+ All Categories
Home > Documents > THE APPLICABILITY OF SODA ASH MANUFACTURING A THESIS …

THE APPLICABILITY OF SODA ASH MANUFACTURING A THESIS …

Date post: 18-Jan-2022
Category:
Upload: others
View: 11 times
Download: 2 times
Share this document with a friend
94
THE APPLICABILITY OF SODA ASH MANUFACTURING WASTES TO FLUE GAS DESULFURIZATION by JAMES EDWIN SCROGGINS, B.S. A THESIS IN CIVIL ENGINEERING Submitted to the Graduate Faculity of Texas Tech University in Particial Fulfillment of the Requirements for the Degree of MASTER OF SCIENCE IN CIVIL ENGINEERING May, 1984
Transcript

THE APPLICABILITY OF SODA ASH MANUFACTURING

WASTES TO FLUE GAS DESULFURIZATION

by

JAMES EDWIN SCROGGINS, B.S.

A THESIS

IN

CIVIL ENGINEERING

Submitted to the Graduate Faculity of Texas Tech University in Particial Fulfillment of

the Requirements for the Degree of

MASTER OF SCIENCE IN

CIVIL ENGINEERING

May, 1984

1 •

^>''" ACKNOWLEDGEMENTS

The research from which this report was derived was financed

by Texasgulf Chemicals Co. of Raliegh N. C. Sincere appreciation is

extended to this organization and their support is gratefully

acknowledged.

Sincere appreciation is expressed to Dr. R. H. Ramsey for his

guidance throughout the project, in addition to Dr. L.V. Urban, Dr.

B. J. Claborn, and Dr. R. M. Bethea for their helpful criticism.

n

TABLE OF CONTENTS

ACKNOWLEDGEMENTS i i

ABSTRACT v

LIST OF TABLES vi

LIST OF FIGURES vi 11

I. INTRODUCTION 1

II. LITERATURE REVIEW 5

Sulfur Dioxide Scrubbing Process 5

Rate of Contact 6

SOp Absorption 6

Scrubber Design * 10

Spray Tower 12

Sieve Tray 12

Corrosion Protection 13

Cost Information 16

Capital 18

Operations 19

Maintenance 20

Availability 20

III. EXPERIMENTAL PROCEDURES 22

Test Facil ities 22

Test Procedures 25

Test Series One 25

111

Test Series Two 30

Test Series Three 33

Test Series Four 34

Cost Data 36

Capital 36

Operating Costs 37

Maintenance 38

Aval 1 abl i ty 38

Cost Equation 39

IV. PRESENTATION OF DATA 40

Test Series One 40

Reagent Usage 40

SO^ Removal Efficiency 44

Corrosion 44

Seal ing 48

Test Series Two 50

Reagent Usage and SOp Removal Efficiency... 50

Scaling 54

Test Series Three 54

Test Series Four 59

Cost Comparisons 63

Capital Cost Estimates 63

Operating and Maintenance Costs 67

Availablity Costs 72

Limestone vs Sodium Alkali Comparison 73

IV

V. CONCLUSIONS AND RECOMMENDATIONS 78

LIST OF REFERENCES oi

V

ABSTRACT

The objectives of this study were two-fold: 1) to determine the

efficacy of low grade sodium alkalies in wet SOp scrubber

applications and 2) to use a simplified method for comparing the

costs of sodium alkali SOp scrubbing systems and limestone scrubbing

systems. The materials tested were waste streams produced in the

manufacturing of soda ash. Experiments were conducted on full scale

industrial SOp scrubbers to determine optium scrubber operating

conditions for each of the streams that were tested. All the

materials tested performed well in the test scrubber removing up to

97 per cent of the SOp from the gas stream. Crystallizer purge

liquor was found to cause silica scale in the test scrubber which

could be controlled by chemical treatment.

A standardized method of cost estimation was used to obtain

relative costs for sodium alkali and limestone scrubbers. The tests

which were conducted showed that capital, operating and maintenance

costs of sodium alkali scrubbers should be less than a limestone

scrubber of the same size and type. The availablity of the test

scrubber was higher than the availablity given in the literature for

a limestone scrubber. This information was then used to develop

generalized descriptions of SOp scrubbing applications where sodium

carbonate chemicals were more cost effective than limestone systems.

It was found that sodium alkalies could be competitive with

limestone in applications where the sulfur content of the coal is 3

per cent and the cost of the sodium alkali was less than $75/ton.

vi

LIST OF TABLES

Table Page

1. Corrosion Rates for Various Alloys Exposed to

Flue Gas Desulfurization Environment 14

2. Reported Capital and Annual Costs 17

3. Scrubber Scale Analysis 28

4. Evaporator Purge Liquor Analysis 30

5. Mine-water Analysis 33

6. Typical Analysis of Run-off 35 7. Test Series One, Test One: Reagent Usage and SOp

Removal Efficiency Results 41

8. Test Series One, Test Two: Reagent Usage and SOp Removal Efficiency Results 42

9. Test Series One, Test Three: Reagent Usage and SOp Removal Efficiency Results 43

10. Stoichiometric Ratio at Various pH Values 46

11. Comparison of Coupon Corrosion Weight and Rate Changes 47

12. Increases in Scrubber Back Pressure from Calcium Sulfite Scale Formation at 80% Boiler Load 49

13. Test Series Two, Test One: Reagent Usage and SOp Removal Efficiency Results 51

14. Test Series Two, Test Two: Reagent Usage and SOp Removal Efficiency Results 52

15. Test Series Two, Test Three: Reagent Usage and SOp Removal Efficiency Results 53

16. Incease in Scrubber Back-Pressure as a Result of Silica Scale Formation at 80% Boiler Load 55

v n

Table Page

17. Test Series Three, Test One: Reagent Usage and SO^ Removal Efficiency Results 56

18. Test Series Three, Test Two: Reagent Usage and SO^ Removal Efficiency Results. 1 57

19. Test Series Three, Test Three: Reagent Usage and SO^ Removal Efficiency Results 58

20. Test Series Four, Test One: Reagent Usage and SOp Removal Efficiency Results 7. 60

21. Test Series Four, Test Two: Removal Usage and SOp Removal Efficiency Results ;. 61

22. Test Series Four, Test Three: Removal Usage and Removal Efficiency Results 62

23. Limestone Scrubbing Process Capital Investment 64

24. Acid Soda Ash Scrubbing Process Capital Investment. 65

25. Basic Soda Ash Scrubbing Process Capital Investment 66

26. Limestone Scrubbing Process Annual Operating Costs Less Reagent Costs 68

27. Acid Soda Ash Scrubbing Process Annual Operating Costs Less Reagent Costs 69

28. Basic Soda Ash Scrubbing Process Annual Operating Costs Less Reagent Costs 70

29. Sodi um Al kal 1 Reagent Costs 71

30. Annualized Costs Less Reagent Cost 74

31. Total Annual Costs for a Limestone Scrubber 74

32. Competitive Sodium Alkali Reagent Costs 75

VT n

LIST OF FIGURES

Figure Page

1. Diagram of Test Scrubber 23

2. General Arrangement of Air Pollution Control Equipment 26

3. Scrubber Efficiency vs pH 45

4. Annual Costs as a Function of Percent Sulfur in Coal and Reaoent Costs 77

IX

CHAPTER I

INTRODUCTION

Most fossil fuels contain sulfur and produce sulfur oxides when

burned. Sulfur dioxide, the most abundant of the oxides, has been

found to cause morphological changes in the cells of the upper

respiratory tract of humans similar to the cellular changes caused

by cigarette smoking. It has also been determined that high ambient

SOp concentrations can cause the premature death of individuals with

weakened cardio-pulmonary systems [11.

In addition to adverse health effects, SOp can cause costly

corrosive damage to vegetation and manmade structures. Atmospheric

SOp combines with water and oxygen to form sulfuric acid which comes

in contact with the surface of the earth either as "dry deposition"

or "acid rain" [2]. Much attention has recently been given to the

effect of acid rain on oligotrophic mountain lakes. Because such

lakes are characteristically low in alkalinity, acid rains lower the

pH of the lake water below the point at which fish can survive. In

Canada near the Sudbury smelters the ph in 33 near by lakes have

dropped 100 fold [26]. Thousands of these lakes in the Adirondack

Mountains are being threatened by the effects of acid rains [3].

In an effort to protect the public health and to mitigate the

deleterious effects of acid rain, federal and state regulations have

been promulgated to limit the release of SOo. The federal standard

2

for new large coal-fired boilers has been set at 1.2 pounds SOp per

million British thermal units (MB) of heat input to the boiler [27],

Many western states, where low-sulfur coal is abundant and the

alpine aquatic habitat is highly prized, have enacted SOp regula­

tions which are much more stringent than the federal regulations.

For example, the state of Wyoming has enacted an emission limit of

0.2 lb SO^/MB of heat input [4]. While certainly providing

protection for the environment, this stringent standard poses a

sizeable technical and economic obstacle to the construction of any

large coal-burning facility in the state of Wyoming.

Currently, the best available control technology for SOp

removal is to contact the SOp laden gases with water containing

soluble alkali chemicals. This contact is achieved in various ways

in wet or dry SOp scrubbers [8]. To meet the Wyoming standard, a

power generation facility which uses western sub-bituminous coal

with a sulfur content of 1 per cent would require an SOo scrubber

that was 90 per cent efficient. It is often necessary to bypass a

portion of the stack gas to keep the temperature in the exit flue

above the dew point to prevent corrosion and stack rain-out. In

those situations, the SOp scrubber would have to be 95 per cent

efficient.

Conventional lime or limestone SOp scrubbers typically operate

at 80 per cent to 85 per cent efficiency [5]. As a result of the

low efficiencies obtained with these materials, some coal-fired

power plants, and industrial facilities, are equipped with SOp

scrubbers which use sodium carbonate (soda ash) as the absorbent.

These units have demonstrated removal efficiencies of 95 per cent

3

and greater [9]. Additionally, the delivered cost of soda ash in

Wyoming is relatively low because there are five soda ash

manufacturing facilities located in the southwestern part of the

state of Wyoming. In the processing of trona ore

(Na^CO^+NaHCO^'H^O) to refined soda ash, the ore is dissolved in

water and the insoluble impurities are separated from the sodium

carbonate liquor. This liquor is then fed to multi-stage

crystallizers which drive off water and re-crystalize the sodium

carbonate. The levels of soluble impurities in the crystal 1izers

which were introduced with the sodium carbonate liquor are

controlled by purging from the last crystallizer effect. One power

generating facility currently buys this purge liquor from a soda ash

manufacturer, which further reduces cost [6]. There is

approximately 1.3 million tons of this material produced each year

and 30 per cent by weight is sodium carbonate. In a raw state it is

only practical to use this material within a 100 mile radius of a

soda ash facility. The purge liquor has to remain heated to prevent

the precipitation of sodium carbonate and the freight penialty from

shipping 70 per cent water increases with distance. Trona ore or

process intermediates could be more cost effective over a wider

range of applications. As a result of requiring less auxiliary

equipment and a less corrosive scrubbing environment, soda ash

scrubbers could be less expensive than lime or limestone scrubbers

in some applications.

