Sunteți pe pagina 1din 132

Optimization of Combustion in Boilers and Furnaces

Projects Documents Home

ARPEL Environmental Guide No. 29 Optimization of Combustion in Boilers and Furnaces June 30, 2000

Funding
This document has been exclusively prepared for the ARPEL Environmental Project Phase 2. The Project was funded by the Canadian International Development Agency (CIDA) and co-managed by the Environmental Services Association of Alberta (ESAA) and the Regional Association of Oil and Natural Gas Companies in Latin America and the Caribbean (ARPEL). Environmental Services Association of Alberta #1710, 10303 Jasper Avenue Edmonton, Alberta T5J 3N6 Tel: (780) 429-6363 Fax: (780) 429-4249 E-mail: info@esaa.org Regional Association of Oil and Natural Gas Companies in Latin America and the Caribbean Javier de Viana 2345 CP 11200 Montevideo, URUGUAY Tel.: (598-2) 410-6993 410-7454 Fax: (598-2) 410-9207 E-mail: arpel@arpel.org.uy

Copyright
The copyright in this document or product, whether in print or electronically stored on a CD or diskette or otherwise (the "Protected Work") is held by the Environmental Services Association of Alberta (ESAA). The Regional Association of Oil and Natural Gas Companies in Latin America and the Caribbean (ARPEL) has been granted a license to copy, distribute and reproduce this Protected Work on a cost-recovery and non-commercial basis. This Protected Work shall not, in whole or in part, be copied, photocopied, reproduced, translated or reproduced to any electronic means or machine-readable form without prior consent in writing from ESAA. Any copy of this Protected Work made under such consent must include this copyright notice.

Authors
These Guidelines have been prepared upon request of ARPEL and its Environment, Health and Safety Committee by: Optimum Energy Management Inc. 921 - 18 Avenue S.W. Calgary, Alberta T2T 0H2 Tel.: (403) 215-6580 Fax: (403) 215-6599 E-mail: briant@optimum.energy.com Consultants were assisted in detailed drafting and revision by the ARPEL Energy Efficiency Project Working Group.

Reviewers
Jaime F. George Jorge Velasco Urquiza Winston Charles Manuel Olivares Paez Miguel Moyano Oscar Gonzlez ECOPETROL CUPET PETROTRIN PEMEX ARPEL Executive Secretariat Environmental Services Association of Alberta

Disclaimer
Whilst every effort has been made to ensure the accuracy of the information contained in this publication, neither ARPEL, nor any of its member, nor ESAA, nor any of its member companies, nor CIDA, nor the consultants, will assume liability for any use made thereof.

The Optimization of Combustion in Boilers and Furnaces

ABSTRACT
This guideline has been prepared to assist the members of ARPEL in optimizing the operation of their fired heaters, boilers and incinerators. It is not intended to be an engineering design manual. Chapters 1 and 2 discuss the two principal components of combustion fuel and air. Data are supplied so that the readers can prepare heat balances around their fired equipment. Chapter 3 presents methods for estimating thermal efficiency, and where energy losses are being experienced. Examples are included in order to provide an indication of the impact of various energy management initiatives. Heater/boiler configuration and burner construction are extremely important factors to consider when evaluating fired equipment. Brief descriptions of the major classifications of heaters and burners are provided in Chapter 4 to illustrate the wide range of options that are available. Chapters 5, 6 and 7 demonstrate ways of improving the efficiency of heaters and boilers. Chapter 5 looks at the daily operation and discusses analyses, monitoring and firing practices that should be conducted as a part of normal operating procedures. Heater optimizations should not be confined to reducing the fuel consumed in order to supply heat to the process. Steps should be taken to minimize process heat requirements. With this in mind, Chapter 6 deals with ways to improve heat transfer using exchangers and options for reducing heat input into distillation columns, which ultimately leads to a reduction in heater duty. The chapter also lists energy management items concerning steam use in order to reduce fired boiler loads. Chapter 7 discusses opportunities for reducing flue gas temperatures by using waste heat recovery to preheat combustion air; boiler feedwater and circulating heat media. Heat sinks for cyclic operations are another option for capturing unused heat. Flue gas emissions are a major source of air contaminants. Chapter 8 lists factors so that ARPEL members can prepare emission inventories resulting from the operation of their fired heaters and boilers. Factors for emission-control technologies are also provided. Chapters 9, 10 and 11 describe options for controlling emissions by changing combustion (Chapter 9), by treating flue gas (Chapter 10) and by improving fuel quality (Chapter 11). The air contaminants of concern are NOx, SOx, particulates and carbon monoxide. The impact of switching from oil to gas-firing, of upgrading heavy fuel oil through blending and of treating fuel are presented. The guideline closes with a discussion on the feasibility of recovering flare gas for use as fuel. Note any mention made to manufacturers names or trademarks of equipment and/or processes in this document does not represent an endorsement neither by the authors nor by ARPEL.

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

TABLE OF CONTENTS
1.0 FUEL PROPERTIES..........................................................................................................1 1.1 1.2 1.3 2.0 Heating Values ..........................................................................................................1 Contaminants ............................................................................................................7 Flame Temperature ...................................................................................................8

AIR REQUIREMENTS .....................................................................................................13 2.1 Stoichiometric Air ....................................................................................................17

2.1.1 Excess Air ............................................................................................................19 2.2 2.3 3.0 Relationship Between Excess Oxygen and Excess Air............................................19 Oxygen Enrichment .................................................................................................21

HEATER EFFICIENCY ....................................................................................................23 3.1 3.2 3.3 3.4 Effect of Process Variables......................................................................................24 Energy Losses.........................................................................................................27 Boiler Efficiency.......................................................................................................30 Examples ................................................................................................................31

4.0

TYPES OF FIRED EQUIPMENT......................................................................................35 4.1 4.2 4.3 Direct-Fired Heaters ................................................................................................35 Fired Boilers ............................................................................................................37 Incinerators .............................................................................................................40

4.3.1 Incinerators for Process Wastes ...........................................................................40 4.3.2 Incinerators for Medical Wastes ...........................................................................41 4.4 4.5 4.6 5.0 Other Heaters..........................................................................................................41 Burners....................................................................................................................42 Design Criteria.........................................................................................................44

MONITORING PERFORMANCE .....................................................................................45

iii

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

5.1

Fuel Analysis ...........................................................................................................45

5.1.1 Wobbe Index ........................................................................................................45 5.2 5.3 6.0 Stack Gas Monitoring ..............................................................................................46 Firing Practices........................................................................................................47

ENERGY REDUCTION TECHNIQUES............................................................................57 6.1 Minimizing Heater Duties.........................................................................................57

6.1.1 Heat Exchange.....................................................................................................57 6.1.2 Effective Distillation ..............................................................................................58 6.2 6.3 7.0 Steam Systems .......................................................................................................64 Insulation.................................................................................................................66

ENERGY RECOVERY TECHNIQUES.............................................................................67 7.1 7.2 Modifying Radiant and Convection Sections............................................................68 Conversion to Forced-Draft Air Supply ....................................................................71

7.2.1 Forced-Draft Air Supply and Low-NOx Burners.....................................................71 7.3 Waste Heat Recovery..............................................................................................72

7.3.1 Combustion Air Preheat .......................................................................................73 7.3.2 Boiler Feedwater Preheat.....................................................................................75 7.3.3 Fuel Preheat.........................................................................................................75 7.3.4 Circulating-Liquid Heat Exchangers......................................................................75 7.3.5 Generation of Steam ............................................................................................76 7.3.6 Waste Heat Recovery from Incinerators ...............................................................76 7.4 8.0 Heat Sinks for Cyclic Operations .............................................................................76

EMISSIONS FROM HEATERS AND BOILERS...............................................................79 8.1 8.2 8.3 Emission Factors.....................................................................................................79 Effect of Heater Size on Emissions..........................................................................82 Effect of Controls on Emissions...............................................................................83

ARPEL Environmental Guideline No. 29

iv

The Optimization of Combustion in Boilers and Furnaces

9.0

CONTROLLING EMISSIONS COMBUSTION CONTROLS .........................................85 9.1 9.2 9.3 Fuel Switching .........................................................................................................85 Excess Air vs. CO, NOx, Efficiency ..........................................................................86 NOx Abatement .......................................................................................................88

9.3.1 Fuel NOx ...............................................................................................................89 9.3.2 Thermal NOx.........................................................................................................89 9.3.3 Flue Gas Recirculation .........................................................................................89 9.3.4 Staged-Burners and Combustion..........................................................................91 9.3.5 Low NOx Burners..................................................................................................94 9.3.6 Diluent Injection....................................................................................................96 10.0 CONTROLLING EMISSIONS-POST-COMBUSTION CONTROLS .................................99 10.1 NOx ........................................................................................................................99 10.1.1 NOx Abatement Considerations .................................................................... 101

10.2 SOx ...................................................................................................................... 103 10.2.1 SOx and NOx ................................................................................................. 105

10.3 Particulates ........................................................................................................... 106 11.0 FUEL IMPROVEMENT .................................................................................................. 107 11.1 Sulfur Removal...................................................................................................... 107 11.1.1 11.1.2 11.1.3 Blending ....................................................................................................... 107 Hydrotreating ................................................................................................ 108 Fuel Gas Sweetening ................................................................................... 110

11.2 Flare Gas Recovery............................................................................................... 110 12.0 BIBLIOGRAPHY............................................................................................................ 113

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

LIST OF TABLES
TABLE 1.1: HEATS OF COMBUSTION AT 25C OF LIGHT HYDROCARBONS, GJ/TONNE ..2 TABLE 1.2: HEATS OF COMBUSTION OF LIQUID FUELS......................................................2 TABLE 1.3 ANALYSES OF COALS .........................................................................................7 TABLE 1.4: ADIABATIC FLAME TEMPERATURES OF COMMON LIGHT TABLE 1.5 CALCULATED FLAME TEMPERATURE OF COMMON LIGHT FUELS,F ........9 FUELS,F.......10

TABLE 2.1: COMPOSITION OF DRY AIR ...............................................................................13 TABLE 2.2: ENTHALPY OF DRY AIR .....................................................................................13 TABLE 2.3: RELATIONSHIP BETWEEN ELEVATION AND BAROMETRIC PRESSURE ......16 TABLE 2.4: ENTHALPY OF WATER VAPOR..........................................................................17 TABLE 2.5: STOICHIOMETRIC AIR REQUIREMENTS FOR GASEOUS FUELS ...................18 TABLE 3.1: ENTHALPY VALUES OF FUELS .........................................................................25 TABLE 3.2: ENTHALPY OF ATOMIZING STEAM ...................................................................26 TABLE 3.3: BOILER BLOWDOWN VS. LOST BOILER EFFICIENCY .....................................30 TABLE 3.4: ENTHALPY OF FLUE GAS COMPONENTS ........................................................33 TABLE 4.1: TYPICAL BATH HEATERS ..................................................................................42 TABLE 4.2: BURNER SPECIFICATIONS ................................................................................43 TABLE 5.1: RECOMMENDED EXCESS AIR LEVELS, % .......................................................48 TABLE 6.1: TYPICAL TRAY EFFICIENCIES...........................................................................62 TABLE 7.1: HEAT FLUX RATES AND MASS VELOCITIES FOR REFINERY PROCESS HEATERS ............................................................................................................68 TABLE 8.1: EMISSION FACTORS FOR FIRED HEATERS (NATURAL GAS), LBS/MMBTU..79

ARPEL Environmental Guideline No. 29

vi

The Optimization of Combustion in Boilers and Furnaces

TABLE 8.2: EMISSION FACTORS FOR FIRED HEATERS (REFINERY GAS), LBS/MMBTU 80 TABLE 8.3: EMISSION FACTORS FOR AIR CONTAMINANTS FROM UNCONTROLLED RESIDUAL OIL COMBUSTION............................................................................81 TABLE 8.4: EFFECTIVENESS OF NOX CONTROL MEASURES FOR RESIDUAL FUEL COMBUSTION.....................................................................................................83 TABLE 8.5: POST-COMBUSTION SO2 CONTROLS FOR RESIDUAL OIL COMBUSTION ....84 TABLE 8.6: REMOVAL EFFICIENCY OF TECHNOLOGIES FOR CONTROLLING EMISSIONS OF PARTICULATES........................................................................84 TABLE 9.1: EFFECT OF FUEL SWITCHING ON HEATER EMISSIONS, LBS/HOUR ............86 TABLE 11.1: EFFECT OF FUEL OIL BLENDING ON EMMISSIONS, LBS/DAY ................... 108 TABLE 11.2: EFFECT OF FUEL OIL HYDROTREATING ON EMISSIONS, LBS/DAY.......... 110

vii

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

LIST OF FIGURES
FIGURE 1.1: BASE CASE HEATS OF COMBUSTION OF LIQUID PETROLEUM FUELS........3 FIGURE 1.2: BASE CASE SULFUR CONTENT IN LIQUID FUELS ..........................................4 FIGURE 1.3: BASE CASE CONTENT OF INERT MATERIAL IN LIQUID FUELS .....................4 FIGURE 1.4: CARBON TO HYDROGEN WEIGHT RATIO........................................................5 FIGURE 1.5: EFFECT OF EXCESS AIR AND AIR TEMPERATURE UPON ADIABATIC FLAME TEMPERATURE .....................................................................................9 FIGURE 1.6: DISSOCIATION OF CARBON DIOXIDE AND WATER VAPOR .........................11 FIGURE 2.1: WATER VAPOR CONTENT IN AIR, LBS OF WATER PER LB OF DRY AIR.....14 FIGURE 2.2: WET-BULB TEMPERATURE AS A FUNCTION OF AIR TEMPERATURE AND RELATIVE HUMIDITY .......................................................................................15 FIGURE 2.3: WATER CONTENT OF AIR-CORRECTION FACTOR FOR BAROMETRIC PRESSURE .......................................................................................................16 FIGURE 2.4: EXCESS AIR AS A FUNCTION OF OXYGEN IN THE FLUE GAS.....................21 FIGURE 3.1: THERMAL EFFICIENCY OF A GAS-FIRED HEATER........................................28 FIGURE 3.2: STACK TEMPERATURE AS A FUNCTION OF FLUE GAS FLOW RATE.......28

FIGURE 3.3: RADIATION AND CONVECTIVE HEAT LOSSES ..............................................29 FIGURE 4.1: TYPICAL DIRECT FIRED PROCESS HEATERS...............................................36 FIGURE 4.2: COMMON BOILER CONFIGURATIONS............................................................39 FIGURE 4.3: PERFORMANCE OF A PACKAGED BOILER ....................................................40 FIGURE 5.1: TYPICAL DRAFT PROFILE IN A DIRECT FIRED HEATER...............................50 FIGURE 5.2: EMISSIONS OF HYDROGEN AND CARBON MONOXIDE UNDER SUBSTOICHIOMETRIC CONDITIONS..............................................................50

ix

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

FIGURE 5.3: THERMAL EFFICIENCY (LHV) FOR A HEATER FIRING NATURAL GAS ........51 FIGURE 5.4: ENERGY LOSSES WHEN BURNING NATURAL GAS ......................................52 FIGURE 5.5: ENERGY LOSSES WHEN BURNING NO. 6 OIL ...............................................52 FIGURE 5.6: EFFECT OF SOOT DEPOSITS ON FUEL COMBUSTION.................................54 FIGURE 5.7: EFFECT OF SCALE DEPOSITS ON FUEL COMBUSTION ...............................55 FIGURE 5.8: AIR LEAKS INTO HEATERS..............................................................................55 FIGURE 6.1: EFFECT OF PRESSURE ON REBOILER DUTY ...............................................61 FIGURE 7.1: HEAT TRANSFER IN THE RADIANT SECTION OF DIRECT FIRED HEATER .70 FIGURE 7.2: ACID GAS DEW POINT AS A FUNCTION OF FUEL SULFUR/H2S CONTENT .73 FIGURE 7.3: SCHEMATIC ARRANGEMENT OF AN AIR PREHEATER.................................75 FIGURE 9.1: CARBON MONOXIDE EMISSIONS FROM HEATERS ......................................87 FIGURE 9.2: EFFECT OF OXYGEN ON HEATER EMISSIONS .............................................88 FIGURE 9.3: NOX EMISSIONS FROM HEATERS...................................................................88 FIGURE 9.4: EFFECT OF FLUE GAS RECIRCULATION ON NOX EMISSIONS.....................90 FIGURE 9.5: STAGED BURNERS ..........................................................................................92 FIGURE 9.6: EFFECT OF LOW-NOX BURNERS ON NOX EMISSIONS (GAS FIRING) ..........95 FIGURE 9.7: EFFECT OF NITROGEN CONTENT IN FUEL OIL ON NOX EMISSIONS USING STANDARD, HIGH INTENSITY / LOW NOX HIGH INTENSITY BURNERS.......95 FIGURE 9.8: EFFECT OF AIR PREHEAT ON NOX EMISSIONS (OIL FIRING) USING STANDARD HIGH INTENSITY AND LOW NOX HIGH INTENSITY BURNERS .96 FIGURE 9.9: EFFECT OF STEAM INJECTION ON NOX EMISSIONS ....................................97 FIGURE 10.1: NOX ABATEMENT TECHNOLOGIES............................................................. 102

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

FIGURE 11.1: HYPOTHETICAL FLARING PATTERNS ........................................................ 111

xi

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

1.0

FUEL PROPERTIES

The overwhelming majority of fired boilers and heaters, or furnaces, that are operated by the oil and gas industry use gas or heavy fuel oil. In a few cases, crude oil is burned as a fuel and in still fewer cases, coal is burned. The properties of these various fuels will influence the operation of the fired heaters and boilers and the emissions resulting from combustion.

1.1

Heating Values
During complete combustion, hydrocarbons will be converted to carbon dioxide and water. Heat will be released. For example, when methane is burned, the following reaction occurs: CH4 + 2O2 CO2 + 2H2O + Heat The heat of combustion represents the heat of formation of the products of the reaction. The amount will depend upon whether the water produced by the combustion is in the form of water vapor or liquid water. In the first case, the amount of heat released at 77F (25C) and constant pressure is 21,502 BTU/lb of methane. In the second case, the amount of heat is 23,861 BTU/lb. The smaller value (21,502 BTU/lb) is called the low heating value (LHV) or net heating value and the greater value (23,861 BTU/lb) is called the high heating value (HHV) or gross heating value. The difference between the two heating values (2,359 BTU/lb) is the heat of vaporization of the water formed during the combustion of one pound mass of methane. It is extremely important to be consistent when using heating values. They have a profound effect upon heater and boiler efficiency (see Chapter 3). By convention Canada and the United States sell natural gas in terms of its high heating value. Much of the rest of the world uses low heating value. Table 1.1 lists high and low heating values of a number of gases typically found in natural gas and refinery fuel gas streams. Note that the heat of combustion of propane and heavier compounds will depend upon whether the material is in gaseous or liquid form. Also there are no high heating values for the combustion of carbon and carbon monoxide because there is no hydrogen present to form water. Compounds sometimes found in gas streams, such as carbon dioxide, nitrogen, helium and oxygen have no heating value and have been omitted from Table 1.1. On the other hand, these compounds do help reduce the efficiency of the heater/boiler, albeit in a relatively minor way, by requiring enthalpy that otherwise would be absorbed by process fluid being heated. Liquid fuels can be classified as either distillates or residuals. Table 1.2 lists heats of combustion of these fuels.

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Table 1.1: HEATS OF COMBUSTION AT 25C OF LIGHT HYDROCARBONS, GJ/TONNE HHV Methane (g) Ethane (g) Propane (g) Propane (l) n-Butane (g) n-Butane (l) iso-Butane (g) iso-Butane (l) n-Pentane (g) n-Pentane (l) iso-Pentane (g) iso-Pentane (l) n-Hexane (g) n-Hexane (l) 55.5007 51.8791 50.3486 49.9857 49.5275 49.1577 49.4089 49.0693 49.0135 48.6460 48.9042 48.5599 48.6785 48.3134 g LHV 50.0137 47.4876 46.3549 45.9943 45.7408 45.3710 45.6222 45.2826 45.3547 45.1942 45.2430 44.8988 45.1035 44.7360 gas l liquid s solid
2

HHV n-Heptane (g) n-Heptane (l) Ethane (g) Propylene (g) 1-Butene (g) cis-2-Butene (g) Trans-2-Butene (g) iso-Butene (g) Hydrogen (g) Carbon (s) Carbon monoxide (g) Hydrogen sulfide (g) 48.4390 48.0738 50.2998 48.9204 48.4576 48.5436 48.2575 48.1854 141.7876 ----16.5076

LHV 44.9244 44.5592 47.1620 45.7827 45.3198 45.1965 45.1174 45.0476 119.9551 32.7659 10.1032 15.2051

Table 1.2: HEATS OF COMBUSTION OF LIQUID FUELS Fuel No. Type Density
3

Heat of Combustion GJ/m

1 2 4 5 6

Distillate Distillate Residual Residual Residual

0.82-0.85 0.86-0.90 0.90-0.91 0.92-0.95 0.96-0.97

37.91-38.74 38.91-40.28 40.47-40.64 40.98-41.70 42.07-42.43

No. 3 oil was a distillate fuel but has been superseded by No. 2 fuel oil. Viscosity and sulfur content are important factors when using liquid fuels. Even though Nos. 4, 5 and 6 are residual oils, they frequently have distillate cutterstock added in order to meet the sulfur and viscosity specifications.
1 2

All except H2S from Perry; page 3-142. H2S data from North American Combustion Handbook, page 8. Perry; page 9-6.

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

The American Petroleum Institute has issued graphs for estimating the heating value of liquid fuels. Figure 1.1 shows how heating value is a function of the density of the fuel. These values are typical. That is, they presuppose typical contents of sulfur, inert material and water. These contaminants reduce the heating value of the fuel oil. Note that the curves for densities of 0.93 and higher, are average values of cracked and virgin fuel oils and crude oils. For densities less than 0.93, the data are for crude oils only. Figures 1.2 and 1.3 show typical levels of sulfur and inert material in liquid fuels. Figure 1.4 lists the carbon-to-hydrogen weight ratio. This latter graph is useful for determining stoichiometric air requirements. Correction for sulfur and inert material content in fuels with a density of 0.825 or less is negligible and the curves in Figure 1.1 at those low densities represent pure petroleum liquids.1 Note that there are no curves for typical water content. The curves in Figure 1.1 assume no water in the fuel.
Figure 1.1: BASE CASE HEATS OF COMBUSTION OF LIQUID PETROLEUM FUELS
20500

20000

High Heating Value


19500

Heating Value, BTU/lb

19000

18500

Low Heating Value


18000

17500

17000

16500 0.70

0.75

0.80

0.85

0.90

0.95

1.00

1.05

1.10

Density of Fuel, kg/litre

1 2

Maxwell; page 180. API Technical Data Book, Figure 14.A1.1.

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Figure 1.2: BASE CASE SULFUR CONTENT IN LIQUID FUELS


3.5

3.0

2.5

Weight % Sulphur

2.0

1.5

1.0

0.5

0.0

0.80

0.85

0.90

0.95

1.00

1.05

1.10

Density of Fuel, kg/litre

Figure 1.3: BASE CASE CONTENT OF INERT MATERIAL IN LIQUID FUELS


1.2

1.1

Weight % Inert Material

1.0

0.9

0.8

0.7

0.6 0.80

0.85

0.90

0.95

1.00

1.05

1.10

Density of Fuel, kg/litre

1 2

API Technical Data Book, Table 14-0.1. API Technical Data Book; Table 14-0.1.

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Figure 1.4: CARBON TO HYDROGEN WEIGHT RATIO


9.0

8.5

C:H Weight Ratio

8.0

7.5

7.0

6.5

6.0 0.80

0.85

0.90

0.95

1.00

1.05

1.10

Density of Fuel, kg/litre

In most cases, the liquid fuel burned at a facility does not have the characteristics listed in Figures 1.2 and 1.3. The following equations can be used to adjust the heating values provided in Figure 1.1. High Heating Value2 H = H0 0.01 * H0 * (% H2O + % S + % Inerts) + 40.5 * % S where: H H0 % H2O %S % Inerts = = = = = Adjusted high heating value, in BTU/lb of fuel High heating value as provided in Figure 1.1, in BTU/lb of fuel Weight % water in the fuel Weight % sulfur in the fuel minus the base case weight % sulfur given in Figure 1.2 Weight % inert material in the fuel minus the base case weight % inert material given in Figure 1.3

1 2

API Technical Data Book, Table 14-0.1. API Technical Data Book, page 14-4.

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Low Heating Value1 L = L0 0.01 * L0 * (% H2O + % S + % Inerts) + 40.5 * % S 10.53 * % H2O

where: L L0 % H2O %S % Inerts = = = = = Adjusted low heating value, in BTU/lb of fuel Low heating value as provided in Figure 1.1, in BTU/lb of fuel Weight % water in the fuel Weight % sulfur in the fuel minus the base case weight % sulfur given in Figure 1.2 Weight % sulfur in the fuel minus the base case weight % sulfur given in Figure 1.2

As an example, consider a fuel oil of density 0.95, with 1.50 weight % sulfur, 1.00 weight % inert material and 0.25 weight % water. From Figure 1.1, the high heating value is 18,710 BTU/lb and the low heating value is 17,655 BTU/lb. The base case heats of combustion, or heating values, in Figure 1.1 assume no water in the fuel. Therefore, the actual heat of combustion must be adjusted for the entire amount of water in the fuel in this example, 0.25 weight %. The actual sulfur content of the fuel is 1.50 weight % and the base case content, from Figure 1.2 is 1.20 weight %. The difference, % S, is 1.50 1.20 = 0.30 wt %. Similarly, the actual content of inert material is 1.00 weight % and the base case, from Figure 1.3, is 0.80 weight %. Thus, % Inerts = 1.00 0.80 = 0.20 weight % H = 18,710 0.01 * 18,710 * (0.25 + 0.30 + 0.20) + 40.5 * 0.30 = 18,582 BTU/lb L = 17,655 0.01 * 17,655 * (0.25 + 0.30 + 0.20) + 40.5 * 0.30 10.53 * 0.25 = 17,532 BTU/lb The heating value of coal is quite dependent upon the type of coal and its location. The following values in Table 1.3 are for coals mined in the USA. In the absence of sitespecific data, use those in Table 1.3.

API Technical Data Book, page 14-4

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Table 1.3 C Anthracite Anthracite 86.7 80.9 H 2.2 3.3

ANALYSES OF COALS O 2.9 4.2 S 0.5 0.5 N 0.8 1.0

Moisture 2.3 2.1

Ash 6.9 10.1

BTU/lb 13,540 13,480

Bituminous

84.0 76.6 79.2 68.4 71.5 62.8 63.4

4.8 5.2 5.7 5.6 5.8 5.9 5.7

5.6 6.2 10.0 16.4 14.3 17.4 18.6

0.6 1.3 0.6 1.2 2.6 4.3 2.3

1.1 1.6 1.5 1.4 1.6 1.0 1.3

3.5 2.6 3.1 8.2 7.2 12.1 12.4

3.9 9.1 3.0 7.0 4.2 8.6 8.7

14,550 13,610 14,290 12,160 12,950 11,480 11,420

Subbituminous

54.6

6.4

33.8

0.4

1.0

23.2

3.8

9,420

Lignite

42.4

6.7

43.3

0.7

1.7

34.8

6.2

7,210

In Table 1.3 the contents of moisture and ash are from the proximate analyses, whereas the values for carbon, hydrogen, oxygen, sulfur and nitrogen are from the ultimate analyses. All values are in weight %.

1.2

Contaminants
The impact of a number of contaminants on fuel quality has been discussed in the previous section. Strictly speaking, carbon monoxide and hydrogen sulfide are not contaminants because they do have heating values, although they are very low compared with the hydrocarbon components of the fuel. Using the criterion from the previous paragraph, the carbon dioxide, nitrogen, helium, and inert materials in fuels are contaminants because they provide no heat. In fact they reduce the efficiency of the fired equipment by absorbing a portion of the heat that is released during combustion. Wherever possible, site-specific analyses should be used to determine more accurate estimates of the heats of combustion of the facilitys fuel(s). The appropriate ASTM test procedures (or equivalent) should be used to determine the levels of fuel contaminants.

