />

Figure 8.1 shows the effect of adsorption temperature on mercury<br />

breakthrough profiles after the carbon bed. As expected, faster mercury breakthrough<br />

is obtained at higher adsorption temperature. This pronounced effect of temperature<br />

on the mercury adsorption capacity of the activated carbon evidences a physical<br />

adsorption mechanism between the mercury and Darco Hg carbon. Physical<br />

adsorption <strong>from</strong> the gas phase is accompanied <strong>by</strong> a decrease in free energy of the<br />

system [3,4]. The gaseous molecules in the adsorbed state have fewer degrees of<br />

freedom than in the gaseous state. This results in a decrease in entropy during<br />

adsorption. Using the thermodynamic relationship:<br />

�G ��H �T� S<br />

(8.1)<br />

It follows that the term ΔH, which is the heat of adsorption, must be negative<br />

indicating that adsorption is always an exothermic process, respective of the nature of<br />

the forces involved in the adsorption process.<br />

177


Gaseous Hg, C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1 2 3<br />

0 0.5 1 1.5 2 2.5 3 3.5 4<br />

Time (hour)<br />

178<br />

1: 150 o C<br />

2:120 o C<br />

3:75 o C<br />

Figure 8.1. Effect of adsorption temperature on mercury breakthrough of 10 mg<br />

Darco Hg mixed with 2 g sand using 2.75 Nl/min simulated flue gas with 160-170 µg<br />

Hg 0 /Nm 3 ,1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.%<br />

H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

The effect of temperature on the extent of mercury oxidation <strong>by</strong> the Darco Hg<br />

carbon is present in figure 8.2. The adsorption temperature does not affect the<br />

oxidation of mercury <strong>by</strong> the Darco Hg carbon. The mercury oxidation is always<br />

larger than 92% and the average mercury oxidation percentage is about 97% in the<br />

studied temperature range of 75-250�C.


<strong>Mercury</strong> oxidation (%)<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

50 100 150 200 250 300<br />

Temperature ( 0 C)<br />

Figure 8.2. Effect of adsorption temperature on mercury oxidation <strong>by</strong> 10 mg Darco<br />

Hg mixed with 2 g sand using 2.75 Nl/min simulated flue gas with 160-170 µg<br />

Hg 0 /Nm 3 ,1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.%<br />

H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

8.3 Isotherm tests<br />

In order to simulate the performance of a given sorbent, the adsorption<br />

equilibrium information such as the isotherm and characteristics of the sorbent must<br />

be known. The adsorption isotherm is the most extensively employed method for<br />

representing the equilibrium states of an adsorption system [3,4]. It can give useful<br />

information regarding the adsorbate, the adsorbent, and the adsorption process. It<br />

helps in the determination of the heat of adsorption, and the relative absorbability of a<br />

gas on a given adsorbent.<br />

<strong>Sorbent</strong> equilibrium data can be generated <strong>by</strong> conducting adsorption<br />

breakthrough tests in the fixed-bed reactor. Figure 8.3 illustrates the mercury<br />

breakthrough curves of 10 mg Norit Hg activated carbon tested at 120�C with<br />

different elemental mercury inlet levels. The time necessary for saturation of 10 mg<br />

carbon is in the order of 0.6-1.2 h for the elemental mercury inlet level of 27-95<br />

µg/Nm 3 . It takes longer time to reach the complete breakthrough when the mercury<br />

inlet level is lower. This is in agreement with the observation <strong>by</strong> Karatza et al. [5] that<br />

179


the saturation time decreased when the inlet mercury level was increased <strong>from</strong> 1 to<br />

5.5 mg/m 3 . The driving force of mercury adsorption is the difference between the<br />

amount of adsorbed mercury <strong>by</strong> unit carbon at a particular mercury inlet<br />

concentration and the theoretical amount of mercury that could be adsorbed <strong>by</strong> unit<br />

carbon at that concentration and this driving force disappears when the adsorption<br />

gradually approaches its equilibrium state. Initially the rate of adsorption is large as<br />

the whole carbon surface is bare but as more and more of the surface becomes<br />

covered <strong>by</strong> the mercury molecules, the available bare surface decreases and so does<br />

the rate of adsorption. The driving force theory can therefore explain the sigmoidal<br />

shape of the breakthrough curve. The driving force for higher mercury inlet level is<br />

larger at the initial stage of the adsorption (the first 12 min as shown in figure 8.3)<br />

and becomes smaller and similar for all the applied mercury inlet levels due to the<br />

accumulation of mercury in the carbon. As a result, faster mercury breakthrough is<br />

obtained for larger mercury inlet concentration.<br />

Gaseous Hg, C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Time (hour)<br />

180<br />

1 2 3<br />

1: 95 �g/Nm 3 Hg 0<br />

2: 57 �g/Nm 3 Hg 0<br />

3: 27 �g/Nm 3 Hg 0<br />

Figure 8.3. Effect of elemental mercury inlet level on mercury breakthrough of 10 mg<br />

Darco Hg mixed with 2 g sand tested at 120�C using 2.75 Nl/min simulated flue gas<br />

with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21<br />

vol.% CO2, and 6 vol.% O2.<br />

The isotherm studies are conducted at 75, 100, 120, and 150�C. As shown in<br />

figure 8.4 there is a linear correlation between the amounts of mercury adsorbed on<br />

the unit mass of carbon and the inlet mercury concentrations. For the mercury levels


applied in this work (


plots lnk as a function of the reciprocal of temperature. Since adsorption is<br />

exothermic, the Henry constant decreases with temperature. There is a linear relation<br />

between lnk and 1/T and <strong>from</strong> the slope and intercept the calculated value for k0 and<br />

�H ads is 0.869 m 3 /g and -8543 J/mol, respectively. In the work of Karatza et al. [5] a<br />

heat of adsorption of -22000 J/mol was found for Darco G60 activated carbon tested<br />

in nitrogen. The Darco G60 carbon has surface area of 600 m 2 /g and is typically used<br />

for treating fine chemicals and pharmaceutical intermediates [9]. Calculated binding<br />

energy of elemental mercury on activated carbon at room temperature using density<br />

functional theory and fused-benzene ring cluster approach is -18100 J/mol [10].<br />

Effects of other flue gas constituents have not been considered in the simulations. The<br />

derived heat of adsorption for Darco Hg carbon in simulated cement kiln flue gas is<br />

about half of both the experimental data for Darco G60 in nitrogen and theoretical<br />

calculation of binding energy of elemental mercury on activated carbon in nitrogen at<br />

room temperature.<br />

ln(k)<br />

-4<br />

-4.2<br />

-4.4<br />

-4.6<br />

Data<br />

Y=1027.4881X-7.0484<br />

R 2 =0.93<br />

-4.8<br />

0.0022 0.0024 0.0026 0.0028 0.003<br />

1/T (1/K)<br />

Figure 8.5. Plot of lnk as a function of 1/T. The Henry’s constants k are derived <strong>from</strong><br />

isotherms at 75, 100, 120, and 150�C. 10 mg Darco Hg mixed with 2 g sand is tested<br />

using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv<br />

HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

The oxidation of mercury <strong>by</strong> the Darco Hg carbon seems not to be affected <strong>by</strong><br />

the elemental mercury inlet level, as shown in figure 8.6. The mercury oxidation is<br />

182


always larger than 90% and the average mercury oxidation percentage is about 96%<br />

in the studied elemental mercury inlet level of 18-180 µg/Nm 3 . While almost no<br />

mercury oxidation takes place on the activated carbon tested in nitrogen. It is<br />

expected that the heat of adsorption is different for elemental mercury and oxidized<br />

mercury adsorption <strong>by</strong> the carbon. This might explain the difference between the<br />

derived heat of adsorption <strong>from</strong> this work and both the experimental data for Darco<br />

G60 in nitrogen and theoretical calculation of binding energy of elemental mercury<br />

on activated carbon in nitrogen.<br />

<strong>Mercury</strong> oxidation (%)<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 40 80 120 160 200<br />

Hg inlet concentration (�g/Nm 3 )<br />

Figure 8.6. Effect of elemental mercury inlet level on mercury oxidation <strong>by</strong> 10 mg<br />

Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min simulated flue gas<br />

with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21<br />

vol.% CO2, and 6 vol.% O2.<br />

8.4 Effect of carbon particle size<br />

The Darco Hg carbon has a mean diameter of only 16 µm. To be able to<br />

observe the influence of carbon particle size on mercury adsorption, the Norit RB4<br />

pellets are crushed and sieved to fractions having mean diameter of 38, 98, 165, and<br />

325 µm. Figure 8.7 presents the mercury breakthrough curves for crushed Norit RB4<br />

carbon with different particle sizes tested at 150�C using simulated cement kiln flue<br />

gas. Figure 8.8 further illustrates the effect of carbon particle size on the percentage<br />

of mercury oxidation and initial adsorption rate. Faster mercury breakthrough is<br />

183


observed for smaller carbon particle. The final mercury adsorption capacity is the<br />

same at a value of 0.725 µg Hg/mg_carbon, which is about 65% of the Darco Hg<br />

adsorption capacity at 150�C. Higher mercury oxidation and initial adsorption rate are<br />

also observed for smaller carbon particles.<br />

Gaseous Hg (C out /C in )<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1 2 3<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6<br />

Time (hour)<br />

184<br />

1: 38 �m<br />

2: 98 �m<br />

3: 325 �m<br />

Figure 8.7. Effect of particle size on mercury breakthrough for 10 mg crushed Norit<br />

pellets in 2 g sand tested at 150�C using 2.75 Nl/min simulated flue gas with 160-170<br />

µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.%<br />

H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

<strong>Mercury</strong> oxidation (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

Thiele modulus �<br />

0 2 4 6 8<br />

10<br />

Hg oxidation<br />

Adsorption rate for first 25 min<br />

0<br />

0<br />

0 50 100 150 200 250 300 350<br />

Carbon particle size (�m)<br />

2<br />

1.5<br />

1<br />

0.5<br />

<strong>Mercury</strong> adsorption rate (�g Hg/mg_carbon/h)


Figure 8.8. Effects of particle size on mercury oxidation and initial adsorption rate for<br />

10 mg crushed Norit RB4 pellets in 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10<br />

ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

The observation on effect of carbon particle size can be quantified using the<br />

well-known Thiele modulus, expressed for the first order kinetics as [11]:<br />

k<br />

S<br />

�<br />

� � R p<br />

ox a<br />

De<br />

p<br />

(8.3)<br />

where Rp is the carbon particle radius, kox is the oxidation rate constant for the first<br />

order reaction, Sa is the surface area per unit mass of carbon, �p is the carbon particle<br />

density, and De is the effective diffusivity. The Thiele modulus is defined as the ratio<br />

of an intrinsic reaction rate in the absence of mass transfer limitations to the rate of<br />

diffusion into the particle under specified conditions. When the carbon particle size<br />

increases, Thiele modulus becomes larger. For the smallest particle size of 38 µm the<br />

calculated Thiele modulus is about 1. For the larger particles the Thiele modulus are<br />

much larger than 1, indicating that the mercury oxidation <strong>by</strong> the large particles might<br />

be limited <strong>by</strong> the internal diffusion resistance. Similar trends are observed for the<br />

mercury oxidation percentage and initial adsorption rate, which again indicates that<br />

mercury oxidation is an important step in elemental adsorption <strong>by</strong> the activated<br />

carbon.<br />

8.5 Effect of flue gas flow rate<br />

The effect of flue gas flow rate on mercury breakthrough profile is presented<br />

in figure 8.9. With larger flue gas flow rate the active carbon is saturated and reach<br />

the equilibrium capacity in shorter time. The initial mercury breakthrough time,<br />

which is defined as time when the mercury concentration after the carbon bed starts<br />

to increase, decreases when the flow rate is increased. The final adsorption capacities<br />

for all tested flow rates are almost the same.<br />

A higher superficial velocity is associated with a higher total mercury input,<br />

resulting in a faster consumption of the sorption capacity of the activated carbon and<br />

185


corresponding higher mercury outlet concentration. While a higher superficial<br />

velocity enhances the mass transfer rate and the corresponding mercury sorption rate.<br />

Gaseous Hg C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.4 0.8 1.2 1.6 2<br />

Time (hour)<br />

186<br />

2750 Nml/min<br />

1830 Nml/min<br />

1100 Nml/min<br />

Figure 8.9. Effect of flue gas flow rate on mercury breakthrough of 10 mg Darco Hg<br />

mixed with 2 g sand tested at 150�C using 2.75 Nl/min simulated flue gas with 160-<br />

170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1<br />

vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

8.6 Effects of flue gas compositions<br />

Most of the previous work conducted the tests either in nitrogen or added<br />

single gas to the baseline gas of either only N2 or a mixture of CO2, O2, H2O and N2<br />

[5,11-19]. Some studies use simulated flue gas that does not contain all the relevant<br />

gases, especially the acid gases [20-26]. It is more reasonable to change ranges of the<br />

relevant gases instead of completely removing the gases <strong>from</strong> the baseline to simulate<br />

the real flue gas. In this work, the concentration of relevant gas component is varied<br />

while the concentrations of the other flue gas components remain at the baseline<br />

values.<br />

8.6.1 Effect of CO2<br />

The effects of CO2 concentration in the flue gas on mercury adsorption<br />

capacity of Darco Hg carbon at 150�C are illustrated in figure 8.10. The mercury


adsorption capacity slightly decreases when the CO2 level in the gas is increased <strong>from</strong><br />

1 to 31 vol.%. The negative effects of CO2 in the flue gas on mercury adsorption <strong>by</strong><br />

the virgin activated carbon were also observed <strong>by</strong> Yan et al. [13]. The decrease of the<br />

adsorption capacity in the presence of CO2 is probably due to the reduction in the<br />

active sites for mercury adsorption due to the competitive adsorption of CO2 and<br />

mercury on the carbon. The weak effect of CO2 on mercury adsorption might due to<br />

the fact that significant CO2 adsorption on the activated carbon only occur at low<br />

temperature (


simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10<br />

ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, and 6 vol.% O2.<br />

8.6.2 Effect of O2<br />

Figure 8.11 shows the effects of oxygen concentration in the flue gas on the<br />

mercury adsorption capacity of Darco Hg at 150�C. The mercury adsorption capacity<br />

hardly changes with increasing the oxygen level <strong>from</strong> 1 vol.% to 16 vol.% when<br />

taking the experimental uncertainty into account. In all the cases the mercury<br />

oxidation <strong>by</strong> the carbon is about 97%.<br />

<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 4 8 12 16 20<br />

O 2 level in gas (%)<br />

Figure 8.11. Effect of O2 concentration in the flue gas on mercury adsorption capacity<br />

of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min simulated<br />

flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl,<br />

1000 ppmv SO2, 1 vol.% H2O, and 21 vol.% CO2.<br />

The effects of oxygen level in the gas on mercury adsorption <strong>by</strong> activated<br />

carbon were investigated using a gas mixture of Hg, N2, and O2 at 140�C in the<br />

literature [24]. When the oxygen concentration was increased <strong>from</strong> 0 to 3 vol.%, the<br />

mercury adsorption capacity on the studied activated carbon remained almost<br />

unchanged. The mercury adsorption capacity increased <strong>by</strong> 16 and 33% when the<br />

oxygen level was further increased to 6 and 9 vol.%, respectively.<br />

188


The possibility of carbon-oxygen complexes formation during the fixed-bed<br />

test and their impact on mercury adsorption was investigated [24]. Pretreatment of the<br />

unoxidized carbon <strong>by</strong> air for 7 days had no impact on the mercury adsorption<br />

performance of the carbon. The air can oxidize carbon surface and increases its acidic<br />

surface functional group content. However, these changes have no impact on the<br />

performance of active carbon for mercury adsorption.<br />

Thermodynamic calculations of mercury-oxygen reactions suggest that about<br />

30% of the mercury could be present as HgO(g) at 200�C, while at lower<br />

temperatures HgO(s) is the dominant form, when acid gases such HCl and NOx are<br />

not present in the gas [33,34]. This could lead to very high mercury adsorption<br />

capacity as it is not adsorption of mercury but precipation of HgO(s) on the carbon<br />

that takes place. However, the exact temperature range at which HgO(s) is the<br />

dominant form was not reported. Homogenous gas phase reaction of mercury with<br />

oxygen in an atmosphere of N2, O2, and Hg was investigated <strong>by</strong> Hall et al [34].<br />

Results suggest that a homogeneous gas phase reaction between oxygen and<br />

elemental mercury is not an important factor in flue gas reaction processes. The<br />

enhanced mercury adsorption in the presence of oxygen can be explained <strong>by</strong> the<br />

conversion of mercury to mercuric oxides as there is no reaction between oxygen and<br />

mercury in the absence of activated carbon surface.<br />

When HCl is present in the gas even at ppmv level, compared to vol.% level<br />

of oxygen, elemental mercury will be mainly oxidized to HgCl2 instead of HgO as<br />

shown <strong>by</strong> the thermodynamic calculations presented in chapter 2 [35,36]. Oxygen<br />

was found to be a weak oxidant of mercury [37]. This could explain the weak effect<br />

of oxygen on the mercury adsorption <strong>by</strong> Darco Hg carbon tested in simulated flue gas<br />

in this work.<br />

8.6.3 Effect of H2O<br />

The effect of water in the flue gas on the mercury breakthrough profile and<br />

mercury adsorption capacity of Darco Hg at 150�C is shown in figure 8.12 and 8.13,<br />

respectively. The presence of water in the flue gas generally accelerates the mercury<br />

breakthrough and therefore decreases the amount of mercury adsorbed on the carbon.<br />

189


The mercury adsorption capacity is increased significantly when water is removed<br />

<strong>from</strong> the simulated flue gas. The mercury adsorption capacity of Darco Hg carbon<br />

tested without water in the flue gas is about 5.5 times of that with 1 vol.% water in<br />

the gas. However, this result is not practically important since full-scale flue gas<br />

always contains water in percentage level [38]. Compared to CO2, the effects of H2O<br />

in the flue gas on mercury adsorption are more pronounced. The mercury oxidation<br />

percentage is about 97% for water level in the range of 0-15 vol.%, however, the<br />

mercury oxidation decreases to 68% when 25 vol.% of water is added to the flue gas.<br />

As a result the mercury adsorption capacity of Darco Hg with 25 vol.% water is only<br />

about 44% of that with 1 vol.% water in the flue gas.<br />

Gaseous Hg, C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1 2 3<br />

0 1 2 3 4 5 6<br />

Time (hour)<br />

190<br />

1: 8% H 2O<br />

2:1% H 2 O<br />

3:0% H 2O<br />

Figure 8.12. Effects of water concentration in the flue gas on mercury breakthrough<br />

profile of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10<br />

ppmv HCl, 1000 ppmv SO2, 6 vol.% O2, and 21 vol.% CO2.


<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 5 10 15 20 25 30<br />

H 2O level in gas (%)<br />

191<br />

Data<br />

Y=1.1277X -0.261<br />

R 2 =0.98<br />

Figure 8.13. Effects of water concentration in the flue gas on mercury adsorption<br />

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10<br />

ppmv HCl, 1000 ppmv SO2, 6 vol.% O2, and 21 vol.% CO2. The regressed equation<br />

of mercury adsorption capacity is not valid for H2O level of zero.<br />

Tests in N2, Hg, and H2O at 140�C show that the mercury adsorption capacity<br />

of sulfur impregnated activated carbon does not change significantly when the water<br />

content in the gas is increased <strong>from</strong> 0 to 5 vol.% [24]. While the adsorption capacity<br />

decreases about 25% when the water content is further increased to 10 vol.%. This is<br />

in contrast to tests with simulated flue gas in this work, as the mercury adsorption<br />

capacity always decreases with water addition in the flue gas.<br />

The negative effects of water presence in the gas on mercury adsorption <strong>by</strong><br />

the carbon could be due to the competitive adsorption of water on the carbon. Water<br />

adsorption on carbon has been studied <strong>by</strong> several researchers [39-42]. Due to the<br />

strong chemisorption of water molecules with the acidic oxygen functional group on<br />

the carbon, the initial water adsorption occurs at the functional groups, and further<br />

water adsorption will occur on top of the chemisorbed water molecules via hydrogen<br />

bonding [42].


