interaction between a supersonic jet and … · interaction between a supersonic jet and tubes in...

184
INTERACTION BETWEEN A SUPERSONIC JET AND TUBES IN KRAFT RECOVERY BOILERS by Ameya Pophali A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Graduate Department of Chemical Engineering and Applied Chemistry University of Toronto © Copyright by Ameya Pophali 2011

Upload: vokhanh

Post on 20-Apr-2018

222 views

Category:

Documents


2 download

TRANSCRIPT

INTERACTION BETWEEN A SUPERSONIC JET AND

TUBES IN KRAFT RECOVERY BOILERS

by

Ameya Pophali

A thesis submitted in conformity with the requirements for the degree of

Doctor of Philosophy

Graduate Department of Chemical Engineering and Applied Chemistry

University of Toronto

© Copyright by Ameya Pophali 2011

Interaction between a Supersonic Jet and Tubes in Kraft

Recovery Boilers

Doctor of Philosophy 2011

Ameya Pophali

Department of Chemical Engineering and Applied Chemistry

University of Toronto

ABSTRACT

Sootblowing is a process in which supersonic steam jets are used to periodically blast deposits off

heat transfer tubes in kraft recovery boilers. However, sootblowing significantly consumes the

valuable high pressure steam generated by the boiler, hence it should be optimized. A recovery

boiler consists of three convective sections - superheater, generating bank and economizer. The

tube arrangement in these sections, particularly the tube spacing is different from each other.

Moreover, tubes in an economizer are finned. A sootblower jet will interact differently with these

tube arrangements, potentially affecting its strength, and hence deposit removal capability.

The objective of this work was to characterize jet/tube interaction in the three sections of

a recovery boiler. Lab-scale experiments were conducted in which these interactions were

visualized using the schlieren technique coupled with high-speed video, and were quantified by

pitot pressure measurements. This work is the first to visualize the interactions. The offset

between the jet and tube centrelines, the nozzle exit diameter relative to the tube diameter, and the

distance between the nozzle and tube were varied to examine their effects on jet/tube interaction.

Results showed that due to the very low spreading rate of a supersonic jet, a jet (primary

jet) stops interacting with a superheater platen when the jet is only a small distance away from it.

When the jet impinges on a tube, the jet deflects at an angle, giving rise to a weaker ‘secondary’

ii

jet. Due to the large inter-platen spacing, a secondary jet has an insignificant impact in a

superheater. In a generating bank, the primary jet weakens between the closely spaced tubes due

to increased mixing. However, a secondary jet impinges on the adjacent tubes exerting a high

impact pressure on those tubes. The primary jet also weakens between finned economizer tubes,

but remains stronger for a greater distance than in a generating bank. As in the case inside a

generating bank, a secondary jet also impinges on adjacent rows of tubes in an economizer.

The results imply that in a superheater, a sootblower jet must be directed close to the

platens to yield useful jet/deposit interactions, and to avoid wasting steam by blowing between

the platens. In a generating bank, deposits beyond the first few tubes of a row experience a

weaker sootblower jet, and thus may not be removed effectively. However, secondary jets may

contribute to removing deposits from the first few adjacent tubes. They may also induce erosion-

corrosion of those tubes. Secondary jets may also help remove deposits from adjacent rows in a

finned tube economizer. In an economizer, the strength and hence, the deposit removal capability

of a sootblower jet diminish only slightly beyond the supersonic portion of the jet.

A mathematical model was also developed to determine the feasibility of using inclined

sootblower nozzles in recovery boiler superheaters, and suggests that it may be possible to clean

superheater platens more effectively with slightly inclined nozzles.

iii

ACKNOWLEDGEMENT

This work is a result only of the Grace of my Lord Sri Sadguru Sainath Maharaj Ji upon

me. Without His Grace and Blessings, this work would not have been possible at all by me. He

blessed me with strength and patience, because of which I could complete this work. My humble

prostrations before Him.

First, I would like to extend my sincere gratitude to my supervisors Prof. Honghi Tran

and Prof. Markus Bussmann for their continuous guidance, support and encouragement over the

course of my PhD study here at the University of Toronto, Canada. Their excellent supervision

and invaluable advice have benefited me immensely. I have continuously learnt from them about

scientific and technical subjects, as well as leadership and effective communication.

I would like to thank Prof. Kortschot and Prof. Jones for being my committee members

and for guiding me throughout my study. Their constructive feedback has helped shape my final

thesis. I thank my senior colleagues Dr. Andrei Kaliazine and Sue Mao for their help throughout

my study period. I had many fruitful discussions with Dr. Kaliazine on different related and

unrelated topics to my research work. Sue Mao helped me in many of my experiments. My thanks

to Dr. Babak Emami who was a great companion in the lab, and from whom I have learnt a lot.

My sincere thanks to Paul Jowlabar, who always helped me in my experimental endeavours, and

offered his expertise related to machining. I am truly grateful to the entire staff of the Chemistry

Machine Shop, who fabricated the different apparatus which made my research possible.

This work is part of the research consortium on “Increasing Energy and Chemical

Recovery Efficiency in the Kraft Pulping Process” in the Pulp & Paper Center at the University of

Toronto. I deeply acknowledge the financial support of all the members of the consortium.

Finally, I am sincerely and heartily thankful to my family for their tremendous support

and patience. Without them, this thesis would not have been possible at all.

iv

TABLE OF CONTENTS

ABSTRACT ii

ACKNOWLEDGEMENT iv

TABLE OF CONTENTS v

LIST OF TABLES ix

LIST OF FIGURES x

LIST OF APPENDICES xv

NOMENCLATURE xvi

1.0 INTRODUCTION 1

1.1 Fouling in Kraft Recovery Boilers and Sootblowing 4

1.2 Recovery Boiler Tube Arrangements 8

1.3 Thesis Objectives 10

2.0 LITERATURE SURVEY 12

2.1 Fouling and Sootblowing in Kraft Recovery Boilers 12

2.1.1 Deposits in recovery boilers 13

2.1.2 Deposit removal by sootblowing and other techniques 15

2.1.3 Sootblower jet dynamics 21

2.2 Basic Compressible Flow Theory 25

2.2.1 One-dimensional isentropic flow relations 25

2.2.2 Shock and expansion waves 26

2.2.3 Supersonic nozzle flow 30

2.3 The Schlieren Technique 33

2.4 Supersonic Free Jets 36

2.4.1 Jet structure 36

v

2.4.2 Jet oscillation 42

2.5 Impinging Jets 43

2.5.1 Incompressible jet impingement on a cylinder 43

2.5.2 Supersonic jet impingement on a flat surface 47

2.5.3 Supersonic jet impingement on a cylinder 50

2.6 Conclusions from the Literature Survey 51

3.0 EXPERIMENTAL DESIGN AND METHODOLOGY 53

3.1 Scaled-down Nozzle and Tube Bundles 54

3.1.1 Similarity of the lab air jet to an actual sootblower jet 56

3.2 High-Speed Schlieren Flow Visualization System 59

3.3 Pitot Probe and Positioning System 60

3.3.1 Repeatability of measurements 62

3.3.2 Accuracy of measurements 65

3.4 LabVIEW Control and Data Acquisition (DAQ) System 66

3.5 Image Processing 68

4.0 FREE JET CHARACTERIZATION 70

4.1 Jet Structure 71

4.2 Centreline Peak Impact Pressure 72

4.3 Radial Peak Impact Pressure and Jet Spread 74

5.0 INTERACTION BETWEEN A JET AND A SINGLE TUBE 78

5.1 Experimental Parameters 78

5.2 Effect of Offset between Jet and Tube Centrelines 79

5.2.1 Secondary jet angle versus offset 82

5.3 Effect of Tube Size and Distance between Nozzle and Tube 84

5.3.1 Formation of secondary jets and their failure to form 84

5.3.2 Alternate rise and fall of secondary jet angle with distance 87

vi

5.3.3 Unsteadiness of flow around tube 88

6.0 INTERACTION BETWEEN A JET AND TUBE ARRANGEMENTS 90

6.1 Interaction with Model Superheater Platens 91

6.1.1 Model superheater platens 91

6.1.2 Effect of offset 92

6.1.3 Jet midway between platens 95

6.2 Interaction with a Model Generating Bank 97

6.2.1 Model generating bank 97

6.2.2 Effect of offset 99

6.2.3 Jet midway between two rows of tubes 102

6.3 Secondary Jets 108

6.3.1 Experimental apparatus and procedure 109

6.3.2 Secondary jet structure 110

6.3.3 Secondary jet peak impact pressure 112

6.4 Interaction with Model Economizer Tubes (Finned Tubes) 115

6.4.1 Model economizer section 115

6.4.2 Effect of offset 116

6.4.3 Jet midway between two rows of tubes 119

6.5 Practical Implications – Effects of Formation of Secondary Jets and

Closer Tube Spacing 122

7.0 SCHLIEREN VISUALIZATION OF SYNTHETIC DEPOSIT BREAKUP

BY JET IMPINGEMENT 127

7.1 Synthetic Deposit and Experimental Procedure 128

7.1.1 Synthetic deposit 128

7.1.2 Experimental procedure 129

7.2 Deposit Breakup Images 130

vii

8.0 FEASIBILITY OF USING INCLINED SOOTBLOWER NOZZLES IN 134

RECOVERY BOILER SUPERHEATERS

8.1 Loss in Jet Penetration versus Nozzle Inclination Angle, 137

8.2 Schlieren Visualization of Inclined Jet Impingement 138

9.0 CONCLUSIONS, CONTRIBUTIONS, AND RECOMMENDATIONS 140

9.1 Conclusions and Practical Implications 140

9.2 Contributions of this Work 142

9.3 Recommendations for Future Work 144

REFERENCES 145

APPENDICES 154

Appendix A: Schlieren Images of Jet Temporal Development 155

Appendix B: LabVIEW Graphical Program to Control the Data Acquisition System 156

Appendix C: Interaction between a Jet and a Single Tube: Supplementary Results 157

Appendix D: Typical Kraft Recovery Boiler Tube Arrangements: Results of a Recent

Survey (August 2010) 163

Appendix E: Schlieren Images of a Jet Midway between Two Rows of Finned

Economizer Tubes 165

viii

LIST OF TABLES

Chapter 2

Table 2.1 Brittle deposit breakup mechanisms [68]. 16

Chapter 3

Table 3.1 Performance of pitot probe – comparison of calculated and measured nozzle exit PIP.

66

ix

LIST OF FIGURES

Chapter 1

Figure 1.1 A two-drum kraft recovery boiler. 3

Figure 1.2 Plugging of the flue gas passages at the generating bank inlet in a recovery boiler [107].

4

Figure 1.3 A sootblower removing deposits from a row of tubes. 5

Figure 1.4 Schematic showing the internal structure of a sootblower head. 6

Figure 1.5 A long retractable sootblower [54]. 6

Figure 1.6 Typical tube arrangements in the superheater, generating bank, and economizer sections of a recovery boiler.

8

Figure 1.7 Arrangement of superheater platens in a recovery boiler [2]. 9

Chapter 2

Figure 2.1 Asymmetric deposits formed on superheater platens by carryover impaction [107].

13

Figure 2.2 Massive deposit buildup between superheater platens. 13

Figure 2.3 Jet PIP required for deposit removal versus deposit thickness; (a) head-on impingement; (b) impingement at 90° relative to the head-on case [53].

18

Figure 2.4 Deposit thermal shock caused by supersonic jet impingement [38]. 20

Figure 2.5 Computed peak impact pressure along the centerline of subsonic and supersonic jets with the same mass flow [21].

22

Figure 2.6 (a) Normal shock wave; (b) oblique shock wave; (c) expansion wave.

27

Figure 2.7 Ratio of pitot pressure across an oblique shock wave as a function of shock angle and upstream Mach number [28].

29

Figure 2.8 Reflection of an oblique shock wave from (a) a solid wall, and (b) a constant pressure free boundary.

30

Figure 2.9 Supersonic jets; (a) fully expanded; (b) overexpanded; (c) underexpanded.

32

x

Figure 2.10 Planar refraction of light passing through a region with a negative vertical refractive index gradient [84].

34

Figure 2.11 A conventional 2-mirror z-type schlieren system [84]. 35

Figure 2.12 A mildly underexpanded jet [18]; (a) schematic of jet structure; (b) sonic jet with PR = 1.59.

39

Figure 2.13 A strongly underexpanded jet [18]; (a) schematic of jet structure; (b) sonic jet with PR = 4.09.

39

Figure 2.14 Acoustic feedback mechanism producing screech noise and oscillations in a rectangular jet (instantaneous schlieren image, [76]).

42

Figure 2.15 Oscillation modes exhibited by a screeching jet generated using a Mach 1.41 convergent-divergent nozzle [83].

44

Figure 2.16 Parameters governing the flow field of a jet impinging on a cylinder.

44

Figure 2.17 Flow field of a jet impinging on a cylinder placed away from the nozzle; the level of turbulence in the impinging jet is high.

45

Figure 2.18 The Coanda effect; (a) a free jet in the absence of a curved surface; (b) a jet attached to a cylinder due to the Coanda effect.

46

Figure 2.19 Supersonic jet impinging on a flat surface. 48

Figure 2.20 Bubble of recirculating fluid between the plate and plate shock [18]. 48

Figure 2.21 The shock layer, transonic zone, and beginning of the near wall jet during supersonic jet impingement on a flat plate [15].

49

Chapter 3

Figure 3.1 Experimental apparatus; (a) schematic; (b) photograph. 55

Figure 3.2 (a) Parabolic mirror used in the schlieren system, with custom-designed stand; (b) high-speed camera.

60

Figure 3.3 Pitot probe designed to measure supersonic jet PIP; (a) photograph; (b) schematic showing probe internal design.

61

Figure 3.4 Typical voltage signal of jet peak impact pressure, obtained using the pitot probe.

62

Figure 3.5 Pitot probe positioning system (inset shows the magnetic stand used to clamp the probe to reduce vibration).

63

Figure 3.6 (a) LabVIEW control and data acquisition (DAQ) system diagram; (b) photograph of the control system hardware.

67

xi

Figure 3.7 Sample of a contrast-enhanced average image. 69

Chapter 4

Figure 4.1 Supersonic jet used in this work. 71

Figure 4.2 Centreline PIP variation in the supersonic jet. 73

Figure 4.3 Radial PIP profiles of the jet at different axial locations. 75

Figure 4.4 (a) Definition of jet radius; (b) jet radius versus axial distance from nozzle exit (jet spread); data from [47] is shown for comparison.

77

Chapter 5

Figure 5.1 Jet impinging on a tube at different offsets. 80

Figure 5.2 Formation of secondary jets. 81

Figure 5.3 Secondary jet angle versus offset. 83

Figure 5.4 Effect of nozzle-tube distance on jet/tube interaction for three tube sizes (offset = 0).

85

Chapter 6

Figure 6.1 Typical layout of boiler tubes. 90

Figure 6.2 Model superheater platens. 92

Figure 6.3 Jet impingement on a platen at different offsets. 93

Figure 6.4 Jet PIP exerted near the surface of a model superheater platen as a function of offset and distance; probe at (a) 72 mm; (b) 103 mm; (c) 151 mm from nozzle (po is the same for all cases) [39].

94

Figure 6.5 (a) Jet midway between two platens – no interaction; (b) jet touching one platen – interaction can be seen (flow is from right to left in both cases).

96

Figure 6.6 Jet midway between superheater platens. 96

Figure 6.7 Model generating bank. 98

Figure 6.8 Jet flow into a model generating bank, at different offsets. 100

xii

Figure 6.9 Flow field inside a model generating bank, away from the nozzle, for jet impingement at an offset (ε/R = 1.05, image e in Figure 6.8); the nozzle is to the right hand side.

101

Figure 6.10 Peak impact pressure profiles of a jet midway between model generating bank tubes.

102

Figure 6.11 Flow of a jet midway between model generating bank tubes; the dashed arrows indicate the local flow direction.

104

Figure 6.12 Jet midway between generating bank tubes. 106

Figure 6.13 Flow midway between two rows of tubes farther from the nozzle. 108

Figure 6.14 Experimental module with tube oriented vertically in front of the nozzle for visualizing secondary jets and measuring their centreline PIP.

109

Figure 6.15 Secondary jet at 0.75R offset for a 13 mm (0.5”) OD tube (de/D = 0.58); (a) tube horizontal in front of the nozzle; (b) tube vertical in front of the nozzle.

111

Figure 6.16 Secondary jets at different offsets for a 13 mm (0.5”) OD tube (de/D = 0.58).

113

Figure 6.17 Centreline peak impact pressure of secondary jets at different offsets, for a 13 mm (0.5”) OD tube (de/D = 0.58); the primary jet peak impact pressure is shown for comparison.

114

Figure 6.18 Model economizer tubes: (a) schematic of a row; (b) tube assembly. 116

Figure 6.19 Jet impinging on economizer tubes at different offsets. 118

Figure 6.20 Secondary jet angle versus offset for the superheater, generating bank and economizer tube arrangements.

119

Figure 6.21 Peak impact pressure profiles of a jet midway between model economizer tubes.

120

Figure 6.22 Impingement of a secondary jet on a tube behind the first tube of a generating bank row.

125

Chapter 7

Figure 7.1 Entrained Flow Reactor (EFR) at the University of Toronto; (a) schematic; (b) photograph (tube is located near the EFR exit); (c) carryover deposit formed on a tube using the EFR.

128

Figure 7.2 Breakup of a synthetic deposit by jet impingement, visualized using the schlieren technique (continued on the next page).

132

xiii

Chapter 8

Figure 8.1 Centreline peak impact pressure of primary and secondary jets; the primary jet can be considered as exiting from an inclined nozzle, whereas the secondary jets result from the impingement of a jet from a straight nozzle (data shown is the same as in Figure 6.17).

135

Figure 8.2 (a) Sootblower jet from inclined nozzles; (b) loss in jet penetration between platens due to inclination angle α.

136

Figure 8.3 Behaviour of h as a function of α. 137

Figure 8.4 Effect of nozzle inclination angle α on jet/platen interaction. 139

xiv

LIST OF APPENDICES

Appendix A Schlieren Images of Jet Temporal Development 155

Appendix B LabVIEW Graphical Program to Control the Data Acquisition System

156

Appendix C Interaction between a Jet and a Single Tube: Supplementary Results

157

Appendix D Typical Kraft Recovery Boiler Tube Arrangements: Results of a Recent Survey (August 2010)

163

Appendix E Schlieren Images of a Jet Midway between Two Rows of Finned Economizer Tubes

165

xv

NOMENCLATURE

a Local speed of sound

A Area

c Speed of light in a medium

co Speed of light in vacuum, 3 × 108 m/s

cp Specific heat of gas at constant pressure

cv Specific heat of gas at constant volume

d Diameter

dc Characteristic length

ddeposit Outer diameter of a model deposit

dj Jet diameter at an axial location along the jet

D Outer diameter of tube

f Frequency of vortex shedding just downstream of a cylinder submerged in unlimited parallel flow

Fjet Sootblower jet force

h Length of a superheater platen that is exposed to a sootblower jet from an inclined nozzle

h Deposit thickness

H Side-spacing between boiler tubes

k Coefficient for turbulent transfer of momentum

K Gladstone-Dale constant

l Tube length

L Length of a typical row or platen of boiler tubes

ṁ Mass flow rate

xvi

Mj Fully expanded jet Mach number, in [50]

MW Gas molecular weight

Ma Mach number, Ma = u/a

Mac Convective Mach number, in [63]

Man Normal component of Mach number

n Local refractive index, n = co/c; also, sample size (number of measurements in a sample or data set)

p Pressure

ppit Pitot pressure

q Member (or a single measurement) of a sample or data set (of measurements)

q Average of a sample or data set (of measurements)

r Radial coordinate

R Outer radius of tube

Rgas Specific gas constant, Rgas = RU/MW

RU Universal gas constant, RU = 8.314 kJ/kg-K

Re Reynolds number

S Front-to-back spacing between boiler tubes

St Strouhal number, St = fd/2u

t Time

T Temperature; also, fin thickness

u Axial velocity

ou Axial velocity normalized by nozzle exit velocity, in [44]

W Fin width

x Axial coordinate

xc Jet core length

xvii

x Axial coordinate normalized by nozzle exit radius, in [44]

Greek Symbols

α Nozzle inclination angle

Oblique shock wave angle

Offset between jet and tube centerlines; also, deflection angle of refracted light rays

Specific heat ratio of a gas, = cp/cv

Dynamic viscosity

Rotational speed of sootblower

Prandtl-Meyer expansion wave angle

max Maximum value of a dimensionless deposit ‘shape function’, in [35]

Density

e Density normalized by nozzle exit density, in [44]

adh Deposit adhesion strength

t Deposit tensile strength

Secondary jet angle defined in this thesis; also, flow deflection angle

Prandtl-Meyer function,

1tan)1(1

1tan

1

1)( 2121

MaMaMa

Subscripts

a Ambient conditions

e Nozzle exit

i Measurement number in a sample or data set (of measurements)

j Local jet conditions

xviii

xix

o Total or stagnation conditions

t Nozzle throat

x Axial location

1 Conditions upstream of a shock or expansion wave; also, initial state of a process

2 Conditions downstream of a shock or expansion wave; also, final state of a process

Abbreviations

EFR Entrained Flow Reactor

ID Inner diameter

OD Outer diameter

PIP Peak impact pressure

PR Pressure ratio, PR = pe/pa

SD Standard deviation

SE Standard error

CHAPTER 1

INTRODUCTION

The pulp and paper industry is one of the most important industries of Canada’s manufacturing

sector, and is responsible for the production of pulp, paper, paperboard, and other paper related

products. This industry was the most energy-intensive manufacturing sub sector in Canada in

2008 [91], with an energy consumption of almost 26% (588 petajoules) of the manufacturing

sector’s total consumption. As a result, energy efficiency is a crucial requirement of this industry.

Paper is manufactured from wood pulp, which is obtained by separating wood fibres from

their binding agent, lignin. There are several different methods of producing pulp, out of which

the kraft process is the most widely used. About 67% of the total pulp produced in a year globally

is generated by kraft pulp mills. The kraft process can accommodate a wide range of wood

species, and paper manufactured using this process is strong.

In the kraft process, wood chips are cooked with sodium hydroxide (NaOH) and sodium

sulphide (Na2S) (a mixture called white liquor) at high temperature and pressure, to produce pulp.

A by-product of this process is a mixture of the organic and inorganic process chemicals called

black liquor, from which the cooking chemicals are recovered. Consequently, the kraft process

1

2

consists of two main cycles - pulping and chemical recovery. The recovery boiler is one of the

most important components of the recovery cycle, and is used to burn black liquor for two main

purposes - to recover the inorganic cooking chemicals used in the pulping process, and to make

use of the chemical energy in the organic portion of the liquor to generate steam for the mill

[105]. Figure 1.1 schematically shows a two-drum kraft recovery boiler. Black liquor is burned in

the furnace region at the bottom of the boiler. At the top, the boiler contains three convective heat

transfer sections consisting of tubes – economizer, generating bank, and superheater sections.

Feedwater to be converted to steam flows through the tubes in these sections, from the

economizer to the superheater. The hot flue gases resulting from black liquor combustion travel

upwards from the furnace region and between the tubes, transferring heat to the feedwater and

converting it into high pressure superheated steam. This steam is sent through a turbine to

generate electricity for the mill, and the low pressure steam exiting the turbine is used in process

applications around the mill.

The kraft process produces about 1.5 tons of black liquor dry solids (black liquor without

water) per ton of pulp produced. For every ton of dry solids fired in the boiler, about 3.5 tons of

high pressure steam are generated by the recovery boiler. Depending on the steam quality and

type of turbine, a 1000 ton per day kraft pulp mill can generate 25 to 35 MW of electricity by

burning 1500 tons per day of black liquor dry solids in its recovery boiler [105]. The recovery

boiler typically generates about 60% of the electricity needed by the mill. Increasing the

efficiency of the boiler is always desired to improve the energy efficiency and self-sufficiency of

the mill, because the boiler is the main bottleneck in pulp production, and one of the main reasons

for this is fouling.

3

Superheated steam

Figure 1.1. A two-drum kraft recovery boiler.

Furnace region

Feedwater

Boiler tubes

Hot flue gases with carryover

and fume

Black liquor

Air

Generatingbank

Economizer

Superheater

Smelt

Superheated steam

Furnace region

Feedwater

Boiler tubes

Generatingbank

Superheater

Economizer

Hot flue gases with carryover

and fume

Black liquor

Air

Smelt

4

1.1 Fouling in Kraft Recovery Boilers and Sootblowing

Recovery boiler flue gases are laden

with two basic types of fly ash particles,

carryover and fume, resulting from

black liquor combustion. Carryover

particles result from the mechanical

entrainment of black liquor droplets or

fragments of burning droplets into the

flue gases, and are relatively large (20

μm - 3 mm). Fume particles form when

vapors of sodium or potassium

compounds in the flue gases condense,

and are much smaller than carryover

particles (0.1 μm - 1 μm) [107]. Both

types of particles rise with the flue gases from the furnace region to the boiler tubes where they

form deposits by different mechanisms. These deposits have a low thermal conductivity, and so

restrict heat transfer from the hot flue gases to the boiler tubes, and lower the boiler thermal

efficiency. If their accumulation on tubes is not controlled, these deposits grow with time and

may completely block the flue gas passages (Figure 1.2). The boiler must then be taken off-line

for a water wash, which stops production and is very costly. Recovery boiler fouling is a

persistent and serious problem for pulp and paper mills, and mills strive to avoid or delay a water

wash.

deposit tubes (6.4 cm OD)

Thus, to prevent boiler plugging and increase boiler runtime between shut-downs, it is

absolutely necessary to remove deposits at the same rate at which they form. To achieve this,

sootblowers are operated continuously. Figure 1.3 shows a sootblower removing deposits from a

Figure 1.2. Plugging of the flue gas passages

at the generating bank inlet in a recovery boiler

[107].

deposit tubes (6.4 cm OD)

5

sootblower

supersonic jet

Figure 1.3. A sootblower removing deposits from a row of tubes.

row of tubes. A sootblower consists of a long, hollow steel tube called the lance, with supersonic

nozzles at its working end. Superheated steam generated by the boiler is supplied to the

sootblower. Many such sootblowers are continuously operating within the boiler at different

locations. They rotate as they move into and out of the boiler (see Figure 1.3), and the supersonic

steam jets generated by the nozzles knock deposits from the boiler tubes. In simple terms, a jet is

a stream of fluid which exits from some kind of nozzle or orifice, with a velocity greater than the

surroundings into which it exits. A supersonic jet has velocity greater than the speed of sound.

Figure 1.4 shows the internal structure of a sootblower head. The lance typically has an

outer diameter of 9-10 cm (3.5”-4”). The nozzles are convergent-divergent or de Laval nozzles,

oriented in opposing directions to balance the hydrodynamic forces of the jets on the lance. The

throat diameter typically varies from 2.2-3.2 cm (0.875”-1.25”). Figure 1.5 shows a long

retractable sootblower used in recovery boilers.

Sootblowers are used not only in kraft recovery boilers but also in coal and biomass-fired

power boilers and in waste incinerators to prevent fouling. However, sootblowing requirements

vary depending on the type of boiler. In coal-fired power boilers, plugging is less of a concern

boiler tubes

deposit

rotational speed,

sootblower

supersonic jet

boiler tubes

deposit

rotational speed,

sootblower

supersonic jet

boiler tubes

deposit

rotational speed,

6

Steam

Lance

Supersonicnozzle

Jet

Steam

Lance

Supersonicnozzle

Jet

Figure 1.4. Schematic showing the internal structure of a sootblower head.

4” OD

housingsootblower

head

4” OD

housingsootblower

head

Figure 1.5. A long retractable sootblower [54].

because of the low fuel ash content, and because deposits are relatively weak and not tenacious;

thus sootblowers are not operated continuously. On the other hand, the ash content of the black

liquor is high, and deposits in recovery boilers are much stronger due to their low melting

temperature. In some boiler locations, deposits are sticky and tenacious, but over large portions of

the boiler, deposits are hard and brittle [107]. Sootblowers must usually be operated continuously

(in a cycle). Consequently, a substantial amount of costly high pressure steam generated by the

boiler, typically between 3-12% is used by the sootblowers, that otherwise would contribute to

steam generation and hence, power generation. As a result, optimizing sootblowing to minimize

steam consumption and to maximize deposit removal is important.

A sootblower jet may remove a deposit either by creating internal stresses in the deposit

which exceed its tensile strength (brittle breakup), or by creating stresses at the interface between

7

the deposit and the tube which exceed the adhesion strength of the deposit (debonding), or by

both ways. Conventionally, the performance of the jet in removing deposits has been correlated

with the jet peak impact pressure (PIP) [32], which is the pressure a pitot tube would measure

when inserted into the jet at its centerline. It is also the pressure that the jet would exert on a

deposit. However, the jet force is also an important quantity. The PIP is the pressure at a single

point, whereas the force is the integral of the pressure across the jet cross-section. As a result, the

PIP decreases faster with distance than the force, because the axial momentum is conserved in the

jet.

Due to the fast decay of PIP with distance from the nozzle, the interaction between the jet

and a tube/deposit is the strongest when the jet impinges on the tube/deposit head-on (i.e. when

the jet impinges on the tube/deposit orthogonally), and hence, this interaction is the most

important. This interaction weakens as the sootblower rotates, because the effective distance

between the nozzle and tube/deposit increases as the jet impinges on the tube/deposit at an angle.

Based on trigonometric calculations and decrease of jet force with distance, Tandra [97] proposed

that the most effective zone of cleaning for rotating sootblowers is between 45° with respect to

the head-on or strongest impingement position. The probability of deposit removal decreases

outside this zone.

During operation, a sootblower jet propagates between different tube arrangements. Since

the jet is supersonic, it is sensitive to any obstacle or disturbance in its flow. An obstacle in a

supersonic flow creates a series of complicated shock and expansion waves, which, in the case of

a jet, can directly affect the jet structure, and hence jet strength (PIP) and penetration. If the PIP is

reduced, then the jet may not be able to remove deposits, particularly those away from the nozzle.

As a result, just as it is important to understand how a sootblower jet breaks and removes

deposits, it is also important to understand how the jet interacts with the tubes in different

arrangements found in a recovery boiler, so that such information could be used to direct the jet

onto deposits to yield maximum impact. However, a sootblower jet and its interaction with tubes

8

have never been visualized to date, mainly because of the hostile conditions inside the boiler, and

because these jets cannot be seen by the naked eye, or via regular photographic process.

1.2 Recovery Boiler Tube Arrangements

The tube arrangements in the three

convective sections of a recovery boiler

are very different. Figure 1.6 illustrates

these tube arrangements.

The superheater section is the

section closest to the furnace. Here, the

temperature is very high, around 900°C in

the portion closest to the furnace.

Consequently, tubes are arranged in

platens, where a platen is a tube sheet with

in-line tubes of zero front-to-back spacing.

The tubes typically have an outer diameter

(OD) of 50 mm (2”), and are 10-20 m

long. The side spacing between two

platens is large, typically 254-305 mm

(10”-12”). This alleviates some of the

problems associated with fouling (such as

plugging of flue gas passages with

deposits), and establishes a flow path for the flue gases. These platens are held on supports

suspended by hinge joints bolted to the boiler ceiling (Figure 1.7), and are free to swing. As a

result, when a sootblower jet impinges on the first tube of a platen, the jet force causes the platen

Figure 1.6. Typical tube arrangements in the

superheater, generating bank, and

economizer sections of a recovery boiler.