Perry; page 9-3.

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Combustion will generate acid gases, which leave the heater/boiler via the flue gas. The most obvious of these are sulfurous and sulfuric acid and nitric acid, resulting from the burning of sulfur and nitrogen compounds in the fuel. However, there could also be hydrochloric and hydrobromic acids. It is extremely important that the facility know the analyses of its flue gas streams. These data can be used to estimate the acid gas dew point temperatures. These temperatures determine the maximum amount of waste heat recovery that is possible. To reduce stack exit temperatures below the dew point will lead to condensation of the acid gas and therefore, corrosion of the stack metallurgy. In practice, the minimum stack temperature should be set higher than the actual dew point in order to avoid localized corrosion. As an example, one manufacturer of waste heat recovery equipment sets a minimum stack temperature of 177C (350F), even though the dew point of the fuel gas is in the range of 93-149C (200-300F).1 For a procedure for estimating acid gas dew points, the reader is referred to the article Compute Dew Point of Acid Gases by V. Ganapathy, Hydrocarbon Processing, February 1993, page 93. In the meantime, as a rule of thumb, gaseous fuels normally have acid gas dew points below 150C (302F). One site, burning crude oil with approximately 6 weight % sulfur, had an acid gas dew point of 230C (446F).

1.3

Flame Temperature
Each fuel and its constituents will generate a flame temperature. If there were no heat loss to the surroundings and no dissociation of the products of combustion, the flame temperature would be at its maximum. All the enthalpy entering the burner in the form of sensible heat of the combustion air, the sensible heat of the fuel and the heat of combustion would be transferred to the products of combustion. The result is the theoretical flame temperature, or the adiabatic flame temperature (AFT). Table 1.4 lists reported values of adiabatic flame temperatures. The AFT on a heater or boiler is an important parameter for the facility to know because it enables the facility to more accurately estimate emissions of NOx.

Maxim Heat Recovery Application Manual; page 7.

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Table 1.4: ADIABATIC FLAME TEMPERATURES OF COMMON LIGHT FUELS ,F Kunz et al Hydrogen Methane Propane Carbon Monoxide Natural gas H2 Plant PSA gas 4,056 3,698 3,818 4,311 3,700 3,273

The adiabatic flame temperature (AFT) is a function of the excess air rate, the gas composition and the inlet conditions of the air and fuel. It can be calculated assuming that the enthalpy of the products of combustion equals the sum of the enthalpy of the incoming air and fuel plus the heat of combustion. It is also assumed that kinetic and potential energy terms are negligible and that there is no shaft work. Calculation of the AFT is a trial and error process. The heat capacity of the product streams from the combustion temperature (frequently the heat of combustion is given at 25C) to the AFT is a function of the final temperature, i.e., the estimated AFT.2 As an example of the effect of excess air conditions upon the AFT, see Figure 1.5.
Figure 1.5: EFFECT OF EXCESS AIR AND AIR TEMPERATURE UPON ADIABATIC FLAME 3 TEMPERATURE

4200

1400C
4000

1100C 900C

3800

AFT, F

3600

60F
3400

3200

3000

2800 60 70 80 90 100 110 120 130 140 150

% o f S t o i c h i o m e tric Air

The adiabatic flame temperature is an important variable for two reasons. There is a direct relationship between the production of NOx and the excess oxygen and the AFT.
1 2

Kunz et al, Hydrocarbon Processing; November 1996; page 66. Smith and Van Ness; page 147. 3 North American Combustion Handbook; page 10.

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Kunz et al1 generated the following formulae from their own work with steam-methane reformers (Formulae 1 and 2) and another study in the literature (Formula 3). In all cases, the x-axis is 10,000 divided by the adiabatic flame temperature in Rankine. The y-axis is the natural logarithm of the concentration of NOx in ppm divided by the concentration of oxygen in %. In summary, x = 10,000 / AFT, R y = ln (NOx, ppm / O2, %)
Low-NOx Burners y= y= y= 12.2-3.60x 12.6-3.58x 21.0-6.43x Conventional General (1) (2) (3)

Secondly, high flame temperatures also cause a phenomenon called dissociation. In this process, the products of combustion break down in a form of reverse combustion: Heat + CO2 CO + O Heat + H2O H2 + O The heat originally liberated in combusting carbon monoxide and hydrogen is now reabsorbed. The net result is that an equilibrium is reached, usually at temperatures in the range of 3,400-3,800F (1,870-2,090C). The equilibrium temperature is also known as the calculated flame temperature. It is also a theoretical value because the effect of heat removal and losses has not been taken into account. Calculated flame temperatures (with air at standard conditions, 100% of stoichiometric airflow and taking into consideration dissociation) are listed in Table 1.5. Because heaters and boilers typically operate with excess air, the values in Table 1.5 are maximum values.
Table 1.5: CALCULATED FLAME TEMPERATURE OF COMMON LIGHT FUELS,F Hydrogen Carbon monoxide Methane Ethane 4,010 3,542 3,484 3,540 Propane n-Butane Natural Gas 3,573 3,583 3,525
2

Dissociation is relatively minor at flame temperatures less than 3000F (1649C). However, as indicated in Figure 1.6, the amount of dissociation, particularly for CO2 climbs rapidly. When estimating the enthalpy of the exiting flue gas, it is necessary to take into account dissociation of the products of combustion.
1 2

Kunz et al; pages 74, 76. North American Combustion Handbook; page 12.

ARPEL Environmental Guideline No. 29

10

The Optimization of Combustion in Boilers and Furnaces


1

Figure 1.6: DISSOCIATION OF CARBON DIOXIDE AND WATER VAPOR


70

60

Amount of Dissociation, %

50

40

Pressure at 1 atm.
30

CO2
20

H2O
10

0 3000

3500

4000

4500

5000

Temperature, F

Perry; page 9-40.

11

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

2.0

AIR REQUIREMENTS

Before discussing the effect that air has upon combustion and heater/boiler efficiency it is necessary to discuss several properties of air. The composition of air is listed in Table 2.1.
Table 2.1: COMPOSITION OF DRY AIR 78.03 volume % 75.46 weight % 20.99 0.94 0.03 0.01 Trace 23.20 1.30 0.04 Trace Trace
1

Nitrogen Oxygen Argon

Carbon dioxide Hydrogen Others

The molecular weight is 28.964 and the density at standard temperature and pressure is 1.22 kg/m3 or 0.0763 lbs/ft3. The enthalpy of dry air is given in Table 2-2. A temperature of 60F has been chosen as enthalpy = zero.
Table 2.2: ENTHALPY OF DRY AIR Temp. F 0 20 40 60 80 100 120 140 160 Enthalpy BTU/lb -14.37 -9.58 -4.79 0.000 4.863 9.719 14.58 19.44 24.30 Temp. F 180 200 250 300 400 500 600 700 800 Enthalpy BTU/lb 29.16 34.03 46.21 58.43 83.05 107.9 129.5 158.4 184.1
2

Temp. F 900 1,000 1,100 1,200 1,300 1,400 1,500 1,600

EnthalpyB TU/lb 210.2 231.8 256.2 280.5 305.4 330.2 355.5 380.7

1 2

North American Combustion Handbook; page 1. Ibid; page 344 for temperatures above 60F and Energy Management Handbook; page 567 for temperatures below 60F.

13

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Air will contain moisture. The amount will depend upon the temperature and the relative humidity. Figure 2.1 provides data for air at 29.92 Hg (1 atmosphere) air pressure. At different barometric pressures the water content of air will change. To determine this effect, use the following procedure. 1. Determine the wet-bulb temperature. This can be found by knowing the air temperature (sometimes called the dry-bulb temperature) and the relative humidity. See Figure 2.2. 2. With the wet-bulb temperature and the barometric pressure, determine the moisture content additive correction factor for air saturated at the wet-bulb temperature when the barometric pressure differs from the standard pressure of 29.92 Hg (760 mm Hg). See Figure 2.3. If the barometric pressure is not known, it can be estimated by relating it to the elevation. See Table 2.3. However, this method is only approximate. Barometric pressure can vary from day to day and from season to season. 3. Calculate the difference between the dry-bulb (air) temperature and the wet-bulb temperature, i.e., td-tw. 4. Adjust the moisture content additive correction factor (ACF) by 1% if td-tw equals 24F. If tdtw does not equal 24F, adjust proportionately, using the formula: [(td-tw) / 24] * 0.01 * ACF 5. The moisture content at barometric pressure p is estimated using the formula WP = W29.92 + ACF [((td tw) / 24) * 0.01 * ACF]
Figure 2.1: WATER VAPOR CONTENT IN AIR, lbs of water per lb of dry air
0.06

100%

0.05

Parameter is Relative Humidity, %


0.04

80%

lb Water/lb Dry Air

60%
0.03

40%
0.02

20%
0.01

0.00 10 20 30 40 50 60 70 80 90 100 110 120

Air Temperature, F

Perry; page 15-3.

ARPEL Environmental Guideline No. 29

14

The Optimization of Combustion in Boilers and Furnaces

As an example consider the following: The air temperature is 90F and the relative humidity is 80%. The barometric pressure is measured at 27.92 Hg. From Figure 2-1, the water content of air at a barometric pressure of 29.92 Hg is 0.025 lbs water/lb of dry air. The dry-bulb temperature is the same as the air temperature measured by a conventional thermometer. Using Figure 2.2, the wet-bulb temperature corresponding to an air temperature of 90F and a relative humidity of 80% is 85F. The moisture content additive correction factor for air saturated at the wet-bulb temperature is read from Figure 2.3. For a wet-bulb temperature of 85F and a barometric pressure of 27.92, the correction factor is 0.002 lbs water/lb of dry air. This factor must be adjusted. The difference between the dry-bulb and wet-bulb temperatures is 90-85 = 5F. The adjustment is therefore, 5/24 * .01 * 0.002 = 0.0000042 lbs water/lb dry air. The effect of reduced barometric pressure is to raise the water content of the air by 0.002 0.0000042 = 0.001996 lbs/lb dry air. The water content is therefore, 0.025 + 0.001996 = 0.027 lbs/lb dry air.
Figure 2.2: WET-BULB TEMPERATURE AS A FUNCTION OF AIR TEMPERATURE AND RELATIVE 1 HUMIDITY
120

100% Parameter is Relative Humidity


100

80% 60%

Wet-Bulb Temperature, F

40%
80

20% 0%

60

40

20

0 10 30 50 70 90 110

Air Temperature, F

Perry; pages 15-3 to 15-5.

15

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Table 2.3: RELATIONSHIP BETWEEN ELEVATION AND BAROMETRIC PRESSURE Barometric Pressure Inches Hg 29.92 28.92 27.92 26.92 25.92 24.92 23.92 Feet 0 900 1,800 2,700 3,700 4,800 5,900 Elevation Metres 0 274 549 823 1,128 1,463 1,798

The enthalpy values of water vapor are listed in Table 2.4. They are adapted from Steam Tables and have been adjusted to an enthalpy value of zero at 60F.
Figure 2.3: WATER CONTENT OF AIR-CORRECTION FACTOR FOR BAROMETRIC PRESSURE
0.02

Parameter is Barometric Pressure, inches Hg


0.016

23.92"

24.92" Additive Correction Factor, lb Water / lb Dry Air


0.012

25.92" 26.92"

0.008

27.92"
0.004

28.92"
0 0 20 40 60 80 100

29.92"
120 140

30.92"
-0.004

Wet-Bulb Temperature, F

1 2

Perry; page 15-8. Perry; page 15-8.

ARPEL Environmental Guideline No. 29

16

The Optimization of Combustion in Boilers and Furnaces


1

Table 2.4: ENTHALPY OF WATER VAPOR Temperature, F Temperature, F Enthalpy, BTU/lb Enthalpy, BTU/lb 0 20 40 60 80 100 120 140 160 180 200 212 250 300 350 -26.1 -17.3 -8.6 0.0 8.6 17.2 26.0 34.4 42.7 50.9 58.6 63.2 82.0 104.8 128.2 400 450 500 600 700 800 900 1,000 1,100 1,200 1,300 1,400 1,500 1,600 151.7 174.9 198.2 245.8 294.2 343.3 393.2 443.9 495.5 547.9 601.8 655.9 711.6 766.4

The effect of humidity upon heater efficiency is significant. As indicated in the example, the water content in air can amount to 2.7 weight %, and in some cases even higher. The amount of heat absorbed by water vapor in the air could exceed 1-2% of the heat of combustion in some heaters. Another way of looking at this is to say that heating the humidity in the air accounts for as much as 5% of the energy losses from a heater or boiler.

2.1

Stoichiometric Air
Stoichiometric air means the amount of air theoretically required to completely burn a fuel. This amount is usually referred to in terms of volume of air per volume of fuel. Both volumes must be at the same temperature and pressure. The type of fuel is a determinant factor. See Table 2.5

Combustion Engineering Steam Tables for temperatures up to 1200F. For higher temperatures, adapted from API Technical Data Book; page 14-22 and Energy Management Handbook; page 573.

17

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Table 2.5: STOICHIOMETRIC AIR REQUIREMENTS FOR GASEOUS FUELS Compound Stoichiometric Air, volume/volume 2.39 2.39 7.16 3.58 -4.78 9.54 16.70 14.32 23.86 Compound Stoichiometric Air, volume/volume 21.48 31.02 28.63 38.18 35.79 45.34 52.50 59.65

Hydrogen Carbon Monoxide Hydrogen Sulfide Ammonia Oxygen Methane Ethane Ethylene Propane

Propylene Butanes Butylenes Pentanes Pentenes Hexanes Heptanes Octanes

Stoichiometric air requirements for mixtures of these gases can be calculated using the gas composition. For example, consider a fuel gas with 85% methane, 10% ethane and 5% propane. The stoichiometric air requirement is: 0.85 * 9.54 + 0.10 * 16.70 + 0.05 * 23.86 = 10.97 volumes of air/volume of gas Note that if the gas contains oxygen the amount of air required decreases. One must be careful to ensure that oxygen measured in the sample is not due to air entrapment while taking the sample. Gases such as nitrogen, carbon dioxide sulfur dioxide and helium do not require combustion air. Stoichiometric air requirements are often listed in terms of volume of air/unit of mass of liquid or solid fuel. An example is SCF air/lb of fuel. It is necessary to know the ultimate analysis of the fuel, i.e., the amount of carbon, hydrogen, sulfur and oxygen. The formula2 is: SCF (60F, 14.7 psia) = wt % C * 1.514 + wt % H * 4.54 + wt % S * 0.568 wt % O * 0.568 This provides the stoichiometric air in SCF/lb fuel. Non-combustibles such as CO2, N2, ash and water require no air.

1 2

GPA Publication 2145; Figure 16-1, 1981 Revision. North American Combustion Handbook; page 47.

ARPEL Environmental Guideline No. 29

18

The Optimization of Combustion in Boilers and Furnaces

It is important to remember that stoichiometric air assumes perfect (total) combustion. In some cases, it is desirable to not burn a fuel to complete combustion, but in the overwhelming majority of heaters/boilers total combustion is desired. This raises the concept of excess air.

2.1.1

Excess Air

To ensure that there is complete combustion, a quantity of air in excess of the stoichiometric amount is added to the air/fuel mixture. This quantity is known as excess air and is expressed in % of the stoichiometric air requirement. As a rule of thumb, gasfired heaters should have 10-15% excess air. Oil-fired heaters should have 15-25% excess air. There are two main reasons for operating heaters and boilers with excess air: The amount of stoichiometric air is the theoretical quantity of air to achieve complete combustion. This would require perfect air/fuel mixing. However, to ensure that there is intimate contact of the air and fuel and that all the air entering the firebox comes in contact with the fuel is difficult. Fuel quality is often changing. To operate at 0% excess air means that sometimes there is insufficient combustion air. This results in poor energy efficiency, air quality problems and operational upsets. While the 10-15% excess air for gas-fired heaters is common, there are facilities that operate with 5-10% excess air. One company installed very sophisticated feed forward control of its fuel quality on the boilers and achieved as low as 2.5% excess air. See Section 5.1.1 for more detail. In actual practice, excess air rates can climb considerably above the rule-of-thumb limits mentioned earlier in this section. As discussed in Chapters 3 and 5, control of excess air rates can result in considerable improvement in the heater efficiency.

2.2

Relationship Between Excess Oxygen and Excess Air


Excess air is the operating parameter that must be controlled but it cannot be measured directly. The oxygen in the flue gas is measured and from this, the excess air is estimated. The oxygen in the flue gas is occasionally referred to as excess oxygen. This is somewhat of a misnomer. The existence of oxygen in the stack does confirm that excess oxygen has been added but the measured quantity of oxygen is not proportional to the quantity of excess air. See Figure 2.4. Figure 2.4 is based upon calculations for natural gas, a refinery fuel gas that has approximately 80% hydrogen and a sulfur-rich crude oil. The difference between the curves is due to the fact that the flue gas analyses do not report the water content. Nevertheless, the combustion of hydrogen does consume oxygen. This changes the relationship between the reported gases oxygen, nitrogen, carbon dioxide and sulfur dioxide and produces a different curve. The solid dots are data from the literature for a No. 6 Residual Fuel Oil and the solid triangles show the relationship for a natural gas, they are taken from the literature. The

19

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

solid squares are data for coke. This curve is essentially the same as for coal. Two conclusions can be drawn from Figure 2.4: 1. Liquid fuels have a slightly higher excess air rate for the same concentration of oxygen in the flue gas. Solid fuels such as coke or coal have a still higher rate. 2. Fuel streams typically consumed in oil and gas facilities are virtually identical in the relationship between excess air and oxygen in the flue gas. It is suggested that each facility calculate an oxygen/excess air curve for its site-specific fuel. The following procedure is suggested: Determine the hydrogen to carbon ratio of the fuel. When burned in heaters and boilers, it can be assumed that the fuel is completely combusted, with the exception of very minor amounts of contaminants. Determine emission factors for CO, particulates, unburned hydrocarbons, SOx and NOx. They can be obtained from the ARPEL Guidelines for Atmospheric Emissions Inventory Methodologies in the Petroleum Industry. Calculate the amount of oxygen consumed in the formation of these combustion products. Calculate the amount of carbon in the CO and particulates (assume particulates are pure carbon). The remainder of the carbon in the feed is burned to CO2 and the hydrogen is burned to H2O. Calculate the amount of oxygen consumed in the formation of CO2 and H2O. Estimate the amount of oxygen added in the combustion/excess air. This is done by performing a nitrogen balance around the combustion/excess air and the flue gas. (It may be necessary to account for any free nitrogen in the fuel.) The nitrogen can be considered to consist of two streams: the nitrogen associated with the combustion air and the nitrogen associated with the excess air. The amount of nitrogen associated with the excess air is determined using the ratio of nitrogen to oxygen in air and the amount of oxygen in the flue gas. The oxygen in the flue gas is associated with the excess air (all the oxygen associated with the combustion air is consumed in the heater/boiler). The ratio of the nitrogen associated with the excess air and the nitrogen associated with the combustion air gives the ratio of excess air. Note that at low concentrations of oxygen in the flue gas the ratio of oxygen to excess air is close to the ratio observed in air roughly 1:5. However, as the amount of oxygen increases this ratio changes quickly. At its extreme, 21% oxygen corresponds to an infinite excess air rate.

ARPEL Environmental Guideline No. 29

20

The Optimization of Combustion in Boilers and Furnaces


1

Figure 2.4: EXCESS AIR AS A FUNCTION OF OXYGEN IN THE FLUE GAS


350

300

Natural Gas No. 6 Oil

250

Coke Excess Air, %


200

Lines , from left to right, are calculated for crude oil, hydrogen-rich refinery fuel gas and natural gas
150

100

50

0 0 2 4 6 8 10 12 14 16 18

Oxygen in Flue Gas, mol %

2.3

Oxygen Enrichment
The addition of combustion air- even if only the stoichiometric quantity means that a considerable quantity of nitrogen (and argon, carbon dioxide and other natural components of air) must be added to the combustion process. This essentially inert gas (note, nitrogen will form thermal NOx, but for the purposes of combustion it can be considered as being inert) must be heated and therefore, reduces the efficiency of the process. One way of reducing the amount of carrier nitrogen is to use oxygen enrichment. As a rule of thumb, raising the oxygen content of the combustion air by x% will increase heater efficiency by 0.25x % at a flue gas temperature of 600F and by 0.5x% at a flue gas temperature of 1200F.2 Oxygen enrichment will also improve flame velocity and widens the flammability limits of the fuel. Oxygen has been injected directly into the firebox using lances but better mixing control is achieved by injecting the oxygen through the burners. Even more preferable is to premix the oxygen and the combustion air. Typically, the oxygen content is held to 2530% in the combustion air.

1 2

North American Combustion Handbook; pages 54-55. North American Combustion Handbook, page 77.

21

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

The use of oxygen enrichment is usually difficult to justify on economic grounds unless there are oxygen-generating facilities available. Moreover, there are safety factors that must be considered.1

North American Combustion Handbook, page 76.

ARPEL Environmental Guideline No. 29

22

The Optimization of Combustion in Boilers and Furnaces

3.0

HEATER EFFICIENCY

Unless specifically mentioned, the term heater refers to fired process heaters, boilers and steam generators. There are some differences in calculating efficiency of boilers. They are discussed in Section 3.3. The efficiency (%) of a heater is defined as the Useful Heat Transferred * 100 Heat Input This equation1 is exceedingly simple to state but is often difficult to use. First of all, there is the question of heat input. This is the amount of fuel multiplied by the heat of combustion. But does one use the high heating value (HHV) or the low heating value (LHV)? In the formula stated above the useful heat transferred is the quantity of heat transferred to the process fluid or steam, etc. It could also be heat transferred to the combustion air or to the fuel (see Section 3.1). This amount is constant, regardless of whether high or low heating value is used. Therefore, since the high heating value is a larger number, the efficiency (HHV basis) is lower than the efficiency calculated using the low heating value. As long as one is consistent, either heating value is appropriate. A problem may arise when comparing actual performance with design, if the basis of calculation of the design efficiency is not clearly stated. The second problem with the formula for heater efficiency is that, very often, it is difficult to calculate the amount of heat transferred to the process. This is especially true if there is partial vaporization occurring in the heater. There are manual methods, as well as computer simulation software, available to estimate the amount of vaporization but even with good field data there is still a reasonably large error. One method of eliminating this potential source of error is by modifying the equation so that the useful heat transferred is equal to the heat input less all the total losses. Efficiency (%) = (Heat Input Total Losses)/Heat Input * 1002 It is recommended that the efficiency be calculated using both the useful heat transfer and total loss methods. The two calculations will act as cross-checks to each other. The availability and quality of the field data will determine which efficiency estimate is more reliable. It is strongly recommended that individual facilities prepare site-specific efficiency calculation sheets that list the required field and laboratory data and the calculation procedure to be followed. The procedures range from nomographs to graphs to rigorous calculations using computer simulation packages. The philosophies outlined in this document provide a basis for these calculation sheets.

1 2

North American Combustion Handbook; page 56. Ibid, page 56.

23

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

3.1

Effect of Process Variables


To more fully understand the effect of process variables upon heater efficiency it is necessary to expand the total loss efficiency formula: Efficiency (%) = (Heat Input Total Losses)/Heat Input * 100 into an enthalpy balance. This will allow one to quantify the effect of the process variables. In the present formula, the effect is buried in the term total losses. Note, even though the process variables are input variables to an enthalpy balance, the term heat input in the formula above refers to the heat generated from the combustion of the fuel. The process variables of concern are on the fuel side of the heater: heating value, air temperature, humidity, fuel temperature, excess air and atomizing steam. Heating Value: If the high heating value of the fuel is used, the heat lost, by not condensing the water vapor in the flue gas, must be included in the losses. See Section 3.2. If the low heating value of the fuel is used, the latent heat of the water vapor in the flue gas is omitted from the losses. The choice of which heating value to use is at the discretion of the site, but bear in mind the following: The site should preferably use the same basis as the design heater efficiency value and/or the basis used in any previous comprehensive heater studies. This will allow for a more direct comparison of present operation with past results. The same rationale applies if the site will be comparing the performance of several heaters, i.e., in order to determine the most efficient one. If the stack exit temperature is approaching or even lower than the dew point of the flue gas it will be necessary to use the high heating value for the fuel. Air Temperature: Heat from the combustion of the fuel is required to bring the air up to the combustion temperature. Therefore, the hotter the inlet air (combustion air plus excess air) the greater the amount of heat available for transfer to the process fluid. This results in a higher efficiency. Enthalpy values for air are provided in Table 2.2. Note that the table is for dry air only. Humidity: The amount of water contained in air is discussed in Chapter 2. See Figures 2.1, 2.2 and 2.3. The enthalpy of water vapor is listed in Table 2.4. Fuel Temperature: As with air, heat from the combustion of the fuel is required to bring the fuel up to the combustion temperature. Therefore, the hotter the fuel entering the heater the greater the amount of heat available for the process. The enthalpies of the more common gaseous fuels and a heavy fuel oil are provided in Table 3.1. The data for the light hydrocarbons (gases) are from the API

ARPEL Environmental Guideline No. 29

24

The Optimization of Combustion in Boilers and Furnaces

Technical Data Book. They are the values for 0 psia. Unless the actual inlet fuel pressures are very high and this is unlikely the errors in not using the actual pressure and in not accounting for the heat of mixing are minor and most likely within the margin of error of the field data. The enthalpy values have been corrected to a base case enthalpy equal to zero for the gas at 60F. This has been done purely for convenience. It is the change in enthalpy that is important. If this convention is used, it is necessary to properly account for phase changes. For example, if liquid normal-butane at 60F were heated by flue gas and vaporized, the enthalpy of the incoming liquid would have to be given an enthalpy value of 161 BTU/lb (i.e., the difference between saturated liquid and saturated vapor at 0 psia and 60F). For heavy fuel oil (20 API, specific gravity 0.934), enthalpy values have been provided for 0 psia. On the basis of the procedures outlined in Chapters 4 and 7 of the API Technical Data Book, for determining the critical properties of the fuel oil, there is no need to apply pressure correction factors as long as the fuel oil temperature is less than 800F and the pressure is less than 300 psia. The enthalpy values have been corrected to zero for liquid at 60F. A Watson K factor of 11.55 has been used. This is based upon a fraction-by-fraction analysis of a crude oil with a Watson K factor of 11.8.
Table 3.1: ENTHALPY VALUES OF FUELS 1 BTU/lb
Temperature, F C1 C2 C3 iC4 nC4 iC5 nC5 nC6 = C2 = C3 = 1C4 cis C4= = trans C4 = iC4 H2S H2 N2 O2 CO2 CO 20API HFO 0 -31.0 -24.0 -22.0 -21.0 -23.0 -22.0 -22.2 -22.5 -20.8 -20.5 -20.5 -18.7 -20.7 -20.7 -14.4 -203.1 -14.9 -13.1 -11.7 -14.9 -24.0 50 -6.0 -4.2 -4.5 -3.5 -4.0 -3.8 -3.7 -4.3 -3.5 -4.0 -3.0 -3.4 -3.7 -4.2 -2.4 -33.9 -2.5 -2.2 -2.0 -2.5 -4.5 60 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 100 22.0 16.3 15.5 16.5 16.0 16.0 15.8 15.5 14.5 14.5 14.7 13.1 15.0 15.0 9.7 135.8 9.9 8.8 8.1 9.9 17.3 150 48.7 38.8 37.3 37.5 38.0 36.5 37.5 36.5 34.0 33.5 34.0 30.6 35.0 35.3 22.0 306.1 22.4 19.8 18.6 22.4 40.2 200 78.0 61.8 60.0 61.0 61.5 59.5 60.1 59.5 54.5 54.5 55.0 50.3 56.3 56.8 34.4 476.8 34.8 30.9 29.4 34.8 64.2 250 107.0 86.3 85.4 85.3 86.0 84.0 84.3 83.2 75.5 76.0 77.5 70.6 78.7 79.5 47.0 648.0 47.2 42.1 40.6 47.3 89.6 300 137.5 112.3 110.0 111.5 112.0 110.0 110.3 108.5 99.5 100.0 101.0 93.1 102.3 104.3 59.7 819.8 59.7 53.4 52.1 59.9 117.6 350 168.5 139.8 137.0 138.2 139.5 136.5 137.6 135.7 122.5 124.5 125.5 116.1 126.3 129.3 72.6 992.0 72.2 64.7 63.8 72.5 146.2 400 201.0 168.8 165.0 167.0 168.0 165.0 165.3 164.5 147.5 150.0 152.2 140.8 153.1 154.8 85.6 1164.7 84.7 76.2 75.7 85.1 175.2 450 234.5 199.5 194.0 197.0 198.0 195.0 195.0 193.5 174.0 176.0 178.0 166.9 179.3 182.3 98.8 1338.0 97.4 87.7 87.8 97.8 206.3 500 270.0 230.8 225.5 228.0 229.5 226.0 226.3 224.5 200.5 203.5 207.0 194.6 208.3 210.3 112.1 1511.7 110.0 99.4 100.0 110.5 238.3

Excess Air: Excess air is required in order to ensure complete combustion and stable operation. Depending upon the fuel and the amount of firing control instrumentation, recommended excess air rates typically range from 10-30%. However, experience has shown that rates considerably above these levels are common.