The purpose of this study is to determine 1) the efficacy of

various low grade sources of sodium alkalies which could be used as

SO^ absorbants and 2) the cost effectiveness of operating a sodium

alkali scrubber compared to a limestone scrubber. This information

is then used in a cost comparison of sodium alkali and

lime/limestone scrubbers for new installations using cost estimating

procedures developed by the Tennessee Valley Authority. All cost

analysis is based on the facilities needed for a 500 MW power

generation facility.

CHAPTER II

LITERATURE REVIEW

As an aid to accomplishing the objectives of this study, a

review of available literature was conducted. A search was first

made for information relative to SOp scrubbing processes. Specif­

ically, sources were sought that gave information about the opera­

ting parameters, chemistry and process layout of various SOp

scrubbing techniques. Design information for several SOp scrubbers

has been included in this review. Emphasis was given to those

design aspects which may be affected by operating the scrubber in

the alkaline pH range. Finally, cost information for several

existing scrubber operations was examined.

Sulfur Dioxide Scrubbing Process

At the present, the vast majority of SO^ scrubbers are "wet"

scrubbers [7]. Wet scrubbers circulate alkali liquors in a contact

chamber through which the SO^-laden flue gas is forced. The SOp

removal efficiency of a particular scrubber is primarily dependent

upon three factors:

1. Rate of contact of the flue gas with the recirculating

1iquor,

2. The dissolution rate and concentration of the

absorbent, and

3. Liquor saturation with respect to sulfite and sulfate

ions.

These factors will be discussed in the following sections.

Rate of Contact

The physical configuration of the interior of the contacting

vessel and liquor recirculation rate will determine the rate of

contact in the scrubber. The internals of the contacting vessel are

designed to induce intimate contact of the flue gas and the

scrubbing liquor. This is achieved primarily by increasing the

surface area of the scrubbing liquor and promoting turbulent flow.

Provided the concentration of alkali remains constant,

increasing the recirculation rate of a scrubber will increase the

rate of contact. Eventually the relationship becomes asymptotic as

other factors become rate limiting.

SOp Absorption

After the SOp comes into contact with the alkali liquor, it

dissolves and combines with oxygen to form sulfite or sulfate ions.

The chemistry and reaction products will vary with each absorbent.

Lime or Limestone

In a lime or limestone scrubber, the sulfite and sulfate ions

combine with calcium to form calcium sulfite and calcium sulfate,

which precipitate out of the solution and are blown down or purged

from the system [8].

H^O + CaO—»Ca(0H)2

Ca(0H)2 + CO2—>CaC03 + HpO

COp + HpO + CaC03—»Ca(HC03)2

SOp + Ca(HC03)p " ^ CaS03 + 200^ + HpO

CaS03 + iOp - ^ CaSO^

1-1)

1-2)

1-3)

1-4)

1-5)

The dissolution of CaC03 '" ^ ^ presence of water to calcium

and bicarbonate ions is a relatively slow process and therefore, it

is difficult to achieve a removal efficiency greater than 80 per

cent in lime/limestone scrubbers [9].

In an attempt to increase the solubility of the limestone in

the recirculation liquor and consequently the rate of reaction, a

scrubbing process has been developed which incorporates weak organic

acids (primarily adipic acid) [10]. Although removal efficiencies of

over 95 per cent have been reported from full scale field tests,

there are some serious drawbacks associated with this process. One

major deficiency is calcium sulfite "blinding." As a result of the

Increased absorption rate, the recirculation slurry becomes

super-saturated with calcium sulfite. As the excess calcium sulfite

begins to precipitate, the undissolved limestone particles in the

slurry act as seed crystals and become coated with calcium sulfite.

The result of this phenomenon is a drastic reduction in limestone

utilization. One study reports that reagent utilization will drop

to 47 per cent during these events [10].

Other shortcomings of this system are the objectionable odor of

adipic acid and increased heavy metal mobilization in the flue gas

desulfurization (FGD) wastes which are caused by the addition of

8

organic acids. As as result of these problems. Industry, for the

most part, has not accepted organic acid addition as a viable

alternative for flue gas desulfurization.

Soda Ash

The absorption reaction begins with SOp dissolving in the

soda ash liquor. Then the following reactions are thought

to occur [8]:

Na2C03 + SOp—^NapS03 ^ ^^2 ' " ^

Na2S03 + iOp-^^NapSO^ (1-7)

Na2S03 + SOp + HpO —>2NaHS03 (1-8)

2NaHS03 + Op—>2NaHS04 ' " ^

All of these reaction products can be found in solution and

their relative concentrations will vary with pH and the dissolved

oxygen content in the scrubbing solution.

Soda ash processes for SOp removal are characteristically

more efficient than limestone scrubbers due to the greater solubil­

ity and dissolution rate of soda ash in comparison with lime­

stone [9]. Most manufacturers of soda ash scrubbers guarantee 95

per cent removal efficiency [5].

The rate limiting factor in a soda ash scrubber is the rate of

contact of the SOp with the soda ash liquor. The efficiency of a

soda ash scrubber is, therefore, dependent upon both the scrubber

design and the concentration of the recirculation liquor.

Scrubber feed water often contains calcium ions. These

combine with SOp in the scrubber to form calcium sulfite which

precipitates to form a hard, tenacious scale which can plug spray

9

heads, sieve trays, mist eliminators, etc. Given the calcium

concentration in the make-up water and the solubility of calcium

sulfite at a given pH, it is possible to calculate the blow-down

rate from the system needed to prevent the formation of CaSO^. It

may also be feasible in some processes to use condensate or softened

water in the scrubber, thereby greatly reducing the potential for

calcium scaling [11].

Calcium sulfite scale also can be controlled with water treat­

ment chemicals. Several companies have proprietary chemicals which

cause distortions in the crystalline structure of calcium sulfite,

preventing the scale accumulation [12]. A thin layer may form on

the scrubber's surface but the scale will not continue to grow. The

calcium sulfite continues to precipitate but is in a non-adhering

form that stays in suspension and can be blown down from the system.

These proprietary scale inhibitors are surface active and,

therefore, are affected by particulate matter in the recirculation

liquor. These inhibitors are of little use in the combination

scrubbers which remove both SOp and particulate or in systems with

inefficient particulate control equipment preceeding the scrubber

[12].

Dual-Alkali

In this process the recirculation liquor is a sodium sulfite

solution that is regenerated with lime. The primary chemical

reactions which occur in the dual-alkali scrubber are probably the

same as those in a soda ash scrubber. The resulting sodium

sulfite/sodium bisulfite liquor is then regenerated with lime in a

10

clarifler where the following reactions occur [8]:

Na2S03 + Ca(OH)p — > 2NaOH + CaS03 (1-10)

NapSO^ + Ca(0H)2 — ^ 2NaOH + CaSO^ (1-11)

NaHS03 ^ ^^i^^)2 — ^ ^ ^ " ^^03 + HpO (1-12)

NaHSO^ + Ca(0H)2 — ^ NaOH + CaSO^ + HpO (1-13)

The soluble sodium compounds leave with the overflow of the clari­

fler and are pumped back to the SOp scrubber. The calcium precipi­

tates are discharged with the clarifier underflow, dewatered and

then disposed as a semi-dry solid. Most of the sodium values are

conserved with this system. Only a small amount of sodium carbonate

make-up is required to replace the sodium which is lost with the

dewatered solids and from mist eliminator carry-over [13]. This

system is superior to a limestone system for the following reasons

[9]:

1. scaling, plugging and erosion potentials are reduced,

2. SOp removal efficiency is greater than for limestone,

3. lower recirculation rates are required,and

4. greater utilization of lime and fev/er solid wastes.

Scrubber Design

A number of SOp scrubber designs are presently used in

industry. The current trend is away from combination scrubbers

which remove both particulate and SOp from the flue gas to

separate units for removal of SOp and particulates. The pressure

drop across combination scrubbers is high and the entrained

particulate matter aggravates scaling problems. Also, when these

11

scrubbers malfunction, both particulate and SOp emission control

is lost. Most of the utility and large industrial boilers being

built today are designed with electrostatic precipitators or bag-

houses preceeding SOp scrubbers [5].

SOp scrubbers can be classified into two broad categories:

throwaway and regenerative [8]. Regenerative processes regenerate

the absorbing solution and recover elemental sulfur or sulfuric acid

in the process. Generally, these processes are more capital

intensive and more difficult to operate than throwaway processes.

The economics of regenerative processes generally are not very

attractive and as a result, few of these processes are in current

use [5].

Throwaway processes are categorized into two basic types:

so-called "dry" scrubbers, and wet scrubbers. Dry scrubbers are

elementary spray drying units. An alkali liquor is forced through

an atomizing spray head which discharges into the flue gas stream.

The resulting fine mist contacts the flue gas and the SOp is

absorbed by the dissolved alkali. The heat in the flue gas

evaporates the water associated with the alkali forming dry

particulate matter. There is no liquid recirculating nor blowdown

as a waste stream from this type of scrubber. Since particulate

matter is generated by this scrubber system, it is necessary for the

particulate pollution control device to follow the SOp scrubber.

This is usually a baghouse since it has been found that additional

SOp removal occurs on the surface of the bags [14].

Wet scrubbers are the most numerous SOo removal devices [7].

There are several wet scrubber designs which are currently in use

12

such as weir, cascading, venturi and constant level contact tank

scrubbers, but the most common are spray towers and sieve tray

scrubbers.

Spray Tower

In these scrubbers, several levels of spray heads are mounted

in the scrubber vessel which produce an alkali liquor "rain" through

which the flue gas is forced. The two main advantages of spray

towers are that they have a low pressure drop and scaling is not a

major problem. The main disadvantages are that they are not as

efficient as other designs and require high liquid/gas (L/G) circu­

lation ratios (e.g., 30 to 40 gpm of recirculating liquor: 1000

actual cubic feet per minute (ACFM)) [7].

Sieve Tray

Flue gas is forced through a perforated plate which is flooded

with alkali liquor. As the flue gas flows through the holes in the

trays, it is forced into intimate contact with the liquor. Sieve

tray scrubbers are usually more efficient than spray towers but the

pressure drop across this type of scrubber is relatively high.

Sieve tray scrubbers typically require liquid/gas ratios of about

10:1. Fan horsepower requirements are higher than for spray towers

than sieve tray units. Scaling can seriously affect any scrubber's

on stream availability [7].

13

Corrosion Protection

The environment inside SOp scrubbers is extremely corrosive as

a result of low pH values and high chloride concentration (chloride

is Introduced to the scrubber from entrained fly ash). As a result,

special-heat resistant and acid resistant materials are needed for

scrubber construction [16].

The inlet of the SOp scrubber is a difficult area to protect

against corrosion. The SOp concentration in this area is greater

than anywhere else in the system, and the temperature is also

higher.The only effective methods of dealing with this problem is to

construct the inlet section of the scrubber using a stainless steel

with a high molybdenum content. Tests have shown that stainless

steel which contains at least four percent molybdenum withstands

acid/chloride attack exceptionally well [16]. Table 1 gives the

corrosion rates for various alloys in an acid/chloride environment.