9

to swing. A platen also swings when the jet blows between

two platens. This is because the jet creates low pressure

between the platens by its entrainment, whereas the

pressure on the other side of the platens is higher. This

causes the platens to swing toward each other and oscillate

at their natural frequency.

Recently conducted sootblowing trials [80] showed

that platens swing with a very low frequency, around 0.2

Hz during sootblowing, but that the amplitude of their

swing is large. Quantitative information about the

amplitude was not obtained, but the swinging was found to

affect the jet force. However, the sootblower was fixed in

those tests, due to which the swinging of platens could affect the sootblower jet. During

operation, a sootblower translates and rotates with adjustable speeds. The effect of platen

swinging on the jet then depends on the speed of the sootblower relative to the swinging speed of

the platen. As a result, platen swinging is not expected to affect the jet frequently. Moreover, the

effect will be the greatest only on the sootblower jet at the lowest elevation, because the

amplitude of swinging is directly proportional to the length of the platen tubes. Sootblower jets at

higher elevations will experience weaker effects, if any.

Figure 1.7. Arrangement of

superheater platens in a

recovery boiler [2].

Tubes in the generating bank section are arranged in an array with much smaller tube

spacing compared to the spacing between superheater platens. These tubes run between the steam

drum (upper drum of the generating bank containing a mix of steam and water) and the mud drum

(lower drum containing water). Due to heat transfer requirements and lower carryover that could

cause fouling, the tubes are closely spaced; the tube spacing between two generating bank tubes is

typically only 5 cm (2”), comparable to the sootblower jet size (nozzle exit diameter is usually

slightly greater than 2.5 cm (1”)).

10

The tubes in the economizer section are arranged in a manner similar to that in the

generating bank. However, due to the lower temperature in this section, modern generating bank

and economizer tubes have fins to increase the heat transfer area. These fins may alter the jet

impingement flow field strongly. As a result, the interaction between a sootblower jet and

generating bank and economizer tubes is expected to be stronger than that with superheater tubes.

Understanding the interaction between a supersonic jet and tubes is the first important

and necessary step in developing improved sootblowing strategies. ‘Seeing’ this interaction will

yield valuable information about the flow field during sootblowing. The only way to visualize

such flows, is by taking advantage of flow characteristics such as shock and expansion waves,

which create density gradients, and hence, refractive index gradients in the jet fluid. Such

refractive index gradients can be captured by special optical techniques such as the schlieren

technique, and thus the jet and its interaction with tubes can be made visible.

1.3 Thesis Objectives

The main objectives of this thesis are:

(1) To visualize and document the interaction of a supersonic jet with a single tube, and

determine the effects of the governing parameters on this interaction. This is important

because the sootblower jet always interacts with the first tube of a given row of tubes during

operation, and this interaction determines the subsequent flow of the jet.

(2) To visualize and document the interaction of a supersonic jet with models of the typical tube

arrangements in the three convective sections of a recovery boiler, and to quantify the effects

of this interaction on jet strength by measuring the jet peak impact pressure between tubes.

These three sections are the main regions of the boiler where fouling occurs. This

information is useful in determining the sootblowing effectiveness in these sections.

11

Just as understanding the flow around a tube is the first important step before deposition

on a tube can be understood, understanding the flow of a supersonic jet between and its

interaction with tube bundles is the first necessary step before its interaction with deposits can be

understood. Therefore this thesis investigated the interaction of a supersonic jet with clean round

tubes instead of the interaction of a particle-laden jet with fouled tubes having an irregular cross-

section. Information obtained from the former study will provide a strong foundation for the latter

study.

The objectives described above were achieved through lab-scale experimentation. As will

be evident later in this thesis, this work has helped evaluate the effectiveness of current

sootblowing practices, and has shed light on how the design of the superheater, generating bank,

and economizer sections of a boiler affects sootblowing effectiveness in those sections. It has also

provided potentially useful data for improving boiler tube arrangements with the purpose of

increasing sootblowing effectiveness and reducing sootblower-assisted boiler tube erosion and

corrosion.

CHAPTER 2

LITERATURE SURVEY

This chapter reviews previous work related to sootblowing optimization, as well as free and

impinging supersonic jets. As sootblower jets are supersonic, basic supersonic flow theory,

including nozzle flow, is central to understanding these jets and their interaction with tubes, so

the relevant portions of this theory are also presented. Finally, conclusions are drawn from the

literature survey.

2.1 Fouling and Sootblowing in Kraft Recovery Boilers

The effectiveness of sootblowing depends upon many factors: those that characterize the

sootblower, such as steam flow rate, supply pressure, and nozzle design, and those that

characterize the deposits, such as size and strength. Understanding the effects of these parameters

on deposit removal effectiveness is important in order to devise effective sootblowing strategies.

Most research on fouling in various boilers (for e.g. [9, 22, 33, 72, 79, 108]) has focused on

measuring and modeling deposit formation and growth, and studying deposit characteristics.

Sootblowing optimization, and more specifically sootblower jet dynamics and jet-tube/deposit

12

13

interaction have received far less attention. Most of the research related to these topics has been

performed at the University of Toronto, and has focused on deposit characterization, as well as

sootblower jet dynamics and deposit removal mechanisms. The main results of this research are

summarized next.

2.1.1 Deposits in recovery boilers

Recovery boiler deposits form due to the deposition of carryover and fume on the boiler tubes

[107]. Carryover deposits tend to be hard, and accumulate mainly on the first few tubes of the

superheater platens (Figure 2.1). Figure 2.2 shows massive deposit buildup between superheater

platens. Fume deposits, on the other hand, form due to the condensation of vapours of Na and K

compounds in the flue gas, and are usually powdery and soft. They form as thin coatings on tube

surfaces. Deposits are a mixture of carryover and fume in proportions that vary with location in

the boiler [107].

Figure 2.1. Asymmetric deposits formed

on superheater platens by carryover

impaction [107].

Figure 2.2. Massive deposit buildup between

superheater platens.

passage between sets of platens (typically 0.5 m); also the sootblower lane

deposit accumulation

passage between sets of platens (typically 0.5 m); also the sootblower lane

deposit accumulation

platens (hidden under deposits)

boiler tubes

carryover

deposit

flue gas

boiler tubes

carryover

deposit

flue gas

platens (hidden under deposits)

14

The melting behaviour of deposits plays a vital role in boiler fouling [7, 106, 110]. Being

chemical mixtures, deposits have several characteristic temperatures. The “sticky temperature” is

defined as the temperature at which the deposit contains 15-20% liquid phase, and becomes

sticky to the tube. The “radical deformation” temperature is the temperature at which the material

contains about 70% liquid phase, with enough fluid that it can run off due to its own weight.

Between these two temperatures, deposits are sticky and massive deposit accumulation can occur.

These temperatures are strong functions of the deposit chloride and potassium contents. For a

given level of potassium, increasing the chloride content from 0 to 10 mol% Cl/(Na+K) can

decrease the sticky temperature by 280°C. The radical deformation temperature also decreases

with increase in the chloride content, but less drastically. For a given level of chloride, increasing

the potassium content decreases the sticky temperature but the effect is much less pronounced

[107].

The mechanical behaviour of deposits is closely linked to their thermal behaviour. The

tensile and adhesion strengths of deposits depend strongly on the flue gas and tube temperatures

respectively, in addition to many other parameters [35]. At low temperatures (less than 300°C),

both these strengths are very low. As the flue gas and tube temperatures increase, these strengths

increase due to sintering and reach a maximum. With further increases in temperature, the

strengths decrease due to the formation of a liquid phase in the deposit. The deposit adhesion

strength is generally lower than the tensile strength. Piroozmand [65] showed that deposit tensile

strength increases exponentially with density.

Deposits in most sections of the boiler are brittle. Generally, deposits such as those in the

superheater section are brittle at temperatures lower than 500C and they completely melt above

about 820C [107]; deposits are generally big and hard as they form due to inertial impaction of

big carryover particles. Deposits in the generating bank section are also brittle, but weaker than

those in the superheater. Deposits in the economizer section are mainly fume deposits, so they are

15

soft and powdery. They are much thinner and weaker than those in the other sections.

The chemical composition of deposits also has a great effect on fouling. To date,

experimental studies have been carried out to investigate the effects of deposit chemistry on the

amount and rate of carryover deposition, and also on the removability of carryover deposits [12,

30, 52, 78, 86, 87]. The adhesion efficiency of deposit particles is primarily a function of their

chloride content, temperature, and size. Deposition occurs only when the chloride content

increases beyond a critical value; this critical value decreases with increasing temperature, and

increases with increasing particle size. The amount of liquid phase in carryover deposits is also a

strong function of the chloride content and temperature; deposit liquid phase content must exceed

18-20% for strong adhesion to occur. The jet peak impact pressure (PIP) required to remove a

deposit increases with the deposit chloride content and tube temperature.

2.1.2 Deposit removal by sootblowing and other techniques

Deposits may be removed from boiler tubes in at least four different ways – (i) brittle breakup

due to internal stresses created by jet impingement, (ii) debonding due to jet impingement, (iii)

thermal shock, and (iv) tube vibration or bending. Of these, brittle breakup and debonding are

most frequently encountered. As it is difficult to study these mechanisms in an operational boiler,

mainly theoretical and laboratory scale experimental studies have been performed to date.

A body immersed in fluid flow is subjected to pressure acting normal to its surface, and

shear stresses acting tangential to its surface. When a sootblower jet impinges on a deposit, it

exerts these forces on the deposit, creating mechanical stresses inside the deposit as well as at the

interface between the deposit and the tube. If the stresses created inside the deposit exceed the

tensile strength, it breaks up into pieces. If the interfacial stresses exceed the deposit bond

strength, the deposit debonds from the tube. Usually, both breakup and debonding occur together

during a deposit removal process.

16

Brittle breakup due to internal stresses. In relation to sootblowing in recovery boilers,

Kaliazine et al. [37] were the first to investigate the effects of different operating parameters on

the breakup of deposits impinged by a supersonic jet. They conducted deposit breakup

experiments using model deposits made from gypsum, and a supersonic air jet. First, they

observed that breakup of brittle deposits occurs rapidly, within a few milliseconds. Noting that

the jet-to-deposit exposure time in recovery boilers is on the order of 100 ms, their finding

implies that where deposits are thin and brittle, there is room for reducing sootblowing steam

consumption by appropriately increasing the lance speed. Second, based on a theoretical analysis,

they proposed the following criterion for brittle breakup:

tPIP 2 … (2.1)

where PIP is the value required for breakup and t is the deposit material tensile strength. This

criterion was found to agree reasonably well with their experimental results.

Recently, Eslamian et al. [23, 24] and Pophali [69] performed experiments similar to

those of Kaliazine et al. [37], but studied the brittle breakup mechanism in detail via high-speed

photography. Their work identified three different deposit breakup mechanisms, which correlated

with the jet-to-deposit diameter ratio (see Table 2.1). Crack formation was found to be vital for

Table 2.1: Brittle deposit breakup mechanisms [68].

Observed Breakup

Mechanism

Jet-to-deposit

Diameter Ratio Breakup Image

Axial crack formation dj/ddeposit > 0.51

Surface erosion + axial

crack formation 0.36 < dj/ddeposit < 0.51

Surface erosion + spalling dj/ddeposit ≤ 0.36

17

fast deposit breakup. Cracks form easily and quickly in thin deposits, whereas in thick deposits,

crack formation can only occur if the jet has drilled into the deposit.

Debonding. This mechanism is particularly important for very hard and thick deposits such as

those in the superheater region of the boiler. If a strong deposit is weakly attached to the tubes, a

strong sootblower jet may not be able to break it, but may remove it by debonding. Debonding

occurs when a jet exerts a moment on the deposit about its interface with the tube, creating

interfacial stresses.

Sabet [19] studied deposit removal by debonding. He measured the mean and fluctuating

drag and lift forces exerted by a jet on different types of deposits, and found that the flow-induced

vibrations caused by the lift force fluctuations may be the dominant lateral forces responsible for

debonding deposits.

Kaliazine et al. [35] developed a theoretical model of the deposit mechanical structure,

and used it to estimate the stress distribution in the adhesion layer between the deposit and the

tube. Their criterion for debonding is:

W

dPIP c

adh max

1 … (2.2)

where PIP is the value required for debonding a deposit of adhesion strength adh, dc is the deposit

characteristic length, W its thickness, and max a dimensionless ‘shape function’ of order unity,

which characterizes the deposit shape. Equation (2.2) shows that the jet PIP required for

debonding decreases with deposit thickness. Thus, thin deposits are more likely to break by brittle

breakup, whereas thick deposits by debonding.

Mao et al. [53] studied the effects of deposit thickness and jet impact angle on deposit

removal in conditions more representative of boiler conditions, using an Entrained Flow Reactor

(EFR) and an air jet apparatus. Synthetic deposits were prepared by mixing different chemicals

and burning them in the EFR. Deposits of varying thicknesses were impinged by a jet in two

18

ways: head-on (Figure 2.3a), simulating brittle breakup, and at 90 relative to the head-on case

(Figure 2.3b), simulating debonding. The jet PIP required for deposit removal increases with

deposit thickness for the head-on case, whereas it decreases when the deposit is rotated 90°.

These results support the theory presented above (equation 2.2). This indicates that debonding is a

more efficient way of removing thick, asymmetric deposits. This then raises the question of

whether inclined nozzles could be used to more effectively clean recovery boiler superheater

platens.

0.00

0.05

0.10

0.15

0.20

0 2 4 6 8 10 12 14

Deposit thickness h (mm)

PIP

(M

Pa)

Figure 2.3. Jet PIP required for deposit removal versus deposit thickness; (a) head-on

impingement; (b) impingement at 90 relative to the head-on case [53].

Inclined sootblower nozzles. Current designs of recovery boiler steam sootblowers have two

opposing nozzles aligned perpendicular to the axis of the lance tube. Sootblowers with nozzles

inclined relative to the lance tube are also available, but used mainly to clean the furnace wall

[115], and are not generally used for on-line cleaning of boiler tubes. A general consensus is that

such nozzles decrease the jet penetration in between the superheater platens, and hence, reduce

the cleaning radius of the sootblower. However, it was mentioned in [8] that such nozzles

improve the penetration of the jet into the tubes behind the first tube of a platen. Massive deposit

accumulation takes place mainly on these tubes. Jets from inclined nozzles might also exert

5%Cl, 5%K

5%Cl

jet

0.00

0.05

0.10

0.15

0.20

0 2 4 6 8 10 12 14

Deposit thickness h (mm)

PIP

(M

Pa

)

5%Cl, 5%K

5%Cl

jet

a b

0.00

0.05

0.10

0.15

0.20

0 2 4 6 8 10 12 14

Deposit thickness h (mm)

PIP

(M

Pa)

0.00

0.05

0.10

0.15

0.20

0 2 4 6 8 10 12 14

Deposit thickness h (mm)

PIP

(M

Pa

)

5%Cl, 5%K 5%Cl, 5%K

5%Cl 5%Cl

jet

jet

a b

19

greater debonding force on these massive deposits. However, they may also increase platen

swinging to their inclined impact on the platens [80]. Sootblowers with such inclined nozzles

have been used successfully in coal-fired utility boilers (lead-lag nozzles), but in these boilers, the

inter-platen spacing is larger than in recovery boilers [102]. Only very recently have these nozzles

been introduced in recovery boilers [101], and their performance is currently being evaluated.

Their feasibility remains to be determined.

Thermal shock. During operation, deposits are subjected to two different types of thermal shocks

- one arising from a change in fuel firing, and the other arising from the impingement of the

sootblower jet, which is relatively cooler than the deposit surface.

Deposits in the superheater section typically have a linear thermal expansion coefficient

of about 46 m/mC, which is 3 - 4 times higher than that of the carbon steel used for making

boiler tubes [107]. So, during a thermal shock or 'chill-and-blow' event, the black liquor flow is

reduced or turned off in order to rapidly cool the deposits. This causes the deposits to contract

faster than steel, and consequently crack and detach from the tubes. Then, they either fall due to

their weight or can be easily blown off by sootblowers. A thermal shock event in a recovery

boiler typically requires about 8 hours to complete [107]. Laboratory experiments carried out at

the University of Toronto have shown that thermally shocked deposits are much easier to remove

using a jet than are deposits not subjected to a thermal shock [38].

As sootblower jets are much cooler than deposits inside a boiler, the deposits undergo

momentary thermal shocks every time the jet impinges on them. However, these thermal shocks

are much weaker compared to those described above. Kaliazine et al. [38] have performed

laboratory tests and theoretical heat transfer calculations which have shown that such thermal

shocks do not contribute to deposit removal.

20

Figure 2.4. Deposit thermal shock caused by supersonic jet impingement [38].

Figure 2.4 shows their calculated variation of the deposit-tube wall interface temperature

with time, caused by sootblower jet impingement. The theoretical model used for this

computation was validated by the researchers using laboratory experimental data. Since the

chemical composition of a typical deposit is similar to that of smelt, the typical thermal

conductivity of smelt, around 1 W/mK was used in the theoretical calculations [38]; however, a

much higher value was also used for comparison. The figure clearly shows that the drop in the

deposit temperature for a deposit conductivity of 1 W/mK is negligible, and that the drop for 12

W/mK is also very small, and practically cannot contribute to deposit removal.

Tube vibration or bending. Industrial experience has shown that flue gas flow and sootblower

jet impingement cause boiler tubes to swing, vibrate, and bend. Bending creates stresses in the

deposits accumulated on these tubes. These stresses continuously change with time from tensile

to compressive and vice-versa depending on the direction of bending, and fatigue the deposits.

The deposits eventually crack, and become easier to remove during sootblowing.

21

Sabet [19] and Kaliazine et al. [37] performed artificial deposit blow-off experiments, in

which they found that the lift force fluctuations cover a large frequency range from zero to the so-

called Strouhal frequency. They theoretically and experimentally showed that if the affected

structure has a natural frequency within this range, the structure interacts with the jet in a resonant

way, drastically increasing the effect of the fluctuating force. The increase in magnitude is

inversely proportional to the square root of a vibration decay coefficient. This coefficient is that

part of the system mechanical energy that dissipates during one cycle of vibration. Kaliazine et al.

[37] also obtained a criterion for vibrational deposit removal during sootblowing.

2.1.3 Sootblower jet dynamics

There have been very few studies on sootblower jet dynamics and jet/tube interaction to date.

Before those studies are reviewed however, it is important to understand why sootblowers used in

kraft recovery boilers are supersonic and not simply subsonic. Jets are typically generated using

convergent nozzles. For any convergent nozzle, the jet supply pressure necessary to generate a

fully-expanded sonic jet (most efficient sonic jet) is given by the following relation derived using

isentropic flow theory –

1

1

2

ao pp … (2.3)

where po is the required jet supply pressure, pa is the ambient pressure, and is the specific heat

ratio of the gas. For air, po = 1.9pa = 192 kPa (abs) = 27.8 psia. If the jet supply pressure is greater

than this value for pa = 101.325 kPa (abs) = 14.7 psia, the subsonic jet automatically converts into

a less efficient supersonic jet. Special convergent-divergent nozzles must be used to obtain the

most efficient supersonic jet, and the supply pressure of these nozzles is relatively much larger.

That is why the supply pressure of subsonic jets is much lower than that of supersonic ones.

Emami [21] used a semi-empirical model to calculate the peak impact pressure along the

22

centerline of subsonic and supersonic jets with the same mass flow. Figure 2.5 shows the results.

The centerline PIP in the core region of a supersonic jet is much higher than that of a subsonic jet

because of the large difference in the jet supply pressures. Both the subsonic and supersonic jets

have approximately the same core length. However, higher PIP is desired in recovery boilers

because large deposits usually accumulate on the leading tubes of superheater platens and may

sinter and become hard with time. A high PIP is needed in order to break and remove them. These

deposits are at a distance shorter than the core of the supersonic jet, and hence lie within the jet

core. Moreover, supersonic jets spread much more slowly than subsonic jets and therefore decay

much slower. As a result, they also penetrate farther between rows of tubes than subsonic jets.

Due to these reasons, supersonic jets are utilized in recovery boilers instead of subsonic jets.

Figure 2.5. Computed peak impact pressure along the centerline of subsonic and

supersonic jets with the same mass flow [21].

The study by Jameel et al. [32] is one of the first ones, in which the authors developed a

mathematical model using Kleinstein’s theory on mixing in turbulent jets [44] and Witze’s

experimental correlations [118] to predict the axial variation of sootblower jet PIP. They used this

23

model to compare conventional nozzles to full expansion nozzles, and concluded that full

expansion nozzles are more efficient than conventional ones. The use of full expansion nozzles

increases the jet energy available for deposit removal, and their cleaning area is much larger than

that achieved by the conventional ones. The findings of this study had a major impact on the pulp

and paper industry due to which about 90% of the recovery boilers world-wide adopted

sootblowers with full expansion nozzles [109].

Kaliazine et al. [36] also investigated the feasibility of using low pressure steam

exhausting from the turbine, for sootblowing instead of the more valuable high pressure steam

generated by the boiler, normally used for sootblowing. The low pressure steam exiting the

turbine has a lower monetary value than the high pressure steam generated by the boiler, and so it

may be possible to achieve substantial monetary savings by switching to low pressure

sootblowing. However, due to the lower pressure, the steam flow rate must be increased above

that required in high pressure sootblowing to obtain a comparable deposit removal capability. As

a result, the feasibility of this technology depends mainly on the differential cost between high

pressure and low pressure steam and the amount of additional steam required to compensate for

the lower pressure [111].

Tandra [99] developed a modified k- turbulence model to numerically simulate a

sootblower jet and its interaction with recovery boiler superheater platens. He also used this

model to investigate the feasibility of low pressure sootblowing [97]. He showed that by using a

larger nozzle and a slightly greater steam flow rate for low pressure sootblowing, it is possible to

exert a drag force on a deposit which is comparable to that exerted by a high pressure sootblower

jet, and thus attain comparable deposit removal capability.

Fouling monitoring and location for targeted sootblowing. Sootblowing timing and strategy

are as important for deposit removal as the jet strength and dynamics. A boiler location which is

greatly fouled must be subjected to sootblowing for a longer duration of time than a location

24

which is relatively cleaner. However, the hostile conditions inside a boiler make it almost

impossible to determine which location is fouled. Presently, mill operators use infrared cameras

for this purpose, but even those cannot provide this information in most cases. As a result, current

research efforts in the industry are directed at finding ways to locate the fouled region in the

boiler.

Very recently, Adams [3] developed a method to determine the local deposition rate in

the vicinity of each sootblower in a fleet of sootblowers cleaning a boiler. Using boiler operating

data, he showed that the rate of deposit removal is a good representative of the rate of deposition

or fouling. In this method, the change in the generating bank outlet temperature with time is

continuously monitored during a stroke of a given sootblower, and reflects the rate of fouling in

the vicinity of that sootblower as well as in the entire superheater and generating bank sections.

An increase in the outlet temperature indicates increased fouling. By applying appropriate

corrections, the rate of fouling in the vicinity of that particular sootblower is isolated, thus

identifying the locations of the boiler which are prone to fouling and plugging.

Another method was presented by Tandra et al. [100], in which the authors used changes

in heat flow from the combustion gases to the water/steam in the heat transfer tubes to identify

when a certain section of the boiler was fouled. In this method, mass and energy balances are

performed on the different sections of a boiler and on the entire boiler, to determine the heat

transfer efficiency of a given section. This efficiency is monitored continuously, and when the

efficiency drops to a pre-determined low value due to fouling, sootblowing is initiated to restore

the efficiency. In this way, only that much steam is consumed for sootblowing as much is truly

needed.

As sootblower jets are supersonic, a review of basic compressible flow theory is

necessary in order to understand their interaction with a tube and tube arrangements.

25

2.2 Basic Compressible Flow Theory

In compressible flow, moderate to strong changes in pressure give rise to substantial changes in

density. The Mach number Ma, is an indicator of flow compressibility, and is defined as the ratio

of the flow velocity u, to the speed of sound a, in the fluid: Ma = u/a. Generally, flow with Ma ≥

0.3 is treated as compressible [5]; flow at Ma < 1 is subsonic, and that at Ma > 1 is supersonic.

Due to the high velocity in supersonic flow, shock and expansion waves form to adjust to abrupt

disturbances. Supersonic flow is conveniently described using one-dimensional isentropic flow

theory.

2.2.1 One-dimensional isentropic flow relations

A flow is isentropic if it is both adiabatic and reversible (i.e. there is no heat transfer with

surroundings and no increase in entropy). The following relation holds for an isentropic process:

1

1

2

1

2

1

2

T

T

p

p … (2.4)

where p, , and T are the static pressure, density, and static temperature respectively, is the ratio

of specific heats of the gas, and subscripts 1 and 2 refer to the initial and final states in the

process. Applying conservation of energy to one-dimensional isentropic flow and considering a

calorically perfect gas, the following relation is obtained:

2

2

11 x

x

o MaT

T

… (2.5)

where To is the total or stagnation temperature of the flow and x is the spatial coordinate. This

relation enables us to calculate the ratio of total to static temperature at any point in the flow as a

function of the Mach number and . From equations (2.4) and (2.5), we obtain a similar relation

for the ratio of total to static pressure:

26

12

2

11

x

x

o Map

p … (2.6)

where po is the total or stagnation pressure of the flow. Similar relations can be obtained for other

thermodynamic quantities, for isentropic flow.

2.2.2 Shock and expansion waves

A flow scenario that occurs frequently during sootblowing is the formation of a normal shock

wave. When a supersonic jet impinges on a tube or deposit (or in general, when fluid flows past a

blunt body at supersonic speed), a normal shock wave forms just upstream of the tube or deposit.

This shock wave is oriented perpendicular to the flow direction, and creates a sudden change in

properties (Figure 2.6a). Applying conservation of mass, momentum, and energy across a normal

shock wave yields the following useful relations:

2/)1(

]2/)1[(12

1

212

2

Ma

MaMa … (2.7)

1

1

21

1

21

21

1

2

)1(2

1

)1(2

)1(

MaMa

Ma

p

p

o

o … (2.8)

Subscripts 1 and 2 refer to conditions upstream and downstream of the shock wave,

respectively. Equation (2.7) shows that supersonic flow always decelerates to subsonic speeds

across a normal shock wave. From equation (2.8), note that for a given gas (), po2 only depends

on po1 and Ma1, and po2 decreases strongly with Ma1. This equation enables us to calculate the PIP

exerted by a sootblower jet on a deposit, if the upstream Mach number and total pressure are

known, because the PIP exerted on the deposit is the total pressure of the jet after a normal shock

wave.

27

Figure 2.6. (a) Normal shock wave; (b) oblique shock wave; (c) expansion wave.

Two other important flow structures that form in supersonic flow fields are oblique

shocks and expansion waves. Oblique shock waves usually form when supersonic flow is ‘turned

into itself’, that is, when it is obstructed (Figure 2.6b); expansion waves form when supersonic

flow is ‘turned away from itself’ (Figure 2.6c). Analysis of an oblique shock wave applying

conservation of mass, momentum, and energy yields the following relation:

2)2cos(

1sincot2tan

21

221

Ma

Ma … (2.9)

is the flow deflection angle and the oblique shock angle. This relation is known as the --Ma

relation, and can be used to calculate either explicitly or implicitly, any one of , or Ma if the

other two quantities are known. Property changes across an oblique shock wave can be calculated

using the normal shock relations and the normal component of the upstream Mach number, Man1.

Expansion in supersonic flow normally occurs across a fan centered at a point, called the

Prandtl-Meyer expansion wave. Analysis of an expansion wave allows calculation of the wave

28

angle, as:

)()( 12 MaMa … (2.10)

where 1tan)1(1

1tan

1

1)( 2121

MaMaMa

… (2.11)

is the Prandtl-Meyer function. Subscripts 1 and 2 again denote the conditions upstream and

downstream of the expansion wave. Expansion through such a wave is isentropic.

From the point of view of sootblowing, it is necessary to determine how the pitot pressure

changes across oblique shocks and expansion waves, because these waves form during jet/tube

interaction. Using the basic one-dimensional flow relations, Graham and Davis [28] calculated

the following relations for the ratio of pitot pressure across an oblique shock wave:

For Ma2 > 1:

1

)1(sin2

)1(2

)1(2 221

1

1

22

21

1

21

22

1

2

Ma

Ma

Ma

Ma

Ma

p

p

pit

pit … (2.12)

For Ma2 < 1:

1

1

221

21

1

221

21

2

1

2

)1(sin2

)1(2

2sin)1(

]2)1[(sin

Ma

Ma

Ma

Ma

p

p

pit

pit … (2.13)

The relation between Ma1 and Ma2 is provided in [28] and not presented here. Figure 2.7

(reproduced from [28]) shows the variation of the ratio of pitot pressures behind and ahead of an

oblique shock as a function of the shock angle, for different upstream Mach numbers. Quantities

in the graph have been denoted by symbols used in the present work, and are different than those

used in the original paper. The figure clearly shows that the pitot pressure increases across an

oblique shock, whether weak (M2 > 1) or strong (M2 < 1). On the other hand, the flow across an

expansion wave accelerates, increasing the Mach number, while the total pressure remains

29

0

1

2

3

4

5

6

0 10 20 30 40 50 60 70 80 90

Oblique shock angle, [deg]

Pit

ot p

res

sure

beh

ind

sho

ck

Pit

ot

pre

ssu

re a

hea

d o

f sho

ck

Ma2 > 1(weak shocks)

oblique shock

Figure 2.7. Ratio of pitot pressure across an oblique shock wave as a function of shock

angle and upstream Mach number [28].

constant. As a result, the normal shock wave at the orifice of a pitot tube is stronger downstream

of the wave than upstream, and the pitot pressure decreases across an expansion wave.

Wave reflection. Another important phenomenon in supersonic flows is that of wave reflections

from solid and free boundaries. When an oblique shock is incident on a solid wall (Figure 2.8a),

the subsequent flow depends on the boundary condition at the wall – i.e. that the flow

immediately adjacent to the wall must be parallel to it [5]. This is satisfied by the formation of a

reflected oblique shock wave. Thus, oblique shock waves and similarly expansion waves reflect

as shock and expansion waves respectively from a solid wall. The reverse occurs in the case of a

pp

it2

pp

it1

= 1.4

10

6

3

Ma1 =

4

2

1.4

Ma2 < 1(strong shocks)

1

2

0

1

2

3

4

5

6

0 10 20 30 40 50 60 70 80 90

Oblique shock angle, [deg]

Pit

ot p

res

sure

beh

ind

sho

ck

Pit

ot

pre

ssu

re a

hea

d o

f sho

ck

Ma2 > 1(weak shocks)

oblique shock

pp

it2

pp

it1

Pit

ot p

res

sure

beh

ind

sho

ck

Pit

ot

pre

ssu

re a

hea

d o

f sho

ck

Pit

ot p

res

sure

beh

ind

sho

ck

Pit

ot

pre

ssu

re a

hea

d o

f sho

ck

pp

it2

pp

it1

= 1.4

10

6

3

Ma1 =

4

2

1.4

oblique shock

2

Ma2 < 1(strong shocks)

1

2

1

30

Figure 2.8. Reflection of an oblique shock wave from (a) a solid wall, and (b) a constant

pressure free boundary.

free boundary. When an oblique shock wave is incident on a free boundary at constant pressure

(Figure 2.8b), the shock wave reflects as an expansion wave to maintain the pressure along that

boundary. Similarly, an expansion wave reflects as an oblique shock. Due to the formation of

dense and rare regions in supersonic flow because of shock and expansion waves, often slip lines

form between these regions across which momentum transfer takes place.