For hydrocarbon gases: API Technical Data Book; pages 7-52 to 7-85. For heavy fuel oil: API Technical Data Book; pages 7-121 and 7-128. For CO, H2, CO2, N2, 02: Energy Management Handbook; pages 570-574.

25

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

The presence of excess air represents a waste of energy. Heat is required to raise this quantity of air to the combustion temperature. Enthalpy of the excess air (dry basis) is listed in Table 2.2. Atomizing Steam: In order to improve the mixing of viscous fuels (such as heavy fuel oil) with the combustion air, and to bring about vaporization of the fuel (a necessary condition for combustion) the fuel is atomized. This can be done mechanically but a common practice is to use air or steam. The enthalpy of atomizing air can be found in Table 2.2. The enthalpy of atomizing steam can be approximated using Table 3.2. For more precise calculations, the reader should consult steam tables. Note that steam tables frequently have a base enthalpy (i.e., equaling zero) for liquid water at 32F (0C). At the same time, tables/graphs listing the enthalpy of water vapor have a base of zero for the vapor at different temperatures. For instance the API Technical Data Book uses 60F (15.6C); the Energy Management Handbook, by Wayne C. Turner, uses 77F (25C). If using the steam tables, the reader has two options: 1. Use the steam table values, but segregate the atomizing steam in the mass balance around the fuel side of the heater. This creates a third source of H2O in the stack gas (the other two being the humidity in the air and the product of combustion of hydrogen). There is one problem in that steam tables may not extend to the temperatures experienced in the flue gas or in the firebox. 2. Convert the steam table values to the same basis as the rest of the streams entering on the fuel side. This document uses a base of zero for the gas/vapor at 60F. Table 3.2 has been prepared on this basis.
Table 3.2: ENTHALPY OF ATOMIZING STEAM
Pressure psia 30 45 60 75 90 105 120 135 150 165 180 195 210 Saturated Steam Temp Enthalpy F BTU/lb 250.3 76.8 274.5 84.8 292.7 90.3 307.6 94.7 320.3 98.2 331.4 101.0 341.3 103.4 350.2 105.5 358.4 107.2 366.0 108.7 373.1 110.0 379.7 111.2 385.9 112.2 Enthalpy of Superheated Steam, BTU/lb Temperature, F 300F 350F 400F 450F 102.0 126.2 150.2 173.4 98.7 123.9 148.6 171.9 94.6 121.3 146.8 170.5 118.6 145.0 169.0 116.0 142.8 167.3 112.5 140.4 165.7 138.2 164.2 135.5 162.5 132.8 160.8 130.2 159.3 127.4 157.5 124.5 155.5 121.6 153.4
1

Corrected to same base as water vapour (Enthalpy of vapour = 0 at 60F)

500F 197.0 195.8 194.6 193.4 192.1 190.8 189.5 188.2 186.9 185.6 184.3 182.8 181.3

600F 244.9 244.1 243.2 242.4 241.5 240.4 239.6 238.6 237.7 236.6 235.6 234.6 233.5

700F 293.6 292.9 292.3 291.6 290.9 290.2 289.5 288.7 287.9 287.3 286.5 285.7 284.9

It is important to remember that the enthalpy of the input streams is estimated at the point prior to receiving any heat input from heater flue gas. For example, if the heater is equipped with combustion air preheat (using flue gas), the input enthalpy is the inlet to the air preheater, not the inlet to the heater.

Combustion Engineering Steam Tables

ARPEL Environmental Guideline No. 29

26

The Optimization of Combustion in Boilers and Furnaces

3.2

Energy Losses
An energy loss is any escape/release of heat that is not recovered for use either in the heater or elsewhere. There are six main sources of loss in typical oil and gas industry heaters and boilers: Too much excess air Hot flue gas Latent heat of the water vapor Radiation and convective heat losses from the heater walls Leaks and openings in the heater construction Boiler blowdown. Excess Air: The air entering the heater consists of combustion air (up to the stoichiometric amount) and excess air. As explained above, a certain amount of excess air is required in order to ensure total combustion. The impact of excess air on heater efficiency is shown in Figure 3.1. It is especially significant on heaters with poor heat recovery in the convection section. It could be argued that the nitrogen component in the combustion air is analogous to excess air. This is true in that heat must be supplied to the nitrogen in order to raise it to the furnace temperature. On the other hand, the amount of excess air is a process variable over which the operator can exercise control. Unless the heater uses oxygenenriched air, the operator has no control over the amount of nitrogen in the combustion air. Hot Flue Gas: In many cases, the energy leaving the heater via the flue gas is the largest source of loss. Unlike excess air, which can be changed easily and at no cost, flue gas temperatures cannot be changed significantly without capital investment. The impact of flue gas temperature on heater efficiency is shown in Figure 3.1. The enthalpy values of flue gas components are listed in Table 3.4 at the end of this chapter. In the previous paragraph, it was implied that flue gas temperatures can be changed to some extent. This is shown in Figure 3.2. It is based on a study of two identical boilers, except that one is equipped with a waste heat recovery unit. These boilers are at the same facility so that they use the same fuel and are subject to the same operating practices. Note that the steam production in Figure 3.2 influences the stack temperature on both boilers. The reason for this is presumably limiting heat flux. The amount of steam can be considered a surrogate measurement of the amount of flue gas. The surface area (and condition) of the boiler tubes and the temperatures inside the boiler radiant and convection sections dictate the amount of heat that can be transferred into the process (i.e., steam). When this heat transfer reaches its upper limit, the excess energy passing over the tubes leaves via the stack and results in higher stack exit temperatures.

27

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Figure 3.1 shows the thermal efficiency of a natural gas with a high heating value of 1,000 BTU/SCF. The thermal efficiency is thus a gross (high heating value basis) efficiency. Note that the effect of heat losses from radiation and convection have not been included.
Figure 3.1: THERMAL EFFICIENCY OF A GAS-FIRED HEATER (Parameter is Percentage of Excess Air)
1

100 90 80

Thermal Efficiency (HHV), %

70 60 50 40 30

0% 20% 40% 60% Energy Datum 60F 80% 100% HHV of Gas 1000 BTU/SCF 150%
0 200 400 600 800 1000 1200 1400 1600 1800

20 10 0

Stack Exit Temperature, F

Figure 3.2: STACK TEMPERATURE AS A FUNCTION OF FLUE GAS FLOW RATE

600

No Waste Heat Recovery


550

Flue Gas Exit Temperature, F

500

450

400

With Waste Heat Recovery


350

300

250 40 60 80 100 120 140 160

Steam Production, 000's lb/hour

1 2

Gas Processors Suppliers Association (GPSA) Engineering Data Book, 9 Edition; page 8-13. From Authors files.

th

ARPEL Environmental Guideline No. 29

28

The Optimization of Combustion in Boilers and Furnaces

Latent Heat of Water Vapor: Fuels commonly used in the oil and gas industry consist of about 15-25% hydrogen by weight. Combustion will therefore, generate a considerable quantity of water vapor. In addition, there will be the humidity in the combustion air/excess air and possibly atomizing steam. The thermal properties of water are such that the overwhelming amount of enthalpy of steam is due to the latent heat of vaporization. Since most heaters must operate at flue gas temperatures of 300F and higher, in order to avoid acid gas corrosion, there is no opportunity to capture the latent heat of the water vapor in the flue gas. It is for this reason that the industry adopted the concept of lower heating value. However if the flue gas exit temperature can be in the range of 150F, or lower, a portion of the latent heat of vaporization can be captured and the higher heating value of the fuel (and resultant efficiency calculations) must be used. Radiation and Convective Heat Losses: These losses are a function of the surface temperature of the heater relative to the surrounding air (causing radiation) and the wind velocity. See Figure 3.3. Typically, radiation and convective heat losses amount to 2-3% of the heater firing. However, the condition and thickness of the refractory and insulation will greatly affect this number. As shown in Figure 3.3, the surface area of the heater is a factor.
Figure 3.3: RADIATION AND CONVECTIVE HEAT LOSSES
1

15 MPH Parameter is Wind Speed 10 MPH

Heat Loss Coefficient, BTU/(hr-ft2-F)

5 MPH

0 MPH
4

1 0 100 200 300 400 500 600

Temperature of Surface minus Temperature of Air, F

Leaks and Openings: The amount of heat lost through leaks in the heater construction and through openings will vary from heater to heater. Openings include the gap between the heater wall and incoming/outgoing heater coil piping. It also includes firebox viewing ports left open. No quantification of these losses was found in the literature but it is opined that the energy loss via this route is usually small. However, leaks between heater box panels and in the ductwork and access panels can be substantial. See Section 5.3.
1

GPSA Engineering Data Book, 9 Edition; pages 8-11.

th

29

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Boiler Blowdown: Boiler efficiency calculations are discussed in Section 3.3. In terms of energy losses, there is no difference between heaters and boilers except the question of blowdown. Even with treating of the boiler feedwater, there will be a steady accumulation of dissolved minerals and salts in the water. It is therefore, necessary to blow down some of the water from the boiler and replace it with fresh make-up water. The concentration of the dissolved solids is greatest just below the water level, where the steam bubbles and water separate. A continuous blowdown arrangement is usually installed to withdraw water from this location. It is sometimes called a top blowdown.1 In addition, sediment and salts accumulate in the form of a sludge in the water space. This must be periodically removed using an intermittent bottom blowdown.2 The amount of blowdown will depend upon the incoming water quality after treatment and the boiler pressure (the higher the steam pressure, the greater the need for cleaner boiler feedwater). Typically, blowdown rates of 5-15% are required. There will be an energy loss if the heat in the blowdown water is not recovered for use either in the boiler system or elsewhere. See Table 3.3, which shows the impact of blowdown on boiler efficiency, if no heat recovery from the blowdown is achieved.
Table 3.3: BOILER BLOWDOWN VS. LOST BOILER EFFICIENCY Boiler Pressure, psig Blowdown, % 200 400 600 800
3

% Efficiency Lost 10 5 3.3 1.7 4.0 2.0 4.5 2.2 5.1 2.5

3.3

Boiler Efficiency
Boiler efficiency can mean thermal efficiency, which does not account for radiation and convective losses, or it can mean fuel-to-steam efficiency, which does account for them. Therefore, fuel-to-steam efficiency is a true indication of overall boiler efficiency.4 The ASME Power Test Code 4.1 specifies that the fuel-to-steam efficiency can be calculated using either the Input-Output Method or the Heat Loss Method. The Input-Output Method is calculated by dividing the boiler output (in energy units) by the boiler fuel input (in energy units). The Heat Loss Method subtracts the losses via the stack and radiation/convective losses, as well as blowdown. In summary, boiler efficiency is calculated in the same manner as heater efficiency, with the additional energy loss term to account for blowdown.

1 2

Sauselein; page 55. Ibid; page 59. 3 Garcia-Borras; page 36. 4 Cleaver-Brooks, Efficiency Facts.

ARPEL Environmental Guideline No. 29

30

The Optimization of Combustion in Boilers and Furnaces

3.4

Examples
The examples presented below illustrate the effects of altering the main factors concerning energy management of heaters and boilers. The analytical procedures and data provided earlier in this document form the basis of this discussion. Many of the graphs and tables can be converted to formulae using regression techniques. The following conditions exist as the base case:
Fuel Gas CO2 C1 C2 C3 iC4 nC4 iC5 nC5 H2S 1.00 vol% 85.00 7.00 3.00 1.50 2.00 0.20 0.29 0.01 Density Sulfur Inerts Water Nitrogen Fuel Oil 0.96 1.50 wt % 0.40 0.10 0.15

Ambient air 90F Fuel gas temperature 90F Fuel oil temperature 175F Atomizing steam 2.50 lbs/USG of fuel (steam at 150 psig, 500F) Process heat required 40.00 MM BTU/hr 50% supplied by gas, 50% by oil Oxygen content of flue gas (dry basis) 6.50 wt % Temperature of flue gas exit, 700F Scenario 1 Base Case: The oxygen content of the flue gas is equivalent to an excess air rate of 41.7%. This is considerably above the recommended 15-25%. Also, the stack exit temperature (700F) is extremely high, although quite common in older heaters. The combination of high excess air and high stack exit temperature result in a heater efficiency of 78.6% (lower heating efficiency). Scenario 2: By reducing the air flow to the burners so that the oxygen content of the flue gas falls to 3.0% (dry basis), the excess air rate will be reduced to 15.5%. This

31

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

action, while involving no capital outlay, will increase the heater efficiency to 83.3%. Note that, by reducing the excess air flow, the stack exit temperature will fall. See Figure 3.2 above. For this exercise, it is assumed that the stack outlet temperature (in this case, the outlet of the convection section) will decrease 4F for every decrease in flue gas flow of 1,000 lbs/hour. This is in line with the data presented in Figure 3.2. Because the heater is more efficient, there will be less flue gas for the same amount of heat absorbed by the process. Therefore, the stack outlet temperature will fall to approximately 645F. The improvement in heater efficiency (4.7%) means that fuel consumption is reduced by 2.9 MM BTU/hr, while still maintaining 40.0 MM BTU/hr of process heat. At a fuel value of $2.00 (US) per MM BTU, the fuel savings total over $50,000 annually. Scenario 3: Significant energy savings will be achieved by recovering the heat lost via the flue gas. Waste heat recovery will involve capital investment. In this scenario, it is assumed that the temperature of the stack exit gas is reduced to 350F. In actual applications it will be necessary to determine the acid gas dew point and then add a safety margin to account for changes in sulfur content of the fuel and variations in heat transfer in the waste heat recovery unit. See Sections 5.2 and 7.3 for further details. Based on the data presented there, a fuel oil with a sulfur content of 1.50 wt % will have a dew point of less than 300F. Fuel gas with low H2S content will also have a dew point less than 300F. Therefore, the assumed stack exit temperature of 350F should be suitable. Reduction of the stack gas temperature to 350F can be realized by a number of ways: By heating another stream (Scenario 3) By heating the process stream going to the heater (Scenario 4) By heating the combustion air/excess air and the fuels and the process feed to the heater (Scenario 5). Reducing the stack temperature to 350F, with the excess air rate held at 15.5% will raise the heater efficiency to 91.4%. The base case load of 40.0 MM BTU/hr is still maintained in the radiant and convection sections of the heater and an additional 3.9 MM BTU/hr is recovered from the flue gas. The fuel savings are dependent upon the efficiency of the heater/boiler that originally supplied the 3.9 MM BTU/hr. Assuming that the efficiency of the other heater was less than the heater in these scenarios, the heat savings are at least 3.9/0.914 = 4.27 MM BTU/hr, with an annual value of nearly $75,000 (US). Scenario 4: In this case, the waste heat in the flue gas is used to preheat the process fluid going to the heater itself. Since only 40.0 MM BTU/hr of process heat is required, it means that the heater duty can be reduced. The efficiency remains the same 91.4%. The estimated fuel savings are 4.26 MM BTU/hr compared with the results in Scenario 2, i.e., the same as in Scenario 3. However, the heater is partially unloaded. Scenario 5: The third option is to use the waste heat to preheat the combustion air, the fuel gas, the fuel oil and/or the process fluid. Preheating the air and/or the fuel puts more energy into the heater firebox, thereby leaving more heat available for transfer to

ARPEL Environmental Guideline No. 29

32

The Optimization of Combustion in Boilers and Furnaces

the process. For this exercise, the fuel gas, fuel oil and the air are preheated to 300F. The remainder of the available heat (1.0 MM BTU/hr) is used to preheat the process fluid. Again, the fuel savings, compared with Scenario 2, are 4.26 MM BTU/hr. The difference between this scenario and the previous two is that the adiabatic flame temperature increases from 3,309F to 3,471F. This has implications regarding emissions of NOx. See Chapter 9. Implementing Scenario 3, 4 or 5 involves a capital expenditure. The choice will depend upon a number of factors: The ability to find a heat sink for the waste heat in the stack gas. The required size of the waste heat recovery equipment. The need to unload process heaters so that problems can be eliminated or unit charge rates can be increased. The need to avoid air quality problems. The presence of potential safety issues.
Table 3.4: ENTHALPY OF FLUE GAS COMPONENTS BTU/lb
Enthalpy of Stack Flue Gases
Temperature, F SO2 H2O (vapour) N2 O2 Air CO2 200 16.5 58.6 34.8 30.9 34.0 29.4 300 34.8 104.8 59.7 53.4 58.4 52.1 400 53.1 151.7 84.7 76.2 83.1 75.7 Outlet Conditions 200-1600F 500 600 700 71.7 90.3 109.0 198.2 245.8 294.2 110.0 135.4 161.1 99.4 122.9 146.8 107.9 129.5 158.4 100.0 125.6 151.2 800 127.9 343.3 187.1 171.1 184.1 177.3 900 145.7 393.2 213.3 195.7 210.2 204.1 1000 165.7 443.9 239.8 220.6 231.8 231.4 1100 185.7 495.5 266.6 245.8 256.2 259.4 1200 205.6 547.9 293.6 271.2 280.5 287.9

1300 225.6 601.8 320.9 296.8 305.4 317.1

1400 245.6 655.9 348.4 322.7 330.2 346.8

1500 265.6 711.6 376.3 348.7 355.5 377.2

1600 285.6 766.4 404.4 374.9 380.7 408.1

Enthalpy of Firebox Gases


Temperature, F SO2 (est'd) H2O (vapour) N2 O2 Air CO2 2400 457.4 1255.3 642.4 594.1 614.9 659.3 2600 503.1 1384.0 702.0 649.1 694.4 722.4 2800 548.9 1514.9 762.2 704.7 750.5 786.3

Firebox Conditions 2400-4400F 3000 3200 3400 594.6 640.4 686.1 1647.9 1782.8 1919.5 822.9 884.0 945.6 760.7 817.1 873.9 807.4 865.5 925.0 850.8 915.9 981.7

3600 731.8 2058.1 1007.6 931.2 1014.3 1048.0

3800 777.6 2198.3 1070.0 988.7 1070.9 1114.9

4000 823.3 2340.2 1132.8 1046.7 1140.5 1182.3

4200 869.1 2483.7 1195.9 1104.9 1210.9 1250.2

4400 914.8 2628.6 1259.4 1163.5 1282.1 1318.6

Energy Management Handbook

33

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

4.0

TYPES OF FIRED EQUIPMENT

There are many designs of fired heaters and boilers. Descriptions of these lie outside the scope of this document. The major classifications will be briefly discussed in order to provide a background for the remaining chapters of this guideline. Similarly, there are numerous burner designs. Again, only the major classifications will be presented.

4.1

Direct-Fired Heaters
Direct-fired heaters use the heat from the flue gas and/or the radiant heat from the flame to heat the process fluid. The heaters are classified primarily according to their shape with sub-classifications according to the path followed by the flue gas and the orientation of the coils. The classifications are cylindrical, cabin and box. The cabin heater has a sloping roof between the radiant and convection sections. The radiant section of a box heater has a flat roof. The coils in cylindrical heaters are frequently vertical but are sometimes helical. The coils in box heaters can be horizontal, vertical or arbour-shaped (in the form of an arc at the top of the radiant section). The flue gas usually flows upward but there are designs where there is downflow to the stack and others where the flue gas flows lengthwise along the radiant section. Figure 4.1 shows examples of the cabin and cylindrical classifications. Cylindrical heaters are typically 10-15% cheaper than cabin heaters for the same process duty except when the duty is less than 10 MMBTU/hr. They also require less plot space for installation. Fewer internal tube supports and soot blowers are required. Air preheat facilities are smaller. Large cylindrical heaters have more natural draft and higher convection heat transfer coefficients than cabin heaters. However, cabin heaters have fewer problems with two-phase flow. (Cylindrical heaters are prone to problems with slug flow.) Cabin heaters can have upward-firing, but they can also have end and side-firing. Therefore, they can be built closer to the ground than cylindrical heaters, which must have upward firing burners. By installing a bridgewall down the center of the cabin, a cabin heater can have dual service. The two main heat transfer compartments in a heater are the radiant section and the convection section. Some cylindrical heaters consist of only a radiant section. They are characterized by low thermal efficiency. There are also heaters that have only convection sections. The radiant section is the portion of the heater that contains the burners. Heat transfer is by radiation. A well-designed and operated radiant section will have no flame impingement on the tubes. The flame length is generally 60% of the firebox and the edge of the flame should be at least 1.5 feet from the tubes. Flue gas leaving the radiant section should be 1,500-1,900F.

35

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Figure 4.1: TYPICAL DIRECT FIRED PROCESS HEATERS

For small cabin heaters the fire box should be of equal dimensions (width, height and length). For larger units the relative dimensions (W:H:L) are typically 1:2:4. For small cylindrical heaters the height of the radiant section is equal to the diameter of the tube

GPSA Engineering Data Book, 10 Edition; page 8-15.

th

ARPEL Environmental Guideline No. 29

36

The Optimization of Combustion in Boilers and Furnaces

circle. (See Figure 4.1). For large cylindrical heaters the length of the radiant section should be twice the diameter of the tube circle.1 The convection section is downstream of the radiant section (the physical location depends upon the heater construction). Heat is transferred from the flue gas. Depending upon the surface area of the tubes installed there, the flue gas leaving the convection section will have temperatures from 300-1,600F. Older heaters tended to have less surface area, i.e., higher flue gas exit temperatures and therefore, lower efficiency. Plants have resorted to a number of options for increasing heat transfer in the convection section and in the lower stack: Adding more tubes to the convection section Installing studded tubes, or even better, finned tubes in the convection section Installing waste heat recovery units into the stack. For more detail, see Chapter 7.

4.2

Fired Boilers
Fired boilers consist of two main types: fire-tube and water-tube. In the fire-tube type, the fuel is fired in the furnace and the flue gas travels through the tubes. The furnace and tubes are installed within a vessel that contains the steam and water. The flue gas may travel through 2-4 passes before exiting via the stack. The boiler may have a dry back, wherein a refractory-lined chamber directs the flue gas from the furnace to the tubes. If the boiler has a wet back, the refractory is replaced by a water-cooled jacket. A very common example of this form of boiler is the Scotch Marine boiler. It is a packaged unit, meaning that the burner, controls and auxiliary equipment are designed as a single engineered package and ready for on-site installation. A fire tube boiler is generally limited to 350 psig steam.2 They are rated in boiler horsepower (BHP), where 1 BHP = 33,475 BTU/hr.3 Ratings for fire tube boilers range from 15-1,500 BHP (0.5 MMBTU/hr to 50 MMBTU/hr).4 Water-tube boilers are generally used for steam pressures greater than 350 psig. In this type of boiler, the tubes are filled with water/steam, while the flue gases flow around the tubes. A steam drum is mounted at the top of the tubes and a mud drum is mounted below the tubes. Packaged water-tube boilers can generate steam up to 200,000 lbs/hr at 1,000 psig and 850F.5 Common configurations for packaged water-tube boilers are the D-type and the O-type. See Figure 4.2. In the D-type, the convection tubes may contain a superheater. It is

1 2

GPSA Engineering Data Book, 10 Edition; page 8-15. Cleaver-Brooks; http://www.cleaver-brooks.com/Boilersa/.html and /GlossFP.html 3 Ibid, http:/www.cleaver-brooks.com/GlossAE.html 4 Ibid, http:/www.cleaver-brooks.com/Boilersa/.html 5 V. Ganapathy, Understand Boiler Performance Characteristics; Hydrocarbon Processing; August 1994; page 131.

th

37

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

installed in the convection section at a point where the flue gas is hot enough to deliver the desired steam temperature. If only 20-50F of superheat is required, the superheater is usually mounted between the evaporator and the economizer. The flue gas leaving the convection section is used to preheat boiler feedwater in an economizer. Air preheaters on packaged boilers are not common due to cost considerations, large gas/air pressure drops and increased formation of NOx.1 The O-type boiler is not as easy to fit with an economizer or superheater. The design is symmetrical so the least tube surface is exposed to radiant heat.2 The A-type has two smaller lower drums or headers and a large upper drum for steam and water separation.3 Packaged boilers usually have forced-draft air supply. This allows pollution abatement equipment such as flue gas recirculation, staged-fuel or staged-air burners. Pressures in the furnace can reach 10-30 inches of H2O. This could lead to emissions of carbon monoxide if the partition wall, which separates the furnace from the convection section, develops leaks.4 It is very important to match the design of a packaged boiler with its normal operating conditions. As the load on the boiler is decreased, the steam temperature, the flue gas exit temperature and the boiler feedwater temperature leaving the economizer also decrease. This phenomenon is discussed in Section 3.2. For the effect on the operation of a packaged boiler, see Figure 4.3. Note that the efficiency of the boiler has a maximum value. This is the result of three factors. The flue gas temperature increases with increasing load, and as seen in Chapter 3, this leads to a loss of thermal efficiency as more heat is escaping out the stack. Another loss is due to radiation and convective losses. Packaged boilers typically have radiation and convective losses of 0.5% of the heat released at full load.5 Since the surface temperature of the boiler remains relatively constant over the operating range, the heat loss in quantitative terms is also relatively constant. However, in terms of the percentage of heat release, the radiation loss increases to approximately 2% at 25% load. Moreover, in order to maintain good firing conditions the excess air may have to be increased at low boiler loads.

V. Ganapathy, Understand Boiler Performance Characteristics; Hydrocarbon Processing; August 1994; page 131. 2 J. Makansi; Managing Steam; page 19. 3 Ibid. 4 V. Ganapathy; Hydrocarbon Processing; August 1994; page 132. 5 Ibid, page 132 and H. Hendry (John Inglis Equipment); Boiler Efficiency and Testing; Publication B623; no date.

ARPEL Environmental Guideline No. 29

38

The Optimization of Combustion in Boilers and Furnaces


1

Figure 4.2: COMMON BOILER CONFIGURATIONS

V. Ganapathy; Hydrocarbon Processing; August 1994; page 132.

39

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Figure 4.3: PERFORMANCE OF A PACKAGED BOILER

800

87

Steam Temperature
700 86

Temperature, F

600

85

Efficiency
500 84

Boiler Feedwater Leaving Economizer


400 83

300

82

Flue Gas Exit Temperature


200 20 30 40 50 60 70 80 90 81 100

Percent Load or Steam Flow

4.3

Incinerators 4.3.1 Incinerators for Process Wastes

Incinerators are used to destroy hazardous wastes and pollutant gases such as sulfur plant tail gas. Depending upon the heating value and quantity of the waste, supplemental fuel may be used. There are burners capable of burning three fuels: oil, gas and low-BTU waste gas. There is a tip for each fuel, which can be burned simultaneously. The single most important factor that determines whether supplemental firing is required is the hydrogen content of the waste gas. Normally, if the fuel contains 10-15 volume % hydrogen and the heating value of the gas is at least 80-100 BTU/ft3, no supplemental fuel is required. However, a gas containing 25% methane and 75% CO2 has a heating value of 227 BTU/ft3 but will require either supplemental firing or a special combustion chamber.2 Descriptions of incinerators for waste solids and liquids are provided in the ARPEL Guidelines for the Management of Petroleum Refinery Solid Wastes (Section 5.5.3) and Guidelines for the Management of Petroleum Refinery Liquid Wastes (Section 5.5.3). Incineration of the tail gas from sulfur recovery units is very common. The oxygen in the flue gas is often in the range of 5%. Use of high intensity burners can significantly reduce supplemental fuel consumption by providing better air-fuel mixing. This results in a shorter, hotter flame, which ensures more complete destruction of the pollutants (H2S, COS, CS2 and S). Reduction of the oxygen in the flue gas to as low as 2-3% is possible, resulting in fuel savings of up to 30%. However, it is important to review the impact of these changes upon downwind air dispersion. With less flue gas going up the stack the
1 2

V. Ganapathy; Hydrocarbon Processing; August 1994; page 135. John Zink Company; Combustion and Industrial Burner Application and Design; no date; page 42-44.

