testing and prediction of erosion-corrosion for corrosion ... dissertations/hernan... · testing...

277
THE UNIVERSITY OF TULSA THE GRADUATE SCHOOL TESTING AND PREDICTION OF EROSION-CORROSION FOR CORROSION RESISTANT ALLOYS USED IN THE OIL AND GAS PRODUCTION INDUSTRY by Hernan E. Rincon A dissertation submitted in partial fulfillment of the requirements for degree of Doctor of Philosophy In the discipline of Mechanical Engineering The Graduate School The University of Tulsa 2006

Upload: hakhue

Post on 07-Aug-2018

227 views

Category:

Documents


1 download

TRANSCRIPT

T H E U N I V E R S I T Y O F T U L S A

THE GRADUATE SCHOOL

TESTING AND PREDICTION OF EROSION-CORROSION FOR CORROSION

RESISTANT ALLOYS USED IN THE OIL AND GAS PRODUCTION

INDUSTRY

byHernan E. Rincon

A dissertation submitted in partial fulfillment of

the requirements for degree of Doctor of Philosophy

In the discipline of Mechanical Engineering

The Graduate School

The University of Tulsa

2006

ii

T H E U N I V E R S I T Y O F T U L S A

THE GRADUATE SCHOOL

TESTING AND PREDICTION OF EROSION-CORROSION FOR CORROSION

RESISTANT ALLOYS USED IN THE OIL AND GAS PRODUCTION

INDUSTRY

by

Hernan E. Rincon

A DISSERTATION

APPROVED FOR THE DISCIPLINE OF

MECHANICAL ENGINEERING

By Dissertation Committee

_______________________________, Co-ChairJohn R. Shadley, Ph.D.

_______________________________, Co-ChairEdmund F. Rybicki, Ph.D.

_______________________________Kenneth P. Roberts, Ph.D.

_______________________________Dale C. Teeters, Ph.D.

iii

COPYRIGHT STATEMENT

Copyright © [2006] by [Hernan, E, Rincon]

All rights reserved. No part of this publication may be reproduced, stored in a

retrieval system, or transmitted, in any form or by any means (electronic, mechanical,

photocopying, recording or otherwise) with the prior written permission of the author.

iv

ABSTRACT

Rincon, Hernan Enrique (Doctor of Philosophy)

TESTING AND PREDICTION OF EROSION-CORROSION FOR CORROSIONRESISTANT ALLOYS USED IN THE OIL AND GAS PRODUCTION INDUSTRY(337pp.- Chapter 10)

Directed by Dr. John Shadley and Dr. Edmund Rybicki

( 328 Words)

The corrosion behavior of CRAs has been thoroughly investigated and

documented in the public literature by many researchers; however, little work has been

done to investigate erosion-corrosion of such alloys. When sand particles are entrained in

the flow, the degradation mechanism is different from that observed for sand-free

corrosive environment. There is a need in the oil and gas industry to define safe service

limits for utilization of such materials.

The effects of flow conditions, sand rate, pH and temperature on the erosion-

corrosion of CRAs were widely studied. An extensive experimental work was conducted.

using scratch tests and flow loop tests using several experimental techniques.

At high erosivity conditions, a synergistic effect between erosion and corrosion

was observed. Under the high sand rate conditions tested, erosivity is severe enough to

damage the passive layer protecting the CRA thereby enhancing the corrosion rate. In

most cases there is likely a competition between the rates of protective film removal due

to mechanical erosion and protective film healing.

v

Synergism occurs for each of the three alloys examined (13Cr and Super13Cr and

22Cr); however, the degree of synergism is quite different for the three alloys and may

not be significant for 22Cr for field conditions where erosivities are typically much lower

that those occurring in the small bore loop used in this research. Predictions of the

corrosion component of erosion-corrosion based on scratch test data compared

reasonably well to test results from flow loops for the three CRAs at high erosivity

conditions. Second order behavior appears to be an appropriate and useful model for

representing the repassivation process of CRAs.

A framework for a procedure to predict penetration rates for erosion-corrosion

conditions was developed based on the second order model behavior observed for the re-

healing process of the passive film of CRAs and on computational fluid dynamics (CFD)

simulations for erosion conducted for a direct impingement flow geometry. Reasonably

good agreement between the experimental and predicted erosion-corrosion penetration

rates was found.

vi

ACKNOWLEDGMENTS

I express my deep gratitude to my advisor Dr. Shadley, who consistently has

shown a great dedication and support, even while facing very difficult and painful times.

His guidance and encouragement really helped me overcome the challenges of this

research. I also give special thanks to Dr. Rybicki for his guidance and timely suggestions

which helped me throughout this research as well as in all aspects of my academic life. I

express sincere gratitude to Dr. Shadley and Dr. Rybicki for having provided me an

opportunity to work for a Ph.D. degree.

I am grateful for timely suggestions of Dr. Roberts. I also thank Dr. Teeters for

being part of my committee and providing his expertise. Special thanks goes to Senior

Technician Mr. Bowers for his expertise and support in the laboratory. I extend my

gratitude to the member companies of the Erosion/Corrosion Research Center. Special

thanks goes to Dr. McLaury and my colleague Yongli for their help throughout the CFD

work. I also express my fraternal appreciation to my friends Mauricio Papa, Jesus

Gonzalez and Rodrigo Chandia for their valuable encouragement and motivation since

the very beginning of this wonderful and challenging experience, as well as for providing

me with state of the art computers to speed up the CFD simulation. Thanks to all my

friends. This also gives me a great opportunity to express my deep appreciation, to my

wife Raicelina, to my son Ricardo, to my parents Eudelys, and Hernan, to my sister,

brother and nephew, Rossana, Enrique and Gabriel, and my coming baby for being part

of my life.

vii

DEDICATION

To my loving wife, my son Ricardo and my coming baby

viii

CONTENTS

COPYRIGHT STATETMENT……..………………..……….…….………......... iii

ABSTRACT…………………..………………………………………….……..... iv

ACKNOWLEDGMENTS…………………………………………………..……. vi

DEDICATION………………..……………………………………….………..... vii

CONTENTS……………………………………………………………..……….. viii

TABLE INDEX……………………………………………………………..……. xiii

FIGURE INDEX………………………………………………………………….. xv

CHAPTER 1 : INTRODUCTION…………..…………………………….…… 1

CHAPTER 2 : BACKGROUND AND LITERATURE REVIEW....…….….. 5

Basic Corrosion Concepts…..…………….……………………….…... 5

Concept and forms of corrosion…………………………….…... 5

Review of the Electrochemical Basis of Corrosion………… …...……. 7

Theory behind polarization measurements……………………… 8

Linear polarization resistance….…………..………...……..……. 11

Potentiodynamic polarization and Tafel constants……………….. 16

Active Passive Metal Behavior…………………………..……….…….. 17

Passive film……...………………………………………….……. 23

Maintenance and breakdown of passivity………………………... 26

Stainless Steel……………………...…………………………………….. 27

Alloying elements……………………………………………….... 27

ix

Flow velocity effects………………………………………….... 29

Flow pattern effects………………………………………….… 31

CO2 Corrosion Resistance of 13Cr Alloy………………….………... 33

Basic Erosion Concepts……………..…………………………….…. 36

Solid particle erosion………………………………………….. 37

Erosion of ductile materials…………….……………………... 39

Erosion of brittle materials………….……….……………...…... 39

Variables influencing erosion…….…………………………….. 40

Erosion-Corrosion……………..………….…….…………………...… 42

13Cr Alloy and Erosion-Corrosion………………………………..….. 46

Single phase liquid flow loop testing (high sand rates)……….... 51

Multiphase flow loop testing (high sand rates)……………...….. 52

CHAPTER 3: OBJECTIVES AND APPROACH……………………….…… 57

Research Objectives………….……………………………………….. 57

Research Approach…………………….……………………………… 58

CHAPTER 4: EXPERIMENTAL PROCEDURE AND TESTING

CONDITIONS…………………………………………………. 61

Scratch Test Experimental Setup……………………………………... 61

Test matrix..………………………………………………….….. 64

Erosion-Corrosion -Loop (Gas/Liquid/Sand Multiphase

Flow Loop)…………………………………………………………….. 64

Test cell…………………………………………………….….... 67

Test conditions…………………………………………….…..... 67

x

Erosion-Corrosion Liquid/Sand Loop (Microloop)…..………….….. 68

Direct impingement test cell…………………………………….. 71

Test conditions for liquid/sand loop test……………………… 74

Material tested…………………………………………………... 75

CHAPTER 5: SCRATCH TEST AS A SIMPLIFIED EROSION-

CORROSION TEST……………………………………….…. 76

Motivation for Doing Scratch Test…………………………………… 76

Data Reduction Technique……………………………………….…….. 77

Scratch Test Results………………………….……………………….. 81

Effect of pH on Scratch Test results………………………….…. 81

Effect of temperature on Scratch Test results............................... 85

Effect of type of material (CRA) on Scratch Test results……….. 90

Cumulative Thickness Loss and Repassivation Time…………..…... 92

CHAPTER 6: MULTIPHASE GAS/LIQUID/SAND FLOW LOOP

TESTING RESULTS…………………..…………………….... 100

Low Sand Rates (Multiphase Flow Loop Testing)………………….. 101

High Sand Rates (Multiphase Flow Loop Testing)…………………. 108

CHAPTER 7: COMPARISON OF SCRATCH TEST RESULTS WITH

FLOW LOOP TEST RESULTS…….………………………… 114

Scratch Test Data vs. Single Phase Liquid Flow Data………………. 116

Prediction of Erosion-Corrosion of CRAs using the Scratch Test….. 117

Single phase liquid flow…………………………..………………. 117

xi

Validation of Scratch Test Predictions of Erosion-Corrosion of

CRAs……………………………………………………………….…… 119

Multiphase flow……………………………………….………… 119

CHAPTER 8: SUBMERGED DIRECT IMPINGEMENT TEST:

SINGLE PHASE LIQUID FLOW…..……………………….... 122

Erosion-Corrosion Liquid/Sand Loop (Microloop)…………..…..….. 122

CHAPTER 9: EROSION-CORROSION MODEL……………………………. 136

General Approach……………………………………………..…….….. 136

Proposed Procedure for Estimating Ce-c……………..……………….. 139

Determination of the indented open area………………………… 142

Implementation of the second order model to determine the total

current……………………………………………………………. 143

Validation of the Erosion-Corrosion Prediction Model……………….. 150

Adjustment of the erosion prediction……………………….……. 150

Comparison between experimental data and predictions……….. 151

Some trends of predicted values………………………………..… 154

Effect of sand rate: comparison between experiments and

prediction trends……………………………………………..…. 154

Effect of temperature: comparison between experiments and

prediction trends…………………………………………….….. 159

Effect of material: comparison between experiments and

prediction trends…………………………………..……………. 160

xii

CHAPTER 10: SUMMARY, CONCLUSIONS AND

RECOMMENDATIONS………………………………………… 163

Summary……………..………………………………………………….. 163

Conclusions………………………………………………………………. 165

Scratch Tests…………………………………………….……….. 165

Multiphase gas/liquid/sand flow loop tests………………………. 166

Single phase liquid/sand flow loop testing (submerged direct

impingement test)……………………………………………… 168

Erosion-corrosion predictive procedure and model……………. 171

Recommendations………………………………………………………. 172

REFERENCES………………………………………………………………… 177

APPENDIX A………………………………………………………………….... 183

APPENDIX B………………………………………………………………….. 196

APPENDIX C……………………………………………………………….…. 198

APPENDIX D……………………………………………………………….…. 212

APPENDIX E…………………………………………………………………. 228

APPENDIX F …………………………………………………………….…… 233

xiii

LIST OF TABLES

Table 4-1. Test conditions for erosion and erosion-corrosion tests. .............................. 68

Table 4-2. Test conditions for erosion and erosion-corrosion tests ............................... 74

Table 4-3. The chemical composition of 13Cr. .............................................................. 75

Table 4-4. The chemical composition of Super 13Cr. ................................................... 75

Table 4-5. The chemical composition of the 22Cr. ........................................................ 75

Table 5-1. Repassivation times in minutes for 13Cr at different test conditions

using current data time series approach. ..................................................... 95

Table 5-2. Repassivation times in minutes for 13Cr at different test conditions

using second order kinetics approximation. ................................................ 96

Table 5-3. Repassivation times in minutes at pH = 4 and the three temperatures

for Super 13Cr and 22Cr using both approaches, current data time

series approach and second order kinetics approximation. ......................... 96

Table 6-1. Erosion-corrosion (EC), pure erosion (E) and corrosion components

(Ce-c) of the erosion-corrosion penetration rates for 3 alloys tested at

similar conditions. 76oF, pH 4, Vsg = 20 ft/s, Vsl = 1.4 ft/s, 30 lb/day

of sand. ...................................................................................................... 111

Table 6-2. Erosion-corrosion (EC), pure erosion (pureE) and corrosion

components (CEC) of the erosion-corrosion penetration rates for 3

xiv

alloys tested at similar conditions. 150oF, pH 4, Vsg = 20 ft/s, Vsl =

1.4 ft/s, 30 lb/day of sand . ........................................................................ 112

Table 7-1. Erosion-corrosion penetration rates and corrosion components (Ce-c)

of the erosion-corrosion penetration rate for 13Cr at different

temperatures. (Vl = 15 ft/s (4.6 m/s), PCO2=50 psig (344.7 kPa),

Brine 3%, about 3,500 lb sand /day (1,587.6 kg sand /day)). ................... 116

Table 7-2. Comparison of relative severity of the corrosion component of the

erosion-corrosion between single phase liquid loop tests and scratch

tests for 13Cr. ............................................................................................ 117

Table 7-3. Actual and predicted values of the corrosion component (Ce-c) of the

erosion-corrosion penetration rates for 3 alloys tested at similar

conditions (76oF, pH 4) ............................................................................. 121

Table 7-4. Actual and predicted values of the corrosion component (CEC) of the

erosion-corrosion penetration rates for 3 alloys tested at similar

conditions (150oF, pH 4). .......................................................................... 121

Table 8-1. Summary of corrosion component rates of erosion-corrosion Ce-c;

erosion-corrosion rates, E-C; and pure Erosion rates E, obtained

experimentally. 150F and 4.3 pH. ........................................................... 135

Table 8-2. Summary of experimental values obtained for the current Io. 150F

and 4.3 pH. ................................................................................................ 135

xv

LIST OF FIGURES

Figure 2-1. Corrosion process showing anodic and cathodic current

components.33 ................................................................................................ 9

Figure 2-2. Standard anodic polarization curve (430 stainless steel in 1 N

H2SO4) showing a typical active passive transition behavior. .................... 19

Figure 2-3. Schematic of active-passive transition. Potentiostatic anodic curve

“jklm”; hydrogen evolution reaction, curve “hi”; low concentration

of dissolved oxygen, curve “defg”; high concentration of dissolved

oxygen, curve “abc”. ................................................................................... 21

Figure 2-4. Theoretical and actual potentiodynamic plots of active passive

metals. ......................................................................................................... 23

Figure 2-5. Decay of passive corrosion rate measured by potentiostatic current.

31 .................................................................................................................. 25

Figure 2-6. Log-log plot of data from Figure 2-5 at extended times. 31 ......................... 26

Figure 2-7. Effect of chromium content on anodic polarization of Fe-Ni alloys ............ 29

Figure 2-8. Major categories of wear based on their fundamental mechanisms. 49 ....... 37

Figure 2-9. Schematic of the effect of the impingement angle on erosion rate of

ductile and brittle materials......................................................................... 40

Figure 2-10. Schematic of polarization curves for type 316 stainless steel

showing the effect of percent solids sand slurry.38...................................... 46

xvi

Figure 2-11. Plot of the mass loss rate vs. normalized distance from the pipe

expansion at different flow rates. (60oC, 3 bar CO2 and 1000 ppm of

sand (Re number 5.7 105 correspond to a flow velocity of 5.7 m/s).

52 .................................................................................................................. 47

Figure 2-12. Influence of various sand concentrations on the mass loss rate along

the pipe length at a flow rate of 3.5 m/s. 52 ................................................. 48

Figure 2-13. Thickness loss versus time for a 13Cr alloy exposed to CO2

saturated brine containing sand particles. 53................................................ 49

Figure 2-14. Comparison of penetration rates for pure erosion and erosion-

corrosion at different temperatures by the weight-loss method

(single phase liquid flow, Vl =15 ft/s). 58..................................................... 52

Figure 2-15. Penetration rates for pure erosion and erosion-corrosion tests of

13Cr in multiphase flow conditions (Vsg = 97 ft/s, Vsl = 0.2 ft/s).57........... 53

Figure 2-16. Comparison among penetration rates for pure Corrosion, pure

Erosion and Erosion-Corrosion processes (Vsg= 97 ft/s, Vsl= 0.2

ft/s). 59 .......................................................................................................... 54

Figure 4-1. Layout of the electrochemical cell for Scratch Test.................................... 62

Figure 4-2. Set-up for electrochemical measurements for Scratch Test...........................63

Figure 4-3. Schematic of the Erosion-Corrosion Flow Loop ...........................................66

Figure 4-4. Photograph of the test section of the Erosion-Corrosion loop with

the ER probe set in place (left). Sensor element of the 13Cr ER

Probe (right). ................................................................................................67

Figure 4-5. Schematic of the Microloop. ........................................................................72

xvii

Figure 4-6. Schematic of the test cell section for the Microloop indicating

positions of target working electrode (WE1) and auxiliary working

electrode (WE2). ...........................................................................................73

Figure 4-7. Schematic of the direct impingement test cell for the Microloop ................73

Figure 5-1. Anodic current decay during the scratch repassivation process. .................. 77

Figure 5-2. Data showing a linear relation between 1/I and time.................................... 78

Figure 5-3. Comparison between TLactual and TLcal for 13Cr alloy. ........................... 80

Figure 5-4. Effect of pH on the decay of the anodic current after a scratch has

been made on the surface of a 13Cr Alloy.................................................. 83

Figure 5-5. Effect of pH on the decay of the anodic current after a scratch has

been made on the surface of a Super 13Cr Alloy........................................ 83

Figure 5-6. 1/I vs. t for 13Cr from the raw scratch test data in Figure 5-4...................... 84

Figure 5-7. 1/I vs. t for 13Cr from the raw data in Figure 5-5......................................... 85

Figure 5-8. Effect of temperature on the decay of the anodic current after a

scratch has been performed on the surface of a 13Cr Alloy. ...................... 86

Figure 5-9. Effect of temperature on the cumulative thickness loss experienced

by 13Cr after being scratched at pH=4.0..................................................... 87

Figure 5-10. Effect of temperature on the decay of the anodic current after a

scratch has been performed on the surface of a Super 13Cr Alloy. ............ 87

Figure 5-11. Effect of temperature on the decay of the anodic current after a

scratch has been performed on the surface of a 22Cr Alloy. ...................... 88

Figure 5-12. 1/I vs. t for 13Cr from the raw data in Figure 5-8. ...................................... 89

Figure 5-13. 1/I vs. t for 13Cr from the raw data in Figure 5-10..................................... 89

xviii

Figure 5-14. 1/I vs. t for 13Cr from the raw data in Figure 5-11..................................... 90

Figure 5-15. Decay of the anodic current for three different alloys................................. 91

Figure 5-16. Comparison of 1/I vs. t for 13Cr and Super 13Cr at 150°F, 22Cr at

pH 4 is also included. .................................................................................. 91

Figure 5-17. Comparison of the cumulative thickness loss after 200 seconds for

the three CRAs at pH = 4 and different temperatures. ................................ 93

Figure 5-18. Effect of temperature on the ratios of the Thickness Loss of 13Cr to

Super 13Cr and to 22Cr............................................................................... 94

Figure 5-19. Effect of pH and temperature on the repassivation times of 13Cr at

different test conditions. Data for Super 13Cr and 22Cr at pH 4 and

the three temperatures are also included. .................................................... 98

Figure 5-20. Effect of pH and temperature on the repassivation times of Super

13Cr at different test conditions. Data for 22Cr at pH 4 and the three

temperatures are also included. ................................................................... 98

Figure 5-21. Comparison of the repassivation times for the three CRAs at pH 4

and different temperature. ........................................................................... 99

Figure 5-22. Comparison between repassivation times obtained from actual data

and estimated from the second order kinetic approximation. ..................... 99

Figure 6-1. Pure erosion test for 13Cr under sand-N2-distilled water flow system

(Vsg=60 ft/s, Vsl= 0.2 ft/s) at 150oF and low sand rate (15 lb/day). .......... 103

Figure 6-2. Erosion-corrosion test for 13Cr under sand-CO2-Brine flow system

(Vsg=60 ft/s, Vsl= 0.2 ft/s) at 150oF and low sand rate (15 lb/day). ........ 104

xix

Figure 6-3. Penetration rate for pure Erosion and Erosion-Corrosion tests of

13Cr in multiphase flow testing at 76oF. (The errors bars on the

average represent the 95% confidence interval on the mean value.) ........ 105

Figure 6-4. Penetration rate for pure Erosion and Erosion-Corrosion tests of 13Cr

in multiphase flow testing at 150oF. (The errors bars on the average

represent the 95% confidence interval on the mean value.)...................... 107

Figure 6-5. Erosion-corrosion (EC), pure erosion (pure E) and corrosion

component (CEC) of the erosion-corrosion penetration rates for data

shown in Table 6-1. ................................................................................... 111

Figure 6-6. Erosion-corrosion (EC), pure erosion (pureE) and corrosion

component (CEC) of the erosion-corrosion penetration rates for data

shown in Table 6-2.................................................................................... 112

Figure 7-1. Corrosion component of the erosion-corrosion process for 13Cr

(actual flow liquid loop data), Super 13Cr and 22Cr (data

extrapolated with scratch data) at pH 4 and 3 different temperatures....... 118

Figure 8-1. Current response for Super 13Cr exposed to CO2 saturated brine

containing sand.......................................................................................... 125

Figure 8-2. Comparison between current responses for Super 13Cr exposed to

CO2-saturated brine at similar environmental and flow conditions

but different sand rates. ............................................................................. 126

Figure 8-3. Comparison between current responses for Super 13Cr exposed to

CO2-saturated brine at similar environmental conditions but

different flow velocities and sand rates..................................................... 127

xx

Figure 8-4. Anodic current decay for Super 13Cr alloy after sand is removed

from the test cell loop. ............................................................................... 128

Figure 8-5. A linear behavior between 1/I and time after sand is removed from

the test cell loop......................................................................................... 129

Figure 8-6. Actual data and second order model for the anodic current decay after

sand is removed from of the test cell loop. ............................................... 130

Figure 8-7. Effect of sand rate on the anodic current decay of Super 13Cr alloy. ........ 131

Figure 8-8. Effect of sand rate on the linear behavior for 1/I with time after sand

is removed from the test cell loop. ............................................................ 131

Figure 8-9. Effect of pH on the anodic current decay of 13Cr alloy. ............................ 132

Figure 8-10. Comparison of the resultant metallic surfaces of the target specimen

of 13Cr and Super 13Cr exposed to different erosion-corrosion

conditions. ................................................................................................. 134

Figure 9-1. Schematic of a particle impingement on a passive alloy. ........................... 140

Figure 9-2. Generic plot for the total current and individual current transients of

6 different impacts taking place in a numerical cell of the target

specimen.................................................................................................... 145

Figure 9-3. Generic plot for the total current generated in a cell due to total

particle impact. .......................................................................................... 146

Figure 9-4. Typical behavior of the total current produce in a numerical cell due

to the particle impact. The cell is hit at an impact frequency of 74

particles per second. .................................................................................. 147

Figure 9-5. Flow diagram to estimate E, Ce-c, & EC .................................................. 149

xxi

Figure 9-6. Comparison between measured erosion and erosion-corrosion

penetration rates (by weight loss) and those predicted by the

proposed model. ........................................................................................ 151

Figure 9-7. Comparison between measured corrosion component penetration

rates (by LPR) and those predicted by the proposed model...................... 153

Figure 9-8. Comparison between measured total current values (A) and those

predicted by the proposed model. ............................................................. 153

Figure 9-9. Prediction trends with sand rate of pure erosion, E for several flow

velocities.................................................................................................... 155

Figure 9-10. Prediction trends with sand rate of pure erosion, E and corrosion

component of erosion-corrosion, Ce-c for Super 13Cr at two flow

velocities.................................................................................................... 156

Figure 9-11. Comparison between 13Cr and Super 13Cr prediction trends with

sand rate of pure erosion, E and corrosion component of erosion-

corrosion, Ce-c. ......................................................................................... 157

Figure 9-12. Effect of pH and temperature on the prediction trends with sand rate

of the corrosion component of erosion-corrosion, Ce-c for 13Cr. ............ 158

Figure 9-13. Comparison between experimental data and predicted trends with

temperature of the erosion-corrosion of 13Cr. All data normalized to

room temperature values. .......................................................................... 160

Figure 9-14. Comparison between experimental data and predicted trends with

temperature of the erosion-corrosion of Super 13Cr. All data

normalized to room temperature values. ................................................... 161

xxii

Figure 9-15. Comparison between experimental data and predicted trends with

temperature of the erosion-corrosion of 22Cr. All data normalized to

room temperature values. .......................................................................... 161

Figure 9-16. Comparison between experimental data and predicted trends of the

erosion-corrosion, EC of the three CRAs at two different

temperatures. All data normalized to 13Cr at room temperature

value. ......................................................................................................... 162

1

CHAPTER 1

INTRODUCTION

Corrosion has an important impact on the total cost of oil and gas production. Its

direct and indirect costs in the US were estimated to be several billions of dollars,

according to a 1999 U.S. Congressional study.1 In particular, initial purchase costs and

maintenance costs for controlling corrosion of production tubing, pipelines and other

equipment, is one of the oldest material performance problems facing the oil and gas

industry.

From the diversity of corrosion related problems, carbon dioxide (CO2) corrosion

of carbon steel is probably the material degradation mechanism most extensively

assessed in this industry for the last 30 years. During this time, many models directed

towards predicting the physics involved in the CO2 corrosion process of carbon steel have

been established. Empirical laboratory models, empirical field models and mechanistic

models have been developed in this area, and many parameters have been taken into

account, such as effects of CO2 pressure, temperature, pH, chloride content, and

hydrodynamics among others. Today, CO2 corrosion is a well understood phenomenon;

however, the dynamics of the oil business have led the oil and gas industry to the

2

production of wells of greater and greater depths with increasing severity of the service

conditions.

Sometimes, it may be more economical in the long term to use corrosion

resistance alloys (CRA) instead of the normally used carbon steel, for which expensive

chemical treatment programs are often required. Recently, we have witnessed the

increasing use of CRAs in the oil and gas industry. In a sweet environment,2 the most

widely used corrosion resistance alloy is 13Cr and its modified types. The main reasons

are its excellent corrosion resistance in a CO2 corrosion environment and its low cost

compared with other CRAs such as duplex stainless steel.3

Numerous research papers were published in recent years to investigate the

corrosion behavior of 13Cr or its modified types at different service conditions.4-21 The

13Cr alloy has an excellent CO2 corrosion resistance. In fact, the effects of high gas

velocity and corrosion by CO2 experienced with carbon steels have been reduced to low

levels or virtually eliminated by alloying with 12 percent or more chromium.22 However,

when sand particles are entrained in the flow, the metal loss mechanism is different.

Recently, the need has arisen to define safe service limits for utilization of such materials

in corrosive oil and gas environments which may contain sand particles. However, little

work has been done to investigate the joint effect of erosion and erosion-corrosion of

such alloys.

Erosion by solid particle impingement, without any additional chemical

degradation components, is a complex problem by itself. The American Petroleum

Institute standard, API RP-14E, has been used for many years as the main guide for

designing and operating oil and gas production piping systems based on an estimation of

3

limiting erosional velocities. Its applicability and limitations have been widely discussed

and published.23-30 Researchers have done extensive experimental and modeling work to

generalize and improve estimates of erosional velocities when solid particles are

suspended in the flow. When CO2 corrosion and solid particle impingement are acting

together on the metallic surface, the metal loss rates experienced are often significantly

higher than those observed when pure corrosion and pure erosion are taken separately.

The mixed degradation mode inherent in the sand-erosion/CO2-corrosion

mechanism does not allow the study to focus on only one mechanism. Besides,

dominating parameters driving the material degradation may be different from the CO2

corrosion or pure erosion mechanisms by themselves. Hence, extensive research work is

required to address this problem. Some work also has been directed toward obtaining a

better understanding of the combined erosion-corrosion process.

Addressing these needs is challenging, especially when taking into account both

the complexity of the chemi-mechanical mechanism presented and the diversity of

service conditions found in oil and gas production. The complexity of the mechanisms

involved in erosion-corrosion of metals, based on the large amount of variables involved,

has been remarked on by several authors.

This research work has been directed towards the study of the effect of sand on

the erosion-corrosion of CRAs, including 13Cr, Super 13Cr (S13Cr) and 22Cr. The

effects of flow conditions, sand rate, pH and temperature on the characterization of the

erosion-corrosion of 13Cr were widely studied, and a direct comparison of 13Cr vs.

Super 13Cr and 22Cr under similar erosion-corrosion conditions was also made. To

characterize the erosion, corrosion and erosion-corrosion behavior of the mentioned

4

CRAs, several experimental techniques were utilized such as weight loss (WL), electrical

resistance (ER), linear polarization resistance (LPR), potentiodynamic scan (PD), and

scratch test (ST) which may be seen as a modified electrochemical noise technique. The

latter, along with computational fluid dynamics (CFD) simulations, provided the

foundation for a procedure built to predict erosion-corrosion rates of CRAs in brine flows

containing sand. Lastly, experimental results and predicted values were compared for a

broad range of conditions with encouraging results.

5

CHAPTER 2

BACKGROUND AND LITERATURE REVIEW

Basic Corrosion Concepts

Concept and forms of corrosion

The material degradation of a metal as a consequence of a chemical reaction

between a metal and its environment is called corrosion. Metal atoms are normally

present in nature as minerals (chemical compounds). The amount of energy needed to

extract metals from their minerals during chemical reactions is the same as that involved

in the corrosion process. Most of the time, corrosion occurs spontaneously, returning the

metal to its combined state in chemical compounds that are similar to the mineral from

which the metals were extracted.31

Several forms of corrosion have been widely studied and classified by different

authors. A full discussion of all forms of corrosion is beyond the scope of this

dissertation, and particular attention will be given to research efforts in the area of

erosion-corrosion phenomena. However, some of these forms are listed below.

Uniform Corrosion Galvanic Corrosion Crevice Corrosion Pitting Corrosion

6

Environmentally Induced Cracking Intergranullar Corrosion Dealloying Erosion-Corrosion

In general, uniform corrosion may be considered as the more common type of corrosion

found as well as the easiest to predict and control. The different types of localized

corrosion are more insidious and difficult to predict and control. While localized

corrosion may not consume as much material, penetration and failure are more rapid and,

hence, is generally considered a more severe type of corrosion.

Uniform loss of metal is also a common form of corrosion observed in the oil

industry. However, in oilfield operations, metal loss is frequently localized in the form of

discrete pits or larger localized areas. Additionally, stages of incipient corrosion in metals

often provide suitable conditions for metal cracking without perceptible loss of material,

and then metallurgical factors become predominant.32 Some alloys often used in the oil

industry owe their corrosion resistance to the formation and persistence of a protective

layer. Removal of this layer at local areas can lead to accelerated attack. Extremely high

velocity flow and intense turbulence may erode away the protective layer to expose fresh

metal which then can be corroded at a faster rate. Whether localized or uniform

corrosion, nearly all metallic corrosion processes involve transfer of electronic charge in

aqueous solutions so their electrochemical nature is the same for all forms of corrosion. It

is thus important to discuss this process.

7

Review of the Electrochemical Basis of Corrosion

Most metal corrosion occurs by means of electrochemical reactions taking place

at the interface between the metal and an electrolyte solution. Corrosion normally occurs

at a rate determined by equilibrium between opposing electrochemical reactions. The

first is the anodic reaction, in which a metal (M) is oxidized, releasing electrons (e-) and

is represented by,

neMM n (2-1)

The other is the cathodic reaction, in which a solution species (often H+ or O2) is reduced,

removing electrons from the metal. Equilibrium between hydrogen gas and acid solutions

is represented by,

222 HeH (2-2)

or a reaction equivalent to Eq. (2.2) in neutral or alkaline solutions.

OHHeOH 222 22 (2.3)

As the potential becomes more positive, another reaction involving water becomes

thermodynamically feasible. In acid solutions, oxygen reduction is represented by

OHeHO 22 244 (2.4)

while in neutral or alkaline solutions

.442 22 OHeOHO (2.5)

8

When these two anodic (equation (2-1)) and cathodic (equations (2-2), (2-3), (2-4) or

(2.5)) reactions are in equilibrium, the flow of electrons from each reaction is balanced,

and no net electron flow (electrical current) occurs. Both reactions can take place on a

single metal or on two different electrically connected metals.31

Electrochemical reactions like those described by equations (2-1) and (2-2)

proceed only at limited rates. If electrons are made available in abundance, according to

equation (2-1), the potential at the surface becomes more negative. This means that an

excess of electrons (with their negative charges) has accumulated at the metal/solution

interface. Not all the available electrons waiting at the surface can be accommodated

because the reduction reaction is too slow. This negative potential change is called

cathodic polarization. Similarly, a deficiency of electrons in the metal/solution interface

produced by a positive potential change is called anodic polarization. As the deficiency

(polarization) increases, the tendency also increases for anodic dissolution. Anodic

polarization can be thought of as a driving force for corrosion by the anodic reaction,

equation (2-1). When the surface potential measures more positive, the oxidizing power

of the solution increases because the anodic polarization is greater.31

Theory behind polarization measurements

Figure 2-1 illustrates the process mentioned above. The ordinate represents

potential and the abscissa represents the logarithm of absolute current. According to the

mixed potential theory, the theoretical currents for the anodic and cathodic reactions vary

linearly with surface potential and are shown as straight lines in Figure 2-1.33 The curved

line is the actual current as would be experimentally measured, which represents the total

current that is the sum of the anodic and cathodic currents. This is the current measured

9

during a sweep of the potential of the metal using a potentiostat. The sharp point in the

curve is actually the point where the current changes sign as the reaction changes from

anodic to cathodic, or vice versa. Because of the passivity phenomenon, the current may

change by six orders of magnitude during a corrosion experiment using a metal that

exhibits an active-passive transition.33

The equilibrium potential achieved by an electrically isolated metal immersed in a

solution electrolyte is called the open circuit potential, or free corrosion potential, or rest

corrosion potential, (Eoc or Ecorr).

Figure 2-1. Corrosion process showing anodic and cathodic current

components.33

The value of the current at Ecorr is called the corrosion current, Icorr. Notice at

Ecorr the anodic current equals the cathodic current, thus the measured current tends to

zero as shown in Figure 2-1. Unfortunately, Icorr cannot be measured directly. If one could

measure Icorr, estimation of the corrosion rate of the metal by means of Faraday’s Law

10

would be straight forward. However, Icorr can be estimated using electrochemical

techniques.

Icorr and corrosion rates are a function of many variables including type of metal,

metal history, surface finishing, solution composition, solution pH, temperature,

dissolved gases, hydrodynamics of the system, and many others. In practice, many metals

form an oxide layer on their surface as they corrode. If the oxide layer significantly

inhibits the corrosion process, the metal is said to passivate. In some cases, local areas of

the passive film break down allowing significant metal corrosion to occur in a small area

leading to localized corrosion with higher localized penetration rates.

Reactions in the corrosion process are electrochemical. Thus electrochemical

techniques are ideal for the study of the corrosion processes. In electrochemical studies,

a metal sample with a surface area of a few square centimeters is used to model the metal

in a corroding system. This metal sample is often called the working electrode. The metal

sample is immersed in a solution that simulates the real environment of the system

studied. For the standard three-electrode system, two more electrodes are immersed in the

solution, namely the reference electrode and the counter electrode (also called the

auxiliary electrode). The reference electrode is commonly a low polarizable, and very

stable, electrode which serves as a reference for the metal sample potential

measurements. The counter electrode is commonly an inert electrode that closes the

electrical circuit. It serves as the substrate where the cathodic reaction occurs while the

anodic reaction is taking place in the working electrode (metal sample) or vice versa. All

the electrodes are connected to a device called a potentiostat. A potentiostat allows you

11

to change the potential of the metal sample in a controlled manner and measure the

current that flows as a function of the applied potential.

Controlled potential experiments, such as linear polarization resistance (LPR) and

potentiodynamic polarization scan (PD), are used to perturb the equilibrium corrosion

process by polarizing the sample. The response (current) of the metal sample due to the

polarization is measured. Some models of the sample’s current behavior have been

developed to estimate corrosion rates from the current response.33

Linear polarization resistance

In the previous section, it was pointed out that Icorr cannot be measured directly.

In many cases, one can estimate it from current versus voltage data. As mentioned before,

current can be measured as potential increases, and a log current versus potential curve

can be plotted over a range of about one half volt. The voltage scan is centered on Ecorr.

Then the measured data can be fitted to a theoretical model of the corrosion process.

The model usually used for the corrosion process assumes that the rates of both

the anodic and cathodic processes are controlled by the kinetics of the electron transfer

reaction at the metal surface. An electrochemical reaction under kinetic control obeys the

Tafel equation as follows.33

)(303.2

exp 00

EEII (2-6)

where, I is the current resulting from the reaction, Io is a reaction dependent constant

called the exchange current, E is the electrode potential, Eo is the equilibrium potential

12

(constant for a given reaction) and is the reaction's Tafel constant with units of

volts/decade (constant for a given reaction).

The Tafel equation describes the behavior of one isolated reaction. In a corrosion

system, there are two opposing reactions (anodic and cathodic). The Tafel equations for

the anodic and cathodic reactions in a corrosion system can be combined to generate the

Butler-Volmer equation (equation. (2-7)).

c

corr

a

corrcorr

EEEEII

303.2

exp303.2

exp (2.7)

where I is the measured cell current in amps; Icorr is the corrosion current in amps; E is

the electrode potential; Ecorr is the corrosion potential in volts; a and c are the anodic

an cathodic beta Tafel constants respectively.

When E = Ecorr, the value of each exponential term is unity, therefore the

exponential terms do not contribute to the cell current. But, near Ecorr, both exponential

terms can contribute significantly to the overall current. Finally, at potentials that are far

from Ecorr, one exponential term predominates and the other term can be ignored. For this

case, a plot of potential versus log current approaches a straight line. If a log current

versus potential curve is linear on both sides of Ecorr then it is likely that the system under

study is under kinetic control which is the case for many corrosion systems seen in

industrial and laboratory applications. However, deviations from this behavior have been

reported, such as the following: 33

13

Concentration polarization: This is the condition for which the rate of a reaction is

controlled by the rate at which reactants arrive at the metal surface. Cathodic reactions

often show concentration polarization at higher currents for which diffusion of the

oxygen or hydrogen ion is slower than the kinetically controlled reaction rate.

Oxide formation: Formation of oxides on the surface of the metal, which may or

may not lead to passivation, can alter the surface of the sample being tested. The original

surface and the altered surface may have different values for the Tafel constants.

Mixed control process: where more than one cathodic, or anodic, reaction occurs

simultaneously may complicate the model. An example of this is the simultaneous

reduction of oxygen and hydrogen ion in aerated acid solutions.

Other effects that alter the surface, such as preferential dissolution of one alloy

component, or adsorption of charged chemical species to the metallic surface, can also

cause problems in interpreting electrochemical measurement data.

In classic Tafel analysis, linear portions of a potential versus log current plot are

extrapolated back to their intersection (see Figure 2-1). The value of either the anodic or

the cathodic current at the intersection is Icorr. Unfortunately, for many real world

corrosion systems, a sufficient linear region to permit accurate extrapolation is not

observed. Therefore, this type of analysis is not often used to estimate corrosion rates but

rather to learn something about the mechanism driving the corrosion process.

Equation (2-7) can be further simplified by restricting the potential to be very near

to Ecorr. For potential E sufficiently close to Ecorr, the current versus voltage curve

approximates a straight line. The slope of this line is called the polarization resistance,

14

Rp, because it has the units of resistance (ohms). An estimate of the corrosion current can

be obtained by combining the Rp value with estimates of the beta coefficients.

Stern and Geary were the first to approximate the exponential terms in equation

(2-7) with the first two terms of a power series expansion (y = 1+ x +x2/2...). Simplifying,

the equation becomes:

ca

ca

pcorr R

I

303.2

1(2-8)

The ASTM Standard-G-5928 has standardized the estimation of Rp by

electrochemical measurements. This standard provides a “Standard Test Method for

Conducting Potentiodynamic Polarization Resistance Measurements.”

In a polarization resistance experiment, the specimen (working electrode)

potential is swept over a small range around Ecorr (generally 10 mV) while the response

of the current is recorded. Icorr is divided by the area of the working electrode to obtain

the current density denoted by icorr. The Polarization Resistance, Rp, of a corroding

system is the slope of the potential vs. current density plot at current density equal to

zero. Notice that data from a polarization resistance experiment do not provide any

information about the values for the beta (Tafel) coefficients. Therefore beta values must

be provided to determine Icorr by using equation (2-8). These beta values are either

obtained from a Tafel curve experiments or often estimates are used based on previous

testing conducted at similar conditions.

Units of rate of penetration are useful for engineering applications: hence, Icorr

units of amperes are converted to penetration rates units by means of Faraday’s Law.

15

Assume an electrolytic dissolution reaction involving a chemical species, S:

neSS n (2-9)

According Faraday's Law, current flow is related to mass via.

nFMQ (2-10)

where, Q is the charge in coulombs resulting from the reaction of species S, n is the

number of electrons transferred per molecule or atom of S, F is Faraday's constant

(96,486.7 coulombs/mole), and M is the number of moles of species S reacting.

The equivalent weight (EW) is the mass of species S that will react with one

96,500 coulombs (Faraday’s constant) of charge. For an atomic species, EW = AW/n

(where AW is the atomic weight of the species). Recall that M = W/AW and substituting

into equation. (2.10) gives:

FEWQW (2-11)

where W is the mass of species S that has reacted.