2.2.3 Supersonic nozzle flow

Applying the differential form of the equation for conservation of mass to an isentropic flow and

manipulating, we obtain the following ‘area-velocity relation’:

u

duMa

A

dA12 … (2.14)

where A is the cross-sectional area normal to the flow and u is the flow velocity. This relation

wall

incident shock

Flow remains parallel to wall via a reflected obliq ue shock wave

reflected shock

constant pressure free boundary

incident shock

reflected expansion wave

To attain free stream pressure, flow turns outwards and expands via a reflected expansion wave

(a)

(b)

wall

incident shock

Flow remains parallel to wall via a reflected obliq ue shock wave

reflected shock

wall

incident shock

Flow remains parallel to wall via a reflected obliq ue shock wave

reflected shock

constant pressure free boundary

incident shock

reflected expansion wave

To attain free stream pressure, flow turns outwards and expands via a reflected expansion wave

constant pressure free boundary

incident shock

reflected expansion wave

To attain free stream pressure, flow turns outwards and expands via a reflected expansion wave

(a)

(b)

31

shows that for Ma < 1 (subsonic flow), the velocity increases if the cross-sectional area decreases.

Surprisingly, for Ma > 1 (supersonic flow), velocity increases only if the cross-sectional area

increases. Consequently, to accelerate a flow to supersonic speeds, the flow must be forced

through a convergent-divergent nozzle. As the initially subsonic flow passes through the

convergent section, it accelerates and reaches Ma = 1 at the nozzle throat (the minimum cross-

sectional area), and then further accelerates to supersonic speeds in the divergent part. Any nozzle

is characterized by only one nozzle exit Mach number Mae, because the Mach number at any

position x along the nozzle centerline Max, depends only on the ratio of the nozzle cross-sectional

area at that position Ax, to the cross-sectional area of the throat At, and the specific heat ratio :

)1(

)1(2

1

2

)1(2

1

)1(2

11

1

x

xt

xMa

MaA

A … (2.15)

For a given nozzle, Mae can be calculated using this relation and the ratio of the exit area

to the throat area Ae/At. Knowing the pressure of the stagnant surroundings (e.g. standard

atmospheric pressure) and Mae, and assuming isentropic flow through the nozzle, the pressure at

the nozzle inlet, i.e. the supply pressure po, can be calculated using equation (2.6). Vice-versa,

knowing the supply pressure po and Ae/At of the given nozzle, the pressure at the nozzle exit pe,

can be calculated.

The static pressure at the nozzle exit pe, relative to the ambient pressure plays a very

important role in controlling the structure of the supersonic jet exiting the nozzle. Depending on

the exit pressure ratio PR = pe/pa, three different types of supersonic jets can form. At PR = 1, the

exit pressure is exactly equal to the ambient pressure, the jet has expanded correctly, and so is

termed a ‘fully expanded’ or ‘correctly expanded’ jet (Figure 2.9a). Such a jet is free of shock

waves and is the most efficient. For PR < 1, the exit pressure is lower than the ambient pressure,

the jet has expanded more than needed, and so is termed an ‘overexpanded’ jet. The jet fluid

32

Figure 2.9. Supersonic jets; (a) fully expanded; (b) overexpanded; (c) underexpanded.

undergoes compression through oblique shocks just outside the nozzle to increase its pressure to

the ambient value (Figure 2.9b). For PR > 1, the exit pressure is greater than the ambient

pressure, the jet has not completed full expansion, and so is termed an ‘underexpanded’ jet. The

jet fluid expands through a Prandtl-Meyer expansion wave at the nozzle exit that lowers its

pressure to the ambient value (Figure 2.9c). Over and underexpanded jets are also termed

‘imperfectly expanded’ or ‘off-design’ jets. The structure of supersonic jets will be further

described in section 2.4.

Although the assumption of isentropic flow is used to describe supersonic nozzle flow

and provides a good approximation to the actual flow, it should be noted that real flow inside any

nozzle is not fully isentropic. Due to unavoidable friction losses and heat transfer, the extent of

the deviation from being isentropic depends on the design of the nozzle – the smoother the

nozzle, the smoother the flow through it, and the closer the flow is to being isentropic.

Another useful equation is that yielding the mass flow rate ṁ, through a choked nozzle:

1

1

1

2

gaso

to

RT

Apm … (2.16)

Rgas is the specific gas constant, and is obtained from the universal gas constant RU as RU/MW,

where MW is the gas molecular weight. RU = 8.314 kJ/kg-K. This relation shows that the mass

flow rate increases with the supply pressure and throat cross-sectional area, or with the square of

the throat diameter, and decreases with the square root of the stagnation temperature.

(a)

nozzle

jet shear layer

(b)

expansion waves

shock wave

(c)

expansion waves

shock wave

(a)

nozzle

jet shear layer

(a)

nozzle

jet shear layer

(b)

expansion waves

shock wave

(b)

expansion waves

shock wave

(c)

expansion wavesexpansion waves

shock wave

(c)

shock wave

33

2.3 The Schlieren Technique

Supersonic gas jets cannot be seen by the naked eye or with regular photographic equipment, but

special techniques are available to visualize them. The schlieren technique is one such method

which has been extensively used for a long time, yet continues to be used today to visualize and

study supersonic flows. The technique was invented by German physicist August Toepler

between 1859 and 1864, who used it to visualize and study flames, convection, shock waves, and

other phenomena invisible until then [45, 84]. This technique was used extensively in the present

thesis, so descriptions of the principles and apparatus are presented here.

The operating principle of the schlieren technique is that parallel light rays refract as they

pass through optical inhomogeneities such as density gradients in a transparent medium.

Supersonic gas jets exhibit such density gradients. For a gas, the local value of the refractive

index1 n, is directly proportional to the local value of the density as per the Gladstone-Dale

relationship [84]:

Kn 1 … (2.17)

where K is the Gladstone-Dale constant and depends on the gas characteristics as well as the

frequency of the light used. Density gradients in the medium (or in a supersonic jet for example)

give rise to refractive index gradients in the medium. When light passes through such an optically

inhomogeneous medium, the refractive index gradients cause the light rays to bend in different

directions, and this refraction can be captured as a schlieren image.

To better explain this technique, consider the situation shown in Figure 2.10. This figure

shows light rays traveling through a medium with a negative refractive index gradient in the

vertical direction (dn/dy < 0)2, and shows their refraction. Due to the negative gradient, density

increases in the downward direction, reducing the speed of light, c. As a result, the distance

1 Refractive index, n = co/c, where co = speed of light in vacuum (3 108 m/s) and c = speed of light in the surrounding medium. 2 This section on geometrical optics is based on Settles [84].

34

covered by the light rays in a given time

duration decreases in the downward

direction, due to which the light wavefront

changes direction, that is, the light rays

bend. Using the geometric theory of optics,

the following relations between the

curvature of the light rays and the refractive

index gradients can be obtained:

0

0

dy

d

dy

dn

rarer medium denser medium

x

n

nz

x

1

2

2

, y

n

nz

y

1

2

2

… (2.18)

These relations show that the bending of the light rays depends on the refractive index

gradient, and thus from equation (2.17), on the density gradient. These light rays (blocked

appropriately using a knife edge) produce a schlieren image. Thus, the schlieren technique

visualizes the first derivative of density.

Figure 2.11 illustrates a conventional 2-mirror z-type schlieren system. A light source LS,

is placed at the focus of a concave mirror M1. Light from LS passes through a pin-hole PH, and a

conical, diverging set of light rays proceeds to M1. The light rays are collimated by M1 and sent

through the test section where some of them encounter the schlieren object with density gradients

SO (for example, within the supersonic jet). These gradients refract the light rays causing them to

bend in different directions. These refracted rays, along with other parallel rays reflect off a

second concave mirror M2. This mirror refocuses the parallel light rays, but the refracted rays are

not focused. A knife edge KE, is placed at this focal plane to control the refracted light rays. The

rays then proceed to a focusing lens FL, which produces a schlieren image of SO on a camera

sensor.

Figure 2.10. Planar refraction of light

passing through a region with a negative

vertical refractive index gradient [84].

z

yx

lower

greater

3

2

1

c3∆t

light rays∆z

c2∆t

c1∆t

1 > 2 > 3c1 < c2 < c3

0

0

dy

d

dy

dn

rarer medium denser medium

z

yx

z

yx

lower

greater

3

2

1

c3∆t

∆zlight rays

c2∆t

c1∆t

1 > 2 > 3c1 < c2 < c3

35

Light source, LPinhole, PH

Parabolicmirror, M1

Parabolicmirror, M2

Region with densitygradients, S

Knife edge, K

Figure 2.11. A conventional 2-mirror z-type schlieren system [84].

Role of the knife edge. The knife edge plays an important role in generating a clear schlieren

image. When the light rays reflect off of the first mirror as parallel rays, and pass through the

region with density gradients, they deflect in different directions at different points within the

region. Due to this, they are not focused by the second mirror; depending on the refractive index

gradient, some rays escape the knife edge, while others are blocked by it. Those which escape,

illuminate a spot on the sensor by means of the focusing lens, and because the knife edge cuts off

some of the rays, some portions of the sensor remain dark. This contrast between the bright and

dark regions produces a schlieren image of the density gradients.

The amount of light that the knife edge cuts off determines the sensitivity of the system -

the more light it cuts off, the better the contrast in the schlieren image. The knife edge cuts off

light rays deflected in a direction perpendicular to it, so it makes visible those density gradients

which are directed normal to its orientation; that is, a horizontal knife edge visualizes vertical

Focusing lens, F

Camera

Light source, LS

Parabolicmirror, M1

Parabolicmirror, M2

Region with densitygradients, SO

Knife edge, KE

Focusing lens, FL

Camera

Light source, LPinhole, PH

Parabolicmirror, M1

Parabolicmirror, M2

Region with densitygradients, S

Knife edge, K

Focusing lens, F

Camera

Light source, LS

Parabolicmirror, M1

Parabolicmirror, M2

Region with densitygradients, SO

Knife edge, KE

Focusing lens, FL

Camera

36

density gradients. The amount of light cut off by the knife edge, that is, the sensitivity, is

restricted by the minimum amount of illumination required to obtain a good image of the overall

flow field.

2.4 Supersonic Free Jets

As described in section 2.2.3, supersonic jets can be either fully-expanded or off-design

(over/underexpanded), depending on the jet static pressure at the nozzle exit. Due to their wide

spread applications in engineering, such as in rocket and aircraft propulsion, the structure of

supersonic jets has been studied in detail over the years.

2.4.1 Jet structure

Fully-expanded jets

The fully expanded jet (PR = 1) has the simplest structure because it is free of any shock or

expansion waves in its core region (although practically, some weak shock waves always exist).

The jet boundary just outside the nozzle is straight and parallel to the direction of jet flow. The

core region of jet extends from the nozzle exit to about 10-15 nozzle diameters downstream, but

the core decreases in diameter with distance from the nozzle, because of turbulent mixing in the

shear layer of the jet. Pressure, temperature, density, total pressure and other variables remain

unchanged in the core. The peak impact pressure and velocity are the highest in the core. Beyond

the core, the jet becomes subsonic and fully developed, with self-similar velocity profiles

characteristic of turbulent jets. In this fully developed region, the velocity and total pressure of

the jet decrease monotonically with distance due to turbulent mixing and radial spreading.

Due to its relatively simple structure, a fully expanded jet is more amenable to theoretical

treatment than its off-design counterparts. The main characteristic of such a jet that must be

determined analytically is the decay of properties along the jet centerline in the subsonic portion

37

of the jet, which is caused by compressible turbulent mixing. Earlier theoretical studies of this

decay by Kleinstein [44], Warren [116], and Witze [118] have been reasonably successful in

obtaining a close agreement between theory and experimental data. Kleinstein [44] was one of the

first to describe a fully-expanded jet by linearizing the basic conservation equations, and by

introducing a modified eddy viscosity expression. He derived the following expression for the

decay of axial velocity along the jet centreline:

7.0

1exp1

2/1e

oxk

u

… (2.19)

in which ou is the axial velocity normalized by the nozzle exit velocity, x is the axial coordinate

normalized by the nozzle exit radius, e is the density normalized by the nozzle exit density and

k is a constant in the expression for the modified eddy viscosity that Kleinstein assumed to be

equal to 0.074. Later, Warren [116] generalized Kleinstein’s model to allow k to depend on the

Mach number, and Witze [118] extended the model further by introducing two expressions for k,

one for subsonic jets (isothermal and heated) and the other for supersonic. For supersonic jets:

15.02 )1(063.0 jMk … (2.20)

where Mj is the nozzle exit Mach number.

Few experimental studies of the structure of fully-expanded jets are available in the

literature. Of these, the earlier studies by Eggers [20] and Lau et al. [50] are important because

they are comprehensive experimental studies of the different aspects of fully-expanded

supersonic jets. Eggers [20] measured the centerline and radial pitot pressure profiles of a Mach

2.22 air jet and converted them into profiles of axial velocity, as well as of eddy viscosity

distributions. He also determined the spreading rate of that jet. Lau et al. [50] measured the

centerline and radial velocity of subsonic and almost fully-expanded supersonic jets using a laser

velocimeter, and developed correlations to fit that data. These correlations can used to predict, for

example, the centerline velocity decay in jets. They also obtained the following relation between

38

the jet core length, xc, and the jet Mach number, Mj:

21.12.4 je

c Md

x … (2.21)

More recent studies involving fully-expanded jets are those of Katanoda et al. [42] and Kweon et

al. [47]. These studies report measurements of the centerline and radial pitot pressure of fully-

expanded and off-design jets, even though the studies focus on the structure of off-design jets.

Although fully expanded jets are the most efficient due to the absence of shock waves,

they are very difficult to generate. Jets are fully-expanded when the nozzle exit static pressure

exactly equals the ambient pressure. This rarely occurs because of unavoidable factors including

supply pressure fluctuations and manufacturing imperfections in the nozzle. Sootblower jets are

no exception, and operate either slightly under or overexpanded (off-design) most of the time.

Off-design jets

For a given nozzle, an underexpanded jet is more powerful than an overexpanded jet, because the

underexpanded jet forms at a higher supply pressure than the overexpanded jet, and so its

diameter is also larger. Underexpanded jets can be tentatively classified into two types depending

on the exit pressure ratio – mildly underexpanded (1 < PR ≤ 2) and strongly underexpanded (PR

> 2) [18, 25]. Underexpanded jets, and specifically strongly underexpanded jets, have been

studied experimentally in great detail (e.g. [4, 6, 16, 18, 25, 31, 41, 42, 47, 51, 57, 60, 114, 119]).

Only the main features of these jets are described here, based on these studies.

The nozzle exit pressure of a mildly underexpanded jet (Figure 2.12a) is only slightly

greater than the ambient pressure, and so the fluid expands as it comes out of the nozzle to match

its pressure to the ambient value (Figure 2.12b). The expansion occurs through a weak Prandtl-

Meyer expansion fan centered at the nozzle lip. Due to this expansion, the jet bulges and grows

radially as it exits the nozzle. The expansion waves reach the opposite side of the jet and reflect

as compression waves back into jet (the jet boundary is a constant pressure boundary). These

39

(a)

(b)

(a)

(b)

Figure 2.12. A mildly underexpanded jet [18]; (a) schematic of jet structure; (b) sonic jet

with PR = 1.59.

(a)

(b)

(a)

(b)

Figure 2.13. A strongly underexpanded jet [18]; (a) schematic of jet structure; (b) sonic

jet with PR = 4.09.

40

waves then merge to form an oblique incident jet shock. The flow is compressed as it passes

through this shock. The shock again reflects from the jet boundary as expansion waves, and the

process repeats itself, forming a diamond-shaped shock cell structure in the core region

(supersonic region of the jet) that decays downstream due to interaction with the shear layer

turbulence. In subsonic and fully-expanded supersonic jets, velocity and other properties are a

maximum at the centerline and decrease towards the jet boundary. On the other hand, at some

locations in an underexpanded jet, these maxima occur not at the centerline but in a ring around

the centerline.

The above behaviour is more pronounced in a strongly underexpanded jet (Figure 2.13a).

The jet exit pressure is now so high that a much stronger expansion wave forms, which in turn

leads to a much stronger normal shock or Mach disk; this creates a barrel-shaped shock cell just

outside the nozzle (Figure 2.13b). As in a mildly underexpanded jet, the same process repeats

itself until the shock cell structure diffuses completely.

Little theoretical work has been done on mildly underexpanded jets. Prandtl [71]

represented these jets in the form of perturbations about a steady parallel fully-expanded jet, with

mean velocity equal to that at the nozzle exit; the jet structure was assumed periodic. The model

breaks down as the underexpansion increases. Pack [58, 59] improvised Prandtl’s solution by

considering the mean velocity in the model as the full-expansion velocity rather than the nozzle

exit velocity. More recently, Tam et al. [94] developed a multiple scales model using the

linearized Navier-Stokes equations to describe the shock cell structure of underexpanded jets.

Their model agreed reasonably well with the measurements of Norum and Seiner [57].

The structure of overexpanded jets has also been studied (see for example [26, 62]). The

structure of an overexpanded jet is similar to that of an underexpanded jet, in that both exhibit the

diamond-shaped shock cell structure. The difference lies at the nozzle exit and in the first shock

cell. As the exit static pressure is lower than the ambient value, the flow undergoes oblique shock

41

waves at the nozzle lip. This causes the flow to turn inwards, contracting the jet just outside the

nozzle, contrary to what happens in an underexpanded jet.

Jet spreading rate. An important characteristic of jets in general is their spreading rate, which is

the rate at which the jet diameter increases with distance from the nozzle exit, due to entrainment

of fluid from the surrounding medium. Supersonic jets spread much slower than subsonic jets.

For example, Eggers [20] calculated the entrainment of a Mach 2.22 air jet, and compared it to

that of a low speed jet (54 m/s, Mach 0.16). He found that the supersonic jet entrained much less

air than the low speed jet, and had a narrower divergence angle. Lau et al. [50] measured the

radial profiles of axial velocity in subsonic and supersonic free jets, and also reported that

supersonic jets spread much more slowly than subsonic jets, and consequently have a longer core.

An important study on this topic is of Papamoschou and Roshko [63], who measured the

growth of compressible shear layers using pitot probes, and compared it to that of incompressible

shear layers, and found the same result. They attributed this reduction in growth rate to the

compressibility of supersonic jets. To explain this, they used the concept of a convective Mach

number (Mac) to isolate the effects of compressibility on the spreading rate from the effects of

density and velocity differences. They defined this convective Mach number as the ratio of the jet

velocity relative to the large scale structures being convected with the jet, to the speed of sound in

the jet. Using an analytical model of a vortex sheet, they showed that a disturbance perturbs the

flow field in its immediate vicinity (which is the jet in the present case) more at lower Mac. As

Mac increases (higher jet Ma), the disturbance perturbs the flow away from the sheet more than it

perturbs the sheet itself. At supersonic Mac, energy is radiated away from the shear layer, which

slows down all processes responsible for entrainment and growth of the shear layer. They showed

that compressible shear layers can spread up to five times less than incompressible shear layers at

high convective Mach numbers.

42

2.4.2 Jet oscillation

Over and underexpanded jets emit loud noise, which consists of three main components – (1)

turbulent mixing noise, (2) broadband shock-associated noise, and (3) screech tones [96]. Screech

tones have discrete frequencies and large amplitudes, and contribute significantly to jet noise.

Powell [70] was the first to observe and describe these screech tones. He proposed an acoustic

feedback mechanism (reproduced in Figure 2.14 from the work of Raman [76]) in which the

boundary layer that forms inside the nozzle produces Kelvin-Helmholtz instability waves in the

shear layer of the jet outside the nozzle (item (1) in Figure 2.14). These instabilities grow as they

travel downstream along with the jet, forming large scale coherent structures [13, 63]. These

structures stretch, coalesce, and amplify, entraining the surrounding air. In over and

underexpanded jets, they interact with the shock cell structure as they travel downstream. This

interaction produces sound waves (item (2) in the figure) which travel upstream outside the jet

(item (3) in the figure), reflect from the nozzle rim (item (4) in the figure), and create new embryo

Figure 2.14. Acoustic feedback mechanism producing screech noise and oscillations in

a rectangular jet (instantaneous schlieren image, [76]).

43

instabilities in the shear layer, thus completing the feedback loop and emitting screech. This

mechanism has been validated by many subsequent studies (e.g. [61, 82, 88]).

An important consequence of the instability/shock cell interaction is that the jet

undergoes strong oscillations when it emits screech [96]. Two types of oscillations are possible

depending on the jet Mach number [95] and nozzle geometry [75] - toroidal mode and helical

mode oscillations. Equal amounts of the left and right helical modes produce a flapping mode that

yields an up and down oscillation, as illustrated in Figure 2.14. In the toroidal mode, the jet

oscillates axisymmetrically. These oscillations destabilize the jet flow, especially towards the end

of the core.

Seiner et al. [83] studied the phenomenon of twin supersonic plume resonance in relation

to supersonic aircraft exhaust plumes. Figure 2.15 (from their paper), shows the variation of the

screech tone wavelength with jet Mach number for a jet generated using a Mach 1.41 convergent-

divergent nozzle. In other words, the figure shows the different modes of oscillation exhibited by

the jet. Under certain conditions, the jet exhibited multiple modes. The toroidal mode was

observed for lower jet Mach numbers, whereas the flapping mode was observed for higher ones

(i.e. the jet exhibited mode switching).

2.5 Impinging Jets

This section briefly describes the main characteristics of the following flow scenarios: an

incompressible jet impinging on a cylinder, and a supersonic jet impinging on a flat surface as

well as on a cylinder.

2.5.1 Incompressible jet impingement on a cylinder

This flow configuration has been studied in the past to understand the heat transfer characteristics

of impinging jet flows; details can be found in [10, 81, 103]. Three main parameters govern this

44

Figure 2.15. Oscillation modes exhibited by a screeching jet generated using a Mach

1.41 convergent-divergent nozzle [83].

flow field (Figure 2.16) – (1) the distance

between the nozzle and cylinder (x), (2)

the nozzle diameter relative to the cylinder

diameter (de/D), and (3) the eccentricity or

offset of the jet with respect to the cylinder

axis (). The offset is simply defined as the

distance between the jet and tube

centerlines. A zero offset implies that the jet impinges on the tube head-on. When a cylinder of a

diameter comparable to the nozzle diameter is placed very close to the nozzle (in the core of a

jet), the impinging jet is narrow and has a very low level of turbulence; the boundary layer in the

impingement region is laminar. As a result, the jet splits into two parts which separate from the

cylinder surface at a certain position near the stagnation region. Due to this, the pressure on the

rear surface of the cylinder remains almost constant at a low value.

Figure 2.16. Parameters governing the flow

field of a jet impinging on a cylinder.

x

offset,

tube(outer diameter D)

x

nozzle(exit diameter de)

xx

offset,

tube(outer diameter D)

x

nozzle(exit diameter de)

45

Figure 2.17. Flow field of a jet impinging on a cylinder placed away from the nozzle; the

level of turbulence in the impinging jet is high.

When the cylinder is located further away from the nozzle, beyond the jet core (Figure

2.17), the level of turbulence in the impinging jet is much higher and the effective diameter of the

jet is larger, but its centerline velocity and impact pressure are lower. The boundary layer on the

cylinder surface is either turbulent, or transitions to turbulent. This causes the flow to remain

attached to the tube further from the impingement region and thus delays separation to further

downstream along the cylinder surface. The Coanda effect also contributes to this. A small wake

forms just behind the cylinder. The delayed separation leads to pressure recovery behind the

cylinder. Effects of the jet diameter relative to the tube diameter, and the eccentricity of the

cylinder position during impingement have also been studied previously and are omitted here for

brevity. For a cylinder submerged in unlimited parallel flow (where the cylinder width is much

smaller than the width of the flow), alternating eddies or vortices continuously form downstream

of the cylinder when the flow Reynolds number is between 102 and 107. These vortices are

commonly referred to as the ‘von Karman vortex street’. The non-dimensional frequency of

vortex shedding is given by the Strouhal number, St, mentioned briefly in section in 2.1.2, which

depends upon the flow velocity u, the cylinder diameter d, and the dimensional frequency at

which vortices are shed from the cylinder f: St = fd/2u. The Strouhal number for many

industrially encountered flows is typically around 0.2 [117, 120]. The alternating vortices lead to

alternating high and low pressure regions just downstream of the cylinder, causing the cylinder to

cylinderwake

free jet

wall jetseparation point

nozzlecylinderwake

free jet

wall jetseparation point

nozzle

46

oscillate or vibrate (‘flow-induced vibrations’).

Coanda effect. When the level of turbulence in an impinging jet is high, separation from the

cylinder surface is delayed because of two main phenomena - the boundary layer on the cylinder

surface is either already turbulent or quickly transitions to turbulent from laminar, and because of

the Coanda effect. The Coanda effect is an important phenomenon in the impingement of a jet on

a tube and more generally, on a curved surface. It is the tendency of a fluid jet directed

tangentially on a curved or angled solid surface to adhere to it. In the absence of a curved surface,

a free jet travels straight, entraining fluid from its surroundings (Figure 2.18a). When a curved

surface is present very close to the jet, the entrainment by the jet creates a low pressure region

between itself and the surface, due to which the jet is pulled towards the surface and attaches to it

Figure 2.18. The Coanda effect; (a) a free jet in the absence of a curved surface;

(b) a jet attached to a cylinder due to the Coanda effect.

47

(Figure 2.18b) [112]. The Coanda effect is experienced more strongly by turbulent jets than by

laminar jets due to the greater entrainment by turbulent jets, and is effective mainly for

incompressible and low subsonic flows. Supersonic flows, which have a high momentum and are

dominated by shock and expansion waves, are mostly not affected by this effect.

In industrial boilers, this effect strongly influences the flow behaviour of combustion

gases between tube bundles. The flow of gases between two rows of tubes is similar to a two-

dimensional jet. Such a jet discharging into a limited space does not flow straight, but deflects to

one side of the flow passage because of the Coanda effect [56].

2.5.2 Supersonic jet impingement on a flat surface

This flow configuration has been studied extensively in the context of vertical takeoff and landing

(VTOL) aircraft, thrust vectoring, and ground erosion due to exhaust plume impingement [46].

Figure 2.19 schematically shows a supersonic jet impinging on a flat surface. Three main

parameters govern this flow [55] – (1) the jet structure (over/underexpanded) characterized by the

nozzle exit pressure ratio (PR), (2) the distance between the nozzle and surface, and (3) the

surface inclination angle. The flow field can be classified into three main regions as indicated in

the figure – (1) the free jet flow, (2) the impingement region, and (3) the radial wall jet. When a

supersonic jet impinges on a flat surface, a normal shock (called the plate shock) forms slightly

upstream of the surface. Flow downstream of this shock is subsonic, changes direction according

to the surface angle, and accelerates. This gives rise to the radial wall jet.

One of the earliest studies of jet impingement on a flat surface is that by Donaldson and

Snedeker [18]. They studied the impingement of subsonic and underexpanded supersonic jets on

a flat plate at different angles, by measuring the pressure distributions on the plate surface. In the

case of a strongly underexpanded jet, they found that the maximum surface pressure occurs in a

ring around the jet centerline, instead of at a stagnation point. Using surface flow visualization to

analyze this phenomenon, they proposed that the ring forms due to a region of separated flow

48

Figure 2.19. Supersonic jet impinging on a flat surface.

between the surface and the plate shock, and

that a bubble of recirculating fluid forms in

this region (Figure 2.20). In the absence of a

bubble, they showed that the stagnation

point location and pressure change as the

surface inclination angle changes. Kalghatgi

and Hunt [34] proposed that the bubbles form due to the interaction of the plate shock with very

weak shock waves in the jet, produced either by small imperfections in the nozzle wall or by

small inaccuracies in the nozzle design. They also presented a criterion for when a bubble will

form.

Following Donaldson and Snedeker, Lamont and Hunt [49] conducted a comprehensive

experimental programme to study jet impingement on a flat surface. They varied all the

governing parameters listed above, and measured the surface pressure distributions and visualized

the flow fields using the shadowgraph technique. One of the major findings was that the peak

pressures on an inclined surface can dramatically exceed those on a perpendicular surface - by a

stagnation ring

bubble with recirculating fluidflat plate

jet

stagnation ring

bubble with recirculating fluidflat plate

jet

Figure 2.20. Bubble of recirculating fluid

between the plate and plate shock [18].

1

Free jet flow

Radial wall jet3Impingement

region

2

nozzle

flat plate

plate shock 1

Free jet flow

Radial wall jet3Impingement

region

2

nozzle

flat plate

plate shock

49

factor of 3 in some cases. In the far field, however, the maximum pressure always decreases with

plate inclination. This phenomenon was confirmed and explored further by later researchers [27,

55].

The radial wall jet has two parts – (1) the near wall jet, which originates from the

impingement region and consists of alternating shock and expansion waves, and (2) the fully

developed wall jet, which is free of any supersonic flow effects. Carling and Hunt [15]

investigated the near wall jet theoretically and experimentally. They showed that the near wall jet

is determined mainly by the jet-edge expansion and its reflections from the sonic line and wall jet

boundaries. Figure 2.21 (redrawn from their paper) schematically shows the shock layer,

transonic zone, and the beginning of the wall jet. The flow first expands downstream of the

intersection of the plate shock, the sonic line, and the jet boundary. These expansion waves reflect

from the sonic line and the wall jet boundaries, forming a network of compression and expansion

waves. Further downstream, these waves disappear due to interaction with turbulence, and the

r

z

plate shock

rJjet edge

constant pressure boundary

expansion wave

compression wave

limiting characteristic

sonic line

pa

wall jet shock layer

jet centreline

r

z

plate shock

rJjet edge

constant pressure boundary

expansion wave

compression wave

limiting characteristic

sonic line

pa

wall jet shock layer

jet centreline

Figure 2.21. The shock layer, transonic zone, and beginning of the near wall jet during

supersonic jet impingement on a flat plate [15].

50

near wall jet is converted into the fully developed wall jet. Donaldson and Snedeker [18] studied

the fully developed axisymmetric radial wall jet in detail, paying particular attention to the radial

velocity gradient. The velocity increases from zero at the wall to a maximum very close to the

wall (i.e. it increases significantly over a very small distance), and then diminishes gradually

towards the jet shear layer.

2.5.3 Supersonic jet impingement on a cylinder

In addition to the three main parameters which govern the flow field of an incompressible jet

impinging on a circular cylinder (the nozzle-cylinder distance, the jet diameter relative to the tube

diameter, and the offset), a fourth important parameter which affects the flow field of a

supersonic jet impinging on a cylinder is the jet structure (over/underexpanded) characterized by

PR. To date, this flow scenario has not been studied in detail, presumably due to a lack of

applications in aeronautical engineering. A survey of the open literature showed only a handful of

studies on this topic. The survey also yielded no studies of the interaction between a supersonic

jet and arrays of tubes. As a result, detailed information about the flow field that forms during

these impingement scenarios is not available. The few related studies are reviewed here.

In addition to studying jet impingement on a flat surface, Donaldson and Snedeker [18]

studied impingement on other surfaces including a convex hemisphere. In the case of the strongly

underexpanded jet, they found that the maximum surface pressure occurred in a ring around the

stagnation point, as occurs on a flat surface, and the pressure distribution was very similar to that

on the flat surface. Umeda et al. [113] experimentally investigated the discrete acoustic tones

generated by a jet impinging on a solid object. They impinged subsonic and supersonic jets on a

slender circular cylinder and found that the interaction between a downstream traveling eddy and

the cylinder is the main cause of the feedback loop responsible for producing jet noise. Thus, the

cylinder aids in noise production. However, they did not study the impingement flow field in

detail. Derbeneva et al. [17] simulated underexpanded jet flow over a sphere, and demonstrated

51

that the flow separates from the surface, with a suspended shock. However, they did not analyze

jet interaction with a cylinder.