ARPEL Environmental Guideline No. 29

40

Efficiency (HHV), %

The Optimization of Combustion in Boilers and Furnaces

plume rise and subsequent dispersion of the SO2 will be altered. It is necessary to confirm that the revised incinerator operations still allow compliance with applicable air regulations.

4.3.2

Incinerators for Medical Wastes

A number of ARPEL facilities also support non-industrial activities such as hospitals. Incineration of medical wastes is therefore, necessary. Wherever possible, they should be sent to an incinerator licensed to handle them. If the incineration is done on-site, a controlled air incinerator, with primary and secondary combustion chambers and a heated hearth is necessary. See Sections 5.5.3.3, 5.5.3.3.1 and 5.5.3.3.2 of the ARPEL Guidelines for the Management of Petroleum Refinery Solid Wastes for further information.

4.4

Other Heaters
Among the less-common types of fired heaters found in the oil and gas industry are the following: Convection heaters. These have no radiant section. They can be equipped with flue gas recirculation and pre-mix burners. As a result, the heaters have greater gas flow (higher heat transfer coefficients) and about 10% excess air. These combine to produce a highly efficient heater. Convection heaters also offer a high degree of safety and are therefore ideal for installations such as offshore platforms. Bath heaters. This is an indirect fired heater. The fuel is combusted in the fire tube and the heat is transferred to the heat medium, which in turn transfers the heat to the process fluid. The desired bath temperature determines the heat medium. See Table 4.1. A common use for bath heaters, especially in the production sector of the oil and gas industry, is for providing heat to circulating heat media such as glycol. The heat media is often used in low-temperature reboilers to separate LPG from solution gas and natural gas for heating buildings. Other uses include the heating of natural gas pipelines from the wellhead to the production facility in order to keep the gas above the hydrate point and to heat oil/water emulsions in order to provide better separation.

41

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Table 4.1: TYPICAL BATH HEATERS Heat Medium Water 50% Ethylene Glycol Low Pressure Steam Hot Oil Molten Salt TEG Reboiler Amine Reboiler Bath Temperature, F 180-195 195-205 245-250 300-550 400-800 350-400 245-270

Thermal Efficiency (LHV), % 76-82 76-80 76-80 71-76 68-74 75-80 75-80

4.5

Burners
Brief descriptions of the various types of burners commonly used in the oil and gas industry are provided. For information regarding the construction and operation of these burners, the reader is referred to the article Combustion and Industrial Burner Application and Design issued by John Zink Company. The burners discussed below are generic types. They have been included to provide the reader with the range of options available and the effects that burner type has on combustion. The wide variety of burner models, even within each generic type, precludes descriptions of individual brand names. Each facility should consult the equipment vendors and manufacturers for their expertise concerning site-specific burner issues. Premix burners. The primary air is pulled into the fuel gas by means of a venturi. Secondary air registers supply the remaining combustion air. The flame is short and dense. The normal mode of operation is to have the primary air registers fully open and the secondary air adjusted to the desired excess air. (The primary air registers are adjusted if there is flashback or flame liftoff.) This burner operates on a natural draft heater. Nozzle-mixing gas burners. The air and gas are kept separate until they leave the nozzle. These are also known as raw gas burners. They have excellent turndown ratios (10 to 1), whereas premix burners have a nominal turndown ratio of 3 to 1. Although the burners require lower gas pressures than premix burners, there still must be sufficient pressure to ensure good air-fuel mixing if low excess air operation is desired. All in all, the mixing is not as good as with premix burners. Nozzle-mixing gas burners also operate with natural draft air supply. Combination natural draft gas and oil burners. Most oil burners in refinery operations have the capability of firing gaseous fuels. The gas firing ports provide the flame pattern. The position of the oil tip is critical. If it is too high there may be

GPSA Engineering Data Book, 10 Edition; page 8-25.

th

ARPEL Environmental Guideline No. 29

42

The Optimization of Combustion in Boilers and Furnaces

flame instability and if it is too low there may be coking on the burner tile and an oil spill. The operation of simultaneous oil and gas firing is discussed in Section 5.3. Low air pressure drop burners. These are forced draft burners and are normally used when low air pressure drops are available or when there is the choice of operating in a natural draft mode. The natural draft mode determines the size of the burner but the materials of construction are designed for preheated air temperatures. High air pressure drop burners. This is a forced draft, dual-fuel burner that is well suited for heaters with a single burner. Rotating vanes in the air register impart a spinning motion to the air, which causes a short, wide flame. There is better mixing than with the low air pressure burner, which allows a lower gas pressure (5 psig versus 15 psig) and possibly less atomizing steam. Large burners have a dual zone air register to ensure that the oil receives sufficient air. There are variations of this burner. High intensity burners. These require forced-draft air supply. The air passes through a series of spin vanes, causing a vortex. Most, (75+%) if not all, of the combustion occurs in the refractory lined combustion chamber. The flame is very compact. Because the heat is retained in the chamber makes it an ideal burner for heavy oils and gases with low heating values. Table 4.2 summarizes the specifications of the above-mentioned burners. intended as a guide. Burner manufacturers should be consulted for their input.
Table 4.2: BURNER SPECIFICATIONS
Burner Type Air Draft Fuel Heat Release MMBTU/hr 0.1 - 15.0 Normal Draft inches WC <0.1 - 1 Normal Minimum Excess Air, % 5
1

This is

Flame Shape Conical Flat Fan Round Flat Conical Flat Fan Conical Flat Fan Conical Flat Fan Conical Conical Compact

Flame Length Various

Flame Diameter Various

Premix

Natural

Gas

Nozzle-Mixing

Natural

Gas

0.5 - 20.0

0.1 - 1

5 - 10

Various

Various

Combination

Natural

Gas Oil Gas Oil Gas Oil Gas Oil Gas Oil

1.0 - 20.0

0.1 - 1

5 - 10 for gas 10 - 15 for oil 5 - 10 for gas 10 - 15 for oil 5 <5 <5

Various

Various

Low Air P High Air P #1 High Air P #2 High Intensity

Forced

1.0 - 30.0

2+

Various

Various

Forced Forced Forced

15.0 - 200.0 5.0 - 40.0 5.0 - 100.0

2 - 10 1-9 6

Various Various

Various Various

Considerable attention has been paid to burners in Claus Sulfur Recovery Units because of the need to maintain very close combustion control and the need to ensure thermal destruction of the tail gas. For further information the reader is referred to the January/February 1993 issue of the magazine Sulfur, for the article Leading Burner Designs for Sulfur Plants, pages 23-34.

John Zink Company; Combustion and Industrial Burner Application and Design; no date; pages 10-40.

43

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Low-NOx burners are becoming very important in light of the increasingly stringent air quality regulations. Among the options for reducing NOx is the use of staged-air and staged-fuel burners. These are discussed in Section 9.3.

4.6

Design Criteria
This chapter has provided an indication of the complexity involved in designing fired heaters and boilers. The same complexity exists when specifying burners and peripheral equipment for improving thermal efficiency and complying with air quality regulations. It is recommended that heater, boiler and burner manufacturers be included into the project teams investigating either new heaters/boilers or ones undergoing modifications and upgrades. These specialists should be contacted as early as possible in the study, as their input will significantly influence the final design. It is incumbent upon the facilities to have sufficient high quality data of past performance that modifications and upgrades can be designed and easily incorporated into the existing equipment. It is also essential that the plant have as much information about future operation as possible for both upgrades and new-fired heaters/boilers. The entire range of operating conditions should be clearly stated. Existing and expected environmental regulations must also be known. The simultaneous, and at times opposing, demands of energy conservation and air quality compliance have reduced the operating region available to staff. Tighter control of fuel and airflow to the heaters is a must if one is to meet these demands while maintaining safe operation. The reader is referred to the article Controlling Fired Heaters by W. Driedger in the April 1997 issue of Hydrocarbon Processing, pages 103118.

ARPEL Environmental Guideline No. 29

44

The Optimization of Combustion in Boilers and Furnaces

5.0

MONITORING PERFORMANCE

Good monitoring of performance and proper operating practices can achieve significant efficiency in heaters and boilers. Options requiring capital outlay are discussed in Chapters 911.

5.1

Fuel Analysis
Analysis of the fuel(s) is necessary to know the heating value, the required air and the resulting adiabatic flame temperature. Also, contaminants such as sulfur/H2S and nitrogen will determine the emissions of SOx and fuel NOx. These in turn dictate the minimum flue gas temperatures that can be achieved in waste heat recovery units. It is recommended that regular testing of fuel(s) consumed at the facility be conducted. This is especially necessary if fuel gas is being supplied by processes subject to catalytic deactivation. For example, catalytic naphtha reformers produce large quantities of hydrogen at the beginning of their cycles. If the unit offgas is sent to fuel, the quantity of hydrogen will have an impact on the fuel characteristics. If the facility imports its fuel, it should ask the supplier for analyses on a regular basis. The term regular testing is defined as being frequent enough to detect normal changes in fuel quality. This will vary from facility to facility and will have to be determined by experience. At least once per week for refineries and gas plants is suggested. Standard fuel testing protocols, as specified by ASTM or equivalent, should be followed. The analysis of a fuel must be sufficient to allow a determination of its heating value and its contaminants. For gaseous fuels, this will consist of a composition at least as far as C6-C8. For liquid fuels, density values and sulfur levels are required frequently and analyses for inert materials, water and nitrogen less frequently.

5.1.1

Wobbe Index

When fuel quality changes, the plant operators must consider the following: The same heat input rate must be maintained for similar process conditions. The heater/boiler equipment must be able to handle the new fuel. Equipment includes the stack, burners, piping valves and controls. The stability of the burner, the heat release pattern and the furnace atmosphere must remain constant. The Wobbe Index enables the operators to evaluate changes to the first two points mentioned above. The index is confined to gas-gas comparison. If the old and new gases have the same index value, there is no need to change valve settings in the fuel gas supply piping.

45

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

The Wobbe index is given by the formula:1 Wo = Ho/(Go)0.5 = Hn/(Gn)0.5 where:


Wo Ho Hn Go Gn = = = = = Wobbe Index heating value of the old fuel heating value of the new fuel gravity of the old fuel, relative to air gravity of the new fuel, relative to air

One company, faced with widely and quickly varying fuel gas quality used the output of a Wobbe Index meter, regression analysis and feed forward control to operate their utility boilers at 1.5-2.0% excess air, or about 0.5% oxygen in the stack.2 The regression analysis was used to correlate to correlate the Wobbe Index with fuel heating value, specific gravity and air requirements. Note that many factors such as fuel quality, burner type, air supply and air quality emission levels must be considered when operating at such low excess air levels. However, with appropriate data and control schemes, it may be possible to achieve those levels. Feedforward air control is especially useful if the heater or boiler is firing, as its entire or partial fuel, waste gases since they are subject to wide variation in composition and flow. This situation is discussed in the article Feedforward Air Control for Fuel BTU Changes by E. Vicknair, in the July 1985 issue of Hydrocarbon Processing, pages 65-66.

5.2

Stack Gas Monitoring


The stack gas from each heater and boiler should be tested as frequently as possible. Many heaters have oxygen analyzers and/or controls on the stack. These should be checked on a regular basis. As discussed in Section 5.3, the greater the draft (the larger the negative pressure) the greater the ingress of air through leaks and other openings in the heater. For this reason, it is suggested that the flue gas be sampled as close to the firebox as possible. (The highest pressure (lowest draft) is recorded at the arch, or entry into the convection section. If the sample is taken at the arch, the probe should be zirconium oxide as the temperature is typically in the range of 1,400-1,800F.) Avoid taking the sample near or downstream of an induced draft fan as the suction of the fan has a very low pressure (i.e., potentially high ingress rates of air). Flue gas analyses can be done using a variety of equipment such as gas chromatographs, ORSAT apparatus, and Bacharach (Fyrite) analyzers. These devices

1 2

North American Combustion Handbook; page 39. T. Mort and I. Verhappen; Improving Fuel Efficiency with Statistics; Chemical Engineering; June 1991; pages 143-146.

ARPEL Environmental Guideline No. 29

46

The Optimization of Combustion in Boilers and Furnaces

vary in sophistication. Some are capable of measuring only one gas some Fyrite analyzers measure oxygen and others measure carbon dioxide while others, such as the ORSAT apparatus can measure oxygen, carbon monoxide and carbon dioxide. Whatever device is used, its accuracy must be confirmed using test standards on a regular basis. As discussed later in this chapter, air leakage into the heater is a common problem. This leakage will provide falsely high readings of the oxygen in the flue gas. This becomes especially critical when operating at very low excess air rates or if the air supply (by a fan) is controlled by the oxygen content in the flue gas. When the amount of combustion air falls below the stoichiometric amount, the generation of hydrogen and carbon monoxide rapidly escalates. See Figure 5.1. To protect the heater/boiler from an unsafe situation, analyzers for combustible gases (hydrocarbons, hydrogen and carbon monoxide) are also installed in the flue gas stack. The reading of this instrument is used to override the reading on the oxygen analyzer. It is suggested that the efficiency of each heater and boiler be calculated at least weekly. This can be done using tables, graphs, nomographs or spreadsheets. For a better understanding of the heater performance, the efficiency should be tracked over time. Key operating parameters should be noted so that the causes of swings in efficiency can be determined. Readings of the oxygen in the stack gas and the stack gas exit temperature should be reported at least daily and preferably each shift. Heater/boiler efficiency is primarily a function of the excess air rate and the flue gas exit temperature. Other factors include the ambient air temperature and the fuel conditions. It is recommended that the monitoring of heater performance include the simultaneous recording of the flue gas analysis and the stack exit temperature, as well as the ambient conditions. It is preferable that the fuel analysis and fuel temperature be recorded at roughly the same time as the stack monitoring. Another important variable to monitor is the draft profile throughout the heater and duct. This is especially necessary if the heater/boiler is suspected of poor operation. When conducting a draft survey, record the damper position and the condition of any equipment in the stack downstream of the convection section. Ensure that the draft gauges are operating properly and that the instrument lines are clear. More intensive surveys such as thermography and analyses of the gases within the firebox and convection sections are recommended if the heater is operating inefficiently and normal practices (see Section 5.3) are not producing the desired effect. The surveys are also required if heater capacity upgrades are planned.

5.3

Firing Practices
The following procedures should be done to optimize heater/boiler operation. Reduce excess air to the levels specified by the burner manufacturer. These levels will vary depending upon the type of burner, the type of air supply (natural draft or forced draft) and the type of fuel. Recommended excess air levels are provided in Table 5.1.

47

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Table 5.1: RECOMMENDED EXCESS AIR LEVELS, % Natural Draft Fuel Gas Light Fuel Oil Heavy Fuel Oil 15-20 20-25 25-30 Forced Draft 10-15 15-20 20-25

Table 5.1 should only be used as a guide if the manufacturers recommended excess air rates are not available. The values in the table presume the heaters are in good condition with minimum air leaks. Moreover, low-NOx burners tend to require more excess air. It is extremely important not to reduce excess air rates to the point where flame instability or incomplete combustion occurs. Theoretically, the stoichiometric amount of air would result in zero percent oxygen in the flue gas. However, air leaks into the heater and stack and poor air-fuel mixing lead to the situation where a measured 515% excess air is needed to achieve stoichiometric conditions at the burner. Airflow to the burner is varied by controlling the draft of the heater. The available draft in a heater is given by the equation:2 Draft = 0.192 * Hs * (g a) + 0.0029 * V2 * g * (4 * f * Hs/D+1) where Draft Hs g a V f D = = = = = = = available draft, in inches of water height of stack, feet density of flue gas, lbs/cubic foot density of air, lbs/cubic foot velocity of flue gas, feet/second Fanning friction factor diameter of stack, feet

The Fanning friction factor can be found in a variety of books, such as Perrys Chemical Engineers Handbook, 4th Edition, page 5-20.

A. Garg; Optimize Fired Heater Operations to Save Money; Hydrocarbon Processing; June 1997; page 103. 2 th GPSA, Engineering Data Book,10 Edition; page 8-10.

ARPEL Environmental Guideline No. 29

48

The Optimization of Combustion in Boilers and Furnaces

For heaters that are natural draft or forced-draft, the draft is controlled by adjusting the damper in the stack. For heaters that are equipped with an induced-draft fan on the outlet of the convection section, the draft is controlled by the dampers on the suction of the induced-draft fan. The draft can also be controlled by the air registers at the burners. The draft equation noted above consists of two terms. The first term is always negative because the density of the flue gas is always less than the density of the air. This provides the total available driving force of the gas. The second term accounts for the pressure drop due to friction and the stack exit velocity.1 Typically, there is a pressure increase in the radiant section of 0.01 inches of water per foot of radiant section height.2 The highest pressure in the heater is at the arch, or top of the radiant section.3 This value is used to control the heater draft. The higher the draft, the greater the amount of air drawn into the heater. Conversely, the lower the draft, the smaller the amount of air drawn into the heater. If the draft is too low, unsafe conditions could exist, wherein flue gas would leak out of the heater. To maintain proper draft control, a pressure reading of 0.1 inches of water should be maintained at the arch.4 This results in safe operation and minimizes air leaks. Closing the stack damper reduces the draft (i.e., the pressure at the arch moves closer to zero). This changes the draft profile across the entire heater. Opening the air registers also reduces the draft but only as far as the damper. To adjust the quantity of excess air, both the damper and the air registers must be adjusted.5 The API Document 535 titled Burners for Fired Heaters in General Refinery Service, 1st Edition (July 1995) published a flow chart for adjusting the draft and oxygen in the flue gas. The contents are summarized here:6 If the draft at the arch is at the target level, check the oxygen in the flue gas. If the O2 is on target, there is good operation. If the 02 is higher than target, close the damper. If the O2 is below the target, open the air registers. If the draft at the arch is higher than the target level (i.e., greater negative pressure), check the oxygen in the flue gas. If the O2 is on target, open the air registers and close the damper. If the O2 is higher than target, close the damper. If the O2 is below target, open the air registers. If the draft at the arch is lower than the target (i.e., smaller negative pressure), check the oxygen in the flue gas. If the O2 is on target, open the damper and close the air registers. If the O2 is higher than target, close the air registers. If the O2 is below target open the damper. Figure 5.1 shows a typical draft profile for a direct fired heater. Note that is slightly different from the profile printed in API Document 535, 1st Edition (July 1995) because it
1 2 th

GPSA Engineering Data Book, 10 Edition; page 8-10. Ibid, page 8-18. 3 A. Garg; Hydrocarbon Processing; June 1997; page 99. 4 Ibid and discussions with John Zink Canada. 5 A. Garg; Hydrocarbon Processing; June 1997; page 101. 6 Ibid.

49

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

has combined the pressure drop and the stack effect across the convection section and it has included the pressure drop across the burners.
Figure 5.1: TYPICAL DRAFT PROFILE IN A DIRECT FIRED HEATER
Negative Pressure
1

Positive Pressure

Stack

Breeching Arch
ooooo ooooo ooooo

Damper Convection Section

Maintain at -0.1 inches of water column

o oo o o o o o o o o

oo o o o o o o o o o

Radiant Section

Burners

If the amount of air into the firebox drops below the stoichiometric requirement, there will be a rapid increase in the amount of carbon monoxide and hydrogen produced. See Figure 5.2. (Note, this is the basis of staged burners. See Section 9.3.)
Figure 5.2: EMISSIONS OF HYDROGEN AND CARBON MONOXIDE UNDER SUBSTOICHIOMETRIC 2 CONDITIONS

9 8

Fuel is Methane
7

Volume % (wet basis)

H2
6 5 4 3 2 1 0 60 70 80 90 100 110

Theoretical Equilibrium Concentration of H2

CO

Theoretical Equilibrium Concentration of CO

% of Stoichiometric Air

1 2

GPSA Engineering Data Book; 10 Edition; page 8-19. Burner Design Parameters for Flue Gas NOx Control; by R. Martin, John Zink Company; no date; Figures 1 and 2.

th

ARPEL Environmental Guideline No. 29

50

The Optimization of Combustion in Boilers and Furnaces

Assuming that the heater is being operated with proper draft control and with sufficient air to achieve good flame shape and complete combustion with the available burners, the efficiency of the heater can be calculated using the procedures and data presented in Chapters 1-3. A quick estimate for heaters firing natural gas can be obtained from Figure 5.3.
Figure 5.3: THERMAL EFFICIENCY (LHV) FOR A HEATER FIRING NATURAL GAS
1

95

90

Thermal Efficiency (LHV), %

Parameter is Excess Air


85

80

15%
75

20% 25%

70

65 300

30% 40% 50%


400 500 600 700 800 900 1000 1100

Stack Exit Temperature, F

If the heater conditions lie outside the range of the parameters in Figure 5.3, the efficiency can be estimated using Figure 3.1. See Figure 2.4 for the relationship between oxygen in the flue gas and excess air rate. Note that Figure 5.3 expresses efficiency in terms of its lower heating value, where as Figure 3.1 is in terms of its high heating value. It is best to work entirely on one graph or the other and to note the basis of the efficiency calculation (LHV or HHV). For plants that measure the carbon dioxide in the flue gas, instead of the oxygen, Figure 5.4 can be used if the heaters burn natural gas and Figure 5.5 if they burn No. 6 Oil. Note that the graphs show the stack loss. To convert to efficiency, subtract the stack loss plus radiation losses from 100%. For additional information regarding optimization of heater operation, the reader is referred to the article Optimize Fired Heater Operations to Save Money by A. Garg in the June 1997 issue of Hydrocarbon Processing, pages 97-104. When firing oil, ensure that there is sufficient amount of atomizing steam to ensure good combustion. Not only does the steam break up the oil into droplets, it provides discharge velocity to ensure good air-fuel mixing. Dry steam should be used: wet steam causes sparking. The atomizing steam pressure should be 20-30 psi above the fuel oil pressure. Steam consumption is about 0.3 pounds of steam per pound of fuel.

Adapted from A. Garg; Hydrocarbon Processing; June 1997; page 101. Some points have been adjusted to smooth the lines.

51

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Figure 5.4: ENERGY LOSSES WHEN BURNING NATURAL GAS


60

Parameter is % CO2 in Flue Gas


50

3%

4% 6%

40

8% 10%

Stack Loss, %

30

12%

20

Radiation and Convection Losses Excluded


10

Gas Heating Value 1000 BTU/SCF


0 200

300

400

500

600

700

800

900

1000

1100

Difference Between Flue Gas Exit and Ambient Temperatures, F

Figure 5.5: ENERGY LOSSES WHEN BURNING NO. 6 OIL

80

Parameter is % CO2 in Flue Gas


70

3%

4%

60

6% Stack Losses, %
50

8%
40

10% 12% 14% 16%

30

20

10

Radiation and Convection Loss Excluded


0 200 300 400 500 600 700 800 900 1000 1100

Difference Between Flue Gas Exit and Ambient Temperatures, F

Steam-assist burners use less steam. However, the fuel orifice is smaller so it is more prone to plug and high oil and steam pressures are required.

1 2

Garcia-Borras, page 27. Ibid, page 26.

ARPEL Environmental Guideline No. 29

52

The Optimization of Combustion in Boilers and Furnaces

Air atomization burners use low pressure air (1-2 psig). With a good design, 10% of the combustion air is injected into the heater as atomizing air. Low fuel pressure can be used but the higher the oil pressure, the lower the amount of air. Mechanical atomizers use extremely high fuel oil pressure to break up the oil. Very high pressures are obtained by having very small orifices at the fuel tip. These devices are usually only found in extremely high heat release burners. The higher the fuel oil pressure at maximum firing the higher the available pressure at turn-down conditions. This results in better atomization even with reduced rates of atomizing steam.1 When firing oil and gas in a single burner, the flame length is always longer, even when there is sufficient air. It is not possible to determine whether the proper fuel-to-air ratio is being maintained. It is suggested that, if two fuels are necessary, fire some burners on oil as the base-load and control the heater by firing gas on the other burners. If the flames on the oil burners are long and smoky, the burner is probably overfired and the fuel oil flow should be adjusted to get good flame quality. If dual-firing on burners must be practiced, all burners should be base-loaded equally with one fuel and controlled by adjusting the other fuel.2 For proper heat distribution in the firebox and burning a single fuel, maintain equal fuel pressure to each burner. Open the air registers of all burners the same amount on a natural draft heater. For heaters with forced-draft air supply, open fully the burner air dampers and control the air flow by using the fan suction damper. If the heater has staged-air low-NOx burners, ensure that all dampers are open the same amount.3 In order to maximize heat transfer across the available surface area, it is necessary to ensure that the tubes are clean. The problem of flame impingement causing internal coking has been discussed elsewhere in this document. External fouling by soot also rapidly leads to poor heat transfer. See Figure 5-6. To remove soot from the convection section tubes, soot blowers are installed if the heaters burn oil. They are typically installed every four or five rows.4 The soot is usually removed by high pressure steam directed onto the tubes. Soot blowing is generally done only once per shift. These devices require considerable maintenance. Among techniques to improve soot blowing is the use of microprocessor based control that directs the soot blowing to areas of poorest heat transfer. Sonic sootblowing uses a nearly continuous sound pressure wave to keep ash and soot suspended in the flue gas. These devices have low maintenance, are easily installed and can be fully automated. However, they can only remove light, friable deposits of slag and soot. Vanadium and sodium in the fuel oil form sticky deposits because of their low melting points. Neither steam nor sonic soot blowing can easily remove such deposits.5
1 2

John Zink Company; Combustion and Industrial Burner Application and Design; no date; pages 21-27. Ibid; pages 27-28. 3 A. Garg; Hydrocarbon Processing; June 1997; page 102 4 th GPSA Engineering Data Book, 10 Edition; page 8-17. 5 A. Garg and H. Ghosh; Chemical Engineering; October 1990; page 222.

53

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Figure 5.6: EFFECT OF SOOT DEPOSITS ON FUEL COMBUSTION

9 8

Increase in Fuel Consumption, %

7 6 5 4 3 2 1 0 0.00

0.02

0.04

0.06

0.08

0.10

0.12

0.14

Thickness of Soot, inches

Scale on the heater tubes has the same effect as soot an increase in fuel consumption for the same heat transfer. See Figure 5.7. Note that the type of scale is a factor. Scales containing silica are particularly troublesome. The tubes should be cleaned of scale during shutdowns in order to improve heat transfer and heater efficiency. (Also during shutdowns, the condition of finned and studded tubes should be noted and burned-off and broken fins/studs repaired or replaced. The quality of the heater refractory and insulation should be checked frequently and repairs made as soon as possible. Typically radiation and convective losses, due to wind, are in the range of 2-3% of the heat release. They can be more accurately estimated using Figure 3.3. The impact of excess air on heater efficiency has been discussed previously. The ingress of air into the heater can take place via unused burners, open sight doors, explosion doors, draft sample tubes and leaks. This air not only reduces efficiency by absorbing heat but also creates a falsely high oxygen reading in the flue gas. See Figure 5.8 for an estimation of air leakage. Note that if the heater has an induced draft fan, the pressure may be as low as 10 inches of water near the suction of the fan. Any leaks in this region will have very high leakage rates.2 Operators should ensure that all open ports are closed if not needed and that the air registers of unused burners are shut. (An exception to this is staged air combustion. See Section 9.3.) Leaks should be repaired as soon as possible.

1 2

Garcia-Borras, page 41. A. Garg; Hydrocarbon Processing; June 1997; page 101.