8.6.4 Effect of CO<br />

Unlike most other combustion processes, organic constituents in the raw<br />

material for clinker production result in CO emissions, even under optimized<br />

combustion conditions. In the preheater, these organic components in the raw<br />

material are liberated, part of them being emitted with the exhaust gas. The carbon<br />

monoxide concentration in the exhaust gas <strong>from</strong> cement rotary kiln systems ranges<br />

between 0.1 and 5 g/Nm³ (80-4000 ppmv) [43]. There is no reported investigation on<br />

possible effect of CO on mercury adsorption <strong>by</strong> the activated carbon.<br />

The effect of CO concentration in the flue gas on the mercury adsorption<br />

capacity of Darco Hg carbon at 150�C is illustrated in figure 8.14. Similar to the<br />

effects of oxygen, the mercury adsorption capacity is not affected <strong>by</strong> the presence of<br />

CO in the range of 0-1000 ppmv.<br />

<strong>Mercury</strong> adsorption capacity (�g Hg/mg_carbon)<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 200 400 600 800 1000 1200<br />

CO level in the gas (ppmv)<br />

Figure 8.14. Effects of CO concentration in the flue gas on mercury adsorption<br />

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10<br />

ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

As shown in figure 8.15, there is a slight decrease of mercury oxidation<br />

percentage when the CO level in the flue gas is increased. The mercury oxidation<br />

decreases <strong>from</strong> 98% when less than 100 ppmv CO is present in the flue gas to 85%<br />

192


with 1000 ppmv CO in the flue gas. This is probably because oxidized mercury is<br />

reduced at higher CO levels.<br />

<strong>Mercury</strong> oxidation (%)<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 200 400 600 800 1000<br />

CO level in flue gas (ppmv)<br />

Figure 8.15. Effects of CO concentration in the flue gas on mercury oxidation <strong>by</strong> 10<br />

mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min simulated flue<br />

gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000<br />

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

8.6.5 Effect of SO2<br />

The effect of SO2 in the flue gas on mercury breakthrough profile and<br />

adsorption capacity of Darco Hg at 150�C is presented in figure 8.16 and 8.17,<br />

respectively. There are strong effects of SO2 in the flue gas on the mercury adsorption<br />

<strong>by</strong> the activated carbon. The mercury adsorption capacity decreases when the SO2<br />

level in the flue gas is increased. The mercury adsorption capacity of Darco Hg<br />

carbon tested with 100 ppmv SO2 in the flue gas is about 4 times of that tested with<br />

1000 ppmv SO2 in the flue gas. Dunham et al. [44] also found that the mercury<br />

adsorption capacity of activated carbon is inversely affected <strong>by</strong> SO2 in the flue gas<br />

with reductions in adsorption capacity noted at concentrations as low as 100 ppmv<br />

SO2. The mercury oxidation is not affected <strong>by</strong> changing the SO2 level in the flue gas<br />

and is about 97% for SO2 added in the range of 100-1000 ppmv.<br />

193


Gaseous Hg, C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1 2<br />

0 0.5 1 1.5 2 2.5 3 3.5<br />

Time (hour)<br />

194<br />

3 4<br />

1: 1000 ppmv SO 2<br />

2: 500 ppmv SO 2<br />

3: 300 ppmv SO 2<br />

4: 100 ppmv SO 2<br />

Figure 8.16. Effects of SO2 concentration in the flue gas on mercury breakthrough<br />

profile of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10<br />

ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Data,Hg<br />

Y=69.176X -0.592<br />

R 2 =0.99<br />

0 200 400 600 800 1000 1200<br />

SO 2 level in gas (ppmv)<br />

Figure 8.17. Effects of SO2 concentration in the flue gas on mercury adsorption<br />

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10<br />

ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. The regressed equation of<br />

mercury adsorption capacity is valid for SO2 level higher than 100 ppmv.


Without the presence of NOx in the gas, the mercury adsorption capacity of<br />

activated carbon is also reported to decrease when the SO2 concentration increases<br />

[38]. This might be explained <strong>by</strong> the oxidation of SO2 to SO3 on the activated carbon<br />

and the inhibiting effect of SO3 on mercury capture <strong>by</strong> activated carbon injection has<br />

been observed in full-scale power plant tests [45,46]. In addition to removing<br />

mercury, activated carbon is also used as catalyst for oxidation SO2 to sulphuric acid<br />

and as SO2 sorbent [47,48]. There is competitive adsorption between Hg and SO3<br />

since both mercury and SO3 bind to the Lewis base sites on the activated carbon<br />

surface [45,49]. Some activated carbon catalysts for converting SO2 to H2SO4 are<br />

self-poisoned <strong>by</strong> SO3 or sulfate buildup on the surface. Therefore, a similar<br />

phenomenon might explain the inhibiting effect of SO3 on mercury capture.<br />

Previous results demonstrated that the oxidation of SO2 on carbon particles<br />

was greatly enhanced <strong>by</strong> the presence of trace quantities of gaseous NO2 [50-53].<br />

NO2 is an efficient oxidant for SO2 sorbed on carbon. According to the mechanisms<br />

of flue gas and mercury interactions on activated carbon proposed <strong>by</strong> Dunham et al.<br />

[44] and Olson et al. [54], sulfurous acid that accumulates <strong>from</strong> the hydration of SO2<br />

converts the previously formed nonvolatile basic mercuric nitrate into the volatile<br />

form. This results in the slow release of previously captured mercury over time in the<br />

presence of NO2 and SO2.<br />

8.6.6 Effect of HCl<br />

Figure 8.18 illustrates the effects of HCl in the flue gas on the mercury<br />

adsorption capacity of Darco Hg tested at 150�C. There are weak effects of HCl<br />

concentration in the flue gas on mercury adsorption capacity when 0.5-20 ppmv HCl<br />

is added to the flue gas. The mercury adsorption capacity increases gradually when<br />

the HCl level is increased <strong>from</strong> 0.5 to 5 ppmv and then it levels off when the HCl<br />

level is further increased. The mercury oxidation percentage is about 97% for all the<br />

tested HCl levels except that only 87% mercury oxidation is obtained when 0.5 ppmv<br />

HCl is added to the gas.<br />

195


<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 4 8 12 16 20 24<br />

HCl level in gas (ppmv)<br />

196<br />

Data<br />

Y=0.9728X 0.071<br />

R 2 =0.82<br />

Figure 8.18. Effects of HCl concentration in the flue gas on mercury adsorption<br />

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 1000<br />

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. The regressed equation of<br />

mercury adsorption capacity is valid for HCl level higher than 0.5 ppmv.<br />

Strong promoting effects of HCl on mercury adsorption <strong>by</strong> activated carbon<br />

in the simulated flue gas without NOx at 135�C have been reported <strong>by</strong> Carey et al.<br />

[38]. Addition of HCl to the simulated gases of 1600 ppmv SO2, 6% O2, 12% CO2<br />

and 7% H2O results in an increase of equilibrium adsorption capacity for elemental<br />

mercury <strong>from</strong> 0 at 0 ppmv HCl to a value approaching 3 µg Hg/mg_carbon in the<br />

range of 50–100 ppmv HCl. The adsorption capacity does not change significantly<br />

above 50 ppm HCl.<br />

When NOx is included the simulated flue gas, the promoting effects of HCl on<br />

adsorption capacity of activated carbon becomes less pronounced. <strong>Mercury</strong> can be<br />

adsorbed <strong>by</strong> the carbon without HCl presence in the gas, provided that NOx is present.<br />

To study whether mercury can be adsorbed <strong>by</strong> the activated carbon in the absence of<br />

HCl, one test was conducted <strong>by</strong> removing HCl <strong>from</strong> the baseline flue gas applied in<br />

this work. Figure 8.19 shows the mercury breakthrough curve for 10 mg Darco Hg in<br />

2 g sand at150�C.


Gaseous Hg, C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.4 0.8 1.2 1.6<br />

197<br />

Time (hour)<br />

10 ppmv HCl<br />

0 ppmv HCl<br />

Figure 8.19. Comparison of mercury breakthrough curves with 0 and 10 ppmv HCl in<br />

the simulated flue gas, 10 mg Darco Hg mixed with 2 g sand and tested at 150�C<br />

using 2.75 Nl/min simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23<br />

ppmv NO2, 1000 ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

Figure 8.19 compares the mercury breakthrough curves of 0 and 10 ppmv<br />

HCl in the simulated cement kiln flue gas. Test without HCl in the gas shows that<br />

large amount of mercury is adsorbed on the carbon and mercury breakthrough is not<br />

obtained after 10 h (not shown in figure 8.19). This might be due to the fact that<br />

HgCl2 will form when HCl is present in the gas and on the other hand HgO or HgSO4<br />

will form when HCl is not present through following reactions:<br />

2Hg O2<br />

2HgO<br />

� � (8R1)<br />

Hg NO � HgO � NO<br />

� 2 (8R2)<br />

Hg � O � SO � HgSO<br />

(8R3)<br />

2<br />

2<br />

4<br />

HgO(s) is easily captured <strong>by</strong> the carbon since it might condense on the carbon at the<br />

applied adsorption temperature of 150�C.<br />

8.6.7 Effect of NO<br />

As shown in figure 8.20, changing of NO concentration in the simulated flue<br />

gas does not affect the adsorption capacity of Darco Hg tested at 150�C. The mercury


oxidation extent <strong>by</strong> the Darco Hg carbon is about 98% for tested NO in the range of<br />

100-1000 ppmv.<br />

<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 200 400 600 800 1000 1200<br />

NO level in gas ( ppmv)<br />

Figure 8.20. Effects of NO concentration in the flue gas on mercury adsorption<br />

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 10 ppmv HCl, 23 ppmv NO2, 1000<br />

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

Strong promoting effects of NO on mercury adsorption <strong>by</strong> the activated<br />

carbon have been reported when 300 ppmv NO was added to baseline gas with 8<br />

vol.% H2O, 6 vol.% O2, and 12 vol.% CO2 at 107�C [5,11-19]. Liu et al. [24] reported<br />

that adding 500 ppmv NO to nitrogen did not change the mercury adsorption <strong>by</strong> the<br />

activated carbon at 140�C. Fan et al. [55] proposed that the promoting effects of NO<br />

on mercury adsorption <strong>by</strong> the activated carbon is due to the reaction of NO with O2 to<br />

form NO2 and active O atoms that could further react with elemental mercury. Thus<br />

the effects of NO on mercury adsorption depend on the presence of other acid gases<br />

in the flue gas. With HCl and NO2 presence in the gas the effects of NO are less<br />

significant.<br />

8.6.8 Effect of NO2<br />

The effect of NO2 in the flue gas on mercury breakthrough profile and<br />

adsorption capacity of Darco Hg at 150�C is presented in figure 8.21 and 8.22,<br />

198


espectively. The mercury oxidation <strong>by</strong> the Darco Hg carbon is about 98% for tested<br />

NO2 in the range of 0-100 ppmv. Dunham et al. [44] also found that the mercury<br />

adsorption capacity of activated carbon is inversely proportional to the concentrations<br />

of NO2 in the simulated flue gas. A decrease in the mercury adsorption capacity of<br />

activated carbon was observed at concentrations as low as 2.5 ppmv NO2. The<br />

negative effects of NO2 are again due to the interaction between NO2 and SO2, as<br />

discussed in section 8.6.5.<br />

Gaseous Hg, C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1 2 3<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Time (hour)<br />

199<br />

1: 100 ppmv NO 2<br />

2: 23 ppmv NO 2<br />

3: 5 ppmv NO 2<br />

Figure 8.21. Effects of NO2 concentration in the flue gas on mercury breakthrough<br />

curves of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 10 ppmv HCl, 1000 ppmv NO, 1000<br />

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.


<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 20 40 60 80 100 120<br />

NO 2 level in gas (ppmv)<br />

200<br />

Data<br />

Y=2.0685X -0.199<br />

R 2 =0.89<br />

Figure 8.22. Effects of NO2 concentration in the flue gas on mercury adsorption<br />

capacity of 10 mg Darco Hg mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 10 ppmv HCl, 1000 ppmv NO, 1000<br />

ppmv SO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. The regressed equation of<br />

mercury adsorption capacity is not valid when NO2 is not presented in the gas.<br />

Table 8.2 summarizes the regressed equations of Darco Hg carbon mercury<br />

adsorption capacity as a function of flue gas composition at 150�C. From these<br />

equations it is possible estimate mercury adsorption capacity in a wide range of glue<br />

gas composition.<br />

Table 8.2 Summary of regressed equations of Darco Hg carbon mercury adsorption<br />

capacity as a function of flue gas composition at 150�C.<br />

Gases Concentration unit Concentration range Equations<br />

CO2 % >0 Y=1.3779X -0.066<br />

H2O % >0 Y=1.1277X -0.261<br />

SO2 ppmv ≥100 Y=69.176X -0.592<br />

HCl ppmv ≥0.5 Y=0.9728X 0.071<br />

NO2 ppmv >0 Y=2.0685X -0.199


8.7 Conclusions<br />

A parametric study of elemental mercury adsorption <strong>by</strong> activated carbon has<br />

been conducted in the fixed-bed reactor <strong>by</strong> mixing 10 mg Darco Hg carbon with 2 g<br />

sand and using simulated cement kiln flue gas. Equilibrium mercury adsorption<br />

capacity, initial adsorption rate and mercury oxidation percentage are evaluated.<br />

Increasing adsorption temperature results in decreased equilibrium mercury<br />

adsorption capacity of the activated carbon. The mercury adsorption isotherm follows<br />

Henry’s law for the applied mercury inlet levels in this project at all tested<br />

temperatures. All these are consistent with a physical adsorption mechanism. The<br />

derived heat of adsorption is -8543 J/mol for elemental mercury adsorption <strong>by</strong> Darco<br />

Hg activated carbon in simulated cement kiln flue gas.<br />

The effects of carbon particle size were investigated using the crushed Norit<br />

RB4 pellets. Higher mercury oxidation and initial adsorption rate are observed for<br />

smaller carbon particles, while the equilibrium mercury adsorption capacity is the<br />

same.<br />

The effects of flue gas composition are investigated <strong>by</strong> varying the<br />

concentrations of relevant gases instead of complete removal of the single gas <strong>from</strong><br />

the baseline to simulate the real flue gas. The mercury adsorption capacity does not<br />

change with changes in the O2, CO, and NO levels in the flue gas. The mercury<br />

adsorption capacity decreases when CO2, H2O, SO2, and NO2 concentrations in the<br />

flue gas increase. The following correlation between mercury adsorption capacity and<br />

these gas concentrations are obtained at 150�C: mercury adsorption capacity is<br />

proportional to CCO2 -0.066 , CH2O -0.261 , CSO2 -0.592 , CNO2 -0.199 . The decrease of mercury<br />

adsorption capacity is due to the competition for active site with mercury <strong>by</strong> CO2 and<br />

H2O, and conversion of the previously formed nonvolatile basic mercuric nitrate into<br />

the volatile form <strong>by</strong> interactions between SO2 and NO2.<br />

Slight promoting effects of HCl on mercury adsorption are observed when<br />

HCl concentration is varied in the range of 0.5-20 ppmv. A larger mercury adsorption<br />

capacity is obtained when HCl is removed <strong>from</strong> baseline gas. This might due to the<br />

fact that HgCl2 will form when HCl is present in the gas while HgO(s) will form<br />

201


when HCl is not present. HgO is more easily captured <strong>by</strong> the carbon since it<br />

condenses on the carbon at the applied adsorption temperature of 150�C.<br />

Significant mercury oxidation is observed <strong>by</strong> the activated carbon. Even for<br />

10 mg carbon typically an oxidation level of 94-97% is found. Increasing CO and<br />

water level in the flue gas causes a slight decrease of mercury oxidation. Larger<br />

mercury oxidation percentage is obtained with smaller carbon particle size. All these<br />

observations indicate that mercury oxidation <strong>by</strong> HCl, when this is present in the gas,<br />

is an important step in elemental adsorption <strong>by</strong> the activated carbon.<br />

8.8 References<br />

[1] A.J. Cruz, J. Pires, A.P. Carvalho, D.C. Brotas, Adsorption of acetic acid <strong>by</strong> activated<br />

carbons, zeolites, and other adsorbent materials related with the preventive conservation of<br />

lead objects in museum showcases, Journal of Chemical & Engineering Data. 49 (2004) 725-<br />

731.<br />

[2] A. Bagreev, T.J. Bandosz, H2S adsorption/oxidation on unmodified activated carbons:<br />

importance of prehumidification, Carbon. 39 (2001) 2303-2311.<br />

[3] D.D. Do, Adsorption analysis: equilibria and kinetics, Imperial College Press, 1998.<br />

[4] R.C. Bansal, M. Goyal, Activated carbon adsorption, CRC Press, 2005.<br />

[5] D. Karatza, A. Lancia, D. Musmarra, F. Pepe, Adsorption of metallic mercury on<br />

activated carbon, Symp. Int. Combust. 26 (1996) 2439-2445.<br />

[6] H.G. Karge, J. Weitkamp, (Eds.), Molecular sieves-science and technology, volume 7,<br />

Adsorption and diffusion,Springer-Verlag Berlin Heidelberg, 2008.<br />

[7] T.C. Ho, N. Kobayashi, Y. Lee, J. Lin, J.R. Hopper, Experimental and kinetic study of<br />

mercury adsorption on various activated carbons in a fixed-bed adsorber, Environ. Eng. Sci.<br />

21 (2004) 21-27.<br />

[8] D.M. Ruthven, Principles of adsorption and adsorption processes, John Wiley & Sons,<br />

Inc., 1984.<br />

[9] D.O. Cooney, A.L. Hines, Extractive purification of activated carbon. 1. Effect of acid<br />

normality, volume, and repeated contact, Industrial & Engineering Chemistry Process Design<br />

and Development. 22 (1983) 208-211.<br />

[10] B. Padak, M. Brunetti, A. Lewis, J. Wilcox, <strong>Mercury</strong> binding on activated carbon,<br />

Environmental Progress. 25(2006) 319-326.<br />

[11] G. Skodras, I. Diamantopoulou, G. Pantoleontos, G.P. Sakellaropoulos, Kinetic studies<br />

of elemental mercury adsorption in activated carbon fixed bed reactor, Journal of Hazardous<br />

Materials. 158 (2008) 1-13.<br />

[12] S.J. Miller, G.E. Dunham, E.S. Olson, T.D. Brown, Flue gas effects on a carbon-based<br />

mercury sorbent, Fuel Processing Technology. 65-66 (2000) 343-363.<br />

202


[13] R. Yan, Y.L. Ng, D.T. Liang, C.S. Lim, J.H. Tay, Bench-scale experimental study on the<br />

effect of flue gas composition on mercury removal <strong>by</strong> activated carbon adsorption, Energy<br />

Fuels. 17 (2003) 1528-1535.<br />

[14] I. Diamantopoulou, G. Skodras, G.P. Sakellaropoulos, Sorption of mercury <strong>by</strong> activated<br />

carbon in the presence of flue gas components, Fuel Process Technol. 91 (2010) 158-163.<br />

[15] R. Yan, D.T. Liang, L. Tsen, Y.P. Wong, Y.K. Lee, Bench-scale experimental<br />

evaluation of carbon performance on mercury vapour adsorption, Fuel. 83 (2004) 2401-2409.<br />

[16] J.A. Korpiel, R.D. Vidic, Effect of sulfur impregnation method on activated carbon<br />

uptake of gas-phase mercury, Environ. Sci. Technol. 31 (1997) 2319-2325.<br />

[17] S. Lee, Y. Park, Gas-phase mercury removal <strong>by</strong> carbon-based sorbents, Fuel Process<br />

Technol. 84 (2003) 197-206.<br />

[18] S.J. Lee, Y. Seo, J. Jurng, T.G. Lee, <strong>Removal</strong> of gas-phase elemental mercury <strong>by</strong> iodineand<br />

chlorine-impregnated activated carbons, Atmospheric Environment. 38 (2004) 4887-<br />

4893.<br />

[19] D. Karatza, A. Lancia, D. Musmarra, C. Zucchini, Study of mercury absorption and<br />

desorption on sulfur impregnated carbon, Experimental Thermal and Fluid Science. 21 (2000)<br />

150-155.<br />

[20] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R.<br />

Chang, Preparation and evaluation of coal-derived activated carbons for removal of mercury<br />

vapor <strong>from</strong> simulated coal combustion flue gases, Energy Fuels. 12 (1998) 1061-1070.<br />

[21] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption <strong>by</strong> activated<br />

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy<br />

Conference, Research Triangle Park, NC, 22-25 April, 1997.<br />

[22] S. Eswaran, H.G. Stenger, Z. Fan, Gas-phase mercury adsorption rate studies, Energy<br />

Fuels. 21 (2007) 852-857.<br />

[23] R. Ochiai, M.A. Uddin, E. Sasaoka, S. Wu, Effects of HCl and SO2 concentration on<br />

mercury removal <strong>by</strong> activated carbon sorbents in coal-derived flue gas, Energy & Fuels. 23<br />

(2009) 4734.<br />

[24] W. Liu, R.D. Vidic, T.D. Brown, Impact of flue gas conditions on mercury uptake <strong>by</strong><br />

sulfur-impregnated activated carbon, Environ. Sci. Technol. 34 (2000) 154-159.<br />

[25] G.E. Dunham, S.J. Miller, <strong>Mercury</strong> capture <strong>by</strong> an activated carbon in a fixed-bed benchscale<br />

system, Environmental Progress. 17 (1998) 203.<br />

[26] S. Lee, J. Lee, T.C. Keener, Novel sorbents for mercury emissions control <strong>from</strong> coalfired<br />

power plants, Journal of the Chinese Institute of Chemical Engineers. 39 (2008) 137-<br />

142.<br />

[27] M. Molina-Sabio, A.M.A. Muñecas, F. Rodríguez-Reinoso, B. McEnaney, Adsorption<br />

of CO2 and SO2 on activated carbons with a wide range of micropore size distribution,<br />

Carbon. 33 (1995) 1777-1782.<br />

203


[28] T.C. Drage, O. Kozynchenko, C. Pevida, M.G. Plaza, F. Rubiera, J.J. Pis, C.E. Snape, S.<br />

Tennison, Developing activated carbon adsorbents for pre-combustion CO2 capture, Energy<br />

Procedia. 1 (2009) 599-605.<br />

[29] M.G. Plaza, S. García, F. Rubiera, J.J. Pis, C. Pevida, Post-combustion CO2 capture with<br />

a commercial activated carbon: Comparison of different regeneration strategies, Chem. Eng.<br />