The corrosion rates presented in Table 1 are an average of five

corrosion tests which were performed on each of these metals in

several different SOp scrubbers. The maximum pit depth which was

observed during these tests is also presented for each metal.

Alloys high in molybdenum are very expensive; Inconel^ 625 or

Hastalloy^C costs approximately 25 times as much as mild steel.

These alloys can be clad in steel but the clads are nearly as

expensive as the alloy alone.

Inside the scrubber vessel, the environment is corrosive but

the temperature is in the range where flake glass polymeric resin

14

Table 1

Corrosion Rates for Various Alloys Exposed

to Flue Gas Desulfurization Environment [16]

Alloy Molybdenum

Content (%)

Average Corrosion Rate

(mils/yr)

Max Pit Depth (mils/yr)

Hastalloy C-276 16.0 0

Inconel 625 9.0

Allegheny AL-GX 6.5

Hastalloy C 6.5

Incoloy 825 3.0

317 Stainless 325

Carpenter 20-3 2.50 0

304 Stainless 0 181 31

AISI 1010 Steel 0 7324 Corroded Away

15

linings can be used. The scrubber is usually constructed of mild

steel with several coats of the flaked glass troweled on the inside

of the vessel.

These linings are effective in protecting the mild steel shell

but require considerable maintenance. High temperature excursions

and years of use will cause the lining to crack and peel, thus

exposing the exterior shell to the corrosive environment in the

scrubber. It is often difficult to detect flaws in the lining until

the outer shell has been perforated with corrosion.

In an attempt to eliminate this problem. Basin Electric Power

Corporation [17] has built two SOp scrubbers of 316 stainless steel.

It is not clear at this time if the extra capital costs will be

justified by lower maintenance costs and increased availability.

The spray headers, spray heads and support beams are also made

of acid-resistant metals or plastics. Sieve trays are typically

made of stainless steel [5]. Mist eliminators are commercially

available in various plastics and acid resistant metals. Here

again, metal is much more costly. If plastic mist eliminators are

installed, they should be made of high temperature material to allow

for occasional recirculation pump failure. Without the

recirculation liquor to quench the flue gas, the temperature in the

scrubber increases.

Auxiliary equipment such as pumps, strainers and piping also

must be constructed of acid-resistant materials. Recirculation

pumps are generally a high-maintenance item. There are varying

reports in the literature describing the success of rubber-lined

(R) pumps [5, 12]. Carpenter 2 0 ^ has been used in recirculation pumps

16

with good success. Pump seals are often a problem, as scale can

form in and around the seal and eventually causes leaks [11].

Scrubber piping is usually made of plastic to prevent corrosion;

however, plastic pipe is easily broken. Periodically, recirculation

lines become clogged with scale. Plastic pipes can be damaged when

attempting to clean them out. [11]

Cost Information

Cost information for a number of scrubbing processes including

limestone, sodium carbonate, and regenerative processes in both

industrial and utility applications is presented in the literature

[7, 19, 20]. A summary of reported costs is given in Table 2 [7].

This summary lists five non-regenerable and two regenerable

processes. The ranges and average reported costs for these two

categories are reported as well as the range and average for each

individual process. The results show capital costs for all existing

lime/limestone, sodium carbonate, and fly ash throw-away systems to

be comparable. Operating costs for the different throw-away

scrubbing systems are competitive. Lime systems are the most

expensive due to the relatively high cost of lime and poor reagent

utilization. Reagent utiliztion averages about 70 per cent for both

lime and limestone scrubbers. Fly ash/limestone systems have the

lowest operating cost since a portion of the scrubbing reagent is

flyash which decreases the amount of limestone needed.

In 1978, 93 per cent of all utility scrubbers were lime or

limestone processes [7] whereas 90 per cent of all industrial

17

Table 2

Reported Capital and Annual Costs [7]

Type

All

New

Retrofit

Non-regenerable

Regenerable

Limestone

Lime

Alkaline fly ash/limestone

Alkaline fly ash/1ime

Sodium carbonate

Magnesium oxide

Wellman-Lord

Capital

Range $/kW

29.3-156.9

47.5-120.7

29.3-156.

29.3-120.7

156.8-156.9

47.5-99.5

29.3-120.7

47.9

77.1-86.0

42.9-113.6

156.8

156.9

Costs

Average $/kW

78.0

78.8

77.2

71.7

156.8

71.4

75.3

47.9

81.6

78.3

156.8

156.9

Annual

Range mills/kWhr

0.27-14.86

0.27-14.35

2.10-14.86

0.27-14.35

14.86

1.61-2.99

2.75-14.35

1.99

0.27

2.10

NA

14.86

Costs

Average milIs/kWhr

5.6

4.3

7.4

5.2

14.9

2.1

9.3

1.99

0.27

2.1

NA

14.86

NA = Not Available

18

scrubbers were sodium based systems [19]. The average capital costs

for an Industrial scrubber as reported by Tuttle are approximately

$30/kW and operating costs are 1.9 mils/kWhr [19]. The average size

of an industrial scrubber is about 20 MW or 40,000 scfm inlet flow.

No explanation was found for the difference in cost per unit between

utility scrubbers and industrial scrubbers.

The cost for a SOp scrubber system can be distributed using

four basic areas: capital, operations, maintenance and system

availability. These costs can vary as a function of the absorbent

for which the system is designed. The Tennessee Valley Authority

(TVA), a major electric utlity located in the southeastern United

States, has developed a scrubber cost estimating procedure which

breaks each of these main categories into component parts[20]. The

TVA method assumes an availability (total operating hours of the

scrubber divided by the total operating hours of the power

generation unit) of 80 per cent. Means of modifying the TVA

procedure to allow for differing availability rates is discussed in

Chapter III.

Capital

Table 2 indicates that the capital investment for limestone

scrubbers and soda ash scrubbers has approximately the range for

similar sized units that are operated in the acid pH range [5].

However, soda ash scrubbers do not require classifiers or vacuum

filters for solids separation. Also, soda ash arc more efficient at

neutral or alkaline pH values. Therefore, smaller units would be

required and the scrubber could possible be constructed of mild

19

steel. Scrubbers operated in the acid pH range must be lined with

an acid-resistant lining such as a glass flake polyester resin

mastic. Since these linings are susceptible to heat and

delamination, some manufacturers now use stainless steel or steel

clad with exotic alloys such as Hastalloy'-'C or InconeV- 625 [16].

Either of these alternatives are much more expensive than mild steel

construction. Although the waste disposal facilities for limestone

and sodium alkali systems are Nevy different the cost of disposal

would be similar. A limestone system would require a land fill and

solids conveying system. The sodium system would require a lined

pond and possibly a seepage control system but the waste stream

would be pumpable.

Operations

Included in operating costs are reagents, electricity, and

labor. Reagent costs are determined by the amount of a particular

reagent needed to remove a given amount of SOp. As shown

previously, one mole of soda ash can absorb two moles of SOp. In

practice, this ratio will vary as a result of a portion of the

NapSO« being oxidized to NapSO^ in the scrubber vessel [8].

Electrical costs would be the same for similar units of the same

size regardless of the absorbent used. However, if scrubber

efficiency is improved, more of the flue gas can be by-passed around

the scrubber. The installation could be scaled down and horsepower

requirements for the system would be reduced. Labor for operations

is primarily a function of the system's complexity.

20

Scrubbers which require solids removal require more service time

than systems which do not require solids removal.

Maintenance

Maintenance costs are affected by 1) the complexity of the

system, 2) extent and rate of corrosion, and 3) degree and rate of

deposition [7]. As each additional piece of equipment is added to a

system, the chances for malfunction increase. Most malfunctions of

scrubbers, however, are caused by corrosion or deposition.

Maintenance costs for sodium-based scrubbers are reported to

be about 2.25 per cent of the initial investment per year and about

4.5 per cent of the initial investments for limestone systems [5].

In the limestone process, a 6 to 12 per cent limestone slurry is

pumped through the scrubber, whereas virtually clear liquor is used

in the sodium system. The abrasion and deposition caused by the

limestone slurry is one cause for the increased maintenance costs.

Availability

These costs are assessed by loss of production and lost oppor­

tunities for profit as a result of scrubber failures. A scrubber

system which requires two overhauls per year, as opposed to once

QMery two years can have significant effects on profitability.

Availability acts as a common denominator for comparing FGD

systems costs. For example, if the total expenditure for a system

over a ten year period is $20 million and its average availability

21

is 70 per cent, 1t is quite possible that a system which costs $30

million and whose availability was 97 per cent percent would be .ore

cost effective.

CHAPTER III

EXPERIMENTAL PROCEDURES

Often it is difficult to extrapolate the data generated by

pilot or bench scale experiments to full scale operations.

Therefore, a full scale industrial flue gas desulfurization system

was used for this study. The following sections describe this

facility and the experiments which were conducted.

Test Facilities

The two scrubbers which were used in this study are identical

Swemco sieve tray scrubbers. These scrubbers are illustrated in

Figure 1. The scrubbers are installed on two Foster Wheeler stoker-

fired boilers. The boilers are rated at 250 million BTU's/hr and

supply steam for the Texasgulf Inc. soda ash production facility

near Green River, Wyoming. The inlet gas flow to each scrubber is

225,000 ACFM at 300°F and a pressure of 24.0 inches of mercury. The

inlet SOp concentration varies with the sulfur content of the coal

which is burned and ranges from 1200 ppm to 5000 ppm. The scrubbers

are guaranteed by the manufacturer to be 90 per cent efficient but

the demonstrated efficiencies are higher.

The main scrubber vessel is preceded by a prequench section

which is constructed of Inconel 625. The prequench is designed

primarily to cool the inlet gas so as to reduce its volume. The

22

A ) PREQUENCH

B) BYPASS DUCT

23

CJ SIEVE TRAYS

(D^ DEMISTERS ^

(E^ COUPON LOCATIONS

Figure 1. Diagram of Test Scrubber [21]

24

water which is added to maintain the liquid level in the

recirculation tank is added at the prequench section. The makeup

flow is approximately 70 gpm at full load. In addition, 200 gpm of

recirculation liquor is routed to the prequench. The inlet gas

temperature to the prequench remains fairly constant at 300°F and

the gas temperature at the exit was constant at 120°F.

The main scrubber vessel is 15 feet in diameter and 25 feet

tall. The scrubber contains a bilevel sieve tray arrangement which

is isolated by seal legs. The upper tray is flooded with a 400 gpm

flow of recirculating liquor which is transferred first to the lower

level and then to the recirculation tank. An additional 400 gpm is

sprayed on the bottom side of the lower sieve tray. The gas flow is

counter-current to liquor flow.

After leaving the sieve trays, the gas stream passes through a

high efficiency mist eliminator before entering the stack. A

portion (up to 20 per cent) of the gas stream can be diverted around

the scrubbers either through a 48-inch manually controlled by-pass

or a 24-inch automated by-pass which controls the temperature in the

stack by regulating the amount of gas flow which is by-passed.