ARPEL Environmental Guideline No. 29

54

The Optimization of Combustion in Boilers and Furnaces


1

Figure 5.7: EFFECT OF SCALE DEPOSITS ON FUEL COMBUSTION


9 8 7 6

Iron and Silica


5

High Iron Content


4 3

Normal Scale
2 1 0 0 0.01 0.02 0.03 0.04 0.05 0.06

Thickness of Scale, inches

Figure 5.8: AIR LEAKS INTO HEATERS

2
700

600

Air Infiltration, ft /hour/in

500

400

300

200

100

-0.40

-0.35

-0.30

-0.25

-0.20

-0.15

-0.10

-0.05

0 0.00

Furnace Pressure, inches water column

1 2

A. Garg; Hydrocarbon Processing; June 1997; page 101. Garcia-Borras, page 54.

55

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

6.0

ENERGY REDUCTION TECHNIQUES

This chapter raises the issue of energy effectiveness. This is related to, but yet quite distinct from, energy efficiency. As an example of these concepts, consider the generation of steam to operate non-condensing turbines. At their best, a boiler will have an efficiency of about 90% and a turbine will be about 75% efficient. Assume 100,000 lbs/hr of 600 psia steam at 700F drives the above-mentioned turbine and exits as 100 psia steam. The turbine exhaust steam is then condensed in aerial coolers for recycling back to the boiler. Assuming that the deaerator operates at atmospheric pressure, the fuel input to the boiler is 130 MM BTU/hr. Only 17.5 MM BTU/hr is consumed in the turbine, where 13 MM BTU/hr of useful work is produced. The remaining energy in the steam 100 MM BTU/hr is dissipated in the aerial coolers (including the energy required to operate the fan motors). In summary, only 10% of the energy input is recovered in the form of useful work. Over 76% of the available energy is wasted in the aerial coolers. While the individual pieces of equipment are very efficient, the overall effectiveness of the system is very low. The object of this chapter is to discuss methods of increasing the effectiveness of heater/boiler systems.

6.1

Minimizing Heater Duties


There are two sides involved with heaters and boilers: the process side and the fuel side. This chapter deals with the minimization of heater duties using technologies and techniques that reduce the amount of process heat that must be absorbed. Chapter 7 discusses methods of ensuring that the amount of fuel input required to achieve the required process heat is as low as possible. The reader is urged to consult the ARPEL Guidelines for the Development of Energy Audits in Upstream and Downstream Oil and Gas Facilities. That document provides a systematic approach for evaluating the impact of process conditions on energy input.

6.1.1

Heat Exchange

Heat exchangers are important pieces of heat transfer equipment. For many process units, the quantity of heat transferred in exchangers is equivalent to, or even greater, than that supplied in heaters. Proper placement and operation of them can significantly reduce the amount of heat that must be supplied by fired heaters. In evaluating the effectiveness of heat transfer by exchangers in a plant, consider the following: The amount of heat transfer using exchangers. Are the expected temperatures being achieved? If not, are the exchangers fouled? Has surface area been decreased due to plugged tubes? Are the fins on tubes with extended surfaces damaged? Do the exchangers have the appropriate internal configuration (i.e., the optimum number of shell/tube passes, correct baffle spacing, tube pitch, etc.). The heat exchanger arrangement in exchanger trains. An exchanger train is a heat exchanger network wherein one common stream (e.g., such as crude oil) is heated by a number of product streams (e.g., the products from a crude distillation unit). To maximize heat transfer, it is desirable to have the common stream heated by

57

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

increasingly hotter streams. The original design usually takes this into account. However, changing product slates, changing feed quality and later modifications/expansions to the exchanger train can combine to reduce the effectiveness of the train. The heat exchanger arrangement in exchanger systems. By this is meant smaller exchanger networks where one stream exchanges heat with only one other. More than one exchanger is involved. It could also describe a subsection of an exchanger train. Multiple exchangers are installed in series or in parallel. Heat transfer is increased by installing in series. However, pressure drop increases significantly. On the other hand, when installed in parallel, the plant has the ability to take an exchanger out of service for cleaning or repair with less disruption to operations. The amount of heat removed from the process by cooling water or air. Water and air for cooling should be utilized only when the product streams can no longer feasibly supply heat to other streams. The use of intermediate storage. In order to prevent excessive evaporative losses and/or to maintain the integrity of the storage tanks, intermediate products from Unit A are usually cooled prior to storage. The heat exchangers in the downstream unit (Unit B) are therefore, required to raise the temperature of the incoming feed, often essentially back to the levels originally experienced in Unit A prior to cooling for storage. Minimization of intermediate storage would make more process heat available for increasing the temperature of the feed to the heater(s) in Unit B and/or providing heat to other locations. The amount of inter-unit heat transfer. Heat transfer between process units can be very effective. However, the improvement in energy use must be balanced against the risk of upsets in one unit causing problems in other units. Heat exchanger simulation programs are useful in evaluating individual exchangers and simple arrangements. Technologies such as PINCH are recommended for analyzing complex arrangements and heat exchanger trains. It must be stated that monitoring and analyzing a heat exchanger network can be complicated. The reader is referred to the article Challenges in Simulating Heat Exchanger Networks, by R. Sigal, in the October, 1996 issue of Hydrocarbon Processing, pages 125-132. Nevertheless experience has shown that significant improvements in heat exchanger performance are possible using these analytical tools.

6.1.2

Effective Distillation

The nature of crude oil, and to a smaller extent, natural gas, is such that distillation is required to manufacture useful products. Proper and effective use of distillation equipment can significantly reduce demands on fired heaters and reboilers. The importance of distillation is illustrated by the fact that in the U.K., distillation-related energy consumption accounts for 13% of the total energy use by process industries.1 At

Chemical Engineering; July 1997; page 72.

ARPEL Environmental Guideline No. 29

58

The Optimization of Combustion in Boilers and Furnaces

the same time, the expanding availability of process simulation computer packages has provided energy management teams with vastly increased analytical capability. Changes to the operation of distillation columns must be undertaken so as to maintain product quality. There will usually be some deterioration in quality but in many cases there is sufficient over-purity or giveaway that the product still meets specification. The following points should be considered when evaluating distillation columns. The amount of overhead reflux. A considerable amount of heat is usually removed in the top reflux exchangers. However, this is the coldest part of the column so there is less opportunity for transferring the heat to another product. By removing heat/more heat using a pumparound stream located further down in the column, a hotter source of heat is obtained and there is less heat removed in the overhead condenser. Pumparound streams are frequently installed on towers where there is no reboiler. A common example is a crude distillation tower. The heat balance around the column dictates that, for constant product draw rates, less heat put into the column means less heat has to be removed. For distillation columns equipped with reboilers, adjustment of the reflux directly affects reboiler firing. (Note, even for reboilers that use a heat medium, the same principle applies because at some point in the process, a fired unit provides the heat to the circulating heat medium.) Operating a propane-propylene splitter at a reflux ratio of 15.6 produces a propylene stream of 99.9% purity. Reducing the reflux ratio to 13.5 causes the propylene purity to fall to 99.7 mol %. However, the reboiler duty is reduced from 3,166 BTU/lb of propylene to 2,740 BTU/lb of propylene. Fuel savings total 13.5%.1 Whether the operators could control the tower closely enough to achieve these savings is another point. The important point of this example is the potential impact on energy use that even small operational changes can cause. Reflux can only be reduced as long as product quality is acceptable. Amount of overflash. This is another variation of reflux. When overflash rates (the liquid flowing down from the tray above the flash zone) are excessive, the flow can be reduced by lowering the feed or reboiler temperature or by raising the hydrocarbon partial pressure in the column. Note that pumparound streams affect the internal reflux flows above the pumparound trays. Overflash is unaffected by the existence of pumparounds.2 Product splits. This is a measure of product purity. For light hydrocarbons, a component balance can determine whether the split between the light and heavy key components is appropriate. For example, on a deethanizer in a gas plant the light key component is ethane and the heavy key component is propane. There are usually specifications on the amount of ethane allowed in propane. Ensure that this specification is being met, but not at the expense of distilling propane into the overhead gas. This not only consumes energy but downgrades product value. The split between petroleum fractions is analyzed by comparing the distillation curves, especially the overlap of the tail end of the heavier product with the front end of the next
1 2

Chemical Engineering; December 1992; page 94. BP Research Project 139; Crude Oil Distillation, Principles and Practice; June 16, 1967; page 6.

59

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

(lighter) product. The greater the fractionation, the smaller the overlap. allowable product quality should determine the size of the overlap.

However,

The use of stripping steam. Steam is often injected into columns in order to reduce the hydrocarbon partial pressure. It is also used to strip out light material that may cause flash point problems. As mentioned above, varying the hydrocarbon partial pressure of a distillation column by using steam injection, for example will change the overflash flow. Stripping steam rates should be set at the minimum required to meet specification. If stripping steam rates are too high, performance may actually deteriorate because of flow distribution problems in the sidedraw strippers, causing channeling and flooding. The points noted above are effective means of achieving energy savings at little or no capital costs. There are also techniques for reducing energy use that involve fundamental changes to the distillation process itself. A common example of this is the installation of preflash columns. By only distilling one product overhead (sometimes two products can be removed in a preflash column), less energy is required than in a conventional column with an overhead product and several product sidedraws. As an example, a preflash column/crude distillation column arrangement was simulated (by computer) at 35,000 bpd of crude. By installing an 1,820 bpd sidedraw from the preflash tower, overall energy requirements would have been reduced by 5.5 MM BTU/hr. This was equivalent to 3.2% of the total fuel input to the unit.1 Studies have been undertaken in Europe regarding the installation of a series of preflash towers on crude units similar to the sequential use in gas plants of demethanizers, deethanizers, depropanizers and debutanizers. Distillation columns/stabilizers are designed for specific operating pressures. When cooling water and even more importantly, aerial coolers, are used for condensing the overhead product, the design pressure is set by the maximum temperature reached by the cooling medium. If the cooling water or ambient air becomes cooler the condenser outlet temperature drops and the column pressure decreases. In older units, to take into account situations where the water/air is cooler, the columns frequently have a bypass around the condensers so that the design pressure can be held constant. Adherence to this operating philosophy results in a lost opportunity to reduce energy costs. Generally, it is desirable to operate at as low a pressure as possible in order to maximize the relative volatility between the key components of the separation. The limit to which the pressure can be reduced is reached when a more expensive cooling medium is required.2 Consideration should be given to reducing the operating pressure of distillation columns. Those columns that utilize cooling water or air in the overhead condensers are potential candidates, especially if significant seasonal and/or diurnal changes in air/water temperatures are experienced. As the temperature of the water/air changes, the pressure also varies, or floats.

1 2

BP Canada; Memo by N.A. Franklin; Preflash Sidedraw; February 28, 1978. th GPSA Engineering Data Book, 10 Edition; page 19-4.

ARPEL Environmental Guideline No. 29

60

The Optimization of Combustion in Boilers and Furnaces

As an example of this principle, floating pressure operation was instituted on a light straight run gasoline stabilizer in a crude unit in a Canadian refinery. The pressure was reduced from the design of 145 psig to an average of 115 psig. Reboiler duty was reduced by approximately 8.5%.1 As a second example, Figure 6.1 shows the effect of pressure on the reboiler duty of a preflash tower handling 34,400 bpd of crude oil.
Figure 6.1: EFFECT OF PRESSURE ON REBOILER DUTY
25

Reboiler Duty, MMBTU/hour absorbed

20

15

10

Constant Reflux Ratio No Sidedraw Product


5

0 15 20 25 30 35 40

Overhead Pressure, psia

Investigate the feasibility of improving the efficiency of distillation column internals. Distillation column calculations deal with theoretical stages. The ratio of the number of theoretical stages to the number of actual trays is the tray efficiency. As indicated in Table 6.1, there is a wide variation in tray efficiencies according to the service. Moreover, tray efficiency varies through the distillation column.

1 2

BP Canada, Memo by N.A. Franklin; Variable Operating Pressure for 21-C-6; December 7, 1976. BP Canada, Memo by N.J. Little; Sidedraw from No. 1 CDU Preflash Column; March 8, 1977.

61

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Table 6.1: TYPICAL TRAY EFFICIENCIES

Demethanizer Deethanizer Depropanizer Debutanizer Butane Splitter Condensate Stabilizer Crude Distillation Column* Top Middle Bottom Crude Preflash Column**

45-60% 50-70 80-90 85-95 90-110 40-60 52 79 32 67 63

* BP Canada, memo by N.A. Franklin, 11-C-1, Primary Column Simulation, April 17, 1978 ** BP Canada, memo by N.J. Little, Sidedraw from CDU Preflash Column, March 8, 1977

The efficiency of distillation columns can be improved using packing. This can consist of random packing such as Pall rings, Raschig rings and Berl saddles or structured packing such as knitted-type mesh and corrugated plates. Packing has several advantages:2 Lower pressure drop per theoretical stage. For this reason packing (especially structured packing) is being used more and more in vacuum columns3 High liquid loading. This is especially so for structured packing Corrosion protection. The packing can be made from ceramics and plastics, whereas trays may have to be fabricated from expensive alloys. Note that one European Company is now manufacturing glass-polytetrafluoroethylene bubble-cap trays.4

1 2

GPSA Engineering Data Book, 10 Edition; page 19-16. th GPSA Engineering Data Book, 10 Edition; page 10-17. 3 Chemical Engineering; November 1997; page 37. 4 Chemical Engineering; November 1997; page 37.

th

ARPEL Environmental Guideline No. 29

62

The Optimization of Combustion in Boilers and Furnaces

Packing also has several disadvantages:1 Limited turndown. Packing is limited to about 50% turndown, while trays can operate at 10-15% of full load. Liquid distribution problems. Channeling is much more prevalent in packed columns unless there is proper distribution of the liquid at the top of the column and periodic redistribution throughout the column. Plugging. Dirt and other substances can more easily plug packing. Higher cost. Unless expensive alloys are required, trayed columns generally cost less than packed columns.2 Rigidity of design. It is very important to match actual performance with design when considering packed columns. In other words, trayed columns are better suited to handle deviations from design.3 Early distillation designs generally favored tray internals. During the energy crises of the 1970s and 1980s, the energy savings achieved by packing resulted in the rapid growth of that technology. Lately, new tray designs are being brought into the market that can handle greater liquid and vapor flows.4 In summary, determine the efficiency of the distillation columns at the facility. When considering means of improving efficiency, look at the question of trays versus packing. Also consult the vendors of both types of internals regarding the latest developments. In closing this section, it is important to note that a considerable amount of work is being undertaken in the area of distillation theory. It is more proper to say that many of these technologies have been known for a long time but concerns about operation and control of these units resulted in little acceptance by industry in the past. These concerns appear to be diminishing. Among the technologies now under development are the following: Fully thermally coupled columns link two or more columns with vapor and liquid streams. However, there are no reboilers or condensers between the columns. This arrangement reduces thermodynamic losses. Divided-wall columns essentially have two fractionation columns mounted within a single shell. This technology is a variant of the fully thermally coupled columns concept. This arrangement reduces the number of trays, column shells, reboilers and condensers. In addition energy consumption is lower. Catalytic distillation combines reactions and distillation within a single column.

1 2

GPSA Engineering Data Book, 10 Edition; page 10-17. Chemical Engineering;November 1997; page 37. 3 Ibid 4 Ibid

th

63

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

For more information on these subjects the reader is urged to read the following articles: Chemical Engineering, Advanced Distillation Saves Energy and Capital, by F. Lestak and C. Collins, July 1997, pages 72-76. Chemical Engineering, Catalytic Distillation Extends Its Reach, by K. Rock, G. Gildert and T. McGuirk, July 1997, pages 78-84. Chemical Engineering, New Horizons in Distillation, by J. Humphrey and F. Seibert, December 1992, pages 86-98. Other distillation technologies mentioned in the literature, such as the spinning cone and the inverted columns appear to have little potential use within the conventional oil and gas industry,1 at least in the near future.

6.2

Steam Systems
Energy management of steam systems is a complex subject. The reader is urged to read the ARPEL Guidelines for Energy Management of Steam Systems. This section will briefly discuss the topic. Unlike hydrocarbon systems, steam systems typically discharge, vent and release a considerable portion of the overall quantity of steam and condensate that is generated. Some of this lost steam/condensate is unavoidable but the energy management team should ensure that losses are minimized and that as much heat as possible is extracted from them prior to ultimate disposal. This will reduce the load on the boilers. Among the potential energy management items are the following. (Efficient operation of the boilers is discussed in Chapters 5 and 7.) Prepare steam balances for the various operating scenarios. These will highlight where steam is used, where it is wasted and where it is lost. On the basis of this information, priorities can be set. Shut off steam going to units that are down or that no longer require steam. Steam tracing is frequently left in service long after it is no longer needed. Ensure that steam traps are appropriate for the service and that they are in working order. This can be a serious source of steam loss. For example, studies have shown that a steam trap that is leaking has been in that condition on average for six months.2 The average defective trap loses 50 pounds of steam per hour.3 While this seems a minimal amount, consider the number of traps in a facility. A survey of the Hls America chemical plant at Theodore, Alabama showed that 12.4% of its 887 traps were leaking. Steam losses totaled $336,000 US annually.4 At a Canadian heavy oil production facility

1 2

Chemical Engineering; Distillation Internal Matters; November 1997; page 37. Predictive Maintenance of Steam Traps: Combining Demand Side Management and Performance Contracting; by F. Hooper, Jr. and R. Gillette; www.trapo.com/idea-2.htm. 3 Back to Basics Steam Traps 101; by D. Fischer; www.powerspecialties.com/Armstrong_back_to_basics_Traps101.htm. 4 Energy-Saving Steam Traps Earn Respect at Hls; by R. Wily; www.powerspecialties.com/EnergySavingSteamTraps.htm.

ARPEL Environmental Guideline No. 29

64

The Optimization of Combustion in Boilers and Furnaces

over 20% of the steam traps were found to be leaking into the condensate return header and causing excess steam venting in the deaerator. Maximize the amount of condensate returned to the boiler feedwater. Loss of condensate not only represents energy waste but it means that water treatment costs must increase because of the additional boiler feed water make-up. Minimize the amount of boiler blowdown. A portion of the water in the boiler must be released in order to control the quantity of salts and minerals in the water. However, excessive blowdown wastes the energy input to raise the water from its inlet temperature to that of the saturated steam, plus the associated pumping costs. Wherever possible, recover the heat in the blowdown before it is discarded. Recover the heat in steam vented at the deaerator. Non-condensable gases, such as oxygen and carbon dioxide, must be vented from steam. This is typically done by injecting low pressure steam into the deaerator and raising the temperature to near that of saturated steam at atmospheric pressure or slightly higher. As a result steam is vented from the deaerator. At these pressures the latent heat of the steam is in the order of 950-970 BTU/lb, whereas the enthalpy of the condensate is only 180-220 BTU/lb (base = zero for water at 32F). Condensing the vented steam from the deaerator can provide heat to incoming boiler feedwater, or for other low-level requirements, such as a circulating heat medium. Minimize steam sent to process. This must be done consistent with acceptable product quality or operating stability. Examples would include stripping steam in distillation columns and steam in FCCU regenerators. Ensure steam-consuming equipment such as turbines and vacuum ejectors are operating as efficiently as possible. Also ensure that the driven equipment (pumps and compressors) is also operating both efficiently and effectively. Determine the optimum steam pressure levels and steam quality (superheated, saturated, wet steam). This is a very complex exercise for the steam system must address a wide variety of needs. A thorough knowledge of the steam system and the process requirements is needed. Consider the use of condensing turbines. High pressure turbine exhaust steam means that the enthalpy drop per pound of steam is low so a large quantity of steam is needed. Condensing turbines greatly increase the enthalpy drop per pound of steam so the quantity of steam is reduced. To improve overall thermal efficiency, recover the heat from the condenser. Evaluate the relative merits of operating electric motors instead of steam-driven turbines. Depending upon the location of the facility and the relative costs of electricity and fuel, motors may be more economical. Reliability of the power supply must be considered. Ensure that the latent/sensible heat of steam is recovered for process use, rather than removed by cooling water or air.

65

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Identify sources of process heat for generating steam in waste heat boilers. Often this heat is supplied by flue gas from heaters (see Section 7.3). Other sources include hot process streams such as vacuum residuum and FCCU regenerators.

6.3

Insulation
Proper installation and maintenance of insulation on piping and equipment is an important component of an energy management program. Wet and damaged insulation should be repaired as soon as possible. Inadequate insulation can have severe repercussions on operations such as downhole steam injection in oil fields, due to the length of pipe involved. For estimates of the heat loss from piping, the reader is referred to Chapter 14 of the ARPEL guideline on Energy Management of Steam Systems. The graphs included therein are applicable to both steam and process lines.

ARPEL Environmental Guideline No. 29

66

The Optimization of Combustion in Boilers and Furnaces

7.0

ENERGY RECOVERY TECHNIQUES

While not an infallible statement, it is generally true that older heaters have lower efficiency than newer ones, unless they have been revamped. The techniques mentioned in Chapter 5 will improve heater operation, but only within the confines of the equipment limitations. To attempt to operate a heater beyond these limits runs the risk of flame impingement, high metal temperatures, improper draft profile in the heater and fans and pumps running at overload conditions. Additional heater duty will require additional equipment. It is stressed that prior to revamping a heater or boiler that a comprehensive analysis of the heater should be undertaken. Among the data that should be collected are the following: Process conditions (flows, temperatures, pressures and pressure drops) Fuel analyses Flue gas analyses Temperature and pressure profiles throughout the heater Tube temperatures, preferably including a survey using infrared thermography Complete operating data of heater peripherals such as waste heat recovery units and fans Known limitations Heater design data Heater inspection report The above-mentioned data should be collected at the same time in order to obtain a snapshot of the heater. There are three main methods of revamping heaters and boilers in order to improve thermal efficiency: Modify and/or expand the radiant and convection sections of the heater for greater process heat Convert natural-draft to forced-draft operation Install waste heat recovery. Each of these methods will be discussed in this chapter. Note that this review will not look at methods of improving heater capacity that do not simultaneously improve heater efficiency.

67

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

7.1

Modifying Radiant and Convection Sections


The performance of the radiant section is gauged by its heat flux and the velocity of the process fluid through the tubes. Typical values for refinery process heaters are provided in Table 7.1.

Table 7.1: HEAT FLUX RATES AND MASS VELOCITIES FOR REFINERY PROCESS HEATERS

Service

Radiant Flux BTU/hr/ft2

Mass Velocity lbs/sec/ft2 174251 61-102 143-251 348-451 45-72 348-451

Atmospheric crude heater Reduced crude vacuum heater Reboilers Circulating oil heaters Catalytic reformer charge reheat Delayed coker Visbreaker heating section Visbreaker soaking section Propane deasphalting Hydrotreater charge Hydrocracker charge Steam superheater

9,82712,046 7,9259,985 9,510-12,046 7,925-10,936 7,449-12,046 9,510-10,936 8,876-9,985 6,023-6,974 7,925-9,034 9,510-12,046 9,510-12,046 9,827-14,899

154-205 154-205 31-76

If the radiant section is not meeting the design specifications, the investigators should: Ensure that the heater tubes are clean and that there are no flame impingement and hot spots, which could indicate coking on the inside of the tubes. Check the amount of air to the burners. As indicated in Figure 7.1, as the amount of air is increased, the fraction of heat absorbed in the radiant section decreases. Remember that, for natural gas firing, stoichiometry requires approximately 10 volumes of air per volume of gas. This means that roughly 17 pounds of air are required for every pound of fuel gas. At 15% excess air (19.5 lbs air/lb gas) and a flux rate of 10,000 BTU/hr/ft2 about 53.5% of the total allowable heat flux occurs in the radiant section. Operating at 25% excess air, the radiant section absorbs only 50.7% of the total heat.
1

H. Ghosh; Improve Your Fired Heaters; Chemical Engineering; March 1992; page 116.

ARPEL Environmental Guideline No. 29

68

The Optimization of Combustion in Boilers and Furnaces

Determine whether there is poor heat distribution within the firebox. When doing this it is necessary to take into consideration the distance of the tubes from the flames and the airflow in the firebox. The maximum radiant temperature is approximately 23 feet from the flame.1 At greater differences the temperature decreases. A long, narrow flame therefore, will have a relatively constant distance between it and the tubes in the radiant section. This results in a more even distribution of heat. The height of the flame should not exceed 50% of the height of the radiant section. As a general rule, the proper flame size is about 15 inches high for every million BTU/hr of heat release. The diameter should be about 30 inches. To prevent flame impingement, the flame edge should be 24-31 inches from the tubes.2 Natural gas burners produce long flames. This is due to the relatively poor mixing of the fuel and air. Low excess air rates at the burner or poor air distribution will exacerbate this situation. The combustion is completed higher up in the firebox where there is more oxygen. (This is the principle used in staged-air firing. See Section 9.3.) Air leaks into the firebox will also change the heat distribution pattern not only by causing changes to the airflow but also by decreasing the temperature of the gases. This last point is important. Radiant heat transfer coefficients are functions of temperature to the fourth power. At the temperatures typically experienced exiting the radiant section (1500-1900F) and the process temperatures (about 300-700F), even a 10F drop in the firebox gas temperature decreases the radiant heat transfer coefficient by nearly 2%. Localized variations in firebox temperature will cause localized variations in heat transfer. The firebox internal construction and air leaks may result in short-circuiting of the airflow, meaning that heat is taken out of the radiant section prematurely. It is very important to remember that heaters and burners are designed to meet the specifications of the client. Depending upon the desired process conditions, and the desired heat transfer rates, there could be long, narrow flames or short, wide flames, and a variety of gas flow regimes as the hot gases travel from the burners to the convection section. Moreover, present firing conditions can range from under-firing to over-firing; both of which could exhibit radically different heat distribution patterns from design. It is recommended that burner and heater manufacturers be consulted during analysis of the radiant section.3 The first two points mentioned above refer back to the discussion in Chapter 5. However, in order to properly assess the performance of the radiant section, it is necessary to optimize conditions, within the limits of the available equipment. The third point lies outside the control of the operating staff and will require a capital outlay in order to rectify it. It is recommended that heater manufacturers or specialists in this field be consulted regarding modifications to the heater internals. There are several options available regarding the convection section. An approach temperature (flue gas outlet minus fluid inlet) of 90F can be used.4 This can be achieved by:

1 2

Discussion with John Zink Canada. H. Ghosh, Improve Your Fired Heaters; Chemical Engineering; March 1992; page 121. 3 Discussion with John Zink Canada. 4 A. Garg; Revamp Fired Heaters to Increase Capacity; Hydrocarbon Processing; June 1998; page 72.

69

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Adding tubes. Flux rates in the convection section are usually in the range of 12,000-24,000 BTU/hr/ft2 of bare tube surface or 2,000-4,000 BTU/hr/ft2 of finned surface.1 (See the next paragraph.) Many heaters have extra space in the convection section to allow the installation of two more rows of tubes. If this extra space is not available, a common solution is to install the new tubes in the breeching leading to the stack. Install tubes with extended surfaces. Extended surfaces in these situations are usually helical fins, although studded tubes were commonly installed in the 1970s and l980s. Studded tubes can transfer 2-3 times more heat than bare tubes. Finned tubes typically transfer 811 times more heat than bare tubes. Fins offer three advantages over studs: they transfer more heat; they cause much less pressure drop and they are cheaper.
Figure 7.1: HEAT TRANSFER IN THE RADIANT SECTION OF DIRECT FIRED HEATER
2

Fraction of Total Heat Absorbed in Radiant Section

70

65

Parameter is the Allowable Heat Flux to the Tubes, BTU/hr/ft2

60

55

50

6,000 Nominal Multiply Pipe Size "G" by inches 2 1.03 3 1.02 4 1.01 6 1.00 8 1.00 10 0.995
12 14 16 18 20 22

9,000 12,000 15,000

45

40

35

30

G, lbs air / lb fuel

24

26

28

30

If considering the installation of tubes with extended surfaces, it is necessary to account for the different sizes of tubes and the tubesheets. If the heater has oil-firing it is necessary to install soot-blowers between every four or five rows of tubes. Note that the first two rows of tubes in the convection section (called shock tubes) receive radiant heat because they see the radiant flame, as well as convective heat transfer. In fact, the first shock row has the highest heat flux in the heater. The shock rows always have bare tubes. The third row of tubes receives radiant heat if it has long radius return bends and it should have bare tubes also. The first row of finned tubes is equipped with fewer, shorter and thicker fins in order to reduce the fin tip temperature.3

1 2

GPSA Engineering Data Book, 10 Edition; page 8-17. th GPSA Engineering Data Book, 10 Edition; page 8-16. 3 th GPSA Engineering Data Book, 10 Edition; page 8-16.

th

ARPEL Environmental Guideline No. 29

70

The Optimization of Combustion in Boilers and Furnaces

Installation of extra tubes will increase thermal efficiency by lowering the flue gas exit temperature. However, as a result, there is a greater flue gas pressure drop and there is less draft. Corrective measures are usually to extend the stack or to install an induced draft fan at the top of the convection section. If the modifications to the heater result in more fuel firing, the ability of the stack to handle the additional flue gas must be checked. With the extra equipment installed, it will be necessary to ensure the structure and foundations can safely support the additional weight.