If corrosion occurs uniformly across a metal surface, the corrosion rate can be

calculated in units of distance per year. Corrosion rate, C can be obtained from weight

loss data in a simple form by applying equation (2-12). The metal density, , and the

sample area, A, are needed as input. For constant current, the charge is given by Q = I t,

where t is the time in seconds and I is a current. The current density is calculated as i =

I/A. Substituting the value of Faraday's constant, Eq. (2-11) becomes.

EWi

kC corr (2.-12)

16

where C is the corrosion rate, icorr is the corrosion current density in A/cm2, EW is the

equivalent weight in grams/equivalent; is density in grams/cm3, k is a constant that

defines the units for the corrosion rate, k = 3.272 x 10-3 for mm/year and k = 0.1288

millinches/year (mpy).

The equivalent weight for a complex alloy undergoing uniform dissolution can be

calculated as the weighted average of the equivalent weights of the alloys components.

Mole fraction is used instead of mass fraction as the weighting factor. The estimation of

corrosion rate by this means has been widely accepted and ASTM Standard-G-10234,

“Standard Practice for Calculation of Corrosion Rates and Related Information from

Electrochemical Measurements”, which can be consulted for further information.

Potentiodynamic polarization and Tafel constants

Potentiodynamic polarization curves are used to determine corrosion behavior of

metal specimens in aqueous environments by studying their current-potential

relationships. The specimen potential is scanned slowly, as in linear polarization

resistance experiments, but over a larger range of potentials. A complete current-potential

plot of a specimen can be measured in a few hours or in some cases in a few minutes

depending on the scan rate and the potential range to be covered.

Suppose the potential is forced from Ecorr to more positive values (anodic region)

by using a potentiostat. That is moving towards the top of the graph in Figure 2-1. This

will increase the rate of the anodic reaction, ia, (that means increase the corrosion rate)

and decrease the rate of the cathodic reaction, ic. Since the anodic and cathodic reactions

are no longer balanced, a net current, imeasured will flow from the counter electrode

17

(through the electronic circuit) into the metal sample (working electrode). The sign of

this current is positive by convention. If the potential is taken far enough from Ecorr the

cathodic reaction current will be negligible and the measured current (imeasured) will

represent the anodic reaction alone. In Figure 2-1, notice that the curves for the measured

current and the theoretical anodic current lie on top of each other at very positive

potentials. Conversely, at strongly negative potentials, cathodic current dominates the

measured current. 33

Theoretically, anodic and cathodic polarization curves are symmetrical about

Ecorr, this meansa= c. However, for a corroding metal, a and c are rarely

equal. In general a for an anodic dissolution reaction is usually half of c or less.31 A

common practice for estimating the Tafel constants in a corroding system is by

comparing the corrosion rate obtained from LPR measurement with the corrosion rate

obtained directly from weight loss data if available. In this way the LPR measurement is

calibrated and the instantaneous corrosion rate can be reliably monitored.

Active Passive Metal Behavior

In certain cases as the potential is increased, the metal will be first passivated as

opposed to corroding faster. Passivation phenomena have been widely studied for many

years. The current use of commercially available potentiostatic systems has contributed

greatly to the understanding of active-passive behavior of certain metals and alloys.

Faraday was among the first to experiment with active-passive behavior of some

metals in the 1840s. He suggested that passivation is caused by an invisible oxide film on

the metal surface, or by an oxidized state of the surface, that prevents contact between the

18

metal and the solution.31 This theory has been supported and extended by a large volume

of experimental evidence accumulated since it was first proposed. However, conflicting

theories have suggested allotropic modifications, bulk oxide, adsorbed oxygen, adsorbed

OH-, and adsorbed anions as the source of passivity. The conflicts in theory have yet to

be totally resolved. 31, 35

Metals and alloys capable of forming a passivation layer display distinctive

behavior as potential and anodic polarization increase as shown in Figure 2-2. The curve

in Figure 2-2 was obtained from a test conducted by following the ASTM Standard G536

(430 stainless steel in 1 N H2SO4). Typically, in deaerated acid solutions, corrosion rates

are high and increase further with potential in the active state. Nevertheless, the passive

film becomes stable at potentials above primary passive potential, Epp, and the corrosion

rate falls to very low values. Passive current is usually lower by as much as 103 to 106

times than Icorr in the active state. At still higher potential, the passive film breaks down

and the anodic rate increases in the transpassive state. Depending on the potential, or

oxidizing power of the solution, an alloy may exist in the passive state above Epp, or in

the active state, below it.

If the steady-state potential (Ecorr) is already higher than Epp then the critical

current icrit and primary passivation potential Epp will not appear on an anodic

polarization curve. In this case, the full polarization curve cannot be constructed from

potentiostat data, and weight loss and solution analysis techniques must be used. 31

The two significant parameters in passivation are Epp and icrit, and although they

are evaluated by means of the potentiodynamically determined anodic E-i curve, they are

19

equally applicable to chemical passivation in which the redox potential and kinetics of

the cathodic reaction determines the potential of the metal/solution interface. 31

Iron (Epp = 0.58 V vs. SHE) can be spontaneously passivated in nitric acid, an

oxidizing acid of very high redox potential (about 1.1 V vs. SHE) and high limiting

current density. The powerful oxidizing agent fuming nitric acid can passivate iron

chemically, but in reducing acids like H2SO4 it can be only passivated by raising its

potential into the passive region by means of an external e.m.f. (anodic protection). 35

-0.8

-0.3

0.3

0.8

1.3

1.8

1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05

i (A/cm2)

Evs

SC

E

Passive region

Active regionEcorr

ip

Ebd

Transpassive region

Epp

icrit

Figure 2-2. Standard anodic polarization curve (430 stainless steel in 1 N H2SO4)

showing a typical active passive transition behavior.

However, it is well known that if iron is alloyed with metals that passivate more readily

(smaller icrit and more negative Epp) such as chromium and nickel, the ease of passivation

of these two metals is imparted to the alloy.31, 32, 35

For passivation, two conditions must be satisfied; (a) the redox potential of the

solution must be more positive than Epp and (b) the rate of the cathodic reaction must be

20

greater than icrit. 31, 35 Figure 2-3 shows diagrammatically how a typical active-passive

metal will corrode when the reduction reaction is spontaneously held at different potential

values (cases a, b and c). Case a) Curve h-i may be a typical reduction reaction case for

an acid solution-stainless steel system at oxygen free conditions (that means hydrogen

reduction, equation (2-2)). In this case, neither of the two criteria for passivation is

satisfied and the metal is active, corroding at high rate. Case b) The presence of a small

amount of oxygen in the solution which would increase the potential for the interface

metal/solution (curve d-e-f-g) could worsen the situation, since iL<icrit. Case c) If the

oxidizing power of the solution is increased (large supply of oxygen is enough for some

systems), complete passivation could result (curve a-b-c). 31

However, additions of dissolved oxygen cannot passivate the ferritic stainless

steels in acid solutions because iL for oxygen reduction is insufficiently large. However,

passivation could be achieved if another oxidant (HNO3, Fe3+, Cu2+) with a high limiting

current density is present in the acid. 35

Figure 2-4 presents again previous cases a and c; but now a comparison is made

between the theoretical scheme and the corresponding experimentally-determined curve.

Suppose the cathodic curve intersects the anodic curve in the active region, as in Figure

2-4a. In this case, the full active-passive transition will be experimentally shown, as

represented in Figure 2-4b. This is the typical case for 430 stainless steel in 1 N H2SO4 as

shown in Figure 2-2. As mentioned before, if the cathodic curve intersects the anodic

curve in the passive region only, as in Figure 2-4c the material will passivate

spontaneously. Experimentally, the anodic curve for a specimen that has already been

21

spontaneously passivated does not exhibit the peak-shape active to passive transition, as

shown in Figure 2-4d.

Figure 2-3. Schematic of active-passive transition. Potentiostatic anodic curve

“jklm”; hydrogen evolution reaction, curve “hi”; low concentration of

dissolved oxygen, curve “defg”; high concentration of dissolved

oxygen, curve “abc”.31

Both oxidizer concentration and solution velocity have similar effects on the

corrosion rate of an active-passive alloy. In particular, if the passive state is stable, and

corrosion rate falls to low values when the rate of cathodic reduction is made sufficiently

22

high, the criterion for passivation is in place, i.e., the passive rate is stable when the rate

(current density) for cathodic reaction, is greater than the critical anodic current density,

icrit. Based on this criterion, alloys having lower icrit and more active Epp are more easily

passivated. 31

Higher acidities and higher temperatures generally reduce the passive potential

range and increase current densities and corrosion rates at all potentials. Usually, alloying

the metal with a noble metal that has a higher exchange current density than that of the

metal to be passivated will promote passivation even in a reducing acid. This was

achieved first by Tomashov who alloyed Fe-18Cr-8Ni stainless steel with Pt, Pd or Cu. 35

Potentiodynamic techniques require that the corrosion potential be constant

during the measurement, which in practice, requires the use of very slow scan rates.

Otherwise, the actual overpotential for any given current measurement would be

erroneously estimated. That is, the potentiodynamic scan must be run slowly enough to

ensure steady-state behavior.31 Mansfeld37 has shown potentiodynamic polarization data

for AISI 430 stainless steel (without Ni) which are dependant on scan rate only in the

passive range and not in the active range.

23

Figure 2-4. Theoretical and actual potentiodynamic plots of active passive metals

(Princeton Applied Research, Application Note, Basics of corrosion

measurements, 1982).

Passive film

The study of corrosion also includes the study of the nature of corrosion products

and of their influence on the reaction rate. These may be formed naturally by reacting

with their environment or as a result of some intentional pretreatment process to enhance

24

the protective properties of films by modifying the nature of existing films. These

corrosion products frequently form a kinetic barrier that isolates the metal from its

environment and thus controls the rate of the reaction. Thickness alone does not provide a

criterion of protection; actually the kinetics of attack is related to a variety of other

factors such a composition, structure, adhesion to substrate, cohesion, mechanical

properties, etc. of the film or scale of reaction products. 35

“A lack of fundamental understanding of passive film properties has delayedthe control and prevention of localized forms of corrosion that result from breakdownof the passive film. The most conveniently determined film properties involvechemical and electrochemical measurements, which have been difficult to interpretobjectively. Unfortunately, ex-situ examinations for direct determination of structureand composition (e.g., in vacuum by electron spectroscopy) are likely to change thefilm structure by dehydration and precipitation of new phases on the surface”. 31

Uhlig and co-workers have emphasized the role of chemisorbed oxygen in

establishing passivity. They suggested that adsorbed films provide a kinetic limitation by

reducing the exchange current density, io, for the anodic reaction instead of acting as a

barrier to dissolution reaction. 31 An initial chemisorbed state on Fe, Cr and Ni has been

postulated by the same authors in which the adsorbed oxygen is abstracted from the water

molecules. A different phase oxide or other film substance emerges at thicknesses within

1-4 nm with a corresponding increase in the anodic potential. Increase in the anodic

potential may lead to the following sequence:35

oxidephasemultilayermonolayer

2 .)( MOOHMOHMM

“Other theories consider that chemisorption of oxygen is favored by thepresence of uncoupled d-electrons in the transition metals. In Fe-Cr alloys, chromiumacts as an acceptor for uncoupled d electrons from iron. When alloyed Cr is less than12%, uncoupled d electron vacancies in Cr are filled from the excess Fe, and thealloys act like unalloyed iron, which is nonpassive in deaerated dilute acid solutions.

25

Above 12% Cr the alloys are passive in such solutions, presumably becauseuncoupled d electrons are available to promote adsorption. During film thickening,metal cations are assumed to migrate into the film from the underlying metal, as wellas protons (H+ ions) from the solution”. 31

Such films usually have low ionic conductivities that restrict cation transport

through the film. Shreir35 proposed that the electronic semiconduction, however, permits

other electrode processes (oxidation of H2O to O2) to take place at the surface without

further significant film growth. However according to Jones the passive film can begin to

grow in thickness once the metal achieves a stable passive potential region above Epp, as

evidenced by decrease in ipass with time, t, asymptotically (Figure 2-5). However, Jones

emphasizes that no true steady state is ever attained since the passive current density still

falls with time, but at continuously decreasing rate, as shown in Figure 2-6.31

Figure 2-5. Decay of passive corrosion rate measured by potentiostatic current. 31

26

Figure 2-6. Log-log plot of data from Figure 2-5 at extended times. 31

Maintenance and breakdown of passivity

Once established, a passivated state can often be maintained by conditions much

less demanding than those required to produce it. Any environment that maintains the

potential of the surface above Epp while supplying the very small passive current assures

passivity. The very large current icrit demand just before passivation is not required. The

passive potential may likewise be maintained by the presence in the solution next to the

passive surface of any oxidizing agent that provides the cathodic reaction.

“Shreir has mentioned that “stainless steels are easily maintained passivebecause both their air-formed oxide films and anodically formed oxide (containing Cr(III)) are ’stronger’ and stable at more negative potentials than the correspondingfilms on iron.” 35

Any factor that produces partial or complete removal of the passivating film can

cause partial or complete breakdown of passivity, and the corrosion rate can be

27

significantly accelerated. Such removal can be provided by a different degradation

mechanisms namely electrochemical reduction or oxidation, undermining by attack on

the underlying metal at film imperfections, or at mechanical disruption. The breakdown

of the passive film due to flows containing solid particles like sand falls into the latter

classification.

Since the passivating oxide films usually consist of brittle material, they are

subject to damage when a mechanical force is applied to them (through bending,

stretching, impact, and scratching, among others). When a passivated material is used to

reduce the corrosion rates in a given corrosive system where an inevitable mechanical

damage will occur, the material can be relied on only under conditions for which it is

chemically or electrochemically self-healing. In general, the uses of passive metals such

as stainless steel are under conditions providing self-healing by chemical processes.

However, many practical cases exist where stainless steel passivated by special

treatments (nitric acid or chromate treatment instead of self-healing) is little better than

the same material with a natural air-formed oxide film. These treatments are commonly

applied when the service condition of the stainless steel involves a corrosive environment

that does not promote self-healing, and slight mechanical damage soon leads to partial or

complete depassivation. 35

Stainless Steel

Alloying elements

Chromium plays a key role in forming the passive film. Other elements can

increase the effectiveness of chromium in forming or maintaining the film, but they

28

cannot, by themselves, create the corrosion resistant properties of stainless steel.31, 38

Protective films have been observed if more than about 10.5% Cr is present in the alloy,

but the film is fairly weak at this composition and provides only mild atmospheric

protection. Increasing the chromium content to higher levels (17 to 20% in austenitic

stainless steels, or 26% in ferritic stainless steels), greatly increases the stability of the

passive film. However, greater amounts of chromium may adversely affect mechanical

properties.38

Figure 2-7 shows anodic polarization for alloys of approximately constant Ni and

variable Cr content in hot sulfuric acid. It should be note that below 12% Cr the passive

potential region is considerably restricted. 31

With chromium levels of about 13% and relatively high carbon levels, it is

possible to obtain austenite at elevated temperatures. By accelerated cooling, the

austenite transforms to martensite. Similar to carbon steels and low-alloy steels, this

strong martensite can be tempered to a favorable combination of high strength and

adequate toughness.33 This is the case of the conventional 13Cr used in the oil industry

for well completion. It has shown excellent CO2 corrosion resistance; however its

protective properties against corrosion are considerably diminished if H2S is present in

the flow.

In recent years, nitrogen, nickel, and molybdenum additions at somewhat lower

carbon levels have produced martensitic stainless steels of improved toughness and

corrosion resistance.

29

Figure 2-7. Effect of chromium content on anodic polarization of Fe-Ni alloys

Flow velocity effects

According to the nature and intensity of the flow parameters like wall geometry,

flow environments lead to corrosion types caused predominantly by either mass transport

or wear. Lotz and Heitz39 made an attempt to classify the various types of flow-dependent

corrosion in relation to existing standards for corrosion and wear.

Fluids acting on the corrosion system material/medium have mainly two effects:

a) They influence the transport of reactants and products by convection (mass transfer by

convective diffusion), and b) They create shear stresses and pressure fluctuations on the

surfaces, and cause wear. The authors distinguish three broad areas of flow-dependent

corrosion based on these effects.

30

1. Absence of forced convection (but natural convection is present).

When no flow is produced, forced convective mass transfer is not present. As a

result enrichment of hydrogen ions by hydrolysis of anodic products and the enrichment

of chloride ions (as a result of migration in the electric field) favor forms of corrosion like

pitting and crevice corrosion. Absence of some type of protective layers typical of mass

transfer processes is also observed in stagnant conditions.

2. Corrosion influenced by convective mass transfer

In this case mechanical flow effects (shear stresses) are negligible. More likely

the corrosion is mainly caused by the transport of reactants and reaction products. If the

process is material transport-controlled the flow-dependency of the corrosion system is

greater than if it is predominantly reaction-controlled.

3. Corrosion influenced by mechanical flow effects

For those conditions where mechanical forces are significant at the interface

metal/fluid, erosion and cavitation corrosion occur. A sudden increment in the corrosion

rates, for flow rates above critical values, usually indicates these types of material

degradation. Hydrodynamic parameters as well as the corrosion system define the critical

flow rate ucrit for the appearance of erosion-corrosion. Therefore, the critical flow value

may change with changes in the fluid flow parameters and changes in the corrosive

environment as well. The fluids also may contain solids or gas bubbles, and very different

metal removal rates may be obtained as a result of differing mechanical effects even for

identical flow rates.

Corrosion failures in the oil industry are normally associated with disturbed flow

conditions as a result of weld beads, preexisting pits, bends, flanges, valves, tubing

31

connections, etc. Due to the complex hydrodynamics of these conditions, attempts at

relating flow effects to steel corrosion have been only partially successful. Estimates of

the effects of the wall shear stress and steady state mass transfer in corrosion testing is

required for estimation of steel corrosion under disturbed flow conditions. The processes

that control corrosion and film formation occur in the turbulent boundary layer and

diffusion boundary layer. In this region, the chemical reactions occur as well as all

movement, to and from the pipe wall, of the chemical species involved in the corrosion

process. As a result, disturbances in the turbulent boundary layer, particularly within the

viscous region, affect the corrosion process.40

Flow pattern effects

A complex interaction of physical and chemical parameters is involved in the

effect of fluid flow on corrosion of steel in oil production. Water wetting the metallic

surface is one of the main requirements for any corrosion to occur. This is strongly

affected by the flow regime. Mass transfer and wall shear stress parameters governing

behavior in the water phase that contacts the pipe wall greatly control the effect of flow

on corrosion (flow accelerated corrosion).40

Flow accelerated corrosion has been defined as “corrosion resulting from the

effect of fluid turbulence due to flow of a fluid that does not contain solid particles.” 40

Flow turbulence occurring in most flows is responsible for accelerating the corrosion of

metal pipe in contact with a flowing electrolyte. These fluctuations increase the mobility

of the corrosive species to the metal surface as well as help to remove the products of the

corrosion reaction from the boundary layer in a more efficient manner. On the other hand,

32

for laminar flow, the effect of fluid flow is inconsequential. The physical structure of

turbulent flow is thus a primary consideration since in most oil and gas situations where

fluid flow accelerates corrosion, the flow is turbulent. 40

In general the wear caused by a flowing single-phase liquid on the surface of a

metal is small, since the wall shear stresses are small too. Significant wear damage may

appear in multi-phase flows, especially if solid particles are present in the flow causing

erosion.39

When both mechanical forces (wear) and material transport are acting together,

there are many processes involved and the interaction between those mechanisms plays a

key role in the understanding of the overall phenomenon. The surface layer may be

removed, leading to an intense local attack. Where the surface layer has been removed,

the material transport may determine the deterioration rate. More complex corrosion

mechanisms can be activated as corrosion cells between the bare surfaces (anode) and the

surfaces with an oxide film (cathode), and consequently a galvanic effect arises.

When solid particles are present in the fluid, erosion problems may arise. The

momentum of the particles carries them into contact with the metal surface. When the

impact kinetic energy is high enough, wear results. Additionally, solid particles entering

the boundary layer increase the turbulence intensity and the mass transfer may be

accelerated. However, particles that are less dense than the liquid may not involve any

additional material transport, since they do not penetrate the laminar boundary layer.39

In most cases, a relationship between the weight loss WL and the flow rate u can

be approximated with an exponential relation of the form WL ~ ua.

33

Usually the value of the exponent, a changes according to the mechanism

involved in the degradation process. For non-flow-dependent corrosion, a is equal to zero

(limiting phase boundary reaction).

Heitmann discussed this subject of erosion-corrosion in water-steam loops and

suggested that plain carbon steel corrodes at a higher rate with WL ~ u2 (erosion-

corrosion); and, low alloy steels corrode very slowly, with WL ~ u0.5 (transport-affected

corrosion). In the case of wear due to solid and liquid impingement, as well as erosion-

corrosion, exponents of 2 < a < 4 are commonly found. 39

It should be mentioned that the type of control of a corrosion process, by transport

alone, by mechanical means alone, or mixed means, depends not only on the

hydrodynamics but also on the corrosion system itself. The electrochemical or chemical

limitations and their influence on the formation and stability of surface layers are of

particular importance.

CO2 Corrosion Resistance of 13Cr Alloy

There have been several studies of the kinetics and mechanism of the cathodic

reaction in slightly acidic oxygen-free CO2-containing solutions of the type encountered

in oil and gas production, and several different reaction mechanisms have been proposed.

41-46 The majority of these studies have been in the pH range of 4 to 4.5, with the pH

controlled by the partial pressure of carbon dioxide. While there are still disagreements

about the precise mechanism, the involvement of potentially complex preceding or

subsequent chemical reactions is known to modify the electrochemical responses.

34

There have been many studies of the cathodic reaction mechanism and several

different processes and rate determining steps have been proposed. These include rate-

determining reduction of carbonic acid with subsequent reformation from the bicarbonate

and protons41 and rate limiting diffusion of bicarbonate.42 Even greater complication

would result from these mechanisms by adding the interaction of the solution with a

chromium rich layer.

The resistance of 13Cr stainless steels to uniform corrosion has been determined

by weight loss by several authors. Crolet and Bonis47 made corrosion rate measurements

after exposures of one month in laboratory autoclaves. The tests were conducted in highly

acidic, relatively concentrated brines, 30 g/L (0.5 M) NaCl, with CO2 pressure within the

range of 14 to 420 psi and at temperatures of 70 to 250oF (20 to 120oC). Despite these

severe conditions, the samples always remained passive or only very slightly active, and

the weight losses never exceeded 0.5 mpy (0.01 mm/y). Moreover the effect of flow was

not determined.

Cayard and Kane3 summarized the corrosion rate data of conventional 13Cr and

modified 13Cr from several authors. The data set represents a range of conditions with

respect to CO2, H2S, temperature and Cl-. The corrosion rates at low pH for conventional

13Cr were in general higher than the corrosion rates for modified 13Cr as expected.

However, above a pH of approximately 4.2, the corrosion rate remains quite low for both

alloys.

Many authors have investigated effects of flow velocity on corrosion behavior in

wet CO2 environments. Several of them used flow loops to establish the effect of the fluid

35

flow in the corrosion mechanism for corrosion resistant alloys (CRAs); also, comparisons

with carbon steel tested at similar conditions are available.

For the case of carbon steel, Denpo et al. showed that the corrosion rate of N80

steel obeys 0.62th-power law under turbulent conditions according to:48

398.062.0Re097.0 ScSh (2-13)

where Sh is the Sherwood number defined by

D

dkSh m (2-14)

where, km is the mass transfer coefficient, d is the diameter (characteristic length) and D

is diffusion coefficient of the specie in solution. The Reynolds number, Re, is defined by

dv

Re (2-15)

where, v is the fluid velocity, is the fluid density, is the fluid dynamic viscosity, and

the Schmidt number, and Sc is defined by

DSc

(2-16)

where, is the kinematic viscosity andD is the diffusion coefficient of the specie

in solution.

De Waard et al.41 showed that the flow dependence of the corrosion rate of carbon

steel is proportional to 0.8th- power law in turbulent flow.

30.080.0Re097.0 ScSh (2-17)

36

Therefore, the effect of fluid flow on corrosion rate could be formulated in the form of an

exponential law.

cb ScaSh Re (2-18)

where a, b, c are constant. However, the b value, which governs the flow dependence of

the corrosion rate, needs to be determined more precisely for carbon steels and low alloy

steels.48

According to Denpo, the corrosion rate of AISI 420 type (13Cr) steel obeys to a

0.5th- power law in a flow velocity range up to 3 m/s (9.84 ft/s), while at flow velocities

above 3 m/s the corrosion rate is constant. Hara et. al.48 found the CO2 corrosion rate of

AISI 420 type (13Cr) steel to be independent of flow velocity at 120 and 150oC. They

claim that even under flowing condition, the protective film is presumed to remain stable.

The flow dependence of corrosion rate for modified 13Cr steels and duplex stainless

steels still needs to be investigated.

Basic Erosion Concepts

To classify the wear on the basis of the fundamental mechanism that is operating

is a difficult task if more than one mechanism is operating at the same time. The

definition of the mechanism and the distinction between different mechanisms sometimes

are not well defined, thus, the semantic aspect further complicates attempts to classify the

phenomena. As a result, different classification schemes based on wear mechanisms have

been developed; however, no one scheme is universally accepted. Most of them have

reasonably similar features. Budinski reduces wear processes into four categories;

abrasion, erosion, adhesion, and surface fatigue as shown in Figure 2-8.49

37

Figure 2-8. Major categories of wear based on their fundamental mechanisms. 49

Solid particle erosion

The loss of material that results from repeated impacts of solid particles, carried

by a fluid, on the target surface material is called solid particle erosion (SPE). If hard

particles are entrained in a fluid flowing at any significant velocity, significant erosion

rates may arise. Fluid flow conditions affect particle acceleration and direction. When

38

particle concentration is very high, the wear phenomenon is called slurry erosion and is

generally treated as a different, though related, subject.

Ludema recently collected and analyzed a large number of analytical models of

the wear processes available in the literature50. His paper illustrated the lack of the ability

of the tribology field to develop a usable model for any wear process, sliding, abrasion, or

erosion. He used solid particle erosion to explain the problem. He reviewed 98

mathematical erosion models and discarded all but 28 of them because they were “not

amenable to conclusive scrutiny.” The 28 remaining equations have a total of 33

variables used by the various authors. No model used more than 7 of them. From 1 to 8

constants are in each model. There was no consistent pattern of use of the variables.

Some used the same variable in the numerator that others used in the denominator. None

of the models could accurately predict material loss rates.50

Because of the complexity and degree of controversy involved in the

modeling of solid particle erosion, a brief description of only the ductile and brittle

mechanism accepted by many authors is given below.

It is a common practice to classify materials as ductile or brittle, based on the

dependence of their erosion rate on , defined throughout virtually all erosion literature

as the angle between the incident particle direction and the tangent to the target material

surface. Commonly, some pure metals classified as ductile materials have a maximum

erosion rate at low angles of incidence (typically 15o to 30o), while ceramics (brittle

materials) present the maximum erosion rate at or near 90o. These cases are illustrated in

Figure 2-9. Intermediate cases between those examples exist and sometimes the same

39

material shows behavior that shifts from one extreme to the other, depending on erosion

conditions.49

Erosion of ductile materials

Hard angular particles impinging a smoother surface may cut the surface. The

theory of erosive cutting for ductile materials by Finnie (1978), distinguishes between

two types of cutting processes: (1) “The particle is stopped during the cutting action at

the depth for which its kinetic energy dissipates, or (2) The particle enters the ductile

surface and subsequently leaves it with some the remaining kinetic energy, together with

surface material.” 51

The maximum rate of material removal for ductile erosion is at a low angle of

impingement as shown in Figure 2-9. The pattern of the eroded surface often shows

evidence of material deformation in the form of particle tracks, ripples, or micro grooves

at low angles and impact craters at high angles. When the impinging particle impact

angles are 90 degrees (= 90o) or graze the surface (= 0o), the weight loss due to

ductile cutting action is negligible.

Erosion of brittle materials

Some authors have proposed that the erosive wear in brittle materials occurs when

the stress generated in the material by impacting high-velocity particles exceeds the

material’s maximum tensile strength, forming cracks on the surface. If the remaining

particle kinetic energy after the initial surface cracking is high enough, additional particle

penetration may occur. This suggested that the removal of material is dependent on the

particle velocity (v). The erosion resistance of brittle materials is nearly proportional to

40

the strain energy of the material at its tensile strength. As mentioned before, brittle

erosion presents a maximum material loss for particles striking the surface at right angles

( = 90o) as illustrated in Figure 2-9. The mechanism of material loss is through

microfracturing and removal of fractured segments by subsequent impacts.51

Figure 2-9. Schematic of the effect of the impingement angle on erosion rate of

ductile and brittle materials

Variables influencing erosion

The variables influencing pure erosion can be classified as one of three types:

impingement variables describing the particle flow, particle variables, and material

variables. The most important impingement variables are particle impact velocity (v),

41

angle of incidence (), and particle flux (particle concentration). Particle variables

include particle shape, size, hardness, and friability (ease of fracture). Material variables

include all the material properties, such as hardness, work hardening behavior, and

microstructure.49

Particle Impact Velocity: Since the kinetic energies of particles moving at high

velocities are also high, they have more potential to degrade the surface than particles

moving at lesser velocities. Hence, particle velocity is directly related to the erosion rate.

Previous work conducted with steel to determine the volume loss per impact due to

changes in velocity, reported that erosion was proportional to a simple power of

velocity.49 That is

nbvE (2.15)

where v is the velocity of erodent and n is a constant. It can be said that the value of n

usually falls in the range of 2 to 2.5 for metals and 2.5 to 3 for ceramics, even though

authors have reported values outside these ranges.

Impact Angle: As mentioned before, incidence angle or impact angle, of the

particle onto the surface of the target is another parameter which greatly influences

erosion. The influence of impact angle can also help in distinguishing between ductile

and brittle materials. For ductile materials the erosion loss is maximum at lower angles of

incidence (around 15-30o) whereas in brittle materials the erosion loss is maximum at

normal incidence around 900.

Particle flux (particle concentration): Previous research at the Erosion/Corrosion

Research Center (E/CRC) at The University of Tulsa has shown the linear proportionality

42

of erosion rates with sand concentration so long as concentrations remain low (£ 2% by

weight in liquid). Kohley and Heitz52 have reported the same behavior for 13Cr alloys.

Particle shape: The shape of the particle is also very important in predicting the

erosion rate due to particle impact. The contact area between the particle and the metal

surface during an impact is determined by the particle shape. More severe erosion

damage is generally associated with angular particles for both ductile and brittle

materials.49

Particle Size: Most theories predict no effect of particle size for metals, although

it is often observed that erosion rate increases strongly with particle size, at least up to

about 100 microns.49, 50

Particle hardness, particle friability and material variables are important variables

to be considered in erosion modeling. In 1995, Levy50 made an effort to combine all his

experimental and theoretical work and correlating it with other theories in a single

volume. In this book, the effect of the erodent characteristics as well as the material

properties of the target are well addressed but the trends very much vary with each target

material-erodent system.

Erosion-Corrosion

Lotz and Heitz39 combined corrosion standard, DIN 50900 with the wear standard

DIN 50320 to make a classification of the types of mechanical/chemical attack. They

claim that in addition to thermal/chemical attack, there is mechanical/chemical attack that

may be classified as follows:

43

Wear at the phase boundary. Stress and fatigue corrosion and forms of mechanical/chemical/thermal

attack in the material. Transport-determined corrosion in the liquid medium (hydromechanical

effects).

In a pure wear situation, whether the weight loss comes from mechanical removal

of the passive layer or removal of the base material itself would not significantly affect

the final erosion rates, whereas under corrosive conditions, damage in the surface layer

and a competition between layer removals versus layer self-healing may signify huge

differences in final weight losses. The term “erosion-corrosion” is defined in DIN 50900

as the “combination of the mechanical removal of a surface (erosion) and corrosion,

where the corrosion is initiated by the destruction of protective layers as a result of

erosion”. This standard only accounts for the mechanical damage caused by wear. The

stimulation of a corrosion process by a fluid-mechanical component is generally also

referred to as erosion-corrosion.39

The research presented in this dissertation deals with the erosion-corrosion of a

13Cr as a consequence of having mechanical wear due to the presence of sand particles

suspended in a CO2 saturated brine fluid. The mechanical wear will be occurring in the

protective layer as an initial stage and if erosivity conditions are severe enough, base

metal also can be removed. Therefore in this research, erosion-corrosion can be

understood as the combination of erosion of the protective layer (and perhaps also base

metal) and corrosion of the underlying metal, which is a common cause of failure in

oilfield equipment.

Some people use the term impingement to refer to a phenomenon similar to

erosion-corrosion but even more localized. Impingement attack occurs when a particle

44

stream impinges upon a metal surface and breaks down protective films in very small

areas (solid particles suspended in a fluid and impacting the inner metallic surface of a

pipeline). The resulting attack is in the form of pits that are characteristically elongated

and undercut in the downstream end. Impingement often results from turbulence

surrounding irregularities in the metal surface. It is particularly a problem in copper and

copper-based alloys.32

Carbon steel and low alloy steels are particularly susceptible in environments

which form scales, such as iron carbonate, that are easily removed. The attack normally

occurs only at certain areas, such as changes of sections or connections where there is

turbulence from flow or at bends and elbows.32 The austenitic stainless steels have high

resistance to both erosion-corrosion and impingement type attack; however, erosion of

the passive film can lead to some acceleration of attack. 31, 32, 38 Under severe conditions

erosion-corrosion of 13Cr alloy has been reported.52, 53

In general, it can be said that corrosion of a metal or alloy can be accelerated

when there is an abrasive removal of the protective oxide layer. The important issue here

would be how significant is the acceleration of the damage. Some authors think this form

of attack is especially significant when the thickness of the oxide layer is an important

factor in determining corrosion resistance. 38

ASTM proposes a “Standard Guide for Determining Synergism between Wear

and Corrosion.”54 The guide provides a means for computing the increased wear loss rate

attributed to synergism or interaction that may occur in a system when both wear and

corrosion processes coexist. The standard proposes four types of tests to obtain values

for: total material loss, T which accounts for pure wear and pure corrosion as well as the

45

synergistic term;. Material loss due to pure erosion, Wo; Material loss due to pure

corrosion, Co; Penetration rate due to corrosion under conditions of corrosion wear Cw.

This can be measured by electrochemical means and is usually higher than Co because of

mechanical wear interaction.

Finally, the standard proposes some calculations to determine S’ as the increase of

mechanical wear due to corrosion, S’’ as the increase of corrosion due to mechanical

wear, and S as the total synergistic effect (S = S’ + S’’). Similar approaches have been

employed by other authors.55, 56 In any case, the analysis of erosion-corrosion problems is

a difficult task. Many variables are involved. Isolating and classifying the effect of each

of them would be tedious work, and even more difficult would be determining the

relationships between the behaviors of the each variables and their integrated effect in the

mass loss rate.

Material selection is also an important consideration for erosion-corrosion

resistance. Alloy hardness has also been shown to be a factor. Generally, soft alloys are

more susceptible to erosion-corrosion than their harder counterparts; but the relative

hardness properties of the alloy can be misleading because the hardening process itself

affects resistance to erosion-corrosion.38

Some researchers38, 52, 56 have used electrochemical techniques to determine the

effects of concentration of sand and flow velocity in the erosion-corrosion of some alloys.

Figure 2-10 shows a schematic of polarization plots for type 316 stainless steel that was

exposed to sand slurries at different sand concentrations and impeller velocities. Notice

how the anodic current density increases as percent solids and speed increase. This is

pointed out by the author as an indication that more of the passive film is removed

46

because of the higher frequency of particle impacts. The oxide film on the stainless steel

is apparently adherent and fast forming. The author also claims that the polarization

curves and corrosion rates of low-alloy steels were not affected to any extent because

they freely corrode and do not form tightly adhering passive films. 38

-0.800

-0.700

-0.600

-0.500

-0.400

-0.300

-0.200

1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02

Log (current density)

Po

ten

tial

0% solids 2% solids 30% solids

Figure 2-10. Schematic of polarization curves for type 316 stainless steel showing

the effect of percent solids sand slurry.38

13Cr Alloy and Erosion-Corrosion

Kohley and Heitz52 also determined the mass loss rates of 13Cr in oil formation

water containing particles, using a loop with a constricted pipe as a test section. Figure 2-

11 shows the influence of flow velocity (Re number) on the corrosion rate along the pipe

length in front and behind the constriction in the pipe. On the x-axis, a normalized pipe

47

length x/D, with D= 50 mm, is plotted. The flow velocity is varied from 1 to 5.7 m/s in

the 50 mm diameter tube. At the entrance of the pipe constriction, the mass loss rates go

through a maximum. For example, the mass loss rate upstream of the constriction (5.7

m/s) exceeds 20 mm/year. At x/D=0.5 just downstream of the pipe expansion, a

minimum occurs, followed by a second maximum.

Figure 2-11. Plot of the mass loss rate vs. normalized distance from the pipe

expansion at different flow rates. (60oC, 3 bar CO2 and 1000 ppm of

sand (Re number 5.7 105 correspond to a flow velocity of 5.7 m/s). 52

The penetration rate at the maximum downstream of the constriction was found to

increase according to the power of 1.85 of the flow speed (WL ~ u1.85). The authors

suggested that the first maximum in Figure 2-11 could be attributed mainly to abrasive

wear and the second maximum to erosion-corrosion. They found at the inlet edge of the

48

constriction, the removal rate was far higher than the electrochemical corrosion rate,

which implies a marked mechanical abrasion of the base material.52

In sand-free flow, no differences between rates along the pipe length can be

observed. Figure 2-12 presents the dependence of the mass loss rate on the sand

concentration. At x/D = 2.5 (second max), a rise in the sand concentration is responsible

for a significant increment in the erosion-corrosion rate.

Figure 2-12. Influence of various sand concentrations on the mass loss rate along

the pipe length at a flow rate of 3.5 m/s. 52

If the mass loss rates at the point of maximum corrosion are plotted against the sand

concentration, a straight line is obtained which means that the erosion-corrosion rate is

directly proportional to the number of particles hitting the metal surface. The findings

were explained by the authors as a breakdown in passivity of 13Cr in sand-containing

formation water due to particle impacts on the metal surface. They claim high kinetic

49

energy particles not only destroy the passive layer, but also the base material, and

mechanical wear is the result. This holds for the entrance region of the constriction of the

segmented tube. Medium energy particles only destroy that passive layer, and if the

kinetics of the healing process is slower than the particle impact kinetics, erosion-

corrosion accelerates. The intensity of the corrosion is therefore a function of the

hydrodynamics of the flow. As a result, the area of maximum erosion-corrosion

downstream of the orifice can be explained as the region in which reattachment of the

flow with high particle impact frequency occurs. 52

Birchenough et al.53 studied the erosion-corrosion phenomena of 13Cr under a

very broad range of flow conditions for an expansion and for an elbow geometry in flows

containing sand. Figure 2-13 shows the results obtained by Birchenough53 et al. (1996).

Figure 2-13. Thickness loss versus time for a 13Cr alloy exposed to CO2 saturated

brine containing sand particles. 53

50

They claim the 13Cr was found to be resistant to corrosion under sand free flow

conditions. Sand was then added after an initial sand-free period according to the history

shown in Figure 2-13. Initially, erosion at a rate consistent with the pure erosion tests was

observed. After approximately 1.5 microns of material had been removed, the removal

rate increased to a value approaching 6 m/d, which was found to persist even though

sand concentration was falling. The authors believed that the sudden change in material

loss rate after approximately 1.5 microns of material loss corresponds to corrosion of the

bare metal following erosion of an oxide film previously established. Finally, they

claimed that high corrosion rates could be sustained after removal of the 1.5 microns of

oxide layer, even when sand is no longer present in the flow since the system is not able

to repassivate in an oxygen free environment. 53

In recent years, research has been conducted at The University of Tulsa, on

erosion-corrosion of 13Cr. In the past, several measurement techniques have been used to

help better understand the different mechanisms involved in the degradation of 13Cr

when exposed to a CO2 corrosive environment containing sand particles. Weight loss,

thickness loss, linear polarization resistance (LPR), and potentiodynamic sweeps were

some of the techniques used. Tests were conducted under several flow conditions. Sand

effect results in single phase liquid flow as well as two phase (gas-liquid) flows, and the

effect of temperature were also addressed. The main findings in this area are highlighted

below.

In general, for both single-phase flow and multiphase flow conditions tested, the

erosion-corrosion penetration rates of 13Cr were higher than those of pure erosion. Both

51

sets of tests were conducted in an elbow geometry where direct impingement is expected.

However, hydrodynamic conditions as well as sand concentrations for both conditions

were quite different and should be considered when trying to draw conclusions from

these results.

Single phase liquid flow loop testing (high sand rates)

Rincon et al.57 and Chen et al.58 conducted single phase liquid tests for erosion-

corrosion of 13Cr alloy at 76, 150 and 200oF. In order to be able to measure some weight

loss in a reasonable period of time, tests were conducted at the maximum liquid velocity

of the miniloop, namely 15 ft/s. However, extremely high concentrations of sand were

needed to create significant weight losses in 13Cr samples in about 7 days. The sand

concentration used was 2% by liquid weight or more (>3,447 lb/day), which might be

much higher than those typical experienced in field conditions.

At 76oF, penetration rates due to erosion-corrosion were found to be from two to

three times higher than penetration rates for pure erosion.57 At higher temperatures (150

and 200oF) the factor increased to values approximately between 4 and 6,58 as can be seen

in Figure 2-14. A synergistic effect where erosion-corrosion rates were higher than the

sum of the penetration rates due to the erosion degradation and corrosion degradation

acting alone over the metallic surface was also evidenced. However, if sand is completely

removed from the system, the oxide layer has been proven to rebuild and provide a

corrosion protection level similar to the protection afforded originally when no sand was

in the system.