Very few sootblowing-inspired studies have been performed to date; they are reviewed

next. More recently, Tabrizi [92] and Rahimi [73] studied the impingement of an underexpanded

jet on a cylinder, in the context of sootblowing in boilers. Tabrizi measured the surface pressure

coefficient around a cylinder impinged by an underexpanded jet, and studied the effects of

nozzle-tube distance and offset on the pressure coefficient. However, he focused mainly on the

interaction between a slot jet and a cylinder, and compared it to limited data on the interaction

with a round jet. He also performed numerical simulations to calculate the shear stress coefficient

around a cylinder impinged by a jet. Assuming that the shear stresses are primarily responsible

for cleaning tubes, he concluded that slot jets are more effective than round jets. Rahimi focused

on the heat transfer characteristics of jet impingement on a cylinder, and carried out temperature

measurements to determine the Nusselt number distributions. He found that the Nusselt number is

a maximum in a ring around the stagnation point on the surface, analogous to the very similar

pressure distribution for underexpanded jets. The Nusselt number decreases very quickly away

from the impingement region, indicating that boiler tubes could be subjected to severe thermal

stresses if a sootblower were placed very close to the tubes.

2.6 Conclusions from the Literature Survey

The following conclusions may be drawn from the literature survey:

(1) The review of the literature related to sootblowing optimization showed that no major study

has been conducted to understand the interaction between a sootblower jet and a single tube,

or an array of tubes. These interactions occur continuously during sootblowing. As the

sootblower jet is supersonic, these interactions may affect jet structure and strength, and

hence, sootblowing efficacy. As a result, understanding these interactions is important.

52

(2) The open literature on supersonic jets is almost entirely related to aerospace applications. As

a result, jet/tube interactions have not been studied in detail due to lack of related

applications. Studies have been conducted to understand the behaviour of free supersonic jets

and supersonic jets impinging on flat surfaces, and these studies provide valuable information

about the impingement flow.

(3) Inclined sootblower nozzles are commonly used in coal-fired boilers, where the inter-platen

spacing is greater than in recovery boilers. Due to concerns over reduced cleaning radius and

increased platen swinging, these inclined nozzles have not been used until very recently to

clean recovery boiler superheater platens. The feasibility of using such nozzles remains to be

demonstrated.

(4) The generating bank and economizer sections of modern recovery boilers are made up of

finned tubes. There is little information in the open literature about the flow of a supersonic

jet impinging on such a geometry.

CHAPTER 3

EXPERIMENTAL DESIGN AND

METHODOLOGY

The interaction of a supersonic jet with models of tube arrangements found in the superheater,

generating bank and economizer sections was investigated by means of flow visualization and

peak impact pressure (PIP) measurements. This work involved the design, fabrication and setup

of the following apparatus, and measurement and data acquisition systems:

scaled-down models of a typical sootblower nozzle, and superheater, generating bank and

economizer tube bundles;

high speed schlieren flow visualization system;

pitot probe and positioning system;

LabVIEW control and data acquisition system; and

image processing software.

These systems will be described in detail in this chapter.

53

54

3.1 Scaled-down Nozzle and Tube Bundles

The main components of the experimental apparatus were ¼ scale models of a typical sootblower

nozzle, and of superheater, generating bank and economizer tube arrangements. The design of

each of the different tube arrangements is described in chapter 6.

A scale-down factor of 4 was selected to keep the experimental apparatus of a reasonable

size and complexity. In the past, the author had performed artificial deposit breakup experiments

using a supersonic jet [66]. He used a ¼ scale apparatus in those experiments, in which air was

supplied to a nozzle from standardized Linde compressed air cylinders. He found that a jet of 2s

duration caused the supply pressure in the cylinder to decrease by about 15%. As a result,

reducing the scale-down factor in the present experiments to 2 for example would have yielded a

larger apparatus, making instrumentation for measurements more convenient, but would have

increased the supply pressure drop much more. Equation (2.16) for the mass flow rate through a

choked nozzle shows that the flow rate is directly proportional to the cross-sectional area of the

nozzle throat, or to the square of the throat diameter. Increasing the throat diameter by a factor of

2 increases the mass flow rate by a factor of 4. Preventing a large supply pressure drop would

have required much bigger equipment such as a large stagnation chamber. On the other hand,

increasing the scale-down factor to greater than 4 would have decreased the supply pressure drop,

but would have yielded a very small apparatus, making instrumentation troublesome. Therefore, a

factor of 4 was selected as a reasonable compromise between maintaining a steady jet supply

pressure and keeping the apparatus simple.

The nozzle was a convergent-divergent nozzle with a throat diameter, dt = 4.5 mm, an

exit diameter, de = 7.4 mm, and a cone half-angle of 6.4° for the conical expansion section.

Where superheated steam is used in actual sootblowers, air was used in these experiments for

reasons of safety and simplicity.

55

Figure 3.1 shows the experimental apparatus that was used for studying jet/tube

interaction. The nozzle was fixed on a two-direction slider arrangement on a workbench. The

sliders used in this arrangement were unislides and elevating tables from Velmex Inc.

Compressed air from a Linde high pressure supply cylinder was stored in a second buffer cylinder

Figure 3.1. Experimental apparatus; (a) schematic; (b) photograph.

(b)

solenoid valve

compressed air

nozzle

wooden clamp for nozzle

(a)

solenoid valve

camera in schlieren system

supersonic nozzlecompressed air

pitot probe pressure transducerDAQ

SYSTEM

dt=4.5mmde=7.4mm

(b)

solenoid valve

compressed air

nozzle

wooden clamp for nozzle

(a)

solenoid valve

camera in schlieren system

supersonic nozzlecompressed air

pitot probe pressure transducerDAQ

SYSTEM

dt=4.5mmde=7.4mm

(a)

solenoid valve

camera in schlieren system

supersonic nozzlecompressed air

pitot probe pressure transducerDAQ

SYSTEM

de=7.4mmdt=4.5mm

56

and supplied to the nozzle through a solenoid valve (series 21EN, manufactured by Granzow Inc.)

such that the pressure at the nozzle inlet was 2.14 MPa gauge (310 psig), similar to that of an

actual sootblower [32, 111]. The nozzle inlet pressure was maintained at this value for all

experiments. In each experiment, the jet was blown for 0.2s. This duration was found to be

optimal considering both the time required for completely pressurizing the pitot probe, and the

amount of air used per experiment. Preliminary experiments were performed in which the

temporal development of the jet from its initial transient stages to its quasi steady-state was

recorded, and showed that the jet attained quasi steady-state in less than about 20 ms, much

shorter than the total jet duration of 0.2s (or 200 ms). Schlieren images of the temporal

development of the jet are presented in Appendix A. The jet impinged on model tubes or tube

bundles placed in front of the nozzle and the resulting interaction was visualized using the

schlieren technique and captured by a high-speed camera.

3.1.1 Similarity of the lab air jet to an actual sootblower jet

For the experiments to reflect actual sootblower jet usage, the lab jet was designed to be

geometrically and dynamically similar to a sootblower jet [32, 111]. First of all, air was used in

these experiments instead of steam for reasons of safety and simplicity mentioned earlier. This is

justified based on the following reasons:

1. Sootblowing steam is superheated; therefore, it is a homogeneous fluid. For dynamic

similarity, the lab jet must also be of a homogeneous fluid, which is air in these experiments.

2. Dynamic similarity of a compressible flow requires that the ratio of specific heats, = cp/cv

should be similar for both the lab jet (air) and a sootblower jet (steam). of steam is ~1.3,

whereas that of air is ~1.4.

3. Recently, field trials were conducted in operating recovery boilers in Sweden, to measure

sootblower jet force by a special probe [80]. The data were found to be consistent with those

57

obtained from lab scale experiments using air as the jet fluid [35, 40].

Geometric similarity was achieved by scaling the nozzle throat and exit diameters four

times smaller than the corresponding diameters of the actual nozzle.

For the lab jet to be dynamically similar to an actual jet, three aspects were considered:

nozzle exit Mach number, Mae, jet structure (under/overexpanded), and jet spreading rate (rate at

which the jet diameter increases with distance from the nozzle exit due to entrainment of the

surrounding fluid; see section 2.4.1). Theoretically, Mae is only a function of the ratio of the

nozzle exit diameter to throat diameter (see equation 2.15). Since the lab nozzle is geometrically

similar to an actual nozzle, the same Mae was obtained for the lab jet as an actual jet. Using the

basic one-dimensional nozzle flow theory summarized in section 2.2.3, the air supply pressure

was set such that an almost fully expanded jet with a nozzle exit Mach number, Mae of 2.5 was

generated.

Actual sootblower jets are always either underexpanded or overexpanded due to supply

pressure fluctuations and other reasons. A given nozzle generates an underexpanded jet when the

supply pressure is greater than the nozzle design pressure, and generates an overexpanded jet

when the supply pressure is lower than the design pressure. As a result, underexpanded jets are

more powerful than overexpanded jets due to the higher steam flow and are preferred over

overexpanded jets. At 2.14 MPa gauge (310 psig), the lab jet was slightly underexpanded, and

hence suitable for experiments.

Three parameters that affect the spreading rate of supersonic jets are (in decreasing order

of importance): (i) Mae, (ii) the ratio of the velocity of the stream surrounding the jet (the co-

flowing stream) to the velocity of the jet, and (iii) the ratio of the density of the co-flowing stream

to the density of the jet [63]. Mae was the same for both the lab and actual jets. The velocity ratio

was effectively 0 in the experiments and inside the boiler. The density ratio is ~4 inside the boiler

whereas it was ~3 in these experiments. As a result, the spreading rate of the lab jet was very

close to that of an actual jet.

58

The field trials in Sweden referred to above [80] showed that the flue gas temperature in

the vicinity of a sootblower does not affect the sootblower jet strength in any significant way.

Two trials were carried out using the same sootblower in a recovery boiler at different times. At

the time of the first trial, the boiler was about to be commissioned and was operating on gas, so

the flue gas temperature in the vicinity of the sootblower was low, between 100°C-300°C. At the

time of the second trial, the boiler was burning black liquor at 50% of its full firing capacity, and

the flue gas temperature was much higher, between 500°C-540°C. The results of the tests

performed during both these trials were similar to and consistent with each other; no significant

difference was observed between them.

In the ideal scenario, the lab jet Reynolds number based on the conditions at the nozzle

exit, Ree, should also be similar to that of an actual jet, to ensure dynamic similarity. Using

Sutherland’s law to calculate viscosity [117], Ree of both the lab jet as well as an actual jet were

determined to be of the order of 106 (around 1.6 x 106 for the lab jet and 1.9 x 106 for an actual

jet). However, it should be noted that for supersonic flow involving obstacles, the Mach number

has a much greater influence on the fluid dynamics than the Reynolds number [117].

Finally, the effect of the scale-down process on jet PIP and force should also be

determined. The PIP at a given location in the supersonic portion of a jet depends only on the

local total pressure and Mach number upstream of a normal shock wave, and the specific heat

ratio, , of the jet fluid (see equation (2.8)). The local upstream total pressure and Mach number

in turn depend on the jet supply pressure and the nozzle exit Mach number, as well as the nozzle

design. As mentioned earlier, both the supply pressure and the exit Mach number of the lab jet

were similar to those of an actual sootblower jet. Any minor differences will be due to differences

in the nozzle design. Moreover, the difference in of air and steam ( of air = 1.4 versus of

steam = 1.3) will contribute only little to the difference between the PIP of the lab jet and an

actual jet. For example, the PIP along the centreline of a Mach 2.5 fully expanded jet core differs

59

by 8% between air and steam jets (with the same supply pressure). As a result, although there will

be a difference in the PIP of the lab and actual jets, the cumulative effect of the difference in the

nozzle design and on the difference in the PIP will not be large.

Control volume analysis just outside the nozzle using appropriate scale-down factors and

isentropic relations showed that the force of the lab jet was approximately 14-15 times smaller

than the force of an actual sootblower jet, mainly due to the reduction in the nozzle exit diameter

by a factor of 4, which leads to a decrease in area of 16 times. However, the present thesis is a

study mainly of the jet flow field and PIP during interaction with tubes, and does not involve a

study of jet force. As a result, the effect of the scale-down process on jet force is not important in

this work.

For the above reasons, the lab air jet was considered similar to an actual sootblower jet,

and the results of this work are applicable to actual sootblowing inside a boiler.

3.2 High-Speed Schlieren Flow Visualization System

The schlieren technique was employed to visualize the invisible air jet and its interaction with

tubes and tube bundles. Due to the very high velocity of supersonic jets, a high-speed camera was

coupled with the schlieren technique.

As described in section 2.3, a conventional two-mirror, z-type schlieren system was used

for flow visualization. The system was supplied by Optikon Inc. and consisted of two parabolic

mirrors, a light source with adjustable aperture, and a knife edge. The mirrors were made of

Pyrex, and had a focal length of 1524 mm (60”), and a diameter and hence field-of-view of 140

mm (5.5”) (Figure 3.2a). The light source was a Schott KL2500 series continuous halogen cold

light source. Custom-designed telescoping metal stands were fabricated for supporting the

mirrors. The knife edge was positioned on an adjustable tripod.

The high-speed camera used was a MegaSpeed 70KS2B2HS from Canadian Photonics

60

(a) (b)(a) (b)

Figure 3.2. (a) Parabolic mirror used in the schlieren system, with custom-designed

stand; (b) high-speed camera.

Labs (Figure 3.2b). It was operated at 6010 frames/s with an exposure time of 150 s. The images

were captured as 504 pixel x 504 pixel greyscale images. All optical components were aligned as

per the schematic in Figure 2.11.

3.3 Pitot Probe and Positioning System

A special pitot probe was designed and fabricated to measure the jet PIP inside the model tube

bundles, and to quantify jet/tube interaction; Figure 3.3a shows this probe. Several references (for

e.g. [11, 14, 64, 77]) were consulted in order to design the probe appropriately for supersonic jet

flow. In supersonic flow, a bow shock wave forms just upstream of the orifice of the pitot probe.

The portion of the bow shock directly in front of the orifice can be treated as a normal shock

described in chapter 2 section 2.2.2.

61

Dimensions in inchesDimensions in inches

(a)

Figure 3.3. Pitot probe designed to measure supersonic jet PIP; (a) photograph; (b)

schematic showing probe internal design.

The length and diameter of the probe were chosen to allow the probe to travel freely even

between the closely spaced tubes of the model generating bank and economizer. The probe

interior (Figure 3.3b) was designed so that the jet could completely pressurize the probe within

0.2s, to yield a stable, reliable, and accurate reading. Figure 3.4 shows a typical voltage signal of

the jet peak impact pressure obtained using the pitot probe and a pressure transducer (to be

described later). The initial part of the signal is the response time of the pitot probe, during which

the jet develops and attains a steady state, and the probe is completely pressurized to the correct

steady-state reading. The middle portion of the signal corresponds to the jet steady- state, which

is averaged to yield the jet peak impact pressure; even though this portion is referred to as the

steady-state reading, note that the signal pressure decreases slightly in this portion due to a

decrease in the jet supply pressure. The final portion of the signal corresponds to the

depressurization of the probe after the jet blow is complete. The response time of the pitot tube

depended on the experimental conditions, but was lower than 80 ms for all experiments. The tip

of the probe was a square-cut orifice, to make the probe insensitive to pitch and yaw

misalignments of up to 5°-7°. The tip was long enough to isolate the measurement region from

OD 3ID 2¼” NPT

fitting

(a)

(b)

25 305

Dimensions in mm

51

OD 10OD 3ID 2¼” NPT

fitting

(b)

25 305

Dimensions in mm

51

OD 10

62

jet steady state signal

Figure 3.4. Typical voltage signal of jet peak impact pressure, obtained using the pitot

probe.

flow disturbances occurring further downstream. The other end of the probe was connected to a

pressure transducer.

The probe was fixed on a three-dimensional slider arrangement (Figure 3.5), and could be

rotated about its own longitudinal axis to measure jet PIP at different locations inside the model

tube bundles. A magnetic stand was used to clamp the pitot probe to reduce vibration (inset in

Figure 3.5).

Pressure transducer. An Omega PX 613 series pressure transducer rated for recording pressures

between 0-2.07 MPa (0-300 psig) was used to record the jet PIP. It was first calibrated and then

used to measure jet PIP.

3.3.1 Repeatability of measurements

For all experimental measurements made in this work, uncertainty (error bars) has been reported

probe depressurizationprobe

response time

jet steady state signal

probe depressurizationprobe

response time

63

Figure 3.5. Pitot probe positioning system (inset shows the magnetic stand used

to clamp the probe to reduce vibration).

in terms of the standard error associated with each obtained data set. Error bars for each data

point shown in a graph are equal to +/- 1 standard error. For a data set or a sample consisting of n

measurements q1, q2, … qn of the same quantity under the same conditions, the average, q ,

standard deviation, SD, and standard error, SE, of that data set are given by the following

relations:

n

iiq

nq

1

1 … (3.1)

)1(

2

11

2

nn

qqn

SD

n

ii

n

ii

… (3.2)

n

SDSE … (3.3)

In the present experiments, the repeatability of measurements was confirmed by first

examining the variation in jet supply pressure and the centreline PIP of a free jet in detail. This

64

was done by repeating those measurements several times. Because those measurements were

found repeatable with only small uncertainty, any subsequent measurements were repeated only

once (that is, two measurements under the same conditions) to confirm their repeatability.

The uncertainty in the jet supply pressure was determined based on 25 measurements;

only for these tests, a transducer other than the one described above was used, since it had a

pressure rating higher than 2.07 MPa (300 psig). The average supply pressure was found to be

2.14 MPa gauge (310 psig) with a standard deviation of only 14 kPa (2 psig) or 0.7% of the

supply pressure, and a standard error of 2.8 kPa (0.4 psig).

Uncertainty in the jet PIP measurements can be introduced by uncertainty in the supply

pressure as well as in the signal generated by the transducer (random errors). The normal shock

relation (equation 2.8) shows that the PIP in the supersonic portion of a jet depends mainly on the

total pressure and Ma upstream of a normal shock wave that forms at the probe tip; the PIP is

directly proportional to the upstream total pressure. The PIP at a subsonic location is the total

pressure of the flow at that location (since no shock waves form). Both the upstream total

pressure and Ma in turn depend on the jet supply pressure. As a result, any uncertainty in the

supply pressure will lead to uncertainty in the PIP measurement. However, this contribution (due

to 0.7% variation in the present experiments) is expected to be very small for a mildly

underexpanded jet, which is the case in the present experiments; the shock cells of such a jet are

weak, yielding mild changes in jet properties such as the total pressure and Ma in the radial and

axial directions.

The uncertainty in the signal recorded at each axial location of the jet was determined

from at least 3 measurements at each location (4 in some cases). The maximum standard

deviation was found to be 70 kPa (10 psig) for a PIP equal to 0.84 MPa gauge (122 psig) at 2.8

cm from the nozzle exit, that is, 8.3% of the PIP. The standard error was 40 kPa (6 psig).

65

3.3.2 Accuracy of measurements

Accuracy of a measurement can be determined by comparing the measured value to that obtained

by some other method for the same conditions. This other method is preferably a theoretical

analysis of the phenomenon governing the quantity being measured. In case the phenomenon is

too complicated for theoretical analysis, the accuracy can be determined by comparison with data

from previous similar experiments. To confirm that the pitot probe measured the jet PIP

accurately in the present experiments, the PIP measured at the nozzle exit was compared to that

calculated theoretically using the normal shock relation (equation 2.8). This theory allows

calculation of the flow inside the nozzle, up to the nozzle exit. Beyond the exit, theoretical

analysis is complicated due to the presence of shock cells and due to shock-turbulence

interaction.

The pitot probe was used to measure the supply pressure at the nozzle inlet as well as the

nozzle exit PIP. The probe directly measured the jet supply pressure at the nozzle inlet, since the

flow in this region is subsonic and free of any shock/expansion waves. The pressure measured at

the nozzle exit was the jet PIP behind a normal shock wave at the probe tip. The measured supply

pressure and the theoretically calculated nozzle exit Mach number (determined from equation

2.15) were inserted in equation 2.8 to calculate the expected nozzle exit PIP. This was carried out

for four different supply pressures, and for each supply pressure, the measurements were repeated

twice, and were found reproducible. The results are shown in Table 3.1.

It is clear from Table 3.1 that the pitot probe functioned properly, and the measured PIP

values were accurate. The measured and calculated nozzle exit PIP values for all the tested supply

pressures were found to be very close to each other. In fact, the error relative to the calculated PIP

decreased as the supply pressure increased, and was only 1% at a supply pressure of 2.15 MPa

gauge (312 psig). At other positions downstream of the exit, the PIP profile of a free jet was

consistent with that available in the literature under conditions similar to the present experiments

[47]. Hence, the probe and measurement system were deemed accurate and capable of measuring

66

the jet PIP between tubes in different arrangements.

Table 3.1. Performance of pitot probe – comparison of calculated and measured nozzle

exit PIP.

No.

Measured supply

pressure (po)

[MPa gauge]

Calculated

nozzle exit PIP

[MPa gauge]

Measured

nozzle exit PIP

[MPa gauge]

Error

%

1 1.55 0.70 0.75 7.1

2 1.86 0.86 0.92 7.0

3 2.15 1.00 1.01 1.0

4 2.49 1.17 1.17 0.0

3.4 LabVIEW Control and Data Acquisition (DAQ) System

A control and data acquisition (DAQ) system was setup to control the solenoid valve, high-speed

camera, and pitot probe pressure transducer for each experiment, and to acquire data from the

transducer. This control system consisted of the following components:

A desktop PC running Windows XP;

National Instruments (NI) LabVIEW version 8.0 software to control all the hardware;

NI PCI-6221 (37 pin) multifunction data acquisition module to sample the analog voltage

signals generated by the pitot probe pressure transducer;

NI SCXI 1000 chassis to house the PCI-6221 module;

NI SCXI 1302 terminal block to make connections between the hardware components; and,

Circuit to control the operation of the solenoid valve.

Operation of the system. Figure 3.6a illustrates the DAQ system and Figure 3.6b shows the

control system hardware and connections. The solenoid valve was controlled by a circuit

consisting of a transistor and relay, and controlled through LabVIEW. Through this arrangement,

the solenoid valve could be opened and closed easily for any duration of time. The control system

67

Figure 3.6. (a) LabVIEW control and data acquisition (DAQ) system diagram; (b)

photograph of the control system hardware.

was used to power the pressure transducer and receive the analog voltage signal. The system was

also used to trigger the high-speed camera used in the schlieren system.

For each experiment, the following steps were performed in sequence:

1. Through the DAQ system, a signal was first sent to operate the pitot probe pressure

transducer and start recording the voltage output from the transducer.

~

In

OutTrigger

High-speed camera

Pitot probe pressure transducer

Solenoid valve control circuit

DAQ

SYSTEM

(a)

(b)