The scrubbers are preceded by cyclone separators and

electrostatic precipitators which together remove 99.2 per cent of

the particulate matter in the flue gas stream. A small amount of

particulate matter is entrained in the SOp scrubber but its

concentration is controlled by purging the system. An induced draft

fan driven by a steam turbine is located between the electrostatic

precipitator and the SOp scrubber. This fan maintains a negative

25

draft in the boiler and forces the flue gas through the scrubber.

Figure 2 illustrates the general arrangement of the air pollution

control equipment. These pollution control devices are common to

the Industry and the information gathered from this system should be

applicable to most other wet scrubber systems.

Test Procedures

The tests which were conducted for this study were designed to

determine the applicability and cost effectiveness of utilizing soda

ash and waste streams produced in the manufacturing of soda ash as

SOp absorbents. Specifically, the purpose of these tests was to

determine if it is more cost effective to operate sodium carbonate

scrubbers at an alkaline or acid pH and to determine if crystallizer

purge liquor, dissolved trona or contaminated storm water run-off

could be used effectively as SOp absorbents without causing severe

operation problems. A total of four tests were conducted to supply

this needed information.

Test Series One

This test was to determine, if an SOp scrubber that was operated

at an alkaline pH could be made of mild steel, and if so, what would

be the cost in terms of additional absorbent usage and scale

prevention. The deposition rate of calcium sulfite was also

measured as a part of this test. Three different conditions were

26

c E o.

o S -r-

o

O Q-

<

O

c

en

i . S-

03

a

CD

CsJ

cu s -CD

a: UJ _i

o cu

27

examined. These were as follows:

1) scrubber make-up water containing 100 ppm calcium

carbonate with no chemical treatment,

2) scrubber make-up water containing 100 ppm calcium

carbonate with chemical treatment, and

3) scrubber make-up water containing 0 ppm calcium carbonate,

The chemical treatment used was a proprietary acrylic acid

derivative (Betz Chemical Company Product 436). The chemical feed

rate was set at 3 gallons per day and the blow-down rate from the

SOp scrubber was set at 5 gpm. The acrylic acid additive is

designed to distort the crystalline structure of CaoSO-, so that

it can not form a crystal lattice [12].

To obtain the goals of this test series, it was necessary to

measure the following parameters:

1) corrosion and deposition rates,

2) sodium carbonate usage, and

3) pH of the scrubber recirculation liquor.

The following sections describe how each of these parameters were

monitored.

Deposition/Corrosion Rates

The deposition and corrosion rates were studied through the

observation of mild steel coupons installed in the scrubber and

through the measurement of differential pressure across the

scrubber. The mild steel coupons, which measured 2 x 6 x 0.125

inches, were installed in the test scrubbers at the location

28

specified In Figure 1. These coupons were first weighed and then

the end which was to be attached to the coupon holder was covered

with a water- resistant tape. The tape was applied to prevent bias

in the data from scale sluffing off the coupon when it was detached

from the coupon holder. The coupons were weighed and replaced with

new coupons at one-month intervals. Scale samples were then scraped

from the coupons. The chemical constituents and the methods used

for analysis are listed in Table 3.

Table 3

Scrubber Scale Analysis

Test Method Calcium Instrumentation Laboratory Inc.

Atomic Absorption/Spec-trophotometer Model 951

Sulfite Standard Methods [22], lodide-lodate

Sulfate Dionex'-' Ion Chromatograph, Model 125

Silica Hatch Co. Silica Test Kit Model SI-5

Alkalinity Standard Methods [22], Titri-metric to methyl orange end point

The proportions of these chemical constituents in the scale

samples varied with changes in scrubber operations but together they

accounted for over 95 per cent of the total mass. A 0.375 inch

diameter hole was drilled through each coupon one inch on center

from the distal end of the coupon. The percent occlusion of these

29

holes was also measured monthly. The purpose of the holes was to

simulate the holes in the scrubber contact sieve trays.

In addition to the corrosion/deposition coupons which were

Installed in the scrubber, another test for deposition was devised

using the differential pressure across the SOp scrubber. A

differential pressure cell had been installed on the SOp scrubber

that measures the static pressure at the inlet of the scrubber

Instantaneously in pounds per square inch. As scale forms on the

sieve trays, the perforations in the trays become restricted. This

subsequently restricts the air flow through the scrubber and in­

creases the scrubber inlet static pressure. This test does not

measure deposition rate as a function of scale thickness but as a

function of back-pressure in the scrubber. In this way the problem

of reduced air flow and reduced boiler capacity are addressed

directly.

Reagent Usage

A 10 per cent by weight sodium carbonate solution was used as

the scrubbing reagent in this test series. The solution was made in

batches so the sodium carbonate concentration remained constant at

10 per cent during this test series. To determine usage rate during

the test period, the reagent flow was measured by magnetic flow

meter.

pH Determinations

The pH in the scrubber during the tests was determined with a

continuous pH meter manufactured by TBI of Carson City, Nevada.

30

Samples were collected every eight hours during the test period and

checked against a laboratory pH meter which was calibrated with

prepared pH standards. Compensation was made for any drift in the

continuous pH monitor.

Test Series Two

This series was designed to determine the suitability of soda ash

crystal lizer train purge liquor as an SOp scrubbing agent.

Parameters of importance included determining reagent usage, removal

efficiency and the cost of controlling silica scale which resulted

from using this material.

A typical analysis of this liquor is presented in Table 4.

Table 4

Evaporator Purge Liquor Analysis

Constituent Wt. %

Sodium Carbonate 30.0

Sodium Cloride 0.2

Sodium Sulfate 1.1

Silica 0.4

Sodium Phosphate 0.07

Water 68.0

31

Reagent Usage

To determine SOp absorbent usage at various pH values, the flow

of crystallizer purge liquor added to the scrubber was measured by

magnetic flow meter and the total alkalinity of this liquor was

measured every two hours by titration [22]. The amount of scrubbing

agent added to the scrubber per unit time was then calculated.

Each test started with the scrubber recirculation liquor at a

pH of 6.5. The reagent feed was increased stepwise until the

recirculation liquor pH reached approximately 7.5. After the

reagent feed was increased, it was not increased again until the

system had reached steady state. The system was considered to have

reached steady state when three consecutive pH measurements taken at

30 minute intervals showed no variation.

Silica Scale

It was anticipated that the silica in this liquor would

precipitate in the scrubber. Coupon weights were used to determine

deposition rates. The coupons were placed and removed monthly and

weighed. The differential pressure across the sieve trays was also

monitored and used as an indication of scale formation. Make-up

water used to maintain operating levels in the scrubber contained no

calcium so as to prevent bias in the data from the formation of

calcium sulfite scale. These parameters were monitored for a period

of four months.

32

SOp Removal Efficiency

The scrubber inlet SOp concentration was measured by two

methods. The first method was the U.S. EPA Stack Sampling Method 5

[4]. The second method entailed shutting down the SOp scrubber

recirculation pump. The scrubbing liquor thus would not contact the

flue gas and would allow untreated gas to flow through the SOp

scrubber.

The gas was measured in both tests at the scrubber outlet with

a Lear Seigler Inc. model SM-810 continuous SOp emissions monitor.

This monitor has been certified by the U.S EPA and the State of

Wyoming for its capibility to meet state and federal accuracy

requirements. Agreement between the two methods was within the ^10

per cent required by these regulations. The results from the

continuous emissions monitor for inlet and outlet SOp concentrations

were used to calculate scrubber efficiencies.

As the feed rate of scrubbing agent to the scrubber was varied,

removal efficiency was calculated using the following formula:

[SOp] J - [SOp] Q

Efficiency = X 100

[SOp] J

The inlet and outlet SOp concentrations were measured every two

hours during each test. Efficiency tests were conducted during all

three repetitions.

33

pH Determinations

The pH determinations for test series two Were made the same as

in the first test series. A continuous pH meter was used and

calibrated once every eight hours.

Test Series Three

The purpose of this test was to determine the suitability of

dissolved trona as a SOp scrubbing agent; specifically, waste water

and water generated in the mining of trona. Trona is a naturally

occurring ore of sodium sesqui-carbonate (NapC03 + NaHCO^ * 2H^0)

[23]. This mineral is water-soluble and in the process of mining

the ore, water containing dissolved trona is generated.

Three samples of this water were collected on three consecutive

days and the analysis of these samples are presented in Table 5.

Table 5 Mine-water Analysis

Weight %

Constituent

Total Alkalinity

Sodium Chloride

Sodium Sulfate

Silica

Sodium Phosphate

Calcium Carbonate

Water

Sample 1

12

> 0.20

> 0.01

> 0.01

? 0.01

> 0.01

87

Sample 2

12

0.20

> 0.01

> 0.01

> 0.01

> 0.01

87

Sample 3

10

0.20

>0.01

>0.01

>0.01

>0.01

89

34

In determining the suitability of this material, scale

deposition did not have to be considered. Since the levels of

calcium and silica in these materials were less than 50 ppm, it was

anticipated that scaling would not be a problem. As a result, only

SOp removal efficiency and reagent usage were measured.

Reagent Usage

As in test series 2, the reagent flow was measured by a

magnetic flow meter during the three individual tests in this

series. The total alkalinity of the scrubber feed liquor was

measured every two hours during this test by titration.

SOp Removal Efficiency

Inlet SO9 concentration was measured in this test series one at

the beginning of each of the three individual tests by shutting down

the scrubber recirculation pumps and allowing untreated flue gas to

go through the scrubber. This gas was then analyzed by the Lear

Seigler continuous emissions monitor. During the actual tests, the

scrubber effluent SOp was measured continuously by the Lear Seigler

monitors.

Test Series Four

The purpose of this test was to determine the efficacy of storm

water runoff collected from a soda ash manufacturing facility as an

SOp absorbent. A typical analysis of this water is presented in

Table 6.

35

Table 6

Typical Analysis of Runoff

Constituent %

NapC03

NaHC03

NaCl

NaSO^

CaC03

Organics

HpO

2.0

>0.1

>0.1

>0.1

>0.1

0.2

97.5

Runoff waters contain sodium carbonate values as a result of

coming in contact with the surface ore stockpile and any soda ash

that is inadvertantly spilled within the site containment area.

Low levels of dissolved silica and calcium were observed in

this water and as a result, no scale inhibitors were added.

Reagent Usage

To establish the amount of reagent used per time interval, the

total alkalinity was measured by titration and the flow of the

scrubbing agent was measured by a magnetic flow meter. This flow

and concentration data was then used to determine the amount of

sodium alkali entering the scrubber per unit time.

36

Scrubber Efficiency

The inlet SOp concentration was measured in this test series

solely by allowing untreated gas to be passed through the SO.

scrubber prior to the start of the three individual tests.

Again the scrubber recirculation pump was turned off and the scrubber

liquor drained out of the scrubber so as not to contact the flue gas.

Cost Data

An objective of this study was to compare the total costs for a

limestone SOp scrubbing system, a soda ash system operated in an

acidic pH range, and a soda ash system operated in a basic pH range.