52

0

2

4

6

8

10

12

14

16

18

20

Temperature

Pene

tra

tion

rate

(mpy

)fr

om

wei

ghtl

oss

Erosion (mpy) 1.3 2.2 3.6

Ero-Corr (mpy) 4.5 15.3 18.5

(E-C)/E 3.5 7.0 5.1

76 °F 150 °F 200 °F

Figure 2-14. Comparison of penetration rates for pure erosion and erosion-

corrosion at different temperatures by the weight-loss method (single

phase liquid flow, Vl =15 ft/s). 58

Multiphase flow loop testing (high sand rates)

Rincon et al.57 also conducted erosion-corrosion tests for 13Cr alloy under

multiphase flow conditions. The multiphase flow facilities allowed increasing the flow

velocity and reducing the sand rates to more reasonable values. Flow conditions tested

were Vsg = 97 ft/s and Vsl = 0.2 ft/s with an annular flow pattern. The average sand

concentration was 2% by liquid weight representing about 48 lb/day of sand at tested

conditions. The penetration rates for erosion-corrosion were also found to be higher than

those observed for pure erosion tests, again by a factor of 2 to 3 as seen in Figure 2-15.

If the removal rate of the passive film is greater than the formation rate of the

passive film, then an accelerated corrosion process takes place. In most cases, there is

53

likely a competition between the protective film removal due to mechanical erosion and

the protective film healing. If erosivity conditions are severe enough, base metal can also

be removed by the mechanical erosion component. The mechanistic model of this

competition may depend on the concentration and distribution of sand as well as the flow

pattern and fluid flow velocities, geometry and environmental factors. The erosion-

corrosion behavior appears to be a function of both erosivity and environmental

conditions.

0

20

40

60

80

100

120

140

160

180

200

Pe

ne

tra

tio

nra

te(m

py

)(w

eig

ht

los

s)

Erosion 23 45 40 37 60 41

Ero-Corr 105 111 81 149 178 125

1 2 3 4 5 Avg.

Figure 2-15. Penetration rates for pure erosion and erosion-corrosion tests of 13Cr

in multiphase flow conditions (Vsg = 97 ft/s, Vsl = 0.2 ft/s). 57

Figure 2-14 and Figure 2-15 indicate that, in general, the erosion-corrosion

penetration rates were 2 to 3 times higher than the pure erosion rates for single-phase

54

flow as well as for multi-phase flow at room temperature. For higher temperatures, this

factor was even larger.

Over the same testing period that the erosion and erosion-corrosion tests were

conducted (6 hours) for multiphase flow conditions, the pure corrosion penetration rate

(without sand) would be so small that it would not be measurable with weight loss

techniques. Longer tests were needed to obtained measurable mass loss values for pure

corrosion conditions (without sand). A longer, (168 hours) pure corrosion test (included

in Figure 2-16) was performed and the corrosion rate was found to be about 2.3 mpy.

That corrosion rate is higher than the corrosion rate obtained for single-phase (below 1

mpy) conducted at lower velocities. 59

2.3 6.1 4.9

41

125

0102030405060708090

100110120130

Pen

etra

tion

Rat

e(m

py)

(wei

gh

tlos

s)

Corrosion EstimatedCorrosion 1

EstimatedCorrosion 2

Erosion (Avg.) Eros-Corr(Avg.)

Type of Test

Figure 2-16. Comparison among penetration rates for pure Corrosion, pure

Erosion and Erosion-Corrosion processes (Vsg= 97 ft/s, Vsl= 0.2 ft/s). 59

Also, a long test was designed to estimate the mass loss of 13Cr due to the sand-

free CO2 environment after being previously exposed to sand impingement for a period

55

of time. Two tests were conducted at these conditions. For the first 6 hours of the test,

erosion-corrosion conditions (2% sand + CO2) in multiphase flow were used (i.e., CO2

and sand acting together on the 13Cr surface). Later, all the sand was flushed out of the

loop and the 13Cr specimen was exposed to a sand-free CO2 flow for about 160 hours

more. Two aspects can be highlighted from results obtained from these two segments of

the tests.

The penetration rate obtained by dividing the total mass loss by the total test time

(166 hours) was quite low (about 10 mpy) compared to the average obtained for 6 hours

erosion-corrosion tests (125 mpy) shown in Figure 2-15. This suggests that the high

corrosion rates are not sustainable once the flow is sand-free, because the oxide layer

rebuilds. Assuming that the mass loss for the first 6 hours was similar to the average

obtained for 6-hour erosion-corrosion tests (125 mpy) shown in Figure 2-15, a corrosion

component for the sand-free condition can be estimated. To do this, the average of the

mass loss for erosion-corrosion tests shown in Figure 2-15 was subtracted from the total

mass loss measured for the two long tests. For calculating the pure corrosion penetration

rate, 160 sand-free hours were considered instead of the total testing time (166 hours).

The calculated penetration rates are shown in Figure 2-16. It should be mentioned that

some traces of sand could have been present for the remaining 160 hours at the presumed

sand-free condition. As a consequence, the estimated corrosion rates for these two tests

(6.1 and 4.9 mpy), shown in Figure 2-16, are a little higher than the strictly pure

corrosion test result (2.3 mpy). However, this small difference does not alter the

conclusions drawn here.

56

Figure 2-16 summarizes the averages of penetration rates for erosion-corrosion

tests and for erosion tests as shown in Figure 2-15 plus the pure corrosion test results and

the estimates of the corrosion component for the two long tests as explained in the

previous paragraph. Notice, the erosion-corrosion rate was higher than each of the

erosion and corrosion processes appraised separately. Furthermore, the erosion-corrosion

rate was higher than the sum of the erosion and corrosion components appraised

separately, e.g., 125 mpy >> 41 mpy + 6 mpy = 47 mpy (using 6 mpy for the pure

corrosion rate which is the worst case scenario).

These results suggest that the 13Cr specimen does not return to a completely

passivated state as long as the system is sufficiently erosive and as a result, a synergistic

effect between erosion and corrosion is exhibited.

57

60, 61,62,63,64

CHAPTER 1 ....................................................................................................................................1

INTRODUCTION .............................................................................................................................1

CHAPTER 2 ....................................................................................................................................5

BACKGROUND AND LITERATURE REVIEW..............................................................................5

Basic Corrosion Concepts .......................................................................................................5

Concept and Forms of Corrosion ............................................................................................................5

Review of the Electrochemical Basis of Corrosion ................................ .................................7

Theory behind Polarization measurements.............................................................................................8

Linear Polarization Resistance .............................................................................................................. 11

Potentiodynamic Polarization and Tafel Constants ............................................................................. 16

Active Passive Metal Behavior ..............................................................................................17

Passive film............................................................................................................................................. 23

Maintenance and Breakdown of Passivity............................................................................................26

Stainless Steel ................................ ................................................................ ........................27

Alloying elements................................................................................................................................... 27

Flow Velocity Effects ............................................................................................................................29

Flow Pattern Effects ..............................................................................................................31

CO2 Corrosion Resistance of 13Cr Alloy ..............................................................................33

Basic Erosion Concepts.........................................................................................................36

Solid Particle Erosion.............................................................................................................................37

Erosion of Ductile Materials.................................................................................................................. 39

Erosion of Brittle Materials ...................................................................................................................39

Variables Influencing Erosion............................................................................................................... 40

Erosion-Corrosion................................ ................................................................ .................42

13Cr Alloy and Erosion-Corrosion.......................................................................................46

Single phase liquid flow loop testing (high sand rates) ....................................................................... 51

Multiphase flow loop testing (high sand rates) ....................................................................................52

REFERENCES ................................................................................................................................59

FIGURE 2-1. CORROSION PROCESS SHOWING ANODIC AND CATHODIC CURRENT COMPONENTS.33 ...........9

58

FIGURE 2-2. STANDARD ANODIC POLARIZATION CURVE (430 STAINLESS STEEL IN 1 N H2SO4) SHOWING

A TYPICAL ACTIVE PASSIVE TRANSITION BEHAVIOR................................................................. .............19

FIGURE 2-3. SCHEMATIC OF ACTIVE-PASSIVE TRANSITION . POTENTIOSTATIC ANODIC CURVE “JKLM”;

HYDROGEN EVOLUTION REACTION, CURVE “HI”; LOW CONCENTRATION OF DISSOLVED OXYGEN, CURVE

“DEFG”; HIGH CONCENTRATION OF DISSOLVED OXYGEN, CURVE “ABC”...............................................21

FIGURE 2-4. THEORETICAL AND ACTUAL POTENTIODYNAMIC PLOTS OF ACTIVE PASSIVE METALS........23

FIGURE 2-5. DECAY OF PASSIVE CORROSION RATE MEASURED BY POTENTIOSTATIC CURRENT. 31 .............25

FIGURE 2-6. LOG-LOG PLOT OF DATA FROM FIGURE II-5 AT EXTENDED TIMES. 31 .................................26

FIGURE 2-7. EFFECT OF CHROMIUM CONTENT ON ANODIC POLARIZATION OF FE-NI ALLOYS .....................29

FIGURE 2-8. MAJOR CATEGORIES OF WEAR BASED ON THEIR FUNDAMENTAL MECHANISMS. 49 .............37

FIGURE 2-9. SCHEMATIC OF THE EFFECT OF THE IMPINGEMENT ANGLE ON EROSION RATE OF DUCTILE

AND BRITTLE MATERIALS .....................................................................................................................40

FIGURE 2-10. SCHEMATIC OF POLARIZATION CURVES FOR TYPE 316 STAINLESS STEEL SHOWING THE

EFFECT OF PERCENT SOLIDS SAND SLURRY .38................................................................ ........................46

FIGURE 2-11. PLOT OF THE MASS LOSS RATE VS. NORMALIZED DISTANCE FROM THE PIPE EXPANSION AT

DIFFERENT FLOW RATES. (60OC, 3 BAR CO2 AND 1000 PPM OF SAND (RE NUMBER 5.7 105CORRESPOND

TO A FLOW VELOCITY OF 5.7 M/S). 52.....................................................................................................47

FIGURE 2-12. INFLUENCE OF VARIOUS SAND CONCENTRATIONS ON THE MASS LOSS RATE ALONG THE PIPE

LENGTH AT A FLOW RATE OF 3.5 M/S. 52................................................................................................48

FIGURE 2-13. THICKNESS LOSS VERSUS TIME FOR A 13CR ALLOY EXPOSED TO CO2 SATURATED BRINE

CONTAINING SAND PARTICLES. 53................................................................ ..........................................49

FIGURE 2-14. COMPARISON OF PENETRATION RATES FOR PURE EROSION AND EROSION-CORROSION AT

DIFFERENT TEMPERATURES BY THE WEIGHT-LOSS METHOD (SINGLE PHASE LIQUID FLOW, VL =15 FT/S).58 52

FIGURE 2-15. PENETRATION RATES FOR PURE EROSION AND EROSION-CORROSION TESTS OF 13CR IN

MULTIPHASE FLOW CONDITIONS (VSG = 97 FT/S, VSL = 0.2 FT/S). 57................................ ........................53

FIGURE 2-16. COMPARISON AMONG PENETRATION RATES FOR PURE CORROSION, PURE EROSION AND

EROSION-CORROSION PROCESSES (VSG= 97 FT/S, VSL= 0.2 FT/S). 59.......................................................54

59

REFERENCES

1 “Corrosion Cost and Preventive Strategies in the United States” A supplement toMaterials Performance Journal, July 2002, pp. 2.

2 J. D. Combes, J. G. Kerr and L. J. Klein, “13Cr Tubulars solve corrosionproblems in Tuscaloosa trend”, Petroleum Engineer International, March, 1983.

3 M. S. Cayard and R. D. Kane, “Serviceability of 13Cr tubulars in oil and gasproduction environments”, Corrosion 98, paper no. 112, 1998 (Houston, TX,USA, NACE International).

4 H. Amaya, K. Kondo and H. Hirata et. al., “Effect of chromium and molybdenumon corrosion resistance of super 13 Cr martensitic stainless steel in CO2

environment”, Corrosion 98, paper no. 113, 1998 (Houston, TX, USA, NACEInternational).

5 G. Fierro, G.M. Ingo and F. Mancia, “XPS Investigation on the CorrosionBehavior of 13Cr Martensitic Stainless Steel in CO2-H2S-Cl- Environments”,CORROSION/88, paper no. 215, 1988, (Houston, TX, USA, NACEInternational).

6 L.J. Klein, “I&S Cracking Resistance of Type 420 Stainless Steel Tubulars”,CORROSION 84, paper no. 211, 1984, (Houston, TX, USA, NACEInternational).

7 A. Tamaki, “A New 13Cr OCTG for High Temperature and High ChlorideEnvironment”, Corrosion 89, paper No.469, 1989 (Houston, TX, USA, NACEInternational).

8 A. Miyasaka and H. Ogawa, “Influence of Metallurgical Factors on CorrosionBehaviors of Modified 13Cr Martensitic Stainless Steels”, Corrosion/90, paperNo. 28, (Houston, TX:NACE, 1990).

9 M. B. Kermani, D. Harrop, M.L.R. Truchon and J.L. Crolet ‘Experimental Limitsof Sour Service for Tubular Steels”, CORROSION/91. Paper no. 21, 1991,(Houston, TX, USA, NACE International).

10 M.F. Galis and J.J. Damman, “13 Percent Chromium Steels for Slightly SourService”, CORROSION 91, paper no. 20, 1991, (Houston, TX, USA, NACEInternational).

11 S. Hashizume, T. Takaoka, Y. Minami, Y. Ishizawa and T. Yamada, “A New 15Percent Steel Developed for OCTG”, Corrosion/91, paper No.21, (HoustonTX:NACE, 1991).

12 M. Ueda, T. Kushida, K. Kondo, and T. Kudo “Corrosion Resistance of 13Cr-5Ni-2Mo Martensitic Stainless Steel”, CORROSION/92, paper no. 55, 1992,(Houston, TX, USA, NACE International).

13 T. Mori, T. Okazawa, M. Ueda and T. Kushidda, “Corrosion Performance ofSuper 13Cr Steel in CO2 Environments Containing a Small Amount of H2S”,

60

Proceedings from the Second NACE Asian Conference -Corrosion Asia 1994, 1994,(Houston, TX, USA, NACE International).

14 S.F. Biagiotti, Jr. and J.S. Reichman, “Justifying the Use of 13Cr Steels forCorrosive CO2 Operations”, CORROSION/95, paper no. 81, 1995, (Houston, TX,USA, NACE International).

15 H. Asahi, T. Hat-a, A. Kawakami and A. Takahashi., ‘Development of SourResistant Modified 13Cr OCTG”, CORROSION/95, paper no. 79, 1995,(Houston, TX, USA, NACE International).

16 M. Ueda, T. Kushida and T. Mori, “Evaluation of SSC Resistance on Super 13CrStainless Steel in Sour Applications”, CORROSION/95, paper no. 80, 1995,(Houston, TX, USA, NACE International).

17 Y. Miyata, Y. Yamane, O. Furukimi, H. Niwa and K. Tamaki., “Corrosion ofNew 13Cr Martensitic Stainless Steel OCTG in Severe CO2 Environment”,CORROSION/95, paper no. 83, (Houston, TX, USA, 1995, NACE International).

18 H. Asahi, T. Hara, and M. Sugiyama, “Corrosion Performance of Modified 13CrOCTG”, CORROSION/96, paper no. 61, 1996, (Houston, TX, USA, NACEInternational).

19 T.F. Illson and R. Greenwood, “Autoclave Testing of 13Cr and Modified 13CrStainless Steels in Simulated Field Conditions”, CORROSION/95, paper no. 87,1995, (Houston, TX, USA, NACE International).

20 S. Hashizume, K. Masamura and Y. Ishizawa, “Corrosion Data on LaboratoryPerformance of Type 420 for OCTG”, CORROSION/95, paper no. 77, 1995,(Houston, TX, USA, NACE International).

21 M. Ueda and A. Ikeda ‘Effect of Microstructure and Cr Content in Steel on CO2Corrosion”, CORROSION/96, paper no. 13, 1996, (Houston, TX, USA, NACEInternational).

22 J. M. A. Van Der Horst, C.R. Sloan, “Erosion Corrosion of Finned HeatExchanger Tubes,” Erosion, Wear and Interfaces with Corrosion, ASTM STP567, American Society for Testing and Material, Philadelphia 1973, pp. 18-29.

23 Salama, M.M and Venkatesh E.S. “Evaluation of API RP 14E Erosional VelocityLimitations for Offshore Gas Well,” Offshore Technology Conference, PaperOTC 4485, 1983.

24 Smart III, J.S. “The Meaning of the API-RP-14E Formula for Erosion Corrosionin Oil and Gas Production,” CORROSION/91, paper 910468, Houston, TX, USA,NACE International.

25 Smart III, J.S. “A Review of Erosion-Corrosion in Oil and Gas Production,”CORROSION/90, paper 90010, Houston, TX, USA, NACE International.

26 Salama, M.M. “Erosional Velocity Limits in Water Injection Systems,”CORROSION/93 paper 93062, Houston, TX, USA, NACE International.

27 Shirazi, S. McLaury, B, Shadley, J. and Rybicki, E. “Generalization of the APIRP 14E Guidelines for Erosive Services,” 69th Annual Technical Conference andExhibition, paper SPE 28518, 1994.

28 Jordan, K. “Erosion in Multiphase Production of Oil & Gas,” CORROSION/98paper 98058, Houston, TX, USA, NACE International.

61

29 Salama, M.M. “An Alternative to API 14E Erosional Velocity Limits for SandLaden Fluids,” Offshore Technology Conference, Paper OTC 8898, 1998.

30 McLaury, B. and Shirazi, S. “An Alternate Method to API RP 14E for PredictingSolids Erosion in Multiphase Flows,” Journal of Energy Resources Technology,ASME, Vol. 122, pp 115-122, Sep 2000.

31 D. A. Jones, “Principle and Prevention of Corrosion” Prentince Hall, 2nd ed.,pp.343 USA, 1996.

32 “Corrosion Control in Petroleum Production”, NACE International, TPC 5publication Texas, 1979.

33 Application Notes, Gamry Instruments website, www.Gamry.com34 ASTM Standard Designation: G 102 “Standard Practice for Calculation of

Corrosion Rates and Related Information from Electrochemical Measurements”03.02

35 L.L.Shreir, “Corrosion: Metal/Environment Reactions” Newnes-Butterworths,London 1976.

36 ASTM Standard Designation: G 5 “Practice for Standard reference Method forMaking Potentiostatic and Potentiodynamic Anodic polarization Measurements”.03.02

37 Mansfeld, F. “Don’t Be Afraid of Electrochemical Techniques,But Use Themwith Care”. Corrosion, Vol44, No12, pp856-868.

38 ASM Specialty Handbook “ Stainless Steels”, J.R.Davis39 U.Lotz, E. Heitz. “Flow-Dependeant Corrosion. I. Current Understanding of the

Mechanisnsm Involved” Werkstoffe und Korrosion, 34, 1983, pp. 454-46140 K. D. Efird, “Disturbed Flow And Flow Accelerated Corrosion In Oil And

GasProduction”, Energy Sources Technology, ASME, 199841 De Waard C. & Milliams D.E., Corrosion, 31, 131 (1975)42 Ogundele g.l. & White W.E., Corrosion ,42, 71 (1986)43 Schmitt G. & Rothmann B. Werkstoffe, u. Korrosion, 28, 816 (1977)44 Wieckowski A., Ghali E., Szklarczyk M. & Sobkowski J., Electrochimica Acta 28

1619 (1983)45 Eriksud E. & Sontvedt T. Advance in CO2 Corrosion (ed Hausler R,Godard H.P.)

NACE p 20 (1984)46 S. Turgoose & R.A. Cottis, “ Modeling of electrode processes and surface

chemistry in CO2 corrosion”47 J. L. Crolet, M. R. Bonis, “Experience in the Use of 13% Cr Tubing in corrosive

CO2 fields”, SPE Production Engineering, sep 1986.48 T. Hara, H. Asahi, H. Kaneta, “Effect of Flow Velocity on CO2 Corrosion and

Galvanic CorrosionBehavior in Oil and Gas Environments” CORROSION/98paper 118

49 J.R. Davis,“Surface Engineering for Corrosion and Wear Resistance” ASMInternational, march, 2001.

50 A.V. Levy, “Solid Particle Erosion and Erosion Corrosion of Material” ASMInternational, Ohio, 1995.

62

51 R. Chattopadhyay, “Surface Wear Analysis, Treatment and Prevention”, ASMInternational

52 T. Kohley, E. Heitz, “Particle Containing Formation Water for the Study ofErosion Corrosion”, The Use of Enviroments for Corrosion Testing, ASTM STP970, American Society for Testing and Material, Philadelphia 1988, pp. 235-245.

53 P. M. Birchenough, S. G. B. Dawson, T. J. Lockett, and P. McCarthy,“Simultaneous Erosion and Corrosion in Multiphase Flow” NACE 7th MiddleEast Corrosion Conference, Bahrain, 26-28 February 1996.

54 ASTM Standard Designation: G 119 “Standard Guide for DeterminingSynerginsm between Wear and Corrosion.” 03.02

55 R.J.K Wood, S.P. Hutton “ The Synergistic Effect of Erosion Corrosion: Trendsin Published Results” Wear, 186-187, pp. 523-532, 1995

56 A. Neville, H. Xu, M. Reyes, “Corrosion and Erosion-Corrosion Behavior of aCo-Based Alloy and a Ni-Containing Austenitic Cast Iron”, CORROSION/2000,paper no. 628, 2000, NACE Congress

57 Rincon, H.E., Chen, J. and Shadley, J.R. (2002). "Erosion-Corrosion Phenomenaof 13Cr Alloy in Flows Containing Sand Particles". Corrosion/2002, paper no.2493, (Houston, TX, USA, NACE International.

58 Chen, J., Shadley, J.R., Rincon, H.E. and Rybicki, E.F. (2003). "Effects ofTemperature on Erosion-Corrosion of 13Cr". Corrosion/2003, paper no. 3320,(Houston, TX, USA, NACE International.

59 Rincon, H.E., "Erosion-Corrosion Phenomena of 13Cr Alloy in Flows ContainingSand Particles". MSc Tesis, The University of Tulsa, 2001.

60 A. McMahon and J. Martin “Simulation Tests on the Effect of MechanicalDamage or Acid Cleaning on CRAs Used for Oil/Gas Production WellTubulars”,CORROSION/04, paper no. 4127, 2004, (Houston, TX, USA, NACEInternational).

61 Rincon, H.E. (2001). "Erosion-Corrosion Phenomena of 13Cr Alloy in FlowsContaining Sand Particles". M.S. Thesis, Department of Mechanical Engineering,The University of Tulsa, Ok, Tulsa.

62 Oka, Y.I, Okamura, K., Yoshida, T., “Practical estimation of erosion damagecaused by solid particle impact. Part 1: Effects of impact parameters on apredictive equation.” Wear 259 (95-101)

63 Oka, Y.I, Yoshida, T., “Practical estimation of erosion damage caused by solidparticle impact. Part 2: Mechanical properties of materials directly associated witherosion damage.” Wear 259 (102-109)

64 M. Stern and E.D. Weisert, Proc. ASTM, Vol. 32, p.1280, 1959.

57

CHAPTER 3

OBJECTIVES AND APPROACH

Research Objectives

The objective of this work is to develop a knowledge base on the roles of scale

formation, passivation, solid particle erosion and their interactions to advance erosion-

corrosion predictive modeling for alloys commonly used in oil production systems.

Completion of this objective would hope to improve and generalize guidelines for the

oil production industry to safely operate by minimizing erosion-corrosion problems in

corrosion resistant alloys (CRAs). These objectives are advanced in this research by

developing an effective and efficient procedure for investigating the erosion-corrosion

behavior of CRAs in oilfield environments. The idea is to reduce the need for expensive

and time consuming loop tests by using a simplified scratch test, and also provide the

basis and framework for an erosion-corrosion predictive model that eventually could be

used to predict erosion-corrosion penetrations rates for oil industry service conditions in

a reliable manner.

58

Research Approach

To address this need, extensive experimental work was conducted. Effects of

pH, temperature, flow conditions and sand concentration in the erosion-corrosion

behavior of 13Cr, Super13Cr and 22Cr alloys were investigated by means of weight

loss (WL), electrical resistance (ER), linear polarization resistance (LPR),

potentiodynamic polarization scans (PD) and scratch tests (ST).

A convenient procedure to predict erosion-corrosion of CRAs using

combination of loop testing and scratch testing is presented. Also, the framework for a

semi-mechanistic model to estimate erosion-corrosion penetration rates for CRAs

exposed to CO2-saturated brine/sand flows conditions was constructed. The procedure

is based on experimental data obtained from scratch tests to address the electrochemical

component of the erosion-corrosion process, and on computational fluid dynamics

(CFD) simulations to address the mechanical component. Experimental and numerical

data are then processed with a VBA code to generate the penetration rates estimation.

The model assumed fixed parameters such a brine concentration, water

chemistry and CO2 pressure. Input parameters are temperature, pH (which are implicit

in the second order scratch test parameters, m and Io), alloy, sand rate, sand size and

liquid velocity. Outputs of the model include the pure erosion rate, E the total erosion-

corrosion penetration rate, EC and the corrosion component rate of the erosion-

corrosion process Ce-c, defined as the sum of the pure corrosion C and the impact-

induced corrosion Ce.

59

Specific objectives of this research programs are:

1. Investigate the erosion and erosion-corrosion behavior of CRAs in

single-phase and multiphase flows.

2. Conduct weight loss and thickness loss measurements (multiphase flow

loop testing) over a range of pure erosion and erosion-corrosion

conditions to identify conditions, if any, for which there is a significant

corrosion component in the erosion-corrosion degradation process of

CRAs.

3. Perform electrochemical tests conducted in the scratch test cell to

investigate the effect of pH and temperature on the re-formation of the

passive film on CRAs

4. Develop a procedure for investigating the erosion-corrosion behavior of

CRAs in an effective and efficient manner using a combination of loop

test data and scratch test data.

5. Perform electrochemical tests conducted in a single liquid phase loop to

investigate the role of sand particle concentration and fluid flow velocity

on the re-formation of the film and repassivation times for CRAs.

6. Conduct computational fluid dynamics (CFD) modeling to predict sand

particle wall impact characteristics (location, impact velocity, impact

angle) and erosion rate as an alternative to loop testing to generate the

mechanical component of erosion-corrosion prediction.

7. Combine the CFD procedures for erosion prediction with scratch test

procedures for prediction of repassivation kinetics to construct a

60

framework for a semi-mechanistic model for prediction of erosion-

corrosion of a CRA in sand-bearing, CO2 saturated liquid flows.

61

CHAPTER 4

EXPERIMENTAL PROCEDURE AND TESTING CONDITIONS

Scratch Test Experimental Setup

Figure 4-1 shows the layout of the electrochemical cell. One of the working

electrodes (WE1) is machined into a special shape to ensure consistent scratch

geometries and surface areas. The lower portion of WE1 is rectangular in shape with the

dimension of 1/4 x 1/4 x 1/2-inches. Only one surface of the electrode is polished, and

the other sides are epoxy coated. The upper portion of WE1 is machined into a

cylindrical shape with 1/8-inch diameter. The side of the cylinder is also epoxy coated,

but the circular surface on the top of the cylinder (scratched area) is initially polished.

The WE1 is screwed into the electrode holder to form an electrical connection with the

wire embedded in the electrode holder. The other working electrode (WE2) is

rectangular in shape with dimensions of 1/4 x 1/4 x 2-inches. The entire surface of the

WE2 is polished. Working electrode area ratios have been selected to control error in

measured current.

The grinding disc is installed above the WE1. It turns at a speed of 500 rpm. To

make a scratch, the grinding disc is lowered to grind the top of the WE1 for about 5

seconds.

62

Figure 4-1. Layout of the electrochemical cell for Scratch Test.

An electrochemical measuring system is used to monitor the potential of the

working electrode system and the current between WE1 and WE2. The anodic current is

considered to be positive in this study. The configuration of the experimental set-up for

electrochemical measurements is shown in Figure 4-2. The corrosion potential of the

working electrode (WE1) with respect to a Ag/AgCl reference electrode is monitored by

a high input impedance, high resolution and low-noise digital multimeter. The current

between the two working electrodes is monitored by a zero resistance picoammeter.

Software was developed to communicate with the instruments and read data via GPIB.

63

The sampling frequencies are synchronized by the software to ensure simultaneous

readings of potential and current. Data are read directly from A/D converters and

relayed immediately to the computer for further processing. The potential and current

signals are saved in the computer so the data can be analyzed in the time domain or in

the frequency domain to examine noise patterns related to erosion-corrosion.

Figure 4-2. Set-up for electrochemical measurements for Scratch Test.

CO2 gas is continuously bubbled into the 3% NaCl brine test solution during the

complete time period of the test. The liquid in the electrochemical cell is gently agitated

by the magnetic stirrer to maintain homogeneity in the solution, and the temperature of

the liquid is maintained at a desired level by a heater. The pH of the solution is

64

monitored by a pH meter and can be adjusted to a desired value by adding HCl or

NaOH to the electrochemical cell. The gas from the electrochemical cell is released to

the atmosphere through two gas trap bottles, which are connected in series and filled

with the same solution as the test cell. A valve installed at the bottom of one of the trap

bottles allows the liquid to flow out when the valve is opened, so the oxygen

concentration can be measured at any time with an oxygen measuring kit. Before

making any scratch on the working electrode, WE1, the system is maintained at this

condition for at least 24 hours during which pH and temperature are adjusted to the

desired levels.

Test matrix

The scratch tests were carried out at three different temperatures; 76°F, 150°F

and 200°F and five different pH conditions; 3.5, 4.0, 5.25, and 6.0 for 13Cr and Super

13Cr testing. 22Cr testing was just conducted at pH 4.0 at the three temperatures. At

each condition, at least 3 scratch tests were conducted to check the reproducibility. The

time between two consecutive scratch tests was at least 2 hours in order for the working

electrode to return to its original state after the scratch. Therefore, at least 81 scratch

tests were conducted to cover all test conditions.

Erosion-Corrosion -Loop (Gas/Liquid/Sand Multiphase Flow Loop)

Tests in multiphase flow are carried out in the Erosion-Corrosion loop. A

schematic of the flow loop is shown in Figure 4-3. The major components of the flow

65

loop are two Corken D490 compressors each with a maximum capacity of about 30

cfm, a 40 gpm diaphragm pump, a large separator tank, several heat exchangers and

scrubber, and a cyclone separator for separating sand from the gas-liquid sand mixture.

Gas from the compressors flows through a heat exchanger and splits into two

branches as shown in Figure 4-3. One branch leads to the test cell section and the other

branch is a bypass going back to the compressor intakes. Before the gas has reached the

test cell, dry sand is added to the system by means of a vibration driven sand feeder.

Downstream of the sand injection point, the liquid is incorporated into the flow-stream

and then the three phases (gas-liquid-sand) flow together in the test section.

Downstream of the test section, sand is separated from the slurry in the cyclone

separator. Sand with some water falls down into the separator to a receiver cylinder

where the sand is collected and removed from the system. Notice the system is a one

through pass for sand. Most of the gas-liquid mixture passes through the top of the

cyclone separator and flows into the main tank (gas-liquid separator). Then the gas

phase is separated from the gas-liquid mixture and the liquid flows back to the pump.

The gas flows to a heat exchanger and scrubber to remove any moisture from it before

flowing back to the compressors. Another heat exchanger at the compressor discharge

is used to cool the gas before it flows back to the test section.

The Erosion-Corrosion loop is capable of producing superficial liquid velocities

up to 15 ft/s, and superficial gas velocities up to 100 ft/s. The maximum working

pressure is 150 psig. The maximum temperature is 200°F. In the CO2 loop, erosion and

erosion/corrosion tests can be conducted on elbow geometries when weight loss

coupons are used. A high resolution ER probe is also used to collect the corrosion,

66

erosion and erosion-corrosion penetration data when the metal loss is not expected to be

high. The probe is set flush mounted at a 90 degree turn at the top of a 6-foot long

vertical section (1/2” piping section) of the test section. The flow geometry is close to a

plugged tee geometry. Figure 4-3 shows the current position of the probe in a schematic

of the Erosion-Corrosion loop.

Figure 4-3. Schematic of the Erosion-Corrosion Flow Loop

67

Test cell

Figure 4-4 shows a photograph of the test section of the CO2 loop with the ER

probe set in place (left) and the spiral-shaped-sensor element of the 13Cr ER probe

(right).

Figure 4-4. Photograph of the test section of the Erosion-Corrosion loop with

the ER probe set in place (left). Sensor element of the 13Cr ER

Probe (right).

Test conditions

Electrical resistance measurements in pure erosion and erosion-corrosion in

multiphase flow conditions were performed to see if there is a significant corrosion

component. For pure erosion tests, distilled water and nitrogen were used; for erosion-

corrosion tests 3% brine and CO2 were used. Test conditions are shown in Table 4-1.

68

Table 4-1. Test conditions for erosion and erosion-corrosion tests.

Low Sand Rates High Sand Rates

Material 13Cr 13Cr, Super 13Cr and 22Cr

Flow Geometry 1” Plugged Tee ½” Plugged Tee

Liquid flow velocity 0.2 ft/s 1.4 ft/s

Gas flow velocity 55 and 60 ft/s 20 ft/s

Sand size (avg) 150 microns

Sand rate 1 and 15 lb/day 30 lb/day

Time with sand 5.5 to 6 hours 1 to 3 hours

Solution 3% NaCl* or Distilled water**

Gas Carbon Dioxide* or Nitrogen**

Pressure 50 psig

Temperature 76 and 150 oF

Dissolved Oxygen Less than 10 ppb

pH 4 at 76oF and 4.4 at 150oF

* erosion-corrosion tests ** pure erosion tests

Erosion-Corrosion Liquid/Sand Loop (Microloop)

A smaller experimental facility was designed similar to the Erosion-Corrosion

Multiphase flow loop but for single phase liquid flow. Thus, the loop is capable of

handling CO2-saturated corrosive liquid containing sand. For ease of operation, this

69

facility is much smaller than the former single phase liquid loop (Miniloop) and is

called the Microloop. The main objective of designing the Microloop was to study the

effect of sand concentration and flow velocity on the erosion-corrosion behavior of

corrosion resistance alloys, and thus be able to link the scratch test cell results to

flowing systems containing sand.

Instrumentation and measurement techniques similar to those used in the static

scratch test were used in this facility. Potential and current flowing between the target

working electrode (WE1) and the auxiliary working electrode (WE2) are recorded

during the pre-sand exposure, sand exposure and post-sand exposure period. The

current and potential between the two working electrodes are monitored by a zero

resistance picoammeter and a high resolution and low-noise digital multimeter

respectively. The same software developed to communicate with the instruments for the

Scratch Test was used for Microloop testing.

After some simple mathematical calculations of the measured current data, the

thickness loss and the repassivation time are estimated. As opposed to the scratch test

cell, in the dynamic liquid/sand loop test mechanical and electrochemical processes

involved are taking place simultaneously and continuously, which may make the

analysis of the results quite complex. However, current responses immediately after

removing the sand from the flow provide useful information to determine repassivation

times and the thickness loss that occurred before the material returned to its passive

state.

This Microloop is mainly constructed of 316L SS steel because of its excellent

resistance to corrosion. The major components of this loop are a 10 gallon tank,

70

circulation pump, cyclone separator, test section, sand feeder and sand filter bypass.

Various piping and fittings connect these components to form a closed loop. The safe

working pressure for the Microloop is around 50 psig. Tests were conducted at 20 psig.

A vacuum pump is used to de-aerate and remove dissolved oxygen from the solution. A

band heater around the tank is used to increase the temperature of the solution. This

temperature is maintained in the operating ranges of 100oF to 200oF by using an

electronic controller. The pump is a Wanner Engineering D-3 Hydra-cell diaphragm

pump with a capacity of 3 gpm flow under operating pressures. The pump is the

principal driver and circulates the solution through the loop and back to the tank. The

velocity of the flow through the Microloop is controlled by a variable speed drive for

the pump. This controller allows varying the average flow velocity at the exit of the

direct impingement jet from 10 to 20 ft/s. Lower flow velocities at the jet outlet may be

achieved by bypassing part of the liquid mass directly to the reservoir tank. The

velocity of the flow impinging the test specimen could not be directly measured due to

the small length of the test section which prevented the use of the ultrasonic flow-

meter. However, the velocities were estimated using Bernoulli’s equation. The pressure

differential between the total and static pressure was measured by using a special

straight pitot-tube and a digital manometer.

A schematic of the Microloop is shown in Figure 4-5. The test cell section,

where most of the sand and some liquid are re-circulated, is shown in Figure 4-6. The

sand feeder is de-aerated and pressurized to a pressure slightly higher than the system

pressure (20 psi) to introduce the sand into the system. Sand is introduced into the test

section just upstream of the cyclone separator, where it gets separated from the liquid

71

phase and drops down to the injector nozzle. The sand free liquid flows back to the

tank. The nozzle at the based of the cyclone separator reintroduces the sand into the

flow stream towards the direct impingement test cell. The sand remains recirculating in

the test section until the filter bypass section is engaged and the valve at the top of the

cyclone separator is closed which forces the entire flow to circulate through the 5

micron cartridge filter.

A sand sampling port is provided for measuring the concentration of sand in the

test section as desired. The pH of the solution is monitored using a pH probe housed in

another by-pass section. A pH meter with digital readout is connected to the probe for

accurately measuring the pH during the test. The pH can be controlled by using a

dosing pump which injects acidic or basic solution into the tank as per requirement.

Direct impingement test cell

A direct impingement configuration of the test specimen is used for testing in

the Microloop. Figure 4-7 shows a schematic of the test cell. Two round specimens

made of the same CRA with ½” diameter and ¼” thickness are inserted into a standard

½” cross fitting such that they are perpendicular to each other. The flow coming from

the pump is first accelerated through 1/8” nozzle which discharges into a 5/16” ID

tubing (jet) extended into the cross fitting and terminating very close to the target

specimen (WE1). The flow then, enters the test cell (cross fitting) horizontally through

the jet and directly impinges the target working electrode (WE1), as shown in Figure 4-

7. Flow is diverted vertically upward after impinging the target specimen, leaving the

test cell and moving towards the cyclone separator.

72

Figure 4-5. Schematic of the Microloop.

Corrosion measurements can be taken by weight loss and LPR. To enable LPR,

the direct impingement specimen is used as the working electrode. The other cylindrical

specimen placed at the bottom-port of the cross fitting is used as the reference electrode

and the loop, which is made of 316 stainless steel, is connected as the counter electrode.

The specimens are mounted on threaded phenolic holders, which are inserted into the

test cell. Threaded rods attached to the specimens and sealed with O-rings, protrude out

of the test cell and are connected to the potentiostat leads for LPR measurements.

73

Figure 4-6. Schematic of the test cell section for the Microloop indicating

positions of target working electrode (WE1) and auxiliary working

electrode (WE2).

Figure 4-7. Schematic of the direct impingement test cell for the Microloop

FF

PhenolicHolder

Direct ImpingementSpecimen, WE1

Reference ElectrodeFor LPR, RE

ImpingementJet

Cross Fitting Test

74

A second (or auxiliary) working electrode (WE2, made of the same CRA 2.75”

x 1.425” x 0.275”) is placed upstream of the nozzle and thus is not exposed to the

flowing sand at any time during the test (see test cell section in Figure 4-6). The current

and potential between the two working electrodes are monitored with the same

instrumentation used for the scratch test cell as described above.

Test conditions for liquid/sand loop test

Table 4-2 shows the range of environmental and flow conditions for testing

conducted in the Microloop.

Table 4-2. Test conditions for erosion and erosion-corrosion tests

Erosion-Corrosion Conditions

Material 13Cr and Super 13Cr

Geometry Direct Impingement

Liquid flow velocity 10, 15, 17 and 20 ft/s

Sand size Range (212-272) microns

Sand rate 10 to 860 lb/day

Time with sand 10 min to 240 hours

Solution 3% NaCl

Gas Carbon Dioxide

Pressure 20 psig

Temperature 150 oF

Dissolved Oxygen Less than 10 ppb

pH 4.3 at 150oF

75

Material tested

The material of the working electrodes under investigation are API 5CT L80-

13Cr, referred to as “13Cr” in this report, and a KO-HP2-13Cr-95 referred to as a

“Super 13Cr” (Super 13Cr). The chemical composition of the 13Cr and Super 13Cr are

shown in Table 4-3 Table 4-4, respectively. For comparison purposes, the UNS 2205

alloy referred to as “22Cr” in this report was also tested at some of the most aggressive

conditions and its chemical composition is shown in Table 4-5

Table 4-3. The chemical composition of 13Cr.

C Si Mn P S Cr Ni Mo Cu Fe

min 0.15 - 0.25 - - 12.0 - - - -

max 0.22 1.00 1.00 0.020 0.010 14.0 0.50 - 0.25 -

Ladle 0.18 0.28 0.45 0.020 0.002 12.91 0.11 0.00 0.01 Bal

Table 4-4. The chemical composition of Super 13Cr.

C Si Mn P S Cr Ni Mo Cu Fe

min - - 0.30 - - 12.00 3.50 1.50 - -

max 0.040 0.50 0.60 0.020 0.005 14.00 5.50 2.50 0.25 -

Ladle 0.020 0.19 0.44 0.018 0.001 12.85 5.35 2.14 0.10 Bal

Table 4-5. The chemical composition of the 22Cr.