~

In

OutTrigger

High-speed camera

Pitot probe pressure transducer

Solenoid valve control circuit

DAQ

SYSTEM

~~~~

In

OutTrigger

High-speed camera

Pitot probe pressure transducer

Solenoid valve control circuit

DAQ

SYSTEM

(a)

(b)

68

2. The high-speed camera was then triggered to start recording.

3. Then, the solenoid valve was opened for 0.2s during which a supersonic air jet exited from

the nozzle and impinged on a model tube or tube bundle placed in front of the nozzle. The

resulting interaction was captured by the high-speed camera. The jet PIP voltage signal

generated by the pressure transducer was simultaneously recorded by the DAQ system.

4. The valve was then closed and recording of the transducer output signal was stopped. By the

end of the experiment, the camera was also stopped.

5. The captured images and transducer signal were processed and analyzed. For each

measurement, a fixed number of samples from the steady-state region of the recorded voltage

signal was averaged, and the average value was converted to gauge pressure using the

transducer calibration equations.

The delay associated with the DAQ system was measured to be about 0.14s and was

taken into account when programming the system, as it was unavoidable. The LabVIEW

graphical program developed to operate the system described above is included as Appendix B.

3.5 Image Processing

For most of this work, image processing was used to enhance the brightness and contrast of

individual schlieren images captured by the high-speed camera. However, in the work related to

secondary jets, for each experiment, image processing was first used to average a large number of

images showing the steady-state flow field. This averaged image was contrast-enhanced and then

used to measure the secondary jet angle. Figure 3.7 shows a sample contrast-enhanced average

image. ImageJ software was used to process and analyze the images [29].

69

Figure 3.7. Sample of a contrast-enhanced average image.

CHAPTER 4

FREE JET CHARACTERIZATION

The main objective of this work was to study the interaction between a supersonic jet and tube

bundles with different tube arrangements. A necessary prerequisite to studying this interaction is

the characterization of the supersonic jet itself, so that the interaction with a tube bundle can be

characterized by comparison with the free jet. Jet flow between two rows of tubes is an important

flow scenario, and knowledge of the jet size (radius or diameter) is very important to correctly

understand this jet/tube interaction. Jet size determines if strong interaction will occur between a

jet and a tube bundle (depending on the spacing between the tubes), and if interaction occurs, jet

size determines the location of the initial point of interaction. As a result, one of the objectives of

characterizing the free jet was determining the jet radius as a function of the axial distance from

the nozzle exit, that is, to determine the jet spreading. This chapter describes the structure and

characteristics of the supersonic jet used in this work.

70

71

4.1 Jet Structure

As described in section 3.1, the jet was generated using a supersonic nozzle with a throat diameter

of 4.5 mm and an exit diameter of 7.4 mm, and the supply pressure was set to 2.14 MPa gauge

(310 psig). Using one-dimensional isentropic flow relations and nozzle flow theory summarized

in sections 2.2.1 and 2.2.3, the nozzle exit Mach number, Mae was calculated to be 2.5 and the jet

was slightly underexpanded. Theoretically, the nozzle exit pressure ratio, PR was calculated to be

around 1.2 (20% underexpansion), but as indicated in section 2.2.3, the flow inside any nozzle is

not fully isentropic. Due to friction losses and heat transfer, the flow is non-isentropic, and the

extent of its deviation from isentropic conditions depends on the design of the nozzle. The

smoother the nozzle, the smoother and more nearly isentropic the flow. As a result, the static

pressure at the nozzle exit was expected to be lower than that calculated theoretically. Hence, PR

was expected to be lower than 1.2 (1 < PR < 1.2), and the jet was expected to be even weakly

underexpanded (< 20% underexpansion), and therefore, closer to being fully-expanded.

nozzle (de=7.4mm)

shock cell

nozzle (de=7.4mm)

shock cell

Figure 4.1. Supersonic jet used in this work.

Figure 4.1 is a schlieren image of the jet used in this work, and shows that the jet was

indeed very slightly underexpanded. Consistent with previous supersonic jet studies reviewed in

section 2.4, the jet contains diamond shaped shock cells indicating a pressure mismatch at the

nozzle exit, and the slight increase in jet diameter just downstream of the nozzle exit indicates

72

that the static pressure at the nozzle exit is higher than the ambient pressure. The size and strength

of the shock cells decrease with distance from the nozzle due to interaction with the turbulence

generated in the shear layer; this is evident from the gradually decreasing intensities of the shock

cells.

4.2 Centreline Peak Impact Pressure

The pitot probe described in section 3.3 was used to measure the centreline peak impact pressure

(PIP) of the jet. Based on the jet structure visible in the schlieren images and the sizes of the

shock cells, measurements were taken every 2 mm in the supersonic portion of the jet, to resolve

the shock cell structure. In the transition and fully developed regions of the jet, where the PIP

decreases monotonically, the spacing between measurements was increased to 5 mm and 10 mm

respectively. In the far field of the jet, where the jet has diffused almost completely, measurement

resolution was further decreased to every 30 mm and 50 mm. Measurements were repeated three

times and were found to be reproducible.

Figure 4.2 presents the centreline PIP variation of the jet. The PIP is normalized by the jet

supply pressure, po, and the distance along the jet centreline, x, is normalized by the nozzle exit

diameter, de. x = 0 at the nozzle exit plane. The standard error is included in the plots. The nozzle

exit PIP calculated theoretically using the normal shock relation (equation 2.8), with Mae = 2.5

and po = 2.14 MPa gauge (310 psig), is also shown on this plot (by the single circle at x/de = 0);

as described in chapter 3 section 3.3.2, the measured value agrees very well with the calculated

one (difference of 1% relative to the calculated value).

The potential core of the jet, in which the flow is supersonic, is around 18 nozzle

diameters long. The PIP just outside the nozzle exit first decreases, indicating that the jet is

underexpanded. This is because an underexpanded jet completes its expansion outside the nozzle

by means of isentropic expansion waves, across which PIP decreases. As the jet was

73

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40

x/de

PIP/po

50

potential core(supersonic)

theoretical nozzle exit PIP assuming isentropic flow

oscillations due to expansion-compression waves

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40

x/de

PIP/po

50

potential core(supersonic)

theoretical nozzle exit PIP assuming isentropic flow

oscillations due to expansion-compression waves

Figure 4.2. Centreline PIP variation in the supersonic jet.

underexpanded, the PIP can be seen to oscillate in this region because of the conical expansion

and compression waves forming the shock cells. Approximately eight shock cells were captured

by the pitot probe measurements; this is consistent with the number of shock cells captured in the

schlieren image of Figure 4.1. The PIP oscillations decay towards the end of the core as the shock

cells diffuse. Thereafter, the PIP decreases monotonically because of turbulent mixing. The jet

structure and its centreline PIP profile were found to be very similar to that reported by Kweon et

al. [47], who used a jet under operating conditions similar to the conditions in the present

experiments.

74

4.3 Radial Peak Impact Pressure and Jet Spread

The term peak impact pressure (PIP) generally refers to the pressure measured at the centreline of

a fully-expanded jet using a pitot tube, and is termed ‘peak’ because this pressure is the maximum

pressure across the jet cross-section at that axial location. However, in this thesis, PIP will also be

used to refer to jet pressure measured off the centreline using a pitot probe, to avoid using

additional terms. Radial PIP profiles were obtained at five axial locations downstream of the

nozzle exit, at x/de = 1.4, 5.4, 10.8, 16.2, and 21.6. At each axial location, the PIP was measured

every 1 mm across the jet (the minimum spacing possible with the apparatus), until the PIP

decreased to zero across the jet boundary or shear layer. Each profile was measured at least three

times and found to be reproducible. Figure 4.3 presents the radial PIP profiles of the jet at the five

axial locations. Similar to Figure 4.2, the PIP is normalized by po and r is normalized by de. The

standard error is also included in the plots.

At x/de = 1.4 (Figure 4.3a), the PIP profile is only as wide as the nozzle. The maximum

PIP does not occur at the jet centreline, but towards the outer part of the jet, in a ring around the

jet centreline. As described in chapter 2 sections 2.2.3 and 2.4.1, the air around the jet centreline

at this axial location passes through an expansion fan originating at the nozzle lip, whereas the air

in the outer region of the jet passes through an incident oblique shock, and is compressed. As a

result, the PIP increases radially outward from the jet centreline. It then rapidly drops to zero

across the jet shear layer because the shear layer is very thin at this axial location.

The profile at x/de = 5.4 (Figure 4.3b) is similar to that at x/de = 1.4, because these axial

locations are very similar positions within the first and second shock cells; so the flow at x/de =

5.4 is similar to that at x/de = 1.4. However, the shear layer at x/de = 5.4 has grown, as indicated

by the relatively slower rate of decrease of PIP across the shear layer; consequently, the jet has

spread radially.

At x/de = 10.8 (Figure 4.3c), the PIP profile is flat near the centerline, and then decreases

monotonically. This indicates that the shock cells are weaker at this axial location compared to

75

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3r/de

PIP/po

Figure 4.3. Radial PIP profiles of the jet at different axial locations.

locations nearer the nozzle. The rate of decrease of PIP across the jet is slower than in the

previous cases, indicating more jet spread.

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3r/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3

r/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3

r/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3

r/de

PIP/po

(a)

(c)

(b)

(d)

(e)

x/de = 1.4 x/de = 5.4

x/de = 16.2x/de = 10.8

x/de = 21.6

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3r/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3r/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3

r/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3

r/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0 1 2 3

r/de

PIP/po

(a)

(c)

(b)

(d)

(e)

x/de = 1.4 x/de = 5.4

x/de = 16.2x/de = 10.8

x/de = 21.6

76

The profiles at x/de = 16.2 and 21.6 (Figures 4.3d and 4.3e respectively) are similar; the

PIP is maximum at the jet centreline and decays gradually towards the jet boundary. The

maximum PIP is lower than that in the upstream profiles, but the radial spread is larger, reflecting

conservation of axial momentum. x/de = 21.6 lies beyond the jet core region; the flow at this

location is subsonic and the PIP profile is characteristic of fully developed jet flow.

As indicated in chapter 2 section 2.4.1, the majority of studies of supersonic free jets have

been of strongly underexpanded jets, which contain a normal shock wave (or Mach disk) just

outside the nozzle. Far fewer studies have been carried out on mildly underexpanded jets, and

these do not provide data similar to the data presented above. Nevertheless, the above results are

qualitatively consistent with the results of these studies [18, 42, 47, 60].

Jet spreading. Jet radius was determined from the radial PIP profiles via the method of

Papamoschou and Roshko [63]. Jet radius at any axial location was defined as the radial distance

from the jet centreline to a point where the PIP is 5% of the maximum value in the cross-section

at that axial location (Figure 4.4a). This is a commonly used definition in supersonic jet literature

[47, 74].

Figure 4.4b shows the variation of the jet radius in the axial direction, as well as the jet

radius data of Kweon et al. [47], who studied the effect of a nozzle exit reflector on the structure

of over and underexpanded jets. Their study involved a slightly underexpanded jet from a nozzle

similar to the one used in the present study, under similar operating conditions (supply pressure,

etc.) as mentioned above; however, their jet was slightly greater underexpanded than the one used

in this work. Again, the agreement between the two is good. The jet radius first increases because

of the expansion of the jet outside the nozzle. Then, the radius increases in the axial direction due

to the entrainment of the surrounding air and turbulent mixing. It should be noted that the

spreading rate is much smaller than that of a subsonic jet, as explained in section 2.4.1 [20, 50,

63].

77

Figure 4.4. (a) Definition of jet radius; (b) jet radius versus axial distance from nozzle

exit (jet spread); data from [47] is shown for comparison.

0

0.5

1

1.5

2

2.5

0 5 10 15 20 25

x/de

r/de

Present experiments

Kweon et al [47]

(a)

0

0.5

1

1.5

2

2.5

0 5 10 15 20 25

x/de

r/de

Present experiments

Kweon et al [47]

(a)

CHAPTER 5

INTERACTION BETWEEN A JET AND A

SINGLE TUBE

As described in the Introduction (chapter 1), tubes in a recovery boiler are arranged in rows.

During sootblowing, the jet almost always interacts with the first tube of a given row. Depending

on the spacing between rows, this interaction can occur for a significant portion of the

sootblowing time. As a result, as a first step in studying the flow and interaction of a sootblower

jet with tube bundles, we consider the interaction between a supersonic jet and a single tube.

5.1 Experimental Parameters

As mentioned in chapter 2 section 2.5.3, the main parameters governing the interaction between a

supersonic jet and a tube are the jet structure (over/underexpanded), as characterised by the

nozzle exit pressure ratio, PR, the distance between the nozzle and tube, x, the size of the jet (dj)

relative to the size of the tube (D), and the offset between the jet and tube centrelines, . In this

work, the same jet described in chapter four was used in all the experiments. Consequently,

experiments were performed to investigate the effects of offset, nozzle-tube distance, and tube

size on jet/tube interaction.

78

79

In the experiments, the offset was increased incrementally from zero to a value at which

the jet was so far from the tube that no interaction occurred. The nozzle-tube distance was

measured from the nozzle exit plane to the windward surface of the tube, and was varied between

1de – 38de. Tubes of three outer diameters were used – 12.7 mm (1/2”), 19.1 mm (3/4”), and 25.4

mm (1”); their sizes normalized by the nozzle exit diameter (de = 7.4 mm) are de/D = 0.58, 0.39,

and 0.29 respectively. These tubes were denoted as small, medium, and large respectively. The

small tube (1/2”) was a ¼ scale typical superheater tube (2”); the medium and large tubes served

two purposes – (1) they enabled investigation of the effects of tube size on the jet/tube

interaction, and (2) they represented tubes with big deposits.

5.2 Effect of Offset between Jet and Tube Centrelines

For a given nozzle-tube distance and tube size, the jet was directed at the tube at different offsets

and the interaction was visualized. Figures 5.1a and 5.1b show schlieren images of jet/tube

interaction for the small and medium tubes (de/D = 0.58 and 0.39) at a nozzle-tube distance of 50

mm, which at a ¼ scale corresponds to the distance inside a boiler. The offset is normalized by

the tube outer radius R. Similar results were obtained for the large tube, and so are omitted here;

they are included in Appendix C for reference.

Figures 5.1a and 5.1b show that upon impingement on a tube at an offset, the supersonic

jet deflects at an angle that depends on the offset. At zero offset (image a in Figure 5.1a), when

the jet impinges on the tube head-on, the jet splits into two, small symmetric jets (the lower jet

cannot be seen because of the tube stand). As the offset increases (image b onwards in Figures

5.1a and 5.1b), the interaction between the jet and the tube weakens, and the upper jet deflects

less and becomes stronger, whereas the lower jet becomes weaker. Beyond a certain offset (image

i in Figures 5.1a and 5.1b), there is no interaction between the jet and the tube, and no jet

deflection occurs.

80

Figure 5.1. Jet impinging on a tube at different offsets.

81

Figure 5.2. Formation of secondary jets.

Figure 5.2 schematically shows the interaction between the jet and a tube. The impinging

jet is termed ‘primary’ jet, whereas the deflected jet is termed ‘secondary’ jet; from this point

onwards in this thesis, these terms will be used to refer to these jets. When the primary jet

impinges on a tube, a shock wave forms upstream of the tube across which the total pressure of

the flow decreases and static pressure increases. The flow accelerates from this impingement

region, and separates from the tube surface as a secondary jet some distance downstream.

The secondary jets show the presence of compression and expansion waves, indicating

they are supersonic jets. These waves arise from the interaction between the shock wave upstream

of the tube, the sonic line of the flow in the impingement region, the tube surface, and the

constant pressure boundary of the secondary jet, similar to how they form in the wall jets that

develop when a supersonic jet impinges on a flat surface (chapter 2 section 2.5.2). However, in

the latter case, the jets remain attached to the flat surface, whereas upon impinging on a tube, the

jets may or may not remain attached, depending on the nozzle-tube distance and the tube size. In

other words, the secondary jets described here can be considered, in a way, as separated wall jets.

Wave reflections in the wall jets occur between a constant pressure boundary on one side (jet

boundary) and a rigid wall on the other; in the secondary jets considered here, the reflections will

also occur until the flow separates from the tube; then, the reflections occur between the

boundaries of the jet, which are all at constant pressure.

secondary jet angle,

nozzle

primary jet

secondary jet

tube

tube radius, R

offset,

secondary jet angle,

nozzle

primary jet

secondary jet

tube radius, R

offset,

tube

82

As described in chapter 2 section 2.5.1, earlier studies of jet impingement on a cylinder

[10, 81] have shown that the impinging jet splits into two parts when the cylinder is located in the

potential core of the jet, and that these two parts separate from the surface of the cylinder beyond

a certain distance along the surface downstream of the impingement region. Those studies

involved incompressible jets. The results of the present work involving supersonic jets are

consistent with the results of those studies.

Returning to Figure 5.1, another important observation can be made by comparing the

interaction at 0 offset for the two tubes (image a in Figures 5.1a and 5.1b). Secondary jets form in

the case of the small tube at 0 offset, whereas they do not in the case of the medium tube. For the

medium tube, they are apparent at an offset of 0.25R, implying they form at an offset somewhere

in between. Below that offset, wall jets exist. Air accelerating from the impingement region first

undergoes compression and expansion processes for some distance downstream along the surface

of the tube; a closer examination of the schlieren images reveals these features. This air then

remains attached to the tube as turbulent wall jets due to the Coanda effect (section 2.5.1), and

creates a wake behind the tube.

Images of the interaction with the large tube (25.4 mm OD, de/D = 0.29, Appendix C)

showed similar phenomena, and the offset at which secondary jets form was found to be even

greater (0.59R). Thus, for tubes larger than the jet, secondary jets do not form at small offsets; the

offset at which they first appear increases with the tube size, and as will be shown in section 5.3,

also depends on the nozzle-tube distance.

5.2.1. Secondary jet angle versus offset

As described in chapter 3 section 3.5, schlieren images of the jet/tube interaction were used to

measure the secondary jet angle, (defined in Figure 5.2) at each offset, . Figure 5.3 shows the

variation of versus for the three tubes considered in this study, for a nozzle-tube distance of 50

83

mm (6.8de). The offset is normalized by the tube radius so that at the same non-dimensional

offset, the slope of the tube surface at the jet impingement point is independent of tube size.

Figure 5.3 shows that for all three tubes, the secondary jet angle decreases almost linearly

with offset. For the smallest tube, the maximum angle occurs at zero offset, whereas for the larger

tubes, secondary jets only appear at non-zero offsets; as already mentioned, the offset at which

they appear increases with tube size. At small offsets, the angles for the medium tube are larger

than those for the small tube because of surface curvature (curvature = 1/radius); the medium tube

surface turns away from the flow less than the small tube surface. One can also see that the offset

at which secondary jets stop forming decreases with tube size; this is, at least in part, because of

the normalization of the offset by tube radius.

0

10

20

30

40

50

60

70

0 0.5 1 1.5 2 2.5

Offset, /R

Se

co

nd

ary

je

t a

ng

le,

[d

eg

]

smallest tube

biggest tube

intermediate tube

0

10

20

30

40

50

60

70

0 0.5 1 1.5 2 2.5

Offset, /R

Se

co

nd

ary

je

t a

ng

le,

[d

eg

]

smallest tube

biggest tube

intermediate tube

Figure 5.3. Secondary jet angle versus offset.

84

5.3 Effect of Tube Size and Distance between Nozzle and Tube

The effects of tube size and distance between nozzle and tube on the formation of secondary jets

were found to be coupled. As a result, they are discussed together in this section. The three tubes

were placed at different distances downstream of the nozzle and the jet was directed at them at

zero offset (head-on impingement). Figure 5.4 shows the jet/tube interaction for these cases; for

each tube size, the images are arranged vertically in sequence, where the nozzle-tube distance

increases from 1de in the topmost image to 38de in the bottom most image. For the larger tubes

(de/D = 0.39 and 0.29), fewer images are presented to avoid repetition, because the flow field for

the larger tubes at distances far from the nozzle was not found to change significantly.

The images showed several interesting flow phenomena during jet/tube interaction -

formation of secondary jets or their failure to form (flow attached to the tube), alternating rise and

fall of the secondary jet angle at distances close to the nozzle, and unsteadiness of the flow

around the tube at distances far from the nozzle. The effects of tube size and nozzle-tube distance

on jet/tube interaction are described in the light of these phenomena.

5.3.1 Formation of secondary jets and their failure to form

For the small tube, secondary jets not only form when the tube is located close to the nozzle, but

also when it is located far from the nozzle. Close to the nozzle (images a-d in Figure 5.4a), the

secondary jets propagate at a large angle, whereas farther away (images e-g in Figure 5.4a), the

angle decreases, and they become weaker, as indicated in the schlieren images by their smaller

length. This is because the oncoming jet is weaker (in terms of its peak impact pressure and

velocity) when the tube is located beyond the core of the primary jet (region in which the jet

velocity and impact pressure are the highest, and remain unchanged, see section 2.4.1). The jet

spreads radially with distance from the nozzle, and the level of turbulence in the jet increases with

distance. As a result, the flow around the tube tends to attach to the surface due to the Coanda

85

Figure 5.4. Effect of nozzle-tube distance on jet/tube interaction for three tube sizes

(offset = 0).

86

effect [10, 81], but inertia prevents it from attaching, resulting in an unsteady flow field around

the tube. When the tube was placed very far from the nozzle (images h and i in Figure 5.4a), the

flow attached to the tube, creating a wake behind it.

With the medium tube, the effect of distance on the formation of secondary jets becomes

visible. A secondary jet forms at zero offset only when the tube is located at distances x ≤ 3de.

Beyond 3de, the flow from the impingement region first undergoes compression and expansion on

the tube, and then remains attached to the tube as turbulent wall jets, instead of separating and

forming secondary jets (as described in section 5.2). A similar flow field exists around the tube

when the tube is located at distances farther from the nozzle.

Closer examination of the jet/tube interaction at x = 3de showed that the secondary jet

formed after a much longer time after impingement, compared to that at x = 1de and 2de, as well

as at all distances for the smallest tube. In all these cases, it was found that the secondary jet

forms as the primary jet develops, that is, the secondary jet forms before the flow field reaches

quasi steady-state. At x = 3de, the flow field reaches a ‘first’ quasi steady-state in which the flow

around the tube remains attached to the tube for a considerable length of time. The flow then

separates from the tube forming a secondary jet, which maintains itself continuously thereafter,

and thus forms a ‘second’ quasi steady-state of the flow field. And, secondary jets do not form

beyond x = 3de at zero offset for this tube. This suggests that this nozzle-tube distance (between

2de and 4de) for this tube size (de/D = 0.39) is a transition region in terms of separation of flow

from the tube. Time-resolved schlieren images of this phenomenon are presented in Appendix C,

where the time required for flow to separate from a tube and form a secondary jet is compared for

the medium tube placed 2de and 3de away from the nozzle. It should be noted that secondary jets

do form for the medium tube at greater offsets, as already shown in Figures 5.1b and 5.3. As

reported in section 5.2, Figures 5.4a and 5.4b show that the secondary jet angle at zero offset is

greater for the medium tube than for the small tube because of the slightly lower surface

87

curvature. Unlike for the small and medium tubes, the flow remains attached to the large tube at

all distances.

As described above, when a supersonic primary jet impinges on a small tube, secondary

jets not only form when the tube is located near the nozzle but also when it is placed away.

However, incompressible jet flow attaches to the tube when the tube is placed much closer to the

nozzle as compared to supersonic jet flow. This is because the supersonic jet has a much higher

momentum than the incompressible jet and so is unaffected by subtle pressure gradients such as

those originating from the entrainment of the jet and causing the Coanda effect. However, such

gradients do affect incompressible flows due to which phenomena such as the Coanda effect

affect such flows.

5.3.2 Alternate rise and fall of secondary jet angle with distance

Another interesting phenomenon observed for both the small and medium tubes (images a-d in

Figure 5.4a and images a-c in Figure 5.4b) is an alternate slight rise and fall in the secondary jet

angle with distance. For example, is actually larger when x/de = 2 (image b in Figure 5.4a) than

when x/de = 1 (image a). decreases slightly at x/de = 3 (image c), and then increases again at

x/de = 4 (image d). At distances greater than 4de, this phenomenon was not observed for the small

tube. A complete explanation of this phenomenon would require a detailed study of the

impingement region, which was not the focus of this work. However, a brief explanation can be

provided in terms of the jet structure at the different axial locations.

Image a in Figure 5.4a shows that at x/de = 1, the tube is located in an expansion region

(also see Figure 4.2 in chapter 4 showing the jet centreline PIP profile). The radial PIP profile at

this location is similar to that presented in chapter 4 Figure 4.3a, where the maximum PIP occurs

not at the jet centreline but in a ring around it. At x/de = 2, the tube is located in a compression

region; here, the maximum PIP occurs at the jet centreline. As a result, it is easier for the flow

88

from the impingement region to separate from the tube earlier than that at x/de = 1, thus

increasing the secondary jet angle. The same phenomenon repeats at x/de = 3 and 4. Beyond 4de,

this alternate rise and fall could not be detected in the schlieren images. Farther away, the

secondary jet angle decreases monotonically.

5.3.3 Unsteadiness of flow around tube

When the small tube was located away from the nozzle, particularly towards the end of the jet

core, the flow around the tube was unsteady. This was identified by the continuous shifting of the

secondary jet separation point slightly upstream and downstream of a mean position on the tube

surface. One reason for this is the counteracting effects of secondary jet inertia creating

conditions favouring separation, and the Coanda effect creating conditions favouring attachment,

and was described earlier.

Another cause of this unsteadiness is the instability of the supersonic jet itself. As

described in chapter 2 section 2.4.2, over- and underexpanded jets undergo toroidal and/or

helical/flapping oscillations due to the interaction between large scale coherent structures (eddies)

and shock cells in the jet. Previous studies [83, 93, 96] have shown that flapping is the preferred

mode of oscillation at high jet Mach numbers (typically greater than 2), characteristic of the jet

used in the present work. A high-speed movie of the jet taken at 20,000 frames/s revealed jet

flapping. This flapping further from the nozzle caused the flow field around the tube to be

unsteady. This phenomenon was observed for all tube sizes, though it was not as remarkable as in

the case of the small tube, because the flow was attached to the larger tubes.

As is evident from the results presented above, varying the offset strongly influenced the

interaction between the jet and a single tube. The flow of secondary jets changed significantly

with offset. Practical implications of the formation of secondary jets will be discussed in the next

chapter, where it will be shown that secondary jets assume practical importance only when the

tube-spacing is small; they do not have a significant impact when the tube spacing is large.

89

A note about the implications of the present results in light of actual flow scenarios and

conditions inside a recovery boiler deserves mention. Inside an operating recovery boiler, jet/tube

interaction takes place between a particle-laden jet and fouled tubes with irregular cross-sections

formed by deposits, whereas in the present work, the interaction takes place between a jet without

particles and clean round tubes. The present results (particularly Figure 5.4) imply that even if a

tube is covered with a layer of deposit making the jet impingement surface very rough, secondary

jets will still form if the tube is of the same size as the jet or smaller; the roughness will most

likely increase the turbulence level in the secondary jets. The breakup of a thin asymmetric

deposit impinged by a supersonic jet is presented and discussed in chapter 7; the results there

provide evidence of this behaviour. On the other hand, if the tube is covered with a deposit bigger

than the jet, secondary jets will not form and the flow will remain attached to the deposit. The

rougher the deposit surface is, the flow will remain attached to the deposit for a greater length

along the deposit surface from its windward side.

In a recovery boiler, the offset between a sootblower jet and a row of tubes changes

continuously, because of sootblower translation. The distance between the first tube of a row of

tubes and a sootblower is fixed, because the sootblower is inserted through a small opening in the

boiler wall. Consequently, this distance cannot be varied in the boiler. Similarly, all tubes in a

given section of a boiler are of the same size. As a result, though the effects of all these

parameters were studied to obtain a general understanding of the interaction and have been

presented in this chapter, only the offset was varied in most of the experiments carried out in this

work using model tube arrangements and reported subsequently in this thesis. The distance

between the nozzle exit and the first tube of any tube arrangement was kept fixed at 50 mm (6.8

nozzle diameters), which at ¼ scale corresponds to the distance inside a boiler. The tube size for

each model tube arrangement was also scaled-down appropriately, and used for all the tubes in

that arrangement, but the size varied slightly between the different arrangements.

CHAPTER 6

INTERACTION BETWEEN A JET AND TUBE

ARRANGEMENTS

This work examined the interaction between

a supersonic jet and models of typical

superheater, generating bank, and

economizer sections of a recovery boiler.

Using the conventional terminology of boiler

tube arrangements illustrated in Figure 6.1,

the tube arrangements in these sections can

be classified into three types – (1) un-finned tubes arranged in platens with zero front-to-back

spacing and large side-spacing (superheater), (2) un-finned tubes arranged in an array with small

front-to-back spacing and small side spacing (un-finned tube generating bank), and (3) finned

tubes arranged in an array with zero or very small front-to-back spacing and small side spacing

(finned tube generating bank and economizer). The interaction of a supersonic jet with these three

tube arrangements was visualized using the schlieren technique, and in some cases quantified by

peak impact pressure measurements. This chapter presents the results of these experiments.

front-to-back spacing

side spacing

fin

flue gas

front-to-back spacing

side spacing

fin

flue gas

Figure 6.1. Typical layout of boiler tubes.

90

91

Section 6.1 describes the interaction of the jet with model superheater platens and section

6.2 the interaction with model generating bank tubes. In the previous chapter, the formation of

secondary jets was described. Experiments with the model generating bank showed that these

secondary jets impinge on tubes in the adjacent rows due to the small side spacing; the results

were found to be consistent with industry corrosion experience. As a result, the structure and

strength of these secondary jets was also studied. These results are presented in section 6.3. The

interaction of the jet with a model finned-tube economizer is described in section 6.4. In each

section, the model of the tube bundle is described first along with any experimental details,

followed by the results. Finally, section 6.5 discusses the practical implications of the

experimental findings.

6.1 Interaction with Model Superheater Platens1

6.1.1 Model superheater platens

Two ¼ scale model superheater platens were constructed and mounted on supporting stands, as

shown in Figure 6.2. Each platen consisted of five 12.7 mm (0.5”) OD steel tubes welded together

in a straight line (actual superheater tubes are typically 2”). The tubes were sufficiently long to

eliminate any end effects during jet/tube interaction (l/de = ~ 23 and l/D > 13, where l is the tube

length, de is the nozzle exit diameter, and D is the tube outer diameter). The stands were

adjustable so that the platen could be positioned at different offsets relative to the nozzle.

It should be noted that the platens were restricted from vibrating and swinging in the

present experiments, though they may swing freely inside a boiler under the action of an external

force, as mentioned in chapter 1 section 1.2. This was done to reduce the number of parameters

affecting jet/tube interaction, and to identify the effects of the main parameters clearly. This is

justified based on the reasons also described in section 1.2, that (1) the effect of platen swinging

1 Some portion of this section is based on article [67].

92

platen(12.7mm OD tubes)

stand

nozzle

Figure 6.2. Model superheater platens.

on the jet depends on the jet speed relative to the platen swinging speed, and so the jet will not be