As has been previously discussed, the FGD costs can be expressed as

capital, operating, maintenance and availability costs. The follow­

ing paragraphs present mathematical descriptions of these costs and

factors which may affect them.

Capital

Capital expenditures were derived by assuming a ten-year loan

at 10 per cent compound interest. The 10 per cent interest figure

was selected because it currently does not appear that the interest

rate will fall below this mark in the near future. The capital cost

estimating procedures used by the Tennessee Valley Authority were

used in estimating the cost/kw. Assuming monthly payments, capital

costs are then expressed as [24]:

C ^ = (3-1)

1-(1 + i) -n

37

where:

C^ = Annual Capital Costs

C = Total Investment

1 = Interest Rate/pay period

n = Number of Pay Periods

Operating Costs

Annual operating cost is primarily the sum of the reagent

costs, labor costs and energy costs. Labor costs and energy costs

are dependent upon the physical size of a particular unit. Reagent

costs are a function of reagent utilization and amount of SOp which

is to be removed.

The amount of SOp which is to be removed is dependent upon the

type of coal used and the SOp emission standards which a facility

must meet. The following equation was used for determining reagent

costs:

where:

R^ = (M) (I^) (U^) (MW^) (KJ / 2000 lb/ton (3-3)

R = Reagent Cost/Year c

M = Moles SOp Produced/Year

I = Required Removal Efficiency

U = Reagent Utilization in Tons r ^

MW = Molecular Weight of Reagent

K = Cost/Ton

38

Total operating costs can then be expressed as:

C„ = C + C +C + C-, (1-^) 0 r e w 1 ^'^ ^^ where

CQ = Annual Operating Costs

C^ = Annual Reagent Costs

Cg = Annual Energy Costs

C^ = Annual Water Treatment Costs

C-j = Annual Labor Costs

Maintenance

Maintenance costs are a function of physical size of the unit

and the type of operating system which is employed. For use in this

paper, the maintenance costs were calculated as follows:

where

c^ = (C,) (S,)

C = Annual Maintenance Costs m C = Total Annual Maintenance Cost s

for a limestone Scrubber

S^ = System Factor

The system factor is based on the findings of Smith [5] that were

discussed in Chapter II. The system factor for a limestone scrubber

is 1 and the system factor for a soda ash scrubber is 1/2.

Availability

The TVA estimating procedure which is used in this study

assumes 1700 operating hours per year or 80 per cent availability of

the generation unit annually. In comparing costs with this method,

39

variances in availability are calculated in the following manner:

Cg = (T^ - 1700)(C^ + Cp) (3-6)

where: C^ = Availabilitv a

Cr = Fixed Production Costs

C = Lost Profit Costs

T • = Downtime

If a system availability is better than the 80 per cent, then

the system will receive a credit equal to the above equation.

Cost Equation

The culmination of the various costs can be expressed in the

following equation as a total annual cost.

where:

C. = C^ + C^ + C^ + C^ (3-7) t c 0 m a

C. = Total Annual Costs

C = Annual Capital Costs c

C = Annual Operating Costs

C = Annual Maintenance Costs m

C = Annual Availability Costs a

The following equation will then give the total cost for the

life of the facility:

C^ = Y (C^) + Y^ (C^ - C^) (3-8)

where:

C. = Total Facility Lifetime Costs

Y = Years of Loan

Y^ = Life of Facility in Years

CHAPTER IV

PRESENTATION OF DATA

During the study period, the test boiler operated at between

70 and 78 per cent of its rated capacity. The sulfur content of the

coal burned during the study period varied between 0.4 and 0.6 per

cent. Water was added to the scrubber as needed to maintain the

level in the scrubber recirculation tank. The water flow during the

test was approximately 70 gpm. The results from the tests which

were conducted are presented in the following sections.

Test Series One

The goal of the first series of tests was to determine the

effects of operating a sodium carbonate SOp scrubber in an alkaline

pH range with respect to reagent usage, corrosion, and scaling. The

results of the tests which were conducted are presented in the

following sections.

Reagent Usage

The data presented in Tables 7, 8 and 9 show that as pH

increases, the amount of reagent used per unit time increases. This

was caused by three factors:

1) The removal efficiency increased with increased pH. Since

more SOp was removed, more reagent was required.

40

41

Table 7

Test Series One, Test One:

Reagent Usage and SOp Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

gpm

2.7 2.7 2.7 2.9 2.9 2.9 3.4 3.4 3.4 3.6 3.6 3.6 3.7 3.7 3.7 3.9 3.9 3.9 4.0 4.0 4.0 4.1 4.1 4.1

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.7 6.7 6.7 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

SOo

lb/MB

0.33 0.30 0.32 0.29 0.28 0.29 0.19 0.19 0.19 0.18 0.18 0.18 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15

Efficiency

% Removed

80 83 81 84 85 84 93 93 93 95 95 95 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0% Scrubber Purge Flow = 10 gpm Scrubber Feed Total Alkalinity = 15.1% as Na2C03 Inlet SOp = 1.06 Ib/MB Water Flow = Added as Needed

MHHI

42

Table 8 Test Series One, Test Two:

Reagent Usage and SOp Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

gpm

2.6 2.6 2.6 2.8 2.8 2.8 3.4 3.4 3.4 3.5 3.5 3.5 3.6 3.6 3.6 3.8 3.8 3.8 3.9 3.9 3.9 4.0 4.0 4.0

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.8 6.8 6.8 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

SOp

Ib/MB

0.31 0.29 0.29 0.28 0.28 0.28 0.18 0.17 0.18 0.16 0.16 0.16 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15

Efficiency

% Removed

80 83 83 84 84 84 93 93 93 95 95 95 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0% Scrubber Purge Flow = 10 gpm Scrubber Feed Total A l ka l i n i t y = 15.2% as NapCU3 In le t SOp = 1.03 Ib/MB Water Flow = Added as Needed

wmmaaask

43

Table 9 Test Series One, Test Three

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

% Sen ibbe Scrubber Scrubber Inlet Water

SO. Fl^

!r By-

^ " ^ ^^jwy«_ ui

Feed Flow

gpm

2.2 2.2 2.2 2.4 2.4 2.4 2.7 2.7 2.7 2.9 2.9 2.9 3.0 3.0 3.0 3.1 3.1 3.1 3.2 3.2 3.2 3.3 3.3 3.3

-Pass = 11.0% Purge Flow = 10 Feed , = 0 5W = i

Alkalinity .85 Ib/MB

gpm = IE

Mded as Needed

.p .x mw

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.7 6.7 6.7 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

5.2% as

v a 1 L. 1 1 1 c- 1 ciic-y r\t

SOp

Ib/MB

0.29 0.30 0.29 0.25 0.26 0.25 0.17 0.17 0.17 0.15 0.14 0.14 0.14 0.14 0.14 0.14 0.14 0.13 0.14 0.14 0.14 0.14 0.14 0.14

NapC03

:3U 1 Li

Efficiency

% Removed

80 79 80 84 84 84 92 92 92 96 97 97 97 97 97 97 97 98 97 97 97 97 97 97

mm

44

2) Less sulfite was converted to bisulfite ion in the

process. As described previously, one mole of NapCO^ can

react with two moles of SOp to form two moles of NaHS03;

however, as pH Increases, the reaction has a tendency to

stop after the Na2C03 has reacted with only one mole of

SOp to form NapS03.

3) Increased NaHC03 concentration was noted in the scrubber

purge liquor with an increase in pH. The concentration of

NaHC03 *" ^^^ scrubber purge also increases with pH.

Therefore, the loss of unreacted reagent from the scrubber

per unit time increases.

SOp Removal Efficiency

The data in Figure 3 shows the effect pH has on SOp removal.

The graph shows that efficiency Increases up to a pH of 6.5. The

removal efficiency then becomes constant. The data in the Table 10

shows that as pH Increases from 5.8 to 7.5 the ratio of the amount

of reagent used to the amount of SOp absorbed, increased by 27 per

cent.

Corrosion

The data presented in Table 11 shows the corrosion rates which

were observed at the various locations in the scrubber. The pH was

varied from 7.2 to 7.8 during these tests. The data shows that

there was no metal loss from the coupons which were placed in the

main scrubber vessel. These coupons were covered by a thin layer of

100-1

A TEST I

O TEST 2

n TEST 3

45

95-

o 90 LU

O

u. LU

^ 85

80

e-g

75 5.5 6.0 6.5

pH

7.0 7.5

Figure 3. Scrubber E f f i c iency vs pH.

46

Table 10

Stoichiometric Ratio at Various pH Values

pH

5.8

6.1

6.3

6.5

6.7

7.0

7.2

7.5

Moles of

SOp

Collected/Hr

2.13

2.21

2.45

2.55

2.55

2.55

2.58

2.58

Moles of

Na2C03

Used/Hr

1.91

2.10

2.42

2.62

2.68

2.78

2.87

2.95

Moles of Na2C03 used/

Moles of SO2

Collected

0.90

0.95

0.99

1.03

1.05

1.09

1.11

1.14

IBOBEtt

47

Table 11 Comparison of Coupon Corrosion Weight and Rate Changes

MONTHS AFTER START OF TEST

Initial 1 2 3

Location # Wt.(g) Wt.(q) Rate(Q/y) Wt.(g) Rate(g/y) Wt.(g) Rate(g/y) Wt.(q) Ratefg/y^

4 158.3 156.7 7.7 155.0 8.2 152.8 10.5 151.5 6.3

5 158.0 156.2 8.7 154.9 6.3 153.0 9.2 150.7 11.1

6 159.0 158.9 0.0 158.9 0.0 158.8 0.0 158.9 0.0

7 157.8 157.9 0.0 157.9 0.0 157.6 0.0 157.6 0.0

3 160.5 160.5 0.0 160.4 0.0 160.5 0.0 160.^ 0.0

48

silica scale and underneath the scale, the metal showed no signs of

pitting or crevice corrosion. The environment immediately above the

mist eliminators is highly corrosive as a result of the SO2 and

oxygen-laden by-pass gas coming in contact with the carry-over from

the SO2 scrubber. The coupons which were placed in the section

above the mist eliminator section were highly oxidized and pitted.

Metal loss was projected to be as high as 11.1 g/yr for coupons in

this area. Coupons placed at the inlet to the scrubber showed signs

of general corrosion and some pitting but less oxidation than at the

outlet.

Sealing

It was anticipated that operating the scrubber in an alkaline

pH would cause the deposition of calcium scale. This was demon­

strated in the first trial of the test, where 100 ppm CaCO.. make-up

water was used. Calcium sulfite scale formed on the sieve trays and

in the holes of the trays. As indicated in Table 9, the induced

draft fan back-pressure increased 6.0 inches of water in five days.

Next the boiler feed water supply which contained no calcium

was used as the source of makeup water to the scrubber. As can be

seen from the data in Table 12, the fan back pressure increased 0.6

inches of water over a four month period. The calcium scale

deposition rate seemed to have been greatly reduced. This was

confirmed by a later visual inspection. Coal fly ash in the in­

fluent gas to the scrubber contained some calcium and analysis of

scrubber purge water showed 50 ppm calcium carbonate. This

49

Table 12

Increases in Scrubber Back Pressure

From Calcium Sulfite Scale Formation

at 80% Boiler Load

Boiler Scale Inhibitor

Make-Up Water Condensate Addition

Diff. Press Diff. Press Diff. Press.