C Si Mn P S Cr Ni Mo Cu Fe

min - - - - - 21.0 4.5 2.5 - -

max 0.030 1.00 2.00 0.030 0.020 23.0 6.5 3.5 - -

Ladle 0.018 0.40 0.480 0.026 0.015 22.62 5.51 3.38 0.20 Bal

76

CHAPTER 3 ..............................................................................................................................57

OBJECTIVES AND APPROACH................................................................................................ ....57

Research Objectives................................................................................................ ...........57

Approach................................ ............................................................................................58

CHAPTER 4 ..............................................................................................................................61

EXPERIMENTAL PROCEDURE AND TESTING CONDITIONS ..........................................................61

Scratch Test Experimental Setup .......................................................................................61

Test Matrix ......................................................................................................................................... 64

Erosion-Corrosion-Loop (Gas/Liquid/Sand Multiphase Flow Loop) ...............................64

Test cell ............................................................................................................................................... 67

Test conditions ................................................................................................................................... 67

Erosion-Corrosion Liquid/Sand Loop (Microloop) ...........................................................68

Direct impingement test cell.............................................................................................................. 71

Test conditions for Liquid/Sand Loop Test ...................................................................................... 74

Material tested .................................................................................................................................... 75

TABLE 4-1. TEST CONDITIONS FOR EROSION AND EROSION-CORROSION TESTS......................................68

TABLE 4-2. TEST CONDITIONS FOR EROSION AND EROSION-CORROSION TESTS .....................................74

TABLE 4-3. THE CHEMICAL COMPOSITION OF 13CR. ................................................................ .............75

TABLE 4-4. THE CHEMICAL COMPOSITION OF SUPER 13CR. ..................................................................75

TABLE 4-5. THE CHEMICAL COMPOSITION OF THE 22CR. ................................................................ ......75

FIGURE 4-1. LAYOUT OF THE ELECTROCHEMICAL CELL FOR SCRATCH TEST. ................................ ....62

FIGURE 4-2. SET-UP FOR ELECTROCHEMICAL MEASUREMENTS FOR SCRATCH TEST...............................63

FIGURE 4-3. SCHEMATIC OF THE EROSION-CORROSION FLOW LOOP................................ ......................66

FIGURE 4-4. PHOTOGRAPH OF THE TEST SECTION OF THE EROSION-CORROSION LOOP WITH THE ER

PROBE SET IN PLACE (LEFT). SENSOR ELEMENT OF THE 13CR ER PROBE (RIGHT). ............................67

FIGURE 4-5. SCHEMATIC OF THE MICROLOOP....................................................................................72

FIGURE 4-6. SCHEMATIC OF THE TEST CELL SECTION FOR THE MICROLOOP INDICATING POSITIONS OF

TARGET WORKING ELECTRODE (WE1) AND AUXILIARY WORKING ELECTRODE (WE2). .....................73

FIGURE 4-7. SCHEMATIC OF THE DIRECT IMPINGEMENT TESTCELL FOR THE MICROLOOP .................73

77

76

CHAPTER 5

SCRATCH TEST AS A SIMPLIFIED EROSION-CORROSION TEST

Motivation for Doing Scratch Test

A laboratory method previously introduced by McMahon and Martin60 was used

to study the effect of pH and temperature on the repassivation behavior of CRAs in CO2-

saturated brines. The laboratory method was conducted in a glass cell. Data were

collected based on an electrochemical technique. The working electrode (CRA) is

scratched while the potential and current are recorded. After some simple mathematical

calculations, the thickness loss and the repassivation time can be estimated. The scratch

made on a well-defined area of the working electrode may be interpreted as similar to the

removal of the passive layer due to a single particle impingement, and it is believed that

this approach can be used to study the effect of erosion on the corrosion response. With

scratch tests, the mechanical process of making the scratch and the repassivation process

of a CRA can be viewed as nearly separate events. The sand erosion process bears some

similarity to a simple scratch test except that mechanical and electrochemical processes

involved are taking place simultaneously and continuously. The goal of investigating the

simple scratch test was to shed some light on the erosion-corrosion behavior of CRAs.

77

The effects of pH and temperature on the repassivation of 13Cr, Super 13Cr and

22Cr were thoroughly examined, and a summary of the results are presented later in this

chapter. First, the data reduction techniques for the measurements obtained from the

scratch test are described in the following section.

Data Reduction Technique

For a typical scratch test, the initial anodic current is very high immediately after

making the scratch. Figure 5-1 shows a typical current decay with time during a scratch

healing process for a 13Cr alloy exposed to 3.5% brine solution at pH 4 and temperature

of 76oF. The three CRAs show similar behavior for all conditions tested. How fast the

current decays with time is determined by the healing rate of the oxide film which

depends on many variables such as temperature, pH, flow velocity, and material tested.

0

5

10

15

20

25

30

35

40

45

50

55

0 20 40 60 80 100 120 140 160 180 200

Time (sec)

Cu

rren

t,I

( mA

)

76 oF, pH = 4, 13Cr

Figure 5-1. Anodic current decay during the scratch repassivation process.

78

The anodic current decay for CRAs was characterized as approximately following

the second order behavior at least for the first 200 seconds of the scratch repassivation

process expressed by the equation (5-1).

mIdtdI 2 (5-1)

where, I is the measured current, t is time, and m is a constant that depends on material

and environmental conditions. The “second order” is used here to denote that the rate of

change of current, I, with respect to time is proportional to I2. Integration of equation (5-

1) yields:

mtII

I0

0

1 or mt

II

0

11 (5-2)

where, Io is the initial current at t = 0.

Equation (5-2) suggests Io and m can be conveniently obtained from a plot of 1/I

vs. time as shown in Figure 5-2, which contains the same data shown in Figure 5-1.

1/I = 1,774 t + 36,067

0.0E+00

1.0E+05

2.0E+05

3.0E+05

4.0E+05

5.0E+05

0 50 100 150 200

Time (sec)

1/I

(1/A

)

76 oF, pH = 4, 13Cr

Figure 5-2. Data showing a linear relation between 1/I and time.

79

Equation (5-2) can be integrated and presented in a convenient form to

characterize the healing process as a protective film growth process. Estimation of the

cumulative thickness loss (TL) at any time t = t* is given by:

)1ln( 0

*

0

tmImCdtICTL

t

(5-3)

where, C accounts for units conversion factor and some material properties, t* is the

repassivation time defined as the time required for the current to return to a specific value

of I* (e.g., I* is the current at the passive state corresponding to a corrosion rate of about

1 mpy), then from equation (5-2), the repassivation time, is given by:

0

11ImIm

t

(5-4)

This means that knowing m and Io, the time required for the corrosion rate to

return to one mpy, or other appropriate limit, can be predicted.

An expression for the TL accumulated by the time the CRAs repassivate can be

derived by substituting equation (5-4) into equation (5-3), yielding:

II

mC

ImImIm

mCTL 0

00 ln111ln (5-5)

Figure 5-3 shows the comparison of the actual cumulative thickness loss (TLactual)

and the calculated cumulative thickness loss (TLcal) for 13Cr from the raw data displayed

in Figure 5-1. Good agreement is obtained.

80

0.0E+00

2.0E-06

4.0E-06

6.0E-06

8.0E-06

0 50 100 150 200Time (sec)

Cu

mu

lati

veT

hic

knes

sL

oss

(mm

) Integration of Current Data

Second Order Model

76 oF, pH = 4, 13Cr

Figure 5-3. Comparison between TLactual and TLcal for 13Cr alloy.

In fact, the second order model was found to be a very good approximation to describe

the repassivation process of 13Cr, Super 13Cr and 22Cr under all conditions tested at

least for the first 200 seconds. This allows the use of equation (5-5) to compare two

different materials exposed to the same environmental conditions, or the same material

exposed to two different environments, according to the TL accumulated by the time the

corrosion rate reduces to zero as follows:

BA

AB

AAA

BBB

A

BAB mC

mCIImCIImC

ITLTL

TLR

lnln

0lim/

(5-6)

where the subscripts A and B refer to materials (or environments) A and B, TLRB/A is the

thickness loss ratio of B with respect to A, and IA and IB refer to initial (at t = 0) corrosion

currents for materials (or environments) A and B respectively. Equation (5-6) is

important and very useful. It indicates that the severity of the corrosion component of

erosion-corrosion is reflected in the slope of the second order repassivation process. The

81

larger the slope m, the faster the repassivation process is, and therefore the lower the

contribution of the corrosion component to the total damage due to the erosion-corrosion

mechanism. Details of the estimation of the constant C for each material may be found in

Appendix B, however due to the similarity in composition of the materials the ratio CB to

CA in equation (5-6) is approximately 1 and can be neglected for these alloys.

Equation (5-6) suggests that the ratio of the slopes may be a useful indicator of

the relative severity of the erosion-corrosion of two different alloys exposed to erosive-

corrosive flow conditions, provided that the erosivities of the two systems being

compared are similar. Support for this hypothesis is provided in Chapter 7.

Scratch Test Results

Details of the results obtained by means of the scratch test for the three CRAs are

presented in Appendix A. This section just summarizes some of the most important

results. Here, the effect of the temperature and pH are highlighted. Comparisons among

the behaviors of the current versus time response for the three alloys are made. Data is

also conveniently displayed as the inverse of the current versus time. In this type of plot

differences between the expected erosion-corrosion behaviors may be easier to grasp

based on the comparison of the slopes m. Besides current transients, other interesting

parameters such as cumulative thickness loss and repassivation times are discussed.

Effect of pH on Scratch Test results

Figure 5-4 shows the current decay curves for 13Cr at 150°F for different pH

conditions. A couple of trends can be seen in this figure. The current decays faster at

82

higher pH as indicated by the greater initial slopes shown for pH 5.25 and 6 as compared

to lower pH values (3.5 or 4). Also, the value of the current observed is lower at higher

pH (5.25 or 6) than for lower pH (3.5 or 4) over the entire interval of 200 seconds shown

in the figure. Since the corrosion rates are proportional to the current, these results

suggest very high corrosion rates immediately after making the scratch and exposing the

bare metal to the corrosive solution. Then corrosion rates decay with time while the

healing process takes place on the metallic surface. The passivated state is achieved faster

for higher pH’s. For the other temperatures, namely 76°F and 200°F, similar trends with

pH are observed (results are included in Appendix A). However the initial currents

obtained immediately after making the scratch for these temperatures are lower at 76oF

and higher at 200oF than the current obtained for 150oF.

Figure 5-5 shows the current decay curves for Super 13Cr at different pH

conditions at 150°F. Trends observed in this figure for Super 13Cr seem to be similar to

those observed for 13Cr in Figure 5-4, i.e., the current decays at faster rates at higher pH

than it does it for low pH. Also the value of the initial current is lower for higher pH.

However, variations in the Super 13Cr current response due to changes in pH are less

evident than those showed by 13Cr. Current transients trends with pH for Super 13Cr

obtained at 76°F and 200°F are similar to that observe for 150oF (see Appendix A). They

all seem to have same trends with pH as those shown for 13Cr in Figure 5-4; but, current

transients for Super 13Cr for all pH are very close to each other and overlap in a very

narrow range which makes it difficult to observe the trend of the response of the Super

13Cr with pH. However plots of 1/I vs. time for the same data included in Appendix A

showed the trends more clearly.

83

0

20

40

60

80

100

120

140

0 50 100 150 200Time (sec)

I( m

A)

pH = 3.5

pH = 4.0pH = 5.25

pH = 6.0

13Cr AlloyT = 150ºF

Figure 5-4. Effect of pH on the decay of the anodic current after a scratch has

been made on the surface of a 13Cr Alloy.

0

20

40

60

80

100

120

140

0 50 100 150 200Time (sec)

I(mA

)

pH = 3.5pH = 4.0

pH = 5.25, and pH = 6.0

Super13Cr AlloyT = 150ºF

Figure 5-5. Effect of pH on the decay of the anodic current after a scratch has

been made on the surface of a Super 13Cr Alloy.

Figure 5-6 shows 1/I versus time t for the same raw data used in Figure 5-4. Good

linearity between 1/I and time t in all pH conditions at room 150°F indicates that the

84

repassivation processes in these conditions all follow approximately second order

behavior at least for the first 200 seconds. According to equation (5-6), the thickness loss

ratio between two different conditions for the same material is the reciprocal of the ratio

of corresponding slopes in the 1/I versus time t charts. As shown in Figure 5-6, the slope

at pH = 3.5, is lower (mpH=3.5= 485) than the slope at pH =4.0 (mpH=4= 1063) which is

much lower than the slope for pH = 6 (mpH=6.0= 2159). Therefore, the corrosion

component of the erosion-corrosion damage at pH 3.5 should be about 2.2 times more

severe than at pH 4.0, and about 4.5 times more severe than at pH 6.0. These results

indicate how significant the effect of pH is on the repassivation process of 13Cr at tested

conditions. Similar plots for 76 and 200oF are found in Appendix A.

0.E+00

2.E+05

4.E+05

6.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

)

pH=4.0

pH=5.25

pH=3.5

pH=6.0T= 150ºF13Cr

Figure 5-6. 1/I vs. t for 13Cr from the raw scratch test data in Figure 5-4.

Figure 5-7 is the 1/I versus time t chart for Super 13Cr at 150°F from the raw data

in Figure 5-5. The trend of the slope with pH observed for this alloy is similar to that

observed for 13Cr, mpH=6.0 > mpH=5.25> mpH=3.5, which predicts a lower corrosion

85

component of the erosion-corrosion process for higher pH values. Notice that slopes for

Super 13Cr at different pH values are clustered closer together as compared with those

for 13Cr; hence, the effect of pH on the erosion-corrosion damage of Super13Cr is

expected to be smaller than that experienced by 13Cr. Again trends for 76 and 150oF are

found in Appendix A.

0.E+00

2.E+05

4.E+05

6.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

)

pH=4.0, & 3.5

pH=5.25pH=6.0

T= 150ºFSuper 13Cr

Figure 5-7. 1/I vs. t for 13Cr from the raw data in Figure 5-5.

Effect of temperature on Scratch Test results

Figure 5-8 shows the comparison of the current decays for 13Cr scratch tests

conducted at different temperatures, 76, 150 and 200oF but for a fixed pH of 4. As

expected, at higher temperatures, higher healing rates were found as suggested by the

steeper slopes shown by the current vs. time curves at 200oF. However, the higher

temperature always causes a current that is initially higher and remains higher over the

entire test period than for a lower temperature. As a result, at higher temperatures, it takes

86

a longer time for the current to return to its original level. Therefore, the cumulative

thickness loss after making the scratch is expected to be higher for higher temperatures,

as shown in Figure 5-9. Similar trends are obtained for other pH values (3.5, 5.25 and

6.0), and results are included in Appendix A.

0

20

40

60

80

100

120

140

160

180

200

0 50 100 150 200Time (sec)

I(mA

)

13Cr AlloypH= 4.0

200oF

150oF

76oF

Figure 5-8. Effect of temperature on the decay of the anodic current after a

scratch has been performed on the surface of a 13Cr Alloy.

Figure 5-10 and Figure 5-11 show the effect of temperature on the current decays

for Super 13Cr and 22Cr, respectively. The trend of the current decay with temperature

for Super 13Cr and 22Cr is similar to the one observed for 13Cr. The higher the

temperature, the higher the currents were for the three alloys. However, results suggest

that the effect of temperature on the current transients is significantly stronger for 13Cr

than it is for Super 13Cr and 22Cr. In fact, for 22Cr no clear distinction among the

behaviors of the current transients with temperature after the first 20 seconds was found.

87

13Cr AlloypH=4.0

0.0E+00

4.0E-06

8.0E-06

1.2E-05

1.6E-05

2.0E-05

2.4E-05

2.8E-05

3.2E-05

0 50 100 150 200Time (sec)

Thic

knes

sLo

ss(m

m)

13Cr, T=76oF

13Cr, T=150oF

13Cr, T=200oF

Figure 5-9. Effect of temperature on the cumulative thickness loss experienced by

13Cr after being scratched at pH=4.0.

0

20

40

60

80

100

120

140

160

180

200

0 50 100 150 200Time (sec)

I( m

A)

Super 13Cr AlloypH= 4.0

200oF

150oF76

oF

Figure 5-10. Effect of temperature on the decay of the anodic current after a

scratch has been performed on the surface of a Super 13Cr Alloy.

88

0

5

10

15

20

25

30

35

40

45

50

0 50 100 150 200Time (sec)

I (mA

)

200oF

150oF76oF

22Cr AlloypH= 4.0

Figure 5-11. Effect of temperature on the decay of the anodic current after a

scratch has been performed on the surface of a 22Cr Alloy.

At the elevated temperatures, the 1/I slopes (for a fixed pH) are smaller than

slopes found at lower temperatures, which indicates the longer repassivation times

needed to reform the passive film at higher temperatures. Trends of the 1/I slopes with

temperature at pH 4 for the three alloys are clearly seen in Figure 5-12 to Figure 5-14. At

a given pH, mT=76 > mT=150> mT=200 which predicts higher corrosion components of the

erosion-corrosion process for higher temperatures. Notice that slopes for 22Cr at different

temperatures are closer together than those for Super 13Cr and 13Cr; hence, the effect of

temperature on the erosion-corrosion damage of 22Cr is expected to be smaller than that

experienced by Super13Cr and 13Cr. The effect of temperature at other pH values

showed trends similar to those shown for pH = 4 (Appendix A)

89

0.E+00

1.E+05

2.E+05

3.E+05

4.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

)

13CrpH= 4.0

T=200oF

T=150oF

T=76oF

Figure 5-12. 1/I vs. t for 13Cr from the raw data in Figure 5-8.

0.E+00

1.E+05

2.E+05

3.E+05

4.E+05

0 50 100 150 200Time (sec)

1/I

,(1

/A)

Super 13CrpH= 4.0

T=76oFT=150oF

T=200oF

Figure 5-13. 1/I vs. t for 13Cr from the raw data in Figure 5-10.

90

0.E+00

1.E+05

2.E+05

3.E+05

4.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

)

22CrpH= 4.0T=200oF

T=76oF

T=150oF

Figure 5-14. 1/I vs. t for 13Cr from the raw data in Figure 5-11.

Effect of type of material (CRA) on Scratch Test results

Figure 5-15 shows a comparison of current response among the three CRAs at a

temperature of 150oF and pH of 4. For a given pH and temperature, the current response,

and thus the erosion-corrosion rate, for Super 13Cr is much lower than that for 13Cr, but

not as low as for 22Cr. As expected, the Super 13Cr is performing better than 13Cr, but

not as well as 22Cr. Scratch tests made at 76oF and 200oF indicate that the ranking of the

alloys is the same regardless of the temperature.

As stated earlier, the second order model was found to be a very good

approximation to describe the repassivation process of CRAs under all conditions tested

at least for the first 200 seconds. Hence, the plot of 1/I vs. time shows roughly linear

91

behavior as shown in Figure 5-16 (for several materials at different pH values) and can be

represented by Equation (5-2).

0

20

40

60

80

100

120

140

0 50 100 150 200Time (sec)

I (mA

)

13Cr_pH

22CrSuper_13Cr

T = 150ºFpH = 4.0

Figure 5-15. Decay of the anodic current for three different alloys.

0.E+00

2.E+05

4.E+05

6.E+05

0 50 100 150 200Time (sec)

1/I ,

(1/A

)

13Cr_pH=4.0

13Cr_pH=5.25

13Cr_pH=3.5

13Cr_pH=6.0

T= 150 ºF

22Cr_pH=4.0Super13Cr_pH=3.5Super13Cr_pH=4.0

Super13CrpH=6.0

Super13CrpH=5.25

Figure 5-16. Comparison of 1/I vs. t for 13Cr and Super 13Cr at 150°F, 22Cr at pH

4 is also included.

92

For a given pH and temperature, the Super 13Cr slope is larger than the

corresponding slope found for 13Cr, which indicates a shorter repassivation time is

needed for Super 13Cr to reform the passive film at similar conditions. At a pH of 4,

m22Cr > mSuper 13Cr> m13Cr at 76, 150 or 200oF, which indicates the highest corrosion rate

of the erosion-corrosion process is for 13Cr and the lowest is for 22Cr, while the Super

13Cr corrosion rate lies between.

Cumulative Thickness Loss and Repassivation Time

Appendix A contains the cumulative thickness loss data based on the integration

of all the current data presented in the previous three sections. From these charts, the

trends are very clear. For a fixed temperature, the corrosion component of the erosion-

corrosion damage is more severe at lower pH values where cumulative thickness losses

are higher. Also notice that the thickness loss for a fixed pH is higher for higher

temperatures. In general, for a given pH and temperature, the cumulative thickness losses

for Super 13Cr are much lower than those for 13Cr, but not as low as for 22Cr. This

relation would predict less erosion-corrosion damage on 22Cr than on Super 13Cr or

13Cr.

As a convenient form of summarizing all the scratch test data in a single plot the

cumulative thickness loss after 200 seconds is shown in Figure 5-17 for the three CRAs

using two different approaches. The first approach is the estimation of the cumulative

thickness loss obtained from integration of the current data time series up to 200 seconds.

93

The second approach, the estimation of the cumulative thickness loss at time t = 200 s

was made by means of the second order approximation using equation (5-3).

As can be seen from Figure 5-17 the higher the temperature the higher the metal

thickness loss due to the corrosion component of the erosion-corrosion process. The three

alloys showed the same trend with temperature. Regardless of temperature the 13Cr

presented the highest thickness loss and the 22Cr the lowest. Figure 5-18 shows the effect

of temperature on the ratios of the thickness loss of 13Cr to Super 13Cr and 22Cr at pH =

4 (from data shown in Figure 5-17). Notice the thickness loss for 13Cr at 76 oF is about

twice the thickness loss of the Super 13Cr and about 5 times the thickness loss of the

22Cr at the same temperature. These factors increase slightly with increasing temperature

as shown in Figure 5-18.

0.0E+00

5.0E-06

1.0E-05

1.5E-05

2.0E-05

2.5E-05

3.0E-05

Thic

knes

sL

oss

(mm

)

13Cr, I dataIntegration

13Cr2ndOrder

S13Cr, I dataIntegration

S13Cr2ndOrder

22Cr, I dataIntegration

22Cr2ndOrder

76

150200

Corrosion Resistant Alloy

Temp.

(oF)

76150200

200

15076

pH = 4

Figure 5-17. Comparison of the cumulative thickness loss after 200 seconds for the

three CRAs at pH = 4 and different temperatures.

94

0

2

4

6

8

10

12

14

16

18

50 100 150 200 250

Temperature

Th

ickn

ess

Lo

ssA

lloy

Rat

io

TL (13Cr/Super13Cr) TL (13Cr/22Cr)

pH = 4

Figure 5-18. Effect of temperature on the ratios of the Thickness Loss of 13Cr to

Super 13Cr and to 22Cr.

The repassivation time may be defined as the time needed for the anodic current

caused by the scratch, to return to a small value “passive current”. Here the repassivation

time has been defined as the time needed for the corrosion rate at the scratch site to

decrease to a value which is low enough to be considered as an acceptable corrosion rate.

Two approaches were used to define the repassivation time. Using the current data time

series, the repassivation time was the time needed for the current value to return to its

original value observed before making the scratch (or its noise level). For applying the

second order model, the approach considered was the time needed for the corrosion rate

to decrease to 1 mpy. Based on this definition, and the assumption that the repassivation

process follows second order kinetics for the entire healing period, equation (5-4) can be

used to estimate the repassivation time for each condition tested. Table 5-1 summarizes

the repassivation time in minutes (trp) for 13Cr alloy obtained for every condition tested

using the current data time series. Similarly, Table 5-2 summarizes the repassivation time

95

in minutes (trp) for 13Cr at every condition tested estimated using the second order

kinetics approximation. Trends are similar for the two approaches. The repassivation

times are longer for higher temperatures and shorter for higher pH. These results suggest

that the corrosion component of the erosion-corrosion phenomenon on 13Cr is more

severe for higher temperatures and lower pH.

Repassivation times in minutes, for Super 13 Cr and 22Cr, for both approaches

are shown in Table 5-3. In general, the repassivation times for Super 13Cr and 22Cr were

much lower than those for 13Cr at similar conditions; 22Cr having the lowest

repassivation times. The trend of repassivation time with temperature for Super 13Cr is

similar to that seen for 13Cr: the higher the temperature, the longer the time needed to

obtain the original low current values. The effects of pH on repassivation times are also

similar for both 13Cr and Super 13Cr; but, Super 13Cr data are just shown for pH 4. No

clear trend was observed for the effect of temperature on the repassivation times of 22Cr

since the data points for different temperatures are so close to each other for 22Cr.

Table 5-1. Repassivation times in minutes for 13Cr at different test conditions using

current data time series approach.

T

pH

76°F 150°F 200°F

3.50 112 148 169

4.00 34 136 262

4.75 19 102 185

5.25 16 66 78

6.00 3 34 30

96

Table 5-2. Repassivation times in minutes for 13Cr at different test conditions using

2nd order kinetics approximation.

T

pH

76°F 150°F 200°F

3.50 155 187 339

4.00 43 104 209

4.75 33 90 187

5.25 22 71 55

6.00 5 43 84

Table 5-3. Repassivation times in minutes at pH = 4 and the three temperatures for

Super 13Cr and 22Cr using both approaches, current data time seriesapproach and 2nd order kinetics approximation.

Material 76°F 150°F 200°F

From current data time

series. Super13Cr 17 25 51

From 2nd Order kinetics

approx. Super 13Cr 18 26 74

From current data timeseries. 22Cr 8 12 5

From 2nd Order kinetic

approx. 22Cr 8 11 9

Trends for repassivation times of 13Cr and Super 13Cr using both approaches are

also clearly seen on Figure 5-19 and Figure 5-20 where data for 22Cr at pH = 4 and the

three temperatures are also included for comparison purposes. Trend lines as straight

lines are included to aid the viewing of trends; however, pH repassivation time and

temperature repassivation time relations may not be linear. As can be seen, the

97

repassivation times for Super 13Cr and 22Cr are much shorter than those for 13Cr for a

fixed pH and temperature, which indicates that Super 13Cr and 22Cr heal much faster

than 13Cr at the same conditions, with 22Cr being the fastest. Comparisons of

repassivation times among the three alloys are more clearly seen for pH 4 in Figure 5-21.

Figure 5-22 shows the comparison between repassivation times provided by the

current data time series and the repassivation times estimated from the second order

kinetics approximation. As can be seen, the agreement is very good, particularly for

repassivation times of 50 minutes and below. The most scattered data points (diamonds)

belong to 200oF testing and may be explained as due to experimental limitations in

keeping the test solution from evaporating in a bubbling test cell setup (for 13Cr testing).

A new closed experimental setup was designed to keep the pressure slightly above

atmospheric pressure reducing liquid evaporation and reducing data scatter (Super 13Cr

data).

98

0

40

80

120

160

200

240

280

3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5pH

Rep

assi

vati

on

tim

e(m

in) 200

150

76

200 (S13Cr)

150 (S13Cr)

76 (S13Cr)

200 (22Cr)

150 (22Cr)

76 (22Cr)

T (oF)

13Cr 200 oF13Cr 150 oF

13Cr 76 oFS13Cr 200 oF

} 22Cr All Temp.

S13Cr 150 oFS13Cr 76 oF

Figure 5-19. Effect of pH and temperature on the repassivation times of 13Cr at

different test conditions. Data for Super 13Cr and 22Cr at pH 4 and

the three temperatures are also included.

0

10

20

30

40

50

60

3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5

pH

Rep

assi

vatio

ntim

e(m

in) 200 (S13Cr)

150 (S13Cr)

76 (S13Cr)

200 (22Cr)

150 (22Cr)

76 (22Cr)

T (oF)S13Cr 200 oF

S13Cr 150 oF

S13Cr 76 oF

22Cr 200 oF

22Cr 150 oF

22Cr 76 oF

Figure 5-20. Effect of pH and temperature on the repassivation times of Super

13Cr at different test conditions. Data for 22Cr at pH 4 and the three

temperatures are also included.

99

0

50

100

150

200

250

300

Rep

asiv

atio

nTi

me

(min

)

13Cr, I datatime series

13Cr2ndOrder

S13Cr, I datatime series

S13Cr2ndOrder

22Cr, I datatime series

22Cr2ndOrder

76

150

200

Corrosion Resistant Alloy

Temp. (oF)

76150200

200150

76

pH = 4

Figure 5-21. Comparison of the repassivation times for the three CRAs at pH 4 and

different temperature.

0

50

100

150

200

250

300

0 100 200 300

2ndOrder model Repassivation Time (min)

Rep

asiv

atio

nT

ime

(min

)fr

omI

data

tim

ese

ries 76 (13Cr)

150 (13Cr)200 (13Cr)76 (S13Cr)150 (S13Cr)200 (S13Cr)

Temp oFPerfect AgreementLine

Figure 5-22. Comparison between repassivation times obtained from actual data

and estimated from the second order kinetic approximation.

100

CHAPTER 5 ..................................................................................................................................76

SCRATCH TEST AS A SIMPLIFIED EROSION-CORROSION TEST .....................................................76

Motivation for Doing Scratch Test ................................................................ ........................76

Data reduction technique ................................ ................................................................ ......77

Scratch Test Results................................ ................................................................ ...............81

Effect of pH on Scratch Test results ..................................................................................................... 81

Effect of temperature on Scratch Test results....................................................................................... 85

Effect of type of material (CRA) on Scratch Test results.................................................................... 90

Cumulative Thickness Loss and Repassivation Time ............................................................92

FIGURE 5-1. ANODIC CURRENT DECAY DURING THE SCRATCH REPASSIVATION PROCESS. ..........................77

FIGURE 5-2. DATA SHOWING A LINEAR RELATION BETWEEN 1/I AND TIME. ................................ ...............78

FIGURE 5-3. COMPARISON BETWEEN TLACTUAL AND TLCAL FOR 13CR ALLOY. ................................ ......80

FIGURE 5-4. EFFECT OF PH ON THE DECAY OF THE ANODIC CURRENT AFTER A SCRATCH HAS BEEN MADE

ON THE SURFACE OF A 13CR ALLOY................................................................................................. ....83

FIGURE 5-5. EFFECT OF PH ON THE DECAY OF THE ANODIC CURRENT AFTE R A SCRATCH HAS BEEN MADE

ON THE SURFACE OF A SUPER 13CR ALLOY..........................................................................................83

FIGURE 5-6. 1/I VS. T FOR 13CR FROM THE RAW SCRATCH TEST DATA IN FIGURE 5-4. ...............................84

FIGURE 5-7. 1/I VS. T FOR 13CR FROM THE RAW DATA IN FIGURE 5-5. .......................................................85

FIGURE 5-8. EFFECT OF TEMPERATURE ON THE DECAY OF THE ANODIC CURRENT AFTER A SCRATCH HAS

BEEN PERFORMED ON THE SURFACE OF A 13CR ALLOY........................................................................86

FIGURE 5-9. EFFECT OF TEMPERATURE ON THE CUMULATIVE THICKNESS LOSS EXPERIENCED BY 13CR

AFTER BEING SCRATCHED AT PH=4.0. ..................................................................................................87

FIGURE 5-10. EFFECT OF TEMPERATURE ON THE DECAY OF THE ANODIC CURRENT AFTER A SCRATCH HAS

BEEN PERFORMED ON THE SURFACE OF A SUPER 13CR ALLOY.............................................................87

FIGURE 5-11. EFFECT OF TEMPERATURE ON THE DECAY OF THE ANODIC CURRENT AFTER A SCRATCH HAS

BEEN PERFORMED ON THE SURFACE OF A 22CR ALLOY........................................................................88

FIGURE 5-12. 1/I VS. T FOR 13CR FROM THE RAW DATA IN FIGURE 5-8. ..................................................89

FIGURE 5-13. 1/I VS. T FOR 13CR FROM THE RAW DATA IN FIGURE 5-10. ................................................89

FIGURE 5-14. 1/I VS. T FOR 13CR FROM THE RAW DATA IN FIGURE 5-11. ................................................90

FIGURE 5-15. DECAY OF THE ANODIC CURRENT FOR THREE DIFFERENT ALLOYS................................. ....91

FIGURE 5-16. COMPARISON OF 1/I VS. T FOR 13CR AND SUPER13CR AT 150°F, 22CR AT PH4 IS ALSO

INCLUDED................................. ............................................................................................................91

FIGURE 5-17. COMPARISON OF THE CUMULATIVE THICKNESS LOSS AFTER 200 SECONDS FOR THE THREE

CRAS AT PH = 4 AND DIFFERENT TEMPERATURES................................................................. ...............93

101

FIGURE 5-18. EFFECT OF TEMPERATURE ON THE RATIOS OF THE THICKNESS LOSS OF 13CR TO SUPER 13CR

AND TO 22CR........................................................................................................................................94

FIGURE 5-19. EFFECT OF PH AND TEMPERATURE ON THE REPASSIVATION TIMES OF 13CR AT DIFFERENT TEST

CONDITIONS. DATA FOR SUPER 13CR AND 22CR AT PH 4 AND THE THREE TEMPERATURES ARE ALSO

INCLUDED................................. ............................................................................................................98

FIGURE 5-20. EFFECT OF PH AND TEMPERATURE ON THE REPASSIVATION TIMES OF SUPER 13CR AT

DIFFERENT TEST CONDITIONS. DATA FOR22CR AT PH 4 AND THE THREE TEMPERATURES ARE ALSO

INCLUDED. 98

FIGURE 5-21. COMPARISON OF THE REPASSIVATION TIMES FOR THE THREE CRAS AT PH 4 AND DIFFERENT

TEMPERATURE......................................................................................................................................99

FIGURE 5-22. COMPARISON BETWEEN REPASSIVATION TIMES OBTAINED FROM ACTUAL DATA AND

ESTIMATED FROM THE SECOND ORDER KINETIC APPROXIMATION................................. ........................99

TABLE 5-1. REPASSIVATION TIMES IN MINUTES FOR 13CR AT DIFFERENT TEST CONDITIONS USING

CURRENT DATA TIME SERIES APPROACH. ..............................................................................................95

TABLE 5-2. REPASSIVATION TIMES IN MINUTES FOR 13CR AT DIFFERENT TEST CONDITIONS USING 2ND

ORDER KINETICS APPROXIMATION. .......................................................................................................96

TABLE 5-3. REPASSIVATION TIMES IN MINUTES AT PH = 4 AND THE THREE TEMPERATURES FOR SUPER 13CR

AND 22CR USING BOTH APPROACHES, CURRENT DATA TIME SERIES APPROACH AND 2ND ORDER KINETICS

APPROXIMATION. ..................................................................................................................................96

60 A. McMahon and J. Martin “Simulation Tests on the Effect of Mechanical Damage orAcid Cleaning on CRAs Used for Oil/Gas Production Well

Tubulars”,CORROSION/04, paper no. 4127, 2004, (Houston, TX, USA, NACE

International).

100

CHAPTER 6

MULTIPHASE GAS/LIQUID/SAND FLOW LOOP TESTING RESULTS

Pure erosion (N2-distilled water-sand) and erosion-corrosion (CO2-Brine-sand)

tests were conducted in a plugged tee geometry. A high resolution electrical resistance

(ER) technique was used to measure the penetration rate at both conditions.

Two different set of conditions were tested. All testing shown in this chapter

refers to erosion and erosion-corrosion tests conducted for multiphase flow conditions

(gas/liquid/sand). The main difference between the two sets of data was the sand rate.

Given the synergistic effect seen for 13Cr alloys at high erosivity conditions59, oil

companies showed interest in checking for the erosion-corrosion behavior of 13Cr alloy

at very low sand rate, similar to those levels referred to in the field as “background sand

levels”. Therefore, the first set of tests was performed at low sand concentrations lying

between 10 and 15 lb/day for 76 and 150oF. A second set of data from multiphase flow

loop tests was for a high sand rate level with two main objectives. The first, was to

compare the erosion-corrosion behavior of 13Cr at high erosivity condition with the

behavior of other CRAs tested at high sand rate. The second was to provide comparison

data for the predictions made by means of the scratch tests and the proposed erosion-

corrosion model.

101

Low Sand Rates (Multiphase Flow Loop Testing)

Previous estimation of erosion and erosion-corrosion rates for 13Cr exposed to

multiphase flow conditions had mainly relied on the weight-loss method for an elbow

geometry.59 Even though the weight-loss method yields accurate erosion-corrosion

penetration rates, it is generally very time consuming, especially when the erosion or

corrosion rates are very low. In an attempt to evaluate the background sand level

conditions experienced in the field, low sand rate loop testing for erosion and erosion-

corrosion conditions was performed at both 76 and 150oF under multiphase flow

conditions. Previous testing at similar conditions was done by means of the weight loss

data, but a high sand rate (48 lb/day) was needed to obtain measurable amounts of mass

loss. The high resolution electrical resistance, ER, probe allows reducing the sand rate to

as low as 10 lb/day while keeping reasonable accuracy on the estimation of the

penetration rates in 5.5-hour tests.

Testing conditions were slightly different from those tested before for an elbow

geometry. The ER probe was set flush mounted at the top of a vertical pipe run at a 90

degree turn for a 1” piping. Flow geometry for this configuration is closer to a 1” plugged

tee than an elbow.

Typical output readings of the ER measurements for both types of tests (erosion

and erosion-corrosion) are displayed on Figure 6-1 and Figure 6-2. Temperature readings

during the test period are also included in plots. The procedures for testing both erosion

and erosion-corrosion conditions were the same. Sand-free conditions with both liquid

102

and gas phases flowing through the test cell were run long enough to stabilize system

temperature and metal loss readings. At least 1 hour of testing was conducted under

stable temperature conditions. During this period of the test, just the corrosion component

would cause a metal loss increment when a CO2-brine test is performed. However, no

measurable metal loss increments were observed in this period (sand-free) as shown by

the plateau for the first and last hour of erosion-corrosion testing shown in Figure 6-2.

This suggests that the pure corrosion component at tested conditions is very low,

probably below 1 mpy. After the initial, sand-free test period, dry sand was injected into

the system at the desired rate by means of a DC-motor driven vibratory sand feeder

system. Sand injection time is indicated on both Figure 6-1 and Figure 6-2 for pure

erosion and erosion-corrosion conditions, respectively. The sand was run for similar

times for both conditions (~5.5 hours) and then stopped. Later, the system was kept

running at sand-free conditions to again check the no-metal-loss increment response

under these conditions. Penetration rates were estimated from the slope of the metal-loss

vs. time plots. As can be observed at the end of the test shown in Figure 6-2, the metal-

loss output readings are pretty sensitive to changes in temperature. Once the temperature

was stable the metal loss noise level reduced to levels low enough to allow the

penetration rates to be measured with reasonable accuracy.

Some sharp spikes are observed in the output readings for both figures during the

stable temperature period. Notice that those spikes are caused by a sudden change in

temperature due to recharging of the condensed (colder) water back to the main liquid

tank separator from which the test cell is being fed. However, once the temperature

transients abated, the metal-loss reading returned back to similar previous values. This

103

few minutes transition period is not expected to affect the total slope much; therefore

spikes due to temperature transients were neglected in the estimation of the penetration

rates. The raw data was down-loaded to an Excel spread sheet and the changes in metal

loss due to changes in temperature were corrected before estimating the slope of the

metal-loss vs. time plots. Actually no large differences were found between spreadsheet

values and those provided by the commercial software.

y = 29.3x - 21.0pure E = 10.1 mpy

0

50

100

150

200

250

0 1 2 3 4 5 6 7 8Time (hr)

Met

alLo

ss(n

m)

0

20

40

60

80

100

120

140

160

Tem

po F

StartSand

StopSand

MetalLoss

Temperature

1” plugged T

Figure 6-1. Pure erosion test for 13Cr under sand-N2-distilled water flow system

(Vsg=60 ft/s, Vsl= 0.2 ft/s) at 150oF and low sand rate (15 lb/day).

The total thickness loss obtained for a low sand rate test was as low as 100 nm for

a 5.5-hour test. It is well known that the ER probes compensate for temperature effects.

However, at nanometer levels of precision, extremely small changes in temperature can

cause large changes in the output reading of the ER probe. Therefore, maintaining a

uniform sand rate during the 6-hour test at low levels of sand was also difficult to

achieve. Non-uniformity of sand rates during the 6-hour tests may be another source of

104

uncertainty in the results. Therefore, a statistical analysis of the results obtained was used

to draw some conclusions.

0

50

100

150

200

250

0 1 2 3 4 5 6 7 8Time (hr)

Met

alL

oss

(nm

)

0

20

40

60

80

100

120

140

160

Tem

per

atu

re.

(oF

)y = 30.1x - 7.9E-C = 10.4 mpy

StartSand

StopSand

MetalLoss

Temperature

1” plugged T

Figure 6-2. Erosion-corrosion test for 13Cr under sand-CO2-Brine flow system

(Vsg = 60 ft/s, Vsl = 0.2 ft/s) at 150oF and low sand rate (15 lb/day).

Several tests were run at 76 and 150oF, for the 1” plugged tee geometry at low sand rates.

Pure erosion (N2-distilled water) conditions and erosion-corrosion (CO2-brine) conditions

were assessed.

Figure 6-3 summarizes the penetration rate for 5 pure erosion tests and 5 erosion-

corrosion tests of 13Cr in a flush mounted plugged tee geometry tested at room

temperature (76oF). The averages for both pure erosion and erosion-corrosion processes

are also included in the plot.

In general, it can be seen that the erosion-corrosion penetration rates are slightly

higher than the pure erosion penetration rates. When a small sample of data is considered

(N < 31), a “student’s t” statistical analysis can be made to interpret the results and

105

estimate how well the obtained average is. A student’s t analysis to estimate the interval

within which the population average, X, will be contained approximately 95% of the time

was performed for both erosion and erosion-corrosion data. As can be noticed, the 95%

confidence intervals for both the erosion-corrosion average and the pure erosion average

overlap to a large extent. An additional F-distribution based analysis made on the data

shown in Figure 6-3, indicate with 95 % confidence that the erosion-corrosion rate is not

higher than the pure erosion rate. In any case the synergistic effect seen in previous

results conducted in an elbow geometry at higher sand rates was not observed for these

low sand rate data sets at 76oF.

0

1

2

3

4

5

Test #

Pen

etra

tion

rate

(mpy

)(E

Rpr

obe

)

E (mpy) 3.0 3.1 2.0 4.1 3.0 3.0

E-C (mpy) 3.1 4.6 3.9 3.5 3.0 3.6

1 2 3 4 5 Avg

Figure 6-3. Penetration rate for pure Erosion and Erosion-Corrosion tests of

13Cr in multiphase flow testing at 76oF. (The errors bars on the

average represent the 95% confidence interval on the mean value.)