affected frequently, and (2) the effect will be significant only on the sootblower jet at the lowest

elevation.

6.1.2 Effect of offset

To study the flow of a sootblower jet over a superheater platen, images of a jet impinging on a

platen at different offsets were recorded; these are presented in Figure 6.3. The offset was

incremented by 2 mm every time. As described in the previous chapter, secondary jets form when

the primary jet impinges on the first tube of the platen, up to an offset of 0.95R (images a-d).

Beyond that, only the primary jet remains, and interacts with all of the tubes of the platen,

forming a complicated sequence of shock and expansion waves visible in the schlieren images

(images e-g). The interaction in image g is the weakest; for even larger offsets, the jet ceases to

interact with the tubes (images h and i).

These images may be corroborated with the PIP measurements of Kaliazine et al [39].

Since they used a nozzle and a supply pressure different from the ones used in this work, the

results can only be evaluated qualitatively. They measured the PIP exerted by a supersonic jet

jet

platen(12.7mm OD tubes)

stand

nozzle

jet

93

Figure 6.3. Jet impingement on a platen at different offsets.

near the surface of a model superheater platen at different offsets. The nozzle was positioned 50

mm from the front of the platen, at various offsets. A pitot probe was placed near the platen

surface to measure the jet PIP at three distances, x = 72, 103, and 151 mm from the nozzle exit

(22, 53, and 101 mm from the leading surface of the first tube of the platen). Figure 6.4 shows the

measurement setup and the PIP profiles from [39]. The results show that the PIP exerted by a

b

0.32

c

0.63

d

0.95

e

1.26

f

1.58

g

1.90

h

2.21

i

2.68

/R = 0

a

b

0.32

c

0.63

d

0.95

e

1.26

/R = 0

a f

1.58

g

1.90

h

2.21

i

2.68

94

All dimensions in mm

50

72

103

151

pitot probe nozzleplaten

(a) (b) (c)

All dimensions in mm

50

72

103

151

pitot probe nozzleplaten

Figure 6.4. Jet PIP exerted near the surface of a model superheater platen as a function

of offset and distance; probe at (a) 72 mm; (b) 103 mm; (c) 151 mm from nozzle (po is

the same for all cases) [39].

sootblower jet near the platen surface (where deposits accumulate) is the strongest when the

nozzle is slightly offset relative to the platen [98].

The flow illustrated in Figure 6.3 is in many cases consistent with the PIP measurements

of Figure 6.4. For instance, at /R = 1, the PIP near the platen surface is negative at 72 mm from

the nozzle (Figure 6.4a); zero at 103 mm (Figure 6.4b); and slightly positive at 151 mm (Figure

6.4c). Now consider image d in Figure 6.3, for which /R = 0.95 (close to 1). x = 72 mm

corresponds approximately to the windward side (side directly facing the jet) of the second tube

of the platen. The probe tip was located in the region between the platen and the edge of the

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-0.2 0 0.2 0.4 0.6

PIP/po

Off

se

t,

/R

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-0.2 0 0.2 0.4 0.6

PIP/po

Off

se

t,

/R

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-0.2 0 0.2 0.4 0.6

PIP/po

Off

se

t,

/R

x = 72 mm x = 103 mm x = 151 mm

vacuum

All dimensions in mm

50

72

103

151

pitot probe nozzleplaten

(a) (b) (c)

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-0.2 0 0.2 0.4 0.6

PIP/po

Off

se

t,

/R

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-0.2 0 0.2 0.4 0.6

PIP/po

Off

se

t,

/R

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-0.2 0 0.2 0.4 0.6

PIP/po

Off

se

t,

/R

x = 72 mm x = 103 mm x = 151 mm

vacuum

(a) (b) (c)

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-0.2 0 0.2 0.4 0.6

PIP/po

Off

se

t,

/R

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-0.2 0 0.2 0.4 0.6

PIP/po

Off

se

t,

/R

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

-0.2 0 0.2 0.4 0.6

PIP/po

Off

se

t,

/R

x = 72 mm x = 103 mm x = 151 mm

vacuum

95

separated secondary jet, which will be a suction region with low pressure (vacuum), and so a

negative PIP was measured at this location. At 103 mm, which corresponds to the windward side

of the fifth tube, the secondary jet has spread further, and so the edge of the jet reaches the probe

tip, corresponding to a near-zero PIP measurement. Finally, even though the schlieren images do

not show the flow at 151 mm (corresponding to the windward side of the ninth tube), the

secondary jet has spread even further and now covers the probe tip, and so a small positive PIP

was measured at this location.

Figure 6.3 shows another interesting phenomenon: once the jet is offset only a small

distance from the platen, interaction between the jet and the platen ceases. Images g and h show

that beyond an offset of only about 12 mm, there is no jet/platen interaction. This is because of

the very low spreading rate of a supersonic jet [20, 63]; the jet diffuses very little in the core

region.

6.1.3 Jet midway between platens

Figure 6.5a shows the flow of a jet exactly between two platens and reinforces the point made in

the above discussion, that there is no interaction between the jet and the platens, or deposits on

the platens when sootblowing between platens (unless the deposits are so big that they block the

flue gas passage); the jet propagates undisturbed between them. Noticeable interaction takes place

only when the jet actually ‘touches’ a platen (Figure 6.5b; similar to image g in Figure 6.3).

This can also be shown by comparing the typical spacing between superheater platens to

the radial spread of a jet. Figure 6.6 shows the measured jet radius as a function of axial distance

from the nozzle (also see Figure 4.4b in chapter 4). Half of the typical inter-platen spacing is also

shown in Figure 6.6. The figure shows that there will be no interaction between the jet and the

platens.

From the point of view of experimentation, the above results imply that building scaled-

down superheater platens did not provide any striking new information compared to that obtained

96

76 mm 76 mm (3(3””))

Figure 6.5. (a) Jet midway between two platens – no interaction; (b) jet touching one

platen – interaction can be seen (flow is from right to left in both cases).

Figure 6.6. Jet midway between superheater platens.

0

1

2

3

4

5

6

0 5 10 15 20 25x/de

r/de

(nozzle exit)

6.8de

(50 mm)

jet radius

half inter-platen spacing

centreline of jet and passage between platens

superheater tubes

r0

1

2

3

4

5

6

0 5 10 15 20 25x/de

r/de

(nozzle exit)

6.8de

(50 mm)

jet radius

half inter-platen spacing

centreline of jet and passage between platens

superheater tubes

r

(a) (b)

interaction between jet and tube

76 mm 76 mm (3(3””))

interaction between jet and tube

(a) (b)

97

from the experiments with a single tube. As a result, any further experiments on sootblower jet

flow in the superheater section should take this into account. Such experiments could possibly use

only one tube, saving considerable time and cost. Building scaled-down platens for experiments

would be useful only if, for example, the jet structure and strength along the platen surface is to

be determined as a function of the axial distance from the nozzle, or if the effect of platen

swinging on jet structure and strength is to be determined.

6.2 Interaction with a Model Generating Bank

A typical recovery boiler generating bank consists of tubes with an outer diameter of 64 mm

(2.5”) and an inter-tube spacing of 51 mm (2”) [2]. This spacing is much smaller than that

between typical superheater platens, which is 254-305 mm (10”-12”), and is comparable to the

exit diameter of a typical sootblower nozzle (slightly greater than 1”). As a result, the main

objective of these experiments was to examine the effect of the smaller spacing on jet/tube

interaction.

6.2.1 Model generating bank

A ¼ scale model of a generating bank was designed and built, consisting of 40 aluminium tubes

of outer diameter 14.3 mm (9/16”) (Figure 6.7). The tubes were arranged in a 4x10 inline array,

in which the 10 tubes were positioned in the direction parallel to the direction of propagation of

the jet. The inter-tube spacing (surface-to-surface) was 12.7 mm (1/2”) in each direction of the

array. Specifically 10 tubes were included because this corresponds to the length of a jet, that is,

the distance downstream of the nozzle exit at which the peak impact pressure of the jet decreases

to zero; this was determined from the free jet centreline PIP profile of Figure 4.2 in chapter 4.

To allow optical access for the schlieren system, the tubes were rigidly fixed between two

specially designed and fabricated quartz plates mounted to steel frames. These plates were

98

quartz window

Al tubes(4x10)

Figure 6.7. Model generating bank.

fabricated by Lasalle Scientific Inc. in Guelph, Ontario, Canada. Despite the challenges

associated with fabrication, quartz was selected as the plate material over plastic, because tests

showed that drilling holes in plastic to fix the tubes overheated the plastic around the holes, and

so changed the local density of the plastic. In the schlieren field-of-view, this region appeared as a

dark ring around the tube, thus preventing flow visualization in this region. The tubes were

sufficiently long (406 mm (16”), or l/de = ~ 55 and l/D = 28.4) to eliminate edge effects.

The distance between the nozzle and the surface of the first tube of the bank was set at 50

mm. The nozzle was fixed on an adjustable stand, to yield different offsets between the nozzle

and the tube. The pitot probe described in section 3.3 was used to measure the jet impact pressure

between the tubes, and to quantify jet/tube interaction. The schlieren mirrors and the tube bank

were moved relative to each other to obtain images near and far from the nozzle. This allowed

visualization of the jet flow inside the generating bank away from the nozzle.

steel frame

473

330

Dimensions in mm

quartz window

Al tubes(4x10)

steel frame

473

330

Dimensions in mm

99

6.2.2 Effect of offset

In these experiments, the offset was varied by 2 mm increments. At zero offset, the jet impinged

directly on the first tube of the tube bank (head-on impingement), whereas at maximum offset

(/R = 1.9 or = 13.5 mm), the jet propagated midway between two tubes. Figure 6.8 shows

images of the jet impinging on a model generating bank tube at different offsets. Image a shows

the flow field at zero offset, whereas image h is at maximum offset.

Images f-h show that at large offsets, only the primary jet exists and interacts with the

tubes; secondary jets form only at small offsets (images a-e) as expected. Figure 6.8 shows an

interesting and important phenomenon - because the tubes are close to each other, the secondary

jets that form during impingement of the primary jet, in turn impinge on the neighbouring tubes

in the adjacent rows. The impinged tube depends on the offset. In image a, two secondary jets

impinge on the sides of the first tubes in the adjacent rows. In images b and c, a single stronger

secondary jet flows between the first two tubes of the adjacent row. In images d and e, a still

stronger secondary jet impinges on the second and third tubes in the adjacent row respectively. In

these images, even the supersonic portion of the secondary jet, that is, its core region impinges on

the adjacent tubes.

This impingement has two potential consequences. First, these jets may exert a large

impact pressure on the adjacent tubes, or on deposits clinging to those tubes, and may even break

and remove these deposits. On the other hand, secondary jet impingement could lead to tube

erosion or corrosion caused by pieces of deposit entrained by the secondary jet. In fact, there is

evidence of industrial corrosion which supports this hypothesis [89]. This phenomenon is further

described in section 6.5 on the practical implications of this work. Therefore, understanding

secondary jets is important. Consequently, the structure and strength of the secondary jets were

also studied in this work, and are presented in section 6.3.

100

Figure 6.8. Jet flow into a model generating bank, at different offsets.

hh

1.901.90

gg

1.611.61

ff

1.331.33

ee

1.051.05

aa

/R = 0/R = 0

ee

1.051.05

aa

/R = 0/R = 0

bb

0.210.21

bb

0.210.21

ff

1.331.33

cc

0.500.50

cc

0.500.50

gg

1.611.61

dd

0.770.77

hh

1.901.90

dd

0.770.77

101

6th 5th column7th8th

Figure 6.9. Flow field inside a model generating bank, away from the nozzle, for jet

impingement at an offset (/R = 1.05, image e in Figure 6.8); the nozzle is to the right

hand side.

Examination of the flow field further downstream in the tube bank shows that there is no

jet flow in this region when the offset is small, because the jet is fully consumed upstream in the

form of a secondary jet. Figure 6.9 illustrates a jet impacting the topmost tube in column five;

little flow can be observed downstream of that tube. This flow is of the secondary jet that forms

due to the impingement of the primary jet on the first tube. Only when the offset is large (1.3R

onwards in Figure 6.8) can the jet flow between two rows of tubes (without forming secondary

jets), and thus penetrate further inside the tube bank. Maximum penetration of a jet will only

occur when the jet is exactly midway between the tubes. However, image h of Figure 6.8 shows

that even in this position, there is some interaction between the jet and the tubes, which may

affect jet strength. Consequently, this particular flow scenario was studied in detail.

flow from secondary jet

no flow in this region

1st row

2nd row

3rd row

4th row

6th 5th column7th8th

flow from secondary jet

no flow in this region

1st row

2nd row

3rd row

4th row

102

6.2.3 Jet midway between two rows of tubes

For this scenario, the nozzle was fixed such that the jet was midway between two rows of tubes.

The pitot probe was used to measure the impact pressure at the jet centreline and along the edge

of one of the tube rows. The measurements were repeated twice and were found reproducible.

Figure 6.10 shows these impact pressure profiles. The centreline impact pressure profile of a free

jet is shown for comparison, along with the approximate position of the tubes. The following

observations can be drawn from this plot.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50x/de

PIP/po

Figure 6.10. Peak impact pressure profiles of a jet midway between model generating

bank tubes.

free jet centreline PIP

jet between tubes – centreline PIP

decrease in centreline PIP from second tube onwards

jet between tubes – PIP along edge of lower row

tubes

cross-over of PIP

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50x/de

PIP/po

free jet centreline PIP

jet between tubes – centreline PIP

decrease in centreline PIP from second tube onwards

jet between tubes – PIP along edge of lower row

tubes

cross-over of PIP

103

Centreline peak impact pressure. The centreline impact pressure of the jet between the tubes

remains the same as that of the free jet until the second tube (x/de = 12.4); beyond the second

tube, PIP decreases relative to that of the free jet. The decrease is seen in both the supersonic and

subsonic portions of the jet; in the supersonic portion, the shock cells are weaker than those in the

free jet, and in the subsonic portion, the rate of decrease of the impact pressure is higher than that

in the free jet. This decrease can be attributed to a number of causes as explained next.

First, the interaction of the jet boundary or shear layer with the tube surfaces leads to the

formation of boundary layers on the surfaces (a 'no-slip' boundary condition exists on the

surfaces, that is, the fluid velocity is zero at the surfaces). Boundary layers are dissipative regions

of the flow, in which the kinetic energy of the flow is dissipated as heat through the viscous

motion of the fluid layers and vorticity. The dissipation of energy leads to a decrease in the axial

momentum of the jet, which manifests itself as a decrease in the jet centreline PIP.

To further investigate the cause of this decrease, a schlieren movie of the jet flow

between the tubes was carefully studied. Figure 6.11 illustrates the flow field. The dashed arrows

in the figure indicate the local direction of flow. PIP measurements along the centreline of a free

jet and along the centreline of a jet midway between tubes are also shown to corroborate the

results. The figure and PIP/po axis are approximately to scale. It should be kept in mind that the

PIP measurements shown in this figure were obtained using a pitot probe, and are values of the

impact pressure after the effect of a normal shock wave at the tip of the probe, even though the

probe is not present in the flow field shown in the figure.

At steady-state, inclined wave-like structures form on the tubes in the first two columns,

as indicated in the figure. From Figure 6.11, these waves form at an angle between 115°-120°

from the windward side of the tubes (side directly facing the jet), that is, they form on that portion

of the tube surface which turns away from the flow direction of the jet. Supersonic flow adjusts to

such a situation by turning in the same direction and expanding simultaneously via an expansion

wave. The impact pressure ahead and behind of one such wave was also measured. It was found

104

Figure 6.11. Flow of a jet midway between model generating bank tubes; the dashed

arrows indicate the local flow direction.

that for the wave that formed on the first tube in the third row, the upstream peak impact pressure

was 441 kPa gauge (64 psig), while the downstream pressure was 303 kPa gauge (44 psig). As

described in section 2.2.2, the peak impact pressure always drops across an expansion wave and

rises across an oblique shock wave [28], and so the observed wave structures are expansion

waves.

As air passes through the expansion waves on the first tubes of the rows, a small amount

of that air deviates outwards from the axial flow direction and impinges on the tube immediately

downstream (i.e. on the second tube). This weaker flow remains attached to the tube via the

x/de

PIP/po

PIP/po Free jet centreline

Jet centreline midwaybetween tubes

6.87.88.79.510.511.412.413.214.115.115.917.3 3.0

primary jet

expansion wave

Coanda-induced re-entrainment

1st row

2nd

3rd

4th

1st column2nd3rd

6.87.88.79.510.511.412.413.214.115.115.917.3 3.0

0.42

0.420.570.450.480.540.420.520.400.450.390.390.350.33

0.590.440.470.560.430.510.370.410.350.340.260.20

x/de

PIP/po

PIP/po Free jet centreline

Jet centreline midwaybetween tubes

6.87.88.79.510.511.412.413.214.115.115.917.3 3.0

primary jet

expansion wave

Coanda-induced re-entrainment

1st row

2nd

3rd

4th

1st column2nd3rd

primary jet

expansion wave

Coanda-induced re-entrainment

1st row

2nd

3rd

4th

1st column2nd3rd

6.87.88.79.510.511.412.413.214.115.115.917.3 3.0

0.42

0.420.570.450.480.540.420.520.400.450.390.390.350.33

0.590.440.470.560.430.510.370.410.350.340.260.20

105

Coanda effect and is re-entrained into the jet. This process repeats itself but intensifies further

downstream because the jet spreads radially and its interaction with the tubes becomes stronger,

and then this process weakens far away from the nozzle because the overall flow slows down

substantially (flow velocity decreases and turbulence increases). Since the second tube lies in the

supersonic portion of the jet, the decrease in PIP observed in Figure 6.10 starts in the supersonic

portion of the jet, and weakens the shock cells; Figure 6.11 shows that noticeable decrease in PIP

starts from about x/de = 12.4, which corresponds to a location just downstream of the second tube.

This Coanda-induced re-entrainment is believed to increase the mixing in the jet from the second

tube onwards, to a level more than that in the free jet. Due to this, the kinetic energy of the jet is

viscously dissipated through increased vorticity leading to a decrease in the centreline PIP.

Moreover, the jet spreads laterally due to confinement by the tubes, and its momentum is

redistributed over a greater cross-sectional area. Due to all these reasons, the peak impact

pressure decreases more rapidly than in the free jet. However, the spreading of the jet between the

tubes is limited by the presence of the tubes.

Although high momentum flows such as supersonic jets are not significantly affected by

subtle pressure gradient-driven effects such as the Coanda effect, the flow situation described

above is believed to involve the Coanda effect because in this situation too, this effect apparently

affects the weak air flow originating from the jet shear layer and not the jet core. The jet shear

layer is the main cause of dissipation of the jet's energy. The vorticity and turbulence in the shear

layer propagate radially inwards towards the jet centreline as the distance away from the nozzle

exit increases. Coanda-induced re-entrainment affects the jet shear layer and hence the main

dissipation mechanism of the jet.

As was done for the superheater platens, Figure 6.12 illustrates the radial spread of the jet

superimposed on a schematic of the typical spacing between generating bank tubes. The figure

shows that for a jet directed midway between tube rows, the jet boundary just touches the first

tube of the row, and is interrupted by the second tube. As a result, jet/tube interaction is expected

106

0

1

2

3

4

5

6

0 5 10 15 20 25x/de

r/de

superheater tubes

Figure 6.12. Jet midway between generating bank tubes.

to begin from the first tube and become stronger from the second tube onwards. The interaction

with the first tube creates the expansion waves, and the interaction with the second tube leads to

Coanda-induced re-entrainment, as described above.

Peak impact pressure along edge of row. For the jet midway between the tubes and for

distances close to the nozzle, Figure 6.10 shows that the impact pressure along the edge of a row

is lower than that along the centreline. This is expected, because along the edge, the probe

averages the PIP mainly across the shear layer of the jet and over some internal portion; the

velocity of the jet in these regions is lower than that closer to the centreline.

The impact pressure along the edge fluctuates until the end of the potential core, although

the fluctuations are not periodic like those along the centreline. This may be because of one or

more of the following possible reasons. First, the core shrinks with axial distance from the nozzle;

as a result, the region near the tubes consists of the interaction between the shock cells and

nozzle exit

6.8de

(50 mm)

jet radius

half inter-platen spacing

centreline of jet and passage between platens and rows of tubes

half generating bank tube spacing

generating bank tubes

0

1

2

3

4

5

6

0 5 10 15 20 25x/de

r/de

superheater tubes

nozzle exit

6.8de

(50 mm)

jet radius

half inter-platen spacing

centreline of jet and passage between platens and rows of tubes

half generating bank tube spacing

generating bank tubes

(nozzle exit)

r0

1

2

3

4

5

6

0 5 10 15 20 25x/de

r/de

superheater tubes

nozzle exit

6.8de

(50 mm)

jet radius

half inter-platen spacing

centreline of jet and passage between platens and rows of tubes

half generating bank tube spacing

generating bank tubes

0

1

2

3

4

5

6

0 5 10 15 20 25x/de

r/de

superheater tubes6.8de

(50 mm)

jet radius

half inter-platen spacing

nozzle exit

generating bank tubes

centreline of jet and passage between platens and rows of tubes

half generating bank tube spacing

r

(nozzle exit)

107

turbulence, which makes the flow unsteady in this region. Second, the air re-entrained in the jet

downstream of any given tube may also disturb the flow locally in that region, making it

unsteady. This may be further amplified by the interaction with the shock cells. Third, the

instability of the jet itself may contribute to this. As described in section 5.3, the jet used in this

work is known to undergo a flapping instability. As a result, the interaction of the jet with the

tubes could have been unsteady, and this would be reflected in the PIP measurements. Finally, the

size of the pitot probe orifice relative to the size of the jet and tube spacing may have also

contributed to this, by averaging the PIP over some small portion of the cross-section of the jet.

However, the contribution from this source is expected to be smaller than from the other sources,

because the first two oscillations along the edge between the first and second tubes were clearly

captured by the probe.

Cross-over of peak impact pressure profiles. Beyond the jet core (approximately 18 nozzle

diameters), in the region where the jet turns subsonic from supersonic, Figure 6.10 shows that the

PIP profile along the edge of the row of tubes crosses over the centreline profile; its PIP beyond

this distance is higher than that at the jet centreline between the tubes, and its rate of decrease is

also slower. At a larger distance, the profile along the edge also crosses over the centreline profile

of a free jet. This is considered to happen because the tubes restrict the entrainment and spreading

of the jet. Some amount of the air exiting the jet stream around the tubes is entrained back into the

jet as described above, whereas that air would mix with the ambient air in the case of a free jet.

The air below the stagnation point on the upper tube (or in the case of the lower tube, above the

stagnation point) accelerates from the stagnation zone around the lower (or upper) surface of the

tube, increasing its velocity. The jet also spreads laterally due to the confinement, redistributing

its momentum. Due to these reasons, the PIP along the edge increases. Because a free jet spreads

unrestricted with distance, the centreline PIP of a free jet decreases continuously.

108

Finally, Figure 6.13 shows the flow in the middle of the tube bank between columns 5

and 8 (further from the nozzle than in Figure 6.11). In this location, the flow is no longer

supersonic, and hence no compression/expansion waves are seen in this region.

7th8th 6th 5th column

1st row

2nd

3rd

4th

7th8th 6th 5th column

1st row

2nd

3rd

4th

Figure 6.13. Flow midway between two rows of tubes farther from the nozzle.

6.3 Secondary Jets

Results presented in the previous sections have shown secondary jets to be unimportant in the

superheater section, but that they may contribute to both deposit removal and tube corrosion in

the generating bank section. Therefore, understanding secondary jets is important. Knowledge of,

for example, the centerline PIP variation of these jets may yield opportunities to optimize boiler

tube spacing in relation to sootblowing. This motivated a study of the structure and strength of

these secondary jets.

109

Secondary jets identified in this work form as a result of the interaction between a round

supersonic jet and a cylindrical tube. The structure of these jets is more complicated than that of

the primary jet itself, with three-dimensional effects having a greater influence on the structure of

the secondary jets. Experiments with a single tube described in chapter five provided images of

secondary jets in the plane through the cross-section of a tube. Visualising these jets in another

orthogonal plane will improve our understanding of their structure. The PIP of these jets was also

measured as an indicator of jet strength. The variation of PIP with distance along the secondary

jet centreline will indicate the extent to which these jets maintain a high level of PIP. The results

of this study are described in this section.

6.3.1 Experimental apparatus and procedure

In the experiments described in chapter 5, the tube was

oriented horizontally in front of the nozzle, with the

longitudinal axis aligned in the direction of the parallel

schlieren light rays; this allowed visualization of the

secondary jets in the plane of the tube cross-section. To

visualize the secondary jets in an orthogonal plane, the

tube was oriented vertically in front of the nozzle at

every offset, and the apparatus was re-arranged using

the measured secondary jet angle corresponding to that

offset (Figure 5.3 in section 5.2.1 in chapter 5). A

special experimental module was designed and

constructed for this (Figure 6.14), which allowed the

offset between the nozzle and tube to be varied, so that

secondary jets at different offsets could be studied. The

nozzle

Figure 6.14. Experimental module

with tube oriented vertically in

front of the nozzle for visualizing

secondary jets and measuring

their centreline PIP.

offset relative to nozzle

nozzle

offset relative to nozzle

110

same module was used to measure the centreline PIP of the secondary jets.

As in the previous experiments, the tube diameter was 13 mm, and the distance between

the nozzle and tube was fixed at 50 mm. For each experiment, the offset between the nozzle and

tube was changed, and the jet was directed at the tube; the resulting secondary jet was visualized

using the schlieren technique. The centreline PIP of each secondary jet was measured twice and

the results were found reproducible.

The variation of secondary jet angle with offset has already been described in chapter

five, section 5.2.1. Images of secondary jets in the orthogonal plane, and the centreline PIP

measurements, are presented next.

6.3.2 Secondary jet structure

Figure 6.15 shows the secondary jet that forms at an offset of 0.75R for a 13 mm (0.5”) OD tube

(de/D = 0.58). Figure 6.15a shows the secondary jet in the plane passing through the tube cross-

section; Figure 6.15b shows the secondary jet in the plane that passes through the tube axis; in

this image, the jet is flowing behind the tube. The nozzle is to the right of the tube in both images.

Figure 6.15a shows that the secondary jet has a defined width that increases with distance

(as indicated by the white lines) along the secondary jet centreline. Shock cells, which are

alternating zones of compression and expansion, are visible in the secondary jet, as described in

section 5.2, indicating that the secondary jets are also supersonic. The same shock cells can be

seen in Figure 6.15b (alternating light and dark bands just downstream of the tube), where it is

now apparent that these cells are geometrically very different than those in the primary jet. In the

primary jet, the cells are formed by waves arranged conically, whereas in the secondary jet, the

waves are curved and semi-circular. The size of these cells decreases downstream due to

interaction with turbulence. These cells identify the core region of the jet.

The air in the primary jet mainly flows in the axial direction, and the jet is round, whereas

Figure 6.15b (and the associated movie) shows that the air in the secondary jet spreads out much

111

a

secondary jet

Figure 6.15. Secondary jet at 0.75R offset for a 13 mm (0.5”) OD tube (de/D = 0.58); (a)

tube horizontal in front of the nozzle; (b) tube vertical in front of the nozzle.

more, forming a sheet-like or a fan-like jet. This is indicated by the arrows on the figure. The

cross-section of the primary jet is circular, whereas that of the secondary jet is oblong and oval.

Based on the structure and behaviour of secondary jets, a subtle issue may arise here: can these

secondary flows be termed jets? Abramovich [1] carried out one of the first comprehensive

theoretical analyses of turbulent jets, and defined a jet as follows:

“In many cases of motion of a liquid or gas, so-called tangential separation surfaces

arise; the flow of fluid on either side of this surface is termed a jet. The jets may be moving in the

same direction or in opposite directions. …”

b tube

a

secondary jetprimary jet

b tubeprimary jet

112

Based on this definition, the secondary flows can be termed jets, because as seen in the

schlieren images, these flows are bounded on all sides by tangential separation surfaces or shear

layers, with faster flow inside and stagnant air outside.

Figure 6.16 (a-f) shows secondary jets at six different offsets (from zero to 1.25R) for a

13 mm (1/2”) OD tube. In each image, the secondary jet is flowing in the plane of the image,

while the primary jet is out of the plane at the secondary jet angle corresponding to that offset.

The secondary jet at zero offset (image a) is the weakest and spreads out the most; shock cells are

not visible in this jet. As the offset increases and the jet/tube interaction weakens (images b-f), the

jet spreads less, and the shock cell structure in the secondary jets becomes stronger, and begins to

appear in the schlieren images. At large offsets, the shock cells appear to have the same geometric

structure as those in the primary jet.

6.3.3 Secondary jet peak impact pressure

Figure 6.17 (a-f) shows the peak impact pressure along the centreline of the secondary jets at

different offsets. Plots a-f correspond to images a-f in Figure 6.16. For comparison, the centreline

PIP profile of a free primary jet is also shown. For each secondary jet, the PIP was measured

along its centreline beginning from the rear (or downstream) side of the tube. Measurements on

the front (or impingement) side could not be performed, because the flow in the impingement

region is very complex, and a pitot probe is not suitable for measurements in this region. To be

consistent in presenting the PIP for all secondary jets, PIP is plotted in Figure 6.17 along the axis

defined at the top of that figure. x represents the distance first along the primary jet centreline

(x/de = 0 at the nozzle exit), then through the impingement region, and finally along the

secondary jet centreline. As a result, the PIP profile until just upstream of the tube is the same as

that of the primary jet.

Figure 6.17a shows that at zero offset, the secondary jet PIP is the lowest; the PIP profile

shows no shock cell in the jet, which is consistent with the schlieren image of Figure 6.16a. As

113

offset = 0R 0.75R

0.25R 1.00R

0.50R 1.25R

a

b

c

d

e

f

offset = 0R 0.75R

0.25R 1.00R

0.50R 1.25R

a

b

c

d

e

f

Figure 6.16. Secondary jets at different offsets for a 13 mm (0.5”) OD tube (de/D = 0.58).

114

Figure 6.17. Centreline peak impact pressure of secondary jets at different offsets, for a

13 mm (0.5”) OD tube (de/D = 0.58); the primary jet peak impact pressure is shown for

comparison.

nozzle

primary jet

secondary jet

axis of PIP measurement

x=0

nozzle

primary jet

secondary jet

axis of PIP measurement

x=0

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

a d

b

c

primary jet

secondary jet

offset = 0R = 60°

0.75R = 34°

1.00R = 18°

1.25R = 8°

0.50R = 42°

0.25R = 50°

e

f

nozzle

primary jet

secondary jet

axis of PIP measurement

x=0

nozzle

primary jet

secondary jet

axis of PIP measurement

x=0

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

a d

b

c

primary jet

secondary jet

offset = 0R = 60°

0.75R = 34°

1.00R = 18°

1.25R = 8°

0.50R = 42°

0.25R = 50°

e

f

115

the offset increases, the interaction between the jet and tube weakens, and the maximum PIP in

the secondary jet increases. The shock cell structure becomes appreciable in the secondary jets

beyond an offset of 0.25R and strengthens with offset; this is indicated by the PIP oscillations.

The number of shock cells in the secondary jets increases with offset, as indicated by the PIP

measurements; the number captured in the measurements is consistent with the number that

appears in the corresponding schlieren images. At large offsets (1.00R and 1.25R, plots e and f),

the secondary jet PIP close to the tube is very high, comparable to that of the primary jet itself.

However, the PIP decreases more rapidly with distance along the centreline of the secondary jet

because the secondary jets spread much more than the primary jet, redistributing the jet

momentum over a larger area.

6.4 Interaction with Model Economizer Tubes (Finned Tubes)

The objective of these experiments was to determine the effects of a fin on jet/tube interaction. A

survey of typical recovery boiler tube arrangements showed that while the economizer sections

usually have finned tube arrangements, even the generating bank sections of modern recovery

boilers consist of finned tubes. As a result, even though the results here are presented in the

context of economizer tube arrangements, the results are more generally applicable to any finned-

tube arrangements which have fins of a similar design.

6.4.1 Model economizer section

Quantitative data about typical economizer tube arrangements in recovery boilers is not available

in the open literature. Specialized texts such as [90] were found to contain only general

information about the layout of tubes in the economizer section (such as inline versus staggered

arrangements), but no dimensions. Consequently, a survey of typical economizer tube

arrangements in recovery boilers was conducted, and information was also collected about typical

116