J. 163 (2010) 41-47.<br />

[30] C. Shen, C.A. Grande, P. Li, J. Yu, A.E. Rodrigues, Adsorption equilibria and kinetics<br />

of CO2 and N2 on activated carbon beads, Chem. Eng. J. 160 (2010) 398-407.<br />

[31] Y. Sun, Y. Wang, Y. Zhang, Y. Zhou, L. Zhou, CO2 sorption in activated carbon in the<br />

presence of water, Chemical Physics Letters. 437 (2007) 14-16.<br />

[32] B. Guo, L. Chang, K. Xie, Adsorption of carbon dioxide on activated carbon, Journal of<br />

Natural Gas Chemistry. 15 (2006) 223-229.<br />

[33] B. Hall, O. Lindqvist, E. Ljungstroem, <strong>Mercury</strong> chemistry in simulated flue gases related<br />

to waste incineration conditions, Environ. Sci. Technol. 24 (1990) 108-111.<br />

[34] B. Hall, P. Schager, J. Weesmaa, The homogeneous gas phase reaction of mercury with<br />

oxygen, and the corresponding heterogeneous reactions in the presence of activated carbon<br />

and fly ash, Chemosphere. 30 (1995) 611-627.<br />

[35] C.L. Senior, A.F. Sarofim, T. Zeng, J.J. Helble, R. Mamani-Paco, Gas-phase<br />

transformations of mercury in coal-fired power plants, Fuel Process Technol. 63 (2000) 197-<br />

213.<br />

[36] R.N. Sliger, J.C. Kramlich, N.M. Marinov, Towards the development of a chemical<br />

kinetic model for the homogeneous oxidation of mercury <strong>by</strong> chlorine species, Fuel Process<br />

Technol. 65-66 (2000) 423-438.<br />

[37] Niksa, S. and J. J. Helble, Interpreting laboratory test data on homogeneous mercury<br />

oxidation in coal-derived exhausts. EPA-DOE-EPRI Combined Power Plant Air Pollution<br />

Control Symposium: The Mega Symposium and the A&WMA Specialty Conference on<br />

<strong>Mercury</strong> Emissions: Fate, Effects, and Control, Chicago, Illinois, August 2001.<br />

[38] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors<br />

affecting mercury control in utility flue gas using activated carbon, Journal of the Air &<br />

Waste Management Association. 48 (1998) 1166.<br />

[39] M.M. Dubinin, V.V. Serpinsky, Isotherm equation for water vapor adsorption <strong>by</strong><br />

microporous carbonaceous adsorbents, Carbon. 19 (1981) 402-403.<br />

[40] M.M. Dubinin, Water vapor adsorption and the microporous structures of carbonaceous<br />

adsorbents, Carbon. 18 (1980) 355-364.<br />

[41] F. Stoeckli, T. Jakubov, A. Lavanchy, Water adsorption in active carbons described <strong>by</strong><br />

the Dubinin-Astakhov equation, Journal of the Chemical Society, Faraday Transactions. 90<br />

(1994) 783-786.<br />

[42] D.D. Do, H.D. Do, A model for water adsorption in activated carbon, Carbon. 38 (2000)<br />

767-773.<br />

204


[43] ECRA (European cement research academy), Carbon monoxide formation and burn-out<br />

during the clinker burning process, News letter 2, 2008.<br />

[44] G.E. Dunham, E.S. Olson and S.J. Miller, Impact of flue gas constituents on carbon<br />

sorbents, McLean, VA, September 19-21, 2000.<br />

[45] A.A. Presto, E.J. Granite, Impact of sulfur oxides on mercury capture <strong>by</strong> activated<br />

carbon, Environ. Sci. Technol. 41 (2007) 6579-6584.<br />

[46] J. Jarvis, F. Meserole, SO3 Effect on <strong>Mercury</strong> Control, Power Eng. 112 (2008) 54-60.<br />

[47] E. Raymundo-Piñero, D. Cazorla-Amorós, C. Salinas-Martinez de Lecea, A. Linares-<br />

Solano, Factors controlling the SO2 removal <strong>by</strong> porous carbons: relevance of the SO2<br />

oxidation step, Carbon. 38 (2000) 335-344.<br />

[48] E. Raymundo-Piñero, D. Cazorla-Amorós, A. Linares-Solano, Temperature programmed<br />

desorption study on the mechanism of SO2 oxidation <strong>by</strong> activated carbon and activated<br />

carbon fibres, Carbon. 39 (2001) 231-242.<br />

[49] A.A. Presto, E.J. Granite, A. Karash, Further investigation of the impact of sulfur oxides<br />

on mercury capture <strong>by</strong> activated carbon, Ind Eng Chem Res. 46 (2007) 8273-8276.<br />

[50] W.R. Cofer III, D.R. Schryer, R.S. Rogowski, The enhanced oxidation of SO2 <strong>by</strong> NO2<br />

on carbon particulates, Atmospheric Environment (1967). 14 (1980) 571-575.<br />

[51] W.R. Cofer III, D.R. Schryer, R.S. Rogowski, The oxidation of SO2 on carbon particles<br />

in the presence of O3, NO2 and N2O, Atmospheric Environment (1967). 15 (1981) 1281-1286.<br />

[52] W.R. Cofer III, D.R. Schryer, R.S. Rogowski, Oxidation of SO2 <strong>by</strong> NO2 and O3 on<br />

carbon: Implications to tropospheric chemistry, Atmospheric Environment (1967). 18 (1984)<br />

243-245.<br />

[53] J.A. Rodriguez, T. Jirsak, J. Dvorak, S. Sambasivan, D. Fischer, Reaction of NO2 with<br />

Zn and ZnO: Photoemission, XANES, and density functional studies on the formation of NO3,<br />

The Journal of Physical Chemistry B. 104 (2000) 319-328.<br />

[54] E.S. Olson, B.A. Mibeck, S.A. Benson, J.D. Laumb, C.R. Crocker, G.E. Dunham, et al.,<br />

The mechanistic model for flue gas-mercury interactions on activated carbons: The oxidation<br />

site, Prepr. Pap. -Am. Chem. Soc. , Div. Fuel Chem. 49 (2004) 279-280.<br />

[55] X. Fan, C. Li, Zeng Guangming, Z. Gao, L. Chen, W. Zhang, et al., <strong>Removal</strong> of gasphase<br />

element mercury <strong>by</strong> activated carbon fiber impregnated with CeO2, Energy & Fuels. 24<br />

(2010) 4250-4254.<br />

205


Fundamental investigation of mercury<br />

chloride adsorption <strong>by</strong> activated carbon in<br />

simulated cement kiln flue gas<br />

This chapter deals with a fundamental investigation of mercury chloride<br />

adsorption <strong>by</strong> the Darco Hg activated carbon in simulated cement kiln flue gas. The<br />

results are compared with tests using elemental mercury.<br />

9.1 Introduction<br />

Compared to research on elemental mercury adsorption [1-13], there are few<br />

studies on mercury chloride adsorption <strong>by</strong> activated carbon [1,2,14-16]. Some of<br />

these studies conducted tests using simulated flue gases containing 1600 ppmv SO2,<br />

50 ppmv HCl, 12 vol.% CO2, 7 vol.% H2O, 6 vol.% O2, but without NOx [1,2,14].<br />

Carey et al. [15] performed tests using simulated flue gas with 1600 ppmv SO2, 1-50<br />

ppmv HCl, 10-12 vol.% CO2, 8 vol.% H2O, 6 vol.% O2, and 200-400 ppmv NOx. The<br />

research [15] focused on a comparison of mercury adsorption capacity obtained in a<br />

fixed-bed reactor using simulated flue gas and real power plant flue gas. Mibeck et al.<br />

[16] investigated the effects of acid gases <strong>by</strong> adding 1600 ppmv SO2, 50 ppmv HCl,<br />

400 ppmv NO, and 20 ppmv NO2 alone or in combination to baseline gas of 12 vol.%<br />

CO2, 8 vol.% H2O, 6 vol.% O2. Only breakthrough curves are presented, neither<br />

adsorption capacity nor kinetics is reported in their work.<br />

In this work, mercury chloride adsorption capacity and kinetics are<br />

investigated <strong>by</strong> varying the relevant gas concentrations and operating parameters.<br />

206<br />

9


9.2 Effect of temperature<br />

The effect of temperature on mercury chloride adsorption is investigated <strong>by</strong><br />

applying the same conditions as for the study with elemental mercury source reported<br />

in chapter 8. The breakthrough curves using HgCl2 source are compared with those<br />

obtained with elemental mercury at 100, 120, and 150�C, as illustrated in figure 9.1-3.<br />

In contrast to the breakthrough curve obtained with elemental mercury, the<br />

breakthrough curve of mercury chloride often has an introduction period. It takes<br />

some time to reach the lowest outlet mercury concentration after switching the flue<br />

gas with mercury chloride to the carbon bed. While the outlet mercury decreases to<br />

the lowest value almost simultaneously after switching the flue gas with elemental<br />

mercury to the carbon bed. This phenomenon is also observed <strong>by</strong> Mibeck et al. [16].<br />

They also reported that about 90-95% of mercury after the carbon bed is oxidized<br />

mercury. Measurement in this work shows that all the mercury after the carbon bed is<br />

oxidized mercury, i.e., no reduction takes place.<br />

Gaseous Hg, C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Time (hour)<br />

207<br />

Hg 0 , 170�g/Nm 3 Hg,<br />

1.183 �g Hg/mg_carbon<br />

HgCl 2, 183�g/Nm 3 Hg,<br />

1.224 �g Hg/mg_carbon<br />

Figure 9.1. Comparison of breakthrough curves obtained using mercury chloride and<br />

elemental mercury source for 10 mg Darco Hg carbon mixed with 2 g sand tested at<br />

150�C using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2,<br />

1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.


Gaseous Hg, C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

Time (hour)<br />

208<br />

Hg 0 , 166�g/Nm 3 Hg,<br />

1.335 �g Hg/mg_carbon<br />

HgCl 2, 156�g/Nm 3 Hg,<br />

1.375 �g Hg/mg_carbon<br />

Figure 9.2. Comparison of breakthrough curves obtained using mercury chloride and<br />

elemental mercury source for 10 mg Darco Hg carbon mixed with 2 g sand tested at<br />

120�C using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2,<br />

1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

Gaseous Hg, C out /C in<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6<br />

Time (hour)<br />

Hg 0 , 167�g/Nm 3 Hg,<br />

1.506 �g Hg/mg_carbon<br />

HgCl 2, 159�g/Nm 3 Hg,<br />

1.429 �g Hg/mg_carbon<br />

Figure 9.3. Comparison of breakthrough curves obtained using mercury chloride and<br />

elemental mercury source for 10 mg Darco Hg carbon mixed with 2 g sand tested at<br />

100�C using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2,<br />

1000 ppmv SO2, 10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2.<br />

The equilibrium mercury adsorption capacities are very similar for tests using<br />

elemental mercury and mercury chloride under same gas conditions at different<br />

temperatures. This is probably due to the catalytic oxidation of elemental mercury <strong>by</strong>


the activated carbon. As shown in chapter 8, almost all elemental mercury is oxidized<br />

<strong>by</strong> the activated carbon. Thus it is not surprising to observe similar adsorption<br />

behavior of elemental mercury and mercury chloride adsorption <strong>by</strong> the activated<br />

carbon for the present flue gas containing HCl. For sorbents with poor mercury<br />

oxidation ability, the adsorption behavior of elemental mercury and oxidized mercury<br />

is expected to be different. Increasing adsorption temperature also decreases the<br />

adsorption capacity of mercury chloride.<br />

Similarly to tests with elemental mercury, the adsorption constants of HgCl2<br />

are also derived. Figure 9.4 shows that there is a linear relation between lnk and 1/T<br />

and <strong>from</strong> the slope and intercept the calculated value for k0 and �H ads is 1.595 m 3 /g<br />

and -6587 J/mol, respectively. The corresponding k0 and �H ads for tests with<br />

elemental mercury is 0.869 m 3 /g and -8543 J/mol, respectively.<br />

ln(k)<br />

-4<br />

-4.2<br />

-4.4<br />

-4.6<br />

Hg 0 data<br />

Y=1027.4881X-7.0484<br />

R 2 =0.93<br />

HgCl2 data<br />

Y=792.2246X-6.4409<br />

R 2 =0.93<br />

-4.8<br />

0.0022 0.0024 0.0026 0.0028 0.003<br />

209<br />

1/T (1/K)<br />

Figure 9.4. Plot of lnk as a function of 1/T. The Henry’s constants k are derived <strong>from</strong><br />

isotherms at 100, 120, 150, and 180�C. 10 mg Darco Hg mixed with 2 g sand is tested<br />

using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv<br />

HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. Data for tests<br />

using elemental mercury are shown for comparison.


9.3 Effect of flue gas composition<br />

Since negligible effects of CO2, O2, CO, and NO on elemental mercury<br />

adsorption <strong>by</strong> the activated carbon are observed and preliminary tests of mercury<br />

chloride adsorption <strong>by</strong> the activated carbon shows similar behavior as elemental<br />

mercury, only effects of H2O, SO2, HCl, and NO2 on mercury chloride adsorption <strong>by</strong><br />

the Darco Hg activated carbon are investigated. The effects of H2O, SO2, HCl, and<br />

NO2 on equilibrium adsorption capacity of mercury chloride are presented in figure<br />

9.5-8, respectively. Data for tests using elemental mercury are shown for comparison.<br />

<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 5 10 15 20 25 30<br />

H 2O level in gas (%)<br />

210<br />

Data,Hg 0<br />

Y=1.1277X -0.261<br />

R 2 =0.98<br />

Data, HgCl2<br />

Y=1.2389X -0.240<br />

R 2 =0.98<br />

Figure 9.5. Effects of water in the flue gas on mercury chloride adsorption capacity of<br />

10 mg Darco Hg carbon mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 10 ppmv<br />

HCl, 6 vol.% O2, and 21 vol.% CO2.


<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 200 400 600 800 1000 1200<br />

SO 2 level in gas (ppmv)<br />

211<br />

Data,Hg 0<br />

Y=69.176X -0.592<br />

R 2 =0.99<br />

Data,HgCl2<br />

Y=58.23X -0.565<br />

R 2 =0.99<br />

Figure 9.6. Effects of SO2 in the flue gas on mercury chloride adsorption capacity of<br />

10 mg Darco Hg carbon mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 6 vol.% O2, 1<br />

vol.% H2O, and 21 vol.% CO2.<br />

<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 4 8 12 16 20 24<br />

HCl level in gas (ppmv)<br />

Data,Hg 0<br />

Y=0.9728X 0.071<br />

R 2 =0.82<br />

Data,HgCl2<br />

Y=0.9162X 0.1158<br />

R 2 =0.76<br />

Figure 9.7. Effects of HCl in the flue gas on mercury chloride adsorption capacity of<br />

10 mg Darco Hg carbon mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2, 6 vol.% O2,<br />

1 vol.% H2O, and 21 vol.% CO2.


<strong>Mercury</strong> adsorption capacity (�g_Hg/mg_carbon)<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

Data, Hg 0<br />

Y=2.0685X -0.199<br />

R 2 =0.89<br />

Data, HgCl2<br />

Y=2.0732X -0.197<br />

R 2 =0.88<br />

0 20 40 60 80 100 120<br />

NO 2 level in gas (ppmv)<br />

Figure 9.8. Effects of NO2 in the flue gas on mercury chloride adsorption capacity of<br />

10 mg Darco Hg carbon mixed with 2 g sand tested at 150�C using 2.75 Nl/min<br />

simulated flue gas with 1000 ppmv NO, 1000 ppmv SO2, 10 ppmv HCl, 6 vol.% O2,<br />

1 vol.% H2O, and 21 vol.% CO2.<br />

The adsorption capacity of mercury chloride is slightly larger than the<br />

elemental mercury when the water content in the flue gas is above 1 vol.%. Almost<br />

the same tendency of adsorption capacity as a function of SO2, HCl, and NO2<br />

concentration in the flue gas is observed for mercury chloride and elemental mercury.<br />

This is again due to the high oxidation rate of elemental mercury <strong>by</strong> the Darco Hg<br />

carbon.<br />

9.4 Conclusions<br />

Similar adsorption behaviors of mercury chloride and elemental mercury <strong>by</strong><br />

Darco Hg activated carbon are observed using simulated cement kiln flue gas at<br />

different temperatures. Increasing adsorption temperature also decreases the<br />

adsorption capacity of mercury chloride.<br />

The effects of H2O, SO2, HCl, and NO2 on mercury chloride adsorption <strong>by</strong><br />

Darco Hg are investigated <strong>by</strong> varying their concentrations in the baseline gas.<br />

Compared to elemental mercury adsorption, a slightly larger adsorption capacity of<br />

mercury chloride is obtained when the water content in the flue gas is above 1 vol.%.<br />

212


The dependence of mercury chloride adsorption capacity on SO2, HCl, and NO2<br />

concentrations in the flue gas is the same as elemental mercury adsorption capacity.<br />

The similar behavior of mercury chloride and elemental mercury is due to the<br />

effective catalytic oxidation of elemental mercury <strong>by</strong> the activated carbon in the<br />

presence of HCl.<br />

9.5 References<br />

[1] H.C. Hsi, S. Chen, M. Rostam-Abadi, M.J. Rood, C.F. Richardson, T.R. Carey, R. Chang,<br />

Preparation and evaluation of coal-derived activated carbons for removal of mercury vapor<br />

<strong>from</strong> simulated coal combustion flue gases, Energy & Fuels. 12 (1998) 1061-1070.<br />

[2] M. Rostam-Abadi, S.G. Chen, H.C. Hsi, M. Rood, R. Chang, T.R. Carey, B., Hargrove, C.<br />

Richardson, W. Rosenhoover, F. Meserole, Novel vapor phase mercury sorbents,<br />

Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant Control, Washington,<br />

DC, Aug 25–29, 1997.<br />

[3] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang and F.B. Meserole, Factors<br />

affecting mercury control in utility flue gas using sorbent injection, Proceedings of the Air &<br />

Waste Management Association's 90 th Annual Meeting, Toronto, Ontario, Canada, June 8-13,<br />

1997.<br />

[4] C.X. Hu, J.S. Zhou, Z.Y. Luo, S. He, G.K. Wang, K.F. Cen, Effect of oxidation treatment<br />

on the adsorption and the stability of mercury on activated carbon, Journal of Environmental<br />

Sciences-China. 18 (2006) 1161-1166.<br />

[5] B. Ghorishi and B.K. Gullett, Fixed-bed control of mercury: Role of acid gases and a<br />

comparison between carbon-based, calcium-based, and coal fly ash sorbents, Proceedings of<br />

the EPRI/DOE/EPA Combined Utility Air Pollutant Control Symposium, Washington, DC,<br />

August 25–29, 1997.<br />

[6] S.B. Ghorishi and C.B. Sedman, Combined mercury and sulfur oxides control using<br />

calcium-based sorbents, Proceedings of the EPRI-DOE-EPA Combined Utility Air Pollutant<br />

Control Symposium, Washington, DC, August 25–29, 1997.<br />

[7] S.B. Ghorishi and B.K. Gullett, An experimental study on mercury sorption <strong>by</strong> activated<br />

carbons and calcium hydroxide, The Fifth Annual North American Waste-to-Energy<br />

Conference, Research Triangle Park, NC, 22-25 April, 1997.<br />

[8] N. Hutson, C. Singer, C. Richardson, J. Karwowski and C. Sedman, Practical applications<br />

<strong>from</strong> observations of mercury oxidation and binding mechanisms, EPRI-DOE-EPA-AWMA<br />

Combined Power Plant Air Pollutant Control Mega Symposium, Washington DC, August 30<br />

- September 2, 2004.<br />

[9] E.J. Granite, M.C. Freeman, R.A. Hargis, W.J. O'Dowd, H.W. Pennline, The thief process<br />

for mercury removal <strong>from</strong> flue gas, J. Environ. Manage. 84 (2007) 628-634.<br />

213


[10] W.J. O’Dowd, H.W. Pennline, M.C. Freeman, E.J. Granite, R.A. Hargis, C.J. Lacher, A.<br />

Karash, A technique to control mercury <strong>from</strong> flue gas: The thief process, Fuel Processing<br />

Technology. 87 (2006) 1071-1084.<br />

[11] R. Bhardwaj. Impact of temperature and flue gas components on mercury speciation and<br />

uptake <strong>by</strong> activated carbon sorbents . Master thesis. Master, University of Pittsburgh, 2007.<br />

[12] M.M. Maroto-Valer, Y. Zhang, E.J. Granite, Z. Tang, H.W. Pennline, Effect of porous<br />

structure and surface functionality on the mercury capacity of a fly ash carbon and its<br />

activated sample, Fuel. 84 (2005) 105-108.<br />

[13] S. Eswaran, H.G. Stenger, Z. Fan, Gas-phase mercury adsorption rate studies, Energy &<br />

Fuels. 21 (2007) 852-857.<br />

[14] T.R. Carey, O.W. Hargrove Jr, C.F. Richardson, R. Chang, F.B. Meserole, Factors<br />

affecting mercury control in utility flue gas using activated carbon, Journal of the Air &<br />

Waste Management Association. 48 (1998) 1166.<br />

[15] T.R. Carey, C.F. Richardson, R. Chang, F.B. Meserole, M. Rostam-Abadi, S. Chen,<br />

Assessing sorbent injection mercury control effectiveness in flue gas streams, Environ. Prog.<br />

19 (2000) 167-174.<br />

[16] B.A.F. Mibeck, E.S. Olson, S.J. Miller, HgCl2 sorption on lignite activated carbon:<br />

Analysis of fixed-bed results, Fuel Process Technol. 90 (2009) 1364-1371.<br />

214


215<br />

10<br />

Simulation of mercury adsorption <strong>by</strong> fixed<br />

carbon bed<br />

To properly understand an adsorption process, two basic ingredients, i.e.,<br />

equilibrium and transport processes must be investigated. Understanding of the<br />

adsorptive capacity is within the domain of equilibrium, and understanding of the<br />

diffusion resistance is within the domain of transport process. These two aspects are<br />

first introduced in this chapter. The remaining part of this chapter deals with<br />

mathematical models describing the behavior of isothermal adsorption of mercury in<br />

a carbon particle and a fixed carbon bed including adsorption isotherm, mass balance<br />

for the gas phase and mass balance inside the adsorbent.<br />

10.1 Adsorption equilibrium<br />

A summary of commonly used isotherm equations for pure gas adsorption is<br />

given in the following table 10.1. Considering the typical mercury level in the flue<br />

gas of ppb level, the simple Henry’s law is able to describe the isotherm well, as<br />

shown in chapter 8.