7.2

Conversion to Forced-Draft Air Supply


Natural draft burners are limited to low pressure drop across the burner - 0.3-0.6 inches of water column (WC). Therefore, the air must be induced at low velocity. This leads to relatively poor air-fuel mixing. Excess air rates must be in the range of 15-20% for gasfiring and 30-40% for oil-firing. Flame lengths are higher for natural-draft burners than for forced-draft burners. Gasfired flames are about one foot per MMBTU/hr of firing. Oil-fired flames are about two feet per MMBTU/hr. Conversion of the air supply to forced-draft fans will increase heater efficiency by improving air-fuel mixing so that there is better combustion with shorter flames that are more stable. Forced-draft fans are a necessity if air preheating equipment is installed. Forced-draft burners are discussed in Section 4.4. Suffice it to say, because of the better air-fuel mixing, use of a high pressure air burner will enable the heater to operate at considerably lower excess air rates (5-10% for gas-firing and 10-15% for oil-firing). See Section 3.4 for the effect of reducing excess air on heater efficiency. In addition, the flames are shorter and wider. This results in a more uniform heat distribution within the firebox, as well as less danger of flame impingement. It is important to remember that under normal circumstances, the conversion of burners from natural draft to forced draft per se cannot be justified in terms of energy savings. However, by taking into account the ability to run the heater harder or at higher capacity, greater safety due to a lower risk of flame impingement and the ability to install air preheating facilities, the conversion often becomes viable.1 A project to convert burners to forced-draft air supply must consider the space required for the fan and ductwork. Included in this consideration is the space beneath the heater. Ducting to the burners (under the heater) and deeper windboxes will be needed.2

7.2.1

Forced-Draft Air Supply and Low-NOx Burners

Flame lengths for burners using natural draft were mentioned in Section 7.2. If the burners were replaced with low-NOx burners (still under natural draft) the flame lengths would be 50-100% longer. Converting to a forced-draft air system will lead to shorter

1 2

A. Garg; Hydrocarbon Processing; June 1998; page 76. A. Garg; Hydrocarbon Processing; June 1998; page 78.

71

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

flames. However, low-NOx burners in a forced-draft system will still have long flames and low excess air rates (i.e., below design levels) will lead to operating problems.1 Conversion of burners to the low-NOx type may be necessary in order to meet air quality regulations. To lessen the risk of flame impingement that may occur because of the longer flame length, more burners can be installed. As discussed above, flame length is a function of the burner heat release. Therefore, more burners, for the same overall heater duty, will result in shorter flames.2

7.3

Waste Heat Recovery


Waste heat recovery is a common method if increasing the efficiency of a heater or boiler. The impact of such equipment is illustrated in Section 3.4. Five possible streams can be heated using waste heat recovery: Combustion air Boiler feedwater Fuel gas Fuel oil Process fluid Of these, the first two combustion air and boiler feedwater are the most common for process heaters and boilers, respectively. The temperature of flue gas leaving the convection section will vary, frequently according to the age of the heater. Temperatures in the range of 350-1500F can be found. The purpose of waste heat recovery is to reduce the flue gas temperature, thereby increasing the thermal efficiency of the heater. Every 35-40F decrease in flue gas temperature is equivalent to a 1% increase in efficiency. The minimum temperature to which a flue gas may be cooled is dependent upon the concentration of acid gases in the stream. Of particular importance is the concentration of SOx, which in turn influences the amount of SO3 in the flue gas. Therefore, the amount of sulfur in the fuel oil and/or H2S in the fuel/natural gas is a parameter that must be monitored closely. Figure 7.2 shows the relationship between sulfur/H2S content and the acid gas dew point. It is drawn from two sources. As indicated in the drawing the literature values for fuel oil (at least) show a diversity. In the absence of fuel-specific dew point data, it is suggested that the highest dew point - gas or oil for the reported sulfur/H2S content be used, in order to be conservative. To prevent corrosion due to localized cooling of the flue gas it is recommended that the metal temperature be kept at least 25F higher than

1 2

Ibid, page 76. Ibid, page 78.

ARPEL Environmental Guideline No. 29

72

The Optimization of Combustion in Boilers and Furnaces

the acid gas dew point.1 Manufacturers of waste heat recovery equipment may have their own design criteria.2
Figure 7.2: ACID GAS DEW POINT AS A FUNCTION OF FUEL SULFUR/H2S CONTENT
330

Maximum Flue Gas Dewpoint for Gas Excess Air = 5-20% (Garg)

310

290

Acid Gas Dewpoint, F

270

250

Maximum Flue Gas Dewpoint for Oil Excess Air = 10-30% (Garg) Fuel Oil (Garcia-Borras)

230

210

190

170

150 0 1 2 3 4 5

H2S in Fuel Gas or Sulphur in Fuel Oil, %

One method of avoiding dew point problems is to heat the cold fluid using a warm stream prior to it entering the preheater. This, of course, requires a warm stream close by. Another option is to have the inlet of the cold fluid (A) at the bottom of the preheater (where the flue gas is the hottest). Once stream A has been warmed to a certain extent so as to avoid dew point problems, it is then routed to the top of the preheater (where the flue gas is coldest).

7.3.1

Combustion Air Preheat

Air preheaters consist of several types: Regenerative Recuperative Circulating liquid The regenerative type is a technology adapted from boilers. A common example of this technology is the Ljungstrom heater. It consists of a slow-moving rotor packed with metal plates or wires. Partitions, and radial/circumferential seals separate the hot flue gas and the combustion air. Flue gas temperatures up to 1500F (1800F, if special alloys are used) and gas velocities of 500 ft/minute can be handled. The effectiveness of these units can reach 85-90%.

1 2

A. Garg; Hydrocarbon Processing; June 1998; page 80. Maxim Heat Recovery Applications Manual; as an example. 3 A. Garg; Hydrocarbon Processing; June 1998; page 78 and Garcia-Borras, page 32

73

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

The main drawbacks to this equipment are there are moving parts, leakage and power consumed turning the rotor. Leakage is due to: Entrainment in the rotor passages (this can be reduced by installing a blow-out section between the hot and cold streams) The gap between the rotor shell and its housing Gaps in the radial seals.1 New types of regenerative units have improved seals. Large preheaters have automatically-activated deflecting sector plates at the hot end.2 Recuperative waste heat recovery units are gas/gas static exchangers. They are often equipped with fins. Some have fins on the inside and outside of the tubes and some on the outside only. The type and location of the fins are determined by the possibility of acid gas dew point problems. A technology that is relatively new to the oil and gas industry, although it has been used in petrochemical plants, is the heat pipe. This is usually a chrome-coated sealed copper tube filled with a volatile liquid. Usually it is ammonia but low boiling point hydrocarbons have been used. A number of these tubes are mounted in a tubesheet that separates the hot and cold fluids. The tubes are mounted at a non-horizontal angle, with the hot end mounted at the lower elevation. There are no moving parts. The hot flue gas passing over the tube vaporizes the ammonia, which flows to the top (cold end) of the tube. The cold air condenses the amount, thereby extracting the heat of vaporization. The liquid ammonia then flows by gravity down the tube and the cycle repeats itself. Projects to install air preheat facilities must take into account the space that will be required for the ducting, the induced-draft fan on the flue gas outlet of the preheater, the forced-draft fan on the air inlet to the preheater, new burners and the preheater itself. The induced-draft fan can be eliminated by installing a circulating-liquid preheater. The liquid is heated in the convection section and heats the air in a heat exchanger. See Section 7.3.4. Recuperative air preheaters are larger and more expensive than regenerative units. However, they require less maintenance. See Figure 7.3 for a typical air preheat arrangement. It is very important to remember that by preheating the air (or fuel), the adiabatic flame temperature increases. This means that the firebox and the tube walls will be hotter. It will be necessary to rerate the heater.3

1 2

Perry; pages 9-63, 9-64. A. Garg and H. Ghosh; Make Every BTU Count; Chemical Engineering; October 1990; page 218. 3 A. Garg; Hydrocarbon Processing; June 1998; page 78.

ARPEL Environmental Guideline No. 29

74

The Optimization of Combustion in Boilers and Furnaces

Figure 7.3: SCHEMATIC ARRANGEMENT OF AN AIR PREHEATER

Cold Flue Gas Stack Damper

Hot Flue Gas

Hot Air

Air Preheater Fired Heater Cold Air

Induced-Draft Fan Forced-Draft Fan

7.3.2

Boiler Feedwater Preheat

This is a very common option in steam plants. They are normally called economizers. For every 10F rise in boiler feedwater temperature, the boiler duty is decreased by about 1%. The tubes, usually arranged in a series-flow bundle are installed in the boiler breeching. The tubes are generally finned. Temperature control is achieved by opening/closing a damper that is installed in a by-pass duct. Boiling in the economizer must be avoided.1

7.3.3

Fuel Preheat

Preheating of heater fuel (oil or gas) inputs energy into the firebox, similar to preheating air. Therefore, the flame temperature will increase and the heater must be rerated. There is the additional safety precautions that must be taken since hot flue gas, which contains some air could come in contact with combustible material if the preheater leaked.

7.3.4

Circulating-Liquid Heat Exchangers

Heat exchangers are one way of supplying heat to combustion air from the flue gas. They are also useful for supplying heat to circulating heat media and even to process and utility streams. (As with fuel preheaters, safety controls must be installed to prevent direct contact of flue gas and hydrocarbon streams. The use of a circulating medium provides a buffer between the two streams.) Circulating-liquid exchangers are commonly used in place of air preheaters if the plant already has a circulating heat medium or if space is limited. The option does not require an induced draft fan or flue gas ducting.
1

Energy Management Handbook, 3 Edition; page 215.

rd

75

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Exchangers are being designed using transfer surfaces made of Pyrex. This allows the exchanger to cool the flue gas below its dew point. Plate exchangers can also cool flue gas to very low temperatures. Both sensible and latent heat is recovered from the water vapor in the flue gas. With proper design, the flue gas leaving the exchanger can still be slightly superheated so that there is no corrosion in the stack.1

7.3.5

Generation of Steam

Waste heat from flue gas is often used to generate steam, by installing tubes in the convection section of the heater. The flue gas can be cooled to within 50F of the boiler feedwater temperature when generating low and medium pressure steam.2 The boiler feedwater temperature must be hot enough to prevent condensation of the acid gas. In many cases an induced-draft fan and flue gas ducting is required to install the waste heat boiler. If using the waste heat to heat boiler feedwater, generally only minor modifications to the process heater convection section are needed. Waste heat boilers are often installed on pyrolysis heaters and steam-reforming heaters. Steam is consumed in both processes.3

7.3.6

Waste Heat Recovery from Incinerators

The ability to recover heat from an incinerator will depend upon the size of the incinerator, the type of waste being burned and the need to have good plume rise and downwind dispersion. Incinerators that completely destroy their wastes have more options available to them. On the other hand, tail gas incinerators in sulfur recovery units destroy H2S but still produce SOx so the plant must retain the ability to emit the stack gases in accordance with the air quality regulations. Nevertheless, energy management opportunities may exist. The reader is referred to the following articles: Chemical Engineering; Capture Heat From Air Pollution Control; by J. Straitz; October 1993; pages 6-14. Hydrocarbon Processing; Recover Heat from Waste Incineration; by V. Ganapathy; September 1995; pages 51-56.

7.4

Heat Sinks for Cyclic Operations


As discussed in Chapter 4, hot oil and molten salt heaters are used to provide heat for services such as circulating heat media and heat regeneration gas for desiccant dryers. When these services are intermittent, there may be opportunities for saving energy. Regeneration of molecular sieve beds and desiccant beds is a cyclic procedure involving normal operation, heating to regenerate the bed followed by cooling prior to switching back to normal operation. Even with two or three beds, there will probably be times when no bed is being heated. It is not practical to shut down the regenerator gas heater

1 2

A. Garg, H. Ghosh; Chemical Engineering; October 1990; page 220. A. Garg; Hydrocarbon Processing; June 1998; page 78. 3 Ibid.

ARPEL Environmental Guideline No. 29

76

The Optimization of Combustion in Boilers and Furnaces

for these relatively short periods of time. Therefore, the regeneration gas is heated and passed over a heat-retentive sink such as lava rock. The stored heat is then used to start the regeneration of the bed in the next cycle. Since the regeneration requires a fixed amount of heat input, the overall firing on the heater can be reduced by the amount of stored heat transferred to the sieve/desiccant bed. Energy use in such cyclic operations can be also reduced by extending the cycle length to the maximum possible, consistent with good operation. For instance, on desiccant beds, the normal operation cycle should be extended until there is water breakthrough. In this case the savings are realized by the reduced number of times that the bed itself (i.e., the molecular/sieve/desiccant) and the bed shell must be heated to the regeneration temperature. The energy savings are relatively minor, compared with the reduced cost of replacing the desiccant or molecular sieve. There are also energy savings in circulating the regeneration gas because a lower flow rate is required.

77

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

8.0

EMISSIONS FROM HEATERS AND BOILERS

Heaters and boilers in the oil and gas industry primarily use three fuels: natural gas, heavy fuel oil and fuel gas. The term fuel gas refers to the off gases produced by many processes. In most circumstances in the upstream industry, the fuel gas is so similar to natural gas that it could be considered to be natural gas. However, the internally-produced fuel gas in refineries can be quite different. Often, it contains alkenes and much more hydrogen. Many heaters have dual-fuel capability. The environmental impact of emissions resulting from combustion is of increasing concern. As a result, initiatives such as energy management, acid rain reduction and ozone abatement are being developed in many jurisdictions in order to improve air quality. This chapter provides emission factors for the fuels commonly used in heaters and boilers. Chapters 9-11 look at emission control technologies for reducing the major air contaminants associated with combustion. It must be stated that these control technologies will, at times, have an adverse effect upon the energy efficiency of the facility. Thus the need for effective energy management is even greater.

8.1

Emission Factors
The published emission factors are generally stated as functions of either the amount of fuel (especially for residual fuels, in the form of weight/volume of oil) or heating value. The data presented here are a synopsis of the information provided in the ARPEL Guidelines for Atmospheric Emissions Inventory Methodologies in the Petroleum Industry. It is important to note that these factors are based upon typical operating conditions and equipment. Wherever possible heater-specific emission factors or direct calculations/measurements should be taken. Table 8.1 lists emission factors for heaters firing natural gas. Table 8.2 deals with emissions resulting from the firing of refinery fuel gas. Two scenarios are provided: low hydrogen-content and high hydrogen-content fuels. The implied compositions are based upon a two-train refinery that had separate fuel gas headers. Note that refinery fuel gas can vary significantly in hydrogen content if there is a catalytic reforming unit, due to the effects of catalyst deactivation.
Table 8.1: EMISSION FACTORS FOR FIRED HEATERS (NATURAL GAS), lbs/MMBTU Heater Duty, MM BTU/hr CO2 CO CH4 NMHC NOx no air preheat 0.0996 0.00267 <9.9 116.93 9.9-99 116.93 0.035 0.00298 0.00279 0.1393 0.5473 0.00298 >99 116.93
1

CAPP Voluntary Challenge for CO2, CH4, NOx; Canadian Ministry of Energy, Mines and Resources for NMHC, CO, N2O, SO2, US-EPA 4/93, Table 1.3-10, for particulates.

79

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

with air preheat N2O Particulates SO2

0.1661 0.00008 0.1000 0.00047

NMHC

Non-methane hydrocarbons
1

Table 8.2: EMISSION FACTORS FOR FIRED HEATERS (REFINERY GAS), lbs/MMBTU Heater Duty, MM BTU/hr Low Hydrogen Content Gas CO2 CO CH4 NMHC NOx no air preheat with air preheat N2O Particulates SO2 High Hydrogen Content CO2 CO CH4 NMHC NOx no air preheat with air preheat N2O Particulates 0.0996 0.00045 69.73 69.74 0.035 0.00050 0.01863 0.1393 0.1661 0.00008 0.1000 0.5473 0.00050 69.74 0.0996 0.00061 119.62 119.64 0.035 0.00068 0.03311 0.1393 0.1661 0.00008 0.1000 0.00037 0.5473 0.00068 119.64 <9.9 9.9-99 >99

Adapted from CAPP Voluntary Challenge for CO2, CH4, NOx; Canadian Ministry of Energy, Mines and Resources for NMHC, CO N2O, SO2; US-EPA 4/93, Table 1.3-10 for particulates.

ARPEL Environmental Guideline No. 29

80

The Optimization of Combustion in Boilers and Furnaces

SOx

Trace

NMHC

Non-methane hydrocarbons

Note that NOx emissions will be a function of the adiabatic flame temperature, which is in turn a function of the fuel composition and the oxygen content of the flue gas. See Section 1.3. Wherever possible actual values should be calculated. Similarly, emissions of carbon monoxide can be estimated using the data presented in Section 9.2. Emission factors for oil firing will depend upon the type of oil (i.e., No. 4, No. 5, or most often, No. 6 Oil). See Table 8.3. Note that some of the factors are expressed in terms of lbs/barrel of fuel oil and others are expressed in terms of lbs/MM BTU. The category heavy metals consists of at least eleven metals, such as antimony, arsenic, cadmium, chromium, cobalt, lead, manganese, nickel and selenium. Approximately 75% of the emissions of heavy metals are nickel. For further information on this subject and others pertaining to emissions from heaters and boilers, see Section 6.1 of the ARPEL Guidelines for Atmospheric Emissions Inventory Methodologies in the Petroleum Industry. Emission factors for distillate fuels have been included. Normally, distillate fuel would not be fired alone. However, it is used as a cutterstock. The values can be used when blending No. 6 oil with cutterstock to make improved fuel oils. See Section 11.1.
Table 8.3: EMISSION FACTORS FOR AIR CONTAMINANTS FROM UNCONTROLLED RESIDUAL 1 OIL COMBUSTION
Units Utility Boilers No. 2 Oil No. 5 Oil No. 4 Oil No. 6 Oil

SO2
SO3 NOx*

Lbs/bbl Lbs/bbl Lbs/bbl

6.66S 0.2418S 2.804 (1.753)** 0.210 0.421

6.31S 0.2418S 2.804 (1.753)** 0.210 0.294

6.66S 0.2418S 2.804 (1.753)** 0.210 0.3926S + 0.130 0.0119 0.0319

CO PM

Lbs/bbl Lbs/bbl

CH4 NMHC

Lbs/bbl Lbs/bbl

0.0119 0.0319

0.0119 0.0319

Industrial Boilers SO2 SO3 NOx*


1

Lbs/bbl Lbs/bbl Lbs/bbl

5.96S 0.0841S 0.841

6.66S 0.0841S 2.313

6.31S 0.0841S 0.841

6.66S 0.0841S 2.313

US-EPA 4/93, Chapter 1.3.

81

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

CO PM

Lbs/bbl Lbs/bbl

0.210 0.084

0.210 0.421

0.210 0.294

0.210 0.3926S + 0.130 0.0421 0.0119

CH4 NMHC All Boilers

Lbs/bbl Lbs/bbl

0.0021 0.0084

0.0421 0.0119

0.0021 0.0084

N2O POM

Lbs/bbl Lbs/MMBTU

0.00456 0.000022

0.00456 0.0000074 0.0000084 0.000161 0.000405 0.00109 0.00329

Lbs/MMBTU Formaldehyde Heavy Metals PM NMHC POM S * ** Lbs/MMBTU

0.000233 0.000405 0.00011 0.00013

particulate matter non-methane hydrocarbons polycyclic organic matter wt % S in fuel If nitrogen content is known, use 0.864 + 4.39 * N, where N is the weight % of nitrogen in the fuel. For vertically-fired utility boilers at full load and 15+% excess air, use 4.416. Bracketed numbers are for tangentially fired utility boilers.

8.2

Effect of Heater Size on Emissions


Note in Tables 8.1 and 8.2 that heater size affects the emissions of methane and NOx. US-EPA Document 4/93, Chapter 1.3, Fuel Oil Combustion, lists emission factors for utility boilers: and industrial boilers. The document defines neither term. It is assumed that a utility boiler is a large fired unit in a steam plant or a public electrical generation facility. The term industrial boiler is assumed to be a smaller boiler in a steam plant, and also applicable to process heaters. Despite the confusion regarding the terminology, it is clear that the size of the boiler is a parameter affecting the quantity of emissions. In their study of NOx emissions from heaters in steam reforming units, Kunz et al found that the US-EPA AP-42 factors overestimated emissions of NOx and CO.1 In earlier publications on this subject (NPRA Paper AM-92-56; Control NOx from Gas-Fired Hydrogen Reformer Furnaces; March 1992 and Hydrocarbon Processing; Control NOx from Furnaces; August 1992) they attribute this over prediction to the fact that the AP-42 factors were derived from studies done on utility boilers and are thus not valid for typical (and smaller) boilers and process heaters. In summary, no definitive and comprehensive relationship between heater/boiler size and emissions was found in the literature. However, studies have found size to have a

Kunz et al; Hydrocarbon Processing; November 1996; page 76.

ARPEL Environmental Guideline No. 29

82

The Optimization of Combustion in Boilers and Furnaces

bearing on the quantity of several categories of emissions. Whenever possible, heaterspecific emission factors should be prepared.

8.3

Effect of Controls on Emissions


The emission factors listed in Section 8.1 are for uncontrolled emissions. That is, no equipment or operating procedures are in place to reduce emissions. The most common control equipment involves measures to reduce emissions of NOx, SOx and particulates. These are listed in Tables 8.4, 8.5 and 8.6, respectively. Descriptions of technologies and procedures are provided in Chapters 9 and 10.

Table 8.4: EFFECTIVENESS OF NOX CONTROL MEASURES FOR RESIDUAL FUEL COMBUSTION
Technique Combustion Modifications Low excess air Staged combustion Reduce excess 02 to 2.5% Fuel-rich burners, secondary combustion air ports Some burners fuel rich, some burners air only 25-30% of the fuel gas recycled to burners 0-28 20-50 Process Reduction Efficiency, %

Burners-Out-Service

10-30

Flue gas recirculation

15-30

Flue gas recirculation plus staged combustion Load reduction Low-NOx burners Reduced air preheat Post-Combustion Modifications Ammonia injection Urea injection Thermal DeNOx Air heater baskets NH3 injected into flue gas Urea injected into furnace NH3 injected into furnace Baskets of catalyst to promote reaction of ammonia with NOx Catalyst in flue gas stream Reduce air and fuel to all burners New burner design Bypass air preheater

25-53

33 to an increase of 25 20-50 5-16

40-70 30-60 30-60 40-65

Selective catalytic reduction (SCR) Duct (SCR)

90

Small version of SCR in existing duct

30

US-EPA 4/93, Table 1.3-12.

83

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Activated carbon on SCR

Carbon catalyst downstream of air preheater

Not available

Table 8.5: POST-COMBUSTION SO2 CONTROLS FOR RESIDUAL OIL COMBUSTION Technology and Chemical Agent Wet scrubber Lime/limestone Sodium carbonate Magnesium oxide/hydroxide Dual alkali Spray Drying Calcium hydroxide slurry Furnace injection Dry calcium carbonate hydrate Duct injection Dry solvent injection 25-50+ 25-50 70-90 80-95+ 80-98 80-95 90-96 Reduction Efficiency %

Table 8.6: REMOVAL EFFICIENCY OF TECHNOLOGIES FOR CONTROLLING EMISSIONS OF 2 PARTICULATES Technology Electrostatic precipitator Scrubber Multiple cyclones Removal Efficiency, % 99.2 94.0 80.0

1 2

US-EPA 4/93, Table 1.3-14. US-EPA 4/93, Tables 1.3-5 and 1.3-6.

ARPEL Environmental Guideline No. 29

84

The Optimization of Combustion in Boilers and Furnaces

9.0
9.1

CONTROLLING EMISSIONS COMBUSTION CONTROLS


Fuel Switching
The emission factors listed in Tables 8.1, 8.2 and 8.3 indicate the impact that fuel type has on emissions. As an example, consider Scenario 5 in Section 3.4. For that case, 40 MM BTU/hr of process heat input is achieved using primarily direct firing, with a minor amount of waste heat recovery. The flue gas was also used to preheat the air and fuels. In Scenario 6, the fuel is switched to 100% gas firing and in Scenario 7, to 100% oil firing. (Fuel qualities are listed in Section 3.4). Table 9.1 lists the emissions for all three cases. The use of gas, because of its higher content of hydrogen, results in considerably less emission of all the listed air contaminants. The efficiency of combustion using gas is slightly lower because full utilization of the heat content of the water vapor formed during combustion cannot be realized. However, no atomizing steam is required, thereby reducing fuel firing in the steam plant. Note that the values in Table 9.1 are for the process heater only. Other benefits of switching from heavy fuel oil to gas are that no corrosion inhibitors, heating of fuel oil lines and tanks and viscosity reducing additives are required. The net effect is that switching to gas reduces fuel firing. Emissions of NOx have been estimated using the factors for oil provided in Chapter 8 and the factors for gas provided in Section 1.3. See Section 9.3 for further details. Emissions of carbon monoxide are based upon Figure 9.1. Carbon dioxide and SOx emissions are calculated using carbon and sulfur balances around fuel and flue gas streams. Table 9.1 is prepared on the presumption that the heater could operate using either gas or oil or both. If the heat release per burner is 250,000 BTU/hr or greater, a dual-fuel burner is usually possible. Oil-derived flames have different heat release patterns because their flames are more luminous than those that are gas-derived. Oil flames require more combustion space because additional time is required to ensure fuel atomization and vaporization. Most conventional general-purpose heaters are designed conservatively, (i.e., large firebox) so there is usually not a problem with excessive flame radiation.1 Another form of fuel switching involves the conversion of No. 6 Oil to No. 5 Oil or even No. 4 Oil. However, there are product yield implications with this choice. The reader is referred to Chapter 3 of the ARPEL Guidelines on the Impact of Fuel Switching on Refinery Operations and Atmospheric Emissions.

North American Combustion Handbook; page 42.

85

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Table 9.1: EFFECT OF FUEL SWITCHING ON HEATER EMISSIONS, lbs/hour Scenario 5 50% Gas, 50% Oil 6911 0.017 0.220 1.819 37.205 0.470 24.235 0.106 4.612 46.756 Scenario 6 100% Gas 5917 0.004 0.130 1.380 0.679 0.009 16.440 0.121 4.376 0 Scenario 7 100% Oil 7906 0.033 0.311 2.277 73.707 0.930 32.315 0.088 4.846 93.476

CO2 N2O CH4 CO SO2 SO3 NOx NMHC Particulates Heavy Metals

NMHC

Non-methane hydrocarbons

9.2

Excess Air vs. CO, NOx, Efficiency


The importance of maintaining good control of the amount of excess air has been discussed in Chapters 3 and 5. There is also an important reason concerning emissions of NOx and carbon monoxide. The decrease in heater efficiency brought about by the heat requirements of the excess air creates an incentive to operate with as low an excess air rate as possible. At the same time, there is not always good air-fuel mixing at low excess air rates. This causes incomplete combustion and there is a steady accelerating increase in emissions of carbon monoxide with falling excess air rates. See Figure 9.1 for the relationship between oxygen in the flue gas and emissions of CO. Fuel type is a factor. Instrumentation exists to control CO emissions in the 150-250 ppm range.