superheater and generating bank tube arrangements. Three boiler manufacturers - Babcock &

Wilcox, Andritz, and Metso Power, provided data on their typical tube arrangement dimensions.

Results of the survey are summarized in Appendix D, and were used to design two identical ¼

scale rows of economizer tubes with fins.

Figure 6.18a schematically shows one of these rows. It consisted of six 11 mm (7/16”)

OD tubes with fins welded on both the windward and leeward sides, with zero front-to-back

spacing. The fin width was the same as the tube outer diameter. Similar to the model superheater

platens, these tubes were sufficiently long to prevent any flow edge effects. The model was

supported on the same stands used for the model superheater platens, which allowed the offset

between them and the nozzle, as well as the spacing between the two rows, to be varied. Figure

6.18b shows the assembly. Based on the survey results, the spacing between the two rows was

fixed at 12.7 mm (0.5”), corresponding to 2” spacing in a boiler.

Figure 6.18. Model economizer tubes: (a) schematic of a row; (b) tube assembly.

6.4.2 Effect of offset

The offset was varied by 2 mm increments and the jet/tube interaction was visualized at each

offset; Figure 6.19 shows these images. Consistent with results presented earlier in this thesis,

OD 1111 1.222

All dimensions in mm

(a) (b)

standrow of finned tubes

nozzle

OD 1111 1.222

All dimensions in mm

OD 1111 1.222

All dimensions in mm

(a) (b)

standrow of finned tubes

nozzle

117

secondary jets form when the jet impinges on the first tube of a row. However, in this case the

secondary jets that form at 0 offset are strongly affected by the leading fin. Image a in the figure

shows that upon impingement on the fin, the primary jet splits into two parts (one above the fin

and one below) which deviate slightly from the original axial flow direction. Due to this, they

impinge on the first tube at a point slightly higher than the location of the fin. Upon impingement

on the tube, the jets undergo further deflection. Due to multiple interactions with the tip of the fin,

the fin surface and the tube, these secondary jets are weaker than those observed in the

superheater and generating bank.

As the offset increases from 0, the interaction of the primary jet with the fin weakens, and

a single stronger secondary jet forms which impinges on the adjacent row of tubes (images b and

c). With further increase in offset (images d and e), the secondary jet angle decreases as usual and

the secondary jet becomes stronger. Beyond /R = 1.79 (image f), only the primary jet remains,

which interacts weakly with the tubes. As the jet approaches the position midway between the

two rows of tubes (images g and h), expansion waves become visible on the tubes, similar to

those observed in the generating bank.

The effect of the fin on the direction of a secondary jet is shown in Figure 6.20. The

figure shows how the secondary jet angle varies with offset, for secondary jets that form when a

jet impinges on the first tube in the superheater, generating bank, and economizer sections.

Unlike for the superheater and generating bank, the secondary jet angle at 0 offset for the

economizer is much lower. The angle first increases, and then decreases with offset because of

the multiple reflections mentioned above.

118

a

b

c

d

e

f

g

h

/R = 0

0.36

0.71

1.07

1.43

1.79

2.14

2.32

fin

tube (11 mm OD)

stand

Figure 6.19. Jet impinging on economizer tubes at different offsets.

expansion wavesshock wave

a

b

c

d

e

f

g

h

/R = 0

0.36

0.71

1.07

1.43

1.79

2.14

2.32

fin

tube (11 mm OD)

stand

expansion wavesshock wave

119

Figure 6.20. Secondary jet angle versus offset for the superheater, generating bank and

economizer tube arrangements.

0

10

20

30

40

50

60

70

0 0.5 1 1.5 2

Offset, /R

Sec

on

dar

y je

t an

gle

, θ [

deg

]

economizer

superheater

generating bank

0

10

20

30

40

50

60

70

0 0.5 1 1.5 2

Offset, /R

Sec

on

dar

y je

t an

gle

, θ [

deg

]

economizer

superheater

generating bank

6.4.3 Jet midway between two rows of tubes

Figure 6.21 presents the PIP profiles of a jet positioned midway between two rows of economizer

tubes. The PIP along the jet centreline, as well as along the edge of the lower row of tubes, is

presented. The centreline PIP profiles of a free jet as well as of a jet midway between two rows of

generating bank tubes are also shown for comparison. The following observations can be drawn

from Figure 6.21.

Centreline peak impact pressure. Figure 6.21 shows that the centreline PIP in the supersonic

portion of a jet midway between the economizer tubes is unaffected by the presence of the tubes;

the centreline PIP in the jet core is the same as that in a free jet. However, the PIP decreases

compared to the free jet in the subsonic portion of the jet (beyond 18 nozzle diameters). This is

due to the increased level of mixing in the jet, particularly just upstream and downstream of the

tubes (which are regions of recirculation and wake respectively), and because the jet spreads in

120

Figure 6.21. Peak impact pressure profiles of a jet midway between model economizer

tubes.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

fins

the lateral direction. The finned tubes form walls confining the jet between them, and restrict the

spreading of the jet in the direction normal to the walls. The jet spreads initially until its

boundaries interact with the first tube in the upper and lower rows, forming expansion waves on

the tubes. The flow of air along the fin surfaces and around the tubes increases the mixing in the

jet. This phenomenon intensifies as the jet spreads further downstream, but weakens at large

distances from the nozzle as the overall flow slows down substantially. The jet spreads laterally to

adjust to the almost planar confinement, its momentum is redistributed over a greater cross-

sectional area, and from about the third tube onwards where the jet turns subsonic, the jet PIP

free jet – centreline PIP

jet in economizer –centreline PIP

jet in generating bank –centreline PIP

tube

jet in economizer –PIP along edge of lower row

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

fins

free jet – centreline PIP

jet in economizer –centreline PIP

jet in generating bank –centreline PIP

tube

jet in economizer –PIP along edge of lower row

121

decreases. Appendix E contains schlieren images of the quasi steady flow of a jet midway

between two rows of finned economizer tubes, showing the flow near and far fields.

Although the economizer and generating bank tubes are equally spaced (12.7 mm or 0.5”

in the present experiments), the jet centreline PIP between the economizer tubes remains stronger

for a greater distance because of the restricted entrainment and spreading of the jet. In the

generating bank, the open gaps between the tubes allow some small portion of the jet air to flow

around the tubes, forming wall jets (attached to the tubes) due to the Coanda effect, and

entraining air from the surrounding. The fins in the economizer, on the other hand, prevent

entrainment of surrounding air. Moreover, the jet reaches a tube sooner in the generating bank

than in the economizer, as the leading fin precedes the first tube in the economizer. As a result,

the jet centreline PIP diffuses more slowly in the economizer, and essentially, the jet is stronger

than in the generating bank.

Peak impact pressure along edge of row. The PIP along the edge of a row of tubes is much

lower than the centreline PIP, because at the edge, the pitot probe measures the PIP in the

outermost part of the jet. The PIP initially oscillates because of the compression and expansion

waves within the jet. In the outermost and weakest part of the jet, the measured oscillations are

much weaker than those captured along the centreline. The PIP then increases with distance up to

about 18 nozzle diameters (transition region from supersonic to subsonic) because of the

spreading of the jet. The PIP measured close to the first tube and fin is lower because the probe

only senses the outermost part of the jet. Further downstream, the jet spreads and the probe senses

a greater area of the jet with a higher PIP. From the transition region onwards (x/de = 18), the PIP

decreases continuously with distance due to mixing.

Cross-over of peak impact pressure profiles. As was observed in the PIP profiles of a jet in a

generating bank (Figure 6.10), the PIP along the edge of a row of tubes in the economizer

122

eventually exceeds the PIP along the jet centreline. Again, this occurs because the jet is restricted

between the tubes. Air accelerates locally around the tubes, and it also spreads laterally. This

causes the momentum to redistribute, increasing the PIP along the edge. The increase in PIP is

greater for the economizer than the generating bank, because the entrainment and spreading is

restricted much more by the fins in the economizer.

6.5 Practical Implications – Effects of the Formation of Secondary Jets

and Closer Tube Spacing

For all three tube bundles (superheater, generating bank, and economizer), secondary jets form

when a primary jet impinges on the first tube of a row of tubes (or on the leading fin in the case of

the economizer). A secondary jet becomes weaker and deviates more from the original flow

direction as the offset (distance between the jet and tube centrelines) decreases, that is, as the

interaction between the jet and tube intensifies. Consequently, there is little or no jet flow at all

beyond the first few tubes of these tube bundles whenever a secondary jet forms (third or fourth

tube in the superheater and generating bank sections and the second tube in the economizer

section). The primary jet impinges on the deposit formed on the first tube of the tube row in these

circumstances. Hence, any deposits beyond the first few tubes do not experience the sootblower

jet for a significant amount of time. These deposits include those clinging to the sides of

superheater platens, those accumulated between generating bank tubes as well as jammed

between them after being debonded by sootblowers, and those formed on economizer tubes and

fins.

Secondary jets practically become more important at closer tube spacing. Since the core

length of the secondary jet is much smaller than the spacing between two superheater platens, the

secondary jet cannot effectively remove a deposit on an adjacent platen. However, due to the

close spacing between the tubes in the generating bank and economizer, secondary jets impinge

123

on the tubes in the adjacent rows. In some situations, even their core region impinges on the

tubes. Thus, these jets may exert a large impact pressure on the adjacent tubes, or on deposits

clinging to those tubes, and may even break and remove these deposits. And even if the deposits

are too hard to break, the secondary jets may erode those deposits as they impinge on them. This

in turn would increase the effective exposure time of the deposits to the jet, which is necessary to

break big and hard deposits [69]. This is because as the primary jet impinges on the first tube of

the preceding row of tubes and continues to translate forward, the deposit on the adjacent tubes is

impinged by a secondary jet with continuously varying angle and strength; after a short period of

time, this same deposit is impinged by the primary jet that has translated forward. Thus, the

deposit is subjected to steam flow twice. Exact determination of whether secondary jets can

remove deposits in the generating bank and economizer will require information about the

physical structure of deposits (their shape and size) as well as their mechanical properties

(adhesion and tensile strengths, and porosity), which is difficult to obtain because of the

continuously changing nature of deposits. Another consequence of secondary jet impingement is

that it could lead to tube erosion as well as under-deposit corrosion.

To address these issues, secondary jet PIP was measured along the jet centreline. The

measurements showed that secondary jets retain a strong PIP for appreciable distances

downstream of the tube only at large offsets, although not as far as a primary jet. As the offset

decreases and the interaction between the jet and tube intensifies, the secondary jets become

weaker. Secondary jet PIP data shows that the PIP exerted by the secondary jets on the adjacent

tubes in the model generating bank (see Figure 6.8 c, d, and e) is slightly less than half the

average PIP in the core of the primary jet. This implies that in the generating bank and

economizer sections where the tube spacing is small, secondary jets will impinge on the tubes of

the adjacent rows, but will not exert a PIP on those tubes (or on deposits on those tubes) that is

more than half the average PIP in the primary jet core region. Nevertheless, these jets may help

erode deposits accumulated on and between the adjacent tubes. Deposits in the generating bank

124

are a mixture of carryover and fume deposits, and in some locations may be hard and brittle. But

those in the economizer are primarily fume, and are thin and powdery. As a result, secondary jets

although weaker because of the fin, may nevertheless help remove deposits even from adjacent

rows. In the superheater section, secondary jets will not be effective at all because of the large

inter-platen spacing.

On the other hand, there is evidence of industrial corrosion which supports the hypothesis

that secondary jets lead to under-deposit corrosion of tubes. This is described next.

Industry corrosion experience and relevance of the present findings

Over the years, North American recovery boiler operators have consistently reported thinning of

generating bank tubes, very near the mud-drum of the bank (lower drum containing water). Tube

thinning at sootblower elevations (slightly higher elevations than the mud drum) has also been

reported occasionally [89].

The most severe metal loss is found on the tubes closest to the sootblower lanes, and the

severity decreases in tube rows further away. The tube thinning is very localized, and only small

areas around the tube circumference are affected – at 10:00 and 2:00 with respect to the direction

from which the sootblower passes by the tube (0° in Figure 6.22). Studies conducted by the Pulp

and Paper Research Institute of Canada (PAPRICAN) [89] showed this thinning to be most likely

caused by a repeated cycle of under-deposit corrosion followed by removal of accumulated

corrosion product by sootblowers located close to the drum surface.

Although the present experiments were conducted for a different objective, their findings

appear consistent with this industry experience. As Figure 6.8 showed, secondary jets impinge

only on the first few tubes of the generating bank rows, that is, on the tubes closest to the

sootblower. Moreover, the secondary jets impinge on these tubes at an angle anywhere between

0° and 90° (Figure 6.22), which corresponds to the 10:00 and 2:00 positions reported in [89].

Although the present results have been obtained for head-on impingement (impingement

on a tube normal to the jet) and the impingement on the tube near the drum will be at an angle,

125

the results are still applicable, and show

that the secondary jets are the means by

which the sootblower steam reaches the

tubes behind the first tube of any given row.

Moreover, the schlieren movie of the

breakup of a synthetic deposit impinged by

a jet shows that deposit particles broken by

a jet are entrained by the secondary jet; this

will be presented in chapter 7. These

particles will also impinge on the adjacent tubes.

90°

90°

Figure 6.22. Impingement of a secondary jet

on a tube behind the first tube of a

generating bank row.

Due to the very low spreading rate of a supersonic jet, jet/tube interaction ceases when

the jet is offset only a small distance from a superheater platen. This implies deposits clinging to

the side of a platen are not exposed to the sootblower jet when it is offset only a small distance

from the surface of the deposit layer. Furthermore, much of the steam that a sootblower blows

between platens is wasted, as only small offsets yield useful interaction between a jet and a

deposit; any sootblowing strategy should take this into consideration. Continuous sootblowing

between platens is justified only if large deposits significantly block the space between platens.

Reducing the steam supply significantly or even stopping the supply when blowing between

platens will help reduce or stop such un-necessary steam consumption. However, depending on

the sootblower travel (length) and the time the blower stays in the high temperature environment

inside the boiler, steam may be required inside the blower lance for cooling the lance, and hence

for preventing excessive lance drooping and the associated large mechanical stresses and fatigue.

This points to the need for identifying ways to minimize such wasteful consumption of steam

(such as increasing the sootblower speed), and to the need for efficient online fouling monitoring

systems to identify fouled regions for targeted sootblowing. The situation is better in the

126

generating bank and economizer sections, because the tube spacing is small and the jet will

interact with the deposit more frequently than in the superheater.

However, the close spacing affects sootblower jet strength (peak impact pressure, PIP)

and penetration between generating bank and economizer tubes. A jet flows unaffected between

superheater platens. The centreline PIP of a jet between generating bank tubes decreases relative

to that of a free jet from the second tube onwards (x/de = 11.4), due to interaction with tube

surfaces and Coanda-induced re-entrainment. This phenomenon decreases the PIP in both the

supersonic and subsonic portions of the jet. For the same spacing between rows of tubes, the

centreline PIP of a jet between finned economizer tubes also decreases, but at a greater distance

from the nozzle (x/de = 18) compared to that in a generating bank. Essentially, a jet is stronger

and penetrates deeper in a tube arrangement consisting of tubes with fins, than in an arrangement

with finless tubes as the fins restrict the entrainment and spreading of the jet.

CHAPTER 7

SCHLIEREN VISUALIZATION OF

SYNTHETIC DEPOSIT BREAKUP BY JET

IMPINGEMENT

The interaction of a supersonic jet with a single tube, as well as model recovery boiler tube

arrangements, was described in previous chapters, and was visualized using the schlieren

technique. The next logical step in this work is a study of the interaction between a supersonic jet

and a deposit, using the schlieren technique. This chapter presents the results of a preliminary

investigation of this interaction.

An experiment was performed in which a synthetic deposit was formed on a tube and

impinged by a supersonic jet, and the resulting breakup of the deposit was visualized using the

schlieren technique. Whereas previous studies of deposit breakup by jet impingement [23, 24]

involved visualization of only the breakup of the deposit, this experiment involved visualization

of both the deposit and the jet. This chapter first describes the synthetic deposit used in the

experiment and the experimental procedure, and then presents schlieren images of the deposit

breakup.

127

128

7.1 Synthetic Deposit and Experimental Procedure

7.1.1 Synthetic deposit

An Entrained Flow Reactor (EFR) at the University of Toronto (mentioned in chapter 2 section

2.1.2) was used to prepare the synthetic deposit for this experiment. The EFR enables a carryover

deposit to form on a tube in a manner similar to that inside a recovery boiler, under similar

operating conditions. The deposit formed is asymmetric, with most of the material on the

windward side of the tube. This EFR has been used in several deposition and deposit removal

studies performed in the past (e.g. [43, 52, 85, 104]). Details of the design and operation of the

EFR can be found in [85].

The EFR (Figure 7.1a) is a 9 m long cylindrical reactor consisting of a natural gas burner

section at the top, and an electrically heated section along its length. Synthetic carryover particles

of known composition, prepared by mixing pure chemicals, are introduced at the top. They are

heated, and melt as they fall downward, and impact a cylindrical probe (with thermocouples)

placed at the exit, forming a carryover deposit on the probe. The probe surface temperature is

Figure 7.1. Entrained Flow Reactor (EFR) at the University of Toronto; (a) schematic;

(b) photograph (tube is located near the EFR exit); (c) carryover deposit formed on a

tube using the EFR.

129

controlled by a regulated flow of air through it [52]. Depending on operating conditions including

particle chemical composition, size and temperature, combustion gas temperature and velocity,

and probe surface temperature, different types of deposits can be formed on the probe (thin/thick,

hard/soft, etc.).

In the present experiment, synthetic carryover particles were made by mixing pure

sodium chloride (NaCl) and sodium sulphate (Na2SO4) salts in a 1:10.9 ratio by weight,

respectively. The particle chemical composition was selected such that the synthetic particles

yielded carryover particles with 10 mol% Cl, which was more than sufficient to make them sticky

and form a reasonably thick deposit on the tube. This mixture of salts was melted, cooled, ground,

and sieved into a 150-300 m range. These particles were then introduced into the EFR, and a ¼

scale tube of 13 mm (0.5”) OD was mounted near the EFR exit (Figure 7.1b). The deposit (Figure

7.1c) was approximately 2 mm thick at its thickest part, with thickness decreasing along the tube

circumference away from this point. The EFR was operated at a temperature of 800°C at

atmospheric pressure. Based on the gas flow rate, the gas velocity was approximately 1.8 m/s.

The tube surface temperature was around 500°C.

7.1.2 Experimental procedure

The main objective of this experiment was to observe the breakup of a deposit impinged by a

supersonic jet using the schlieren technique, and understand the breakup mechanism as well as

the behaviour of the broken deposit particles. For this, a deposit strength was required that could

be broken by the jet. Consequently, after the deposition was complete in the manner described

above, the deposit was subjected to a thermal shock by quickly removing the tube from the hot

exit region of the EFR, into the relatively cool ambient surroundings (20°C). Then the tube was

fixed on the same adjustable stands used in the previous schlieren experiments, so that the jet

could be directed at the deposit at zero offset. The tube was located 50 mm away from the nozzle

130

exit. The jet was then impinged head-on on the thickest part of the deposit, and the resulting

breakup was captured using the schlieren system.

7.2 Deposit Breakup Images

Figure 7.2 shows the breakup of the synthetic deposit upon jet impingement. Time t=0 is

arbitrarily assigned. In some instances, very small broken deposit particles may be difficult to

distinguish from the turbulence in the flow field. This is because the field-of-view was much

larger than those particles to enable capture of the breakup of the entire deposit and the

trajectories of the broken particles. Moreover, the light source used in the schlieren system

limited the maximum achievable illumination. It should also be noted that the images in Figure

7.2 only capture the motion of the particles in the plane of the images, whereas the particles

actually travel in all three dimensions. As a result, most of the particles captured in the images are

also travelling out of the plane of the images. Despite these issues, the following phenomena

during breakup were clearly observed in the breakup movie, and Figure 7.2 presents schlieren

images derived from the movie.

(1) Even before the jet develops fully and reaches steady state, that is, before it attains maximum

strength, the jet starts eroding the surface of the deposit. Very small particles are removed

from the surface by the flow. This may be because the deposit is thin, and was weakened by

the thermal shock.

(2) Secondary jets form even during the interaction of the primary jet with a deposit (image c

onwards), because the effective de/D does not increase much due to the deposit thickness; the

roughness of the deposit surface does not influence the formation of the secondary jets,

although it will influence their turbulence characteristics. Images d-f show that the particles

eroded from the deposit surface are entrained by the secondary jets.

(3) After a certain amount of deposit has been eroded off by the jet in the form of small particles,

131

the jet pressurizes the porous deposit in and around the impingement region to the jet peak

impact pressure at that axial location, and blasts a large amount of deposit (images f-m).

Images i-k show that the broken deposit particles fly away from the tube surface in all

directions (from the upper secondary jet, clockwise all around the front or impingement side

of the tube, to the lower secondary jet) except directly behind the tube. However, because the

maximum deposit is located on the impingement side of the tube due to the asymmetric

nature of the deposit, a large amount of deposit is seen to propagate away from the tube

‘backwards’ (in the direction opposite to that of the jet), towards the nozzle (image m). Image

m also shows that the deposit particles are larger now.

(4) A closer examination of the breakup movie shows that some of the deposit particles that fly

backwards from the tube (toward the nozzle), impact the nozzle rim and rebound; however,

only a few small particles were seen to impact the rim in this particular experiment.

The fact that the eroded deposit particles are entrained by the secondary jets (at least at

the offset in this particular experiment, point 2 above) implies that these particles will impinge on

adjacent platens and tubes inside a boiler. In the generating bank section, they may support the

cyclic process of under-deposit corrosion described in section 6.5, by abrading the accumulated

salt cake and corrosion product off of the tube, thus exposing more tube metal to corrosion.

The above results indicate that such experiments will yield useful qualitative as well as

quantitative data of the deposit breakup process, such as the time required to complete breakup of

a certain type of deposit (strength, thickness), the trajectory of broken deposit particles, and their

velocity. Hence, further work in this direction is recommended.

132

nozzle

tube deposit

stand

a

b

c

d

e

f

g

h

t=0ms

4.16

8.15

10.65

10.82

11.48

11.65

12.15

deposit particles travelling with secondary jet

Figure 7.2. Breakup of a synthetic deposit by jet impingement, visualized using the

schlieren technique (continued on the next page).

secondary jet

start of breakup of large amount of deposit

nozzle

tube deposit

stand

a

b

c

d

e

f

g

h

t=0ms

4.16

8.15

10.65

10.82

11.48

11.65

12.15

deposit particles travelling with secondary jet

secondary jet

start of breakup of large amount of deposit

133

15.14

19.14

36.77

50.59

12.31

12.65

12.81

13.98

i

j

k

l

m

n

o

p

15.14

19.14

36.77

50.59

12.31

12.65

12.81

13.98

i

j

k

l

m

n

o

p

Figure 7.2. Continued.

CHAPTER 8

FEASIBILITY OF USING INCLINED

SOOTBLOWER NOZZLES IN RECOVERY

BOILER SUPERHEATERS

As illustrated in Figure 1.4 in chapter 1, sootblower nozzles currently used in recovery boilers are

oriented normal to the sootblower lance, and the sootblower jets exert a force that tends to push

deposits against the tubes. The resulting stresses induced by normal jet impingement are unlikely

sufficient to break strong deposits. Moreover, jets from straight nozzles travel parallel to the

platens, and cannot impinge directly on the deposits clinging to the platen sides. Only secondary

jets are directed at these deposits at an angle whenever the primary jet impinges on the first tube

of an adjacent platen. However, even the strongest of the secondary jets becomes very weak when

it reaches an adjacent platen. This can be seen in Figure 8.1, which shows the centreline peak

impact pressure of the primary and secondary jets. The figure presents the same data as Figure

6.17, but in one plot.

If a sootblower nozzle were inclined at some angle , relative to the lance as shown in

Figure 8.2a, then a component of the jet force would act normal to the platen centerline and exert

a debonding force (and moment) on a deposit. An inclined nozzle would also yield a jet that

impinged more directly on deposits accumulated on the sides of platens. Its centerline PIP would

134

135

be high, the same as that of a primary jet, as shown in Figure 8.1 (the jet itself is not affected in

any significant way but its direction is simply changed by a certain amount). On the other hand,

inclined jets would penetrate less far between platens, and would also increase platen swinging.

Sootblowers with such inclined nozzles have been used successfully in utility boilers, where the

inter-platen spacing is larger than in recovery boilers [102]. Only very recently have these nozzles

been introduced in recovery boilers [101], and their performance is currently being evaluated.

Figure 8.1. Centreline peak impact pressure of primary and secondary jets; the primary

jet can be considered as exiting from an inclined nozzle, whereas the secondary jets

result from the impingement of a jet from a straight nozzle (data shown is the same as in

Figure 6.17).

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

jet from inclined nozzle (primary jet)

secondary jets resulting from impingement of jet from straight nozzle

tube

1.25R (=8°)

1.00R (18°)

offset

0.75R (34°)

0.50R (42°)

0R (60°)

0.25R (50°)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 10 20 30 40 50

x/de

PIP/po

jet from inclined nozzle (primary jet)

secondary jets resulting from impingement of jet from straight nozzle

tube

1.25R (=8°)

offset

1.00R (18°)

0.75R (34°)

0.50R (42°)

0R (60°)

0.25R (50°)

136

Fjet

Figure 8.2. (a) Sootblower jet from inclined nozzles; (b) loss in jet penetration between

platens due to inclination angle .

Nevertheless, there may be an optimum inclination angle at which a large debonding

force may be obtained, with little loss of penetration. The objective of the following analysis was

to assess the possible use of inclined nozzles to clean recovery boiler superheater platens, by

determining the relationship between and a reduction in jet penetration, and by assessing

whether the loss in penetration for a given is very high. The flow field of a jet impinging on a

scaled-down platen at an angle was also visualized and compared with that of a jet from a straight

nozzle. Variation of the debonding component of the jet force with nozzle inclination angle will

also be interesting and useful data, but was beyond the scope of the present thesis and is a topic

for a subsequent study.

(a)

FjetFjet

(a)

H

R

h

(b)

H

R

h

(b)

137

8.1 Loss in Jet Penetration versus Nozzle Inclination Angle,

A simple measure of the loss of jet penetration as a function of the inclination angle, , is the

reduction in the maximum length of the platen, h, which would be impacted by the jet (Figure

8.2b). A simple geometrical analysis involving the side spacing between two platens H, the tube

radius R, the nozzle inclination angle , and the nozzle exit diameter de, yields the following

expression for h:

245cot

2tan)(coscos

2)cot(

Recd

Hh e … (8.1)

for 0° < < 90°.

The typical side spacing between recovery boiler superheater platens is 250-310 mm

(10”-12”); for the present calculation, H = 250 mm (10”) was assumed. Based on theoretical

considerations [32], the effective diameter of the jet close to the nozzle can be roughly

approximated as the nozzle exit diameter, de. A typical value of de for commercial sootblower

nozzles is 32 mm (1.25”). The typical superheater tube radius is 25 mm (1”). Using these values

in equation (8.1), Figure 8.3 illustrates h as a function of . Obviously, h → ∞ as → 0°, which

corresponds to a straight jet; in practice, the jet diffuses completely within a finite distance. At

large values of , very low values of h are obtained, as expected.

0

1

2

3

4

0 5 10 15 20 25 30 35 40 45

h [m]

Figure 8.3. Behaviour of h as a function of .

138

The typical length of a superheater platen in a recovery boiler is about 1.5 m. Since

sootblowers are operated from both sides of a platen, the required cleaning radius for a

sootblower can be considered as half of the platen length or 750 mm. Figure 8.3 shows that there

is a range of that yields a penetration depth h of this length. Using equation (8.1), h = 750 mm

corresponds to ≈ 17°. Thus, it seems feasible to use a 17° inclination to exert greater debonding

force on deposits, yet still clean half of a platen, without affecting the jet cleaning power or peak

impact pressure. The results imply that the loss in jet penetration is not a serious issue, and using

slightly inclined nozzles to clean superheater platens more effectively may be possible.

8.2 Schlieren Visualization of Inclined Jet Impingement

The schlieren technique was used for a preliminary investigation of the flow field of an impinging

inclined jet. Experiments similar to those described in chapter 6 section 6.1.1 were performed

using the same apparatus, in which a supersonic jet was directed at a model superheater platen. In

the experiments, the platen position was adjusted, rather than the nozzle orientation, to achieve

different inclination angles. At each inclination angle, the jet was impinged on the platen at

different offsets.

Schlieren images of the interaction are presented in Figure 8.4 with the flow from right to

left. Compared to the interaction at 0 (Figure 8.4a, offset = 10 mm), the jet/platen interaction

at 9 and 13 is more direct, but also more complicated. Due to inclination, the jet impinges more

directly on the side of the platen. It is believed that the impingement of the jet at an angle will be

beneficial for deposit removal, for two reasons.

First, as mentioned in section 2.5.2, the maximum pressure applied on a flat surface by a

supersonic jet impinging on that surface at an angle, is higher than that exerted by the same jet on

the surface when the jet is perpendicular to the surface. This effect exists when the surface is

close to the nozzle, mainly in the supersonic portion of the jet not much affected by turbulence.

139

a

α=0°

a

α=0°

b

b

c

13°

c

13°

Figure 8.4. Effect of nozzle inclination angle on jet/platen interaction.

Based on the results presented in chapter 4 for the jet used in this work and the results presented

in the work of Lamont and Hunt [49], this effect may possibly exist when a jet from an inclined

nozzle impinges on the first few tubes of a superheater platen.

Second, the impingement at an angle will likely generate more oblique shock waves on

the deposits (as can be seen in Figure 8.4b). As was described in section 2.2.2, the PIP rises

across oblique shock waves. As a result, these deposits will likely be subjected to a higher PIP

than when a jet from a straight nozzle interacts with the deposits. However, detailed investigation

must be performed to confirm these effects, and is beyond the scope of the present work.

Following impingement, the jet reorients along the direction of the platen via a

complicated shock-expansion system. With increasing, the shock-expansion system becomes

stronger.

CHAPTER 9

CONCLUSIONS, CONTRIBUTIONS, AND

RECOMMENDATIONS

9.1 Conclusions and Practical Implications

Interaction between a jet and a single tube

Upon impingement on a tube, a supersonic ‘primary’ jet deflects at an angle, forming a weaker

‘secondary’ jet. The angle and strength of the secondary jet depend on the position of the primary

jet (centreline) relative to the position of the tube (centreline), that is, on the offset. As the jet

centreline moves away from the tube centreline, jet/tube interaction weakens, and the secondary

jet becomes stronger. Once the primary jet is a small distance away from the tube, interaction

between the jet and the tube ceases.

The secondary jet that forms during jet/tube interaction spreads out much more compared

to the primary jet, and is fan-like or sheet-like. Secondary jets do not form at small offsets for

tubes larger than the jet; the flow remains attached to the tube.

The results imply that during sootblowing, secondary jets will always form when the

sootblower jet impinges on the first tube of a platen or row of tubes. So long as secondary jets

form, there will be little or no sootblower jet flow at all beyond the first few tubes in that platen

or row.

140

141

Interaction between a jet and tube arrangements

Experiments with model superheater platens showed that due to the very low spreading rate of a