Table 10.1. Summary of commonly used isotherm equations for pure gas adsorption<br />

[1].<br />

Isotherm Equation Remarks<br />

Henry law C�� KC<br />

Low pressure range<br />

Langmuir<br />

bC<br />

C� � C�s<br />

1�<br />

bC<br />

Has Henry law limit and<br />

finite saturation limit<br />

Frendlich<br />

1/n<br />

C�KFC � Does not have Henry law<br />

limit and no saturation limit<br />

Langmuir-Frendlich<br />

1/ n<br />

( bC)<br />

C� � C�s<br />

1/ n<br />

1 � ( bC)<br />

Does not have Henry law<br />

limit, but has finite<br />

saturation limit<br />

Toth<br />

bC<br />

C� � C�s<br />

t 1/ t<br />

[1 � ( bC)<br />

]<br />

Has Henry law limit and<br />

finite saturation limit<br />

Unilan s<br />

C�s 1�<br />

be C<br />

C��ln( ) �s<br />

2s 1�be<br />

C<br />

Has Henry law limit and<br />

finite saturation limit<br />

K: Henry constant; C: gaseous adsorbate concentration; C�: adsorbed concentration in<br />

the sorbent; C�s: saturation adsorbed concentration in the sorbent; b: Langmuir<br />

constant; KF: Frendlich constant; s: heterogeneity parameter<br />

10.2 Transport consideration in adsorption process<br />

Adsorption of an adsorbate molecule on to the porous surface of an adsorbent<br />

include following steps [2]:<br />

1. External (or interphase) mass transfer of the adsorbate <strong>from</strong> the bulk fluid <strong>by</strong><br />

convection through a thin film or boundary layer.<br />

2. Internal (intraphase) mass transfer of the adsorbate <strong>by</strong> pore diffusion <strong>from</strong> the<br />

outer surface of the adsorbent to the inner surface of the internal porous structure.<br />

3. Surface diffusion along the porous surface.<br />

4. Adsorption of the adsorbate onto the porous surface.<br />

10.2.1 External transport<br />

Rates of convection mass and heat transfer between the outer surface of a<br />

particle and the surrounding bulk fluid during an adsorption process are given,<br />

respectively, <strong>by</strong> [2]:<br />

216


dN<br />

�kmA�cb � cs�<br />

(10.1)<br />

dt<br />

dQ<br />

�hA�Ts� Tb<br />

�<br />

(10.2)<br />

dt<br />

where km is the external mass transfer coefficient, A is the particle external surface<br />

area, Cb and Cs is the gas concentration in the bulk and at the particle surface,<br />

respectively. h is the heat transfer coefficient, Tb and Ts is the gas temperature in the<br />

bulk and at the particle surface, respectively.<br />

When fluid flows past a single particle, experimental transport data<br />

correlations are usually developed for coefficients averaged over the particle surface.<br />

Some typical correlations published <strong>by</strong> Ranz and Marshall for Nusselt numbers as<br />

high as 30, and Sherwood numbers to 160 are the following [2]:<br />

N �2� 0.60N<br />

N<br />

(10.3)<br />

Nu<br />

1 1<br />

2 3<br />

Re Pr<br />

N �2� 0.60N<br />

N<br />

(10.4)<br />

1 1<br />

2 3<br />

Sh Re Sc<br />

C p�<br />

where Prandtl number NPr= ; Schmidt number NSc=<br />

k<br />

217<br />

�<br />

; Reynolds number<br />

�Di<br />

d pG<br />

NRe= , and G is the fluid mass velocity.<br />

�<br />

When particles are packed in a bed, the fluid flow patterns are restricted, and<br />

the single particle correlations cannot be used to estimate the average external<br />

transport coefficients for particles in the bed. A correlation of 37 sets of mass-transfer<br />

data including Sherwood number corrections for axial dispersion result in an<br />

expression of the form [2]:<br />

k d<br />

N �<br />

m p<br />

0.<br />

5 1<br />

3<br />

Sh � � 2 1.<br />

1N<br />

Re N Sc<br />

(10.5)<br />

Di<br />

This equation covers a Schmidt number range <strong>from</strong> 0.6 to 70600, a Reynolds number<br />

range <strong>from</strong> 3 to 10000. Particle shapes applicable include spheres, short cylinders,<br />

flakes and granules. By analogy, the corresponding equation for fluid-particle<br />

convection heat transfer in packed beds is:<br />

N<br />

Nu<br />

hd p<br />

0.<br />

5 1<br />

3<br />

� � 2 �1.<br />

1N<br />

Re N Pr<br />

(10.6)<br />

k


When these equations are used with beds packed with non-spherical particles, dp, is<br />

the equivalent diameter of a spherical particle.<br />

10.2.2 Internal transport<br />

Porous particles in most cases have a sufficiently high effective thermal<br />

conductivity so that temperature gradients within the particle are negligible. In<br />

contrast, internal mass transfer within the particle must be considered. In sorption<br />

processes, transport is <strong>from</strong> the exterior to the interior for adsorption and <strong>from</strong> the<br />

interior to the exterior for desorption processes. The flux of mercury transported to<br />

the carbon particle can be expressed as:<br />

dC<br />

N A � �De<br />

(10.7)<br />

dr<br />

where De is the effective diffusion coefficient. There are basically three modes of<br />

transport of molecules inside a porous medium: Knudsen diffusion, molecular<br />

diffusion, and surface diffusion [1].<br />

10.2.2.1 Molecular Diffusion<br />

When the adsorbate is in a macropore or in the fluid phase, the frequency of<br />

collision with a surface is minimal and transport of the molecule occurs via<br />

intermolecular collisions only. This mode of transport is due to a partial pressure<br />

gradient of a continuum fluid mixture.<br />

For binary gas mixtures at low pressure (


parameters of the individual species in the system. MWA and MWB are molecular<br />

weights of species A and B. The molecular collision diameter, �AB, is calculated as<br />

the arithmetic average of the two species:<br />

� � ( � � � )<br />

(10.9)<br />

1<br />

AB 2 A B<br />

�D,AB is a dimensionless function of temperature and the intermolecular potential<br />

field for a molecular of A and B. The interaction is described <strong>by</strong> the individual<br />

Lennard-Jones 12-6-potentials, �A and �B, in accordance with following equation:<br />

�AB k �<br />

�A�B k. k<br />

(10.10)<br />

�D,AB can be calculated according to:<br />

1.06036 0.19300 1.03587 1.76474<br />

�DAB , � � � �<br />

0.15610<br />

( kT ) exp(0.47635 kT ) exp(1.52996 kT ) exp(3.89411 kT )<br />

� AB �AB �AB �AB<br />

(10.11)<br />

where k is the Boltzmann’s constant.<br />

The Lennard-Jones potential parameter for N2 can be easily found while for<br />

elemental mercury only few data are available. Table 10.2 lists the Lennard-Jones<br />

potential parameter for N2 and elemental mercury.<br />

Table 10.2. Lennard-Jones potential parameter for N2 and elemental mercury [3].<br />

� (Å) �/k (K)<br />

N2 3.681 91.5<br />

Hg 3.23 627<br />

10.2.2.2 Knudsen Diffusion<br />

This diffusion process occurs when the mean free path of the adsorbate is<br />

much larger than the diameter of the channel in which the diffusing molecules reside.<br />

This normally occurs at very low pressure and channels of small size, usually of order<br />

of 10 nm to 100 nm [1]. The flow is induced <strong>by</strong> collision of gaseous molecules with<br />

the pore wall.<br />

2r<br />

8RT<br />

g<br />

T<br />

DK� � 9700r<br />

(10.12)<br />

3 � MW MW<br />

where r is the pore radius in cm, T in K, MW in g/mol, Dk, in cm 2 /s.<br />

219


10.2.2.3 Surface Diffusion<br />

In most cases, the surface diffusion coefficient is unknown as the heat of<br />

adsorption is not available. Furthermore, the surface diffusivity is a strong function of<br />

the amount of mercury adsorbed and the sorbent surface coverage is low due to the<br />

low mercury level in the flue gas, it is therefore reasonable to assume that the surface<br />

diffusion resistance can be neglected.<br />

10.3 Modeling of adsorption in a single particle<br />

The final aim of this project is to develop a mathematical model that can<br />

simulate mercury adsorption <strong>by</strong> a carbon cake on the fabric filter bags. A single<br />

particle model is the core and starting points of the filter model. The single particle<br />

model can be used to study how an adsorption process would vary with parameters<br />

such as particle size, bulk concentration, pressure, temperature, pore size, and<br />

adsorption affinity. Analytical solution of the single particle model is available when<br />

linear adsorption isotherm is used. However, the single particle model works only at<br />

constant gas atmosphere. The gas concentration changes in time for both the fixedbed<br />

and fabric filter adsorption processes. Therefore a numerical solution of the<br />

single particle model is required in order to incorporate it to fixed-bed and fabric<br />

filter models.<br />

Do [1] has made a detailed description of the single particle adsorption model<br />

using linear isotherm. Both analytical solution and numerical solution using<br />

orthogonal collocation method with subroutines in MATLAB are provided in his<br />

book. Fixed-bed and fabric filter models in this work are further developed on the<br />

basis of the single particle model.<br />

Since the mercury level in the flue gas is very low, the adsorption system can<br />

be treated as isothermal. Mass balance around a thin shell element in the particle<br />

gives [1]:<br />

�C �C�<br />

1 � s �C<br />

� p �(1 �� p) � De ( r )<br />

s<br />

�t �t r �r �r<br />

220<br />

(10.13)


where �p is the porosity of the particle, C is gaseous mercury concentration, C� is the<br />

mercury concentration in the adsorbed phase, De is the pore diffusivity, and s is the<br />

particle shape factor (s=0, 1, and 2 for slab, cylinder, and sphere, respectively).<br />

The free molecules of mercury in the pore space and the adsorbed mercury<br />

molecules at any point within a particle are assumed in equilibrium with each other.<br />

The local linear isotherm takes the form:<br />

C�� KC<br />

(10.14)<br />

where K is the Henry constant.<br />

Substituting the local equilibrium into the mass balance equation, we can<br />

obtain [1]:<br />

�C D 1 � s �C<br />

� D � C �<br />

( r )<br />

�t � � K r �r �r<br />

2<br />

e<br />

app<br />

� p (1 � p)<br />

s<br />

221<br />

(10.15)<br />

with initial condition: t=0, C=Ci, (10.16)<br />

and typical boundary conditions:<br />

�C<br />

r �0, �0<br />

�r<br />

�C<br />

r � Rp, �De �km( C �Cb)<br />

Rp<br />

�<br />

r r�R (10.17)<br />

(10.18)<br />

For slab object R is the half thickness, while for cylindrical and spherical objects, R is<br />

their respective radius. Cb is the concentration of the adsorbate in the bulk<br />

surrounding the particle, km is the external mass transfer coefficient.<br />

An analytical solution of the concentration distribution within the particle is<br />

given in the form of an infinite series [1]. The solution is only valid for a particle<br />

surrounded <strong>by</strong> a gas atmosphere not changing in time. To get a numerical solution,<br />

equation 10.15 is written in a dimensionless form <strong>by</strong> defining following nondimensional<br />

variables and parameters:<br />

C r D t C C<br />

y � ; x� ; � � ; y � ; y �<br />

C R R C C<br />

app b i<br />

2 b i<br />

0 p p<br />

0 0<br />

�y 1 � s �y<br />

� ( x )<br />

(10.19)<br />

s<br />

��x�x �x<br />

Initial condition: �=0, y=yi (10.20)


Boundary conditions:<br />

�y<br />

x �0, �0<br />

�x<br />

222<br />

(10.21)<br />

x 1; y<br />

Bi(<br />

yb<br />

y)<br />

x<br />

� � �<br />

�<br />

(10.22)<br />

�<br />

where Bi is the Biot number � km<br />

R p De<br />

.<br />

The problem has symmetry at x=0, and it is useful to utilize this <strong>by</strong> making<br />

the transformation of u=x 2 , and the differential equation becomes [1]:<br />

2<br />

�y � y �y<br />

�4u �2( s�1)<br />

2<br />

���u � u<br />

(10.23)<br />

The equation is solved <strong>by</strong> the orthogonal collocation method [4]. The domain u�(0,1)<br />

is represented <strong>by</strong> n interior collocation points. Taking the boundary point (u=1) as the<br />

(n+1) -th point, we have a total of n+1 interpolation points. The first and second<br />

derivatives at these interpolation points are related to the functional values at all<br />

points as given below:<br />

�y<br />

n�1<br />

� Aij y j<br />

�u i j<br />

� (10.24)<br />

�<br />

y<br />

2 n�1<br />

� 2 � Bij y j<br />

(10.25)<br />

�u i j<br />

The matrices A and B are constant matrices once n+1 interpolation points have been<br />

chosen. The mass balance equation is valid at any point within the u domain.<br />

Evaluating the equation at the i th interior collocation point we get:<br />

n�1<br />

�yi � Cy ij j<br />

�� j�1<br />

� (10.26)<br />

For i=1, 2,…n, where<br />

C �4uB �2(1 � s) A<br />

(10.27)<br />

ij i ij ij<br />

Numerical calculation of the average gaseous mercury concentration inside the<br />

particle is obtained <strong>by</strong>:<br />

1<br />

() (, ) ( 1) (, ) s<br />

Ct � CtxdV� s� Ctxxds<br />

V<br />

V<br />

1<br />

� � (10.28)<br />

0


1 s�1<br />

2<br />

( s �1)<br />

Ct () � Ctxu (, ) du<br />

2 � (10.29)<br />

0<br />

The integration is evaluated <strong>by</strong> Radau quadrature [1,4]:<br />

1 s�1 1<br />

n�1<br />

2<br />

� �<br />

s �1<br />

Ctxu (, ) du� (1 �u) uCtxdu (, ) ��wC k k;<br />

� �0; � �<br />

2<br />

� � (10.30)<br />

0 0<br />

where the weight factors wk are the Radau quadrature weights.<br />

The calculated Radau quadratue weights <strong>from</strong> the program are normalized [4]:<br />

k Wk ( � , � )<br />

I<br />

k �1<br />

w<br />

� (10.31)<br />

( �� , ) �( � �1) �( � �1) 2 s �1<br />

I � � for � �0and � �<br />

�( � �� �2) s �1<br />

2<br />

2<br />

w � W<br />

s �1<br />

k k<br />

1 s�1 2<br />

n�1<br />

0<br />

k �1<br />

223<br />

(10.32)<br />

(10.33)<br />

( s �1)<br />

Ct () � Ctxu (, ) du��WC k k<br />

2 � (10.34)<br />

The boundary condition at the particle surface becomes:<br />

�y<br />

Bi<br />

u �1; � ( yb� y)<br />

�u<br />

2<br />

Bi<br />

(10.35)<br />

N �1<br />

� An�1, jyj � ( yb � yn�1)<br />

(10.36)<br />

j�1<br />

2<br />

From which we can solve for the concentration at the boundary in terms of other<br />

dependent variables [1]:<br />

y<br />

n�1<br />

y<br />

n<br />

2<br />

� � A y<br />

�<br />

2<br />

1�<br />

An�1,<br />

n�1<br />

Bi<br />

b n�1, j j<br />

Bi j�1<br />

(10.37)<br />

The mass balance equation together with initial and boundary conditions are solved<br />

numerically <strong>by</strong> combination of collocation and Runge-Kutta methods using<br />

MATLAB [1]. Both the analytical solution and numerical solution give the same<br />

results for the single particle adsorption model using local linear isotherm [1].<br />

The developed model is used to simulate mercury adsorption <strong>by</strong> a single<br />

Darco Hg activated carbon particle exposed to elemental mercury at 150�C in


simulated cement kiln flue gas with 1000 ppmv NO, 23 ppmv NO2, 1000 ppmv SO2,<br />

10 ppmv HCl, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. The inputs to the model<br />

are presented in table 10.3. The Henry’s constant is derived <strong>from</strong> fixed-bed<br />

experiments using 10 mg Darco Hg carbon in 2 g sand. Table 10.4 presents the<br />

calculated diffusion coefficients, external mass transfer coefficient and Biot number<br />

<strong>by</strong> the program.<br />

Table 10.3. Inputs to the single particle adsorption model.<br />

Parameters Unit Value<br />

Temperature �C 150<br />

Carbon particle diameter µm 16<br />

Carbon true density kg/m 3 2200 [1]<br />

Carbon particle porosity - 0.73 [1]<br />

Carbon pore radius nm 10 [1]<br />

Bed porosity - 0.5<br />

Reactor diameter mm 18<br />

Flow rate Nl/min 2.75<br />

Hg inlet concentration µg/Nm 3 170<br />

Henry’s constant m 3 /g 10.24<br />

Collocation point number - 10<br />

Table 10.4. Calculated diffusion coefficients, external mass transfer coefficient and<br />

Biot number <strong>by</strong> the single particle adsorption model.<br />

Parameters Unit Value<br />

Diffusion coefficient of Hg 0 in N2, DHg,N2 m 2 /s 2.44e-5<br />

Knudsen diffusion coefficient of Hg 0 , DK m 2 /s 1.41e-6<br />

Pore diffusion coefficient, Dp m 2 /s 1.33e-6<br />

Effective pore diffusion coefficient, De m 2 /s 7.10e-7<br />

Apparent diffusion coefficient, Dapp m 2 /s 1.17e-13<br />

External mass transfer coefficient, km m/s 3.37<br />

Biot number - 37.95<br />

The concentration profiles of elemental mercury inside the particle at different<br />

times are illustrated in figure 10.1. At the beginning there is a sharp concentration<br />

profile inside the particle, indicating larger diffusion resistance inside the particle<br />

compared to the boundary layer. At later stage, the concentration profile becomes flat<br />

224


until the mercury concentration reaches the bulk level at all positions inside the<br />

particle when the adsorption equilibrium is obtained.<br />

Figure 10.1. Simulated mercury concentration profile inside the particle at different<br />

time. The corresponding time for each curve <strong>from</strong> bottom to top is 0.06, 0.29, 0.77,<br />

2.20, 5.40, 14.04, 33.84, 94.32, 323.64, 1425.24, 3600 s, respectively. Inputs to the<br />

model are given in table 10.3.<br />

The amount of elemental mercury adsorbed as a function of time for different<br />

particle sizes is illustrated in figure 10.2. The calculated external mass transfer<br />

coefficient, Biot number, and time for equilibrium adsorption are given in the table<br />

10.5. The larger the particle is, the larger the Biot numbers are. This indicates that the<br />

larger particle has larger internal mass transfer resistance. As a result, it takes the<br />

larger particle longer time to reach the equilibrium. For a 16 µm particle it takes 7.2<br />

min to reach the equilibrium, while for a 100 µm particle it takes more than four<br />

hours.<br />

225


Figure 10.2. Simulated amount of mercury adsorbed in the Darco Hg carbon particle<br />

as a function of time for particles with different diameters.<br />

Table 10.5. Calculated external mass transfer coefficient, Biot number, and simulated<br />

equilibrium approach time for Darco Hg carbon with different particle sizes.<br />

d=5 �m d=16 �m d=50 �m d=100 �m d=200 �m d=300 �m<br />

External<br />

mass transfer<br />

coefficient,<br />

km (m/s)<br />

10.26 3.37 1.18 0.64 0.36 0.26<br />

Biot number 36.14 37.95 41.48 45.15 50.73 55.24<br />

99%<br />

equilibrium<br />

approach<br />

time (h)<br />

0.01 0.12 0.89 4.28 17.04 36.41<br />

10.4 Fixed bed adsorption model<br />

In this project a plug flow model with linear equilibrium isotherm, external<br />

and intraparticle mass transfer resistances is developed. Due to the low level of<br />