Date In. of Water Date In. of Water Date In. of Water

10/13/81 12.2 07/82 12.1 07/82 12.7

10/14/81 12.6 08/82 12.0 08/82 12.9

10/15/81 13.9 09/82 12.7 09/82 12.6

10/16/81 16.1 10/82 12.5 10/82 13.4

10/17/81 18.2 11/82 12.7

50

indicated that some calcium was introduced by fly ash carry-over but

not enough to cause scaling.

Often in Industrial or utility applications, excess condensate

is not available. Proprietary scale inhibitors are used with the

available water source. In this test a proprietary scale inhibitor

(Betz 406) was tested to determine if it stopped the formation of

scale. The data in Table 12 shows that the chemical was effective

when used with make-up water which contained 100 ppm CaCO^. The fan

back pressure increased 0.7 inches of water in a three month period

or 0.2 Inches of water per month. The data for this test and the

condensate test fluctuated to a small degree. This was probably

caused by deposition on the mist eliminators building up and then

sloughing away.

Test Series Two

Two sets of analyses were performed to determine the

suitability of the crystallizer purge liquor as a SOp scrubbing

agent. Reagent usage and removal efficiency were measured in the pH

range between 5.8 and 7.5. Silica scale formation resulting from the

use of crystal lizer purge liquor was also measured as well as the

effectiveness of a proprietary scale inhibitor.

Reagent Usage and SOp Removal Efficiency

The information in Tables 13, 14 and 15 shows that the

crystallizer purge liquor had virtually the same effect in the test

scrubbers as refined soda ash. Reagent usage and removal

efficiencies evidently were not affected by the presence of

51

Table 13 Test Series Two, Test One:

Reagent Usage and SO Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

gpm

4.4 4.4 4.4 4.7 4.7 4.7 5.5 5.5 5.5 5.9 5.9 5.9 6.1 6.1 6.1 6.3 6.3 6.3 6.4 6.4 6.4 6.6 6.6 6.6

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.7 6.7 6.7 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

SOp

1 b/MB

0.29 0.29 0.28 0.24 0.24 0.24 0.16 0.16 0.16 0.15 0.15 0.15 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13

Efficiency

% Removed

78 78 78 84 84 84 93 93 93 95 95 95 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0% Scrubber Purge Flow = 10 gpm Scrubber Feed Total Alkalinity = 10.6 wt.% as NapC03 Inlet SOp = 0.89 Ib/MB Water Flow = Added as Needed

52

Table 14 Test Series Two, Test Two:

Reagent Usage and SO Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

gpm

4.3 4.3 4.3 4.6 4.6 4.6 5.4 5.4 5.4 5.7 5.7 5.7 5.9 5.9 5.9 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.8 6.8 6.8 7.0 7.0 7.0 7.2 7.2 7.2 7.8 7.5 7.5

SO2

Ib/MB

0.26 0.26 0.26 0.22 0.22 0.22 0.18 0.17 0.17 0.14 0.14 0.14 0.12 0.12 0.12 0.12 0.12 0.12 0.12 0.12 0.12 0.12 0.12 0.12

Efficiency

% Removed

81 81 81 86 86 86 91 92 92 95 95 95 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0% Scrubber Purge Flow = 10 gpm Scrubber Feed Total Alkalinity = 10.3 wt.% as Na2CU3 Inlet SOp = 0.88 Ib/MB Water Flow = Added as Needed

53

Table 15 Test Series Two, Test Three:

Reagent Usage and SO Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

qpm

4.3 4.3 4.3 4.6 4.6 4.6 5.4 5.4 5.4 5.7 5.7 5.7 5.9 5.9 5.9 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.8 6.8 6.8 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

SO2

Ib/MB

0.25 0.25 0.26 0.22 0.22 0.21 0.17 0.17 0.17 0.15 0.15 0.15 0.13 0.12 0.12 0.12 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13

Efficiency

% Removed

80 80 81 86 86 86 92 92 92 95 95 95 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0% Scrubber Purge Flow = 10 gpm Scrubber Feed Total Alkalinity = 10.3 wt.% as Na2C03 Inlet SOp = 0.88 Ib/MB Water Flow = Added as Needed

54

dissolved silica or any of the other impurities found in the

crystallizer purge liquor.

Sealing

The crystallizer purge liquor which was used in this experiment

contained between 3000 and 6000 ppm soluable silica. As indicated

on Table 16, the silica precipitated in the scrubber and caused the

fan back pressure to increase at a rate of 8.3 inches of water per

month. It was thought at the outset that the scrubber coupons used

in Test 1 would simulate the silica scale deposition in the

scrubber. However, build-up of scale in the coupon holes did not

simulate the build-up in the sieve tray holes. The most probable

reason for this is that air flow was not being forced through the

coupon holes as it was through the sieve tray holes.

A proprietary surfactant (Nalco 8WT-512) was added to the

scrubber in an attempt to inhibit the silica scale from forming. As

Table 16 indicates, the surfactant did slow down the deposition

rate. In the months following the test period, it was found that

silica scale was causing the back pressure on the fan to increase at

a rate of 0.07 Inches of water per month. The scrubbers need only

to be cleaned every six months which was considered acceptable by

operations personnel.

Test Series Three

The data in Tables 17, 18 and 19 shows that the dissolved trona

behaves identically to the crystallizer purge liquor and soda ash in

removing SOp. During the dissolved trona test, the soluble silica

concentration remained below 150 ppm and the silica deposition rate

55

Table 16

Increase in Scrubber Back Pressure as a Result of Silica Scale Formation

at 80% Boiler Load

No Seal Inhibitor

Date

6/1/81

6/7/81

6/14/81

6/21/81

e Added

Diff. Press. In. of Water

12.2

13.5

15.7

18.0

Seal

Date

07/81

08/81

09/81

10/81

e Inhibitor Added Diff. Press. In. of Water

12.5

12.5

12.7

12.7

56

Table 17

Test Series Three, Test One: Reagent Usage and SOp Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

•qpm

4.4 4.4 4.4 4.7 4.7 4.7 5.5 5.5 5.5 5.7 5.7 5.7 6.0 6.0 6.0 6.2 6.2 6.2 6.4 6.4 6.4 6.6 6.6 6.6

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.7 6.7 6.7 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.8 7.5

SOp

Ib/MB

0.26 0.26 0.26 0.22 0.22 0.22 0.19 0.19 0.19 0.15 0.15 0.15 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13

Efficiency

% Removed

81 81 81 84 84 84 93 93 93 95 95 95 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0% Scrubber Purge Flow = 10 gpm Scrubber Feed Total Alkalinity Inlet SOp = 0.89 Ib/MB Water Flow = Added As Needed

= 10.1% as Na2C03

57

Table 18

Test Series Three, Test Two: Reagent Usage and SOp Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

qpm

4.1 4.1 4.1 4.5 4.5 4.5 5.3 5.3 5.3 5.6 5.6 5.6 5.8 5.8 5.8 6.0 6.0 6.0 6.1 6.1 6.1 6.3 6.3 6.3

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.8 6.8 6.8 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

SOp

Ib/MB

0.30 0.31 0.31 0.26 0.26 0.26 0.19 0.18 0.18 0.15 0.14 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15

Efficiency

% Removed

81 81 81 84 84 84 93 93 93 97 97 97 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0% Scrubber Purge Flow = 10 gpm Scrubber Feed Total Alkalinity = 10.5 Inlet SOp = 0.90 Ib/MB Water Flow = Added as Needed

58

Table 19

Test Series Three, Test Three: Reagent Usage and SOp Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

qpm

5.3 5.3 5.3 5.8 5.8 5.8 6.7 6.7 6.7 7.2 7.2 7.2 7.4 7.4 7.4 7.6 7.6 7.6 7.9 7.9 7.9 8.2 8.2 8.2

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.7 6.7 6.7 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

SOp

Ib/MB

0.33 0.32 0.33 0.31 0.31 0.31 0.19 0.19 0.20 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15

Efficiency

% Removed

81 81 81 84 84 84 93 93 93 97 97 97 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0%

Scrubber Purge Flow = 10 gpm

Scrubber Feed Total Alkalinity = 10.08% as NapCO^

Inlet SOp = 1.10 Ib/MB

Water Flow = Added as Needed

59

was slow. Upon visual inspection of the scrubber, it was observed

that some scale had formed around the edges of the sieve trays.

This was probably due to reduced gas flow in these areas. In the

middle of the tray, the velocity of the flue gas was probably great

enough to prevent scale from formating.

Test Series Four

The data in Tables 20, 21, and 22 shows that the site

run-off water used in this test worked well to absorb SOp. The

removal efficiencies compared closely with the efficiencies observed

during the previous test series. Since the site run-off water was

more dilute with respect to alkalinity than the other absorbants

tested, the liquor flow was relatively higher. The actual NapCO..

consumption was about the same in all four test series.

During the test, large volumes of foam were generated in the

test scrubber. The foam did not cause any operational problems but

was somewhat of an aesthetic problem. After the test was completed,

the scrubber was inspected and it actually has less scale in it than

when the test started. The organics in the water which caused the

foam evidently worked as surfactants to actually remove silica

scale. As evidenced by the efficiency testing, the foaming or the

presence of organics in the scrubber did not diminish or enhance SOp

removal efficiencies or otherwise affect reagent usage.