supersonic jet, the primary jet stops interacting with a platen when the jet is only a small distance

(offset) away from it. Thus, deposits clinging to the side of a superheater platen are not exposed

to the sootblower jet once the jet is beyond this distance from the deposit layer. The jet must be

directed close to the platens to yield useful jet/deposit interactions. Continuous sootblowing

between platens is justified only if large deposits significantly block the space between the

platens. Due to the large spacing, a secondary jet will not be of any use in removing deposit from

an adjacent platen, due to its small core length.

A jet between generating bank tubes becomes weaker than a free jet, because of

interaction with tubes and increased mixing caused by Coanda-induced re-entrainment around the

tubes. The centreline peak impact pressure begins to decrease in the core region of the jet.

Consequently, deposits beyond the first few tubes of a row experience a weaker sootblower jet,

and thus may not be removed effectively. Due to the close tube spacing, secondary jets can

impinge on tubes in adjacent rows. PIP measurements of these secondary jets showed that the PIP

exerted is not negligible. These jets may break and remove weak deposits clinging to adjacent

tubes or accumulated between them. However, this impingement may also cause tube erosion and

corrosion due to entrainment of pieces of deposit. These results are consistent with industry

experience of generating bank tube corrosion.

A jet between finned economizer tubes also decays more quickly than a free jet, but is

stronger than the same jet in the generating bank. The strength (centreline peak impact pressure)

and hence, the deposit removal capability of the jet diminish only slightly beyond the supersonic

portion of the jet. This is because the fins restrict the entrainment and spreading of the jet. As in a

generating bank, secondary jets impinge on adjacent rows of economizer tubes due to the much

closer tube spacing, and may help remove deposits accumulated on those tubes.

142

From the point of view of experimentation, results obtained using the model superheater

platens showed that building scaled-down platens did not provide any striking new information

compared to that obtained from experiments with a single tube. As a result, any further

experiments on sootblower jet flow in the superheater section should take this into account.

Feasibility of inclined sootblower nozzles to clean superheater platens

A mathematical model was developed to study the effect of jet inclination angle on the loss in jet

penetration between kraft recovery boiler superheater platens. The model suggests that the loss in

penetration is not a serious issue, and using slightly inclined nozzles to clean superheater platens

more effectively may be possible. Visualization of a jet impinging on a platen at an angle showed

the interaction to be more direct, but also more complicated than that due to a straight jet.

Schlieren visualization of synthetic deposit breakup by jet impingement

A preliminary experiment was conducted to visualize the breakup of a synthetic deposit using the

schlieren technique. The experiment showed several interesting phenomena during breakup.

Similar experiments will yield further insight into the deposit breakup process during

sootblowing, and research in this direction is recommended.

9.2 Contributions of this Work

The main contributions of this work are the following.

1. The most important contribution is that it has characterized sootblower jet flow inside a

recovery boiler, and determined how the interaction of both the primary and secondary jets

with tubes is affected in the different sections of the boiler, especially by the tube spacing.

This has shed light on how the primary jet’s deposit removal capability may be affected in

those sections. Such an extensive visualization study of jet/tube interaction has been carried

143

out for the first time through this work, yielding valuable insight into sootblower jet flow.

This work also opens doors to new opportunities for research related to sootblowing

optimization, development of advanced sootblowers, and possible sootblower-related erosion

and corrosion.

2. This work is one of the first studies to investigate the fundamental interaction between a

supersonic jet and a cylinder. The effects of most of the important parameters governing this

interaction were studied in this work. Moreover, this work is the first to examine the

behaviour and strength of secondary jets that form when a supersonic primary jet impinges on

a cylinder. Since secondary jets form during sootblowing, this work has provided secondary

jet peak impact pressure data that can be used by boiler manufacturers to design their tube

arrangements taking sootblowing into account.

3. Interaction between a supersonic jet and a finned tube has been studied for the first time. This

study is appropriate and has been conducted at the right time because finned tubes are

commonly used in modern industrial boilers.

4. This work has shown that continuous sootblowing between superheater platens is justified

only when large deposits block the passages between platens; otherwise, that steam is simply

wasted. This work highlights the need for research to minimize this wasteful consumption of

steam, and also to develop methods to monitor and locate fouling online, so that sootblowing

may be performed optimally.

5. This work sheds light on the generating bank tube corrosion problems affecting boiler

operation in many North American pulp mills.

6. Until now, inclined sootblower nozzles have been used mainly in coal-fired utility boilers due

to the large spacing between superheater platens. Such nozzles have been introduced in kraft

recovery boilers only very recently, and are currently being evaluated. This work has shown

that the loss in jet penetration between platens due to these nozzles is not a big issue, and that

144

these nozzles can be used in kraft recovery boiler superheaters to remove large deposits from

leading tubes more effectively.

7. From the point of view of experimentation, this work has shown that building scaled-down

platens to study jet/tube interaction with superheater platens does not yield any new

information compared to that obtained using only a single tube, unless, for example the effect

of platen swinging on jet structure and strength needs to be determined. As a result, any

further experiments should take this into account. Such experiments could possibly use only

one tube, saving considerable time and cost.

9.3 Recommendations for Future Work

The following are recommendations for future work related to sootblowing optimization.

1. A detailed study of deposit breakup using the schlieren technique (the experiment described

in chapter 7 of this thesis) is strongly recommended. Firstly, such a study will shed light on

how deposits are broken and removed by a sootblower jet inside a boiler. Secondly, it will

provide information about the trajectories of broken deposit particles and their velocities.

2. Artificial deposit breakup experiments should be performed to determine the exposure time

required for a supersonic jet of given strength to break a deposit of given strength and

thickness.

3. Emami [21] developed a computational fluid dynamics (CFD) model to accurately simulate

sootblower jets and their interaction with tubes. This model should be used in conjunction

with a finite element analysis (FEA) program to determine the stresses and stress distributions

created in a typical deposit by a supersonic jet impinging on that deposit.

4. Finally, as mentioned in section 6.5, the inability of mills to monitor and locate fouling on-

line is a key bottleneck to optimizing sootblowing. There is a strong need for such methods,

and further research should be carried out in this direction.

REFERENCES

[1] Abramovich, G. N., The Theory of Turbulent Jets, Chapter 1, MIT Press, 1963.

[2] Adams, T. N., “Chapter 1 - General Characteristics of Kraft Black Liquor Recovery Boilers”. In Adams, T.N. (Ed.), Kraft Recovery Boilers, TAPPI Press, 1997.

[3] Adams, T. N., “Sootblowing Control Based on Measured Local Fouling Rate”, Proceedings, 2010 TAPPI PEERS & 9th Research on Recycling Forum, Norfolk, VA, Oct. 17-21, 2010.

[4] Adamson, T. C. and Nicholls, J. A., “On the structure of jets from highly underexpanded nozzles into still air”, Journal of the Aerospace Sciences, 26, p. 16, 1959.

[5] Anderson, J. D., Modern Compressible Flow: With Historical Perspective, 3rd ed., Chapter 1, McGraw Hill, 2003.

[6] Avduevskii, V. S., Ivanov, A. V., Karpman, I. M., Traskovskii, V. D. and Yudelovich, M. Ya., “Flow in Supersonic Viscous Underexpanded Jet”, Mekhanika Zhidkosti i Gaza, 5(3), pp. 63-69, 1970.

[7] Backman, R., Hupa, M. and Uppstu, “Fouling and corrosion mechanisms in recovery boiler superheater area”, TAPPI Journal, 70(6), pp. 123-127, 1987.

[8] Barsin, J. A., “Recovery Boiler Sootblowers”, TAPPI Kraft Recovery Operations Short Course, TAPPI Press, pp. 219-227, 1992.

[9] Baxter, L. L., Abbott, M. F. and Douglas, R. E., “Dependence of Elemental Ash Deposit Composition on Coal Ash Chemistry and Combustor Environment”. In S. A. Benson (Ed.), Inorganic Transformations and Ash Deposition during Combustion, Engineering Foundation Meeting, Palm Coast, FL, pp. 679-698, 1991.

[10] Brahma, R. K., Faruque, O. and Arora, R. C., “Experimental investigation of mean flow characteristics of slot jet impingement on a cylinder”, Wärme- und Stoffübertragung, 26, pp. 257-263, 1991.

[11] Brankovic, A., Ryder, R. C. Jr., Hendricks, R. C., Liu, N., Gallagher, J. R., Shouse, D. T., Roquemore, W. M., Cooper, C. S., Burrus, D. L. and Hendricks, J. A., “Measurement and Computation of Supersonic Flow in a Lobed Diffuser-Mixer for Trapped Vortex Combustors”, NASA TM - 2002-211127, 2002.

[12] Brown, C., “Effect of SO2 on Deposit Removal in Kraft Recovery Boilers”, MASc thesis, Dept. of Chemical Engineering and Applied Chemistry, University of Toronto, Canada, 2005.

[13] Brown, G. L. and Roshko, A., “On density effects and large structures in turbulent mixing layers”, Journal of Fluid Mechanics, 64, pp. 775-781, 1974.

145

146

[14] Bryer, D. W. and Pankhurst, R. C., Pressure-probe Methods for Determining Wind Speed and Flow Direction, H. M. S. O., London, 1971.

[15] Carling, J. C. and Hunt, B. L., “The near wall jet of a normally impinging, uniform, axisymmetric, supersonic jet”, Journal of Fluid Mechanics, 66, part 1, pp. 159-176, 1974.

[16] Crist, S., Sherman, P. M. and Glass, D. R., “Study of the Highly Underexpanded Sonic Jet”, AIAA Journal, 4(1), pp. 68-71, 1966.

[17] Derbeneva, L. I., Kurshakov, M. Yu., Tillyaeva, N. I. and Shishkin, Yu. N., “Solution of the Problem of Interaction between a Supersonic Jet and an Obstacle of Finite Dimensions”, Mekhanika Zhidkosti i Gaza, 5, pp. 179-184, 1986.

[18] Donaldson, C. D. and Snedeker, R. S., “A study of free jet impingement. Part 1. Mean properties of free and impinging jets”, Journal of Fluid Mechanics, 45, part 2, pp. 281-319, 1971.

[19] Ebrahimi-Sabet, A., “A Laboratory Study of Deposit Removal by Debonding and Its Application to Fireside Deposits in Kraft Recovery Boilers”, PhD thesis, Dept. of Chemical Engineering and Applied Chemistry, University of Toronto, 2001.

[20] Eggers, J. M., “Velocity Profiles and Eddy Viscosity Distributions Downstream of a Mach 2.22 Nozzle Exhausting to Quiescent Air”, NASA TN D-3601, 1966.

[21] Emami, B., “Numerical Simulation of Kraft Recovery Boiler Sootblower Jets”, PhD thesis, Dept. of Mechanical and Industrial Engineering, University of Toronto, 2009.

[22] Erickson, T. A., Allan, S. E., McCollor, D. P., Hurley, J. P., Srinivasachar, S., Kang, S. G., Baker, J. E., Morgan, M. E., Johnson, S. A. and Borio, R., “Modeling of fouling and slagging in coal-fired utility boilers”, Journal of Fuel Processing and Technology, 44, pp. 155-171, 1995.

[23] Eslamian, M., Pophali, A., Bussmann, M. and Tran, H. N., “Breakup of brittle deposits by supersonic air jet: The effects of varying jet and deposit characteristics”, International Journal of Impact Engineering, 36(2), pp. 199-209, 2009.

[24] Eslamian, M., Pophali, A., Bussmann, M., Cormack, D. E. and Tran, H. N., “Failure of Cylindrical Brittle Deposits Impacted by a Supersonic Air Jet”, ASME Journal of Engineering Materials and Technology, 130(3), pp. 031002-1-7, 2008.

[25] Ewan, B. C. R. and Moodie, K., “Structure and Velocity Measurements in Underexpanded Jets”, Combustion Science and Technology, 45, pp. 275-288, 1986.

[26] Foster, C. R. and Cowles, F. B., “Experimental Study of Gas-flow Separation in Overexpanded Exhaust Nozzles for Rocket Motors”, Progress report no. 4-103, Jet Propulsion Laboratory, California Institute of Technology, 1949.

[27] Goto, Y., Nonomura, T., McIlroy, K. and Fujii, K., “Detailed Analysis of Flat Plate Pressure Peaks Created by Supersonic Jet Impingements”, AIAA paper 2009-1289, 2009.

147

[28] Graham, W. J. and Davis, B. M., “The Change of Pitot Pressure across Oblique Shock Waves in a Perfect Gas”, Aeronautical Research Council, C. P. No. 783, Ministry of Aviation, H. M. S. O., London, 1965; (downloaded from: aerade.cranfield.ac.uk/ara/arc/cp/0783.pdf on Mar. 1, 2011).

[29] ImageJ - http://rsbweb.nih.gov/ij/ (as of Mar. 1, 2011).

[30] Isaak, P., Tran, H. N., Barham, D. and Reeve, D. W., “Stickiness of Fireside Deposits in Kraft Recovery Units”, Journal of Pulp und Paper Science, 12(3), pp. J84-88, 1986.

[31] Ishii, R., Fujimoto, H., Hatta, N. and Umeda, Y., “Experimental and numerical analysis of circular pulse jets”, Journal of Fluid Mechanics, 392, pp. 129-153, 1999.

[32] Jameel, M.I., Cormack, D.E., Tran, H.N. and Moskal, T.E., “Sootblower Optimization Part I: Fundamental hydrodynamics of a sootblower nozzle and jet”, TAPPI Journal, 77(5), pp. 135-142, 1994.

[33] Jenkins, B. M., Baxter, L. L., Miles Jr., T. R. and Miles, T. R., “Combustion properties of biomass”, Journal of Fuel Processing and Technology, 54, pp. 17–46, 1998.

[34] Kalghatgi, G. T. and Hunt, B. L., “The Occurrence of Stagnation Bubbles in Supersonic Jet Impingement Flows”, Aeronautical Quarterly, 27, pp. 169-185, 1976.

[35] Kaliazine, A., Cormack, D.E., Ebrahimi-Sabet, A. and Tran, H.N., “The Mechanics of Deposit Removal in Kraft Recovery Boilers”, Journal of Pulp and Paper Science, 25(12), pp. 418-424, 1999.

[36] Kaliazine, A., Cormack, D. E., Tran, H. N. and Jameel, I., “Feasibility of Using Low Pressure Steam for Sootblowing”, Pulp & Paper Canada, 107(4), pp. 34-38, 2006.

[37] Kaliazine, A., Piroozmand, F., Cormack, D.E. and Tran, H.N., “Sootblower Optimization Part II: Deposit and sootblower interaction”, TAPPI Journal, 80(11), pp. 201-207, 1997.

[38] Kaliazine, A., Pophali, A. and Tran, H. N., "Thermal Effect of Sootblower Jets on Deposits", Proceedings, Annual Meeting of the Consortium on “Increasing Energy and Chemical Recovery Efficiency in the Kraft Pulping Process", University of Toronto, Canada, Nov. 2007.

[39] Kaliazine, A., Tandra, D. S., Cormack, D. E. and Tran, H. N., “Sootblower Jet Impact Pressure near Platen Surfaces”, Report no. 12, Proceedings, Annual Meeting of the Consortium on “Increasing the Throughput and Reliability of Recovery Boilers and Lime Kilns”, University of Toronto, Canada, Nov. 2003.

[40] Kaliazine, A., Tran, H.N. and Cormack, D.E., “Understanding Deposit Removal Process in Recovery Boilers”, Proceedings, TAPPI Engineering/Finishing & Converting Conference, San Antonio, TX, pp. 703-712, 2001.

[41] Katanoda, H., Handa, T., Miyazato, Y., Masuda, M. and Matsuo, K., “Effect of Reynolds Number on Pitot-Pressure Distributions in Underexpanded Supersonic Freejets”, Journal of Propulsion, 17(4): Technical Notes, pp. 940-942, 2001.

148

[42] Katanoda, H., Miyazato, Y., Masuda, M. and Matsuo, K., “Pitot pressures of correctly-expanded and underexpanded free jets from axisymmetric supersonic nozzles”, Shock Waves, 10, pp. 95-101, 2000.

[43] Khalajzadeh, A., Kuhn, D. C. S., Tran, H.N. and Wessel, R., “Composition of Carryover Particles in Recovery Boilers”, Journal of Pulp & Paper Science, 32(2), pp. 90-94, 2006.

[44] Kleinstein, G., “Mixing in Turbulent Axially Symmetric Free Jets”, Journal of Spacecraft and Rockets, 1(4), pp. 403-408, 1964.

[45] Krehl, P. and Engemann, S., “August Toepler - the first who visualized shock waves”, Shock Waves, 5, pp. 1-18, 1995.

[46] Krothapalli, A., Rajkuperan, E., Alvi, F. and Lourenco, L., “Flow field and noise characteristics of a supersonic impinging jet”, Journal of Fluid Mechanics, 392, pp. 155-181, 1999.

[47] Kweon, Y. -H., Miyazato, Y., Aoki, T., Kim, H. -D. and Setoguchi, T., “Experimental investigation of nozzle exit reflector effect on supersonic jet”, Shock Waves, 15, pp. 229-239, 2006.

[48] LabVIEW User Manual, National Instruments Corp., 2003.

[49] Lamont, P. J. and Hunt, B. L., “The Impingement of Underexpanded, Axisymmetric Jets on Perpendicular and Inclined Flat Plates,” Journal of Fluid Mechanics, 100, part 3, pp. 471–511, 1980.

[50] Lau, J. C., Morris, P. J. and Fisher, M. J., “Measurements in subsonic and supersonic free jets using a laser velocimeter”, Journal of Fluid Mechanics, 93, part 1, pp. 1-27, 1979.

[51] Love, E. S., Grigsby, C. E., Lee, L. P., and Woodling, M. J., “Experimental and Theoretical Studies of Axisymmetric Free Jets”, NASA TR R-6, 1959.

[52] Mao, X., Tran, H.N. and Cormack, D.E., “Effects of Chemical Composition on the Removability of Recovery Boiler Fireside Deposits”, TAPPI Journal, 84(6), p. 68, 2001.

[53] Mao, X., Tran, H. N., Kaliazine, A. and Cormack, D. E., “Effects of Deposit Forming and Jet Blowing Conditions on Deposit Removability”, Report no. 21, Proceedings, Annual Meeting of the Consortium on “Improving Recovery Boiler Performance, Emissions and Safety”, University of Toronto, Canada, Nov. 2000.

[54] “Model RKS Long Retractable Sootblower” brochure, Clyde Bergemann, Inc., Atlanta, 2005. Downloaded on Mar. 1, 2011 from: http://www.clydebergemann.com

[55] Nakai, Y., Fujimatsu, N. and Fujii, K., “Experimental Study of Underexpanded Supersonic Jet Impingement on an Inclined Flat Plate”, AIAA Journal, 44(11), pp. 2691-2699, 2006.

[56] Nishikawa, E., “Chapter 3 - General Planning of the Boiler Gas-side Heat Transfer

149

Surface”. In Ishigai, S. (Ed.), Steam Power Engineering: Thermal and Hydraulic Design Principles, Cambridge University Press, 1999.

[57] Norum, T. D. and Seiner, J. M., “Measurements of Mean Static Pressure and Far-Field Acoustics of Shock-Containing Supersoinc Jets”, NASA TM 84521, 1982.

[58] Pack, D. C., “A Note on Prandtl’s Formula for the Wave-length of a Supersonic Gas Jet”, Quarterly Journal of Mechanics and Applied Mathematics, 3(2), pp. 173-181, 1950.

[59] Pack, D. C., “On the Formation of Shock Waves in Supersonic Gas Jets (Two-dimensional Flow)”, Quarterly Journal of Mechanics and Applied Mathematics, 1(1), pp. 1-17, 1948.

[60] Panda, J. and Seasholtz, R. G., “Measurement of shock structure and shock-vortex interaction in underexpanded jets using Rayleigh scattering”, Physics of Fluids, 11(12), pp. 3761-3777, 1999.

[61] Panda, J., “An experimental investigation of screech noise generation”, Journal of Fluid Mechanics, 378, pp. 71-96, 1999.

[62] Papamoschou, D. and Johnson, A., “Unsteady Phenomena in Supersonic Nozzle Flow Separation”, AIAA paper 2006-3360, 2006.

[63] Papamoschou, D. and Roshko, A., “The compressible turbulent shear layer: an experimental study”, Journal of Fluid Mechanics, 197, pp. 453-477, 1988.

[64] Pinckney, S. Z., “Short Static-pressure Probe Design for Supersonic Flow”, NASA TN D-7978, 1975.

[65] Piroozmand, F., Tran, H. N., Kaliazine, A. and Cormack, D. E., “Strength Recovery Boiler Fireside Deposits at High Temperatures”, Proceedings, TAPPI Engineering Conference, Miami, FL, Sept. 13-17, 1998.

[66] Pophali, A., “Breakup Mechanisms of Brittle Deposits Impinged by a Supersonic Air Jet”, MASc thesis, Dept. of Chemical Engineering and Applied Chemistry, University of Toronto, Canada, 2008.

[67] Pophali, A., Bussmann, M. and Tran, H. N., “Visualizing the Interactions between a Sootblower Jet and a Superheater Platen”, Proceedings, International Chemical Recovery Conference, Williamsburg, VA, Mar. 29-Apr. 1, pp. 196-206, 2010.

[68] Pophali, A., Emami, B., Bussmann, M., and Tran, H. N., “Studies on Sootblower Jet Dynamics and Ash Deposit Removal in Industrial Boilers”, Journal of Fuel Processing and Technology (special ed.), accepted for publication, 2011.

[69] Pophali, A., Eslamian, M., Kaliazine, A., Bussmann, M. and Tran, H. N., “Breakup mechanisms of brittle deposits in kraft recovery boilers - a fundamental study”, TAPPI Journal, 8(9), pp. 4-9, 2009.

[70] Powell, A., “On the Mechanism of Choked Jet Noise”, Proceedings of the Physical Society, Section B, 66(12), pp. 1039-1056, 1953.

150

[71] Prandtl, L., “Über die stationaren Wellen in einem Gasstrahl”, Phys. Zeits, 5, pp. 599-601, 1904.

[72] Raask, E., Mineral Impurities in Coal Combustion, Hemisphere, 1985.

[73] Rahimi, M., Owen, I. and Mistry, J., “Heat transfer between an under-expanded jet and a cylindrical surface”, International Journal of Heat and Mass Transfer, 46, pp. 3135-3142, 2003.

[74] Rajakuperan, E. and Ramaswamy, M. A., “An experimental investigation of underexpanded jets from oval sonic nozzles”, Experiments in Fluids, 24, pp. 291-299, 1998.

[75] Raman, G., “Advances in Understanding Supersonic Jet Screech: Review and Perspective”, Progress in Aerospace Sciences, 34, pp. 45-106, 1998.

[76] Raman, G., “Cessation of screech in underexpanded jets”, Journal of Fluid Mechanics, 336, pp. 69-90, 1997.

[77] Rathakrishnan, E., Instrumentation, Measurements, and Experiments in Fluids, CRC Press, 2007.

[78] Rezvani, K., Mao, X. and Tran, H. N., “Effects of Potassium, Carbonate and Sulfide on Carryover Deposition”, Report no. 2, Proceedings, Annual Meeting of the Consortium on “Improving Recovery Boiler Performance, Emissions and Safety”, University of Toronto, Canada, Nov. 2000.

[79] Romeo, L. M. and Gareta, R., “Fouling control in biomass boilers”, Biomass and Bioenergy, 33, pp. 854-861, 2009.

[80] Saviharju, K., Kaliazine, A., Tran, H. N. and Habib, T., “In-situ measurements of sootblower jet impact in a recovery boiler”, TAPPI Journal, 10(2), pp. 27-32, 2011.

[81] Schuh, H. and Persson, B., “Heat Transfer on Circular Cylinders Exposed to Free-Jet Flow”, International Journal of Heat and Mass Transfer, 7, pp. 1257-1271, 1964.

[82] Seiner, J. M. and Norum, T. D., “Aerodynamic aspects of shock containing jet plumes”, AIAA paper 80-0965, 1980.

[83] Seiner, J. M., Manning, J. C. and Ponton, M. K., “Model and full scale study of twin supersonic plume resonance”, AIAA paper 87-0244, 1987.

[84] Settles, G. S., Schlieren and Shadowgraph Techniques: Visualizing Phenomena in Transparent Media, Springer-Verlag, 2001.

[85] Shenassa, R., “Dynamic Carryover Deposition in an Entrained Flow Reactor”, PhD thesis, Dept. of Chemical Engineering and Applied Chemistry, University of Toronto, 2000.

[86] Shenassa, R., Kuhn, D.C.S. and Tran, H. N., “The Role of Liquid Content In Carryover Deposition In Kraft Recovery Boilers - A Fundamental Study”, Journal of Pulp & Paper Canada, 29(4), pp.132-136, 2003.

151

[87] Shenassa, R., Tran, H.N. and Kuhn, D.C.S., “Dynamic Study of Carryover Deposition using an Entrained Flow Reactor”, Pulp & Paper Canada, 100(10), pp. 56-62, 1999.

[88] Sherman, P. M., Glass, D. R. and Duleep, K. G., “Jet Flow Field During Screech”, Journal of Applied Sciences Research, 32, pp. 283-303, 1976.

[89] Singbeil, D., “Corrosion of Generating Bank Tubes in Kraft Recovery Boilers”, Pulp & Paper Canada, 95(12), pp. 132-135, 1994.

[90] Stultz, S. C. and Kitto, J. B. (Eds.), Steam: Its Generation and Use, 41st ed., The Babcock & Wilcox Co., Barberton, OH, 2005.

[91] “Summary Report of Energy Use in the Canadian Manufacturing Sector 1995-2008”, Industrial Consumption of Energy (ICE) Survey, Natural Resources Canada, Oct. 2010; downloaded on Mar. 1, 2011 from: http://oee.nrcan.gc.ca/publications/statistics/ice08/index.cfm

[92] Tabrizi, S. P. A., “Jet Impingement onto a Circular Cylinder”, PhD thesis, University of Liverpool, 1996.

[93] Tam, C. K. W. and Ahuja, K. K., “Theoretical model of discrete tone generation by impinging jets”, Journal of Fluid Mechanics, 214, pp. 67-87, 1990.

[94] Tam, C. K. W., Jackson, J. A. and Seiner, J. M., “A Multiple-scales Model of the Shock-cell Structure of Imperfectly Expanded Supersonic Jets”, Journal of Fluid Mechanics, 153, pp. 123-149, 1985.

[95] Tam, C. K. W., “On the Noise of a Nearly Ideally Expanded Supersonic Jet”, Journal of Fluid Mechanics, 51, pp. 69-95, 1972.

[96] Tam, C. K. W., “Supersonic Jet Noise”, Annual Review of Fluid Mechanics, 27, pp. 17-43, 1995.

[97] Tandra, D.S., “Development and Application of a Turbulence Model for a Sootblower Jet Propagating Between Recovery Boiler Superheater Platens”, PhD thesis, Dept. of Chemical Engineering and Applied Chemistry, University of Toronto, 2005.

[98] Tandra, D. S., Kaliazine, A., Cormack, D. E. and Tran, H. N., “Interaction between sootblower jet and superheater platens in recovery boilers”, Pulp & Paper Canada, 108(5), pp. 43-46, 2007.

[99] Tandra, D. S., Kaliazine, A., Cormack, D. E. and Tran, H. N., “Numerical simulation of supersonic jet flow using a modified k-ε model”, International Journal of Computational Fluid Dynamics, 20(1), pp. 19-27, 2006.

[100] Tandra, D. S., Manay, A. and Edenfield, J., “The Use of Energy Balance Around Recovery Boiler Heat Exchangers to Intelligently Manage Sootblower Operations: A Case Study”, Proceedings, 2010 TAPPI PEERS & 9th Research on Recycling Forum, Norfolk, VA, Oct. 17-21, 2010.

[101] Tandra, D. S., Manay, A., Thabot, A., and Jones, A., “Mill Trial on New Sootblower Design and Strategy to Combat Plugging in a Recovery Boiler”, Proceedings, 2010

152

TAPPI PEERS & 9th Research on Recycling Forum, Norfolk, VA, Oct. 17-21, 2010.

[102] Tandra, D. S., Personal communication, 2008.

[103] Tawfek, A. A., “Heat transfer due to a round jet impinging normal to a circular cylinder”, Heat and Mass Transfer, 35, pp. 327-333, 1999.

[104] Theis, M., Skrifvars, B.-J., Hupa, M. and Tran, H.N., “Fouling Tendency of Ash Resulting from Burning Mixtures of Biofuels. Part 1: Deposition Rates”, Fuel, 85, pp. 1125-1130, 2006.

[105] Tran, H. N. and Vakkilainnen, E. K., “The Kraft Chemical Recovery Process”, TAPPI Kraft Recovery Operations Short Course, TAPPI Press, pp. 1.1-1 – 1.1-8, 2011.

[106] Tran, H.N., Barham, D. and Reeve, D.W., “Sintering of fireside deposits and its impact on plugging in kraft recovery boilers”, TAPPI Journal, 70(4), pp. 109-113, 1988.

[107] Tran, H.N., “Chapter 9 - Upper Furnace Deposition and Plugging”. In Adams, T.N. (Ed.), Kraft Recovery Boilers, TAPPI Press, 1997.

[108] Tran, H. N., “Kraft Recovery Boiler Plugging and Prevention”, TAPPI Kraft Recovery Operations Short Course, TAPPI Press, pp. 209-217, 1992.

[109] Tran, H. N., Personal communication, 2011.

[110] Tran, H.N., Reeve, D.W. and Barham, D., “Formation of kraft recovery boiler superheater deposits”, Pulp & Paper Canada, 84(1), T7-T12, 1983.

[111] Tran, H. N., Tandra, D. S. and Jones, A., “Development of Low-Pressure Sootblowing Technology”, Proceedings, International Chemical Recovery Conference, PAPTAC & TAPPI, Quebec City, Canada, May 29-June 1, 2007.

[112] Tritton, D. J., Physical Fluid Dynamics, Van Nostrand Reinhold Co., 1977.

[113] Umeda, Y., Maeda, H. and Ishii, R., “Discrete tones generated by the impingement of a high-speed jet on a circular cylinder”, Physics of Fluids, 30(8), pp. 2380-2388, 1987.

[114] Venkatakrishnan, L., “Density Measurements in an Axisymmetric Underexpanded Jet by Background-Oriented Schlieren Technique”, AIAA Journal, 43(7), pp.1574-1579, 2005.

[115] “Wall Blowers”, from: http://www.clydebergemann.com/ (Products>Cleaning Devices>Wall Blowers) on Nov. 30, 2008.

[116] Warren, W. R., “An analytical and experimental study of compressible free jets”, Report no. 381, Aeronautical Engineering Laboratory, University of Princeton, 1957.

[117] White, F. M., Fluid Mechanics, 5th ed., McGraw Hill, 2003.

[118] Witze, P. O., “A Generalized Theory for The Turbulent Mixing of Axially Symmetric Compressible Free Jets”, Fluid Mechanics of Mixing, Proceedings, Joint Meeting of the Fluids Engineering and Applied Mechanics Divs., ASME, Atlanta, GA, p. 63, 1973.

153

[119] Zapryagaev, V. I. and Solotchin, A. V., “Three-dimensional structure of flow in a supersonic underexpanded jet”, J. Appld. Mech. Tech. Phys., 32(4), pp. 503-507, 1991.

[120] Zdravkovich, M. M., Flow Around Circular Cylinders: Vol. I Fundamentals, Oxford University Press, 2003.

APPENDICES

154

155

APPENDIX A

Schlieren Images of Jet Temporal Development

9.32 18.64

17.14

10.32

11.98

a

b

c

d

g

h

f

e

7.32

0 ms

5.49

jet starts developing

nozzle

Figure A.1. Temporal development of the jet used in the experiments.

Figure A.1 presents the schlieren images of the temporal development of the jet used in the

experiments. 0 ms (image a) corresponds to when the first wave of air exited the nozzle, and was

determined from the corresponding schlieren movie. The figure shows that the jet attains quasi

steady-state in less than 20 ms (image h). No change in jet structure was observed beyond 20 ms

until the end of the jet blow.

jet reaches quasi steady-state9.32 18.64

17.14

10.32

11.98

a

b

c

d

g

h

f

e

7.32

0 ms

5.49

jet starts developing

nozzle

jet reaches quasi steady-state

156

APPENDIX B

LabVIEW Graphical Program to Control the Data Acquisition

System

The LabVIEW User Manual [48] was consulted when developing this program.

Figure B.1. LabVIEW graphical program developed to operate the solenoid valve, high-

speed camera and DAQ system during experiments.

157

APPENDIX C

Interaction between a Jet and a Single Tube: Supplementary

Results

This appendix contains supplementary results obtained from experiments performed to study the

interaction between a supersonic jet and a single tube. The main results related to these

interactions are presented in chapter 5. The results presented here are specifically schlieren

images of (1) the interaction when a primary jet impinges on a large tube (25.4 mm OD, de/D =

0.29) located 50 mm (6.8de) from the nozzle exit, at different offsets (section C.1), and (2) the

instability of the flow when a primary jet impinges at 0 offset on a medium tube (19.1 mm OD,

de/D = 0.39) located 3de away from the nozzle (section C.2). In this case, a secondary jet forms

after a much longer time period as compared to all other cases in which it forms. To present this

clearly, the time required for flow to separate from a tube and form a secondary jet is compared

for the medium tube when it is placed 2de and 3de away from the nozzle. To present these images

properly, each figure containing a set of images is presented on a new page, starting from the next

page, since these figures require considerable space.

158

C.1 Interaction between a Jet and a Large Tube (25.4 mm OD, de/D =

0.29): Effect of Offset

a

b

c

d

e

f

g

h

i

j

= 0R

0.20R

0.39R

0.59R

0.79R

1.18R

1.38R

1.58R

2.18R

0.98R

a

b

c

d

e

f

g

h

i

j

= 0R

0.20R

0.39R

0.59R

0.79R

1.18R

1.38R

1.58R

2.18R

0.98R

Figure C.1. Jet impinging on a large tube at different offsets.

159

C.2 Flow Instability: Primary Jet Impinging at 0 Offset on a Medium

Tube (19.1 mm OD, de/D = 0.39) Located 3 Nozzle Exit Diameters from

the Nozzle Exit

Figures C.2 and C.3 show the temporal development of the flow field when a primary jet

impinges on a medium tube (19.1 mm OD, de/D = 0.39) placed 2de and 3de away from the nozzle

respectively. Time 0 is arbitrarily selected, but corresponds to nearly the same stage of flow

development in both cases. The uncertainty in determining the time 0 in these cases is 1 frame

or 0.17 ms.

Figures C.2 and C.3 show that the secondary jet that forms when the medium tube is

placed 3de away from the nozzle, forms after a much longer period of time compared to when the

tube is placed 2de away, as well as in all other situations (for the small and medium tubes, at

different offsets and different distances). In all other cases, the secondary jet forms as the primary

jet develops, that is, the secondary jet forms before the flow field reaches quasi steady-state. For

the case shown in Figure C.2, the time required for the secondary jet to form from time 0 is about

16 ms. Together with the time required for the jet to reach the stage at time 0, the time required

for a secondary jet to form is approximately the same as that required for the primary jet to fully

develop, which is around 20 ms, described in chapter 3 section 3.1 and in Appendix A.

At x = 3de (Figure C.3), the flow field reaches a ‘first’ quasi steady-state in which the

flow around the tube remains attached to the tube for a considerable length of time (images d and

e). Thereafter, the flow separates from the tube forming a stable secondary jet, and hence the flow

field attains a ‘second’ quasi steady-state (images f-h). The total time required for the secondary

jet to form in this case (time required to reach the stage at time 0 + the time from time 0) is about

60 ms, which is much higher than that at x = 2de. Beyond x = 3de, secondary jets do not form at 0

offset for this tube, as indicated in chapter 5 section 5.3.1. This indicates that this nozzle-tube

160

distance (between 2de and 4de) for this tube size (de/D = 0.39) is a transition region in terms of

flow separation from the tube.

161

a

b

c

d

e

f

g

h

time=0ms

3.74

8.33

12.75

13.43

14.11

14.79

15.81

secondary jet

tube

stand

nozzle

25mm

a

b

c

d

e

f

g

h

time=0ms

3.74

8.33

12.75

13.43

14.11

14.79

15.81

secondary jet

tube

stand

nozzle

25mm

Figure C.2. Temporal development of the flow field when a primary jet impinges on a

medium tube placed 2de away from the nozzle; the secondary jet forms as the primary

jet develops.

162

a

b

c

d

e

f

g

h

time=0ms

3.74

7.48

15.13

35.53

55.93

57.63

58.82

secondary jet

flow reaches 1st

quasi steady-stateflow reaches

2nd quasi steady-state

flow separation starts

a

b

c

d

e

f

g

h

time=0ms

3.74

7.48

15.13

35.53

55.93

57.63

58.82

secondary jet

flow reaches 1st

quasi steady-stateflow reaches

2nd quasi steady-state

flow separation starts

Figure C.3. Temporal development of the flow field when a primary jet impinges on a

medium tube placed 3de away from the nozzle; flow first remains attached to the tube for

a long period of time, after which the secondary jet forms.

163

APPENDIX D

Typical Kraft Recovery Boiler Tube Arrangements: Results of

a Recent Survey (August 2010)

To obtain quantitative data about typical kraft recovery boiler tube arrangements, three boiler

manufacturers - Babcock & Wilcox, Andritz, and Metso Power were contacted. The data

provided by them is summarized in the tables below. Where available, data from [2] is also

included as it represents typical tube arrangements from about 13 years ago. It was found that, the

tube arrangements of the different boiler manufacturers are in general similar to each other. The

survey results also clearly showed the trend towards finned tube generating banks and

economizers in modern boilers, from traditional un-finned tube generating banks. The side

spacing of modern generating banks is larger than that of earlier ones. Tube size (OD) in the

different sections varies between 1.67”-2.75” (42.4-70.0 mm).

Survey results

(Dimensions in parentheses are in inches.)

Superheater

ParameterRange/

Typical Values [mm]

D 57-70

(2.24”-2.75”) 241-255

(9.50”-10.04”) H Ref. [2]

178 or 254 (7” or 10”)

S 3-38

(0.12”-1.50”)

L 762-2277

(30.00”-89.65”)

D S

H

L

D S

H

L

164

Generating bank – finned tubes

ParameterRange/

Typical Values [mm]

D 43.2, 63.5

(1.7”, 2.5”)

H 89, 140, 160, 165

(3.5”, 5.5”, 6.3”, 6.5”)

S 0, 12.7, 51, 71

(0”, 0.5”, 2”, 2.8”)

T 4.8, 5.1

(0.19”, 0.20”)

W 28, 38

(1.1”, 1.5”)

L 1067-1334

(42.0”-52.5”)

D S

H

L

W

T

D S

H

L

W

T

D S

H

L

W

T

Generating bank – un-finned tubes

ParameterRange/

Typical Values [mm] 51-63.5

(2.0”-2.5”) D

Ref. [2] 63.5 (2.5”)

63.5 (2.5”)

H Ref. [2]

50.8 (2”)

S 127 (5”)

L 1257

(49.5”) Economizer

ParameterRange/

Typical Values [mm]

D 42.4, 43.9, 44.5, 50.8

(1.67”, 1.73”,1.75”, 2”)

H 50.8-62.5 (2”-2.46”)

S 0-12.7

(0”-0.5”)

T 4.1, 4.8

(0.16”, 0.188”)

W 39.1, 50.8-63.5 (1.54”, 2”-2.5”)

L 859-2921

(33.8”-115”)

D S

H

L

D S

H

L

D S

H

L

D S

H

L

W

T

D S

H

L

W

T

D S

H

L

W

T

165

APPENDIX E

Schlieren Images of a Jet Midway between Two Rows of

Finned Economizer Tubes

primary jet

1st tube (11mm OD)

fin

stand expansion wave

2nd3rd

primary jet

1st tube (11mm OD)

fin

stand expansion wave

2nd3rd

primary jet

1st tube (11mm OD)

fin

stand expansion wave

2nd3rd

3rd tube4th5th6th

(a)

(b)

Figure E.1. Jet midway between two rows of finned economizer tubes; (a) near field; (b)

far field; the nozzle is to the right hand side in these images.

3rd tube4th5th6th 3rd tube4th5th6th

(a)

(b)