ARPEL Environmental Guideline No. 29

86

The Optimization of Combustion in Boilers and Furnaces


1

Figure 9.1: CARBON MONOXIDE EMISSIONS FROM HEATERS

1000

800

CO Emissions, ppm

Coal
600

Oil
400

200

CO Control Range Gas

0 -1 0 1

Oxygen Content of Flue Gas, vol %

As stated in Section 1.3, the formation of NOx in gas-fired heaters and boilers is a function of the oxygen content of the flue gas and the adiabatic flame temperature (AFT). At very low concentrations of oxygen in the flue gas (i.e., less than 1-2%) there is insufficient oxygen to drive the formation of NOx forward. At concentrations of 4% and higher, there is enough excess air to materially affect the adiabatic flame temperature. This quenching of the AFT outweighs the effect of the increased availability of oxygen. However, at middle concentrations of oxygen (2-4%), there are sufficient oxygen and sufficient adiabatic flame temperature present such that the generation of NOx is usually maximized in that range of oxygen content. See Figure 9.2. Figure 9.2 illustrates the relationship between NOx, CO and heater efficiency at low rates of excess air. Natural gas fuel is illustrated. For fuel oil, the heater efficiency line will be essentially unchanged. The lines for NOx and CO will be moved upward (i.e., greater emissions for the same oxygen content in the flue gas. Note also that the conditions that generate the highest concentration of NOx do not generate the highest emission rate. This is because the amount of excess air is increasing faster than the rate that the NOx concentration is decreasing. See Figure 9.3.

Garcia-Borras; page 31.

87

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Figure 9.2: EFFECT OF OXYGEN ON HEATER EMISSIONS

600

92.5

Fuel is Natural Gas Heater Efficiency, % (LHV Basis)


500 92.0

Emissions, ppm v (dry basis)

Heater Efficiency
400 91.5

300

91.0

NOx
200 90.5

100

90.0

CO
0 0 1 2 3 4 5 6 7 89.5

Oxygen in Flue Gas, volume % (dry basis)

9.3

NOx Abatement
Generally speaking, the lighter the fuel the less the NOx emissions, all other thing being equal. An exception to this, however, are fuels rich in hydrogen. High concentrations of hydrogen result in higher adiabatic flame temperatures. Abatement technologies usually involve combustion control - at the burner or post-combustion control - in the flue gas. There are a great number of technologies available for reducing NOx. Some of these have been listed in Table 8.4. The effectiveness of these options will depend upon whether they address the question of thermal NOx, fuel NOx or both.
Figure 9.3: NOx EMISSIONS FROM HEATERS
400
20

Fuel is Natural Gas


360 320
18

16

Emissions, ppm v (dry basis)

280

14

240

12

ppm v
200 160 120 80 40 0 0 1 2 3 4 5 6 7
10

Oxygen in Flue Gas, volume % (dry basis)

ARPEL Environmental Guideline No. 29

88

Emissions, lb/hour

lbs/hour

The Optimization of Combustion in Boilers and Furnaces

9.3.1

Fuel NOx

There are two types of NOx fuel NO x and thermal NO x. The first is the result of combustion of chemically-based nitrogen compounds in the fuel and the second is the result of thermal fixation of nitrogen with the oxygen in the combustion air. Gaseous fuels contain very little nitrogen compounds, so fuel NOx is essentially a problem encountered only when burning fuel oils. Factors for NO x emissions are provided in Section 8.1. The source document for these factors (US-EPA 4/93, Chapter 1.3) estimates that 20-90% of the nitrogen in residual fuel oil is converted to NOx. This means that more than 50% (and in another part of the document, it is stated 60-80%) of the NOx emissions from residual oil firing is fuel NOx. US-EPA 4/93, Chapter 1.3 is not explicit but it is assumed that the factors for the combustion of residual oil include both thermal and fuel NOx. Fuel nitrogen conversion to NOx is highly dependent upon the fuel-to-air ratio in the combustion zone and, in contrast to thermal NOx formation, is relatively insensitive to small changes in combustion zone temperature. Increased mixing of air with the fuel increases combustion efficiency but it also increases fuel NOx. To reduce this problem, initial air injection to the burners is kept below the stoichiometric amount. This results in a reducing atmosphere at relatively high temperature and long residence time. Fuel nitrogen is then converted to N2, rather than NO.1

9.3.2

Thermal NOx

US-EPA Document 4/93, Chapter 1.3 states that the thermal NOx concentration is a function of the peak temperature (adiabatic flame temperature), fuel nitrogen concentration, oxygen concentration and the residence time at the peak temperature (page 1.3-2). Kunz et al (Hydrocarbon Processing, November 1996) prepared formulae for steam-reformer heaters that showed NOx formation to be a function of the adiabatic flame temperature and the oxygen content of the flue gas. (Since their study involved gaseous fuel, the fuel nitrogen content would be extremely low and fuel NOx would not be a significant factor.) Their formulae are presented in Section 1.3 of this guideline. Production of thermal NOx is increased by better air/fuel mixing. Reduction of thermal NOx focuses on reducing the level of oxygen in the primary flame zone, by reducing the adiabatic flame temperature and by reducing the residence time at high flame temperatures.

9.3.3

Flue Gas Recirculation

A portion of flue gas, preferably less than 600F,2 is recycled back to the burner throat. This achieves two goals: it dilutes the combustion air with dilute material, which lowers the adiabatic flame temperature, and it reduces the concentration of oxygen, which inhibits NOx formation. Figure 9.4 shows the theoretical effectiveness of this technology, if it could be used to its maximum extent. Unfortunately, as outlined in the following paragraph, operational problems limit the amount of flue gas recirculation. Regarding this graph, note that the
1 2

US-EPA 4/93, Chapter 1.3; page 1.3-7. A. Garg; Chemical Engineering; November 1992; page 124.

89

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

x-axis is in terms of the volume of flue gas recirculation per volume of natural gas fired. Normally, the recirculation rate is expressed in terms of percentage of flue gas. As an illustration, roughly 10 volumes of air are required for every volume of gas for stoichiometry. With 15% excess air, the flue gas flow is therefore, 11.5 + 1 = 12.5 times the gas flow. A recirculation ratio of 2 on Figure 9.4 is therefore, equivalent to recirculating 2/12.5 = 16% of the flue gas.
Figure 9.4: EFFECT OF FLUE GAS RECIRCULATION ON NOX EMISSIONS
70

60

50

NOx Emissions, ppm

Recirculation Ratio = Volume of Recirculated Flue Gas per Volume of Natural Gas
40

30

20

10

0 0 1 2 3 4 5 6 7 8 9 10 11

Recirculation Ratio

The amount of flue gas recirculation is usually limited to 15-25% of the flue gas on gasfired units, although a rate of 45% is possible on gas-fired boilers.2 At high recirculation rates, flame instability and high emission rates of particulates occur, especially on oilfired boilers. For this reason, recirculation rates are typically limited to 10-12% for oil firing. There are several drawbacks to flue gas recirculation: Although a very successful technology for gas-firing (estimates in the literature range from 50-85%), it is much less effective for oil-firing (10-15%). This is because it addresses the issue of thermal NOx, whereas the majority of NOx with oil-firing is fuel NOx. This option is usually installed on heaters/boilers with forced-draft air supply. It works best in units with few burners, such as vertical, cylindrical heaters.3 Capital costs can be quite high, especially on retrofits. Adequate space for the additional ducting and fans may be an issue of concern. Continuous oxygen and carbon monoxide analysis is recommended. If the heating value of the fuel is highly variable, a flame safeguard system should be installed. Corrosion in the ducting due to particulates, may occur.1
1 2

NPRA Paper AM-92-56, Figure 9. D. Fusselman, D. Lipsher; Oil and Gas Journal; November 2, 1992; page 48. 3 A. Garg, page 126.

ARPEL Environmental Guideline No. 29

90

The Optimization of Combustion in Boilers and Furnaces

Flue gas recirculation does not affect the overall efficiency of the heater if the temperature of the flue gas leaving the convection zone is the same as that of the recirculated flue gas. However, due to the diluent effect of the recirculated flue gas, heat transfer shifts from the radiant section to the convection section of the heater. This may be a problem for heaters operating near their maximum capacity.

9.3.4

Staged-Burners and Combustion

There are three main variations of this technology: staged-air burners, staged-fuel burners and staged-air combustion. The air or fuel, depending upon the variation, is added in two stages, referred to as primary and secondary. (Some literature sources refer to the air added after primary combustion as tertiary air.) Figure 9.5 shows schematic drawings of staged-air and staged-fuel burners. Staged-air and staged-fuel burners are often called Low-NOx burners. Also listed under this category are low excess air burners and ultra-low NOx burners. These are discussed in Section 9.3.5. Staged-air burners inject all the fuel at the nozzle but the quantity of primary air is kept below the stoichiometric amount. Typically, burners are designed to give a residence time in the primary combustion zone of 35-100 milliseconds.

Fusselman, Lipsher, page 48.

91

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Figure 9.5: STAGED BURNERS

John Zink Company; NOx Control in Fired Heaters; paper by R. Martin, W. Johnson; presented at 1984 Winter National Meeting of the American Institute of Chemical Engineers; Tulsa, Oklahoma; March 11-14, 1984.

ARPEL Environmental Guideline No. 29

92

The Optimization of Combustion in Boilers and Furnaces

Approximately 65-70% of the stoichiometric air is added at the primary zone.1 The fuel partially burns and the nitrogen forms ammonia, hydrogen cyanide and nitrogen oxide, which are subsequently reduced to elemental nitrogen. Incomplete combustion produces a lower adiabatic flame temperature, as does the backmixing of the products of combustion. The secondary air is added through the refractory ports to complete the combustion and to stabilize the flame. NOx reductions of 20-50% have been reported in the literature.2 The technology is relatively inexpensive and has been used in forced-draft heaters and with flue gas recirculation installations. The principal drawback is that the flame tends to be longer and wider than in conventional burners. This may cause flame impingement and thermal stresses. Staged-fuel burners operate with all of the air, but only a portion of the fuel being injected through the burner. There are localized high levels of excess air, resulting in low thermal NOx formation. Among the products of incomplete combustion are hydrogen and carbon monoxide, which reduce part of the NOx formed in the primary stage of combustion. Staged-fuel burners produce lower NOx levels than staged air burners particularly in gasfired heaters. Reduction of NOx is 50-70%. The flame is about 1 imes longer than in t standard burners,3 but shorter than the flame produced in a staged-air burner. Staged-air combustion is similar to the staged-air burner, except that the secondary air is injected into the firebox. This means that the burners operate in a fuel-rich region and the firebox in a fuel-lean region. Three variations predominate: overfire, burners-out-ofservice, biased firing. In the overfire air option, air injection ports are installed above the top row of burners. Burners-out-of-service is a commonly-used technique in retrofit situations where there are multiple burners. In this option, the top row of burners receive air only (i.e., no fuel). In biased firing, only selected burners receive air. This is commonly used if the burnersout-of-service option results in firing limitations. These techniques (which in one literature source4 are all referred to as burners-out-of-service) are not practical for heaters with less than eight burners. Removing isolated single burners or a series of burners in small heaters will cause flame pattern problems, reduced thermal capacity and reduced thermal efficiency.5 The feasibility of staged combustion depends upon the number of burners, the physical dimensions of the firebox in order to install over-fire air ports and the residence time in the firebox. The potential benefits, though, are significant: 60-70% reduction in NOx when firing gas and 40-50% when firing oil or coal.6
1

John Zink Company; Burner Design Parameters for Flue Gas NOx Control; paper by R. Martin, no date; page 15. 2 A. Garg; page 128 states 20-35%; US-EPA 4/93, Table 1.2-12 states 20-50% for residual oil firing. 3 A. Garg; page 128. 4 NOx Control Techniques for the CPI; by D. Lambert, T. McGowan; Chemical Engineering; June 1996; page 101. 5 Ibid. 6 R. McInnes, M. Van Wormer; Chemical Engineering; September 1990; page 133.

93

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

9.3.5

Low NOx Burners

Low-NOx burners take advantage of the reaction kinetics of NOx formation and ensure that high adiabatic flame temperatures and adequate oxygen content do not occur simultaneously. This is frequently achieved by staging either the injection of the fuel or the addition of the combustion air. See Section 9.3.4 for more detail. Two other technologies are low-excess-air burners and ultra low-NOx burners. Low-excess-air burners provide excellent mixing of the air and fuel. They can operate at flue gas oxygen concentrations of 0.5-1.5%. As seen in Figure 9.2, this is a region where the oxygen concentration is limiting so NOx production is relatively low. These burners work best in single-burner heaters. With multiple burners there are difficulties ensuring proper and equal air distribution. They are mechanical or forced-draft burners so they are more likely to be installed on boilers than on process heaters, which are more often equipped with natural draft burners. Savings of 0-28% have been reported. Also, because they operate at lower overall excess air rates, fuel savings of 1-2% are realized.1 Ultra-low-NOx burners combine staged-fuel burners with flue gas recirculation. They represent a new technology. Production of NOx is very low. In general, installation of low-NOx burners is an extremely popular and economical option, especially on new boilers. Retrofits are often more costly because of the changes needed to accommodate the new burners. Low-NOx burners operate in a reducing atmosphere in the firebox. This could lead to accelerated slagging and corrosion unless the tube metallurgy is changed. Moreover, the flames are longer so flame impingement and thermal stresses may pose problems. In terms of emissions, the low excess air and incomplete combustion in parts of the firebox could lead to increased generation of carbon monoxide, and smoke. The impact of low-NOx burners upon NOx emissions, and even fuel efficiency, is substantial. See this section and the previous one for examples. This is also shown graphically in Figure 9.6. The curves are from work by Kunz et al. (Hydrocarbon Processing, November 1996, pages 65-79). The exact type of low-NOx burner(s) involved in the study was not mentioned. Also, the curves are for gas-firing in a steammethane reformer heater. Figure 9.7 shows the effect of low-NOx burners on oil firing as a function of the weight fraction of nitrogen in the feed. Figure 9.8 shows the effect as a function of air preheat temperature. In both cases, high intensity burners are compared with low-NOx stagedair high intensity burners.

R. McInnes, M. Van Wormer, Chemical Engineering, September 1990, page 132.

ARPEL Environmental Guideline No. 29

94

The Optimization of Combustion in Boilers and Furnaces

Figure 9.6: EFFECT OF LOW-NOX BURNERS ON NOX EMISSIONS (GAS FIRING)


350

300

NOx Concentration, ppm v (dry basis)

250

Conventional Burners/Boilers
200

150

Conventional Burners in a Steam-Methane Reformer


100

50

Low-NOx Burners in a Steam-Methane Refomer


0 0 1 2 3 4 5 6 7

Oxygen in Flue Gas, volume % (dry basis)

Figure 9.7: EFFECT OF NITROGEN CONTENT IN FUEL OIL ON NOX EMISSIONS USING 1 STANDARD, HIGH INTENSITY / LOW NOX HIGH INTENSITY BURNERS
1000 900

NOx, ppm (dry basis) corrected to 3% O2

800

Conventional High Intensity Burner


700 600 500 400

Low-NOx High Intensity Burner with Staged Air


300 200 100

Air Temperature 90F


0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0

Fuel Nitrogen, wt%

Burner Design Parameters for Flue Gas NOx Control; Figure 18.

95

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Figure 9.8: EFFECT OF AIR PREHEAT ON NOX EMISSIONS (OIL FIRING) USING STANDARD HIGH 1 INTENSITY AND LOW NOX HIGH INTENSITY BURNERS
600

Nitrogen Content of Oil 0.3 weight % NOx, ppm (dry basis) corrected to 3% O2
500

Conventional High Intensity Burner Oil Firing, 10% Excess Air


400

Conventional High Intensity Burner Gas Firing, 5% Excess Air


300

200

Low-NOx High Intensity Burner Oil Firing, 10% Excess Air Low-NOx High Intensity Burner Gas Firing, 5% Excess Air

100

0 0 100 200 300 400 500 600

Air Preheat Temperature, F

9.3.6

Diluent Injection

In this option, water or steam is injected into the combustion zone. Water is added through nozzles in the windbox. This vaporizes it before it enters the combustion chamber. John Zink found that the most effective means of injecting steam was by mixing it with fuel gas. Steam is added directly to the combustion zone. Significant reductions in NOx emissions can be achieved. One literature source states that a water-to-fuel ratio of 0.5 reduces thermal NOx formation by about 40%. Doubling the water-to-fuel ratio reduces NOx by a further 20-30%.2 John Zink Company conducted a number of tests and found that, by injecting a steam weight-to-fuel weight ratio of 0.2-0.3, NOx reductions of 50% were achieved in a standard gas-fired burner.3 Figure 9.9 shows the effect of steam injection for a natural draft burner firing natural gas. Note that this graph shows a diminishing effect for steam injection ratios of greater than 0.3. The Fusselman and Lipsher article quotes water-to-fuel ratios of 0.21 to 1:1 and steam injection ratios of 1:1 to 2:1.4 The John Zink studies did find that the actual reduction will be a function of the burner design and the concentration of NOx in the flue gas prior to the steam injection.5 In view of the diversity of the results, it is suggested that a series of heater/burner-specific tests be conducted and the optimum injection rate be determined empirically.

Burner Design Parameters for Flue Gas NOx Control; Figure 17. D. Fusselman and D. Lipsher; Oil and Gas Journal; November 2, 1992; page 46. 3 Burner Design Parameters for Flue Gas NOx Control; by R. Martin; John Zink Company; page. 8. 4 D. Fusselman and D. Lipsher, page 46. 5 R. Martin, page 8.
2

ARPEL Environmental Guideline No. 29

96

The Optimization of Combustion in Boilers and Furnaces


1

Figure 9.9: EFFECT OF STEAM INJECTION ON NOX EMISSIONS


1.0 0.9 0.8 0.7

Relative NOx Levels

0.6 0.5 0.4 0.3 0.2 0.1 0.0 0.00

0.05

0.10

0.15

0.20

0.25

0.30

Lbs .Steam/lb. of Fuel (CH4)

For retrofits, injection of water or steam is usually a relatively cheap option. However, it is essential that the water be at least at boiler feedwater quality and that it be completely vaporized before it enters the combustion zone. These conditions are necessary to minimize corrosion. Newer boilers and heaters tend to operate at higher firing temperatures in order to improve fuel efficiency. This results in greater NOx generation. At the same time, many jurisdictions are reducing the allowable NOx emission limits. Therefore, in order to reduce emissions using this technology, higher injection rates are required. This has an adverse effect on fuel efficiency and CO emissions increase.2 As a variation of this technology, at least one company is conducting research into the combustion of oil-water mixtures. Initial results are very promising.

1 2

R. Martin, Figure 5. D. Fusselman and D. Lipsher, page 48.

97

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

10.0 CONTROLLING EMISSIONS-POST-COMBUSTION CONTROLS


Considerable attention has been directed towards reducing emissions from heaters and boilers by treating the flue gas. The most common air contaminants of concern are NOx, SOx and particulates. The first pollutant - NOx - can be treated either at the combustion stage or at postcombustion (flue gas) stage. SOx and particulates are commonly treated at the postcombustion stage. Improvement of fuel quality is another option and is discussed in the next chapter. Figure 10.1 shows schematic drawings of three post-combustion techniques for reducing NOx emissions. Non-selective catalytic reduction and selective catalytic reduction are completely post-combustion processes. Flue gas recirculation is partly post-combustion and partly combustion.

10.1

NOx
Post-combustion technologies for reducing NOx emissions are usually either selective non-catalytic reduction (SNCR) or selective catalytic reduction (SCR). These are expensive options to implement but they achieve very high NOx removal rates and are often required where stringent emission regulations are in force. Selective non-catalytic reduction uses ammonia or a urea reagent to combine with NOx to form an intermediate ammonium salt, which decomposes to elemental nitrogen and water. Injection of these chemicals is either into the firebox or the flue gas duct. Urea is becoming more popular because it is safer and easier to handle. The urea decomposes first to carbon dioxide and ammonia, which then reacts with NOx to eventually form nitrogen and water. Very close temperature control is required. If using ammonia injection, the flue gas temperature should be 1600-1750F. Exxons Thermal DeNOx process has a stated operating range of 1600-2200F, when using ammonia alone. The addition of hydrogen plus ammonia extends the operating range down to 1300F.1 If urea is used the temperature range is much wider, 1000-1900F. If the temperature is too high, the ammonia will preferentially react with the oxygen in the flue gas and actually produce NOx. If the temperature is below the optimum range, the NOx conversion reactions slow down. This will lead to poorer NOx removal and a phenomenon called ammonia slip, in which unreacted ammonia is emitted in the flue gas. This in itself will lead to air quality problems. NOx removal is typically 40-70%. It is an option well suited for retrofits. Capital investment is low, compared with the catalytic version, and less space is needed. In a study of SCR and SNCR for an oil-and-gas-fired boiler, the capital cost of the SNCR option (Exxons Thermal DeNOx process) was only 20% of the cost of the SCR

Exxon Thermal DeNOx Process; brochure; Florham Park, New Jersey, USA; no date.

99

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

scenario.1 The reagent can be injected directly into the boiler/heater. There are no hazardous waste concerns. On the other hand, there are problems with SNCR if close operational control is not achieved. Ammonia slip has been kept as low as 10 ppm2 but if the proper ratio of ammonia to NOx is not maintained, slip can reach 50-100 ppm.3 The ammonia will react with a number of contaminants at low temperatures, such as sulfur and chlorine, to form complex salts. At flue gas temperatures below 600F ammonia reacts with sulfur trioxide and water vapor to form ammonium bisulphate and ammonium sulphate. These can result in fouling of downstream waste heat recovery equipment and the bisulphate can cause corrosion. Periodic water washing easily removes the sulphates. At flue gas temperatures below 250F the ammonia reacts with hydrochloric acid to form ammonium chloride. This may result in a visible plume.4 Selective catalytic reduction (SCR) uses the same principle as selective non-catalytic reduction (SNCR). Either anhydrous or aqueous ammonia can be used. However, the presence of the catalyst reduces the required temperature range to 500-950F. The limits are actually smaller than that as they are linked with the catalyst type. (The catalysts are typically vanadium pentoxide, titanium dioxide or a noble metal. Catalyst shapes include honeycomb plates, parallel ridged plates, rings and pellets.) For vanadium pentoxide the temperature range is only 600-750F.5 On the other hand, Cormetech, Inc. (a joint venture by Corning and Mitsubishi, based in Corning, New York) claims that its SCR catalyst has an operating range of 400-850F and can achieve 90% NOx removal. If the flue gas contains sulfur dioxide, the temperature should be kept above 608F. SCR units are prone to the same adverse reactions that occur with SNCR if the operating temperatures are above or below the design range. The fouling and plugging of the catalyst beds associated with poor temperature control led to considerable changes in the early designs: Ensuring a mixing time of 0.5-1.0 second between the ammonia injection and the catalyst bed. This length of time is usually adequate. Controlling the ammonia injection by monitoring and maintaining the ammonia slip at 5 ppm. Using vanadium and titanium-based catalysts to minimize bisulphate production. Using titanium and zeolite catalysts to resist sulfur poisoning. Placing base-metal catalysts downstream of acid-gas control equipment to prevent corrosion by hydrochloric acid. Using catalysts resistant to HCl is another option.

1 2

McInnes and Van Wormer; Chemical Engineering; September 1990; page 134. Exxon Thermal DeNOx Process. 3 A. Garg; Chemical Engineering; November 1992; page 128. 4 Exxon Thermal DeNOx Process. 5 D. Fusselman and D. Lipsher; Oil and Gas Journal; November 2, 1992; page 49.

ARPEL Environmental Guideline No. 29

100

The Optimization of Combustion in Boilers and Furnaces

Installing soot blowing and water wash equipment to remove fouling by fine particulates. Using parallel flow catalysts in high dust scenarios in order to minimize plugging and erosion by fly ash. (Honeycomb catalysts are usually installed in low dust units where full advantage of the high surface area can be realized.) By maintaining ammonia use between 0.9-1.0 mol of ammonia per mol of NOx, SCR can remove 70-90% of the NOx and leave an ammonia slip of only 5-10 ppm.1 This excellent performance is partially offset by the large capital cost and space required for installation. Another potential problem is the narrow temperature range. It may pose problems trying to place the equipment in the flue gas process flow scheme. Disposal of the spent catalyst must also be considered.

10.1.1

NOx Abatement Considerations

The feasibility of NOx reduction technologies is dependent upon many factors: Does the fuel generate predominantly thermal NOx or fuel NOx? Gas generates thermal NOx and oil produces mainly fuel NOx. Flue gas recirculation does not affect fuel NOx. Low-NOx burners operate very well with gas firing. With oil firing there is also good NOx reduction, but the flue gas concentration may still exceed the air quality requirements. This raises the need for post-combustion treatment. Does the heater have forced-draft or natural draft burners? Boilers are commonly equipped with forced draft air but heaters often use natural draft. Low-NOx burners and flue gas recirculation are more common on heaters/boilers with forced-draft air supply. What space is available for retrofits? This is especially true for selective catalytic reduction processes, flue gas recirculation and, to a lesser extent, selective noncatalytic reduction. Can several heaters use a common technology? For instance, selective catalytic reduction could be installed on a common flue gas duct. It is harder with selective non-catalytic reduction because of the narrow feasible temperature range, but still possible. Do the existing flue gas temperature profiles lend themselves easily to the installation of selective catalytic and non-catalytic reduction? The presence (or absence) of flue gas waste heat boilers and economizers may make the placement of these selective reduction technologies more difficult.

A. Garg; Chemical Engineering; November 1992; page 126.

101

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces


1

Figure 10.1: NOx ABATEMENT TECHNOLOGIES

R. Martin and W. Johnson; NOx Control in Fired Heaters; Figures 3, 4 and 5

ARPEL Environmental Guideline No. 29

102

The Optimization of Combustion in Boilers and Furnaces

10.2

SOx
The removal of SOx from flue gas normally involves one of five process categories: Absorption by non-regenerable solutions Absorption by regenerable solutions Adsorption on a solid bed Direct conversion to sulfur Direct conversion to sulfuric acid. The first category is the most predominant and of this category the most common process is wet lime/limestone scrubbing. Limestone has a lower reactivity than lime, but its greater availability and lower cost make it the preferred reagent. The process is a wet chemical process so there will be some ability to remove particulates but generally particulates removal is low and plugging can occur. For this reason, wet scrubbers are usually installed downstream of filters, electrostatic precipitators or cyclones. This has the added advantage of reducing the volume of waste for disposal. Following particulates removal, the flue gas is sent to a prescrubbing step. Waste streams, preferably containing sodium or calcium reagents, are used at this stage. The main scrubber ensures direct contact of the flue gas with a lime/limestone slurry. The scrubber can consist of various technologies. Spray towers are the most efficient for removing SOx and the pressure drops are low. The main drawbacks are that the scrubbing liquid must be free of solids in order to prevent plugging of the nozzles. Spray towers are ineffective for removing particles five microns and smaller. Baffle scrubbers operate at low pressure. Liquid cascades over a series of baffles while the gas stream passes through the falling curtain of liquid. Multiple stages are used to achieve the desired desulfurization. Plugging is not a problem and up to 90% of particles one micron in size and larger can be removed. Poor liquid/flue gas mixing may occur. In tray scrubbers, the flue gas passes through a perforated tray and then strikes an impingement plate. Even with only one stage, about 90% of the SOx and much of the particulate matter can be removed. Multiple stages are installed for greater removal rates. The major problems are plugging of the orifices, scale formation on the impingement plate and the need to have a moveable plate in order to accommodate varying gas flows. Packed-bed towers have high SOx removal capability but generally cannot achieve good removal of particulates. After the scrubber, the flue gas passes through a mist eliminator and is vented. The sulfur-laden liquid waste must be treated using either natural or forced oxidation and dewatered. The resulting calcium sulfate sludge is mixed with lime and fly ash and sent

103

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

to disposal. (With the forced oxidation process, the calcium sulfate is often sold as gypsum.) The wastewater must be treated. Some designs use the dual-alkali process in which the SO2 is removed by a soluble alkali such as a solution of sodium sulfite or aluminum sulfate. The solution is then removed from the scrubber and reacted with lime or limestone to form insoluble calcium sulfite. The regenerated sodium or aluminum solution is returned to the scrubber. Therefore, loss of the sodium/aluminum reagent is small, the waste is easily disposed of and there is less scaling and plugging. Scrubbers using sodium carbonate or sodium hydroxide do not generate solid wastes. The liquids are sent to either wastewater treatment or deep well disposal. The capital costs are lower than for calcium-based options, but the operating expenses are higher. Magnesium hydroxide is cheaper than the sodium reagents and is therefore, becoming popular. Ammonia scrubbers produce ammonium sulfate, which can be used as fertilizer. High SO2 removal efficiencies are achieved and there is no, or little, wastewater. Seawater can be used for SO2 removal because it is slightly alkaline. The flue gas is first quenched to about 400F and then sent to a two-stage seawater scrubber. The seawater leaving the scrubber is now acidic and is neutralized with fresh sea water. The effluent water is then aerated to form sulfates prior to release to the sea. A major drawback to this process is that the flue gas must be reheated in order to provide adequate plume rise. Obviously, this option is limited to coastal facilities. Dry scrubbing involves less complex mechanical equipment than wet scrubbing. The flue gas passes through a spray of lime or soda ash. The heat of the gas evaporates the moisture in the slurry, producing a fine powder of calcium sulfite and calcium sulfate. The particulate matter is collected in a downstream device such as an electrostatic precipitator or fabric filter. The dry waste is either sent to a landfill or a portion is recycled back to the spray dryer in order to completely react the reagent. The most frequent application is for small and medium-sized boilers that use coal with a sulfur content less that 1.5%.1 They would thus be satisfactory for many process heaters firing residual oils. The process involves less capital costs then other forms of wet scrubbing, but operating expenses are higher. Retrofit applications are common. SO2 removal is in the order of 90%, especially if high molal feedrates of reagent are used. Sodium-based reagents (hydroxide and carbonate) are more reactive than calcium-based reagents (lime and limestone) but are much more expensive. SO2 removal rates of 50-70% can be achieved by injecting dry sorbent into the furnace itself or the duct. The most common sorbents are limestone or dolomite (CaCO3/MgCO3). They are injected into the furnace or into the duct upstream of the particulate removal device.