226


mercury applied in this project the system can be treated as isothermal. The plug-flow<br />

model means that the fluid velocity profile is uniform at all radial positions, a fact<br />

which generally involves turbulent flow conditions. In addition, it is assumed that the<br />

fixed-bed adsorption reactor is packed randomly with adsorbent particles. The<br />

adsorption process is supposed to be very fast relative to the convection and diffusion<br />

effects; subsequently, local equilibrium will exist inside the adsorbent particles [5].<br />

If the solid particles are small, the axial diffusion effects can be ignored and<br />

the main mode of transport in the mobile fluid phase is <strong>by</strong> convection [6]. Consider a<br />

section of the fixed bed column with a length of �z, cross section area of A, and bed<br />

porosity of �b, as shown in figure 10.3, a mass balance of the mercury contained in<br />

both phase, we get [6]:<br />

�C<br />

�q<br />

v0 AC(<br />

z,<br />

t)<br />

� v0<br />

AC(<br />

z � �z,<br />

t)<br />

� � b A�z<br />

� ( 1�<br />

� b ) A�z<br />

(10.38)<br />

�t<br />

�t<br />

where v0 is the superficial fluid velocity. Dividing through <strong>by</strong> A�z and taking limit,<br />

we get the overall balance of the mercury [6]:<br />

�C �C �q<br />

v0<br />

��b �(1 ��b) �0<br />

�z �t �t<br />

q is the volume-average mercury loading per unit volume of porous pellet,<br />

Figure 10.3. Sketch of a fixed-bed absorber.<br />

Using the void velocity u, we get:<br />

�C �C 1�<br />

�b<br />

�q<br />

u � � �0<br />

�z �t � �t<br />

b<br />

227<br />

(10.39)<br />

(10.40)


q can be expressed as [2]:<br />

3 P R<br />

� � (10.41)<br />

2<br />

q r qdr<br />

3<br />

RP<br />

0<br />

where Rp is the radius of the carbon particle.<br />

Equation 10.40 gives the concentration of the mercury in the bulk gas as a<br />

function of time and location in the bed. The concentration of mercury in the gas<br />

within the pores of a carbon particle is obtained <strong>by</strong> solving equation 10.13.<br />

The simultaneous solution of equation 10.13 and10.41 is a hard task, which<br />

can be avoided <strong>by</strong> using the tank-in-series method. The fixed-bed is divided into N<br />

equal size well-mixed tanks and the mercury mass balance in the bulk gas phase for<br />

each tank can be written as:<br />

dCbi<br />

,<br />

� (10.42)<br />

bVi� FCbini , �FCbi , �kmNpAs( Cbi , �Csi<br />

, )<br />

dt<br />

where Vi is the volume of each tank, F is the flow rate through the bed, Cbin,i and Cb,i<br />

is the inlet and outlet mercury concentration in tank i, respectively, Np is the particle<br />

number in tank i, As is the outer surface area of one particle, Cs,i is the gaseous<br />

mercury concentration at the particle surface in tank i. Giving the bed cross area A,<br />

bed height h, total mass of sorbent in the bed M, void velocity u, particle radius Rp,<br />

and density �p, the above equation can be expressed as:<br />

Ah dCbi<br />

,<br />

3M<br />

�b �u�bA( Cbin, i �Cb, i) �km ( Cb, i �Cs,<br />

i)<br />

N dt NRp�p(1<br />

��<br />

p)<br />

Further arranging equation 10.43 into:<br />

dC uN 3Mk<br />

� ( C �C ) � ( C �C<br />

)<br />

dt h R Ah<br />

bi ,<br />

m<br />

bini , bi ,<br />

�b p�p(1 ��<br />

p)<br />

bi , si ,<br />

228<br />

(10.43)<br />

(10.44)<br />

Initial condition:<br />

t=0, C=0, Cbin,i=Cb0, all tanks (10.45)<br />

t>0 Cbin= Cb0, tank 1 (10.46)<br />

Boundary conditions:<br />

�C<br />

r �0, �0<br />

�r<br />

�C<br />

r � R, �D �k ( C �C<br />

)<br />

e m R b, i<br />

�r<br />

r�R (10.47)<br />

(10.48)


Dimensionless equation can be written as:<br />

2<br />

dybi , R � p uN 3Mk<br />

�<br />

m<br />

� � ( ybini , � ybi , ) � ( ybi , � y ) 1, i �<br />

d� Dapp �� h �bRp�p(1 ��<br />

p)<br />

Ah ��<br />

Initial condition:<br />

229<br />

(10.49)<br />

�=0, y=0, ybin,i=1, all tanks (10.50)<br />

�>0 ybin= 1, first tank (10.51)<br />

Boundary conditions become:<br />

�y<br />

u �0, �0<br />

�u<br />

(10.52)<br />

�y<br />

Bi<br />

u �1; � ( yb� y)<br />

(10.53)<br />

�u<br />

2<br />

The boundary-value partial differential equation along the particle radius<br />

(equation 10.19) is solved <strong>by</strong> the orthogonal collocation method [4]. The particle<br />

radius is represented <strong>by</strong> n interior collocation points. The boundary point is the<br />

(n+1) th point. For each tank the initial-value ordinary equations contain n+1 equation<br />

for the particle collocation points and another equation for the bulk phase mercury in<br />

the bed:<br />

2<br />

dyn 2 R �uN 3Mk<br />

�<br />

�<br />

m<br />

� � ( ybin, i � yn�2) � ( yn�2 � yn�1)<br />

�<br />

d� Dapp �� h �bR�p(1 ��)<br />

Ah ��<br />

(10.54)<br />

For the whole fixed-bed the resulting system of N(n+2) ordinary differential<br />

equations are solved <strong>by</strong> the MATLAB routine ode15s.<br />

Besides the inputs to the single particle adsorption model, the inputs to the<br />

developed fixed bed adsorption model also include tank number of 20 after parameter<br />

study the effect of tank number, a bed thickness of 5 mm, which corresponds to a<br />

mixture of 10 mg Darco Hg carbon with 2 g sand powder in the reactor, actual and<br />

baseline concentrations of SO2, NO2, and H2O in the simulated cement kiln flue gas,<br />

preexponential factor of Henry’s constant of 0.869 m 3 /g and heat of adsorption of -<br />

8543 J/mol as presented in chapter 8.


A parameter study of the model was first conducted to evaluate the effect of<br />

collocation point number inside the carbon particle and tank number on the mercury<br />

breakthrough curve of the carbon bed. Figure 10.4 illustrates the effects of collocation<br />

point number inside the carbon particle on the mercury breakthrough curve of 10 mg<br />

Darco Hg carbon tested at 75�C with 90 µg/Nm 3 mercury in the simulated cement<br />

kiln flue gas. Reasonable agreement between the simulation and experimental data is<br />

already obtained using one collocation point inside the carbon particle. Simulations<br />

using 2, 5, and 10 collocation points generate the same mercury breakthrough curve.<br />

Generally more accurate solutions can be obtained using more collocation points.<br />

Since the simulation can be done within 30 s, a collocation point of 10 is used as<br />

default input to the program.<br />

Figure 10.4. Effects of collocation point number inside the carbon particle on the<br />

simulated mercury breakthrough curves of 10 mg Darco Hg mixed with 2 g sand<br />

tested at 75�C using 2.75 Nl/min simulated flue gas with 90 µg Hg 0 /Nm 3 , 10 ppmv<br />

HCl, 1000 ppmv NO, 1000 ppmv SO2, 23 ppmv NO2, 1 vol.% H2O, 6 vol.% O2, and<br />

21 vol.% CO2. A tank number of 20 is applied in the simulation.<br />

230


Figure 10.5 presents the effects of applied tank number in the simulation on<br />

the predicted mercury breakthrough curve of 10 mg Darco Hg carbon tested at 75�C<br />

with 90 µg/Nm 3 mercury in the simulated cement kiln flue gas. Better agreement<br />

between the simulation and experimental data is obtained when larger tank number is<br />

applied. When the tank number is above 20, the produced breakthrough profile is<br />

almost the same.<br />

Figure 10.5. Effects of applied tank number on the simulated mercury breakthrough<br />

curves of 10 mg Darco Hg mixed with 2 g sand tested at 75�C using 2.75 Nl/min<br />

simulated flue gas with 90 µg Hg 0 /Nm 3 , 10 ppmv HCl, 1000 ppmv NO, 1000 ppmv<br />

SO2, 23 ppmv NO2, 1 vol.% H2O, 6 vol.% O2, and 21 vol.% CO2. A collocation point<br />

number of 10 inside the carbon particle is applied in the simulation.<br />

Validation and parametric study of mercury adsorption <strong>by</strong> the activated<br />

carbon is conducted <strong>by</strong> simulation and comparison with the experimental data as<br />

shown in figure 10.6-10.12. The developed fixed bed model can reasonably simulate<br />

the effects of temperature, mercury inlet concentration, flow gas rate, carbon particle<br />

size, and SO2, H2O, NO2 level in the flue gas on the mercury breakthrough curve of<br />

fixed bed with 10 mg carbon in 2g sand powder. Isotherm study of crushed Norit<br />

231


RB4 pellets is not performed and the Henry’s constant for Norit RB4 carbon is<br />

calculated <strong>by</strong> comparing the equilibrium adsorption capacity with Darco Hg carbon at<br />

the same conditions.<br />

Figure 10.6. Comparison of simulation and experimental data for effect of adsorption<br />

temperature on mercury breakthrough of 10 mg Darco Hg mixed with 2 g sand using<br />

2.75 Nl/min simulated flue gas with 90 µg Hg 0 /Nm 3 ,1000 ppmv NO, 23 ppmv NO2,<br />

10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

The model can clearly simulate the effect of adsorption temperature on<br />

mercury breakthrough curve of the carbon bed, i.e., faster mercury breakthrough is<br />

obtained at higher adsorption temperature. The Henry constants at each temperature<br />

are calculated <strong>from</strong> the derived preexponential factor and heat of adsorption. The best<br />

agreement between the simulation and experimental data is obtained for adsorption<br />

test at 75�C as shown in figure 10.6.<br />

232


Figure 10.7. Comparison of simulation and experimental data for effect of elemental<br />

mercury inlet level on mercury breakthrough of 10 mg Darco Hg mixed with 2 g sand<br />

tested at 120�C using 2.75 Nl/min simulated flue gas with 1000 ppmv NO, 23 ppmv<br />

NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

Figure 10.7 shows that the best agreement between the simulation and<br />

experimental data regarding the effects of mercury inlet level is the mercury<br />

breakthrough curve of 57 µg/Nm 3 mercury level. The simulation slightly overpredicts<br />

the mercury adsorption with 95 µg/Nm 3 and underpredicts the mercury adsorption for<br />

tests with 27 µg/Nm 3 in the flue gas. When taking the experimental uncertainty into<br />

account, the simulation is acceptable as the average uncertainty of the equilibrium<br />

mercury adsorption capacity is about ±10%.<br />

233


Figure 10.8. Comparison of simulation and experimental data for effect of flue gas<br />

flow rate on mercury breakthrough of 10 mg Darco Hg mixed with 2 g sand tested at<br />

150�C using 2.75 Nl/min simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv<br />

NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and<br />

6 vol.% O2.<br />

Figure 10.8 illustrates that there are good agreements between the simulations<br />

and experimental data regarding the initial mercury breakthrough time when the<br />

mercury concentration after the carbon bed starts to increase and the model correctly<br />

predicts that. The model overpredicts the effect of changing gas flow especially for<br />

high flow rates.<br />

234


Figure 10.9. Comparison of simulation and experimental data for effect of particle<br />

size on mercury breakthrough for 10 mg crushed Norit RB4 pellets in 2 g sand tested<br />

at 150�C using 2.75 Nl/min simulated flue gas with 160-170 µg Hg 0 /Nm 3 , 1000 ppmv<br />

NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and<br />

6 vol.% O2.<br />

There are good agreements between the simulations and experimental data for<br />

the effect of particle size for Norit RB4 over the size range of 38-325 µm, as<br />

illustrated in figure 10.9.<br />

235


Figure 10.10. Comparison of simulation and experimental data for effect of SO2<br />

concentration in the flue gas on mercury breakthrough profile of 10 mg Darco Hg<br />

mixed with 2 g sand tested at 150�C using 2.75 Nl/min simulated flue gas with 160-<br />

170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1 vol.% H2O, 6<br />

vol.% O2, and 21 vol.% CO2.<br />

236


Figure 10.11. Comparison of simulation and experimental data for effect of water<br />

concentration in the flue gas on mercury breakthrough profile of 10 mg Darco Hg<br />

mixed with 2 g sand tested at 150�C using 2.75 Nl/min simulated flue gas with 160-<br />

170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 6<br />

vol.% O2, and 21 vol.% CO2.<br />

237


Figure 10.12. Comparison of simulation and experimental data for effect of NO2<br />

concentration in the flue gas on mercury breakthrough curves of 10 mg Darco Hg<br />

mixed with 2 g sand tested at 150�C using 2.75 Nl/min simulated flue gas with 160-<br />

170 µg Hg 0 /Nm 3 , 10 ppmv HCl, 1000 ppmv NO, 1000 ppmv SO2, 1 vol.% H2O, 6<br />

vol.% O2, and 21 vol.% CO2.<br />

Figure 10.10 and 10.11 show that there are good agreements between the<br />

simulations and experimental data on the effects of SO2 and H2O levels in the<br />

simulated cement kiln flue gas. The model slightly overpredicts the mercury<br />

adsorption rate with 100 and 5 ppmv NO2 in the flue gas and slightly underpredicts<br />

the mercury adsorption with 23 ppmv NO2 as shown in figure 10.12. The effects of<br />

these gases on the mercury adsorption capacity are evaluated <strong>by</strong> the derived<br />

correlations between mercury adsorption capacity and gas concentrations, as<br />

presented in chapter 8.<br />

238


10.5 Conclusions<br />

Mathematical models for mercury adsorption <strong>by</strong> a single carbon particle and a<br />

fixed carbon bed are developed. Local equilibrium within the carbon particle is<br />

assumed and the adsorption system is assumed to be isothermal due to the low<br />

mercury concentration presented in the flue gas. The models account for both the<br />

external and internal mass transfer resistances. The orthogonal collocation method is<br />

used to solve mercury diffusion and adsorption inside a sorbent particle. The fixedbed<br />

model is solved <strong>by</strong> a tank-in-series method.<br />

Henry’s constant obtained <strong>from</strong> fixed-bed investigation of mercury adsorption<br />

<strong>by</strong> activated carbon in the simulated cement kiln flue gas is used as input to the<br />

models. The single particle model can simulate the mercury concentration profile and<br />

amount of adsorbed mercury inside the carbon particle as a function of adsorption<br />

time.<br />

The developed fixed bed model can reasonably simulate the effects of<br />

adsorption temperature, mercury inlet concentration, flow gas rate, carbon particle<br />

size, and SO2, H2O, NO2 level in the flue gas on the mercury breakthrough curve of<br />

fixed bed with 10 mg carbon in 2 g sand powder. The developed models are useful<br />

tools for understanding the mercury adsorption <strong>by</strong> the activated carbon and<br />

interpretation of the experimental results.<br />

10.6 List of symbols<br />

A: carbon particle external surface area (m 2 )<br />

A: cross area of the fixed-bed (m 2 )<br />

A: matrix in equation 10.24<br />

As: outer surface area of one carbon particle (m 2 )<br />

b: Langmuir equilibrium constant (m 3 /g)<br />

B: matrix in equation 10.25<br />

Bi: dimensionless Biot number<br />

C: gaseous mercury concentration (µg/m 3 )<br />

C: matrix in equation 10.26<br />

Cb: gas bulk mercury concentration (µg/m 3 )<br />

Cbo: initial gas bulk mercury concentration (µg/m 3 )<br />

Cb,i: outlet mercury concentration in tank i (µg/m 3 )<br />

239


Cbin,i: inlet mercury concentration in tank i (µg/m 3 )<br />

Ci: initial mercury concentration (µg/m 3 )<br />

Cs,i : gaseous mercury concentration at the particle surface in tank i (µg/m 3 )<br />

cp: specific heat (J/(kg.K)<br />

Cs: gaseous mercury concentration at the particle surface (µg/m 3 )<br />

Cµ: adsorbed mercury concentration in the sorbent (µg/m 3 )<br />

Cµs: saturated concentration of adsorbed mercury in the sorbent (µg/m 3 )<br />

DAB: binary molecular diffusion coefficient (m 2 /s)<br />

Dapp: apparent diffusion coefficient (m 2 /s)<br />

De: effective diffusion coefficient (m 2 /s)<br />

Di: molecular diffusion coefficient (m 2 /s)<br />

Dk: Knudsen pore diffusion coefficient (cm 2 /s)<br />

dp: particle diameter (m)<br />

F: gas flow rate (m 3 /s)<br />

G: fluid mass velocity (kg/(m 2 .s))<br />

h: heat transfer coefficient (W/(m 2 .K))<br />

h: bed height (m)<br />

k: thermal conductivity (W/(m.K))<br />

k : the Boltzmann’s constant (J/K)<br />

K: Henry’s constant<br />

KF: Frendlich constant<br />

km: gas film mass transfer coefficient (m/s)<br />

M: carbon load in the fixed-bed (mg)<br />

MW: mole weight (g/mol)<br />

n: exponent in isotherm equations of Frendlich, and Langmuir-Frendlich<br />

n: number of interior collocation points<br />

N: amount of transported mercury in equation 10.1 (µg)<br />

N: tank number<br />

Np: carbon particle number in the tank<br />

NPr: Prandtl number<br />

NSc: Schmidt number<br />

NRe: Reynolds number<br />

P: pressure (atm)<br />

q: mercury concentration in the sorbent (µg/m 3 )<br />

Q: heat (W)<br />

r: radial coordinate (m)<br />

r: pore radius (cm)<br />

Rg: universal gas constant, 8.314 (J/(mol.K))<br />

Rp: sorbent particle radius (m)<br />

s: heterogeneity parameter in Unilan isotherm equation<br />

240


s: the particle shape factor (s=0, 1, and 2 for slab, cylinder, and sphere, respectively)<br />

t: exponent in isotherm equations of Toth<br />

t: time (s)<br />

T: temperature (K)<br />

Tb : gas temperature in the bulk (K)<br />

Ts: gas temperature at the particle surface (K)<br />

u: dimensionless parameter, u=x 2<br />

u: void velocity (m/s)<br />

v0: superficial fluid velocity (m/s)<br />

Vi: volume of the tank (m 3 )<br />

Vp: volume of the carbon particle (m 3 )<br />

wk: Radau quadratue weight<br />

Wk: normalized Radau quadratue weight<br />

x: dimensionless radius<br />

y: dimensionless mercury concentration<br />

yb: dimensionless gas bulk mercury concentration (µg/m 3 )<br />

yi: dimensionless initial mercury concentration (µg/m 3 )<br />

z: axial coordinate (m)<br />

Greek symbols<br />

�: parameter defined <strong>by</strong> equation 10.31<br />

�: parameter defined <strong>by</strong> equation 10.31<br />

�AB: Lennard-Jones 12-6-potentials for specie A and B<br />

�b: bed void fraction<br />

�p: sorbent particle porosity<br />

�p: sorbent particle density (kg/m 3 )<br />

µ: dynamic viscosity (kg/(m.s))<br />

�AB : molecular collision diameter (Å)<br />

�D,AB: dimensionless parameter in equation 10.8 and 10.11<br />

�Hads: heat of adsorption (J/mol)<br />

�: dimensionless time<br />

10.7 References<br />

[1] D.D. Do. Adsorption analysis: equilibria and kinetics, Imperial College Press, 1998.<br />

[2] J.D. Seader, E.J. Henley, Separation process principles, John Wiley & Sons, Inc. 1998.<br />

[3] P.J. Gardner, P. Pang, S.R. Preston, Binary gaseous-diffusion coefficients of mercury and<br />

of zinc in hydrogen, helium, argon, nitrogen, and carbon-dioxide, J. Chem. Eng. Data. 36<br />

(1991) 265-268.<br />

241


[4] J. Villadsen, M.L. Michelsen, Solutions of differential equation models <strong>by</strong> polynomial<br />

approximation, Prentice-Hall, Inc., 1978.<br />

[5] D.M. Ruthven, Principles of adsorption and adsorption processes, John Wiley & Sons,<br />

Inc., 1984.<br />

[6] R.G. Rice, D.D. Do. Applied mathematics and modelling for chemical engineers, John<br />

Wiley & Sons, Inc, 1995.<br />

242


243<br />

11<br />

Simulation of mercury removal <strong>by</strong> activated<br />

carbon injection upstream of a fabric filter<br />

This chapter deals with the development of a two-stage model for simulation<br />

of mercury removal <strong>by</strong> carbon injection upstream of a fabric filter. First the<br />

development of duct-fabric filter models is presented, and then the models are<br />

compared with available experimental data <strong>from</strong> pilot-scale investigation.<br />

11.1 Common assumptions for mercury removal in the duct and<br />

fabric filter<br />

<strong>Mercury</strong> removal <strong>by</strong> the sorbent injection upstream of a fabric filter consists<br />

of two stages, i.e., the duct and filter sections as illustrated in figure 11.1. Powdered<br />

sorbent such as activated carbon is metered to the injection point at a rate<br />

proportional to the gas stream flow. Once dispersed, mercury species diffuse to the<br />

particle surface and migrate into pores of the activated carbon particle. The carbon<br />

particles remain suspended in the moving gas stream in the duct for periods of one to<br />

three seconds. It then deposits onto the carbon cake formed on the filter bags.<br />

Additional mercury capture takes place when the mercury-containing gas stream<br />

passes through the carbon cake. The carbon cake grows with filtration time and after<br />

a certain time the pressure drop across the filter reaches its threshold value and the<br />

cleaning process is initiated <strong>by</strong> pulse injection of compressed air. A fraction of the<br />

filter bags is periodically cleaned to relieve the pressure drop across the fabric filter.