60

Table 20

Test Series Four, Test One: Reagent Usage and SOp Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

qpm

17.5 17.5 17.5 19.0 19.0 19.0 22.0 22.0 22.0 24.0 24.0 24.0 25.0 25.0 25.0 25.5 25.5 25.5 26.0 26.0 26.0 27.0 27.0 27.0

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.8 6.8 6.8 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

SOp

Ib/MB

0.29 0.28 0.28 0.25 0.24 0.25 0.18 0.17 0.17 0.14 0.14 0.14 0.14 0.14 0.14 0.14 0.13 0.13 0.13 0.13 0.13 0.13 0.13 0.13

Efficiency

% Removed

78 79 79 83 84 83 91 92 92 97 97 97 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0% Scrubber Purge Flow = 10 gpm Scrubber Feed Total Alkalinity = 2.6 wt.% as NapC03 Inlet SO = 0.87 Ib/MR Water Flow = Added as Needed

61

Table 21

Test Series Four, Test Two: Reagent Usage and SOp Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

% Scrubber E

Feed Flow

gpm

18.0 18.0 18.0 20.0 20.0 20.0 23.0 23.0 23.0 25.0 25.0 25.0 25.5 25.5 25.5 26.0 26.0 26.0 27.0 27.0 27.0 28.0 28.0 28.0

ly-Pass = 11.0%

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.8 6.8 6.8 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

SOp

Ib/MB

0.30 0.30 0.30 0.26 0.25 0.26 0.19 0.19 0.18 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15

Efficiency

% Removed

81 81 81 84 84 84 93 93 93 97 97 97 97 97 97 97 97 97 97 97 97 97 97 97

Scrubber Purge Flow = 10 gpm Scrubber Feed Total Alkalinity = 2.6 wt.% as Na2C03 Inlet SOp = 0.90 Ib/MB Water Flow = Added as Needed

62

Table 22

Test Series Four, Test Three: Reagent Usage and SO Removal Efficiency Results

Sample

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

Feed Flow

qpm

17.5 17.5 17.5 19.0 19.0 19.0 22.0 22.0 22.0 24.0 24.0 24.0 24.5 24.5 24.5 25.5 25.5 25.5 26.0 26.0 26.0 27.0 27.0 27.0

pH

5.8 5.8 5.8 6.1 6.1 6.1 6.3 6.3 6.3 6.5 6.5 6.5 6.7 6.7 6.7 7.0 7.0 7.0 7.2 7.2 7.2 7.5 7.5 7.5

SO2

Ib/MB

0.30 0.30 0.29 0.25 0.25 0.25 0.18 0.18 0.18 0.14 0.14 0.14 0.14 0.14 0.14 0.15 0.15 0.15 0.10 0.10 0.10 0.05 0.05 0.05

Efficiency

% Removed

81 81 81 84 84 84 93 93 93 97 97 97 97 97 97 97 97 97 97 97 97 97 97 97

% Scrubber By-Pass = 11.0% Scrubber Purge Flow = 10 gpm Scrubber Feed Total Alkalinity = 2.6 wt.% as Na2C03 Inlet SOp = 0.88 Ib/MB Water Flow = Added as Needed

63

Cost Comparisons

Using the data generated by the four experiments, three

scrubber system cost estimates were made. The first estimate is for

a lime/limestone system, the second for a sodium-alkali system

operated at an acid pH and the third estimate is for a sodium-alkali

system operated at a basic pH. All three estimates are based on the

procedure described in Chapter III. The capital investment was

first Itemized for each system foiled by the annual operating and

maintenance costs. Availability costs were estimated using

maintenance records to determine steam plant down time caused by

scrubber failures.

Capital Cost Estimates

Table 23 is an estimate of the capital cost for a limestone

scrubber done by the Tennessee Valley Authority in 1980. The

capital estimates for the acid and the basic sodium alkali scrubbing

systems are shown in Tables 24 and 25, respectively. These estimates

were derived from the TVA estimate by altering the various component

costs according to the findings of this study. The values presented

fall within the ranges given in Table 2. The total capital costs

for a lime system is greater than either type of soda ash scrubber

due to the required solids handling equipment. These estimates show

that a basic sodium alkali system would have the lowest capital

costs of the three systems. This is due to the fact that a basic

soda ash scrubber would be smaller than either a lime or acid soda

ash scrubber. A basic soda ash scrubber would be more efficient and

64

Table 23

Limestone Scrubbing Process Capital Investment [20]

Direct Investment

Capital Investment $1000

Material handling ^ .p^ Feed preparation yZii Gas handling ,f'^°2 SO^ absorption OOQQQ

Refieat 2^,988 Solids disposal 2*868

Total 47,270

Service, utilities, and miscellaneous 2,836

^^^^^ 50,106

Landfill or pond construction 2,076 Landfill equipment '500

Total 52,682

Indirect Investment (% of D. I.)

Engineering design and supervision 7 3,688 Architect and engineering contractor 2 1,057 Construction expense 15 7,902 contractor fees 5 2,634 Contingency 14 7,375

Total Fixed Investment 75,338

Other Capital Investment

Allowance for startup and modifications ' 5,732 Interest during construction 11,781 Land 641 Working capital 3,161

Total Capital Investment 96,653

$/kW 194

65

Table 24

Acid Soda Ash Scrubbing Process Capital Investment

Direct Investment

Capital Investment $1000

Material handling 3 .gy Feed preparation ^Q_ Gas handling 11~I?P

S0« absorption n \f)r. Reheat 3*30^ Solids disposal IQ_

Total 28,930 Service, utilities, and miscellaneous 2,836

Total 31,766

Landfill or pond construction 2,076 Landfill equipment 500

Total 34,342

Indirect Investment

Engineering design and supervision Architect and engineering contractor Construction expense contractor fees Contingency

Total Fixed Investment 49,109

Other Capital Investment

Allowance for startup and modifications ' 5,732 Interest during construction 8,064 Land 641

Working capital 3,161

Total Capital Investment 66,707

$/kW 134

(% of D. I.)

7 2

15 5

14

2,404 687

5,151 1,717 4,808

66

Table 25

Basic Soda Ash Scrubbing Process Capital Investment

Direct Investment

Capital Investment $1000

Material handling 3 497 Feed preparation IQ_ Gas handling -^-^ -j pg SOp absorption 9*000 Reneat 3^304 Solids disposal IQ.

Total 26,930

Service, utilities, and miscellaneous 2,836

Total 29,766

Landfill or pond construction 2,076 Landfill equipment 500

Total 32,342

Indirect Investment (% of D. I.)

Engineering design and supervision 7 2,262 Architect and engineering contractor 2 647 Construction expense 15 4,851 contractor fees 5 1,617 Contingency 14 4,528

Total Fixed Investment 46,247

Other Capital Investment

Allowance for startup and modifications ' 5,732 Interest during construction 7,752 Land 641 Working capital 3,161

Total Capital Investment 63,533

$/kW 127

67

a portion of the flue gas could be by-passed requiring a smaller

scrubber. To calculate the annual debt service for each of the

systems it was assumed the investment would be amoritized over a 30

year period at 10 per cent interest. Given this information," C c

from equation 3.1 for the three systems discussed would be as

follows:

System Type C^ ($10^/yr)

Limestone 10.1

Sodium Alkali (acid) 7.0

Sodium alkali (basic) 6.8

Operating and Maintenance Costs

The estimated annual operational and maintenance costs

presented in Tables 26, 27, and 28 were also determined by the TVA

method. These costs do not include reagent costs. The effect of

reagent prices will be discussed separately. The annual costs of

each system were changed to reflect the cost variances indicated by

the study. The three catagories which vary between the three

systems are steam for reheat, maintenance, and reagent costs. As

stated in Chapter II, the maintenance costs for soda ash scrubbers

are approximately one half that of lime/limestone scrubbers. In

addition to lower maintenance costs, a basic soda ash scrubber would

not require steam for stack reheat. The higher efficiency of a

basic soda ash scrubber makes it possible to use by-passed flue gas

for reheat instead of steam.

68

Table 26

Limestone Scrubbing Process Annual Operating Costs Less Reagent Costs

Category

Annual

Quantity

Unit

Cost

Total

Amount ($10^)

Operating labor

Steam for Reheat

Process Water

Electricity

Maintenance

28,000 man hrs

489,800 MB

243,400 kgal

53,588,000 kWhr

$20.00/man hr

$2.00/MB

$0.12/kgal

$0,029 kWhr

0.56

0.98

0.03

1.55

Material & Labor 2.04

Total 5.1

Table 27

Acid Soda Ash Scrubbing Process Annual

Operating Costs Less Reagent Costs

69

Category

Annual

Quantity

Unit Total

Cost Amount ($10 )

Operating Labor

Steam for Reheat

Process Water

Electricity

Maintenance

$28,000

489,800 MB

243,400 kgal

53,588,000 kWhr

$20/man hr

$2.00/MB

$0.12/kgal

$0.029/kWhr

0.56

0.98

0.03

1.55

Labor & Material 1.00

Total 4.1

M i a ttmniOB*

Table 28

Basic Soda Ash Scrubbing Process Annual

Operating Costs Less Reagent Costs

70

Category

Annual

Quantity

Unit Total

Cost Amount ($10^)

Operating Labor

Steam for Reheat

Process Water

Electricity

Maintenance

28,000 man hrs

0 MB

243,400 kgal

53,588,000 kWhr

$20/man hrs

$0/MB

$0.32/kgal

$0.029/kWhr

0.56

0

0.08

1.55

Labor & Material 1.00

Total 3.2

71

Reagent costs can be affected by at least three variables:

1) price of the reagent, 2) reagent utilization, 3) and the sulfur

content in the coal that is being burned. A cost comparison is made

in Table 29 based on equation 3-2 of the two sodium alkali scrubbing

systems. Coal containing 1, 3, and 5 per cent sulfur were used for

the comparison. The coal consumption rate is based on the

consumption rate of the test facility which was 5250 tons/yr/MW.

The coal comsumption rate for a 500 MW facility would then be 2.6

million tons/yr. Three reagent prices were used for each of the

sodium alkali systems: $50/ton, $75/ton, and $100/ton. Based on the

Information in Table 10, the reagent utilization values for the acid

sodium alkali system and the basic sodium alkali system were 0.9 and

1.1 respectively.

Table 29 Sodium Alkali Reagent Costs

$loVvr

Sulfur Content in Coal 1% 3% 5% Price/Ton $50 $75 $100 $50 $75 $100 $50 $75 $100

Acidic (R ) 3.9 5.6 7.8 11.6 17.5 23.2 19.4 29.1' 38.8 c

Basic (R ) 4.7 7.1 9.4 14.2 21.3 28.4 23.7 35.6 47.4

72

A cost of $7/ton was assumed for limestone [25] to obtain the

following reagent costs of a limestone system:

% Sulfur in Coal R^ ($10^/yr) c

1 0.8 3 2.4 5 4.1

Availability Costs

The TVA model assumes an 85 per cent availability for limestone

scrubbers. After reviewing the operational history of the test

scrubbers, it was found that the availability while operating at

acid pH values was approximately 92 per cent. Under alkali

conditions and proper scale control, the availability of the test

scrubbers was virtually 100 per cent. Often times maintenance was

performed on the scrubber during boiler outages but during the time

the scrubber was operated at basic pH the test scubber was not sited

as the limiting factor which caused the outage. As the result of

redundency in the system or modifing operating procedures the

scrubber could be kept in operation and emissions within limits

until the boiler was taken out of service.

To calculate the cost of the lack of availablity, the total

number of kWhr's a 500 MW facility would produce in a year was

multiplied by a factor which was the product of the per cent

downtime multiplied by the cost of a kWhr minus the cost of

73

consumables not used when the plant is non-operational. Assuming a

selling price of $0.03/kWhr and a consumables cost of $0.01/kWhr,

the net cost of down time to a utility would be $0.02/kWhr. For a

500 MW utility, the 15 per cent downtime for a limestone scrubber

would result in 12.0 million dollars in lost revenue over a year.

According to the data which was collected on the test scrubbers, a

soda ash scrubber would have an availablity of at least 92 per cent.

The lost revenue for a soda ash scrubber a year would then be 6.4

million dollars. Utility scrubbers are much larger than the test

scrubbers and it may be invalid to assume that a utility scrubber

could attain the availability of an industrial scrubber. Therefore,

these findings should not be considered conclusive.

Limestone vs Sodium Alkali Comparisons

As the data presented in this study indicates, all of the

sodium alkali materials tested worked well to remove SO2. At the

present time, there are no published prices for low grade materials

that were tested in this study. A price which v/ould make a sodium

system competitive with limestone systems can be calculated.