Similar tests were also conducted at a higher temperature (150oF). To be able to

run similar tests at higher temperatures a minor modification of the sand feeder was

106

made. The clear PVC pipe used in the sand feeder system to introduce the dry sand into

the flow-stream was change to a stainless steel pipe. The sand bed on this pipe was no

longer visible, therefore the sand rates were increased a little (from 10 to 15 lb/day

approximately) to ensure sand flow into the system.

Figure 6-4 summarizes the penetration rate for 3 pure erosion tests and 3 erosion-

corrosion tests of 13Cr in a flush mounted plugged tee geometry tested at 150oF and 15

lb/day. The averages for both pure erosion and erosion-corrosion processes are also

included in the plot, as well as the 95% confidence interval error bars. As can be seen, the

scatter of the pure erosion data set of the 150oF data was higher than the scatter found for

the remaining of the data sets. Additionally the number of tests run was lower than the

number of the test run at 76oF which makes the uncertainty of these results larger.

However, similar trends as obtained for 76oF can be seen for 150oF as well. The erosion-

corrosion penetration rates are slightly higher than the pure erosion penetration rates. But

if there is an erosion-corrosion synergistic effect for this temperature at low sand rate

conditions, it is not very significant.

Care must be taken when scaling up the flow loop test results to extrapolate them

to field conditions. Spatial distribution or density of sand over the cross section needs to

be considered to evaluate how erosive a specific condition may be. A sand rate of 10

lb/day represents a sand mass flux of about 12 lb/day-in2 for the 1” piping used in our

tests. By contrast, the same sand mass flux would represent a sand mass rate of about 89,

153 and 600 lb/day for 3”, 4” and 8” pipes diameter (schedule 40) respectively. Or, in

other words, a sand rate of 10 lb/day, which represents a sand flux of about 12 lb/day-in2

107

in 1” piping, would represent sand fluxes of only about 1.4, 0.8 and 0.2 lb/day-in2 for 3”,

4” and 8” pipes diameter (schedule 40) respectively.

This analysis suggests that erosivities for laboratory flow loop tests conducted at

low sand rates may still be considered more aggressive than those seen in field at

“background sand levels” conditions. Hence, in light of the laboratory results, no

significant erosion-corrosion synergistic effect of 13Cr is expected for low erosivity

conditions for either 76 or 150oF.

0

2

4

6

8

10

12

14

16

Pen

etra

tion

rate

(mp

y)(E

Rp

rob

e)

E (mpy) 10.1 7.2 5.0 7.4

E-C (mpy) 10.4 8.8 11.0 10.0

1 2 3 AvgTest #

Figure 6-4. Penetration rate for pure Erosion and Erosion-Corrosion tests of 13Cr

in multiphase flow testing at 150oF. (The errors bars on the average

represent the 95% confidence interval on the mean value.)

108

High Sand Rates (Multiphase Flow Loop Testing)

A set of multiphase flow loop tests for 13Cr, Super 13Cr and 22Cr at high

erosivity conditions was also performed. High resolution electrical resistant probes made

of the three CRAs under study were obtained and used in a gas/liquid/sand flow facility

to measure pure erosion rates and erosion-corrosion rates. For these sets of testing a

special fitting was use to configure the ER probes in a ½” pipe test section for the

multiphase flow loop. With the new arrangement, the same plugged tee flow geometry as

used for low sand rate experiments was tested, but now the clearance between the

cylindrical probe and the ½” piping was much smaller than the one between the probe

and the 1” piped test section used before. Hence, using the ½” pipe system assured that

the probe would be directly exposed to the impingement of the entire sand through put,

obtaining much higher erosivity conditions than those for the 1” geometry. In addition, a

much higher sand flux was used for the high sand rate testing (30-99 lb/day-in2 for the ½”

pipe) as compared to the low sand rate testing (10-12 lb/day-in2 for the 1” pipe).

Therefore, as far as sand concentration is concerned, the “high sand rate” erosivities were

10 times more aggressive than the “low sand rate” erosivities. However, flow conditions

also were different. A superficial gas velocity of 20ft/s and superficial liquid velocity of

1.4 ft/s were used for the ½” pipe tests and superficial gas velocity of 60ft/s and

superficial liquid velocity of 0.2 ft/s were used for the 1” pipe tests, both conditions were

annular flow.

The testing procedure was similar to that used for determination of erosion-

corrosion of 13Cr alloy under low sand concentration presented in the previous section.

109

Hence, the output reading trends of the ER measurements for both types of tests (erosion

and erosion-corrosion) are similar to those displayed on Figure 6-1 and Figure 6-2 in the

previous section, even though the magnitudes of slopes, and thus the erosion and erosion-

corrosion rates were much higher.

Table 6-1 contains the pure erosion and erosion-corrosion penetration rates for the

three alloys tested at similar conditions (76oF, pH 4, Vsg = 20 ft/s, Vsl = 1.4 ft/s, 30 lb/day

of sand). The total corrosion component rates, Ce-c of the erosion-corrosion process, were

estimated by subtracting the pure erosion rates, E from the total erosion-corrosion rates

EC. Trends found are as expected. In general the erosion-corrosion rates were higher than

the pure erosion rate for each of the three materials. Considering that the pure corrosion

rates for these alloys have been proved to be smaller than 2 mpy at the conditions tested,

the current results suggest that a synergistic effect is observed in the erosion-corrosion

processes of the three alloys, provided that the erosivity is high enough. However, the

degree of synergism is quite different for the three alloys: at the much lower sand rates

typically found in the field, the synergism could be negligible for 22Cr, or at least never

observed in the field.

The ratio of EC to pure erosion (E) for 13Cr was the highest, with a value of 2.2

followed by Super 13Cr at 1.6 and 22Cr at 1.4. Slight differences, if any, were seen in the

pure erosion penetration rates between the three alloys, which may fall within the

reproducibility errors of the measurements. Unfortunately, due to lack of time and

resources, a rigorous reproducibility procedure was not followed this time as in previous

data sets. However experience gathered by running the multiphase flow loop and the ER

probes for similar testing procedures has shown the 95% confidence interval of the

110

penetration rate estimation falling within 10 to 30% of the mean value. If the worst case

scenario (30%) is adopted for data shown in Table 6-1, one may get a sense of whether

the erosion-corrosion rate is higher than pure erosion rates for a given condition. Figure

6-5 shows data from Table 6-1 including the assumed error interval of 30%. This plot

strongly suggests that the erosion-corrosion rate is higher than the pure erosion rate for

13Cr at 76oF. For super13Cr this effect is less pronounced, and for 22Cr may not be

significant. Notice, for 13Cr, the corrosion component is even higher than the pure

erosion for what may be defined as corrosion dominated erosion-corrosion process.

Erosion and erosion-corrosion data were also gathered for the 3 alloys exposed to

similar flow conditions but a higher temperature of 150oF (see Table 6-2, and Figure 6-6).

In general, trends found at 150oF are similar to those observed at 76oF. A point to

highlight is that the corrosion component of the Super13Cr, which was lower than the

pure erosion rate at 76oF, is higher than the pure erosion rate at 150oF; hence, at this

temperature the erosion-corrosion to erosion ratio is significant for Super13Cr as it is for

13Cr. For 22Cr, erosion-corrosion rates and erosion-corrosion to erosion ratios are low

and quite similar regardless of the temperature. Meaning the repassivation rates are not

strongly affected by the temperature, as was also seen in scratch test results obtained for

this alloy in Chapter 5.

111

Table 6-1. Erosion-corrosion (EC), pure erosion (E) and corrosion components

(Ce-c) of the erosion-corrosion penetration rates for 3 alloys tested atsimilar conditions. 76oF, pH 4, Vsg = 20 ft/s, Vsl = 1.4 ft/s, 30 lb/day of

sand.

Alloy

EC rate (mpy)

ER Probe

E rate (mpy)

ER Probe

Ce-c rate (mpy)

*Ce-c =EC- E

13Cr 179 80 99

Super 13Cr 123 76 47

22Cr 95 66 29

*Obtained by subtracting the pure erosion from the total erosion-corrosion.

Figure 6-5. Erosion-corrosion (EC), pure erosion (pure E) and corrosion

component (CEC) of the erosion-corrosion penetration rates for data

shown in Table 6-1.

0

50

100

150

200

250

13Cr alloy Super13Cr alloy 22Cr alloyMaterial

Pene

trat

ion

rate

(mp

y)

ER Probeflush mounted

1/2" plugged tee

76oF

112

Table 6-2. Erosion-corrosion (EC), pure erosion (pureE) and corrosion components

(CEC) of the erosion-corrosion penetration rates for 3 alloys tested atsimilar conditions. 150oF, pH 4, Vsg = 20 ft/s, Vsl = 1.4 ft/s, 30 lb/day of

sand .

Alloy

EC rate (mpy)

ER Probe

pureE rate (mpy)

ER Probe

CEC rate (mpy)

*CEC =EC- pureE

13Cr 303 96 207

Super13Cr 175 72 103

22Cr 108 83 25

* Obtained by subtracting the pure erosion from the total erosion-corrosion.

Figure 6-6. Erosion-corrosion (EC), pure erosion (pureE) and corrosion

component (CEC) of the erosion-corrosion penetration rates for data

shown in Table 6-2

In summary can be said that at high erosivity conditions, a synergistic effect

between erosion and corrosion was confirmed since the total metal loss rate was shown to

be higher than the sum of the rates of corrosion (without sand) and erosion measured

0

50

100

150

200

250

300

350

13Cr alloy Super13Cr alloy 22Cr alloyMaterial

Pen

etra

tion

rate

(mpy

)

ER Probeflush mounted

1/2" plugged tee

150oF

113

separately. It appears that under the high sand rate conditions tested, the erosivity is

severe enough to damage the passive layer thereby enhancing the corrosion rate.

Synergism seems to occur for the three alloys; however, the degree of synergism is quite

different for the three alloys and may not be significant for 22Cr under field conditions

where erosivities are typically much lower that those occurring in the small bore loop

used in this testing.

For the other two alloys, 13Cr and Super13Cr, especially at high temperatures, if

the erosion rate of the passive film is great enough, then an accelerated erosion-corrosion

process may take place with a significant contribution from a corrosion component.

114

CHAPTER 7

COMPARISON OF SCRATCH TEST RESULTS WITH FLOW LOOP

TEST RESULTS

An expression to estimate the corrosion component of the erosion corrosion, Ce-c

for the conditions being studied has not been yet constructed, however, information

generated with the scratch test cell method presented in Chapter 5 may provide the basis

for a procedure intended to predict the corrosion component, Ce-c of CRAs exposed to

environments containing CO2 and sand.

This section compares some of the trends obtained from scratch test results with

loop test data assuming that the scratch test results relate to the total corrosion component

of the erosion-corrosion process caused by the scratch (e.g. Ce-c). This suggests that the

scratch test results should be comparable to the estimation of the “total corrosion” in a

loop test conducted under CO2-brine-sand flow conditions, where the “total corrosion”

component is the summation of the pure corrosion (sand-free corrosion) and the corrosion

increment induced by the sand impingement process.

Hence, the scratch test data is being compared here to the single phase liquid flow

loop data presented in previous publications57,58 for 13Cr alloy, as well as to the two-

phase flow loop data gathered for 13Cr, Super13Cr and 22Cr and presented in Chapter 6.

115

In Chapter 5, it was suggested that the ratio of the slopes obtained from scratch

test might be a useful indicator of the relative severity of two different erosion-corrosion

conditions. Equation (5-6) defined the thickness loss ratio for a certain condition B with

respect to baseline condition A as the ratio of the slope at condition A to the slope at

condition B. Where A and B would represent different environmental conditions or

different alloys.

If the relative severity of the erosion-corrosion of a flowing system is defined as

the ratio of the “total corrosion” rate for two different erosive conditions A and B, then

the relative severity of the same system can be also obtained as the inverse of the slopes

obtained from scratch tests at similar conditions, as expressed by equation (7-1)

B

A

AB

ce

ce

ABce m

mTLR

C

CRS

A

B

, (7-1)

where RSe-c is the relative severity of erosion-corrosion of a flowing system, Ce-cA and

Ce-cB are the corrosion components of the erosion-corrosion rates of a flowing system, at

condition A and condition B respectively, TLRB/A is the ratio of thickness loss of B with

respect to the thickness loss of A, and mA and mB represent the slopes obtained from

scratch tests at condition A and condition B respectively. This suggests that the ratio of

the slopes may be a useful indicator of the relative severity of the erosion-corrosion of

two different alloys exposed to erosive-corrosive flow conditions, provided that the

erosivities of the two systems being compared are similar. Notice condition A accounts

for a set of variables such a temperatureA, pHA, in the case of scratch test and

temperatureA, pHA, flow-conditionsA and erosivityA in the case of flow loop testing. The

same variables apply to condition B.

116

Scratch Test Data vs. Single Phase Liquid Flow Data

Data gathered from the erosion-corrosion tests for single phase liquid flow

conditions include measurements of the total corrosion component based on Linear

Polarization Resistance technique (LPR). The total erosion-corrosion rates (weight-loss

method) and the corrosion rate components (LPR method) of the erosion-corrosion

process at different temperatures are listed in Table 7-1.

In order to compare these results with those obtained from the scratch tests, the

relationship between the relative severity of erosion-corrosion, RSe-c (obtained from flow

loop testing) and the thickness loss ratio, TLRB/A (obtained from the Scratch Test)

defined in equation (7-1) is used.

If room temperature is used as the comparing baseline, say condition A, equation

(7-1) can be used to convert the data in Table 7-1 into RSe-c and TLRB/A for different

temperatures at a fixed pH of 4. The results are summarized in Table 7-2.

Table 7-1. Erosion-corrosion penetration rates and corrosion components (Ce-c) of

the erosion-corrosion penetration rate for 13Cr at differenttemperatures. (Vl = 15 ft/s (4.6 m/s), PCO2=50 psig (344.7 kPa), brine 3%,

about 3,500 lb sand /day (1,587.6 kg sand /day)).

Temperature, °FEC rate

weight-loss , mpy

Ce-c rate

LPR , mpy (mm/y)

76 4.5 3.1

150 15.3 12.0

200 18.5 15.3

117

Table 7-2. Comparison of relative severity of the corrosion component of the

erosion-corrosion between single phase liquid loop tests and scratchtests for 13Cr.

Temperature, °F RSe-c,B/A (loop test) TLRB/A (scratch test)

76 1 1

150 3.9 2.5

200 4.9 5.0

Table 7-2 demonstrates that the scratch tests and the loop tests have similar RSe-c

trends, suggesting some similarity between scratch test and sand erosion-corrosion

processes. According to the relative severity parameter from both loop and scratch tests,

the severity of the erosion-corrosion process increases with temperature. Results in Table

7-2 do not perfectly match at 150oF, but they do match at 200oF based on 76oF as a

reference temperature. This suggests that the scratch test may be useful as a

complementary tool in the study of the erosion-corrosion process of CRAs. However,

results shown here in the light of the scratch tests technique, would not account for

differences (if any) in erosion resistance of the oxide layer at different environmental

conditions.

Prediction of Erosion-Corrosion of CRAs using the Scratch Test

Single phase liquid flow

Having more confidence in scratch test results allows one to predict erosion-

corrosion behavior of CRAs exposed to environmental conditions similar to those

experienced in the scratch tests. For a given temperature and pH condition, parameters A

118

and B from equation (7-1) can be redefined as material A and material B, if a comparison

between two different alloys is required. Also, the dimensionless parameter RSe-c (relative

severity of erosion-corrosion) as presented in equation (7-1) can be used to predict flow

loop erosion-corrosion rates from scratch testing obtained at similar environmental

conditions. Data shown as a corrosion component of erosion-corrosion for 13Cr in Table

7-1, and average values of the slopes obtained from scratch testing at pH = 4 and the

three temperatures for the three CRAs were used to perform this prediction. Results are

shown in Figure 7-1. As expected, the best performance is for 22Cr followed by Super

13Cr, while 13Cr showed the highest corrosion rates for a given temperature.

13Cr ActualLoop Data

S13CrPredicted

22CrPredicted

76

150200

0

2

4

6

8

10

12

14

16

Cor

rosi

onC

ompo

nent

(mpy

)

Corrosion Resistant Alloy

Temperature(oF)

76

150200

Figure 7-1. Corrosion component of the erosion-corrosion process for 13Cr

(actual flow liquid loop data), Super 13Cr and 22Cr (data

extrapolated with scratch data) at pH 4 and 3 different temperatures.

Notice that information obtained from the scratch tests does not account for

effects of some variables such as flow conditions and particle impingement (impact

velocity, impact angle). However, these effects are being factored into the prediction

procedure by including the flow loop erosion-corrosion rate data point used in the

119

equation (7-1). The procedure requires that the flow and erosivity conditions to which

EC-predicted alloys (Super 13Cr and 22Cr in this example case) are assumed to be

exposed, are the same conditions to which the reference alloy (13Cr in this example case)

was exposed in the flow loop test. Furthermore, the procedure assumes, that differences

in the erosion responses of the different alloys are negligible as compared to the effects

that environmental conditions such as pH and temperature have in the total erosion-

corrosion process. In other words, differences in the erosion damage in the different

alloys here tested are assumed to be small provided the alloys are all exposed to similar

flow and erosivity conditions (sand concentration, sand size and shape).

Validation of Scratch Test Predictions of Erosion-Corrosion of CRAs

Multiphase flow

In general, predictions of the corrosion component of erosion-corrosion based on

scratch test data compared well to test results from multiphase gas/liquid flow loop for

the three CRAs at high erosivity conditions. Hence, second order behavior appears to be

an appropriate and useful model to represent the repassivation process of CRAs. A

procedure to predict penetration rates for erosion-corrosion conditions was developed

based on this second order model behavior observed by the re-healing process of the

passive film of CRAs under tested conditions. The procedure, as developed thus far,

requires that one flow loop test be run at known environmental conditions and specified

flow and erosivity conditions Once the erosion-corrosion rate for this baseline set of

conditions is established, then, provided erosivity conditions remain unchanged, erosion-

corrosion rates can be predicted for other environments and other CRAs for which

120

scratch test data is available. Good agreement between the actual and predicted

penetration rates was found.

Table 7-3 shows actual and predicted values of the corrosion component (Ce-c) of

the erosion-corrosion penetration rates for 3 alloys tested at similar conditions (76oF and

pH = 4). Actual values are the same as displayed in Table 6-1 of Chapter 6. Predicted

values were obtained by mean of equation (7-1) and the slope ratio of the scratch test data

using 13Cr loop data as a reference (A) material. As can be seen, actual and predicted Ce-c

values match very well, suggesting scratch test data may be useful for predicting erosion-

corrosion penetration rates provided that data from at least one loop test is available.

An analysis similar to that done for the penetration rates of the three CRAs at

room temperature can be done for 150oF with flow loop test data shown in Table 6-2 of

Chapter 6 and proper used of the slopes obtained by means of scratch tests. Results of

this comparison are shown in Table 7-4. In general, trends found at 150oF are similar to

those observed at 76oF. Notice that the scratch test prediction data seem to closely match

the trends and observations obtained from loop test data.

In general, predictions of the corrosion component of erosion-corrosion based on

scratch test data compared well to test results from multiphase gas/liquid flow loop for

the three CRAs at high erosivity conditions. Hence, second order behavior appears to be

an appropriate and useful model to represent the repassivation process of CRAs. A

procedure to predict penetration rates for erosion-corrosion conditions was developed

based on this second order model behavior observed by the re-healing process of the

passive film of CRAs under tested conditions. The procedure, as developed thus far,

requires that one flow loop test be run at known environmental conditions and specified

121

flow and erosivity conditions Once the erosion-corrosion rate for this baseline set of

conditions is established, then, provided erosivity conditions remain unchanged, erosion-

corrosion rates can be predicted for other environments and other CRAs for which

scratch test data is available. Good agreement between the actual and predicted

penetration rates was found.

Table 7-3. Actual and predicted values of the corrosion component (Ce-c) of the

erosion-corrosion penetration rates for 3 alloys tested at similar

conditions (76oF, pH 4)

Alloy

Actual Ce-crate

*Ce-c =EC- E

(mpy)

Scratch Test Data

TLRB/A =(m13Cr/mCRA)

Predicted Ce-c rate

Ce-c,CRA= TLRB/A CEC, 13Cr

(mpy)

13Cr 99 1 (reference)

Super13Cr 47 0.40 40

22Cr 29 0.22 22

*Obtained from subtraction the erosion from the erosion-corrosion

Table 7-4. Actual and predicted values of the corrosion component (CEC) of theerosion-corrosion penetration rates for 3 alloys tested at similar

conditions (150oF, pH 4).

Alloy

Actual CEC rate

*Ce-c =EC- E

(mpy)

Scratch Test Data

TLRB/A =(m13Cr/mCRA)

Predicted Ce-c rate

Ce-c,CRA= TLRB/A CEC, 13Cr

(mpy)

13Cr 207 1 (reference)

Super13Cr 103 0.34 71

22Cr 25 0.12 25

* Obtained from subtraction the erosion from the erosion-corrosion

122

CHAPTER 6 ................................................................................................................................100

MULTIPHASE GAS/LIQUID/SAND FLOW LOOP TESTING RESULTS................................................100

Low Sand Rates (Multiphase Flow Loop Testing)................................ ...............................101

High Sand Rates (Multiphase Flow Loop Testing)..............................................................108

CHAPTER 7 ................................................................................................................................114

COMPARISON OF SCRATCH TEST RESULTS WITH FLOW LOOP TEST RESULTS ..............................114

Scratch test data vs. single phase liquid flow data ..............................................................116

Prediction of Erosion-Corrosion of CRAs using the Scratch Test.......................................117

Single phase liquid flow.......................................................................................................................117

Validation of Scratch Test Predictions of Erosion-Corrosion of CRAs...............................119

Multiphase flow....................................................................................................................................119

123

TABLE 6-1. EROSION-CORROSION (EC), PURE EROSION (E) AND CORROSION COMPONENTS (CE-C) OF THE

EROSION-CORROSION PENETRATION RATES FOR 3 ALLOYS TESTED AT SIMILAR CONDITIONS. 76OF, PH 4,

VSG = 20 FT/S, VSL = 1.4 FT/S, 30 LB/DAY OF SAND...............................................................................111

TABLE 6-2. EROSION-CORROSION (EC), PURE EROSION (PUREE) AND CORROSION COMPONENTS (CEC) OF THE

EROSION-CORROSION PENETRATION RATES FOR 3 ALLOYS TESTED AT SIMILAR CONDITIONS. 150OF, PH

4, VSG = 20 FT/S, VSL = 1.4 FT/S, 30 LB/DAY OF SAND . ................................................................ .........112

TABLE 7-1. EROSION-CORROSION PENETRATION RATES AND CORROSION COMPONENTS (CE-C) OF THE

EROSION-CORROSION PENETRATION RATE FOR 13CR AT DIFFERENT TEMPERATURES. (VL = 15 FT/S (4.6

M/S), PCO2=50 PSIG (344.7 KPA), BRINE 3%, ABOUT 3,500 LB SAND /DAY (1,587.6 KG SAND /DAY))..116

TABLE 7-2. COMPARISON OF RELATIVE SEVERITY OF THE CO RROSION COMPONENT OF THE EROSION-

CORROSION BETWEEN SINGLE PHASE LIQUID LOOP TESTS AND SCRATCH TESTS FOR 13CR. ................117

TABLE 7-3. ACTUAL AND PREDICTED VALUES OF THE CORROSION COMPONENT (CE-C) OF THE EROSION-

CORROSION PENETRATION RATES FOR 3 ALLOYS TESTED AT SIMILAR CONDITIONS (76OF, PH 4) ........120

TABLE 7-4. ACTUAL AND PREDICTED VALUES OF THE CORROSION COMPONENT (CEC) OF THE EROSION-

CORROSION PENETRATION RATES FOR 3 ALLOYS TESTED AT SIMILAR CONDITIONS (150OF, PH 4). .....121

FIGURE 6-1. PURE EROSION TEST FOR 13CR UNDER SAND-N2-DISTILLED WATER FLOW SYSTEM (VSG=60

FT/S, VSL= 0.2 FT/S) AT 150OF AND LOW SAND RATE (15 LB/DAY). ................................ ......................103

FIGURE 6-2. EROSION-CORROSION TEST FOR 13CR UNDER SAND-CO2-BRINE FLOW SYSTEM (VSG=60 FT/S,

VSL= 0.2 FT/S) AT 150OF AND LOW SAND RATE (15 LB/DAY). ................................ .............................104

FIGURE 6-3. PENETRATION RATE FOR PURE EROSION AND EROSION-CORROSION TESTS OF 13CR IN

MULTIPHASE FLOW TESTING AT 76OF. (THE ERRORS BARS ON THE AVERAGE REPRESENT THE 95%

CONFIDENCE INTERVAL ON THE MEAN VALUE.) ................................................................ ..................105

FIGURE 6-4. PENETRATION RATE FOR PURE EROSION AND EROSION-CORROSION TESTS OF 13CR IN

MULTIPHASE FLOW TESTING AT 150OF. (THE ERRORS BARS ON THE AVERAGE REPRESENT THE 95%

CONFIDENCE INTERVAL ON THE MEAN VALUE.) ................................................................ ..................107

FIGURE 6-5. EROSION-CORROSION (EC), PURE EROSION (PURE E) AND CORROSION COMPONENT (CEC) OF

THE EROSION-CORROSION PENETRATION RATES FOR DATA SHOWN IN TABLE 6-1...............................111

FIGURE 6-6. EROSION-CORROSION (EC), PURE EROSION (PUREE) AND CORROSION COMPONENT (CEC)

OF THE EROSION-CORROSION PENETRATION RATES FOR DATA SHOWN IN TABLE 6-2..........................112

FIGURE 7-1. CORROSION COMPONENT OF THE EROSION-CORROSION PROCESS FOR 13CR (ACTUAL FLOW

LIQUID LOOP DATA), SUPER13CR AND 22CR (DATA EXTRAPOLATED WITH SCRATCH DATA) AT PH 4

AND 3 DIFFERENT TEMPERATURES. .....................................................................................................118

124

59 Rincon, H.E. (2001). "Erosion-Corrosion Phenomena of 13Cr Alloy in Flows

Containing Sand Particles". M.S. Thesis, Department of Mechanical Engineering, TheUniversity of Tulsa, Ok, Tulsa..

57 Rincon, H.E., Chen, J. and Shadley, J.R. (2002). "Erosion-Corrosion Phenomenaof 13Cr Alloy in Flows Containing Sand Particles". Corrosion/2002, paper no.2493, (Houston, TX, USA, NACE International.

58 Chen, J., Shadley, J.R., Rincon, H.E. and Rybicki, E.F. (2003). "Effects ofTemperature on Erosion-Corrosion of 13Cr". Corrosion/2003, paper no. 3320,(Houston, TX, USA, NACE International.

122

CHAPTER 8

SUBMERGED DIRECT IMPINGEMENT TEST: SINGLE PHASE LIQUID

FLOW

Erosion-Corrosion Liquid/Sand Loop (Microloop)

The study of the effects of some environmental parameters (temperature and pH)

on the repassivation behavior of CRAs in CO2-saturated brine was extensively discussed

in Chapter 5. This study was performed based on results obtained with the “Scratch

Test”60, a laboratory method conducted in a glass cell. Some similarities between the

scratch test and the sand erosion process observed in pipe-flow conditions have been

previously mentioned as well. The convenience of the experimental setup used for the

scratch test to separate the mechanical process (mechanical removal of material) from the

electrochemical process (corrosion) during repassivation of CRAs has been mentioned.

However, in the sand erosion-corrosion process, the mechanical and electrochemical

processes involved take place simultaneously and continuously during impingement of

solid particles on the target material. Thus, prediction of erosion-corrosion needs to

account for parameters driven by flow field conditions such as flow velocity, flow

geometry and sand rate. For this purpose, a single phase liquid flow loop was used to

123

experimentally determine the effect of flow velocity and sand rate in the erosion-

corrosion mechanism of CRAs.

Data were collected based on an electrochemical technique. The instrumentation

and the configuration of the electrodes used for scratch test were successfully adapted to

the dynamic flow loop testing as described in Chapter 4. Thus, the working electrode

(target specimen) is continuously impinged by sand particles while the potential and

current are recorded.

The main objective of the flow loop testing was to check the effect of certain

parameters such a flow velocity and sand rate on the repassivation behavior of CRAs.

Data collected in the flow loop testing also provided a reference point and target to the

predictions of erosion rates (E), corrosion component rates (Ce-c), and total erosion-

corrosion rates (E-C).

However, obtaining quality data with the low erosivities commonly obtained in

liquid flows testing is challenging and time consuming, if the weight loss technique is

used. Thus, several electrochemical techniques along with a few weight loss tests were

performed

The set of electrochemical tests conducted are based upon the response of the

current flowing between two electrodes made of the same CRA but exposed to different

flow conditions. The working electrode #1 is the target specimen exposed to the sand

laden impinging jet. A larger auxiliary working electrode #2, is placed upstream of the

nozzle and thus is not exposed to the flowing sand at any time during the test. Other

standard electrochemical techniques, such as LPR and potentio-dynamic scans, were also

performed at selected conditions.

124

The flow loop tests for 13Cr and Super 13Cr were carried out at 10, 15, 17 and 20 ft/s, a

fixed temperature of 150°F, a pH of 4.3 and 6, and at several sand concentrations. To

perform a flow loop test, temperature and pH are adjusted while all the fluid flow is

bypassed to the main tank. Once these conditions are adjusted and stabilized, the flow is

diverted to the test cell where the target specimen, auxiliary working electrode, and the

reference electrode are housed. At this time, the current noise flowing between the target

specimen and auxiliary working electrode is measured and recorded for further analysis.

The system is left to run overnight to stabilize the current to very low values similar to

those observed for the static scratch test. Once the current stabilizes at low values, the

required amount of sand is added to the flow. At this point, the current increases with

time at a rate determined by the combined effect of flow velocity, sand rate, and

environmental conditions. Upon stabilization of the new, higher current with direct

impingement of particles on the target, the sand slurry flow is diverted from the cyclone

separator circuit loop to the filter system. Thus, the specimen is then exposed to a clean

flow at the same environmental and flow conditions, but without sand.

Figure 8-1 shows the typical behavior of the current vs. time response for a CRA

exposed to a CO2-saturated brine flow containing sand. This particular test was

conducted for Super 13Cr at 150oF and pH 4.3 with a sand rate of 75 kg/day and flow

velocity of 20 ft/s. As can be seen in the figure, the current rapidly increased after sand

was added, and then stabilized between 5 and 7 microamperes. After engaging the filter,

and the sand from the loop is removed, the current rapidly returned to initial values,

suggesting a passive film was recreated on the surface of the target specimen

125

0

2

4

6

8

10

0 5 10 15 20 25 30 35 40 45 50 55Time (hours)

I( m

A)

Super 13Cr, 150oF, pH 4.320 ft/s, Sand Rate 75 kg/d.

Sand added

Filter engagedSand removed

Flow

ImpingementDirect

Figure 8-1. Current response for Super 13Cr exposed to CO2 saturated brinecontaining sand.

These results are very much in line with those observed for the scratch test cell

results described in Chapter 5. As expected, the extent to which the current increases

depends on experimental parameters of sand rate and flow velocity.

Figure 8-2 shows the same current vs. time data shown in Figure 8-1, including a

new set of data representing the current response of super13Cr exposed at similar

conditions but at a higher sand rate. Figure 8-2 clearly shows how larger amounts of sand

(per unit time) cause the current to increase to higher levels. The current magnitudes are

also affected by the flow velocity.

Figure 8-3 shows the same current vs. time data shown in Figure 8-1, including a

new set of data representing the current response of super 13Cr exposed to similar

conditions, but at a lower velocity of 15 ft/s and with a higher san rate of 235 kg/d. Note,

126

that the current data series for 15 ft/s (235 kg/day) stabilized at lower values than the data

for 20ft/s (75 kg/day) regardless of the higher sand rate.

These results indicate that flow velocity and sand rate play important roles in the

equilibrium between the mechanical removal of material caused by the particle

impingement and the electrochemical re-healing process of the protective film, i.e.,

repassivation. Hence, the current magnitudes, or corrosion component of the erosion-

corrosion process, should be dependent not only on environmental factors (pH and

temperature) as shown in previous chapters, but also on flow-mechanical parameters such

as sand rate and flow velocity.

0

2

4

6

8

10

12

14

16

18

0 5 10 15 20 25 30 35 40 45 50 55Time (hours)

I( m

A)

Super 13Cr, 20 ft/s, 150oF, pH 4.3

Sand added

Filter engagedSand removed

Sand Rate75 kg/day

Sand Rate200 kg/day

Flow

ImpingementDirect

Figure 8-2. Comparison between current responses for Super 13Cr exposed toCO2-saturated brine at similar environmental and flow conditions butdifferent sand rates.

127

0

2

4

6

8

10

0 5 10 15 20 25 30Time (hours)

I( m

A)

Sandadded

Filter engagedSand removed

15 ft/s235 kg/day

20 ft/s75 kg/day

Super 13Cr, 150oF, pH 4.3Flow

ImpingementDirect

Figure 8-3. Comparison between current responses for Super 13Cr exposed toCO2-saturated brine at similar environmental conditions but differentflow velocities and sand rates.

Figure 8-1 to Figure 8-3 have already shown how the current decreases with time

immediately after the sand is removed, suggesting that repassivation taking place on the

target specimen can be accomplished once the mechanical removal of the passive film

has stopped. Once sand impingement is stopped, the high anodic current immediately

decreases, approximately following a second order model similar to that observed for the

static scratch test. This is the reason why this type of testing has been referred to as

Dynamic Scratch Test in this research.

Figure 8-4 shows a typical current decay with time during the healing process of

Super 13Cr alloy exposed to a 3.5% brine solution at pH 4.3, a temperature of 150oF, a

flow velocity of 20 ft/s and after sand (400 kg/day) has been removed.

128

0

5

10

15

20

25

30

0 0.5 1 1.5 2 2.5Time (hours)

I( m

A)

Super 13Cr

Temp: 150oF, pH 4.3Sand rate: 400 kg/dayVelocity: 20ft/s

Flow

ImpingementDirect

Figure 8-4. Anodic current decay for Super 13Cr alloy after sand is removedfrom the test cell loop.

How fast the current decays with time is determined by the healing rate of the

protective layer, and may depend on many variables such as temperature, pH, flow

velocity and material tested. Notice that the healing time frame is different from that

shown by the static scratch test for the same material at the same environmental

conditions (T and pH) showed in Chapter 5. For the scratch test of Super 13Cr at pH 4

and 150oF, after only 200 sec, the current has decreased to a value about 2% of the initial

current. On the other hand, for the dynamic flow loop test conducted at similar

environmental conditions and sand, the current has decreased to about 5% of the initial

current after 2 hours that sand was removed.

129

In any case, the current response for the flow loop test is also following a second

order model as shown in Figure 8-5. However, the parameters Io and m obtained for flow

loop tests typically show a significant difference to those observed for the static scratch

test at similar environmental conditions. Some possible explanations for this observation

are presented in Chapter 9.

1/ I = 229t + 26928

0.0E+00

5.0E+04

1.0E+05

1.5E+05

2.0E+05

2.5E+05

3.0E+05

0 200 400 600 800 1000 1200

Time (sec)

1/I

(1/A

) 1/ I = mt + 1/ I o

Super 13CrTemp: 150oFSand Added: 400 kg/dayVelocity: 20ft/s

Flow

ImpingementDirect

Figure 8-5. A linear behavior between 1/I and time after sand is removed fromthe test cell loop.

Figure 8-6 shows the anodic current decay data previously shown in Figure 8-4

along with the current vs. time obtained by using the second order model approximation.

This figure suggests that the second order model is in fact a very good approximation to

describe the repassivation process of CRAs under dynamic flow conditions tested for a

significant long period of time.

130

0

5

10

15

20

25

30

0 0.5 1 1.5 2 2.5

Time (hours)

I(mA

)

Actual Data2nd Order Model

Io = 28 mAm = 229 (1/A*s)

Super 13CrTemp: 150oFSand rate: 400 kg/dayVelocity: 20ft/s

Flow

ImpingementDirect

Figure 8-6. Actual data and second order model for the anodic current decayafter sand is removed from of the test cell loop.

Figure 8-7 shows the effect of sand rate on the anodic current decay of Super

13Cr at fixed environmental and flow conditions. This figure corroborates the trend

observed and discussed about Figure 8-2, i.e., for given environmental and flow

conditions, the higher the sand rate the higher the current (Io). Figure 8-7 also shows

similar behavior in the current vs. time response regardless of the sand rate. Figure 8-8,

shows how the sand rate only very slightly affects the value of the slope of the second

order model for the same data series seen in Figure 8-7.

131

0

5

10

15

20

25

30

0 1 2 3 4 5

Time (hours)

I( m

A)

400 Kg/day

200 Kg/day

75 Kg/day

Super 13CrT = 150oF, pH 4.3

Flow Velocity: 20 ft/s

Flow

ImpingementDirect

Figure 8-7. Effect of sand rate on the anodic current decay of Super 13Cr alloy.

1/ I = 204 t + 229,65675 Kg/day

1/ I = 198 t + 94,533200 Kg/day

1/ I = 229 t + 26,928400 Kg/day

0.E+00

1.E+05

2.E+05

3.E+05

4.E+05

5.E+05

0 200 400 600 800 1000

Time (sec)

1/I

(1/A

)

1/ I = mt + 1/ I o

Super 13Cr

T = 150oF, pH 4.3Flow Velocity: 20 ft/s

Figure 8-8. Effect of sand rate on the linear behavior for 1/I with time after sandis removed from the test cell loop.

132

Figure 8-9 shows the anodic current decay for 13Cr alloy after sand is removed

from the test cell loop. Initial currents are generally higher than those observed for Super

13Cr alloy at similar environmental and flow conditions. Also, the repassivation time

tends to be longer for 13Cr than for Super 13Cr at given conditions. These particular tests

were conducted at 150oF, sand rate of 60kg/day, at two different pH values of 4.3 and 6,

and two different flow velocities 20ft/s and 17ft/s, respectively. The current for both

conditions approximate the second order model.

0

5

10

15

20

25

30

35

40

0 0.5 1 1.5 2 2.5

Time (hours)

I( m

A)

13Cr

Temp: 150oF, pH 4.3 & 6Sand rate: 60 kg/day

20 ft/s, pH 4.3

17 ft/s, pH 6

Flow

ImpingementDirect

Figure 8-9. Effect of pH on the anodic current decay of 13Cr alloy.

The current magnitude for the test conducted at pH 4.3 is slightly higher than that

for the test conducted at pH 6, at least for the first hour. This difference in current

magnitude may correspond to the combined effect of the higher velocity and lower pH

and in any case it seems to be not too large. However, the time frames during which sand

133

flowed in the test cell loops for the two tests were significantly different, about 3 hours

for the test at pH 6, and just 5 minutes for the test at pH 4.3. If longer periods of time

were used to test 13Cr alloy at severe conditions such a low pH, high flow velocities and

high sand rates, high current magnitudes may be sustained, even after removing the sand,

thus not following the second order repassivation model.

When high currents are maintained for long periods of time in 13Cr, a black

coating grows on the surface of the 13Cr target specimen as seen in Figure 8-10a and

Figure 8-10b shows the metallic surface of 13Cr also exposed to severe environments

(low pH, high flow velocity and high sand rate) but just for a short period of time (about

5 minutes). Hence, the image shown in Figure 8-10b may be assumed to be a typical

initial stage of the black coating. The concentric ring pattern shown on the surface of the

13Cr specimen suggests that the erosion-corrosion damage is not uniform, being worse in

the dark zones. This pattern, matches material degradation patterns obtained with

computational fluid dynamics simulations performed for erosion and erosion-corrosion,

which are discussed in next chapter. If the combined mechanical and environmental

conditions are not that severe, for example, similar high flow velocities and sand rates but

higher pH, the black coating formation is prevented, the second order repassivation

behavior is followed and a typical CRA surface finish is obtained as shown in Figure 8-

10c. This observation also agrees with the common lower current values, corrosion rates

and erosion-corrosion rates obtained for 13Cr tested at pH 6 as compared to those

obtained for 13Cr tested at lower pH values. However, even at high pH values, the

resultant surface for 13Cr is still dull as compared to the bright mirror surface obtained

for super 13Cr at any pH. (see Figure 8-10d).

134

Even though the study of the morphology of the black coating is beyond of the

scope of this dissertation, an effort to better characterize this phenomenon was performed

by scanning electronic microscope (SEM), energy dispersive spectroscopy (EDS), X-ray

diffraction (XRD), and Raman spectroscopy. Also, a few electrochemical tests

(potentiodynamic scans) directed towards the understanding of differences between the

erosion-corrosion behavior of 13Cr and Super 13Cr were performed and are extensively

documented in Appendix C of this dissertation.

a) 13Cr, pH4.3long sand exposure

b) 13Cr, pH 4.3short sand exposure

c) 13Cr, pH 6long sand exposure

d) Super 13Cr, pH4.3long sand exposure

Figure 8-10. Comparison of the resultant metallic surfaces of the target specimenof 13Cr and Super 13Cr exposed to different erosion-corrosionconditions.

Table 8-1 and Table 8-2 summarize the experimental values of current, Io,

corrosion component rates of erosion-corrosion, Ce-c, erosion-corrosion rates, E-C and

135

pure erosion rates E, obtained for 13Cr and Super 13Cr at 150oF and pH 4.3 unless

otherwise specified. These values are compared to predicted values in Chapter 9.

Table 8-1. Summary of corrosion component rates of erosion-corrosion Ce-c;erosion-corrosion rates, E-C; and pure Erosion rates E, obtainedexperimentally. 150F and 4.3 pH.