Figure 11.1. Sketch of the mercury removal process <strong>by</strong> carbon injection upstream of a<br />

fabric filter.<br />

A mathematical model is a useful tool to simulate the mercury capture and<br />

evaluate the mercury removal efficiency for various operational conditions. An<br />

advanced model can provide a rational basis for describing and characterizing the<br />

effectiveness of mercury removal <strong>by</strong> sorbent injection and provide guidelines for<br />

developing new types of sorbents and improve of the process.<br />

To make the mathematics tractable, following assumptions are made:<br />

1. The relevant mercury species in the gas phase is assumed to be either<br />

elemental mercury or mercuric chloride. Elemental mercury is much more difficult to<br />

remove if the sorbent cannot oxidize it. As shown in chapter 8 and 9, similar<br />

adsorption behavior of elemental mercury and mercury chloride <strong>by</strong> activated carbon<br />

is observed since significant oxidation of mercury <strong>by</strong> the activated carbon occurs if<br />

HCl is present in the gas above few ppmv. This is the case in most practical systems<br />

and will be assumed here.<br />

2. Activated carbon particles are spherical, uniform in size and uniformly<br />

dispersed in the duct and filter cake.<br />

244


3. The temperature is constant and uniform through the system. <strong>Mercury</strong><br />

adsorption heat effects are neglected due to the trace level mercury concentrations.<br />

The adsorption equilibrium is described <strong>by</strong> Henry’s law as shown in chapter 8.<br />

4. Both the gas and the solid flow rates are constant. In reality, there are changes<br />

of both differential pressures over the filter bag and cake porosity with time. The<br />

cleaned section of the filter <strong>by</strong> pulse jet would have less hydraulic resistance,<br />

resulting in a larger fraction of the flow diverted to this section. There would be a<br />

dynamic redistribution of the flow as the cake grows on the filter bag surface. Flora et<br />

al. [1] evaluated the effect of the dynamic redistribution of flow on removal of<br />

mercury using activated carbon injection in a fabric filter system. The magnitude of<br />

this impact is small compared with the potential impact caused <strong>by</strong> uncertainties in the<br />

isotherm and mass transfer parameters. When the differential pressure over the bag<br />

increases the system fan speed will be increased correspondingly to maintain the<br />

same filtration velocity. It is therefore reasonable to assume constant gas flow<br />

through the filter cake.<br />

5. <strong>Mercury</strong> adsorption on the duct walls and filter fabric is negligible.<br />

Equilibrium conditions are reached between the gas phase and walls/fabric so that no<br />

net exchange of mercury is present.<br />

6. <strong>Removal</strong> of mercury <strong>from</strong> the bulk gas phase is caused solely <strong>by</strong> adsorption<br />

on the activated carbon.<br />

7. A mass transfer boundary layer causes resistance to mass transfer <strong>from</strong> the<br />

bulk gas phase to the activated carbon particle external surface, and mass transfer<br />

within the carbon particle is controlled <strong>by</strong> pore diffusion.<br />

8. Since the mercury level in the flue gas is very low and the surface diffusivity<br />

is a strong function of the amount of mercury adsorbed, it is reasonable to assume<br />

that the surface diffusion resistance can be neglected.<br />

9. The free gaseous mercury molecules in the pore and the adsorbed mercury<br />

molecules at any point within a particle are in equilibrium with each other. The local<br />

adsorption kinetics is much faster than the diffusion process into the particle.<br />

245


11.2 Duct model<br />

Part of the mercury is removed in the duct section. Simulation of mercury<br />

removal in the duct is presented in this section taking into account relevant<br />

mechanisms.<br />

The flue gas is assumed to travel in plug flow along the duct. To verify<br />

whether the slip velocity between the activated carbon particles and the gas is<br />

relevant, Scala evaluated the terminal velocity of the particles for the particle sizes of<br />

interest, e.g., less than 100 �m [2-4]. Results indicated that terminal velocities are<br />

always more than one order of magnitude lower than typical flue gas velocity so that<br />

it is reasonable to assume that particles travel at the same velocity as the flue gas. The<br />

particle Reynolds number was always smaller than one, justifying the assumption of<br />

Stokes regime.<br />

Mass balance around a thin shell element in the spherical particle gives:<br />

�C �C�<br />

1 � 2 �C<br />

� p ��p(1 �� p) � De ( r )<br />

(11.1)<br />

2<br />

�t �t r �r �r<br />

where �p is the porosity of the particle, C is the gaseous mercury concentration, C� is<br />

the adsorbed mercury per unit mass of the particle, �p is the particle density and De is<br />

the effective diffusivity.<br />

The local linear isotherm takes the form:<br />

C�� KC<br />

(11.2)<br />

where K is the Henry’s constant.<br />

Substituting the local equilibrium into the mass balance equation, we can get:<br />

�C D 1 � �C<br />

� Dapp� C �<br />

( r )<br />

�t � � K r �r �r<br />

2 e<br />

2<br />

� p (1 � p) �p<br />

2<br />

246<br />

(11.3)<br />

Assuming plug flow and no slip velocity, mercury adsorption in the duct can be<br />

treated as a batch adsorber. Assuming perfect mixing, the mass balance of mercury in<br />

the bulk phase is:<br />

dCb<br />

V �� AJ<br />

(11.4)<br />

Rp<br />

dt<br />

where V is the volume of the adsorber, Cb is the concentration of mercury in the<br />

adsorber, A is the total exterior surface area of all carbon particles in the adsorber, and


J is the mass transfer into the carbon particle per unit interfacial area. If the<br />

Rp<br />

particles are spheres, the total exterior surface area is<br />

mp<br />

3<br />

A �<br />

� (1 � � ) R<br />

p p p<br />

where mp is the mass of the particles and Rp is the particle radius. Equation 11.4 can<br />

be rearranged into:<br />

dC m<br />

b p 1 3 3�<br />

3�km<br />

�� J �� J �� ( Cb�C )<br />

Rp Rp Rp<br />

dt V � (1 �� ) R � (1 �� ) R � (1 ��<br />

) R<br />

p p p p p p p p p<br />

247<br />

(11.5)<br />

(11.6)<br />

where � is the carbon load in the flue gas (kg/m 3 ), km is the external mass transfer<br />

coefficient,<br />

C is the gaseous mercury concentration at the carbon particle surface.<br />

Rp<br />

Initial condition: t=0, C=0, Cb=Cb0, (11.7)<br />

Boundary conditions:<br />

�C<br />

r �0, �0<br />

�r<br />

�C<br />

r � R , �D �k ( C �C<br />

)<br />

p e m R b<br />

p<br />

�r<br />

Rp, t<br />

(11.8)<br />

(11.9)<br />

Equations 11.3 and 11.6 are written in a dimensionless form <strong>by</strong> defining the<br />

following non-dimensional variables and parameters:<br />

C r D t C<br />

y � ; x� ; � � ; y � ;<br />

C R R C<br />

app b<br />

2 b<br />

b0 p p b0<br />

where y is the interparticle gas concentration.<br />

�y 1<br />

� 2<br />

��x � 2 �y<br />

( x )<br />

�x �x<br />

dy<br />

d� �<br />

k<br />

� R<br />

R<br />

D<br />

( )<br />

3�k<br />

R<br />

� � D<br />

( )<br />

2<br />

b 3�<br />

m<br />

p m p<br />

�� yb �y �� y 1 b �y<br />

1<br />

p(1 � p) p app p(1 � p) app<br />

(11.10)<br />

(11.11)<br />

Initial condition: �=0, y=0, yb=1 (11.12)<br />

Boundary conditions become:<br />

�y<br />

x �0, �0<br />

�x<br />

(11.13)


�y<br />

kmRp x �1; � ( yb � y) � Bi( yb � y)<br />

�x<br />

D<br />

e<br />

248<br />

(11.14)<br />

The problem of diffusion and adsorption in the carbon particle has symmetry at x=0,<br />

and it is useful to utilize this <strong>by</strong> making the transformation of u=x 2 , and the<br />

differential equation becomes:<br />

2<br />

�y � y �y<br />

�4u�6 2<br />

���u � u<br />

(11.15)<br />

Boundary conditions become:<br />

�y<br />

u �0, �0<br />

(11.16)<br />

�u<br />

�y<br />

Bi<br />

u �1; � ( yb� y)<br />

(11.17)<br />

�u<br />

2<br />

The equation is solved <strong>by</strong> the orthogonal collocation method [5]. The domain<br />

u�(0,1) is represented <strong>by</strong> n interior collocation points. Taking the boundary point<br />

(u=1) as the (n+1) -th point, we have a total of n+1 interpolation points. The first and<br />

second derivatives at these interpolation points are related to the functional values at<br />

all points as given below:<br />

�y<br />

n�1<br />

� Aij y j<br />

�u i j<br />

� (11.18)<br />

�<br />

y<br />

2 n�1<br />

� 2 � By ij j<br />

(11.19)<br />

�u i j<br />

The matrices A and B are constant matrices once n+1 interpolation points have been<br />

chosen. The mass balance equation is valid at any point within the u domain.<br />

Evaluating the equation at the i th interior collocation point we get:<br />

n�1<br />

�yi � Cy ij j<br />

�� j�1<br />

� (11.20)<br />

For i=1, 2,…n+1, where<br />

C �4uB � 6A<br />

(11.21)<br />

ij i ij ij<br />

�y � �<br />

n<br />

i Cy ij j Ci, n�1yn�1 �� j�1<br />

� (11.22)<br />

The boundary condition at the carbon particle surface is:


Bi<br />

n�1<br />

� An�1, jyj � ( yb � yn�1)<br />

(11.23)<br />

j�1<br />

2<br />

From which we can solve for the concentration at the boundary in terms of other<br />

dependent variables [6]:<br />

y<br />

n�1<br />

y<br />

n<br />

2<br />

� � A y<br />

�<br />

2<br />

1�<br />

An�1,<br />

n�1<br />

Bi<br />

b n�1, j j<br />

Bi j�1<br />

Including the equation for the bulk phase mercury,<br />

dy 3�k<br />

2<br />

mR n�<br />

p<br />

�� ( yn�2 �yn�1)<br />

d� � (1 � � ) D<br />

p p app<br />

249<br />

(11.24)<br />

(11.25)<br />

n+1 initial-value ordinary differential equations are solved simultaneously <strong>by</strong><br />

MATLAB routine ode15s.<br />

The developed duct model is very similar to that developed <strong>by</strong> Scala [2-4].<br />

The main difference between the models is that Scale used Langmuir isotherm and<br />

dynamic adsorption, i.e., local equilibrium is not assumed.<br />

The model input parameters are listed in table 11.1 for simulation of mercury<br />

adsorption <strong>by</strong> injection of Darco Hg activated carbon into the duct. The Henry’s<br />

constant is derived <strong>from</strong> fixed-bed investigation as presented in chapter 8. When the<br />

gas composition is different <strong>from</strong> the baseline test, the effect of individual gas on the<br />

mercury adsorption is evaluated using correlations derived <strong>from</strong> chapter 8. Since fullscale<br />

data are not available for comparison it is the intention here to test the model<br />

ability instead of simulating the full-scale application. The simulation results are<br />

analyzed <strong>by</strong> selecting a set of operating variables as a base case for computations and<br />

to assess the influence of the relevant input variables on the process <strong>by</strong> varying them<br />

one at a time.


Table 11.1. Inputs to the duct adsorption model.<br />

Parameters Unit Value<br />

Temperature �C 75-150<br />

Actual SO2 concentration ppmv 1000<br />

Baseline SO2 concentration ppmv 1000<br />

Actual NO2 concentration ppmv 23<br />

Baseline NO2 concentration ppmv 23<br />

Actual H2O concentration % 1<br />

Baseline H2O concentration % 1<br />

Hg inlet concentration µg/Nm 3 170<br />

Carbon particle diameter µm 5-200<br />

Carbon true density kg/m 3 2200<br />

Carbon particle porosity - 0.73<br />

Carbon pore radius nm 10<br />

Carbon injection rate g/m 3 0.05-10<br />

Residence time in the duct s 0-10<br />

Henry’s constant<br />

preexponential factor<br />

m 3 /g 0.869<br />

Heat of adsorption J/mol -8543<br />

Collocation point number - 10<br />

Figure 11.2 illustrates the simulated bulk mercury concentration in the duct as<br />

a function of flight time for different injection rates of 16 µm Darco Hg carbon at<br />

150�C to the baseline gas of 170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2, 10<br />

ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. The gas<br />

bulk mercury concentration is normalized with the inlet mercury concentration. The<br />

mercury concentration decreases with increasing of residence time in the duct and the<br />

carbon load. To obtain 80% mercury removal <strong>by</strong> injection of 16 µm Darco Hg carbon<br />

at 150�C, it needs either a long residence time in the duct, i.e., long duct (> 1 g/m 3<br />

load and 10 s) or large carbon injection rate (10 g/m 3 load and 0.16 s). After 10 s in<br />

the duct the mercury removal efficiency is 77.9% and 97.6% for a carbon injection<br />

rate of 1 and 10 g/m 3 , respectively.<br />

250


Figure 11.2. Simulated gaseous mercury concentration as a function of residence time<br />

in the duct and carbon injection rate. Darco Hg carbon with a diameter of 16 µm is<br />

injected at 150�C to simulated cement kiln flue gas with 170 µg Hg 0 /Nm 3 , 1000<br />

ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.%<br />

CO2, and 6 vol.% O2.<br />

The effects of carbon particle size on mercury concentration in the duct are<br />

illustrated in figure 11.3. The gas bulk mercury concentration decreases faster for<br />

smaller carbon particles during the first 2 s in the duct and a larger mercury removal<br />

is obtained <strong>by</strong> smaller carbon particles at all residence time, indicating that diffusion<br />

resistance is relevant for mercury adsorption on carbon particles. Decreasing the<br />

particle size <strong>from</strong> 16 to 5 µm can increase the mercury removal efficiency <strong>from</strong> 77.9<br />

to 87.6% using an injection rate of 1 g/m 3 at 150�C. The improvement of mercury<br />

removal efficiency <strong>by</strong> further lowering the particle size is less pronounced (not shown<br />

in figure 11.3).<br />

251


Figure 11.3. Simulated gaseous mercury concentration as a function of residence time<br />

in the duct and carbon particle size. 1 g/m 3 Darco Hg carbon is injected at 150�C to<br />

simulated cement kiln flue gas with 170 µg Hg 0 /Nm 3 , 1000 ppmv NO, 23 ppmv NO2,<br />

10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

The effects of particle size on mercury removal can be explained <strong>by</strong> Biot<br />

numbers for different particle sizes as shown in table 10.5. Calculations show that the<br />

larger the particle is, the smaller external mass transfer coefficient, the larger the Biot<br />

numbers are. This indicates that the larger particle has relatively larger internal<br />

transfer resistance. As a result, it takes the larger particle longer time to reach the<br />

equilibrium. In all the cases calculated here, the Biot numbers are much larger than<br />

36, indicating that the internal diffusional resistance is much larger than the external<br />

mass transfer resistance.<br />

The effects of temperature on mercury removal in the duct are presented in<br />

figure 11.4. Similar mercury outlet concentrations are observed for injection 0.5 g/m 3<br />

Darco Hg carbon with a size of 16 µm for the first 2 s in the duct, and then lower<br />

mercury outlet concentrations are obtained with lower flue gas temperature and<br />

longer residence time in the duct.<br />

252


Figure 11.4. Simulated gaseous mercury concentration as a function of residence time<br />

in the duct and flue gas temperature. 0.5 g/m 3 Darco Hg carbon with a diameter of 16<br />

µm is injected at 150�C to simulated cement kiln flue gas with 170 µg Hg 0 /Nm 3 ,<br />

1000 ppmv NO, 23 ppmv NO2, 10 ppmv HCl, 1000 ppmv SO2, 1 vol.% H2O, 21<br />

vol.% CO2, and 6 vol.% O2.<br />

11.3 Model for the filter cake<br />

The filter cake formed on the bags is similar to a fixed-bed with the difference<br />

that new adsorbent is continuously fed. The fixed-bed model with constant bed<br />

thickness developed in chapter 10 is extended here to deal with this situation. During<br />

filtration without cleaning of the bag the carbon cake thickness grows with time. At<br />

the beginning no carbon particles are collected on the bag surface. After a short time<br />

�t a layer of carbon is collected and the corresponding carbon cake thickness is:<br />

�u� b<br />

L � �t<br />

(11.26)<br />

� (1 �� )(1 ��<br />

)<br />

p p b<br />

where � is the carbon load in the flue gas, u is the face velocity on the filter bags.<br />

This thin layer of carbon particles can be treated as a continuous stirred-tank reactor<br />

253


(CSTR). The bed volume is AL, the mass of carbon in the tank<br />

is M � � (1 �� ) V � � (1 �� )(1 � � ) AL.<br />

The mass balance for mercury in the bulk<br />

b p b p p b<br />

phase in the tank becomes:<br />

dCb u 3(1 ��<br />

b) km<br />

� ( Cbin �Cb) � ( Cb� C )<br />

(11.27)<br />

Rp<br />

dt L � R<br />

b p<br />

The dimensionless equation can be written as:<br />

2<br />

dy R �<br />

b p u 3(1 ��<br />

b) k �<br />

m<br />

� � ( ybin � yb) � ( yb� y ) 1 �<br />

(11.28)<br />

d� Dapp ��L �bRp��<br />

n+1 collocation points are used for the carbon particle. The bulk phase mercury<br />

balance equation becomes:<br />

2<br />

dy R �<br />

n�2pu3(1 ��<br />

b) k<br />

�<br />

m<br />

� � ( ybin � yn�2) � ( yn�2 � yn�1)<br />

�<br />

d� Dapp ��L �bRp��<br />

Initial and boundary conditions:<br />

254<br />

(11.29)<br />

�=0, y=0, ybin=1 (11.30)<br />

�y<br />

u �0, �0<br />

�u<br />

(11.31)<br />

1; ( �2 �1)<br />

2<br />

� �<br />

�y<br />

Bi<br />

u � yn<br />

yn<br />

(11.32)<br />

�u<br />

Theses n+1 equations are solved in a time interval of [0 �t] <strong>by</strong> MATLAB<br />

routine ode15s. After another �t, another layer of carbon with thickness of L is<br />

formed on the bag surface and on top of the first layer and is termed as tank 2. Now<br />

the system contains 2(n+1) initial-value ordinary differential equations which are<br />

solved simultaneously in a time interval of [0 �t] <strong>by</strong> MATLAB routine ode15s. The<br />

initial conditions for equations in tank 1 are the calculated concentration <strong>from</strong> last �t<br />

interval. The initial conditions for tank 2 are y=0, ybin,2=1. At �>0, ybin,1= yb,2. The<br />

cycle is conducted to the desired filtration time. Here it is assumed that the carbon<br />

particles are injected just at the filter inlet. The combination of duct injection and<br />

filter cake model will be presented in section 11.5. In the later case the carbon<br />

particles arriving in the filter will have already adsorbed mercury with some radial<br />

profile, i.e., y�0.