Table 30 is a summary of the annual cost that do not vary due

to the sulfur in the coal or to reagent prices. Additions have been

made to the annual power costs for the two sodium systems to

compensate for their higher availabity. These units would be on

line a greater percentage of the time than a limestone scrubber and

would consume more power.

74

Table 30

Annualized Costs Less Reagent Costs

^$10^/yr

Cost Limestone Sodium Alkali Sodium Alkali

(acidic) (basic)

Capital

Operating & Maint. Less Reagent (C +C -R ) ^ 0 m c

Availablity

Total Less Reagent

( t- c

10.1

5.1

12.0

27.2

7.0

4.3

6.4

>17.7

6.8

3.5

0.0

10.3

The total cost for a limestone system burning 1, 3, and

5 per cent sulfur coals are presented in Table 31.

Table 31

Total Annual Costs for Limestone Scrubber

$10Sr

Reagent Total Total Less Availablity

% Sulfur in Coal (R.) (C^) ^^t'^a'

1

3

5

0.8 28.0 16.0

2.4 29.6 17.6

4.1 31.3 19.3

To be competitive then, a sodium alkali system would have to

gave an annual cost less than C^ for a limestone unit. A

75

competitive price for sodium alkalies can be calculated by

substituting C^ of a limestone unit for the C^ of a sodium alkali

unit and solving for R^. Table 32 gives R^ for both types of sodium

alkali scrubbers that would make them competitive with a limestone

system with and without the benefit of increased availablity.

Table 32

Competitive Sodium Alkali Reagent Costs

$10^/yr

85% Availability Increased Availablity

% S in Coal Acidic Basic Acidic Basic

1 4.7 5.7 10.3 17.7

3 6.3 7.3 11.9 19.3

5 8.0 9.3 13.6 21.0

This data along with the data in Table 29 shows that situations

occur in which sodium alkali scrubbers would be more cost effective

than limestone scrubbers. Primarily the soda ash scrubbers would be

competitive when the delivered cost of the sodium alkali reagent is

below $75/ton and the power generation facility is burning 1 per

cent sulfur coal. Sodium alkali scrubbers could also be cost

effective on facilities burning 3 per cent sulfur coal provided the

avalliblity of limestone scrubbers remains low and a utility sodium

alkali scrubber performs as the test scrubber has to this point.

BBS

76

To illustrate the point, consider a 500 MW utility burning 3

per cent sulfur coal. Figure 4 shows that a basic sodium alkali

scrubber would be competitive if reagent costs were approximately

$70/ton or less. For the acidic sodium alkali scrubber to be

competitive the reagent cost would have to be less than $50/ton.

Where a power plant burns 1 per cent coal a sodium alkali scrubber

is more cost effective than limestone with reagent costs at over

$100/ton.

BfiS^SS^^

77

>•>

VD O

-be-

•4-> 0_)

to O C_J

o

60

50

40

30

20

10

0

60

50

40 I-

30

20

10

I 0 0

Limestone Sodium Alkali

Sodium Alkali, Acidic

Sodium Alkali, Basic

$100/ton

$ 75/ton

$ 50/ton

$100/ton

$ 75/ton

$ 50/ton

% Sulfur in Coal

Figure 4. Annual Costs as a Function of Percent Sulfur in Coal and Reagent Costs.

CHAPTER V

CONCLUSIONS AND RECOMMENDATIONS

This study has outlined a means to make SO2 removal less

costly.

The results of this study have shown that:

1. Low grade sodium alkali materials such as trona and

crystallizer purge liquor can be effectively used as an

SO2 scrubbing agent. The SO2 removal efficiency results

were very much the same for each of the materials tested.

In all cases the efficiencies varied from approximately 80

per cent at a pH of 5.8 and stoichiometeric ratio of 0.9

to 97 per cent at pH 6.8 and a stoichiometric ratio of

1.05.

2. Operating a sodium carbonate scrubber at an alkaline pH

reduces corrosion and can lower capital costs by

approximately 6 per cent. However, controlling the pH in

the scrubber in the basic range increases reagent costs.

It was found that a stoichiometric ratio of 1.1 was needed

to maintain the scrubber above pH 7.0.

3. Low grade sodium alkali can be cost competitive with

limestone wet scrubber systems when the sulfur in the coal

78

79

being burned is 3 per cent or less, and the cost of the

sodium alkali is less than $75/ton. These materials could

also be competitive if SOp regulations require greater

removal efficiencies than a limestone system can meet.

4. The main problem observed in operating a sodium alkali

scrubber at an alkaline pH is the formation of calcium

sulfite scale. It was found that this problem could be

controlled either by eliminating the source of calcium

ions or by chemical treatment.

5. Of the materials tested, crystallizer purge liquor was the

only one which demonstrated a significant operating

problem. Silica scale formed in the scrubber as a result

of using this material can also be controlled by chemical

treatment.

It is recommended that companies planning to build either

Industrial or utility scrubbers contact the soda ash producers and

negotiate a price for their low-grade sodium alkali and determine

which system, sodium alkali or limestone, would be the most cost

effective for their situation. A sodium alkali system could be

particularly attractive if a utility has a choice between a higher

priced low sulfur coal or a less expensive, higher sulfur coal and

SOp removal restrictions prohibit the use of a limestone scrubber.

Another situation where a sodium alkali scrubber is cost effective

would be where a plant is to be sited in Wyoming where scrubbers are

required and coal with less than 1 per cent sulfur is abundant.

80

Little information is currently available on the disposal of

sodium sulfite and sodium sulfate wastes produced by SO^ scrubbers.

These sodium salts are very soluble and could be leached into ground

water. More research needs to be done on developing methods of

containing sodium FGD wastes. The cost analysis of these methods

would be valuable in determining economic justification for proposed

sodium scrubbers.

More research is also needed to obtain realistic costs for low

grade sodium alkalies. The price at which these materials can be

sold v/111 be the most important factor in the economic evaluation of

a proposed scrubbing system.

LIST OF REFERENCES

1. Carnow, B.W.; Bouchard E. "Health Effects of SO. and SO,." Proceedings: Symposium on Flue Gas Desulfurization - ^ Las Vegas, March 1979. pages 21l47~

2. United States Environmental Protection Agency. The Acid Deposition Phenomenon and Its Effects. Vol. 1, 1983:

3. Gage, S. J. "Keynote Address for the 1980 EPA Symposium on Flue Gas Desulfurization." Proceedings: Symposium on Flue Gas Desulfurization - Houston", October 1980; Volume 1. page 11.

4. Wyoming Department of Environmental Quality,Air Quality Division.Wyoming Air Quality Standards and Regulations 1982, Cheyenne, Wyoming, pages 5, 6.

5. Smith, M.P.; Melia, M.T.; Laseke, B.A., Jr.; and Kaplan, N. "Recent Trends in Utility Flue Gas Desulfurization" Proceedings: Symposium on Flue Gas Desulfurization - Houston, October 1980, Vol 1, pages 143 - 157.

6. Stark,R. Environmental Engineer for Pacific Power and Light Co., Point of Rocks, Wyoming. Interview, 18 September 1981.

7. Leseke, B. A.; and Devitt, T. W. "Status of Flue Gas Desulfurization in the United States." Proceedings: Symposium on Flue Gas Desulfurization - Las Vegas, March 1979, pages 3,4.

8. Wark, K.; Warner, C. F. Air Pollution, Its Origin and Control. New York, Dun-Donnelley Publisher, pages 327 - 351.

9. Kaplan, N. "Summary of Utility Dual Alkali System" Proceeding: Symposium on Flue Gas Desulfurization - Las Vegas, March 1979.

10. Burbank, D.A.; Wang, S.C; and McKinsey, R.R. "Test Results on Adipic Acid - Enhanced Limestone Scrubbing at the EPA Shawnee Test Facility - Third Report." Proceedings: Symposium on Flue Gas Desulfurization - Las Vegas, March 1979.

11. Whittish, J. Steam Plant Foreman for Texasgulf Inc., Granger, Wyoming. Interview, October 1979.

12. Yantis, M. Sales Engineer for the Nalco Chemical Co. Interview, 15 May 1980.

81

82

^^' "nn^!l:J* -.i ^^"^el, E. R.; Piasecki, E. J.; Phillips, R. J. nni.hi M u History and Present Status of the General Motors uouDie Alkali SOp Control System" Proceedings: Symposium on j-iue Gas Desulfufization - Las Vegli, March 1979.

14. Stevens, N. J. "Dry SOp Scrubbing Pilot Test Results" Kroceedmqs: Symposiufri on Flue Gas Desulfurization - Houston, October 198U. paopc; llTZmv

15. Muela, C. A.; Menzies, W. R. "Stack Gas Reheat Benefits and Associated Energy Penalties." Proceedings: Symposium on Flue Gas Desulfurization-Las Vagas,March 1979;pages 1-16.

16. "The Corrosion Resistance of Nickel Containing Alloys in Flue Gas Desulfurization and Other Scrubbing Processes." The Internation Nickel Company Inc., One New York Plaza, New York 10004, 1978.

17. Rourke, E. T. "Power Plant Scrubber Using Special Stainless Steels is Under Construction in Wyoming." Nickel Topics, Vol. 32, No. 4, 1979, pages 7-12.

18. Hewitt, R. A. "Operating and Maintenance Experience of the World's Largest Spray Tower SOp Scrubbers" Proceedings: Symposium on Flue Gas Desulfurization - Houston, October 1980, page 7.

19. Tuttle, J.; Patkar, A. "The Status of Industrial Boiler Application in the United States." Proceedings: Symposium on Flue Gas Desulfurization-Las Vagas, March,1979.

20. McGlamery, G.G., O'Brien, W.E., Stephanson, C D . , and Veitch, J.D.; "FGD Economics in 1980." Proceedings: Symposium on Flue Gas Desulfurization - Houston, October 1980, pages 49 - 83.

21. Swemco Inc. Drawing Number D-4360-R-19, 1974 New York, New York.

22. American Public Health Association. Standard Methods for Examination of Water and Waste Water, Thirteenth Edition, Washington D.C. American Public Health Association Publica-tions, 1975.

23. Lowenhelm,F. A.; Moran, M. K. Industrial Chemicals, New York, John Wiley and Sons, 1975, page 709.

24. Welch, G. A.; Zlatkovich, C. T.; White, J. A. Intermediate Accounting, Homewood, ILL; Richard D. Irwin, Inc., pages 190, T9T: "

83

25. M m 2 ! ? ^ ^'M^- Chemistry and Technology of Lime and Limestone, New YorK, John Wiley and Sons, 1980, pages 122 and

^ * Fnwr^' ^"\^r Bo^niann, F. H. "Acid Rain: A Serious Regional tnvironmental Problem." Science, Vol. 184, 1974, page 1178.

27. United States Environmental Protection Agency. "Standards of Performance for New Stationary Sources, Title 40 , Chapter 1, Subpart D, 60.43." Fed. Reg. Vol. 36, No. 249, December 23,1971, page 24877.


Recommended