MaterialVelocity

(ft/s)

SandRate

(kg/day)

Ce-c

(mpy)LPR

E-C(mpy)WL

E(mpy)

WL13Cr* 17 60 2.413Cr* 17 132 6.0 6.413Cr* 17 132 2.0 6.413Cr 17 132 1.6

S13Cr 10 100 7.0S13Cr 17 132 5.8 3.5S13Cr 20 200 9.0S13Cr 20 200 5.0

*Tests performed at 150oF and pH 6

Table 8-2. Summary of experimental values obtained for the current Io. 150Fand 4.3 pH.

MaterialVelocity

(ft/s)

SandRate

(kg/day) Io (A)13Cr* 17 60 14.213Cr* 17 132 18.013Cr 20 60 38.013Cr 20 75 20.013Cr 20 108 100.0

S13Cr 10 25 2.1S13Cr 10 100 5.0S13Cr 15 55 4.0S13Cr 15 165 3.3S13Cr 15 165 4.5S13Cr 15 235 10.0S13Cr 20 28 3.2S13Cr 20 75 18.0S13Cr 20 75 6.6S13Cr 20 200 20S13Cr 20 200 18.0S13Cr 20 200 13.0S13Cr 20 200 20.0S13Cr 20 400 28.0

*Tests performed at 150oF and pH 6

136

CHAPTER 8 ..............................................................................................................................................122

SUBMERGED DIRECT IMPINGEMENT TEST: SINGLE PHASE LIQUID FLOW ................................ ..................122Erosion-Corrosion Liquid/Sand Loop (Microloop) ................................................................ ...........122

TABLE 8-1. SUMMARY OF CORROSION COMPONENT RATES OF EROSION-CORROSION CE-C; EROSION-CORROSION RATES, E-C; AND PURE EROSION RATES E, OBTAINED EXPERIMENTALLY. 150F AND 4.3PH. 135

TABLE 8-2. SUMMARY OF EXPERIMENTAL VALUES OBTAINED FOR THE CURRENT IO. 150F AND 4.3 PH.135

FIGURE 8-1. CURRENT RESPONSE FOR SUPER 13CR EXPOSED TO CO2 SATURATED BRINE CONTAININGSAND. 125

FIGURE 8-2. COMPARISON BETWEEN CURRENT RESPONSES FOR SUPER 13CR EXPOSED TO CO2-SATURATEDBRINE AT SIMILAR ENVIRONMENTAL AND FLOW CONDITIONS BUT DIFFERENT SAND RATES . ..............126

FIGURE 8-3. COMPARISON BETWEEN CURRENT RESPONSES FOR SUPER 13CR EXPOSED TO CO2-SATURATED

BRINE AT SIMILAR ENVIRONMENTAL CONDITIONS BUT DIFFERENT FLOW VELOCITIES AND SAND RATES.127

FIGURE 8-4. ANODIC CURRENT DECAY FOR SUPER 13CR ALLOY AFTER SAND IS REMOVED FROM THE TESTCELL LOOP. ................................................................................................................................ .........128

FIGURE 8-5. A LINEAR BEHAVIOR BETWEEN 1/I AND TIME AFTER SAND IS REMOVED FROM THE TEST CELL

LOOP. 129FIGURE 8-6. ACTUAL DATA AND 2ND ORDER MODEL FOR THE ANODIC CURRENT DECAY AFTER SAND IS

REMOVED FROM OF THE TEST CELL LOOP................................................................. ...........................130FIGURE 8-7. EFFECT OF SAND RATE ON THE ANODIC CURRENT DECAY OF SUPER13CR ALLOY................131FIGURE 8-8. EFFECT OF SAND RATE ON THE LINEAR BEHAVIOR FOR 1/I WITH TIME AFTER SAND IS REMOVED

FROM THE TEST CELL LOOP. ................................................................................................................131FIGURE 8-9. EFFECT OF PH ON THE ANODIC CURRENT DECAY OF 13CR ALLOY........................................132FIGURE 8-10. COMPARISON OF THE RESULTANT METALLIC SURFACES OF THE TARGET SPECIMEN OF 13CR

AND SUPER 13CR EXPOSED TO DIFFERENT EROSION-CORROSION CONDITIONS. ................................ ..134

136

CHAPTER 9

EROSION-CORROSION MODEL

The complexity of the chemi-mechanical mechanisms involved in erosion-

corrosion of CRAs under slurry flows has been mentioned on in previous chapters. The

many variables involved in the erosion-corrosion process such a temperature, pH, flow

velocity, sand rate, and material type have been described and discussed. How these

parameters affect the erosion-corrosion mechanism was carefully studied by means of

different experimental techniques during this research period. Some of these parameters

may need further study and many other parameters such water chemistry, chloride

contents, H2S, and material properties may be included in future research work. However,

there is an immediate need to define safe service limits for utilization of such materials in

a great diversity of corrosive oil and gas environments which contain sand particles. As

an initial effort to address this need a framework for an erosion-corrosion prediction

model is presented in this Chapter.

General Approach

The concept of a synergistic effect between chemical and mechanical processes

found in erosion-corrosion conditions was introduced in previous chapters. Even though

no uniform nomenclature has been used to define the terms contributing to the total

137

erosion-corrosion weight loss, many authors agree on the need to consider four terms.64-68

The first two terms usually account for the chemical degradation and the mechanical

degradation as separate processes, and here they are named as pure corrosion, C and pure

erosion, E. The two remaining terms describe the synergism involved in the erosion-

corrosion process; one accounts for the corrosion increment due to erosion, Ce and the

other accounts for the erosion increment due to corrosion, Ec. Hence, the total erosion-

corrosion rate EC is given by equation 9-1.

ce ECECEC (9-1)

Determining whether there is a dominant term is not straight forward, and often

depends on the combined effect of environmental and mechanical variables as well as

material properties. However, existing erosion-corrosion data have provided enough

understanding of the process to make reasonable assumptions in regards to simplifying

the procedures to predict erosion-corrosion under specific conditions.

A mechanistic procedure to estimate erosion-corrosion penetration rates for CRAs

exposed to CO2 saturated brine/sand flows conditions was constructed and is described in

the following paragraphs. The procedure is based on experimental data obtained from

static scratch tests and on computation fluid dynamics (CFD) simulations. Experimental

and numerical data are integrated into the model and processed with a VBA code to

numerically estimate the erosion-corrosion penetration rate. The model assumes some

fixed parameters such a brine concentration, water chemistry, and CO2 pressure. Input

parameters are temperature, pH (which are implicit in the second order scratch test

parameters, m and Io), type of alloy, sand rate, sand size, and liquid velocity.

138

Output of the model includes the pure erosion rate, E, the total erosion-corrosion

penetration rate, EC, and the corrosion rate component of the erosion-corrosion process,

Ce-c, defined as the sum of the pure corrosion C and the impact-induced corrosion Ce

expressed as follow

ece CCC (9-2)

The Ce-c component conveniently groups those terms describing chemical

degradation since they are experimentally measured together when electrochemical

techniques are used to estimate penetration rates for flows containing solid particles.

Hence, direct comparisons between the model and experimental data can be made. For

the purpose of this research, an additional output was considered as the total current, Io,

to be compared with experimental data shown in Chapter 8.

Results of a previous investigation57,59 performed for erosion-corrosion of 13Cr

alloy exposed to CO2 saturated brine/sand flows, suggested that the total erosion-

corrosion penetration rate for this environment and alloy combination equals the sum of

the pure erosion term E, and the corrosion component of the erosion-corrosion term Ce-c.

This means that the corrosion-induced erosion term, Ec, was negligible for these tested

conditions. Based on this finding, the model estimates the total erosion-corrosion

neglecting the term Ec, as follow.

ceCEEC (9-3)

139

The pure erosion term, E, is estimated from the computational fluid dynamics

procedure explained in Appendix D, and Ce-c is estimated with a new procedure

developed in this research.

Proposed Procedure for Estimating Ce-c.

The estimation of the corrosion component of erosion-corrosion, Ce-c, has its

foundation on a model that reasonably describes the competition between the mechanical

removal of the passive film due to solid particle impingement and the re-formation of this

protective film due to the electrochemical repassivation process. The mechanical part of

the model relies on the particle tracking and erosion modeling executed with CFD, while

the electrochemical process relies on the second order model displayed by the

experimental data obtained with the scratch test technique. The physical model to explain

the higher currents observed when sand is introduced into the flow as compared with the

low passive currents typically observed for CRAs exposed to sand-free brine flows, is

shown schematically in Figure 9-1.

140

rR

Base Metal

tT

X

X

IndentationPassive film

Particle

Removed material

rR

Base Metal

tT

X

X

IndentationPassive film

Particle

Removed material

Figure 9-1. Schematic of a particle impingement on a passive alloy.

Figure 9-1 shows an ideal spherical particle of radius R that has impacted a

passive material. The impact resulted in a spherical indentation in the metallic surface

which is shown with a maximum penetration depth x. Also the volume of material

(passive layer and base metal) displaced by the impact is shown in the same figure

forming a platelet shape, as described in an erosion model proposed by Finnie50.

The similarity between the scratch test and the erosion-corrosion process for a

single solid particle impact was discussed in Chapter 5. At given environmental

conditions (pH and temperature), the amount of current or corrosion taking place in the

scratch test can be assumed to be the same as that due to a single solid particle impact

provided that the scratched area and the indented area are identical. In other words, the

increment in current due to removal of the passive layer does not depend on how the

141

passive layer was removed (scratch or particle impact) but on how much exposed (bare,

without passive layer) area was created. Hence, the current density is assumed to be the

same for both processes provided that environmental conditions were the same.

If the current density is known, determination of the amount of the open (bare)

area of the base metal created (due mechanical film removal) by a particle impact is the

only additional quantity needed to estimate the increment in corrosion due to that particle

impact. Then, the remaining key parameters to be considered in the erosion-corrosion rate

determination are the time frame that the open (bare) surface will take to re-form the

passive film after each particle strike, and the particle impact frequency.

At this point the healing rate represented by the slope m of the second order

model figures into the prediction model. The initial current density is given by the scratch

test, and the open area is determined using the solid particle information from the CFD

erosion model. The slope of the second order model describes the rate of decay of the

initial current density for each particle impact. Therefore, if the number of particles that

hit a particular cell is known (computed from the CFD simulation) the total current

generated per cell can be computed. Also, the current for the whole target specimen can

be estimated by adding the currents from each cell, allowing for computation of the total

corrosion component of the erosion-corrosion process.

142

Determination of the indented open area

The determination of the indented surface area created in the base metal is not too

different from the estimation of the volume of material removed due to the particle

impingement. Ultimately, both the indented area and the volume of material removed are

caused by the same impingement. Thus, one can reasonably assume that the variables

affecting the variation of the indented area are similar to those affecting the material

removal and erosion rates.

Furthermore, if the volume of material removed is assumed to be equal to the

volume of the indented space, the volumetric erosion ratio is then equal to the ratio of the

volume of the indented area to the volume of the spherical particle causing the

indentation given by

volumeimpactingsand

materialremoved

sphere

nindentatio ERV

V

VV

_

_ (9-4)

where the volumetric erosion ratio is the mass erosion ratio (kg of removed material/kg of

sand impacting) multiplied by the inverse of the ratio of the density of the target material

and the erodent particle, given by.

material

particlemassvolume ERER

(9-5)

Also a relation between the ratio of the volume of the indentation to the volume of

a sphere and the ratio of the surface area of the indentation to the surface area of a sphere

was found to be closely approximated by equation (9-6). (See details in Appendix E)

143

5026.0

6022.0

sphere

nindentatio

sphere

nindentatio

VV

SS (9-6)

If equations (9-4) and (9-5) are substitute into equation (9-6), the surface of the

indentation caused by a single impact can be computed as a function of the erosion ratio,

the densities of the target material and erodent particle, and the particle size as shown in

equation (9.7)

25026.0

6022.0 pmaterial

particlemassnindentatio DERS

(9-7)

where Dp is the diameter of the particle.

Since the function obtained for determination of the area of the indentation

strongly depends in the erosion ratio, the indented area is expected to vary with impact

velocity and impact angle in a similar fashion as the erosion ratio does. Furthermore, the

same material removal mechanisms considered in the erosion equation by Oka et.al.61,

namely “cutting action” and “repeated plastic deformation” are also being considered in

the estimation of the indented area and therefore in the corrosion component of the

erosion-corrosion process.

Implementation of the second order model to determine the total current

Using equation (9-7) the area of the indentation where the passive layer has been

removed by the particle impact can be estimated. With the current density obtained from

the scratch tests and the known surface area of the indentation, the initial current, Io,

generated for each particle impact can be estimated. Notice, each particular impingement

144

will have its own impact characteristics (impact speed and impact angle) causing a

specific erosion ratio and thus a particular indentation area. Therefore, each impact will

generate a different amount of current.

From the CFD simulation, the impact frequency (# particles per second) for each

cell is also known and can be used to estimate the time between impacts, tbp for that

specific cell as the inverse of the impact frequency. With the estimated current, Io,

produced by the particle impact, the slope, m, from the scratch test and the time between

particles, tbp, the total current generated by all the particle impacts taking place in a

particular cell as a function of time can be estimated. To achieve this, the second order

model is used to compute the current transients caused by each particle impact. The total

current per cell is then estimated by adding the current transients due to each particle

impact which are a time, tbp, apart from each other as seen in Figure 9-2.

The continued addition of the high initial current values as additional sand

particles strike causes a repeating behavior of the total current series. Still, the total

current trend seems to increase with time and tends to stabilized for large times when the

tails of the first impingements are negligible as compared with the value of total current

for the same time. Thus, a range of current, more than a single value is expected when the

stabilized current is reached after some time has past, as depicted by Figure 9-3.

145

0

1

2

3

4

5

6

7

8

9

0.0 0.3 0.5 0.8 1.0 1.3 1.5 1.8 2.0 2.3 2.5 2.8 3.0Time

I( m

A)

1st Imp.2nd Imp.3rd Imp.

4th Imp.5th Imp.6th Imp.Itotal

t bp

Total currentgenerated in acell impactedby 6 particles

t total

Figure 9-2. Generic plot for the total current and individual current transients of

6 different impacts taking place in a numerical cell of the target

specimen.

Figure 9-3 only shows the total current also shown in Figure 9-2 but, now

includes the trends observed for the peaks, and valleys displayed by the cyclic total

current series. If only the values of the peaks and valleys are estimated and averaged a

average trend with a single value for the total Io current can be estimated.

Notice that the size of the range between the peaks and the valleys of the total

current would decrease by decreasing the time between particles, tbp. In fact, the typical

impact frequencies obtained for the simulated conditions considered in this study were

high (hundreds of particles per second), therefore, values of tbp are so small that trend

curves for the peaks and valleys and the average trend lie on top of each other.

146

0

1

2

3

4

5

6

7

8

9

0.0 0.3 0.5 0.8 1.0 1.3 1.5 1.8 2.0 2.3 2.5 2.8 3.0Time

I( m

A)

Itotal

t bp

peaks trend

valleys trend

Avg trendtotal I o

Figure 9-3. Generic plot for the total current generated in a cell due to total

particle impact.

Notice also that for the prediction, the total current will not ever completely

stabilize since the tails of the transients will never reach absolute zero (see Figure 9-2).

Therefore, a stop criterion is needed to determine an appropriate value for ttotal. An

intuitive value for the total time is not easy to determine other than one that is large

enough to yield an acceptably accurate result. For this reason, the stop criterion was

designed in a way that the value will emerge from the same simulation. Hence, the

numerical summation of the transient currents stops when the current added for the last

particle impact causes an increment equal or less of 1/10,000th of the total current

summed up to the previous time step. The criterion seems to work reasonably well and

Figure 9-4 shows the total cell current for a cell with an impact frequency of 74 particles

per second. The continuous curve representing the total current has significantly bent

147

after the first 150 seconds even though continuing to increase with time at a diminishing

rate. This also can be verified by the linear behavior that the cumulative current shows

after the first 150 seconds in the same plot meaning that at that time the current increment

from each new particle impingement is small and nearly constant.

Once the current per each cell is determined, the estimation of the total current for

the specimen is straight forward. Then Faraday’s law is applied to convert the current

units into mass loss rate units so the corrosion component of the erosion corrosion, Ce-c,

can be estimated. Also, the total erosion-corrosion EC is estimated by applying equation

(9-3).

0.000

0.005

0.010

0.015

0.020

0.025

0 50 100 150 200 250 300

Time sec

I( m

A)

0

5

10

15

20

25

30

35

40

45

Cu

mu

lati

ve,I

( mA

)

Instantaneoustotal currentproducedin one cell withImpact Frequency of74part/sec

Cumulativetotal current

same cell

Figure 9-4. Typical behavior of the total current produce in a numerical cell due

to the particle impact. The cell is hit at an impact frequency of 74

particles per second.

148

Figure 9-5 shows the flow chart for the complete prediction procedure of erosion-

corrosion of CRAs in brine flows containing sand. It includes the experimental scratch

test work, the CFD modeling, the mechanistic approach of the chemi-mechanical

degradation process, the erosion loop testing needed to adjust the CFD predicted erosion

rates, the erosion-corrosion loop testing to compare the predicted values for the corrosion

component of the erosion-corrosion Ce-c, and the total erosion-corrosion, EC.

149

Particle Io

Add Individual Particle Current&

Estimate Total Cell Current

Scratch TestOutput,

m, io

AdjustedErosion

IndentationArea

CFD Output,Impact frequency, impact

velocity, impact angle, impactlocation.

Scratch Test,Vary, temp, pH and

Material

Estimate Total Specimen Current&

Faraday law Ce-cEstimate

EC

ErosionWL

Loop Testing

CFD Input,Flow velocity, sand rate

flow geometry

LPR and WLLoop Testing to

compare Ce-c & EC

Figure 9-5. Flow diagram to estimate E, Ce-c, & EC

150

Validation of the Erosion-Corrosion Prediction Model

The procedure to estimate the penetration rates (E, Ce-c and EC) shown in Figure

9-5, was built into a computational code (VBA) to be able to generate prediction results

and compare them to gathered experimental data. A series of prediction cases was run for

a broad range of flow velocity, sand rate, temperature, and pH values. Cases include

predictions for 13Cr, Super 13Cr and 22Cr alloys.

Adjustment of the erosion prediction

Before comparisons between predicted values and experimental results were

done, an adjustment of the erosion model was needed. Appendix D describes the over-

prediction of erosion rates for all cases run following the state-of-the-art in CFD

simulation. A factor F, with a value of 125 units, was used in the denominator of the

erosion equation proposed by Oka and Yoshida62 shown as equation (D-7) in Appendix

D.

The F factor was determined by matching the predicted erosion rate to

experimental flow loop data for pure erosion (obtained by weight loss) shown in Chapter

8. Difficulties obtaining reliable experimental data in erosion by liquid-sand flows were

discussed in Chapter 8. Hence, F=125 should not considered to be a precise number and,

a new adjustment based in more reliable experimental data is recommended for future

purposes. In any case, the present adjustment, allows the prediction of erosion rates to be

within the right order of magnitude.

151

Comparison between experimental data and predictions

Figure 9-6 shows the comparison between the measured erosion and erosion-

corrosion rates obtained by weight loss with those predicted by the proposed model. Only

the pure erosion data point shown in this figure was adjusted. From the two erosion-

corrosion data points included in the plot, it can be seen one of them in perfect agreement

with the prediction while the remaining data point seems to be a little over predicted by

the model. However, considering that no further adjustments or any fitting constants were

used for the estimates of the corrosion component of erosion-corrosion and the total

erosion-corrosion rates, Figure 9-6 suggests reasonably good agreement between

measured and predicted values. And again, it should be noted that all the experimental

data represented in Figure 9-6 was gathered for experimental conditions yielding quite

low weight losses for which significant error is possible.

0

1

2

3

4

5

6

7

0 1 2 3 4 5 6 7

E or EC Experimental

E,o

rE

CM

od

el

ECE

S13Cr, pH=4

13Cr, pH=6

Flow

ImpingementDirect

Figure 9-6. Comparison between measured erosion and erosion-corrosion

penetration rates (by weight loss) and those predicted by the proposed

model.

152

Figure 9-7 shows the comparison between the measured corrosion component

rates Ce-c by linear polarization resistance (LPR) and those predicted by the proposed

model. A reasonable agreement between experimental data and predicted results can be

seen if the experimental data point at 7 mpy is ignored even though some scatter still

exists around the perfect agreement line. However, some scatter of the experimental data

is a fact of life when LPR data is collected. Studies 63 performed to check reproducibility

of the LPR technique have suggested error factors of up to 2 in corrosion rates measured

when using this technique without Tafel slope adjustment. In addition, the LPR values

(and the scale shown for this figure) are small when compared to the scale often

associated with erosion, corrosion and erosion-corrosion data showing much higher mpy

values, generally found at higher flow velocity conditions. Recently a new higher

velocity single phase liquid flow loop capable of handling CO2-saturated brines and sand

has been constructed. Experimental data collected for liquids-sand flows at higher

velocities of 20-40 ft/s would provide more reliable experimental data and perhaps a

better test for the present erosion-corrosion prediction model.

Figure 9-8 shows the comparison between the total current values (A) as

measured with the scratch test instrumentation adapted to the flow system and those

values predicted by the proposed model. Again, some scatter is shown but a good trend is

observed for the model predictions in comparison with experimental data. These

predictions are encouraging considering that no adjustments or fitting constants at all

were used for the corrosion component of the predictive model and the single adjustment

made for the erosion component of the model was based on only one set of experimental

conditions.

153

0

1

2

3

4

5

6

7

8

9

10

0 1 2 3 4 5 6 7 8 9 10

C e-c (mpy) Experimental (LPR)

Ce-

c(m

py)

Mo

del

Flow

ImpingementDirect

Figure 9-7. Comparison between measured corrosion component penetration

rates (by LPR) and those predicted by the proposed model.

0

5

10

15

20

25

30

35

40

45

0 5 10 15 20 25 30 35 40 45I (mA) Experimental

I(

A)

Mod

el

Flow

ImpingementDirect

Figure 9-8. Comparison between measured total current values (A) and those

predicted by the proposed model.

154

Another point to consider when evaluating these comparisons is related to the

number of variables that the model is taking into account to predict erosion-corrosion.

The proposed model considers temperature, pH, flow velocity, sand rate, sand properties,

flow geometry and material type. From this point of view, experimental data and

predicted values may consider to be in reasonably good agreement.

Some trends of predicted values

The following sections of this chapter display the trends of predicted values when

varying some of the variables such as temperature, pH, sand rate, flow velocity and

material type. The experimental data shown in previous plots are inserted into figures

showing trends. In addition, a new set of comparisons between the predictions and other

sets of experimental data discussed in Chapter 2 and Chapter 7 are included.

Effect of sand rate: comparison between experiments and prediction trends

Figure 9-9 shows trends for predicted pure erosion rates with sand rates for

different flow velocities. Trends with sand rates and velocities are as expected, that is,

increasing erosion rate for increasing sand rate and increasing flow velocity. A type of

saturation effect for the erosion rate in direct impingement experiments is usually

observed at high sand rates due to shielding. This phenomenon consists of a blockage

provided by a front of sand particles, which bouncing back from the target, form a shield

in front of it providing some protection from new impingements from trailing particles.

CFD simulations do not take this effect into account so the predicted erosion rates are

linearly increasing with the sand rate regardless of how high is the sand concentration.

Erosion rates seen in Figure 9-9 were obtained for CFD simulations performed for 13Cr,

155

the trends for Super 13Cr are exactly the same, but the values are slightly lower. This

variation is due to the differences in hardness of the two alloys being 25Rc for 13Cr and

30Rc for Super 13Cr. The effect of hardness in erosion has been a controversial topic;

some trends of erosion rates with hardness have been found provided that the

microstructure of the alloys compared are similar. Even so, the effect of material

hardness on erosion rate is only a small one in comparison with the effect of other

parameters such as flow velocity, erodent concentration and erodent properties.

0

1

2

3

4

5

6

7

0 100 200 300 400

Sand Rate (kg/d)

Pen

etra

tio

nR

ate

(mp

y)

E 10ftsE 15ftsE 17fts

E 20fts

TargetSpecimen

Flow

Jet

Figure 9-9. Prediction trends with sand rate of pure erosion, E for several flow

velocities.

Figure 9-10 shows the prediction trends with sand rate of pure erosion, E and

corrosion component of erosion-corrosion, Ce-c for Super 13Cr at two flow velocities.

Predicted trends with sand and velocity are as expected: the erosion rate and the corrosion

156

component of erosion-corrosion increase with both flow velocity and sand rate. Also, the

corrosion component of erosion corrosion rates are higher than the pure erosion rates as

have been seen for experiments with Super 13Cr conducted at relatively low erosivity

conditions presented in Chapter 7. Figure 9-10 also includes 3 experimental data points.

Predictions slightly underestimate the experimental values at both velocities, but

considerable scatter is also shown for experiments conducted at 20ft/s, whose average of

6.5 mpy, accurately matches the predicted value.

0

2

4

6

8

10

12

0 100 200 300 400 500Sand Rate (kg/d)

Pen

etra

tio

nR

ate

(mp

y)

E 20fts

E 17fts

Ce-c 20fts

Ce-c 17fts

Ce-c 20fts (LPR)

Ce-c 17fts (LPR)

Super 13Cr, 150F, pH 4

TargetSpecimen

Flow

Jet

Figure 9-10. Prediction trends with sand rate of pure erosion, E and corrosion

component of erosion-corrosion, Ce-c for Super 13Cr at two flow

velocities.

Figure 9-11 shows a comparison for the same conditions but different material. It

shows the prediction trends with sand rate of pure erosion, E and corrosion component of

157

erosion-corrosion, Ce-c for 13Cr and Super 13Cr at 20ft/s and similar environmental

conditions (pH and 150oF). As expected, and seen in trends for experimental data shown

in Chapter 7, the corrosion component rates of 13Cr are significant higher than those for

Super 13Cr. Also notice how the estimated pure erosion rates for 13Cr and Super 13Cr

are fairly similar.

0

5

10

15

20

25

0 100 200 300 400 500

Sand Rate (kg/d)

Pen

etra

tio

nR

ate

(mp

y)

Ce-c, 13Cr

E, 13Cr

Ce-c, S13Cr

Ce-c, S13Cr (LPR)

E, S13Cr

150oF, pH 4

13Cr, C e-c

Super 13Cr, Ce-c

Erosion, 13Cr & Super 13Cr

TargetSpecimen

Flow

Jet

Figure 9-11. Comparison between 13Cr and Super 13Cr prediction trends with

sand rate of pure erosion, E and corrosion component of erosion-

corrosion, Ce-c.

Figure 9-12 shows a comparison for the same material exposed to different

environmental conditions. It shows the prediction trends of pure erosion, E and corrosion

component of erosion-corrosion, Ce-c for 13Cr at 20ft/s with increasing sand rate and

three different environmental conditions, pH4 at 150oF, pH4 at 200oF and pH6 at 150oF.

158

The corrosion component rates are in general higher than the pure erosion rates; but there

are significant differences among them. The highest corrosion component rates are

predicted for the combination of highest temperature, 200oF and the lowest pH value of 4.

If the pH is fixed at 4 and the temperature is decreased to 150oF, the corrosion component

rates decreases by about 40% (reduction % varies slightly with sand rate). If then, at a

fixed temperature of 150oF, the pH is increased up to 6, the corrosion component rates

further reduce an additional 30%. These trends are in agreement with experimental data

shown in Chapter 7. Figure 9-12 also includes a couple of experimental data point

obtained by using LPR technique for 13Cr exposed to pH 6 and 150oF. For these

conditions, the experimental data and the predicted corrosion component rates are in very

good agreement.

0

5

10

15

20

25

30

35

40

0 100 200 300 400 500

Sand Rate (kg/d)

Pen

etra

tion

Rat

e(m

py)

Ce-c, pH4, T=150 F

Ce-c, pH4, T=200F

Ce-c, pH6, T=150

Ce-c, pH6 T= 150 F, Exp

E

13Cr, 20 ft/s

pH 4, 200 o F, Ce-c

pH 4, 150 o F, Ce-c

pH 6, 150 oF, C e-c

pure Erosion

TargetSpecimen

Flow

Jet

Figure 9-12. Effect of pH and temperature on the prediction trends with sand rate

of the corrosion component of erosion-corrosion, Ce-c for 13Cr.

159

Effect of temperature: comparison between experiments and prediction trends

Unfortunately, most of the data shown in chapter 7 for erosion-corrosion testing

of CRAs were conducted at multiphase flow conditions (gas-liquid-sand) for plugged tee

geometry, while the CFD simulations were conducted for direct impingement in liquid-

sand flows. However the multiphase flow experimental data may be used to check the

trends with temperature and material given by the proposed model. Differences in the

magnitudes of the erosion-corrosion rates for multiphase flow conditions and single

phase liquid conditions are often large. Therefore to be able to compare them, all values

presented will be normalized to the erosion-corrosion value at 76oF.

Figure 9-13 shows the comparison between experimental data and predicted

values for 13Cr exposed to the same environmental conditions (pH and temperature) but

different flow conditions. Erosion-corrosion experimental data was conducted for

multiphase flow conditions (gas-liquid-sand) and plugged tee geometry while predicted

values were obtained for liquid sand flows for direct impingement. It is encouraging to

see in Figure 9-13 that, in spite of the huge differences in flow conditions, the proposed

model still is able to match the trends observed with temperature when the normalized

values are compared. Both model and experimental data suggest that erosion-corrosion

damage at 150oF is about 1.7 larger than that obtained at 76oF. Very good agreement

between experimental data and predicted trends with temperature were also obtained for

the other two alloys under study (Super 13Cr and 22Cr), as displayed in Figure 9-14 and

Figure 9-15. These results suggest that erosion-corrosion damage at 150oF for Super 13Cr

is about 1.4 times that obtained at 76oF, while for 22Cr, the erosion-corrosion damage at

150oF is only about 1.1 times that obtained at 76oF.

160

Effect of material: comparison between experiments and prediction trends

To be able to compare all these experimental data and predicted values shown for

the three CRAs on a single plot, all experimental values were normalized to the

experimental value of the erosion-corrosion rate of 13Cr at 76oF, and all predicted values

were normalized to the predicted value for erosion corrosion of 13Cr at 76oF.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

0 20 40 60 80 100 120 140 160

Temperature (oF)

Eros

ion-

Co

rro

sio

n,E

C(m

py)

Model, Vl=15fts, 235kgd, Direct Impingement

Exp, Vsg=20fts, Vsl=1.4fts, 14.6kgd, Plugged T

13Cr, pH 4

Figure 9-13. Comparison between experimental data and predicted trends with

temperature of the erosion-corrosion of 13Cr. All data normalized to

room temperature values.

161

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

0 20 40 60 80 100 120 140 160

Temperature (oF)

Ero

sio

n-C

orr

osio

n,E

C(m

py)

Model, Vl=15fts, 235kgd, Direct Impingement

Exp, Vsg=20fts, Vsl=1.4fts, 14.6kgd, Plugged T

Super 13Cr, pH 4

Figure 9-14. Comparison between experimental data and predicted trends with

temperature of the erosion-corrosion of Super 13Cr. All data

normalized to room temperature values.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

0 20 40 60 80 100 120 140 160

Temperature (o F)

Ero

sio

n-C

orro

sio

n,E

C(m

py)

Model, Vl=15fts, 235kgd, Direct Impingement

Exp, Vsg=20fts, Vsl=1.4fts, 14.6kgd, Plugged T

22Cr, pH 4

Figure 9-15. Comparison between experimental data and predicted trends with

temperature of the erosion-corrosion of 22Cr. All data normalized to

room temperature values.

162

Figure 9-16 shows the comparison between experimental data and predicted

values for the three CRAs as shown in Figure 9-13 to Figure 9-15, but collected in on

single plot. It is clearly seen how the model was able to completely grasp the trends for

the erosion-corrosion behavior of the three alloys with respect to CRA type and

temperature. These results are encouraging, especially when considering the complexity

of the phenomena under study and the large number of variables taken into account in the

model. It is recognized that there are some opportunities to improve the model and some

of the ideas for this are presented in Chapter 10.

0.00.20.40.60.81.01.21.41.61.8

No

rmal

ized

Ero

sion

Co

rro

sion

,EC

(ref

13C

r,76

F)

13CrExper.

13CrModel

S13CrExper.

S13CrModel

22CrExper.

22CrModel

76 150Temp.

oF

76150

Exp, Vsg=20ft/s, Vsl=1.4ft/s, 14.6 kg/d sand, Plugeed TeeModel, Vl=15ft/s, 235 kg/d sand, Direct Impingement

Figure 9-16. Comparison between experimental data and predicted trends of the

erosion-corrosion, EC of the three CRAs at two different

temperatures. All data normalized to 13Cr at room temperature

value.

163

CHAPTER 9 ................................................................................................................................136

EROSION-CORROSION MODEL ................................ ................................................................ ....136

General Approach ................................ ................................................................ ...............136

Proposed procedure for estimating Ce-c. ................................................................ .............139

Determination of the indented open area............................................................................................142

Implementation of the second order model to determine the total current.......................................143

Validation of the Erosion-Corrosion prediction model .......................................................150

Adjustment of the erosion prediction..................................................................................................150

Comparison between experimental data and predictions...................................................................151

Some trends of predicted values..........................................................................................................154

Effect of sand rate: comparison between experiments and prediction trends ..................................154

Effect of temperature: comparison between experiments and prediction trends .............................159

Effect of material: comparison between experiments and prediction trends....................................160

FIGURE 9-1. SCHEMATIC OF A PARTICLE IMPINGEMENT ON A PASSIVE ALLOY. ........................................140

FIGURE 9-2. GENERIC PLOT FOR THE TOTAL CURRENT AND INDIVIDUAL CURRENT TRANSIENTS OF 6

DIFFERENT IMPACTS TAKING PLACE IN A NUMERICAL CELL OF THE TARGET SPECIMEN. .....................145

FIGURE 9-3. GENERIC PLOT FOR THE TOTAL CURRENT GENERATED IN A CELL DUE TO TOTAL PARTICLE

IMPACT. 146

FIGURE 9-4. TYPICAL BEHAVIOR OF THE TOTAL CURRENT PRODUCE IN A NUMERICAL CELL DUE TO THE

PARTICLE IMPACT. THE CELL IS HIT AT AN IMPACT FREQUENCY OF 74 PARTICLES PER SECOND. .........147

FIGURE 9-5. FLOW DIAGRAM TO ESTIMATE E, CE-C, & EC ................................ .............................149

FIGURE 9-6. COMPARISON BETWEEN MEASURED EROSION AND EROSION -CORROSION PENETRATION RATES

(BY WEIGHT LOSS) AND THOSE PREDICTED BY THE PROPOSED MODEL................................. ...............151

FIGURE 9-7. COMPARISON BETWEEN MEASURED CORROSION COMPONENT PENETRATION RATES (BY LPR)

AND THOSE PREDICTED BY THE PROPOSED MODEL................................................................. .............153

FIGURE 9-8. COMPARISON BETWEEN MEASURED TOTAL CURRENT VALUES (A) AND THOSE PREDICTED BY

THE PROPOSED MODEL. .......................................................................................................................153

FIGURE 9-9. PREDICTION TRENDS WITH SAND RATE OF PURE EROSION, E FOR SEVERAL FLOW VELOCITIES.

155

FIGURE 9-10. PREDICTION TRENDS WITH SAND RATE OF PURE EROSION, E AND CORROSION COMPONENT

OF EROSION-CORROSION, CE-C FOR SUPER 13CR AT TWO FLOW VELOCITIES. ................................ ....156

FIGURE 9-11. COMPARISON BETWEEN 13CR AND SUPER 13CR PREDICTION TRENDS WITH SAND RATE OF

PURE EROSION, E AND CORROSION COMPONENT OF EROSION-CORROSION, CE-C. ...............................157

164

FIGURE 9-12. EFFECT OF PH AND TEMPERATURE ON THE PREDICTION TRENDS WITH SAND RATE OF THE

CORROSION COMPONENT OF EROSION-CORROSION, CE-C FOR 13CR................................. ..................158

FIGURE 9-13. COMPARISON BETWEEN EXPERIMENTAL DATA AND PREDICTED TRENDS WITH

TEMPERATURE OF THE EROSION-CORROSION OF 13CR. ALL DATA NORMALIZED TO ROOM

TEMPERATURE VALUES................................. ................................................................ ......................160

FIGURE 9-14. COMPARISON BETWEEN EXPERIMENTAL DATA AND PREDICTED TRENDS WITH

TEMPERATURE OF THE EROSION-CORROSION OF SUPER 13CR. ALL DATA NORMALIZED TO ROOM

TEMPERATURE VALUES................................. ................................................................ ......................161

FIGURE 9-15. COMPARISON BETWEEN EXPERIMENTAL DATA AND PREDICTED TRENDS WITH

TEMPERATURE OF THE EROSION-CORROSION OF 22CR. ALL DATA NORMALIZED TO ROOM

TEMPERATURE VALUES................................. ................................................................ ......................161

FIGURE 9-16. COMPARISON BETWEEN EXPERIMENTAL DATA AND PREDICTED TRENDS OF THE EROSION-

CORROSION, EC OF THE THREE CRAS AT TWO DIFFERENT TEMPERATURES. ALL DATA NORMALIZED TO

13CR AT ROOM TEMPERATURE VALUE................................................................. ...............................162

1 Rincon, H.E., Chen, J. and Shadley, J.R. (2002). Erosion-Corrosion Phenomena of

13Cr Alloy in Flows Containing Sand Particles. Corrosion/2002, paper no. 2493,

(Houston, TX, USA, NACE International.

2 Levy, A. “Mechanisms of Erosion” in Solid Particle Erosion and Erosion-Corrosion

of Materials, ASM International, Chapter 2, pp 11-25, 1995

3 Oka, Y.I, Okamura, K., Yoshida, T., “Practical estimation of erosion damage caused

by solid particle impact. Part 1: Effects of impact parameters on a predictive

equation.” Wear 259 (95-101)

165

4 Oka, Y.I, Yoshida, T., “Practical estimation of erosion damage caused by solid

particle impact. Part 2: Mechanical properties of materials directly associated with

erosion damage.” Wear 259 (102-109)

5 M. Stern and E.D. Weisert, Proc. ASTM, Vol. 32, p.1280, 1959.

64 Wood, R.J.K. and Hutton, S.P. Wear 140(2): 387-394, 1990.

65 Stack, M.M., Zhou, S. and Newman, R.C. Wear 186-187(Part 2): 523-532. 1995

66 Neville, A. and Hu, X. (). Wear 251(1-12): 1284-1294, 2001.

67 Lu, B.T., Luo, J.L. and Lu, J.F. “Chemo-Mechanical Effect in Erosion-CorrosionProcess of Carbon Steel.” Corrosion/2004, paper no. 4659, (Houston, TX: NACE,2004).

68 Rincon, H., Shadley, J.R., and Rybicki, E.F. “Erosion Corrosion Phenomena of13Cr at Low Sand Rate Levels,” CORROSION/2005, paper no. 05291, (Houston,TX: NACE, 2005).

163

CHAPTER 10

SUMMARY, CONCLUSIONS AND RECOMMENDATIONS

Summary

The general objective of this work was to develop the knowledge base on the

roles of solid particle erosion and repassivation process of CRAs needed to advance and

expand erosion-corrosion predictive modeling for alloys commonly used in oil

production systems. The complexity of the chemical-mechanical mechanisms involved in

erosion-corrosion of metals, considering the large number of variables involved and the

diversity of service conditions found in oil and gas production, was discussed. Hence,

extensive experimental work was conducted. Effects of pH, temperature, flow velocity,

flow pattern, flow geometry, sand concentration, and material properties were studied

with respect to the erosion-corrosion behavior of CRAs.

Several standard techniques, such as weight loss (WL), electrical resistance (ER),

linear polarization resistance (LPR), and potentiodynamic polarization scans (PD) were

used to provide a broad understanding of the mechanisms involved in the erosion-

corrosion of CRAs in flows containing sand. Also, a relatively new experimental

approach and test called the “scratch test” (ST), which may be considered as a modified

164

electrochemical noise test, or modified galvanic test, was used to better understand the

reforming film process taking place during the repassivation of CRAs.

Some characterization of eroded and eroded-corroded surfaces was done by

techniques such as Scanning Electronic Microscope (SEM), Energy Dispersive

Spectroscopy (EDS), X-Ray Diffraction, and Ramman Spectroscopy.

Based on all experimental evidence collected, an effective and efficient procedure

for investigating the erosion-corrosion behavior of CRAs in oilfield environments was

developed. The idea was to reduce the need for expensive and time consuming loop tests

by using the simplified scratch test, and still be able to predict erosion-corrosion rates at

flow conditions.

In addition, a frame-work for a semi-mechanistic model to predict erosion-

corrosion of CRAs has been developed and implemented based on the solid particle

impingement information collected from CFD simulations and experimental data

obtained from scratch tests on the repassivation of CRAs, which was found to

approximate the second order behavior. Estimates of pure erosion, impact-induced

corrosion, and synergistic effects of the erosion-corrosion damage have been provided for

in the model. Only the prediction of erosion rate is performed during the CFD

simulations; the impact-induced corrosion and erosion-corrosion rates are not actually

executed during the CFD simulations, but rather in a separate VBA code built for that

specific purpose.

Experimental results and predicted values were compared for a broad range of

conditions with encouraging results. Hence, the proposed model provides the basis and

framework for a erosion-corrosion predictive model that eventually could be used to

165

predict erosion-corrosion penetration rates for oil industry service conditions in a reliable

manner.

In the initial feasibility investigation of the proposed framework, the model

assumed some fixed parameters such a brine concentration, water chemistry and CO2

pressure. Input parameters are temperature, pH (which are implicit in the second order

scratch test parameters, m and Io), alloy, solid particle rate, particle size, and liquid

velocity, as well as physical properties of the fluid, solid particle, and target material.