Simulation of mercury removal <strong>by</strong> the fixed-bed with moving boundary is<br />

performed using the conditions <strong>from</strong> the Durkee pilot plant study [7,8]. The inputs to<br />

the model are given in table 11.2. The flue gas temperature is taken <strong>from</strong> the field test<br />

report [7,8] and the flue gas compositions are supplied <strong>by</strong> Paone [9]. Other gas<br />

concentrations are the same as the baseline gas. The effects of CO2 and HCl are not<br />

accounted for since the effects of these gases are less pronounced compared to SO2,<br />

NO2 and H2O. Referring to results <strong>from</strong> chapter 8, when the CO2 level in the flue gas<br />

is above 21 vol.%, which is used in baseline test and deriving of the adsorption<br />

kinetics, the mercury adsorption capacity of the carbon is only slightly decreased.<br />

With HCl in the gas up to 15 ppmv, the mercury adsorption capacity is almost not<br />

affected <strong>by</strong> changing the HCl level in the gas. Large adsorption capacity is obtained<br />

without HCl in the gas.<br />

Table 11.2. Inputs to the filter cake model.<br />

Parameters Unit Value<br />

Temperature �C 138<br />

Actual SO2 concentration ppmv 5<br />

Baseline SO2 concentration ppmv 1000<br />

Actual NO2 concentration ppmv 5<br />

Baseline NO2 concentration ppmv 23<br />

Actual H2O concentration % 15<br />

Baseline H2O concentration % 1<br />

Hg inlet concentration µg/Nm3 200<br />

Carbon particle diameter µm 16<br />

Carbon true density kg/m 3 2200<br />

Carbon particle porosity - 0.73<br />

Carbon pore radius nm 10<br />

Carbon injection rate mg/m 3 8-80<br />

Filtration time s 1500<br />

Air to cloth ratio m/min 1.2<br />

Time for new cake layer min 0.5-5<br />

Henry’s constant<br />

preexponential factor<br />

m 3 /g 0.869<br />

Heat of adsorption J/mol -8543<br />

Collocation point number - 10<br />

255


The effect of the time for new carbon layer addition �t on the mercury<br />

removal efficiency of the fabric filter is illustrated in figure 11.5. Generally the more<br />

frequently the new layer is added the larger mercury removal efficiency is obtained<br />

for filtration time less than 1200 s. Smoother mercury removal efficiency curve will<br />

be obtained using a smaller time interval for adding a carbon layer. For short<br />

filtration time the new carbon layer should be added very fast in the simulation,<br />

otherwise the simulated mercury removal efficiency will be smaller due to the delay<br />

of new carbon layer addition. For filtration time larger than 1200 s same mercury<br />

removal efficiency is predicted with 5 min interval for new carbon layer addition as<br />

with smaller time interval. However, the computation time of the program is<br />

considerably reduced using an interval of 5 min compared to 1 min.<br />

<strong>Mercury</strong> removal, %<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

30 s<br />

60 s<br />

120 s<br />

300s<br />

0<br />

0 300 600 900 1200 1500 1800<br />

Time (s)<br />

Figure 11.5. Simulated effects of new cake layer addition frequency on the mercury<br />

removal efficiency of a fabric filter without cleaning of the bags. 16 mg/m 3 Darco Hg<br />

carbon with a diameter of 16 µm is injected at 138�C to simulated cement kiln flue<br />

gas with 200 µg Hg 0 /Nm 3 , 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2,<br />

15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.<br />

256


Figure 11.6 presents the simulated mercury removal efficiency <strong>by</strong> a fabric<br />

filter without cleaning of the bags as a function of filtration time and injection rate of<br />

carbon. Every minute a layer of carbon is added to the bag surface. As expected,<br />

larger mercury removal efficiency is obtained with higher carbon injection rate. The<br />

mercury removal efficiency increases fast with time after initiating carbon injection<br />

up to 600 s and then it slowly increases with filtration time.<br />

Figure 11.6. Simulated mercury removal efficiency <strong>by</strong> a fabric filter without cleaning<br />

of the bags as a function of filtration time and injection rate of carbon. Darco Hg<br />

carbon with a diameter of 16 µm is injected at 138�C to simulated cement kiln flue<br />

gas with 200 µg Hg 0 /Nm 3 , 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2,<br />

15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. �t= 1 min.<br />

11.4 Fabric filter model<br />

In reality the filter bags are periodically cleaned. Therefore the filter cake<br />

model needs to be extended with periodical cleaning of the bags in order to simulate<br />

mercury adsorption in the bag filter. Assuming the cleaning cycle interval is tclean,<br />

fraction of filter cleaned per cycle is fclean. After first tclean, the carbon cakes on the<br />

257


ag surface have different exposure time to mercury and the mercury removal<br />

efficiency is the average of the mercury removal efficiency in different sections of the<br />

filter. Table 11.3 illustrates the carbon cake life time in different sections of the fabric<br />

filter and the calculation of average mercury removal efficiency across the filter as a<br />

function of filtration time. Pulse duration is selected as 0.1 second. Symbol * means a<br />

pulse with 0.1/60 min and indicates that a fraction of the filter bags is cleaned.<br />

Table 11.3 Illustration of carbon cake lifetime for different filter sections due to<br />

periodic cleaning of bags. Here tclean=25 min; fclean=0.1.<br />

Filtration time Exposure time of Average mercury removal efficiency<br />

(min) different filter sections<br />

0 0 0<br />

0-25 10@[0 25] �<br />

25 10@[25]<br />

25* 1@[0], 9@[25]<br />

25*-50 1@[0 25], 9@[0 50]<br />

50 1@[25], 9@[50]<br />

50* 1@[0], 1@[25], 8@[50]<br />

50*-75 1@[0-25], 1@[0 50],<br />

8@[0 75]<br />

75 1@[25], 1@[50], 8@[75]<br />

75* 1@[0], 1@[25], 1@[50],<br />

7@[75]<br />

75*-100 1@[0 25], 1@[0 50],<br />

258<br />

�<br />

[0 25]<br />

[25]<br />

9<br />

[25]<br />

10 �<br />

1<br />

��[0 25] 9�[0<br />

50] �<br />

10 �<br />

1<br />

��[25] 9�[50]<br />

�<br />

10 �<br />

1<br />

�� [25] 8�<br />

[50] �<br />

10 �<br />

1<br />

��[0 25] ��[0 50] � 8�[0<br />

75] �<br />

10<br />

1<br />

��[25] ��[50] � 8�[75]<br />

�<br />

10<br />

1<br />

��[25] ��[50] � 7�[75]<br />

�<br />

10<br />

1<br />

�[025] ��[050] ��[075] � 7�[0100]<br />

10<br />

1<br />

�[25] ��[50] ��[75] � 7�[100]<br />

10<br />

1<br />

�[25] ��[50] ��[75] � 6�[100]<br />

10<br />

1 ��[025] � �[050] ��[075] ��[0100] ��<br />

�<br />

10 �<br />

�<br />

6�<br />

�<br />

� [0 125]<br />

�<br />

1 ��[25] ��[50] ��[75] ��[100] ��<br />

�<br />

10 � �<br />

6�<br />

�<br />

� [125]<br />

�<br />

1 ��[25] ��[50] ��[75] ��[100] ��<br />

�<br />

10 � �<br />

5�<br />

�<br />

� [125]<br />

�<br />

1@[ 0 75], 7@[0 100] � �<br />

100 1@[25], 1@[50],<br />

1@[75], 7@[100] � �<br />

100* 1@[0], 1@[25], 1@[50],<br />

1@[75], 6@[100] � �<br />

100*-125 1@[0 25], 1@[0 50],<br />

1@[0 75], 1@[0 100],<br />

6@[0 125]<br />

125 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

6@[125]<br />

125* 1@[0], 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

5@[125]


Filtration time Exposure time of<br />

(min) different filter sections<br />

125*-150 1@[0 25], 1@[0 50],<br />

1@[0 75], 1@[0 100],<br />

1@[0 125], 5@[0 150]<br />

150 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 5@[150]<br />

150* 1@[0], 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 4@[150]<br />

150*-175 1@[0 25], 1@[0 50],<br />

1@[0 75], 1@[0 100],<br />

1@[0 125], 1@[0 150],<br />

4@[0 175]<br />

175 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

4@[175]<br />

175* 1@[0],1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

3@[175]<br />

175*-200 1@[0 25], 1@[0 50],<br />

1@[0 75], 1@[0 100],<br />

1@[0 125], 1@[0 150],<br />

1@[0 175],3@[0 200]<br />

200 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

1@[175], 3@[200]<br />

200* 1@[0], 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

1@[175], 2@[200]<br />

200*-225 1@[0 25], 1@[0 50],<br />

1@[0 75], 1@[0 100],<br />

1@[0 125], 1@[0 150],<br />

1@[0 175], 1@[0 200],<br />

2@[0 225]<br />

225 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

1@[175], 1@[200],<br />

2@[225]<br />

225* 1@[0], 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

1@[175], 1@[200],<br />

1@[225]<br />

259<br />

Average mercury removal efficiency<br />

1 ��[0 �<br />

10 �<br />

��[0 25] ��[050] ��[0 125] � 5�[0<br />

150]<br />

75] ��[0100] ��<br />

�<br />

�<br />

�<br />

1 ��[25] ��[50] ��[75] ��[100] ��<br />

�<br />

10 � �<br />

�[125] 5�<br />

�<br />

� � [150]<br />

�<br />

1 ��[25] ��[50] ��[75] ��[100] ��<br />

�<br />

10 � �<br />

�[125] 4�<br />

�<br />

� � [150]<br />

�<br />

1 ��[0 �<br />

10 �<br />

��[0 25] ��[050] ��[075] ��[0100] ��<br />

�<br />

125] ��[0 150] �4�<br />

�<br />

[0 175] �<br />

1 �� 10<br />

�� �� �� ��<br />

[25] [50] [75] [100]<br />

�� �<br />

�[125] �[150] 4�<br />

�<br />

� � � [175] �<br />

1 �� 10<br />

�� �� �� ��<br />

[25] [50] [75] [100]<br />

�� �<br />

�[125] �[150] 3�<br />

�<br />

� � � [175] �<br />

1 �� 10<br />

�� �� �� � �<br />

[0 25] [0 50] [0 75] [0 100]<br />

�<br />

�<br />

�<br />

�[0 125] �[0 150] �[0 175] 3�<br />

�<br />

� � � � [0 200] �<br />

1 �� 10<br />

�� �� �� � �<br />

[25] [50] [75] [100]<br />

�� �<br />

�[125] �[150] �[175] 3�<br />

�<br />

� � � � [200] �<br />

1 �� 10<br />

�� �� �� � �<br />

[25] [50] [75] [100]<br />

�� �<br />

�[125] �[150] �[175] 2�<br />

�<br />

� � � � [200] �<br />

��[025] ��[050] ��[075] ��[0100] � �<br />

1 � �<br />

�� �� �� �� ��<br />

10<br />

2<br />

[0 125] [0 150] [0 175] [0 200]<br />

� �<br />

� �[0<br />

225]<br />

�<br />

�� �� �� �� � �<br />

1<br />

10<br />

2<br />

[25] [50] [75] [100]<br />

� �<br />

��[125] ��[150] ��[175] ��[200] ��<br />

�<br />

�<br />

�<br />

� [225]<br />

�<br />

�� �� �� �� � �<br />

1<br />

10<br />

[25] [50] [75] [100]<br />

� �<br />

��[125] ��[150] ��[175] ��[200] ��<br />

�<br />

�<br />

�<br />

� [225]<br />


225*-250 1@[0 25], 1@[0 50],<br />

1@[0 75], 1@[0 100],<br />

1@[0 125], 1@[0 150],<br />

1@[0 175], 1@[0 200],<br />

1@[0 225], 1@[0 250]<br />

��[025] ��[050] ��[075] ��[0100] � �<br />

1 � �<br />

��[0 125] ��[0 150] ��[0 175] ��[0 200] ��<br />

10 � �<br />

��[0225] ��[0250]<br />

�<br />

250 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

1@[175], 1@[200],<br />

1@[225], 1@[250]<br />

��[25] ��[50] ��[75] ��[100] � �<br />

1 � �<br />

��[125] ��[150] ��[175] ��[200] ��<br />

10 �<br />

�[225] �<br />

�<br />

� � [250]<br />

�<br />

250* 1@[0], 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

1@[175], 1@[200],<br />

1@[225]<br />

��[25] ��[50] ��[75] ��[100] � �<br />

1 � �<br />

��[125] ��[150] ��[175] ��[200] ��<br />

10 �<br />

�<br />

�<br />

� [225]<br />

�<br />

250*-275 1@[0 25], 1@[0 50],<br />

1@[0 75], 1@[0 100],<br />

1@[0 125], 1@[0 150],<br />

1@[0 175], 1@[0 200],<br />

1@[0 225], 1@[0 250]<br />

��[025] ��[050] ��[075] ��[0100] � �<br />

1 � �<br />

��[0 125] ��[0 150] ��[0 175] ��[0 200] ��<br />

10 � �<br />

��[0225] ��[0250]<br />

�<br />

275 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

1@[175], 1@[200],<br />

1@[225], 1@[250]<br />

��[25] ��[50] ��[75] ��[100] � �<br />

1 � �<br />

��[125] ��[150] ��[175] ��[200] ��<br />

10 �<br />

�[225] �<br />

�<br />

� � [250]<br />

�<br />

275* 1@[0], 1@[25], 1@[50],<br />

1@[75], 1@[100],<br />

1@[125], 1@[150],<br />

1@[175], 1@[200],<br />

1@[225]<br />

��[25] ��[50] ��[75] ��[100] � �<br />

1 � �<br />

��[125] ��[150] ��[175] ��[200] ��<br />

10 �<br />

�<br />

�<br />

� [225]<br />

�<br />

… … …<br />

The filter cake model is run for time interval of [0 250] min. The calculated<br />

mercury removal efficiencies at different time are used to calculate the corresponding<br />

average mercury removal efficiency across the whole fabric filter.<br />

The input parameters <strong>from</strong> Durkee slipstream tests listed in table 11.2 are<br />

again used as model inputs to the fabric filter model. Other inputs include a bag<br />

cleaning interval of 25 min and a cleaning fraction of 0.1. It is assumed that a new<br />

sorbent layer is accumulated on the filter bag every 5 min.<br />

Figure 11.7 shows the simulated mercury removal efficiency if the filter was<br />

running without periodical cleaning up to 4 h. Compared to the short filtration time of<br />

25 min as shown in figure 11.5, the mercury removal efficiency reaches a stable value<br />

after about 1 h for the applied injection rates of powdered activated carbon. This<br />

260


ehavior is due to the growing thickness of the carbon cake. Fresh carbon is<br />

continuously injected to the filter, providing increased mercury adsorption. At long<br />

times, the inner layers of the carbon cake, consisting of almost fully spent carbon,<br />

gives negligible contribution to the process so that asymptotic conditions are reached.<br />

Figure 11.7. Simulated mercury removal efficiency <strong>by</strong> 1/10 of the fabric filter without<br />

cleaning of the bags as a function of filtration time and injection rate of carbon. Darco<br />

Hg carbon with a diameter of 16 µm is injected at 138�C to simulated cement kiln<br />

flue gas with 200 µg Hg 0 /Nm 3 , 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv<br />

SO2, 15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. �t=5 min, Air to cloth ratio<br />

ufilter= 1.2 m/min.<br />

Figure 11.8 presents the simulated mercury removal efficiency across the<br />

fabric filter with periodical cleaning 10% of the bags every 25 min. When 10% of the<br />

bags is cleaned the mercury removal efficiency decreases. At the beginning the<br />

mercury removal efficiency decreases slightly and it decreases more at later stage.<br />

This is due to the fact that more carbon is collected on the filter bag and is removed<br />

<strong>by</strong> pulse cleaning. The model assumes that all the carbon collected on the bag is<br />

completely removed <strong>from</strong> the bag surface and the corresponding mercury removal<br />

efficiency for this fraction of bags drops to zero when the pulse cleaning is initiated.<br />

The mercury removal efficiency across the whole filter reaches a stable level after all<br />

261


the bags have been cleaned once. The less smooth curve is due to the applied time<br />

interval of 5 min for a new carbon layer addition. Only five data points are used in a<br />

cleaning interval of 25 min. This can be easily improved <strong>by</strong> decreasing the time<br />

interval of new carbon cake layer addition at the expense of longer computation time.<br />

Figure 11.8. Simulated mercury removal efficiency of the fabric filter with cleaning<br />

of the bags as a function of filtration time and carbon injection rate. Darco Hg carbon<br />

with a diameter of 16 µm is injected at 138�C to simulated cement kiln flue gas with<br />

200 µg Hg 0 /Nm 3 , 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15<br />

vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. tclean=25 min, �t=5 min, ufilter=1.2 m/min.<br />

Figure 11.9 shows the simulated effects of flue gas temperature on the<br />

mercury removal efficiency of the fabric filter. When the flue gas temperature is<br />

reduced <strong>from</strong> 138�C to 75�C an improvement of about 8% mercury removal<br />

efficiency is obtained. However, whether this improvement is economical needs to be<br />

compared with additional costs <strong>by</strong> cooling down the flue gas.<br />

262


Overall bag filter mercury removal efficiency, %<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 1 2 3 4 5 6 7<br />

Time (hour)<br />

263<br />

1: 75 degree C<br />

2: 100 degree C<br />

3: 115 degree C<br />

4: 138 degree C<br />

Figure 11.9. Simulated effects of flue gas temperature on mercury removal efficiency<br />

of the fabric filter. 16 mg/m 3 Darco Hg carbon with a diameter of 16 µm is injected to<br />

simulated cement kiln flue gas with 200 µg Hg 0 /Nm 3 , 1000 ppmv NO, 5 ppmv NO2,<br />

10 ppmv HCl, 5 ppmv SO2, 15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. tclean=25<br />

min, �t=5 min, ufilter= 1.2 m/min.<br />

11.5 Two­stage model<br />

Models developed in previous sections simulate mercury removal <strong>by</strong> separate<br />

parts of the full-scale process. In this section mercury adsorption in the duct section is<br />

coupled with the fabric filter section. First the duct model is run. The mercury<br />

concentrations inside the particle and the bulk mercury concentration at the end of the<br />

duct are used as initial conditions for the fabric filter model. Then the fabric filter<br />

model is run to desired filtration time. The model inputs are the same as the fabric<br />

filter model. A flight time of 1 s in the duct is applied.<br />

Figure 11.10 shows the simulated mercury removal efficiency in the duct<br />

section at Durkee cement plant. When a smaller carbon injection rate is applied the<br />

mercury removal efficiency after 1 s in the duct is negligible. As shown in figure<br />

1<br />

2<br />

3<br />

4


11.10, about 2% mercury removal is obtained when 16 mg/m 3 Darco Hg carbon is<br />

injected. However, at larger carbon injection rates the mercury removal efficiency<br />

after 1 s in the duct is noticeable. Therefore, high mercury removal efficiency can be<br />

obtained <strong>by</strong> increasing the residence time of carbon particles in the duct, i.e., <strong>by</strong><br />

applying long duct, provided that there is enough space in the plant and large carbon<br />

injection rate is applied.<br />

Figure 11.10. Simulated mercury removal efficiency in the duct as a function of<br />

residence time and injection rate of carbon. Darco Hg carbon with a diameter of 16<br />

µm is injected at 138�C to simulated cement kiln flue gas with 200 µg Hg 0 /Nm 3 ,<br />

1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15 vol.% H2O, 21 vol.%<br />

CO2, and 6 vol.% O2.<br />

Figure 11.11 compares the simulated and measured mercury removal<br />

efficiency of the fabric filter at Durkee slipstream plant. Generally there is good<br />

agreement between the simulation and pilot-scale data. This indicates that the<br />

adsorption kinetics derived <strong>from</strong> 10 mg Darco Hg carbon in 2 g sand is reasonable<br />

and the developed model is a useful tool to simulate and optimize the carbon injection<br />

process.<br />

264


<strong>Mercury</strong> removal over the filter (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 10 20 30 40 50 60 70 80 90 100<br />

Activated carbon injection rate (mg/m 3 )<br />

265<br />

Model<br />

Data<br />

Figure 11.11. Comparison of simulated and measured mercury removal efficiency of<br />

the fabric filter at Durkee slipstream plant. Darco Hg carbon with a diameter of 16<br />

µm is injected at 138�C to simulated cement kiln flue gas with 200 µg Hg 0 /Nm 3 ,<br />

1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15 vol.% H2O, 21 vol.%<br />

CO2, and 6 vol.% O2. tclean=25 min, �t=5 min, ufilter= 1.2 m/min.<br />

The overall mercury removal efficiency of the sorbent injection system refers<br />

to mercury removal <strong>from</strong> the carbon injection point to the fabric filter outlet and can<br />

be evaluated as following:<br />

� �100 �(1 �(1 �� %)(1 � � %))<br />

(11.33)<br />

total duct filter<br />

Table 11.4 summarizes the calculated mercury removal efficiencies in the duct, fabric<br />

filter and the whole carbon injection system for different carbon injection rates. The<br />

contribution of mercury removal in the duct is much smaller to the mercury removal<br />

in the whole carbon injection system. However, data regarding overall mercury<br />

removal for the Durkee slipstream plant are not available for comparison.