Rounding Up Sulphur; Chemical Engineering; February 1995; page 81.

ARPEL Environmental Guideline No. 29

104

The Optimization of Combustion in Boilers and Furnaces

By humidifying the unreacted CaO from the sorbent injection into the furnace, or by installing both furnace and duct injection facilities, a process configuration called hybrid sorbent is created. It is capable of achieving 90% SO2 removal.1 As mentioned at the beginning of this section, there are five process categories when considering flue gas desulfurization. The first non-regenerable absorption - has been discussed in some detail since it is the most common category. For more information on other sulfur-removal processes and technologies, the reader is referred to the following documents: ARPEL Guidelines for the Reduction and Control of Gaseous Emissions from Petroleum Refineries, Section 10.3. Rounding Up Sulfur; by V. Kwong and R. Meissner; Chemical Engineering; February 1995; pages 74-83. Descriptions of processes specifically related to flue gas treatment are found on pages 80-83. Sulfur Production Continues to Rise; Chemical Engineering; June 1994; pages 3035.

10.2.1

SOx and NOx

Early NOx removal catalysts deteriorated in performance when flue gas containing sulfur was treated. Two reactions occurred. Titanium-based catalysts reacted with SO2 to form SO3, which in turn sometimes led to increased formation of ammonium bisulfate. Alumina-based catalysts reacted with SO3 to form Al2 (SO4)3, which reduced the surface area (and activity) of the catalyst.2 Improved, sulfur-resistant catalysts were developed. The results fall into two types: Processes that remove NOx only, but are resistant to SOx Processes that simultaneously remove NOx and SOx. Of the first type, Mitsubishi Heavy Industries, in a joint venture with Corning, called Cormetech, developed a sulfur-resistant NOx Selective Catalytic Reduction catalyst. Details regarding this process/catalyst are provided in the ARPEL Guidelines for the Reduction and Control of Gaseous Emissions from Petroleum Refineries, Section 9.1. There are at least three NOx/SOx removal processes available. These are NOx (Noxso Corp.),SNOX (Haldor Tops ) and Desonox (Lentjes and Lurgi). Very high removal e rates are achieved. The Noxso and SNOX processes claim to remove more than 95% of the SOx. Noxso also removes 90% of the NOx. No other performance data were available. More information on these processes can be found in the previously mentioned Kwong and Meissner article (Chemical Engineering, February 1995) and Major Stack-Gas Cleanup Process Trial for Ohio in the September 17, 1990 issue of C & EN, pages 35-36.

1 2

Rounding Up Sulfur; Chemical Engineering; February 1995; page 81. High SOx SCR Experience, by M. Yamamura, T. Koyanagi, (both of Mitsubishi Heavy Industries) R. Iskandar (of Cormetech); no date.

105

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

10.3

Particulates
The importance of good particulates removal when desulfurizing flue gas was mentioned in Section 10.2. There are also health reasons for reducing the emission of particulates. Information regarding the types and effectiveness of equipment for removing particulates is provided in Chapter 12 of the ARPEL Guidelines for the Reduction and Control of Gaseous Emissions from Petroleum Refineries.

ARPEL Environmental Guideline No. 29

106

The Optimization of Combustion in Boilers and Furnaces

11.0 FUEL IMPROVEMENT


The third method of controlling emission from heaters and boilers is to treat the fuel so that it is less likely to produce emissions of NOx and SOx in the first place. Fuel switching is a different concept i.e., the substitution of one fuel for another that is completely different in its emission producing properties. This has been discussed in Section 9.1. Fuel improvement, in this chapter, means the changing of the fuels properties without fundamentally changing the fuel itself. The question of fuel improvement not only involves heater and boiler operation but it has a profound influence on the operation and long-term operability of the facility. The decision to improve fuels used within the facility requires analysis of the future product market (buying or selling), existing and upcoming environmental legislation and the financial resources of the facility/company. For a more detailed discussion of this subject, the reader is referred to the ARPEL Guidelines on the Impact of Fuel Switching On Refinery Operations and Atmospheric Emissions.

11.1

Sulfur Removal
The emission factors listed in Chapter 8 clearly highlight the effect of feed sulfur content on flue gas emissions. Pre-combustion controls involve either blending of high-sulfur fuel with sweeter cutterstock, hydrotreating of the fuel oil and fuel gas sweetening.

11.1.1

Blending

Theoretically sour gas streams could be blended with a sweet gas stream but invariably the concept is limited to the blending of residual oils with a lighter gas oil, or even kerosene. No. 6 Oil can be upgraded to No. 5 Oil and even No. 4 Oil. This is done primarily to reduce the viscosity and sulfur content. However, emissions of virtually all air contaminants will be improved. As an example, assume that 1,000 barrels/day of No. 6 Oil at 1.50 wt % sulfur, 0.3 wt % nitrogen and density of 0.96 is upgraded to No. 5 Oil by adding more kerosene (at 0.2 wt % sulfur, 0.02 wt % nitrogen and 0.82 density). The resulting No. 5 Oil has a sulfur content of 1.20 wt % and a fuel nitrogen content of 0.236 wt %. From a yield point of view, an additional 348 bpd of kerosene must be diverted from the diesel (or similar) pool in order to produce the No. 5 oil. This represents a considerable loss in product value. Secondly, the No. 5 Oil is lighter. The heating value will rise from roughly 17,600 BTU/lb LHV to 17,840 BTU/lb. However, in terms of BTU/barrel, the heating value drops by 2.45%. Therefore, if 1,000 bpd of No. 6 Oil were required for fuel originally, 1,025 bpd of No. 5 Oil would be required in order to supply the same heat input. This leaves a surplus of 348 25 = 323 bpd of No. 5 Oil which must be sold. Table 11.1 outlines the change in emissions.

107

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Table 11.1: EFFECT OF FUEL OIL BLENDING ON EMMISSIONS, lbs/day Contaminant SO2 SO3 NOx CO PM CH4 NMHC N2O POM Formaldehyde Heavy Metals PM NMHC POM No. 6 Oil 9,900 126 2,181 210 719 42 12 4.6 0.04-0.05 1.0-2.4 6-19 No. 5 Oil 8,192 103 1,948 215 432 43 12 4.7 0.06-0.07 1.1-2.4 5-15 No. 4 Oil 6,101 88 1,715 221 309 2 9 4.8 0.08-0.09 1.1-2.4 4-11

particulate matter non-methane hydrocarbons polycyclic organic matter

If the No. 6 Oil is upgraded to No. 4 Oil, the additional kerosene required as cutterstock is 938 bpd. The sulfur and nitrogen contents are 0.92 and 0.175 wt % respectively. To supply the heating needs, 1,051 bpd of No. 4 Oil are consumed leaving a net surplus of 887 bpd of No. 4 Oil. The impact upon emissions is shown in Table 11.1. Blending to a higher grade of fuel oil will improve estimated emissions of all air contaminants except carbon monoxide, methane in some cases, N20, polycyclic organic matter and formaldehyde. This may be due to the insensitivity of the published factors to respond adequately to changes in fuel quality. In practice, since No. 5 Oil and No. 4 Oil have progressively lower viscosity values, there will be better air-fuel mixing and therefore, more complete combustion. Also, because the sulfur content of the fuel is less, there will be a decrease in the acid gas dew point. This may allow the site to increase the amount of waste heat recovery from the flue gas. It is however, unlikely that the overall increase in fuel efficiency will offset the lost product value resulting from the downgrading of the additional cutterstock (kerosene in this case). On the other hand, if air quality standards are stringent and/or disposal of No. 6 Oil is difficult, fuel blending may be feasible. In most cases, fuel switching to gas is chosen.

11.1.2

Hydrotreating

Hydrotreating involves the use of high pressure hydrogen, at good purity and in the presence of a catalyst to remove contaminants such as sulfur, nitrogen and metals from liquid fuels. At the same time, saturation of olefinic and aromatic compounds will occur.

ARPEL Environmental Guideline No. 29

108

The Optimization of Combustion in Boilers and Furnaces

The changing production slates and product quality specifications have placed greater emphasis on this technology. Hydrotreating of heavy gas oils, atmospheric residua and vacuum residua is becoming more common as refiners attempt to improve product quality and upgrade the bottom-of-the-barrel. In terms of fuel for on-site heaters and boilers, hydrotreating of residua is the principal concern. It is unlikely that middle distillates would be hydrotreated and then used as cutterstock. Hydrotreating the residua reduces the amount of cutterstock that is required. The degree of hydrotreating will depend upon a great many factors such as catalyst type and activity feed quality, hydrogen partial pressure, reactor temperature and liquid hourly space velocity. Processing a variety of crude oils, one refinery achieved the following hydrotreating results:1
Removal of sulfur Removal of nitrogen Removal of metals (Ni + V) Decrease in residuum density Viscosity of product at 122F 86-91% 40-52% 76-94% 0.025 0.039 30 52.5 SFS

Hydrotreating results in a decrease in the density of the residuum. This is due to saturation of olefinic and aromatic compounds and cracking of long-chain hydrocarbons. The decrease in density is such that a No. 6 Oil residuum inlet would have the same outlet density range as a No. 5 Oil or a No. 5 Oil inlet would have a No. 4 Oil outlet density. This would greatly reduce the amount of cutterstock. Specifications, as given in ASTM D396, show a viscosity range of 45-300 Saybolt Furol seconds (SFS) for No. 6 Oil and 40 SFS maximum for No. 5 Oil. Not all the viscosity values were below 40 SFS2 so cutterstock would still be needed when running some crudes or at certain operating severity levels. However, the amount of cutterstock would be greatly reduced, and in some cases, eliminated. The impact on emissions could be estimated using the same procedure as used to calculate Table 11.1. The effect on the main emissions is shown in Table 11.2. This table assumes 1,000 bpd of No. 6 Oil with 1.5 wt % sulfur and 0.3 wt % nitrogen. The second scenario has a hydrotreated fuel oil of No. 5 Oil quality with a sulfur level of 0.18 wt % (88% sulfur removal) and a nitrogen content of 0.165 wt % (45% nitrogen removal). Because of the change in density, about 1,025 bpd of the hydrotreated fuel oil is required.

J. Hohnholt and C. Fausto; Refinery Maintains Optimum Yields via Resid HDS; Oil and Gas Journal; January 6, 1986; page 66. 2 Perry; page 9-6.

109

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Table 11.2: EFFECT OF FUEL OIL HYDROTREATING ON EMISSIONS, lbs/day No. 6 Oil SO2 SO3 NOx PM 9,990 126 2,181 719 Hydrotreated Fuel Oil 1,229 16 1,628 432

A comparison of Table 11.1 with 11.2 illustrates the benefits of hydrotreating over blending in terms of heater emissions. Both options will result in improved air-fuel mixing and a lower acid gas dew point, but hydrotreating consumes much less cutterstock, thereby increasing product revenue. However, hydrotreating does involve substantial capital costs and operating expenses in the form of fuel and compression horsepower. For more information on this subject, see the ARPEL Guidelines on Impacts of Fuel Switching On Refinery Operations and Atmospheric Emissions, Section 7.1.

11.1.3

Fuel Gas Sweetening

Treating of fuel gas will not achieve the dramatic change in emissions brought about by the blending/hydrotreating of fuel oils, with the exception of reduced emissions of SOx and CO2. By removing SOx, the acid gas dew point will be reduced, thereby allowing greater opportunity for waste heat recovery. CO2 removal will improve the heating value of the fuel gas and reduce the amount of inert material entering the firebox. This will improve heater efficiency, but only marginally, unless CO2 constitutes a significant portion of the gas, as in the case of pressure Swing Adsorber offgas, which can be as much as 40-45 volume % CO2.

11.2

Flare Gas Recovery


Environmental air quality regulations are becoming more stringent. Facilities that burn No. 6 Oil are faced with the decision of either improving the quality of their residual fuels or ceasing to use them altogether. A common response has been to use natural gas in place of fuel oil. This can be costly both in terms of operating expenses but also in terms of capital outlay if natural gas pipelines to the facilities must be built. Flare gas can be considered as a wasted source of internal fuel (and in some cases, a loss of LPG/NGL). It is a necessary component of many oil and gas facilities. Its primary function is to relieve excess pressure within the plant or individual processing units. Although subject to great variation in composition, flare gas is often quite similar to fuel gas, especially if contaminants such as H2S are removed. Under circumstances outlined below, flare gas can be potentially recovered and used as fuel. By replacing residual oil with gas, emissions of greenhouse gases and other air contaminants from the heaters and boilers will be reduced. Overall emissions from the plant will also be reduced because flaring will be largely eliminated.

ARPEL Environmental Guideline No. 29

110

The Optimization of Combustion in Boilers and Furnaces

There are three types of flaring. The first type is incident flaring. This is characterized by greatly variable flow rates due to major and minor upsets in operations. It can also include recurring problems such as passing pressure control valves and safety valves. The second type of flaring is background flaring. In reality, it is composed of the myriad contributors of flaring, such as passing valves. However, in this situation the flow is relatively constant and constitutes the long-term minimum level of flaring. See Figure 11.1. At oil batteries, the incoming solution gas, if it is not being recovered for its NGL and natural gas values, can be considered as part of the background flaring.
Figure 11.1: HYPOTHETICAL FLARING PATTERNS

Major Upset

Volume of Flaring

Minor Upset Passing Control Valve Passing Safety Valve

Background Flaring Time Periods

Background Flaring

The third source of flaring is due to maintenance actions. It includes depressuring process units or equipment prior to maintenance work and turnarounds. The flare gas very often contains significant quantities of steam and nitrogen. Whether these inert gases will seriously deteriorate flare gas quality will depend upon the amount of maintenance work being undertaken. When considering flare gas recovery, it is recommended that a flare gas reduction program be implemented first. There are two reasons for this: By reducing the amount of gas being sent to the flare the recovery facilities will be smaller, and therefore, cheaper. If the flare gas recovery facilities shut down due to mechanical/operating problems, flaring rates will immediately return to their old levels. Experience has shown that repair of flare gas recovery facilities can sometimes entail a lengthy wait of weeks and even months for spare parts. Details of establishing a flare gas reduction program and task force are provided in the ARPEL Guidelines for the Reduction and Control Of Gaseous Emissions from Petroleum Refineries, Section 6.2, and the ARPEL Guidelines for the Establishment of a Product

111

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

Loss and Waste Accounting and Reduction Programme in Petroleum Refineries, Sections 9.3, 9.4 and 9.5 and Appendix C. Assuming that a flare gas reduction program has reduced flaring as far as possible, the flaring pattern will look like that shown in Figure 11.1. It will be infeasible to attempt to recover all flaring including that during operational upsets. The equipment would have to be extremely large and would be operating practically unloaded for the overwhelming majority of time. The recovery equipment should be sized for the background level of flaring. Appropriate instrumentation and controls are required to ensure that gas - during upsets, etc. is by-passed around the recovery facilities and routed to the flare. Flare gas recovery has proven to be economically feasible at many sites. The following points must be considered: Is there a fuel that can be backed out of the fuel pool? Normally, this would be residual fuel oil. The displaced fuel could also be natural gas, and even fuel gas. Assuming that an existing fuel can be replaced with recovered flare gas, the plant must be able to dispose of the displaced fuel. For imported fuel it involves the cessation of purchases and for internally-produced fuels and involves the expansion of existing markets. Is there sufficient long-term background flaring to justify the installation of the facilities? Is the flare gas of appropriate quality? For instance, if there is too much nitrogen in the gas, fuel gas heating values will drop and this may cause firing problems. If there is too much hydrogen there will also be fuel quality problems and the flare gas recovery compressors may have to be modified. Will the flare gas require treating? Contaminants such as H2S, cyanides and ammonia will have to be removed from the fuel. Other contaminants, such as hydrofluoric acid (from alkylation units) and particulates (from coking units), may have be removed prior to recovery facilities or appropriate changes made to the metallurgy of the recovery facilities. New or expanded gas treatment facilities may be required. Do sufficient and appropriate burners exist to handle the switch from oil-firing to gasfiring? Can NGLs and condensate be feasibly recovered from the flare gas?

ARPEL Environmental Guideline No. 29

112

The Optimization of Combustion in Boilers and Furnaces

12.0 BIBLIOGRAPHY
1. American Petroleum Institute; API Technical Data Book; Washington, DC, USA 2. ARPEL; Guidelines for the Reduction and Control of Gaseous Emissions from Petroleum Refineries; Montevideo, Uruguay; July 1992 3. ARPEL; Guidelines for the Management of Petroleum Refinery Liquid Wastes; Montevideo, Uruguay, July 1992 4. ARPEL; Guidelines for the Management of Petroleum Refinery Solid Wastes; Montevideo, Uruguay, July 1992 5. ARPEL; Impacts of Fuel Switching on Refinery Operations and Atmospheric Emissions; Montevideo, Uruguay; April 1997 6. BP Canada; Variable Operating Pressure for 21-C-6; memo by N. Franklin; Oakville, ON, Canada; December 7, 1976 7. BP Canada; Sidedraw from No. 1 CDU Preflash Column; memo by N. Little; Oakville, ON, Canada; March 8, 1977 8. BP Canada; Preflash Sidedraw; memo by N. Franklin; Oakville, ON, Canada; February 28, 1978 9. BP Canada; 11-C-1, Primary Column Simulation; memo by N. Franklin, Oakville, ON, Canada; April 17, 1978 10. British Petroleum; Crude Oil Distillation: Principles and Practice; M. Auckland and E. Horne; Research Project 139; Sunbury-on-Thames; Middlesex, United Kingdom; June 16, 1967 11. Canadian Association of Petroleum Producers; Guide: Global Climate Change Voluntary Challenge; CAPP Guide to Calculating GHG Emissions; Publication # 1997-0001; Calgary, AB, Canada; April 1997 12. C & EN; Major Stack-Gas Cleanup Process Trial for Ohio; J. Haggin; September 17, 1990 13. Chemical Engineering; Cleaning Up NOx Emissions; R. McInnes and M. Van Wormer; September 1990 14. Chemical Engineering; Make Every BTU Count; A. Garg and H. Ghosh; October 1990 15. Chemical Engineering; Improving Fuel Efficiency with Statistics; T. Mort and I. Verhappen; June 1991 16. Chemical Engineering; Improve Your Fired Heaters; H. Ghosh; March 1992 17. Chemical Engineering; Trimming NOx from Furnaces; A. Garg; November 1992

113

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

18. Chemical Engineering; New Horizons in Distillation; J. Humphrey and F. Seibert; December 1992 19. Chemical Engineering; Capture Heat from Air-Pollution Control; J. Straitz; October 1993 20. Chemical Engineering; Sulfur Production Continues to Rise; G. Parkinson, G. Ondrey and S. Moore; June 1994 21. Chemical Engineering; Rounding Up Sulfur; V. Kwong and R. Meissner; February 1995 22. Chemical Engineering; NOx control Techniques for the CPI; D. Lambert and T. McGowan; June 1996 23. Chemical Engineering; Advanced Distillation Saves Energy and Capital; F. Lestak and C. Collins; July 1997 24. Chemical Engineering; Catalytic Distillation Extends its Reach; K. Rock, G. Gildert and T. McGuirk; July 1997 25. Chemical Engineering; Distillation Internal Matters; N. Chopey and E. Culp; November 1997 26. Cleaver-Brooks; Boiler Types; http://www.cleaver-brooks.com/Boilersa1.html 27. Cleaver-Brooks; Efficiency Facts; http://www.cleaver-brooks.com/Efficiency1.html 28. Cleaver-Brooks; Glossary; http://www.cleaver-brooks.com/GlossAE.html http://www.cleaver-brooks.com/GlossFP.html and

29. Combustion Engineering-Superheater Ltd.; Steam Tables: Properties of Saturated and Superheated Steam, 3rd Edition; Montreal, PQ, Canada; 1940 30. Exxon Research and Engineering Company; Thermal DeNOx Process; Florham Park, NJ, USA; no date 31. Garcia-Borras, T.; Manual for Improving Boiler and Furnace Performance; Gulf Publishing Company; Houston, TX, USA; 1983 32. Gas Processors Association; Publication 21-45; Tulsa, OK, USA; 1981 33. Government of Canada, Ministry of Energy, Mines and Resources; Emission Factors for Greenhouse and Other Gases by Fuel Type: An Inventory; Ad Hoc Committee on Emission Factors; December 1990 34. Government of United States of America, Environmental Protection Agency; Chapter 1.3, External Combustion Sources; Publication 4/93; Washington, DC, USA 35. Hendry, H.; Boiler Efficiency and Testing; John Inglis Equipment Division Publication B 623; Toronto, ON, Canada; no date

ARPEL Environmental Guideline No. 29

114

The Optimization of Combustion in Boilers and Furnaces

36. Hooper, F. and Gillette, R.; Predictive Maintenance of Steam Traps: Combining Demand Side Management and Performance Contracting; http://www.trapo.com/idea-2.htm; presented to International District Energy Association, Indianapolis, IN, USA; June 1995 and updated October 1997 37. Hydrocarbon Processing; Feedforward Air Control for Fuel BTU Changes; E. Vicknair; July 1985 38. Hydrocarbon Processing; Compute Dew Point of Acid Gases; V. Ganapathy; February 1993 39. Hydrocarbon Processing; Understand Boiler Performance Characteristics; V. Ganapathy; August 1994 40. Hydrocarbon Processing; Recover Heat from Waste Incineration; V. Ganapathy; September 1995 41. Hydrocarbon Processing; Challenges in Simulating Heat Exchanger Networks; R. Sigal; October 1996 42. Hydrocarbon Processing; Predict NOx from Gas-Fired Furnaces; R. Kunz, D. Smith and E. Adamo; November 1996 43. Hydrocarbon Processing; Controlling Fired Heaters; W. Driedger; April 1997 44. Hydrocarbon Processing; Optimize Fired Heater Operations to Save Money; A. Garg; June 1997 45. Hydrocarbon Processing; Revamp Fired Heaters to Increase Capacity; A. Garg; June 1998 46. John Zink Company; Combustion and Industrial Burner Application and Design; Tulsa, OK, USA; no date 47. Makansi, J.; Managing Steam: An Engineering Guide to Commercial, Industrial, and Utility Systems; Hemisphere Publishing Corp.; New York, NY, USA; 1985 48. Martin, R.; Burner Design Parameters for Flue Gas NOx Control; John Zink Company; Tulsa, OK, USA; no date 49. Martin, R. and Johnson, W.; NOx Control in Fired Heaters; John Zink Company; Tulsa, OK, USA; presented at 1984 Winter National Meeting of American Institute of Chemical Engineers; March 11-14, 1984 50. Maxim Heat Recovery Equipment, Division of Beaird Industries, Inc.; Heat Recovery Application Manual; Shreveport, LA, USA; no date 51. Maxwell; Data Book on Hydrocarbons; 52. Mitsubishi Heavy Industries; High SOx SCR Experience; M. Yamamura, T. Koyanagi and R. Iskander; no date

115

ARPEL Environmental Guideline No. 29

The Optimization of Combustion in Boilers and Furnaces

53. National Petroleum Refiners Association; Control NOx from Gas-Fired Hydrogen Reformer Furnaces; R. Kunz, D. Smith, N. Patel, G. Thompson and G. Patrick; NPRA Paper AM-9256; presented at the 1992 NPRA Annual Meeting, March 22-24, 1992 54. Natural Gas Processors Suppliers Association; Engineering Data Book, 10th Edition; Tulsa, OK, USA; 1985 55. North American Mfg. Co.; North American Combustion Handbook, 2nd Edition; Cleveland, Ohio, USA; 1983 56. Oil & Gas Journal; Refinery Maintains Optimum Yields via Resid HDS; J. Hohnholt and C. Fausto; January 6, 1986 57. Oil & Gas Journal Special; Several Technologies Available to Cut Refinery NOx; D. Fusselman and D. Lipsher; November 2, 1992 58. Perry, J., Editor; Chemical Engineers Handbook, 4th Edition; McGraw-Hill Book Company; 1963 59. Power Specialties; Back to Basics Steam Traps 101; by http://www.powerspecialties.com/Armstrong_back_to_basics_Traps101.htm D. Fischer;

60. Power Specialties; Energy-Saving Steam Traps Earn Respect at Hls; by R. Wily; http://www.powerspecialties.com/EnergySavingSteaTraps.htm 61. Sauselein, T.; Stationary Engineering; Business News Publishing Company; Troy, MI, USA; 1990 62. Smith, J. and Van Ness, H.; Introduction to Chemical Engineering Thermodynamics, 2nd Edition; McGraw-Hill Book Company, 1959 63. Sulfur; Leading Burner Designs for Sulfur Plants; January-February 1993 64. Turner, W.; Energy Management Handbook; 2nd Edition and 3rd Edition; Fairmont Press; Lilburn, GA, USA, 1997 (3rd Edition)

ARPEL Environmental Guideline No. 29

116

Mission It is our mission to generate and carry out activities that will lead to the creation of a more favorable environment for the development of the oil and natural gas industry in Latin America and the Caribbean, by promoting:

* * *

The expansion of business opportunities and the improvement of competitive advantages of its members. The establishment of a framework to favor competition in the sector. The timely and efficient exploitation of hydrocarbon resources and the supply of its products and services; all this in conformity with the principles of sustainable development.

To accomplish this mission, ARPEL works in cooperation with international organizations, governments, regulatory agencies, technical institutions, universities and non-governmental organizations. Vision ARPEL aims at becoming an international level organization that through its guidelines activities and principles exert an outstanding leadership in the development of the oil and natural gas industry in Latin America and the Caribbean. Objectives

* * * * * * * * *

To foster cooperation among members. To study and assess actions leading to energy integration. To participate pro-actively in the process of development of laws and regulations concerning the industry. To support actions that expand the areas of activity and increase business opportunities. To serve as an oil and gas activity information center. To develop international cooperation programs. To promote a responsible behavior for the protection of the environment, thus contributing to sustainable development. To take care of the oil and natural gas industrys public image. To study and disseminate criteria and opinions on the sectors relevant issues.

Regional Association of Oil and Natural Gas Companies in Latin America and the Caribbean Javier de Viana 2345 CP 11200 Montevideo URUGUAY Phone: (598 2) 400 6993* Fax (598 2) 400 9207* E-mail: arpel@arpel.org.uy Internet web site: http://www.arpel.org

S-ar putea să vă placă și