Output of the model includes the pure erosion rate, E, the total erosion-corrosion

penetration rate, EC, and the corrosion component rate of the erosion-corrosion process,

Ce-c, defined as the sum of the pure corrosion C and the impact-induced corrosion Ce.

Conclusions

Scratch Tests

The scratch test results showed how strongly environmental conditions such as

temperature and pH affect the repassivation process of CRAs exposed to erosion-

corrosion conditions.

The initial healing rate of the oxide layer, immediately after the scratch action, is

greater for higher temperatures. However, because initial corrosion rates are also higher

at higher temperatures, the repassivation times are longer for higher temperatures; hence

the per scratch damage increased with increasing temperature. Therefore, greater erosion-

corrosion damage is expected at higher temperatures

At all temperatures the corrosion currents and cumulative thickness loss increased

with decreasing pH. The repassivation process is faster for higher pH values, so the

166

corrosion component of the erosion-corrosion process should be expected to be lower for

higher pH values.

Second order behavior appears to be an appropriate and useful model to represent

the repassivation process of CRAs at least for the first 200 seconds. Also, good

agreement between the repassivation times obtained by means of the current data time

series approach and the second order kinetics approximation was found. This suggests

that for the conditions examined in this research, the second order approach can be

usefully applied beyond the initial repassivation periods.

Repassivation processes for 13Cr and Super 13Cr showed similar trends with

respect to temperature and pH. However Super 13 was found to repassivate much faster

and to a lower corrosion rates as compare to 13Cr. The effect of temperature on the

repassivation process of 22Cr at pH 4 was observed to be not as strong as it is for 13Cr

and super13Cr. The 22Cr alloy showed the fastest repassivation leading to the lowest

corrosion rates of the CRAs.

Multiphase gas/liquid/sand flow loop tests

At high erosivity conditions, a synergistic effect between erosion and corrosion

was confirmed since the total metal loss rate was shown to be higher than the sum of the

rates of corrosion (without sand) and erosion measured separately. It appears that under

the high sand rate conditions tested, the erosivity is severe enough to damage the passive

layer thereby enhancing the corrosion rate.

167

Synergism seems to occur for each of the three alloys; however, the degree of

synergism is quite different for the three alloys and may not be significant for 22Cr under

field conditions where erosivities are typically much lower that those occurring in the

small bore loop used in this testing.

For the others two alloys, 13Cr and Super13Cr, especially at high temperatures, if

the erosion rate of the passive film is great enough, then an accelerated erosion-corrosion

process may take place with a significant contribution from a corrosion component. In

most cases there is likely a competition between the protective film removal due to

mechanical erosion and the protective film healing. If erosivity conditions are severe

enough, base metal can also be removed by the mechanical erosion component. The

mechanistic model of this competition depends, among others, on film characteristics and

the concentration and distribution of sand as well as on the flow pattern and fluid flow

velocities, geometry and environmental factors. The erosion-corrosion behavior appears

to be a function of both material properties and the erosivity of environmental conditions.

At given environmental and flow conditions the ranking of the alloys was always

the same. 13Cr was consistently showing higher erosion-corrosion rates than super 13Cr,

and 22Cr showed the lowest erosion-corrosion rates. Also erosion-corrosion rates and

impact-induced corrosion rates of 13Cr was shown to be much more sensitive to effects

of temperature and flow conditions than those shown by Super 13Cr and 22Cr.

At low sand rates (background sand levels), the penetration rates were in general

much lower than for high sand rates for both pure erosion and erosion-corrosion

conditions. In fact, the pure erosion penetration rates and the erosion-corrosion

penetration rates at low sand rates were not statistically different. This was found at both

168

76 (pure Eavg = 3.0 mpy: E-Cavg = 3.6 mpy) and 150oF (pure Eavg = 7.4 mpy: E-Cavg =

10.0 mpy). The small corrosion component of the erosion-corrosion penetration rates at

low erosivity conditions suggests that the 13Cr oxide layer was healing fast enough to

remain substantially intact against the erosive attack of the impinging sand particles.

Therefore, the corrosion component of the erosion-corrosion process would be similar to

the pure corrosion rate under sand-free conditions. However, the oxide layer structure and

its properties may not be the same as the original layer before adding the sand, and more

research needs to be done in this respect. The synergistic effect of erosion-corrosion on

13Cr appears to have a threshold erosivity condition where the corrosion component of

the erosion-corrosion process starts to play a dominant role in the whole metal loss

process. This threshold would depend on the same number of variables that the erosion-

corrosion process itself depends on; hence its determination is not straight forward. Here

is where the combination of efficient experimental procedures and predictive models,

such as those proposed in this research, appear to have promising potential for guiding

the design and material selection for production systems required to handle sand.

Single phase liquid/sand flow loop testing (submerged direct impingement test)

The scratch test cell conveniently separates the mechanical process (mechanical

removal of material) from the electrochemical process (corrosion). Thus, this technique

provides a good way to the study of the effect of environmental conditions on the

repassivation of CRAs, although is not useful to the study of many other important

parameters affecting the erosion-corrosion process.

169

However, the instrumentation and the configuration of the electrodes used for

scratch test were successfully adapted to the submerged direct impingement flow loop

testing system as described in Chapter 4. Thus, the recording of current transients for

CRAs being continuously impinged by sand particles was possible.

The new experimental set up allowed the study of the mechanical and

electrochemical processes taking place simultaneously and continuously while solid

particles impinged the target material.

For a given condition, corrosion currents for flows containing sand were much

higher than those without sand. In general, once sand was filtered out of the flowing

brine, the corrosion currents immediately decreased with time suggesting that the

repassivation process taking place on the target specimen could be accomplished once the

mechanical removal of the passive film has stopped.

Once sand impingement is stopped, the high anodic current immediately began to

decrease approximately following a second order model, similar to that observed for the

static scratch test. In fact the second order model was proven to be a very good

approximation to describe the repassivation process of CRAs under dynamic flow

conditions tested for a significantly long period of time. Similarities between the behavior

of the anodic current decay for the scratch test and the microloop testing, and the fact

that a single solid particle impingement can be thought as a micro scratch made on the

passive film are the reasons why this type of testing has been referred to in this research

as a “Dynamic Scratch Test.”

170

Submerged direct impingement test results also showed that the extent to which

the current increases strongly depends on parameters like sand rate and flow velocity;

however, the slope in the second order model for these tests was only very slightly

affected by the sand rate. However, it is believed than the slopes may have been more

different if the sand rates were much more different.

An interesting phenomenon, when longer periods of time were used to test 13Cr

alloy at severe conditions such a low pH, high flow velocities, and high sand rates, was

observed. At such aggressive conditions, high current magnitudes were sustained even

after removing the sand, the second order repassivation did not result, and a black coating

grew on the surface of the 13Cr target. No cases like this have been reported to happen

for field conditions in the literature, and its study was beyond the scope of this research.

However, preliminary characterization of the black coating was performed by means of

several techniques and it is known to be rich in Fe and Cr, suggesting it is a combined Fe-

Cr oxide. However, the exact chemical composition and crystalline structure has not

clearly determined yet and more research needs to be done in this matter.

If the combined mechanical and environmental conditions are not that severe, for

example, similar high flow velocities and sand rates but higher pH, the black coating

formation was prevented, the second order repassivation behavior was observed, and a

typical CRAs surface finish was obtained. This is probably a case more representative of

field conditions.

171

Erosion-corrosion predictive procedure and model

Trends of the erosion-corrosion damage obtained from scratch test results were

compared with those obtained from erosion-corrosion flow loop tests conducted at single

phase liquid flows and multiphase flows. Flow loop tests gave trends similar to trends

exhibited by scratch tests results for the three CRAs at all tested conditions; therefore

corroborating our scratch tests observations.

A procedure to predict penetration rates for erosion-corrosion conditions was

developed based on the second order model behavior observed in the re-healing process

of the passive film of CRAs under tested conditions. Good agreement between the actual

and predicted penetration rates was found. Predictions of the corrosion component of

erosion-corrosion based on scratch test data compared well to test results from multiphase

gas/liquid flow loop for the three CRAs at high erosivity conditions. Second order

repassivation behavior appears to be important process in the erosion-corrosion of CRAs.

In spite of the great advances recently achieved in the application of CFD

modeling to the prediction of erosion rates, the existing CFD models still greatly over

predict erosion rates due to particle impingement in single phase liquid flows.

A simple multiplier adjustment, determined by matching the CFD predicted

erosion rate to experimental flow loop data, was made to provide the needed erosion rate

starting point for the current erosion-corrosion prediction model.

No further adjustments or any fitting constants were used for the estimates of the

corrosion component of erosion-corrosion and the total erosion-corrosion rates.

172

Experimental data obtained for corrosion component rates of erosion-corrosion (LPR),

and total erosion-corrosion (WL), were in reasonable agreement with those predicted

values obtained with the proposed model at the same condition. However, significant

scatter in the data was observed and the need for more reliable data to better validate the

model is recognized.

Also trends predicted by the model with trends observed in single phase liquid

testing and a more reliable set of data for multiphase flow testing showed very good

agreement. In general predicted results are encouraging, especially when considering the

complexity of the phenomena under study and the large number of variables taken into

account in the model.

The proposed framework for the semi-mechanistic model not only addresses

prediction of erosion rates, corrosion rate components of erosion-corrosion, and erosion-

corrosion rates, but also predicts localized variation of the erosion-corrosion damaged of

CRAs based on the flow conditions and flow geometry. The erosion-corrosion damage

characterized by the concentric ring pattern obtained on the surface of the 13Cr direct

impingement specimen, was closely reproduced by the material degradation patterns

obtained for erosion using computational fluid dynamics simulations, and also by the

erosion-corrosion patterns obtained by applying the proposed semi-mechanistic model

for the estimation of localized erosion-corrosion damage.

Recommendations

The Second order behavior displayed by transients obtained with scratch tests has

proved to be an appropriate and useful model to represent the repassivation process of

173

CRAs, hence the scratch test has been one of the key factors in the success of the

construction of the framework for the erosion-corrosion predictive model. Limitations on

the number of variables that the proposed model manages depend to a great extent on the

experimental data obtained from scratch tests.

Using the scratch test to study the effects of other variables affecting the

repassivation process of active/passive alloys is recommended. Chloride content, CO2

pressure, water chemistry and H2S on the erosion corrosion of CRAs are some of the

parameters that can be easily added to the model by expanding the data base of Scratch

Testing results. Furthermore it is believed that the technique can be slightly modified and

adapted to the study of the effect of inhibitors on the erosion-corrosion of carbon steel.

This research has provided some insight into how difficult and complex the

erosion-corrosion topic may be. Establishing comparisons between the Stainless Steel

family (13Cr, Super 13Cr, 22Cr), and other CRA families such as high nickel alloys and

titanium alloys, would be extremely useful for making material selection decisions in the

oil and gas industry. Further evaluation with some flow loop testing might be

recommendable to assure Scratch Test is still providing reliable information in the

repassivation process of other CRA families.

Experimental determination of the second order constants m and Io from scratch

tests has been a key component in the procedure used in erosion-corrosion predictive

model framework proposed in this research. This experimental approach might be

replaced in the future with a more mechanistic model for determining the m and Io

parameters.

174

Potentiodynamic scans showed an anodic shift in the primary passivation

potential Epp along with the increment in the passive current density ip, for 13Cr exposed

to the combined effect of high erosivity flows and severe environmental conditions for

the direct impingement testing (see Appendix C). For these conditions 13Cr forms a

black coating and the repassivation is prevented. A more systematic investigation into the

conditions under which 13Cr will developed the black coating is recommended along

with more accurate characterizations of the chemical composition and crystalline

structure of the components of the black layer.

Similar scans conducted for less severe flow conditions in a 13Cr elbow flow

geometry for similar environmental conditions showed similar increment in the passive

current density ipass, but, the anodic shift in the primary passivation potential Epp never

occurred at this conditions. For this condition, no black coating was observed and

repassivation was accomplished. This suggests that the requirements to accomplish

repassivation might depend on the severity of flow conditions and the erosivity level in

addition to environmental conditions such pH, temperature and chloride contents and

water chemistry.

Based on these results a systematic study of the effect of erosivity (flow

velocities, flow geometry, sand concentration, sand shape, and sand size) on the passive

parameters such as primary passivation potential Epp, critical current density, icrit and

passive current density ip is recommended (see Figure 2-2 in Chapter 2).

Differences between the erosion-corrosion behaviors of 13Cr and super 13Cr

22Cr and other CRAs might be revealed in the light of potentiodynamic scans conducted

175

for flows containing sand, thus improve the understanding of the mechanisms under

which the erosion-corrosion process is taking place on active/passive surface alloys.

Better characterizations of particle indentations on the passive surfaces might

improve the accuracy of the erosion-corrosion prediction model proposed. This would

involve collection of data on the shapes and depths of indents so that the amount of new

bare surface created by the erosion component could be better represented in the model.

Based on the need for determining threshold conditions for which a synergistic

erosion-corrosion is seen in active/passive alloys used in the oil industry, further

developed of the erosion-corrosion prediction model proposed here is advised.

The erosion-corrosion prediction model depends, as a starting point, on the

accurate prediction of ersosion rates on the metal surface using a CFD model. But,

because the existing CFD models greatly over predict erosion from sand particles in

single phase liquid flows, a simple multiplier adjustment of the erosion rates estimated by

CFD was made to provide the needed erosion rate starting point for the current erosion-

corrosion prediction model. Need for improvement in this area of CFD modeling has

been already recognized, and research is on-going to improve the erosion equations and

prediction accuracy in single phase liquid flows.

Validation of the proposed erosion-corrosion prediction model suffered partly

because of the lack of reliable experimental data for erosion and erosion-corrosion in

sand bearing single phase flows. More reliable experimental data (higher weight loss

conditions) obtained from single phase liquid erosion testing conducted at higher flow

velocities is needed for future validation and/or adjustment of the erosion-corrosion

prediction model. The Erosion/Corrosion Research Center has a new erosion-corrosion

176

test loop capable of flow velocities in excess of 40 ft/s which could be used to generate

the needed higher weight loss data. This work and validation/improvement of the model

would be a good masters research project, and should be a high priority item in advancing

E/CRC modeling capability in erosion-corrosion.

In the light of the good agreement showed by predicted and measured values of

the corrosion rate component of the erosion-corrosion process, no further adjustments or

fitting constants may be required in the procedure of the prediction model that estimates

the corrosion component of erosion-corrosion and the total erosion-corrosion rates.

However, more reliable erosion-corrosion data to check this procedure is also needed. In

addition to LPR measurements, data based on weight loss and high resolution ER

technique for erosion-corrosion of CRAs in liquid containing sand are advised.

As this research has demonstrated, flow geometry is another important issue to be

considered in erosion-corrosion prediction. The present form of the erosion-corrosion

prediction model proposed here does a good job of providing information on the erosion-

corrosion penetration rate distribution. A systematic study to compare experimental data

and predictions for different flow geometries is also advised. Flow geometries of greatest

interest to the oil and gas production industries would include elbows and direct

impingement geometries such as headers and tees.

177

CHAPTER 10 ..............................................................................................................................163

SUMMARY , CONCLUSIONS AND RECOMMENDATIONS ................................................................163

Summary ..............................................................................................................................163

Conclusions .........................................................................................................................165

Scratch Tests.........................................................................................................................................165

Multiphase gas/liquid/sand flow loop tests.........................................................................................166

Single phase liquid/sand flow loop testing (Submerged Direct Impingement Test) .......................168

Erosion-corrosion predictive procedure and model ...........................................................................171

Recommendations................................................................................................................172

177

REFERENCES

1 “Corrosion Cost and Preventive Strategies in the United States” A supplementto Materials Performance Journal, July 2002, p.2.

2 J. D. Combes, J. G. Kerr and L. J. Klein, “13Cr Tubulars Solve CorrosionProblems in Tuscaloosa Trend,” Petroleum Engineer International, March,1983.

3 M. S. Cayard and R. D. Kane, “Serviceability of 13Cr Tubulars in Oil and GasProduction Environments,” CORROSION/98, paper no.112, (Houston, TX:NACE, 1998).

4 H. Amaya, K. Kondo and H. Hirata, “Effect of Chromium and Molybdenumon Corrosion Resistance of Super 13Cr Martensitic Stainless Steel in CO2

Environment,” CORROSION/98, paper no.113, (Houston, TX: NACE, 1998).

5 G. Fierro, G.M. Ingo and F. Mancia, “XPS Investigation on the CorrosionBehavior of 13Cr Martensitic Stainless Steel in CO2-H2S-Cl- Environments,”CORROSION/88, paper no.215, (Houston, TX: NACE, 1988).

6 L.J. Klein, “I & S Cracking Resistance of Type 420 Stainless Steel Tubulars,”CORROSION/84, paper no.211, (Houston, TX: NACE, 1984).

7 A. Tamaki, “A New 13Cr OCTG for High Temperature and High ChlorideEnvironment,” CORROSION/89, paper no.469, (Houston, TX: NACE, 1989).

8 A. Miyasaka and H. Ogawa, “Influence of Metallurgical Factors on CorrosionBehaviors of Modified 13Cr Martensitic Stainless Steels,” CORROSION/90,paper no.28, (Houston, TX: NACE, 1990).

9 M. B. Kermani, D. Harrop, M.L.R. Truchon and J.L. Crolet “ExperimentalLimits of Sour Service for Tubular Steels,” CORROSION/91. Paper no.21,(Houston, TX: NACE, 1991).

10 M.F. Galis and J.J. Damman, “13 Percent Chromium Steels for Slightly SourService,” CORROSION/91, paper no.20, (Houston, TX: NACE, 1991).

11 S. Hashizume, T. Takaoka, Y. Minami, Y. Ishizawa and T. Yamada, “A New15 Percent Steel Developed for OCTG,” CORROSION/91, paper No.21,(Houston, TX: NACE, 1991).

178

12 M. Ueda, T. Kushida, K. Kondo, and T. Kudo “Corrosion Resistance of 13Cr-5Ni-2Mo Martensitic Stainless Steel,” CORROSION/92, paper no.55,(Houston, TX: NACE, 1992).

13 T. Mori, T. Okazawa, M. Ueda and T. Kushidda, “Corrosion Performance ofSuper 13Cr Steel in CO2 Environments Containing a Small Amount of H2S”,Proceedings from the Second NACE Asian Conference-Corrosion 1994,(Houston, TX: NACE International, 1994).

14 S.F. Biagiotti, Jr. and J.S. Reichman, “Justifying the Use of 13Cr Steels forCorrosive CO2 Operations,” CORROSION/95, paper no.81, (Houston, TX:NACE, 1995).

15 H. Asahi, T. Hat-a, A. Kawakami and A. Takahashi., “Development of SourResistant Modified 13Cr OCTG,” CORROSION/95, paper no.79, (Houston,TX: NACE, 1995).

16 M. Ueda, T. Kushida and T. Mori, “Evaluation of SSC Resistance on Super13Cr Stainless Steel in Sour Applications,” CORROSION/95, paper no.80,(Houston, TX: NACE, 1995).

17 Y. Miyata, Y. Yamane, O. Furukimi, H. Niwa and K. Tamaki., “Corrosion ofNew 13Cr Martensitic Stainless Steel OCTG in Severe CO2 Environment,”CORROSION/95, paper no.83, (Houston, TX: NACE, 1995).

18 H. Asahi, T. Hara, and M. Sugiyama, “Corrosion Performance of Modified13Cr OCTG,” CORROSION/96, paper no.61, (Houston, TX: NACE, 1996).

19 T.F. Illson and R. Greenwood, “Autoclave Testing of 13Cr and Modified13Cr Stainless Steels in Simulated Field Conditions,” CORROSION/95, paperno. 87, (Houston, TX: NACE, 1995).

20 S. Hashizume, K. Masamura and Y. Ishizawa, “Corrosion Data on LaboratoryPerformance of Type 420 for OCTG,” CORROSION/95, paper no.77,(Houston, TX: NACE, 1995).

21 M. Ueda and A. Ikeda ‘Effect of Microstructure and Cr Content in Steel onCO2 Corrosion,” CORROSION/96, paper no.13, (Houston, TX: NACE,1996).

22 J. M. A. Van Der Horst, C.R. Sloan, “Erosion Corrosion of Finned HeatExchanger Tubes,” Erosion, Wear and Interfaces with Corrosion, ASTM STP567, American Society for Testing and Material, Philadelphia 1973, pp. 18-29.

179

23 Salama, M.M and Venkatesh E.S. “Evaluation of API RP 14E ErosionalVelocity Limitations for Offshore Gas Well,” Offshore TechnologyConference, Paper OTC 4485, 1983.

24 Smart III, J.S. “The Meaning of the API-RP-14E Formula for ErosionCorrosion in Oil and Gas Production,” CORROSION/91, paper. 910468,(Houston, TX: NACE, 1991)

25 Smart III, J.S. “A Review of Erosion-Corrosion in Oil and Gas Production,”CORROSION/90, paper 90010, (Houston, TX: NACE, 1990).

26 Salama, M.M. “Erosional Velocity Limits in Water Injection Systems,”CORROSION/93 paper 93062, (Houston, TX: NACE, 1993).

27 Shirazi, S. McLaury, B, Shadley, J. and Rybicki, E. “Generalization of theAPI RP 14E Guidelines for Erosive Services,” 69th Annual TechnicalConference and Exhibition, paper SPE 28518, 1994.

28 Jordan, K. “Erosion in Multiphase Production of Oil & Gas,”CORROSION/98 paper 98058, (Houston, TX: NACE, 1998).

29 Salama, M.M. “An Alternative to API 14E Erosional Velocity Limits for SandLaden Fluids,” Offshore Technology Conference, Paper OTC 8898, 1998.

30 McLaury, B. and Shirazi, S. “An Alternate Method to API RP 14E forPredicting Solids Erosion in Multiphase Flows,” Journal of Energy ResourcesTechnology, ASME, Vol. 122, pp 115-122, Sep 2000.

31 D. A. Jones, Principle and Prevention of Corrosion, Prentince Hall, 2nd ed.,USA, 1996.

32 Corrosion Control in Petroleum Production, NACE International, TPC 5publication Texas, 1979.

33 Application Notes, Gamry Instruments website, www.Gamry.com

34 ASTM Standard Designation: G 102 “Standard Practice for Calculation ofCorrosion Rates and Related Information from ElectrochemicalMeasurements” 03.02.

35 L.L.Shreir, Corrosion: Metal/Environment Reactions, Newnes-Butterworths,London, 1976.

36 ASTM Standard Designation: G 5 “Practice for Standard reference Methodfor Making Potentiostatic and Potentiodynamic Anodic polarizationMeasurements” 03.02

180

37 Mansfeld, F. “Don’t Be Afraid of Electrochemical Techniques, But Use Themwith Care”. Corrosion, Vol. 44, no.12, pp856-868.

38 J.R.Davis, Stainless Steels, ASM Specialty Handbook.

39 U.Lotz, E. Heitz. “Flow-Dependeant Corrosion. I. Current Understanding ofthe Mechanisnsm Involved” Werkstoffe und Korrosion, 34, 1983, pp. 454-461.

40 K. D. Efird, “Disturbed Flow And Flow Accelerated Corrosion In Oil AndGasProduction”, Energy Sources Technology, ASME, 1998.

41 De Waard C. & Milliams D.E., Corrosion, 31, 131 (1975).

42 Ogundele g.l. & White W.E., Corrosion ,42, 71 (1986).

43 Schmitt G. & Rothmann B. Werkstoffe, u. Korrosion, 28, 816 (1977).

44 Wieckowski A., Ghali E., Szklarczyk M. & Sobkowski J., ElectrochimicaActa 28 1619 (1983).

45 Eriksud E. & Sontvedt T. Advance in CO2 Corrosion, (ed Hausler R,GodardH.P.) NACE p. 20 (1984).

46 S. Turgoose & R.A. Cottis, “Modeling of electrode processes and surfacechemistry in CO2 corrosion”.

47 J. L. Crolet, M. R. Bonis, “Experience in the Use of 13% Cr Tubing incorrosive CO2 fields”, SPE Production Engineering, Sep 1986.

48 T. Hara, H. Asahi, H. Kaneta, “Effect of Flow Velocity on CO2 Corrosionand Galvanic Corrosion Behavior in Oil and Gas Environments,”CORROSION/98 paper 118.

49 J.R. Davis, Surface Engineering for Corrosion and Wear Resistance, ASMInternational, march, 2001.

50 A.V. Levy, Solid Particle Erosion and Erosion Corrosion of Material, ASMInternational, Ohio, 1995.

51 R. Chattopadhyay, “Surface Wear Analysis, Treatment and Prevention”, ASMInternational.

52 T. Kohley, E. Heitz, “Particle Containing Formation Water for the Study ofErosion Corrosion”, The Use of Enviroments for Corrosion Testing, ASTM

181

STP 970, American Society for Testing and Material, Philadelphia 1988, pp.235-245.

53 P. M. Birchenough, S. G. B. Dawson, T. J. Lockett, and P. McCarthy,“Simultaneous Erosion and Corrosion in Multiphase Flow” NACE 7th MiddleEast Corrosion Conference, Bahrain, 26-28 February 1996.

54 ASTM Standard Designation: G 119 “Standard Guide for DeterminingSynerginsm between Wear and Corrosion.” 03.02.

55 R.J.K Wood, S.P. Hutton, Wear, 186-187, pp. 523-532, 1995.

56 A. Neville, H. Xu, M. Reyes, “Corrosion and Erosion-Corrosion Behavior of aCo-Based Alloy and a Ni-Containing Austenitic Cast Iron”,CORROSION/2000, paper no. 628, (Houston, TX: NACE, 2000).

57 Rincon, H.E., Chen, J. and Shadley, J.R. “Erosion-Corrosion Phenomena of13Cr Alloy in Flows Containing Sand Particles.” Corrosion/2002, paper no.2493, (Houston, TX: NACE, 2002).

58 Chen, J., Shadley, J.R., Rincon, H.E. and Rybicki, E.F. “Effects ofTemperature on Erosion-Corrosion of 13Cr.” Corrosion/2003, paper no.3320,(Houston, TX: NACE, 2003).

59 Rincon, H.E. (2001). Erosion-Corrosion Phenomena of 13Cr Alloy in FlowsContaining Sand Particles, M.S. Thesis, Department of MechanicalEngineering, The University of Tulsa, Ok, Tulsa.

60 A. McMahon and J. Martin “Simulation Tests on the Effect of MechanicalDamage or Acid Cleaning on CRAs Used for Oil/Gas Production WellTubulars,” CORROSION/04, paper no.4127, (Houston, TX: NACE, 2004).

61 Oka, Y.I, Okamura, K., Yoshida, T., Wear 259 (95-101), 2005.

62 Oka, Y.I, Yoshida, T., Wear 259 (102-109), 2005.

63 M. Stern and E.D. Weisert, ASTM, Vol. 32, p.1280, 1959.

64 Wood, R.J.K. and Hutton, S.P. Wear 140(2): 387-394, 1990.

65 Stack, M.M., Zhou, S. and Newman, R.C. Wear 186-187(Part 2): 523-532. 1995

66 Neville, A. and Hu, X. (). Wear 251(1-12): 1284-1294, 2001.

67 Lu, B.T., Luo, J.L. and Lu, J.F. “Chemo-Mechanical Effect in Erosion-CorrosionProcess of Carbon Steel.” Corrosion/2004, paper no. 4659, (Houston, TX:NACE, 2004).

182

68 Rincon, H., Shadley, J.R., and Rybicki, E.F. “Erosion Corrosion Phenomenaof 13Cr at Low Sand Rate Levels,” CORROSION/2005, paper no. 05291,(Houston, TX: NACE, 2005).

183

APPENDIX A

REMAINING SCRATCH TEST RESULTS

This section presents data obtained from Scratch Test that were not included in

Chapter 5. However, results showed in this Appendix showed similar trends than those

showed in Chapter 5, therefore no additional comments are included here.

Effect of pH

0

5

10

15

20

25

30

35

40

45

0 50 100 150 200Time (sec)

I( m

A)

T= 76ºF

13Cr_pH=4.00

13Cr_pH=5.25

13Cr_pH=3.50

22Cr_pH=4.00

Figure A- 1. Effect of pH on the repassivation process of 13Cr at 76°F, 22Cr at

pH 4 is also included.

184

0

5

10

15

20

25

30

35

40

45

0 50 100 150 200Time (sec)

I( m

A)

22Cr_pH = 4.0

13Cr_pH=3.5 T= 200 ºF

13Cr_ pH=5.25

13Cr_pH=6.0

13Cr_pH=4.0

Figure A- 2. Effect of pH on the repassivation process of 13Cr at 200°F, 22Cr at

pH 4 is also included.

0.E+00

2.E+05

4.E+05

6.E+05

0 50 100 150 200

Time (sec)

1/I,

(1/A

)

T= 76ºF13Cr

pH = 3.5

pH = 4.0

pH = 5.25

pH = 4.022Cr

Figure A- 3. 1/I vs. t from the raw current data in Figure A- 1.

185

0.E+00

1.E+05

2.E+05

3.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

)

13Cr_pH=4.0

13Cr_pH=5.25

13Cr_pH=3.5

T= 200ºF

22 Cr,pH=4.0

13Cr_pH=6.0

Figure A- 4. 1/I vs. t from the raw data in Figure A- 2.

0

5

10

15

20

25

30

35

40

45

0 50 100 150 200Time (sec)

I( m

A)

T= 76ºF

Super13Cr_pH=3.50Super13Cr_pH=4.00Super13Cr_pH=5.25Super13Cr_pH=6.00

Figure A- 5. Effect of pH on the repassivation process of Super13Cr 76°F.

186

0

5

10

15

20

25

30

35

40

45

0 50 100 150 200Time (sec)

I( m

A)

Super13 Cr_pH=3.5

T= 200 ºF

Super 13Cr_ pH=5.25

Super 13Cr_pH=6.0

Super 13Cr_pH = 4.0

Figure A- 6. Effect of pH on the repassivation process of Super 13Cr at 200°F.

0.E+00

2.E+05

4.E+05

6.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

)

T= 76ºFSuper 13Cr

pH=3.5

pH=4.0pH=6.0

pH=5.25

Figure A- 7. 1/I vs. t from the raw data in Figure A- 5

187

0.E+00

1.E+05

2.E+05

3.E+05

4.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

) pH=4.0pH=5.25

pH=3.5

pH = 6T= 200ºFSuper 13Cr

Figure A- 8. 1/I vs. t from the raw data in Figure A- 6

Effect of type of material (CRA) on Scratch Test results

0

5

10

15

20

25

30

35

40

45

0 50 100 150 200Time (sec)

I( m

A)

T= 76ºF

Super _13Cr3.5 pH 6

13Cr_pH=4.00

13Cr_pH=5.25

22Cr_pH=4.00

13Cr_pH=3.50

Super13Cr_pH=3.50Super13Cr_pH=4.00Super13Cr_pH=5.25Super13Cr_pH=6.00

Figure A- 9. Comparison of the repassivation process of 13Cr and Super 13Cr at

76°F, 22Cr at pH 4 and room temperature is also included.

188

0

5

10

15

20

25

30

35

40

45

0 50 100 150 200Time (sec)

I( m

A)

T= 150 ºF13Cr_pH=3.5

13Cr_pH=4.0

13Cr_pH=5.25

13Cr_pH=6.0

22Cr pH4.0

Super_13Cr3.5 pH 6.0

Super13Cr_pH=3.50Super13Cr_pH=4.00Super13Cr_pH=5.25Super13Cr_pH=6.00

Figure A- 10. Comparison of the repassivation process of 13Cr and Super 13Cr at

150°F, 22Cr at pH 4 is also included.

0

5

10

15

20

25

30

35

40

45

0 50 100 150 200Time (sec)

I( m

A)

T= 200 ºF

22Cr_pH = 4.0

13Cr_pH=3.5

13Cr_ pH=5.25

13Cr_pH=6.0

13Cr_pH = 4.0

Super13 Cr_pH=3.5

Super 13Cr_ pH=5.25

Super 13Cr_pH=6.0

Super 13Cr_pH = 4.0

Figure A- 11. Comparison of the repassivation process of 13Cr and Super 13Cr at

200°F, 22Cr at pH 4 is also included.

189

0.E+00

2.E+05

4.E+05

6.E+05

0 50 100 150 200Time (sec)

1/I

,(1/

A)

T= 76ºF

13Cr_pH=3.5

13Cr_pH=4.0

Super13CrpH=5.25

Figure A- 12. Comparison of 1/I vs. t for 13Cr and Super 13Cr at 76°F, 22Cr at pH

4 is also included.

0.E+00

1.E+05

2.E+05

3.E+05

4.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

)

13Cr_pH=4.0

13Cr_pH=5.25

13Cr_pH=3.5

Super13Cr_pH=6.0

T= 200ºF

22 Cr,pH=4.0

Super13Cr_pH=5.25Super13Cr_pH=4.00Super13Cr_pH=3.50

13Cr_pH=6.0

Figure A- 13. Comparison of 1/I vs. t for 13Cr and Super 13Cr at 200°F, 22Cr at pH

4 is also included.

22Cr

pH=4

.0Sup

er13

Cr pH=6.0

Super 13Cr pH

=4.0Sup

er 13Cr pH=3.5

13Cr pH=5.25

190

0.0E+00

2.0E-06

4.0E-06

6.0E-06

8.0E-06

1.0E-05

1.2E-05

0 50 100 150 200Time (sec)

Th

ickn

ess

Lo

ss(m

m)

T=76oF

22Cr_pH=4

13CrpH=3.5

13Cr_pH=4.0

13Cr_pH=5.25

Figure A- 14. Effect of pH on the cumulative thickness loss of the 13Cr at 76°F,

22Cr at pH 4 is also included. Integration of the current raw data.

0.0E+00

5.0E-06

1.0E-05

1.5E-05

2.0E-05

2.5E-05

0 50 100 150 200Time (sec)

Th

ickn

ess

Lo

ss(m

m)

T=150oF 13Cr_pH=3.5

22Cr_pH=4

13Cr_pH=4.0

13Cr_pH=5.25

13Cr_pH=6.0

Figure A- 15. Effect of pH on the cumulative thickness loss of the 13Cr at 150°F,

22Cr at pH 4 is also included. Integration of the current raw data.

191

0.0E+00

1.0E-05

2.0E-05

3.0E-05

4.0E-05

5.0E-05

6.0E-05

0 50 100 150 200

Time (sec)

Thic

kne

ssL

oss

(mm

)

13Cr_pH=3.5

13Cr_pH=4.0

13Cr_pH= 6.0

T=200oF

22Cr_pH=4

Figure A- 16. Effect of pH on the cumulative thickness loss of the 13Cr at 200°F,

22Cr at pH 4 is also included. Integration of the current raw data.

0.0E+00

1.0E-06

2.0E-06

3.0E-06

4.0E-06

5.0E-06

0 50 100 150 200Time (sec)

Th

ickn

ess

Lo

ss(m

m)

T=76oF

22Cr_pH=4

13CrpH=3.5

13Cr_pH=4.0 13Cr_pH=5.25

Super13Cr_pH=3.5

Super 13Cr_pH=5.25

Super 13Cr_pH=6.0

Super13Cr_pH= 4.0

Figure A- 17. Comparison of the cumulative thickness loss of 13Cr and Super

13Cr at 76°F, 22Cr at pH 4 is also included. Integration of the

current raw data.

192

0.0E+00

2.0E-06

4.0E-06

6.0E-06

8.0E-06

1.0E-05

1.2E-05

1.4E-05

0 50 100 150 200Time (sec)

Th

ickn

ess

Los

s(m

m)

T=150oF

22Cr_pH=4

13CrpH=3.5

13Cr_pH=4.0 13Cr_pH=5.25

Super 13Cr_3.5≤pH≤6.0

13Cr_pH=6.0

Figure A- 18. Comparison of the cumulative thickness loss of 13Cr and Super

13Cr at 150°F, 22Cr at pH 4 is also included. Integration of the

current raw data.

0.0E+00

4.0E-06

8.0E-06

1.2E-05

1.6E-05

2.0E-05

0 50 100 150 200

Time (sec)

Thic

kne

ss

Loss

(mm

)

13Cr_pH=3.5 13Cr_pH=4.00

13Cr_pH=5.25 & 6.0

T=200oF

22Cr_pH=4

Super13Cr_pH=3.5 & 4

Super 13Cr_pH=5.25

Super 13Cr_pH=6

Figure A- 19. Comparison of the cumulative thickness loss of 13Cr and Super

13Cr at 200°F, 22Cr at pH 4 is also included. Integration of the

current raw data.

193

0

20

40

60

80

0 50 100 150 200Time (sec)

I( m

A)

pH = 4.013Cr_200oF

13Cr_150oF

13Cr_76oF

22Cr,200, 150 & 76 oF

Super 13Cr_ 76oF

Super 13Cr_200oF

Super 13Cr_ 150oF

Figure A- 20. Comparison of the current decays for 13Cr, Super 13Cr and 22Cr

alloys at pH 4.0 and three different temperatures.

0

20

40

60

80

100

0 50 100 150 200Time (sec)

I( m

A)

pH= 6.0

13Cr, T =200oF

13Cr, T =150oF

Super 13Cr, T =200oF

Super 13Cr, T =150oF

Super 13Cr, T =76oF

Figure A- 21. Comparison of the current decays for both 13Cr and Super 13Cr

alloys at pH 6.0 and three different temperatures.

194

pH=4.0

0.0E+00

5.0E-06

1.0E-05

1.5E-05

2.0E-05

0 50 100 150 200Time (sec)

Th

ickn

ess

Lo

ss(m

m)

13Cr_76oF

13Cr_150oF13Cr_200oF

22Cr_200,

150 & 76oF

Super 13Cr_200oF

Super 13Cr_150oFSuper 13Cr_76oF

Figure A- 22. Comparison of the cumulative thickness loss of 13Cr, Super 13Cr

and 22Cr at pH 4. Integration of the current raw data.

0.0E+00

2.5E-06

5.0E-06

7.5E-06

0 50 100 150 200Time (sec)

Th

ick

nes

sL

oss

(mm

)

pH=6.0 13Cr_150oF13Cr_200oF

Super 13Cr_200oF

Super 13Cr_150oF

Super 13Cr_76oF

Figure A- 23. Comparison of the cumulative thickness loss of 13Cr and Super 13Cr

at pH 6. Integration of the current raw data.

195

0.E+00

1.E+05

2.E+05

3.E+05

4.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

)

pH= 4.0

22Cr,T=200oF

13Cr,T=200oF

13Cr,

T=150oF

13Cr,T=76oF

22Cr,T=76oF

22Cr,

T=150oF

Super 13Cr,

T=76oF

Super 13Cr,

T=150oF

Super 13Cr,T=200oF

Figure A- 24. Comparison of slopes at different temperatures at pH=4.0 for the

three 13Cr, Super 13Cr and 22Cr.

0.E+00

1.E+05

2.E+05

3.E+05

4.E+05

0 50 100 150 200Time (sec)

1/I,

(1/A

)

pH= 6.0

13 Cr,200oF

13 Cr,150oF

Super 13Cr,76oF

Super 13Cr,150oF

Super 13Cr,200oF

Figure A- 25. Comparison of slopes at different temperatures at pH=6.0 for both

13Cr and Super 13Cr.

196

APPENDIX B

DETERMINATION OF THE “C” COSTANT

This appendix contents the mathematical analysis performed to obtain the

constant “C” used in the conversion of current to thickness loss per unit time for the three

CRAs.

From Faraday’s Law31 penetration rates can be obtained from equation (B-1)

provided that the corrosion current Icorr is known.

Penetration Raten

aiK (B-1)

Where, i = corrosion current density Icorr/exposed Area in A/cm2, = material

density in g/cm3, a = atomic weight, n = number of electrons lost (valence change). a/ n

is also known as Equivalent Weight, Weq and K is a conversion factor which changes

depending on the units required. K = 0.129 for mpy, 3.27 for m/y, 0.00327 for mm/y.

If the measured current in scratch test Imeasured is assumed to be Icorr, the constant

“C” used in equations (5-3), (5-5) and (5-6) in Chapter 5 is given by

A

WKC eq (B-2)

197

The area exposed is the scratched area of the 1/8” diameter rod of the working

electrode was given by:

2

2

2

079.04

54.2*81

4cm

DA

Equation (B-2) used for parameters with consistent units gives the C constant for

three CRAs as listed in Table B-1.

Table B- 1. C constants for the three CRAs.

Material Weq CRA (g/cm3) C (mm/A-y) C (mm/A-sec)

13Cr 25.9 7.65 140 x 103 4.439 x10-3

Super 13Cr 26.1 7.65 141 x 103 4.471x10-3

22Cr 24.9 7.80 132 x 103 4.186 x 10-3

The equivalent weight for an alloy can be estimated as the inverse of the

equivalent number given by equations (B-3) and (B-4) respectively31.

eqeq N

W1

(B-3)

i

iieq a

nfN (B-4)

Where, a = atomic weight, n = number of electrons lost (valence change), fi

represents the mass fraction of the alloying element contents in the alloy.

31 D. A. Jones, “Principle and Prevention of Corrosion” Prentince Hall, 2nd ed., pp.343

USA, 1996.

198

APPENDIX C

EROSION-CORROSION OF 13Cr EXPOSED TO VERY SEVERE

EROSIVITY CONDITIONS

Black coating development on 13Cr

An interesting phenomenon was observed when longer periods of time were

used to test 13Cr alloy at severe conditions such a direct impingement, low pH, high

flow velocities, and high sand rates. At such aggressive conditions, high current

magnitudes were obtained as seen in Figure C-1. The high current values were

sustained even after removing the sand, therefore the second order repassivation did

not result, and a black coating grew on the surface of the 13Cr target.

0

25

50

75

100

125

150

175

200

22 22.5 23 23.5 24 24.5 25 25.5 26

Time (hours)

I(m

A)

13Cr, pH 4, 150oF20 ft/s, 200 kg/day

Flow

ImpingementDirect

Figure C-1. Current response for 13Cr exposed to very severe erosivityconditions.