Table 11.4 Simulated mercury removal efficiencies in the duct, fabric filter and the<br />

whole carbon injection system for different carbon injection rates. Darco Hg carbon<br />

with a diameter of 16 µm is injected at 138�C to simulated cement kiln flue gas with<br />

200 µg Hg 0 /Nm 3 , 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15<br />

vol.% H2O, 21 vol.% CO2, and 6 vol.% O2. Carbon particle residence time in the<br />

duct is 1 s. Every five minute a layer of carbon is added to the bag surface. Air to<br />

cloth ration is 1.2 m/min.<br />

Carbon load<br />

(mg/m 3 <strong>Mercury</strong> removal <strong>Mercury</strong> removal in Total <strong>Mercury</strong><br />

) in duct (%) fabric filter (%) removal (%)<br />

8 1.0 54.0 54.5<br />

16 2.0 69.6 70.2<br />

48 5.9 86.0 86.9<br />

80 9.7 90.3 91.2<br />

The applied carbon injection rate for mercury control is much smaller than the<br />

typical dust load in the flue gas for particulate emission control process. Therefore,<br />

the pressure drop over the fabric filter is expected to increase slowly with filtration<br />

time for mercury control process. It is then feasible to extend the bag cleaning<br />

interval, i.e., use less frequent cleaning of the bags. Figure 11.12 illustrates the<br />

simulated effects of bag cleaning frequency on the mercury removal efficiency of the<br />

fabric filter. The mercury removal efficiency slightly increases when the bag cleaning<br />

interval is increased. Extending the bag cleaning interval <strong>from</strong> 25 min to 100 min<br />

results in a 1.3% improvement of mercury removal efficiency. A longer bag cleaning<br />

cycle results in longer retention time of the carbon particles on the bags, which allows<br />

the carbon particles to adsorb more mercury <strong>from</strong> the flue gas.<br />

266


Hg removal efficiency, %<br />

88<br />

87.6<br />

87.2<br />

86.8<br />

86.4<br />

86<br />

25 50 75 100 125<br />

Bag cleaning interval (min)<br />

Figure 11.12. Simulated effects of bag cleaning frequency on mercury removal<br />

efficiency of the fabric filter. 48 mg/m 3 Darco Hg carbon with a diameter of 16 µm is<br />

injected at 138�C to simulated cement kiln flue gas with 200 µg Hg 0 /Nm 3 , 1000<br />

ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2, 15 vol.% H2O, 21 vol.% CO2,<br />

and 6 vol.% O2. tclean=25 min, �t=5 min, ufilter= 1.2 m/min.<br />

The only carbon injection system for mercury control <strong>from</strong> cement production<br />

is operated at Ash Grove’s Durkee cement plant [10]. Instead of continuous injection<br />

of powdered activated carbon, the fabric filter works in fixed-bed adsorber mode.<br />

During the first run period, the activated carbon was added to the system through the<br />

first day and removed <strong>from</strong> the system through the eighth day. The average removal<br />

efficiency during those intervening 6 days was 92.8% [10]. However, the carbon<br />

injection rate and duration in the first day is not reported.<br />

Simulations are performed to simulate the fabric filter fixed-bed operation<br />

mode. Firstly activated carbon is injected at high load for 30 min to form a carbon<br />

cake on the bags. Then the activated carbon injection is stopped and the fabric filter<br />

works as a fixed-bed adsorber. When the initial mercury breakthrough occurs or the<br />

mercury emission limit is reached, the bags are cleaned <strong>by</strong> pulse-jet compressed air<br />

and later activated carbon is injected again. The injection-adsorption-cleaning cycle is<br />

repeated. The carbon injection without bag cleaning period is simulated <strong>by</strong> the filter<br />

cake model and the fixed-bed adsorption period is simulated <strong>by</strong> the fixed-bed model.<br />

267


Figure 11.13 shows the simulated mercury breakthrough curves of the fabric<br />

filter injected with 0.5-2.0 g/m 3 Darco Hg carbon for 30 min. The actual carbon<br />

injection rate at full-scale test is not available and high carbon injection rates are<br />

tested here to show the effect. The initial mercury breakthrough time for carbon load<br />

of 0.5, 1.0, 1.5, and 2.0 g/m 3 is 17.9, 35.7, 53.6, and 71.5 h, respectively. This means<br />

that about 100% mercury removal efficiency is obtained within the initial mercury<br />

breakthrough periods. With an activated carbon injection of 1-2 g/m 3 for 30 min, the<br />

fabric filter can work for 53.6 to 71.5 h before the initial breakthrough occurs. This<br />

means that the activated carbon injection is only required 2-3 times per week. This is<br />

in agreement with the information obtained <strong>from</strong> Paone [9] on practice of the Durkee<br />

sorbent injection plant. However, the actual carbon injection rate and duration at<br />

Durkee plant are unknown and are required for validation of the simulation.<br />

Gaseous mercury outlet (Cout/Cin)<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

A<br />

B C D<br />

0<br />

0 50 100 150 200 250 300 350 400<br />

Time (hour)<br />

268<br />

A: 0.5 g/m3<br />

B: 1.0 g/m3<br />

C: 1.5 g/m3<br />

D: 2.0 g/m3<br />

Figure 11.13. Simulated mercury breakthrough curves of fabric filter injected with<br />

large carbon loads for 30 min and then in fixed-bed operation mode. Darco Hg<br />

carbon with a diameter of 16 µm is injected at 138�C to simulated cement kiln flue<br />

gas with 200 µg Hg 0 /Nm 3 , 1000 ppmv NO, 5 ppmv NO2, 10 ppmv HCl, 5 ppmv SO2,<br />

15 vol.% H2O, 21 vol.% CO2, and 6 vol.% O2.


To compare with mercury removal <strong>by</strong> continuous injection of carbon,<br />

simulations are conducted <strong>by</strong> injection of the same total amount of carbon within the<br />

initial mercury breakthrough period. The corresponding carbon injection rate is 14<br />

mg/m 3 and the simulated mercury removal efficiency is only 66.8%. The fabric filter<br />

works more efficiently for mercury removal in fixed-bed operation model than<br />

continuous injection of carbon for the same total amount of carbon injected. This is<br />

probably due to the larger amount of carbon accumulated on the bags in the fixed-bed<br />

operation mode. The fixed-bed operation model is limited <strong>by</strong> the pressure drop over<br />

the fabric filter. It is expected that a larger pressure drop over the filter and power<br />

consumption of the system is required when a large carbon injection rate is applied in<br />

a reasonable period.<br />

11.6 Conclusions<br />

The developed single particle and fixed-bed adsorption models are further<br />

extended to duct and fabric filter models to simulate mercury removal <strong>by</strong> carbon<br />

injection upstream of a fabric filter. The fabric filter model is accounted for <strong>by</strong> adding<br />

a new carbon layer on the bag surface after a short time and treating each layer as a<br />

well mixed tank. Finally the duct model and fabric filter model are coupled to a twostage<br />

model. The mercury concentrations inside the particle and the bulk mercury<br />

concentration at the end of the duct are used as initial conditions for the fabric filter<br />

model. The models are based on materials balances in both gaseous and adsorbed<br />

phase along the duct length/growing filter cake and inside the carbon particles. The<br />

models account for adsorption kinetics, both the external and internal mass transfer<br />

resistances, accumulation of carbon layer on the bags, and periodical cleaning of the<br />

bags.<br />

Henry’s constant obtained <strong>from</strong> fixed-bed investigation of mercury adsorption<br />

<strong>by</strong> activated carbon in the simulated cement kiln flue gas is used as input to the<br />

models. The effects of SO2, H2O, NO2 levels in the flue gas on mercury removal are<br />

accounted <strong>by</strong> using correlations derived <strong>from</strong> the fixed-bed investigation.<br />

Duct model simulations indicate that large carbon loading in the flue gas are<br />

required to obtain high mercury removal efficiency due to the short residence time.<br />

269


To minimize the carbon feed rate it is advisable to lower the operating temperature.<br />

Improvements in the mercury removal efficiency can be obtained also <strong>by</strong> increasing<br />

the in-duct particle residence time and decreasing the carbon particle size.<br />

In contrast to the in-duct removal process, simulations of mercury adsorption<br />

in the fabric filter show that higher mercury removal efficiency can be achieved with<br />

moderate carbon consumption due to the effective gas/carbon contact on the filter<br />

bags. The effects of carbon load, temperature, frequency of new carbon layer addition<br />

and bag cleaning on mercury removal efficiency are simulated. The fabric filter<br />

model can predict the mercury removal profile with jagged nature because of the<br />

intermittent partial cleaning of the bags. Comparison with simulation and<br />

experimental data <strong>from</strong> Durkee cement plant slipstream tests shows that the<br />

developed two-stage model can reasonably predict the mercury removal <strong>from</strong> cement<br />

plants <strong>by</strong> carbon injection upstream of a fabric filter.<br />

Minor benefits can be obtained <strong>by</strong> increasing the cleaning cycle time of the<br />

fabric filter compartments. The fabric filter works more efficiently on mercury<br />

removal when it is operated as fixed-bed adsorbed <strong>by</strong> injection of high carbon load in<br />

short time and then stopping carbon injection and cleaning of the bags.<br />

11.7 List of symbols<br />

A: total exterior surface area of all particles in the adsorber (m 2 )<br />

A: matrix in equation 11.18<br />

B: matrix in equation 11.19<br />

Bi: dimensionless Biot number<br />

C: gaseous mercury concentration (µg/m 3 )<br />

C: matrix in equation 11.20 and 11.21<br />

Cb: gas bulk mercury concentration (µg/m 3 )<br />

Cbo: initial gas bulk mercury concentration (µg/m 3 )<br />

Cbin: inlet mercury concentration in tank (µg/m 3 )<br />

Cµ: adsorbed mercury concentration in the sorbent (µg/m 3 )<br />

Dapp: apparent diffusion coefficient (m 2 /s)<br />

De: effective diffusion coefficient (m 2 /s)<br />

fclean: fraction of bags cleaned per pulse cleaning<br />

J: mercury flux (µg/m 2 )<br />

K: Henry’s constant<br />

270


km: gas film mass transfer coefficient (m/s)<br />

L: thickness of carbon cake (m)<br />

mp: mass of carbon particle in the adsorber (g)<br />

M: carbon load in the tank (mg)<br />

n: number of interior collocation points<br />

r: radial coordinate (m)<br />

Rp: carbon particle radius (m)<br />

t: time (s)<br />

tclean: time interval for bag cleaning (25)<br />

u: dimensionless parameter, u=x 2<br />

u: face velocity on the filter bags (m/s)<br />

ufilter: air to cloth ratio (m/min)<br />

V: volume of the adsorber (m 3 )<br />

Vb: volume of the adsorber (m 3 )<br />

x: dimensionless radius<br />

y: dimensionless mercury concentration<br />

yb: dimensionless gas bulk mercury concentration (µg/m 3 )<br />

ybin: dimensionless inlet mercury concentration in tank (µg/m 3 )<br />

Greek symbols<br />

�b: bed void fraction<br />

�p: carbon particle porosity<br />

�p: carbon particle density (kg/m 3 )<br />

�: dimensionless time<br />

�: carbon load in the flue gas (kg/m 3 )<br />

�: mercury removal efficiency (%)<br />

11.8 References<br />

[1] J.R.V. Flora, R.A. Hargis, W.J. O'Dowd, A. Karash, H.W. Pennline, R.D. Vidic, The role<br />

of pressure drop and flow redistribution on modeling mercury control using sorbent injection<br />

in baghouse filters, J. Air Waste Manage. Assoc. 56 (2006) 343-349.<br />

[2] F. Scala, Simulation of mercury capture <strong>by</strong> activated carbon injection in incinerator flue<br />

gas. 1. In-duct removal, Environ. Sci. Technol. 35 (2001) 4367-4372.<br />

[3] F. Scala, Simulation of mercury capture <strong>by</strong> activated carbon injection in incinerator flue<br />

gas. 2. Fabric filter removal, Environ. Sci. Technol. 35 (2001) 4373-4378.<br />

[4] F. Scala, Modeling mercury capture in coal-fired power plant flue gas, Ind Eng Chem Res.<br />

43 (2004) 2575-2589.<br />

[5] J. Villadsen, M.L. Michelsen, Solutions of differential equation models <strong>by</strong> polynomial<br />

approximation, Prentice-Hall, Inc., 1978.<br />

271


[6] D.D. Do, Adsorption analysis: equilibria and kinetics, Imperial College Press, 1998.<br />

[7] L. Hayes-Gorman, Regulating mercury emissions: Ash Grove <strong>Cement</strong> in Durkee, Air<br />

toxics summit 2008, Boise, Idaho, 4-7 August, 2008.<br />

[8] Schreiber & Yonley Associates, <strong>Mercury</strong> emissions test report, Ash Grove <strong>Cement</strong><br />

Company Durkee, Oregon, Project No. 060204, 2007.<br />

[9] P. Paone, Personal communication about flue gas compositions and temperature for pilotscale<br />

sorbent injection tests at Ash Grove Durkee plant and FLSmidth Mineral Lab, 2010.<br />

[10] Curtis D. Lesslie, Mail to U.S.EPA about initial results of Ash Grove's Durkee sorbent<br />

injection system, http://www.whitehouse.gov/sites/default/files/omb/assets/ oira_2060/2060_<br />

07292010-3.pdf, visited March 21, 2011.<br />

272


273<br />

12<br />

Concluding remarks<br />

To develop and get a better understanding of mercury removal <strong>from</strong> cement<br />

plant <strong>by</strong> sorbent injection upstream of a pulse jet fabric filter, this project has focused<br />

on four areas: comprehensive review of mercury emission <strong>from</strong> cement plants and<br />

analysis applicability of available technologies for mercury removal <strong>from</strong> cement<br />

plants, test and development of thermal catalytic converters for oxidized mercury<br />

reduction and dynamic total mercury measurement, screening tests and fundamental<br />

investigation of mercury adsorption <strong>by</strong> sorbents in simulated cement kiln flue gas,<br />

and development of mathematic models that can describe mercury removal <strong>by</strong> fixedbed<br />

and carbon injection upstream of a fabric filter.<br />

<strong>Cement</strong> plants are quite different <strong>from</strong> power plants and waste incinerators<br />

regarding the flue gas composition, temperature, residence time, and inherent<br />

material circulation. Instead of fuel, cement raw materials are the dominant sources of<br />

mercury in the cement kiln flue gas. The mercury emissions and speciation <strong>from</strong><br />

cement kilns can vary over time and depend on raw materials and fuels used, and<br />

process operation. Among the available technologies for mercury removal <strong>from</strong> flue<br />

gas, sorbent injection upstream of a polishing fabric filter is considered as the most<br />

promising and suitable technology for cement plant application.<br />

To be able to perform dynamic measurement of mercury adsorption <strong>by</strong><br />

sorbents, a red brass chips and sulfite converter were investigated in simulated<br />

cement kiln flue gas. The red brass converter works only when measuring elemental<br />

mercury in nitrogen and does not work properly even when only elemental mercury is<br />

added to the simulated flue gas. The red brass converter cannot fully reduce HgCl2 to<br />

elemental mercury under any relevant condition.<br />

The sodium sulfite converter material was prepared <strong>by</strong> dry impregnation of<br />

sodium sulfite and calcium sulfate powders on zeolite pellets using water glass as


inder. The sulfite converter works well at 500�C when less than 10 ppmv HCl is<br />

present in the simulated cement kiln flue gas. The response time of the sulfite<br />

converter is short and typically within at most two minutes, which makes it<br />

appropriate for not too fast dynamic measurements.<br />

Inconsistent mercury adsorption capacity of activated carbon is observed at<br />

different carbon loads in 2 g sand. A smaller mercury adsorption capacity is obtained<br />

with larger carbon load. Tests with elemental mercury and mercury chloride, different<br />

carbon type and particle sizes show the same trend. Effects of bed dilution on the<br />

equilibrium mercury adsorption capacity appear to be limited.<br />

Screening tests of sorbents for mercury removal <strong>from</strong> cement plants have<br />

been conducted in the fixed-bed reactor system using simulated cement kiln flue gas<br />

with elemental mercury and mercury chloride source. The tested sorbents include<br />

commercial activated carbons, commercial non-carbon sorbents, and cement<br />

materials. Screening measurements are used to evaluate initial mercury capture rate,<br />

oxidation potential, and capacity for the selected sorbents.<br />

The sorbents don’t adsorb any mercury when tested with elemental mercury<br />

in nitrogen. Tests of a range of 30 mg collected non-carbon based sorbents and<br />

cement materials as sorbents in 2 g sand at 150�C in simulated cement kiln flue gas<br />

with elemental mercury do not show any mercury adsorption or oxidation. Generally<br />

a larger amount of adsorbed mercury is obtained with sorbents that have larger<br />

mercury oxidation capacity. While all the non-carbon based sorbents and cement<br />

materials show some adsorption of mercury chloride. This indicates that mercury<br />

oxidation is an important factor for mercury adsorption <strong>by</strong> the sorbents. Elemental<br />

mercury needs to be oxidized either in the flue gas with HCl or on the sorbent.<br />

Among the tested sorbents the Darco Hg activated shows the best performance of<br />

adsorption of both elemental and oxidized mercury and is recommended as the<br />

reference sorbent for fundamental investigation.<br />

A parametric study of elemental mercury adsorption <strong>by</strong> activated carbon has<br />

been conducted in the fixed-bed reactor <strong>by</strong> mixing 10 mg Darco Hg carbon with 2 g<br />

sand. Increasing adsorption temperature results in decreased equilibrium mercury<br />

adsorption capacity of the activated carbon. The mercury adsorption isotherm follows<br />

274


Henry’s law for the applied mercury inlet levels in this project. The derived heat of<br />

adsorption is -8540 J/mol for elemental mercury adsorption <strong>by</strong> Darco Hg activated<br />

carbon in simulated cement kiln flue gas. Higher mercury oxidation and initial<br />

adsorption rate are also observed for smaller carbon particles, while the equilibrium<br />

mercury adsorption capacity is the same.<br />

The mercury adsorption capacity does not change with O2, CO, and NO levels<br />

in the flue gas, but decreases when CO2, H2O, SO2, and NO2 concentrations increase.<br />

The decrease of mercury adsorption capacity is due to the competition for active site<br />

with mercury <strong>by</strong> CO2 and H2O, and conversion of the previously formed nonvolatile<br />

basic mercuric nitrate into the volatile form <strong>by</strong> interactions between SO2 and NO2.<br />

Slight promoting effects of HCl on mercury adsorption are observed when HCl<br />

concentration is varied in the range of 0.5-20 ppmv. Larger mercury adsorption<br />

capacity is obtained when HCl is removed <strong>from</strong> baseline gas because HgO(s) is<br />

formed on the carbon.<br />

Similar adsorption behaviors of mercury chloride and elemental mercury <strong>by</strong><br />

Darco Hg activated carbon are observed using simulated cement kiln flue gas. This is<br />

due to the effective catalytic oxidation of elemental mercury <strong>by</strong> the activated carbon.<br />

Mathematical models are developed to simulate mercury adsorption <strong>by</strong> a<br />

single carbon particle, fixed carbon bed, in the duct and fabric filter. Orthogonal<br />

collocation method is used to solve mercury diffusion and adsorption inside a carbon<br />

particle. The fixed-bed model is solved <strong>by</strong> tank-in-series method. The fabric filter<br />

model is accounted for <strong>by</strong> adding a new carbon layer on the bag surface after a short<br />

time as a well mixed tank. The two-stage duct-fabric filter model accounts for<br />

adsorption kinetics, both the external and internal mass transfer resistances,<br />

accumulation of carbon layer on the bags, and periodical cleaning of the bags.<br />

Henry’s constant obtained <strong>from</strong> fixed-bed investigation are used as input to<br />

the models. The developed fixed bed model can reasonably simulate the effects of<br />

adsorption temperature, mercury inlet concentration, flow gas rate, carbon particle<br />

size, and SO2, H2O, NO2 level in the flue gas on the mercury breakthrough curve of<br />

the fixed carbon bed.<br />

275


Duct model simulations indicate that a large carbon load is required to obtain<br />

a high mercury removal efficiency due to the short residence time. Simulations of<br />

mercury adsorption in the fabric filter show that higher mercury removal efficiency<br />

can be achieved with moderate carbon consumption due to the effective gas/carbon<br />

contact on the filter bags. The effects of carbon load, temperature, frequency of new<br />

carbon layer addition and bag cleaning on mercury removal efficiency are simulated.<br />

The fabric filter model can predict the mercury removal profile with jagged nature<br />

because of the intermittent partial cleaning of the bags. Comparison with simulation<br />

and experimental data <strong>from</strong> Durkee cement plant slipstream tests shows that the<br />

developed two-stage model can reasonably predict the mercury removal <strong>from</strong> cement<br />

plants <strong>by</strong> carbon injection upstream of a fabric filter.<br />

276


277<br />

13<br />

Suggestions for further work<br />

This work has investigated the mercury removal <strong>by</strong> carbon injection upstream<br />

of a fabric filter under more controlled conditions using a fixed bed reactor. Pilot or<br />

full-scale tests are desired to demonstrate the ability of the studied sorbents and<br />

technology to control emissions of mercury <strong>from</strong> cement plant over a typical range of<br />

operating conditions for an extended period of time and to further validate the<br />

developed models. The condition in full-scale application is much more demanding<br />

than in the lab-scale investigation. Further development and test of the sulfite<br />

converter is required for dynamic measurement of mercury in large scale<br />

investigation. A sampling probe is needed to separate the particles <strong>from</strong> the flue gas<br />

efficiently without plugging. Adsorption of mercury <strong>by</strong> the dust and probe should be<br />

minimized <strong>by</strong> high sampling flow rate and high heating temperature.<br />

The problem of inconsistent mercury adsorption capacity for different carbon<br />

loads could not be solved within the project. More thorough investigation is<br />

necessary to reveal the cause. New analysis technology is required to reveal whether<br />

mercury is adsorbed <strong>by</strong> the sand when it is mixed with activated carbon.<br />

This project investigates only mercury removal <strong>by</strong> the activated carbon. In the<br />

future multipollutants control <strong>by</strong> the activated carbon should be studied <strong>by</strong> measuring<br />

also other harmful species such as SO2 and NOx. When more than one component is<br />

involved in the adsorption system, adsorption equilibrium involving competition<br />

between molecules of different types is needed for the understanding of the system as<br />

well as for the design purposes.<br />

To reduce the sorbent cost, regeneration of used sorbents should be<br />

investigated. Recycling sorbent collected <strong>by</strong> the fabric filter to the injection process<br />

also requires more investigation. Modification of cement materials <strong>by</strong> additives that


can oxidize mercury is attractive. However, the influence of the additives on the<br />

cement quality needs to be investigated.<br />

Models developed in this work assume that all the particles have a uniform<br />

size. It is interesting to take the particle size distribution into account in the more<br />

advanced model. The developed fabric filter model does not include pressure drop<br />

over the filter. It is useful to incorporate the pressure development of the fabric filter<br />

and pulse jet cleaning instead of assuming a constant bag cleaning interval. Current<br />

models assume local equilibrium inside the carbon particle, simulation with a full<br />

kinetics description of the adsorption process is necessary to investigate whether this<br />

assumption is reasonable.<br />

278

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!