199

An effort to characterize the black coating was performed by means of

scanning electronic microscope (SEM), energy dispersive spectroscopy (EDS) and X-

ray diffraction (XRD). Figure C-2a shows the black coating grown on the metallic

surface of the 13Cr target specimen exposed to very severe erosion-corrosion

conditions (pH = 4.3, 150oF, 20ft/s, 200 Kg sand/day, more than 48 hours exposure).

a) 13Cr, pH4.3 long sand exposure b) SEM image of developed blackcoating 1000X

c) EDS of developed black coatingd) SEM image of developed black

coating 200X

Figure C-2. Black coating on metallic surface of the 13Cr target specimenexposed to erosion-corrosion conditions (pH = 4.3, 150oF, 20ft/s,200 Kg sand/day, more than 48 hours exposure).

The black coating seems to be scaly with a rough and cracked texture and

evenly distributed over the complete surface as seen on Figure C-2b and Figure C-2d.

200

The stoichiometry of the compound or compounds present in the black coating could

not be determined with the techniques used in this investigation, however the EDS

shown in Figure C-2c suggests it to be either a mixture of iron oxide and chromium

oxide or a Fe-Cr binary complex oxide. There are always chances for the corrosion

products to be hydrates types of compounds while they are still exposed to the wet

corrosive conditions, and then become de-hydrates oxides compounds showing

cracked texture due to shrinking during the dehydration process, similar to that

observed on Figure C-2b. Of course cracking could come from the continued particle

impingement on the brittle oxide layer.

Figure C-3a shows the metallic surface of the 13Cr target specimen exposed

to very severe erosion-corrosion conditions but for a short sand exposure time (pH =

4.3, 150oF, 20ft/s, 200 Kg sand/day, less than 10 minutes exposure). A pattern with

alternate rings was developed on the surface of the 13Cr. The pattern suggests an

incipient stage of the back coating development. The pattern suggests that the

erosion-corrosion damage is not uniform. Those regions were the erosion-corrosion

damage is greater the black coating grows earlier as a direct consequence of the

severity of the chemi-mechanical degradation process.

Figure C-3b shows a SEM image for the inner dark ring shown in Figure C-

3a. Numerous incipient pits with irregular shapes and a variety of sizes (up to 20

micrometers) can be seen along with some scratch marks from the polish preparation

process. Figure C-3c shows the EDS spectrum for the incipient black coating with the

same elements seen for the developed black coating. However, the intensity of the

201

peaks is different, indicating enrichment of Cr and O in the completely developed

black coating as compare with its incipient stage.

a) 13Cr, pH4.3 short sandexposure b) SEM image of incipient black

coating 1000X

c) EDS of incipient black coating

Figure C-3. Black coating patterns on metallic surface of the 13Cr targetspecimen exposed to erosion-corrosion conditions (pH = 4.3, 150oF,20ft/s, 200 Kg sand/day, less than 10 minutes exposure).

For erosivity conditions similar to those for which the black coating was able

to develop, but at the higher pH of 6, the black coating was inhibited and the passive

film reformed preventing high erosion-corrosion rates for 13Cr. Figure C-4 shows the

resultant metallic surface of 13Cr exposed to severe erosion-corrosion conditions and

pH = 6. The surface is dull but the black coating is not seen. The SEM image seen in

202

Figure C-4b shows a smooth surface with several tiny black spots (1-2 micrometers).

A precise determination of the nature of the black spots was not possible, but they are

thought to be indentations from the particle impingements. Figure C-4c shows the

EDS spectrum indicating the presence of the same elements seen in the black coating

but with different proportions (less enrichment in Cr and O is seen for the passive

film formed at pH = 6).

a)13Cr, pH 6 long sandexposure b) SEM image 13Cr passive film

(pH =6) 1000X

c) EDX 13Cr passive film (pH = 6)

Figure C-4. Resultant metallic surface of the 13Cr target specimen exposed toerosion-corrosion conditions (pH = 6, 150oF, 20ft/s, 200 Kgsand/day, more than 48 hours exposure).

203

Super 13Cr was also tested at erosivity conditions similar to those for which

the black coating was able to develop for 13Cr. Differences between 13Cr and Super

13Cr at these conditions were dramatic. The Super 13Cr showed lower current

magnitudes and second order behavior on repassivation with quick reformation of

passive film at pH of 4.3, resulting in significantly lower erosion-corrosion damage as

compared to 13Cr. Figure C-5 shows the resultant metallic surface of Super 13Cr

exposed to severe erosion-corrosion conditions and pH = 4.3. The surface is very

polished and bright as a mirror surface. The SEM image seen in Figure C-5b shows a

very smooth surface with several tiny darker spots (1-2 micrometers), thought to be

indentations from the particle impingements. Figure C-5c shows the EDS spectrum

indicating the presence of the same elements seen in the passive film of 13Cr but with

different proportions, and in addition, Nickel and Molybdenum peaks are present.

So far differences between the resultant surfaces for 13Cr and Super 13Cr

exposed to different conditions have been described, but the stoichiometric chemical

composition was not determined.

Furthermore, the black coating on the 13Cr was found to be rich in Fe and Cr,

similar to that found in the passive film of the same alloy. In fact no significant

distinction was seen between the spectrums obtained from the X-ray diffraction

analysis performed over the surfaces of the black coating and passive film of a 13Cr

as seen in Figure C- 6.

204

a)Super 13Cr, pH 4.3 long sandexposure b) SEM image Super 13Cr passive film

(pH =4.3) 1000X

c) EDX Super 13Cr passive film (pH = 4.3)

Figure C-5. Resultant metallic surface of the Super 13Cr target specimenexposed to erosion-corrosion conditions (pH = 4.3, 150oF, 20ft/s,200 Kg sand/day, more than 48 hours exposure).

However differences between the black coated 13Cr surface and the air-

passivated 13Cr surface in color, texture and appearance as observed without sight

aid devices as well as through SEM images clearly suggest significant distinctions

205

between the two surfaces. Furthermore, significant differences in the magnitudes of

the current obtained when the black coating is developed as opposed to when the

passive film is reformed, also suggest that the protective properties of the black

coating and the passive film are dramatically different. Thus, a carefully study to

better characterize the link between the erosion-corrosion response of 13Cr with its

surface finish is advised.

Figure C- 6. X-ray diffraction analysis on the black coating grown on a 13Crexposed to erosion corrosion-conditions and a on a air-passive13Cr surface.

Researchers 1,2 have successfully used X-ray Photoelectron Spectroscopy

(XPS) to characterize the passive films of CRAs. Results from this technique are

precise enough to determine the presence of distinguishable multilayers of oxides and

206

hydroxides with thicknesses from 1 to 5 nanometers depending upon the conditions at

which the passive films were formed. Olefjord and Wegrelius1 studied the passive

film of a stainless steel (20Cr18Ni6Mo0.2N) during polarization in 0.1 M HCl + 0.4

M NaCl. The film had two layers, an outer monolayer of Cr(OH)3 and a thicker inner

layer of Fe-Cr oxide.−Mo6+ and Mo4+ were also found enriched in the outer layer,

and their concentration depended on the potential. Ni2+ was also found in the passive

film but at very low concentration.

Vayer et. al. studied the effect of an HNO3 passivation treatment on the

characteristics of the passive film as compared to an air-passivated martensitic

stainless steel (X20Cr13, X30Cr13, X40Cr14). The HNO3 passivated film consisted

of a 3 nanometers thick layer with enhanced pitting resistance, and with oxidized

chromium and iron uniformly distributed, but enriched with oxidized chromium

(about 50% of the metallic elements of the passive layer was Cr). The air-formed

passive film consisted of a 4-5 nanometer double layer oxidized iron was

concentrated at the surface outer layer, and oxidized chromium was concentrated at

the inner layer closer to the metal matrix.

In summary, preliminary characterization of the black coating obtained for

13Cr exposed to certain very severe erosion-corrosion conditions was performed by

means of several techniques and it is known to be rich in Fe and Cr, suggesting it is a

combined Fe-Cr oxide. However, the exact chemical composition and crystalline

structure has not yet been clearly determined, and more research needs to be done in

this matter (XPS technique is advised to be used thoroughly for this purpose).

207

Potentiodynamic Scans for 13Cr and Super 13Cr Exposed to Corrosion

and Erosion-Corrosion Conditions.

A few potentiodynamic scans directed towards the understanding of

differences between the erosion-corrosion behavior of 13Cr and Super 13Cr were

performed. The liquid/sand flow loop (Microloop) was not provided with a

commercial reference electrode, therefore a rigorous analysis of the potential

parameters such as primary passivation potential Epp and corrosion potential Ecorr was

not performed. However, interesting results regarding the critical current density, icrit

and passive current density ip were obtained. Potentiodynamic scans are shown as a

function of the overpotential.

Figure C-7 shows potentiodynamic scans for 13Cr and Super 13Cr exposed to

brine flowing at 17 ft/s with a temperature of 150oF and pH = 4.3. Both alloys show a

broad passive range with low current. The larger passive range and lower passive

current density displayed by Super 13 Cr once again corroborate the better corrosion

resistance of this alloy as compare to 13Cr.

Figure C-8 shows potentiodynamic scans for Super 13Cr exposed to sand-free

brine and brine containing sand flowing at 17 ft/s with a temperature of 150oF and pH

= 4.3. Curves are similar at these conditions. Both shown a similar passive range,

however the passive current obtained with sand is about one half order of magnitude

larger than the curve obtained for sand free conditions.

208

Figure C-9 shows potentiodynamic scans for 13Cr exposed to sand-free brine

and brine containing sand flowing at 17 ft/s with a temperature of 150oF and pH =

4.3. The curves are greatly different at these conditions −for the curve obtained with

sand, the passive range shrunk, and the passive current is about one order of

magnitude larger than the curve obtained for sand free conditions. In addition the

steady state corrosion potential Ecorr seems to become more negative, so the critical

current icrit and primary passivation potential Epp appeared in the scan plot suggesting

that requirements to achieved passivation are more demanding under sand flows at

tested conditions (see Figure 2-3 and Figure 2-4 in Chapter 2).

For comparison purposes Figure C-10 shows the potentiodynamic scans for

Super 13Cr and 13Cr exposed to a brine containing sand shown in Figure C-8 and

Figure C-9 respectively. Again the larger passive range and lower passive current

density shown by Super 13Cr exposed to CO2 saturated brine containing sand also

corroborate the better erosion-corrosion resistance of this alloy as compare to 13Cr.

209

-100

0

100

200

300

400

500

600

1.E-09 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01

Current Density, i (A/cm2)

Ov

erp

ote

nti

al

(mV

)

Super 13Cr Alloy 13Cr Alloy

No Sand, T = 150 oF, pH = 4.3, Vliq = 17 ft/s

Figure C-7. Potentiodynamic scans for 13Cr and Super 13Cr exposed to brineflow at 17 ft/s, 150oF and pH=4.3.

-100

0

100

200

300

400

500

600

1.E-09 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01

Current Density, i (A/cm 2)

Ove

rpo

ten

tia

l(m

V)

No_Sand

Super 13Cr Alloy, T = 150 oF, pH = 4.3, Vliq = 17 ft/s

60 Kg/day of sand

Figure C-8. Potentiodynamic scans for Super 13Cr exposed to brine flow withand without sand at 17 ft/s, 150oF and pH=4.3.

210

-100

0

100

200

300

400

500

600

1.E-09 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01

Current Density, i (A/cm2)

Ove

rpo

ten

tia

l(m

V)

No_Sand

13Cr Alloy, T = 150 oF, pH = 4.3, Vliq = 17 ft/s

60Kg/day of sand

i cri

E pp

Figure C-9. Potentiodynamic scans for 13Cr exposed to brine flow with andwithout sand at 17 ft/s, 150oF and pH=4.3.

-100

0

100

200

300

400

500

600

1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01

Current Density, i (A/cm2)

Ov

erp

ote

nti

al(

mV

)

60Kg/day of Sand, T = 150oF, pH = 4.3, Vliq = 17 ft/s

Super 13Cr Alloy13Cr Alloy

Figure C-10. Potentiodynamic scans for 13Cr and Super 13Cr exposed to brineflow with and without sand at 17 ft/s, 150oF and pH=4.3.

211

In summary potentiodynamic scans showed an anodic shift in the primary

passivation potential Epp along with the increment in the passive current density ipass,

for 13Cr exposed to the combined effects of high erosivity flows and severe

environmental conditions for the direct impingement testing. For these conditions

13Cr forms a black coating and repassivation is prevented. A more systematic

investigation into the conditions under which 13Cr will developed the black coating is

recommended along with more accurate characterizations of the chemical

composition and crystalline structure of the components of the black layer.

Potentiodynamics scans conducted for less severe flow conditions in a 13Cr

elbow flow geometry for similar environmental conditions showed a similar

increment in the passive current density ipass, but, the anodic shift in the primary

passivation potential Epp never occurred at these conditions3. For there conditions, no

black coating was observed and repassivation was accomplished. This suggests that

the requirements for repassivation might depend on the severity of flow conditions

and the erosivity level in addition to environmental conditions such pH, temperature

and chloride contents and water chemistry.

1 I. Olefjord and L. Wegrelius, Corrosion Science, Vol 31, pp. 89-98, 1990.

2 M. Vayer, I. Reynaud, and R. Erre, Journal of Material Science, Vol.35, pp.2581-2587, 2000.

3 Rincon, H.E. (2001). Erosion-Corrosion Phenomena of 13Cr Alloy in FlowsContaining Sand Particles, M.S. Thesis, Department of MechanicalEngineering, The University of Tulsa, Ok, Tulsa.

212

APPENDIX D

COMPUTATION FLUID DYNAMICS PROCEDURES

General Procedure

Erosion–Corrosion of metals by corrosive fluids carrying solid particles is a very

complex phenomenon present in several fields and industries. Research in this field is

very challenging due to the large amount of parameters affecting the degradation

mechanism.

During this research, an experimental approach has been adopted for the study of

many of these variables such as temperature, pH, material composition, material

properties and sand rates. Also, several flow velocities, flow patterns and flow geometries

have been experimentally tested to characterize the effect of such variables on the

erosion-corrosion damage of CRAs. However, the hydrodynamics of the conditions

tested are complex and additional tools are needed to better understand the degradation

mechanism. In addition, experiments to emulate field conditions are very difficult to

perform, expensive and time consuming. Thus, overall lack of predictability is

characteristic of these kinds of studies.

The contributions of computational fluid dynamics (CFD) to more closely

simulate the hydrodynamics of field conditions and to reduce data are encouraging.

213

Significant efforts have been directed towards the use of CFD as a potential tool to

evaluate and predict solid particle erosion. Some researchers have also included the

prediction of erosion-corrosion. The present work does not attempt to develop any new

CFD procedure. Here CFD is used as a tool to predict the hydrodynamics, particle

tracking, and erosion of a material exposed to a flowing slurry system since, erosion is a

necessary component of erosion-corrosion.

Several authors1-6 have followed the same general procedure to obtain the

numerical evaluation of erosion. The procedure can be summarized in three major steps:

1) generation and convergence of flow field domain simulation, 2) computation of the

stream particle trajectories inside the flow field, and 3) computation of the erosion rate by

applying the suitable erosion model equations. According to Bozzini et. al.,Error! Bookmark

not defined. two additional steps are required if the corrosion component needs to be

included. The additional steps are: 4) evaluation of the impact-induced corrosion effect

and 5) evaluation of the synergistic erosion-corrosion damage.

In this research, a semi-mechanistic model to predict erosion-corrosion of CRAs

has been developed and implemented based on the solid particle impingement

information collected from CFD simulations. The estimation of the impact-induced

corrosion and the synergistic effect of the erosion-corrosion damage have been included

in the model. Only the prediction of erosion rates is performed during the CFD

simulations, the impact-induced corrosion and erosion corrosion rates are not actually

executed during the CFD simulations but in a separate VBA (Visual Basic for

Applications) code built for that specific purpose.

214

This appendix briefly summarizes the complete CFD procedures used to predict

erosion rates and produce the solid particle impingement information needed as an input

for the erosion-corrosion prediction model. The impingement information includes, (x, y,

z) coordinates of the cell center (m), Cell Area (m2), total impact number in the cell,

averaged impact speed in a cell (m/s), averaged impact angle in a cell (degree), and

erosion (kg removed metal/m2.s).

Numerical Model Approach

CFD experts agree on the convenience of combining the Eularian CFD modeling

to predict the flow field on the continuous fluid phase with the Lagrangian particle

tracking of the solid or particulate phase2-5. Eularian modeling of the discrete phase is

also possible, however, the particle-wall interactions have been shown to be difficult to

manage.Error! Bookmark not defined.

The model presented here is Eularian-Lagrangian and can be broken down to the

following main steps: build and mesh geometry, fluid flow model, particle tracking

model, and erosion model.

CFD Geometry and Mesh

The geometry of the CFD model was built simulating features and dimensions of

the direct impingement test cell of the single phase liquid flow test whose schematic is

shown in chapter 4. Figure D- 1.a shows the CFD geometry as built with Gambit 2.1.6

software indicating the place of the jet and the target specimen.

215

A grid refinement study was performed to guarantee grid-size-independent

results. Several mesh designs were examined for good convergence of the solution and

reasonable computational time that did not compromise result quality.

Figure D- 1.b shows the exterior mesh grid use for the volume representing the

cross fitting. The biggest volume element size used for the simulation was 0.05% of the

total volume. For the interior of the volume, a size function along an imaginary

centerline, going from the center of the jet to the center of the target, was used to mesh

the entire volume. The size of the volume elements grew from a starting size of 0.005%

for those elements next to the center line, to a maximum size of 0.05% at a rate of 1.1%,

as shown in Figure D- 1.c. Figure D- 1.d shows the 2D elements for the face of the target

specimen. This face has a total of 684 computational cells with areas varying from 8.28E-

09 to 7.68E-07 m2, for a total area of 2.0E-4 m2 or 2 cm2 which matches the actual

impinged surface area of the specimen.

Flow Modeling

Flow simulations were accomplished by means of a commercially available

software (FLUENT version 6.1). First, flow fields were determined by using the kinetic

energy-dissipation rate (-) two-equation model of turbulence. This model is usually a

good starting point since it has been extensively used for several geometries with

reasonably good estimates of the flow field and relatively low computational time (easy

convergence). However, places where recirculation zones are expected may not be

completely resolved with this turbulence model. Hence, the Reynolds Stress Model

(RSM) was also used. Flow simulations obtained with both turbulence models were

216

similar, and as expected, the RSM showed more difficulties to obtain convergence, but a

slightly more resolved flow field. Hence the RSM was used for the rest of the

simulations.

a) b)

c) d)

Figure D- 1. CFD geometry and grid features for the direct impingement jet

simulation.

Flowdirection

ImpingementJet

TargetSpecimen

Flowdirection

ImpingementJet

TargetSpecimen

217

The importance of minimizing the truncation errors in the flow simulation was

also considered. Therefore, a second order upwind scheme was used to solve momentum

equations, turbulent kinetic energy and turbulent dissipation rate equations, and Reynolds

stresses equations. Also, the semi-implicit method for pressure-linked equations

(SIMPLE) proposed by Patankar and Spalding to solve the pressure term was used 7.

Typical contour plots for flow velocity and turbulence intensity of a simulation

conducted at 15 ft/s are shown in Figure D- 2. The typical flow structure of this type of

geometry such as the disruption in the velocity profile caused by the target in front of the

flow jet, were well simulated by the numerical simulation.

a) b)

Figure D- 2. Flow velocity and turbulence intensity contours plot obtained using

RSM 2nd Order, for water at 150oF and outlet jet velocity of 15 ft/s

(4.572 m/s)

Figure D- 2a clearly shows how the velocity gradually reduces as the jet

approaches the target surface. Furthermore, the velocity profile close to the surface

Flow Velocity, m/s Turbulence Intensity, %

218

suggests there is a low flow velocity zone at the center region of the target which has

been commonly referred to as a “stagnation zone”. Surrounding this zone, flow velocities

are still high and streamlines are radially diverted away form the center line, causing

reverse flows and secondary cross-flow vortices around the projected high flow velocity

jet.

Validation of the flow simulation by comparison with actual flow velocity profile

data was not explicitly done during this research. However, Yongli8 recently has

successfully compared simulations performed with the same numerical solution approach

to experimental data collected with a Laser Doppler Velocimetry (LDV) It was found that

the numerical simulation reasonably matched the experimental data.

Particle Tracking

As mentioned before, a Lagrangian approach was taken to conduct the particle

tracking. The equation of dispersed particle motion is essentially a force balance around a

single particle. FLUENT also has the capability to couple the particle equation of motion

with the flow solution. However, at low particle concentrations the particles do not affect

the flow and, thus, coupling is not necessary.

The particle motion comes from Newton’s second law of motion is expressed as:

dt

dVmF p

p D-(1)

where Vp is the particle velocity vector, mp, is the particle mass and F is the resultant

force vector on the particle which may include drag force, FD, exerted by the fluid on the

particle, buoyancy force, FB, due to differences in fluid–particle densities, pressure

219

gradient force, Fp, exerted by pressure gradients on the continuous phase, the added mass

force, FA, which accounts for the inertia of the fluid surrounding the particle (also named

as virtual mass force), the Saffman lift force and the rotating coordinates force, FR, which

accounts for rotational motion and centrifugal effects may contribute significantly to the

particle motion and were included in the particle simulation. Further details of the forces

included in the particle tracking, their mathematical representations and the physics

behind them have been extensively discussed in literature and a good summary may be

found in the user guide of FLUENT 7.

The effect of turbulence on the particle dispersion was also considered based on

the “random eddy lifetime” model to predict the turbulent dispersion of particles as

interacting with a series of fluid phase turbulent eddies. Also Grant and Tabakoff 9

correlations for the restitution coefficients required to predict particle velocity after

impact with a solid surface were included in the simulation.

Erosion Model

Most of currently available models predict erosion based on a direct impingement

mechanism. Many also include the effects of turbulent fluctuations and the random

impingement mechanisms10. In any case, results presented here are not expected to be

significantly affected by this, since erosion predictions were performed for direct

impingement flow geometry. In the direct impingement mechanism, the particles are

driven to the walls primarily by the momentum of the particle provided by mean flow

velocity. Where viscous effects are small or when the particles are large and dense as

compared with the carrier fluid density, direct impingements can be the dominant

220

mechanism. For denser liquids carrying sand, velocity fluctuations due to turbulence may

effect particle motion. Turbulent eddies can transfer radial momentum to the particles

near the wall and cause random impingements resulting in erosion damage; but, this

effect is expected to be small for direct impingement simulations.

If the impingement impact speed and angle are known, the erosion rate can be

computed. Erosion rates may be expressed in different units; most commons used units

are: penetration rate, which is the amount of wall thickness loss per unit time or amount

of mass loss per unit time. Other quantities of interest are erosion ratio, defined as the

amount of mass lost by the pipe wall due to erosion divided by the total mass of particles

impinging the metal.

Some of the erosion models available in the literature provide equations can be

broken down into 3 factor functions 1-3, 8-12 The first function accounts for the effect of

flow velocities and is usually the impact velocity raised to a power, with the power being

from 1.7 to 2.4. The second term refers to the effect of impact angle; this term usually

comes from fitted data obtained from direct impingement erosion tests conducted at

different impact angles. The last term usually accounts for properties of the erodent,

target material or even flow geometries.

Edwards et alError! Bookmark not defined. used relations provided by Ahlert 11 for

carbon steels, developed at The University of Tulsa. These relations depend primarily on

the particle impact speed and impact angle. As seen in the following equation:

fVAFER ns (D-2)

221

where Fs is a particle shape coefficient (Fs=1.0 for sharp (angular), 0.53 for semi-

rounded, or 0.2 for fully rounded sand particles). Ahlert11 used two functional forms of

the angle dependence, with matching conditions applied at some angle . The

dependence on impingement angle, f (), is given in Eqs. (D-3) and (D-4).

baf 2 for (D-3)

zyxf 22 sinsincos for (D-4)

A, a, b, , x, y, z, and n, are all empirical constants.

This model was developed based on direct impingement data obtained from gas-

sand flows testing, and tends to over-predict the estimated penetration rates. A more

recent work published by Oka et al.12 in 2005 uses correlative equations derived from

particle impact energy and indentation behavior as a function of material hardness to

generalize prediction of metal loss regardless of the target material. This model was also

first developed for gas-sand flows. However, it seems to grasp more of the physics

involved in the erosion process and has been successfully used by Yongli et al.8 to

estimate direct impingement erosion rates in liquids by means of CFD.

The Oka model12 assumes that erosion is mainly driven by the impact velocity,

impact angle, size and type of particles and material hardness. The general equation to

estimate material loss is given by equation (D-4):

90)()( EgE (D-4)

222

where, E() represents the removed material volume per mass of particle

(mm3/kg) at any angle ; E90 is the erosion damage at 90 degree angle and g() denotes

the function ratio of erosion ratio at any given angle E() to that at normal angle E90.

The angle function depends on the material hardness (Vickers Hardness, Hv in

Gpa) and accounts for what the authors12 referred to as a “repeated plastic deformation”

linked to the normal component of the impact velocity (first term in Equation D-5). A

“cutting action” proved to be more effective at low impact angles (second term in

equation D-5).

21 ))sin1(1()(sin)( nn Hvg (D-5)

where n1 and n2 are similar functions with constants determined experimentally which

accounts for material hardness and erodent properties. Functions n1 and n2 are given by

equation (D-6)

qHvSnn )(, 21 (D-6)

For sand particle Sn1= 0.71, qn1 = 0.14, Sn2= 2.4 and qn2 = -0.94.

The erosion damage at normal angle E90 has been discussed in detail by Oka and

Yoshida,13 and accounts for impact velocity, V(m/s), particle diameter, D(m), and

material hardness according to equation B-7 as follows.

32

1)(90

kkk

DD

VV

HvKE

(B-7)

where K, k1, k2 and k3 are constants determined experimentally. K and k1 depend on

particles properties, k2 depends on particle properties as well as on material hardness. For

223

sand K = 65, k1 = -0.12, k2 = 2.3(Hv)0.0038 and k3 = 0.19 regardless of the erodent used. V’

(104 m/s) and D’ (326 m) are reference velocity and particle diameter respectively used

in the experiments to obtain the correlations of the erosion damage.

In the numerical calculation of the metal mass loss it is assumed that when a

particle strikes the wall, the mass loss is distributed uniformly over the computational cell

in which the particle impinged. With both sufficiently small grid spacing and a large

number of particles simulated, approximation errors induced by this assumption can be

kept at a minimum.

Once all particle trajectories have been computed and all wall impingement data

gathered, the mass loss for all impingements can be compiled to generate a local

penetration rate for each cell that lies on the surface of the geometry. At grid cell j on the

surface of the geometry, the local Erosion penetration rate, Ecell(j), is given by:

i

ji

p

p

pj

pjcell E

NN

A

dE )(

)()(

3)(

6

(D-8)

where A(j)is the surface area of the computational cell j, E()i(j) denotes the erosion ratio of

impingement i at cell location j,

pN is the number of particles per second flowing, Np is

the total number of particles simulated, and p is the particle density.

CFD simulation sample cases

Table D- 1 shows the combination of flow velocities and sand rates used in the

CFD simulations to generate the solid particle information and erosion rates used as input

in the erosion-corrosion model proposed in this research. Flow velocities and sand rate

224

were selected to approximately match the experimental conditions used in the single

phase liquid flow loop.

Table D- 1 Flow velocity and sand rate combinations for simulations performed

in CFD for 13Cr and Super 13Cr materials.

Velocity(ft/s)

Velocity(m/s)

SandRate

(kg/day)10 3.05 1010 3.05 2510 3.05 10015 4.57 5515 4.57 16515 4.57 23517 5.18 6017 5.18 13220 6.10 6020 6.10 7520 6.10 10820 6.10 20020 6.10 400

As mentioned earlier, the Oka erosion model was based on gas-sand experimental

data tested at much higher flow velocities with higher erosion rates. Hence, some over-

prediction was expected from these simulations as compared to the experimental data

collected for liquid-sand direct impingement. However, impact velocities, impact angles,

and erosion patterns obtained in simulations are reasonable. The erosion pattern, impact

velocity profile and impact angle profile were similar for all simulations regardless of

sand rate and flow velocity, and typical simulation results are shown in Figure D- 3.

225

a) Impact Velocity (m/s) b) Impact Angle (degree)

c) Impact Hits (#/s) d) Erosion (mpy) Avg = 9.9 mpy

Figure D- 3. Impact velocity, impact angle and number of impacts and erosion rate

distributions for CFD simulation performed for 13Cr at 10 ft/s (3.048

m/s) and sand rate of 25 kg/day.

226

1 Edwards J., McLaury B., Shirazi S., “Evaluation of Alternative Pipe Bend Fittings

in Erosive Service”, Proceedings of Fluid Engineering Summer Meeting ASME

2000, June 11-15-2000, Boston, Massachusetts.

2 Edwards J., McLaury B., Shirazi S., “Modeling Solid Particle Erosion in Elbows

and Plugged Tee”, Journal of Energy Resources Technology, ASME, Vol 123,

December 2001.

3 Wallace M., Peters J., “CFD-Based Erosion Modelling of Multi-Orifice Choke

Valves”., Proceedings of Fluid Engineering Summer Meeting ASME 2000, June

11-15-2000, Boston, Massachusetts.

4 Keating, A. and S. Nesic (2000). "Particle Tracking and Erosion Prediction in

Three-Dimensional Bends", Proceedings of Fluid Engineering Summer Meeting

ASME 2000, June 11-15-2000, Boston, Massachusetts.

5 Bozzini, B., M. E. Ricotti, et al. (2003). "Evaluation of erosion-corrosion in

multiphase flow via CFD and experimental analysis." Wear 255(1-6): 237-245.

6 Keating, A. and S. Nesic (2001). "Numerical prediction of erosion-corrosion in

bends." Corrosion 57(7): 621-633.

7 User’s Guide,Fluent 6.1 Documntation.

227

8 Yongli Zhang, Erik Reuterfors, Brenton S. McLaury, Siamack A. Shirazi, and

Edmund F. Rybicki. “Experimental and CFD results for particle velocities and

erosion.” Submitted to Wear journal

9 Grant G. and Tabakoff W., “Erosion Prediction in Turbomachinery Resulting from

Environmental Solid Particles,” J. Aircraft, Vol. 12, No. 5, pp. 471-478

10 Edwards, J., Development, Validation, and Application of a Three-Dimensional,

CFD-Based Erosion Prediction Procedure, PhD Dissertation, Department of

Mechanical Engineering, The University of Tulsa, 2000.

11 Alhert, K., Effects of Particle impingement Angle and Surface Wetting on Solid

Particle Erosion of AISI 1018 Steel, M.S. Thesis, Department of Mechanical

Engineering, The University of Tulsa, 1994.

12 Oka, Y.I, Okamura, K., Yoshida, T., “Practical estimation of erosion damage

caused by solid particle impact. Part 1: Effects of impact parameters on a predictive

equation.” Wear 259 (95-101)

13 Oka, Y.I, Yoshida, T., “Practical estimation of erosion damage caused by solid

particle impact. Part 2: Mechanical properties of materials directly associated with

erosion damage.” Wear 259 (102-109)

228

APPENDIX E

DETERMINATION OF THE ESTIMATING FUNCTION OF THE

INDENTATION SURFACE

This appendix contents the mathematical analysis performed to obtain the

function used to represent the surface area created by a single particle impact. To obtain

equation (9-6) showed in Chapter 9, first the volume and surface area of a cap of a sphere

were obtained.

Volume of a sectioned sphere

Find the volume of a sphere with a slice cut off at rY (see Figure E- 1).

Figure E- 1. Schematic of a circle

229

The volume of the hemisphere above the X axis is ½ the volume of the sphere

33

32

34

21

RRVhemisphere

(E-1)

Below the X axis, the value of dV is the area of a disk 2X times the

infinitesimal height dY

dYXdV 2 (E-2)

From equation of circle 222 RYX

Hence the volume for the section from Y= –r to Y = 0 is given by.

03

020 22

3r

rr

YYRdYYRV

(E-3)

332 03

0 rrRV (E-4)

33

3232 rrRrrRV

(E-5)

The crater volume is the total volume of the sphere below rY or the volume

of the hemisphere minus equation (E-5) given by.

332

332 3

233

23 rrRR

rrRRVcrater

(E-6)

230

Surface of a sectioned sphere

To find the surface area of the crater consider schematic shown in Figure E- 2 .

Figure E- 2. Schematic of a sectioned sphere.

The infinitesimal area for the sphere is XdSdA 2 (E-7)

RddS , sinRX (E-8)

Substitute (E-8) into (E-7)

dRRdRdA sin2sin2 2 (E-9)

The range of Φis from 0 at (0,-R) to rX (E-10)

Rr1cos (E-11)

231

The integral is

Rr

Rr

RdRdAA1 1cos

0

cos

022 cos2sin2 (E-12)

0coscoscos2 12

Rr

RA (E-13)

RrR

RrRA 1212 22 (E-14)

Where r < R.

The questions are:

(1) What is the ratio of the volume below rY , Vcrater, to the sphere volume,

Vsphere?

(2) What is the ratio of the surface area of the sphere below rY , Acrater, to the total

surface volume of the sphere, Asphere?

Answer (1)

3

3

3

323

332

43

34

31

32

Rr

Rr

R

rrRR

VV

sphere

crater

(E-15)

Answer (2)

Rr

RRrR

REq

AA

sphere

crater 121

4

12

418

2

2

2

(E-16)

232

Area Ratio to Volume Ratio Relationship

To obtain equation (9-6) showed in Chapter 9, a regression was made with data

generated from equations (E-15 and E-16). Figure E- 3 shows the fitted curve for the

Area Ratio to Volume Ratio relationship. As can be seen the fitted curve does a very

good match with data generated from equations (E-15 and E-16), for a broad range of

volume ratios, meaning that this relationship can be used for a broad range of erosion

ratios. Typical erosion ratios for a single particle impingement of liquid flows containing

sand are within 10-9 and 10-12 while for gas flows containing sand erosion ratios may fall

within 10-8 to 10-6. Hence this relationship will cover the whole range of erosion ratios

with accuracy.

1.E-07

1.E-06

1.E-05

1.E-04

1.E-03

1.E-02

1.E-01

1.E+00

1.E-12 1.E-10 1.E-08 1.E-06 1.E-04 1.E-02 1.E+00

Vcrater/Vsphere

Acr

ater

/Asp

he

re

Acrater/Asphere= 0.6022 (Vcrater/Vsphere)^0.5026

Figure E- 3. Area ratio to volume ratio relationship.

233

APPENDIX F

VBA CODE FOR THE PROPOSED EROSION-COROSION MODEL

Sub Erosion-Corrosion Model()

'Estimation of the Total current, Corrosion Component Rates and Erosion-CorrosionRates PER EACH CFD NUMERICAL CELL of CRAs in CO2 Saturated BrineContaining Sand

Dim Time(100000, 1) As Double, Dim I(100000) As Double, DimI_high(100000) As Double, Dim I_low(100000) As Double, Dim I_Avg(100000) AsDouble, Dim I_cum(100000) As Double, Dim pi As Double, Dim Part_flow AsDouble, Dim Name_Input_File As String, Dim TotalCell As Integer, Dim FinalWBAs Workbook, Dim AvgInfoWS As Worksheet, Dim QuickCalcSheet As Worksheet

'Reading Input from Source Code WorkBook "QuickCalc Spreadsheet"

m_st = Worksheets("QuickCalc").Cells(2, 2) '1/A*sec

Io_st = Worksheets("QuickCalc").Cells(3, 2) 'microA

Stop_Current = Worksheets("QuickCalc").Cells(4, 2)

D_part = Worksheets("QuickCalc").Cells(2, 4) 'micrometer

f = Worksheets("QuickCalc").Cells(3, 4) 'Kg/m^3

rho_sand = 2650 'Kg/m^3

material = Worksheets("QuickCalc").Cells(1, 24)

Simulated_particles = Worksheets("QuickCalc").Cells(3, 6)

Sand_rate = Worksheets("QuickCalc").Cells(4, 6) 'Kg/day

I_vs_time_info = Worksheets("QuickCalc").Cells(1, 27)

Plots = Worksheets("QuickCalc").Cells(1, 30)

234

'Open Dialog Box to browse and read CFD-txt file with impact information

Application.Dialogs(xlDialogOpen).Show

Set FinalWB = Worksheets.Parent

x = 2

s = 1

y = 2

'Pre-calculations

'Calculate area of tip of WE1 in Scratch Test

pi = Atn(1#) * 4

R_st = (1 / 8) / 2

Stip_st = pi * (R_st * 0.0254) ^ 2 'm^2

'Material Properties

If material = 1 Then '13Cr Alloy

rho_metal = 7650 'Kg/m^3

EW = 25.9 ' Equivalent Weight

Hv = 0.85 ' Material Hardness in GPa

End If

If material = 2 Then 'Super 13 Cr Alloy

rho_metal = 7650 'Kg/m^3

EW = 26.2 ' Equivalent Weight

Hv = 0.95 ' Material Hardness in GPa

End If

If material = 3 Then '22 Cr Alloy

rho_metal = 7880 'Kg/m^3

EW = 24.9 ' Equivalent Weight

Hv = 1 ' Material Hardness in GPa

End If

While Not IsEmpty(AvgInfoWS.Cells(x, 1))

'Conveting Erosion rate units for CFD detailed particle info

'ER in KgMetal/m^2sec, from CFD-exported txt file

235

ER_5 = AvgInfoWS.Cells(x, 14) / f

e = (ER_5 * 1000) * ((86400 * 365) / (rho_metal * 0.0254)) '(mpy)

AvgInfoWS.Cells(x, 16) = e

V = AvgInfoWS.Cells(x, 7) 'm/s, from CFD-exported txt file

Avg_hits = AvgInfoWS.Cells(x, 6) 'from CFD-exported txt file

'Volume for a spherical sand particle

Vol_part = (4 / 3) * pi * (D_part / 2 / 1000000) ^ 3 'm^3

Part_flow = (Sand_rate / (Vol_part * rho_sand)) / (86400) ' particles/sec

'Actual impact frequency for the given sand rate(particles/sec)

Impact_Freq= Round(Part_flow * (Avg_hits / Simulated_particles))

Time_Between_Imp = 1 / Impact_Freq ' sec

'Area of the cell use in CFD simulation for target specimen in m^2, from CFD-exportedtxt file

Cell_area = AvgInfoWS.Cells(x, 9)

'For ER_5 in Kg metal/m^2 sec as given in fluent subroutine and CFD-exported txt file _

'ER estimated using detailed particle info

ER_mass = (ER_5 * Cell_area) / (Impact_Freq* rho_sand * Vol_part)

AvgInfoWS.Cells(x, 18) = ER_mass 'Kg metal removed / Kg of sand that hit

ER_vol = ER_mass * (rho_sand / rho_metal)

AvgInfoWS.Cells(x, 17) = ER_vol

'Surface area for a spherical sand particle

S_part = 4 * pi * (D_part / 2 / 1000000) ^ 2 ' m^2

'Surface of the indentation crater made by particle hit (new area exposed without passivefilm)

S_crater = (0.6022 * (ER_vol) ^ 0.5026) * S_part 'm^2

AvgInfoWS.Cells(x, 19) = S_crater * 10000 'cm^2

236

'Since I(t)=(Io/(1 + m*Io*t) changed of current with time depends on the product m*Iothis product needs to be adjusted by the new area of the crater

m = m_st * (Stip_st / S_crater)

Io = Io_st * (S_crater / Stip_st)

'Estimation of Average Total current PER CELL of the target specimen on CFDsimullation

T = 0

'initial values

Time(T, 0) = T * Time_Between_Imp ' sec

Time(T, 1) = T * Time_Between_Imp / 3600 'hour

I(T) = Io / (1 + Io * (m / 1000000) * T * Time_Between_Imp)

I_high(T) = I(T)

I_low(T) = 0

I_Avg(T) = (I_high(T) + I_low(T)) / 2

T = 1

Do

If x = 2 Then

Time(T, 0) = T * Time_Between_Imp

Time(T, 1) = T * Time_Between_Imp / 3600

End If

'A quickcalc method based I(T) Instantaneous current for the by the 1st hit

I(T) = Io / (1 + Io * (m / 1000000) * T * Time_Between_Imp)

'I_high(T) tracks the instantaneous response of the highs(peaks) of current caused by thenew hits

I_high(T) = I_high(T - 1) + I(T)

'I_low(T) tracks the instantaneous response of the lows(valleys) of current shown beforenew hits

I_low(T) = I_low(T - 1) + I(T)

'Average between highs and lows

I_Avg(T) = (I_high(T) + I_low(T)) / 2

'Cumulative of the I_Avg as a function of time

237

I_cum(T) = I_cum(T - 1) + I_Avg(T)

'ratio is used as stop criterion

ratio = I(T) / I_high(T)

T = T + 1

No_Particles = T

Loop While ratio >= Stop_Current

'Writing final current on summary sheet

AvgInfoWS.Cells(x, 20) = I_Avg(T - 1)

AvgInfoWS.Cells(x, 21) = I_Avg(T - 1) / (Cell_area * 10000)

'Calculating the Corrosion Component of E-C (CRe-c in mpy)

'Using the E Ratio from detailed particle imp info

AvgInfoWS.Cells(x, 22) = (I_Avg(T - 1)) * (3.27 * EW * 39.37) / _

(rho_metal * AvgInfoWS.Cells(x, 9) * 10000)

'Total E-C in mpy

'Using the E Ratio from detailed particle imp info

AvgInfoWS.Cells(x, 23) = AvgInfoWS.Cells(x, 22) + AvgInfoWS.Cells(x, 16)

w = T

Finally, after the total current per each CFD numerical cell has been estimated by

this code the total current for the whole area specimen is computed by adding up the

current value of each numerical cell. Then, the total current is converted to penetration

rate units by using Faraday’s law and the erosion-corrosion is computed as the sum of the

corrosion component rate Ce-c plus the pure erosion rate E obtained from the average of

the CFD simulation.