Materials On Black (Powder
Materials On Black (Powder
Materials On Black (Powder
Processes
A thesis presented to
the faculty of
In partial fulfillment
Master of Science
Martin L. Colahan
April 2017
Processes
by
MARTIN L. COLAHAN
David Young
Dennis Irwin
ABSTRACT
However, despite its impact on downstream and midstream operations, black powder
production is poorly understood. In the present work, black powder formation as a result
of corrosion was investigated by simulating sales gas conditions in a glass cell. Steel
specimens were systematically exposed to a range of CO2, H2S, and O2 partial pressures at
differing water condensation rates. The potential for hygroscopic material assisting black
powder formation was also investigated. Friable corrosion products found in dewing
was identified as a leading cause of black powder production from FeS. The presence of
hygroscopic NaCl crystals facilitated corrosion at relative humidities as low as 33%, but
DEDICATION
ACKNOWLEDGEMENTS
I would like express my deepest gratitude to my family and friends for without their
I would also like to thank my thesis committee members Dr. Marc Singer, Dr.
Rebecca Barlag, Dr. Kevin Crist, and Dr. Srdjan Nesic for their guidance. A special thanks
go to the technical staff at the Institute for Corrosion and Multiphase Technology (ICMT)
namely Mr. Cody Shafer, Mr. Al Schubert, Mr. Alexis Barxias, and Mr. Phil Bullington
for their technical expertise and assistance with my experimental work. I would like to
thank The Petroleum Institute in Abu Dhabi, United Arab Emirates for their financial
support. A special thanks goes to Dr. Bruce Brown of the ICMT, and Dr. Ricardo Nogueira
and Dr. Yves Gunaltun of The Petroleum Institute for their support and insightful
discussions.
Last but certainly not least, I would like to thank my advisor, Dr. David Young.
One could not ask for a better mentor. His steady guidance has been the foundation to all
the work herein and has helped me remember why I fell in love with science and
TABLE OF CONTENTS
Page
Abstract ............................................................................................................................... 3
Dedication ........................................................................................................................... 4
Acknowledgements............................................................................................................. 5
Chapter 1: Introduction................................................................................................. 16
4.6 Safety.................................................................................................................. 56
5.2.5 Black Powder Formation through Hygroscopic Corrosion Processes ...... 103
References....................................................................................................................... 110
LIST OF TABLES
Page
Table 1: Typical sales gas specifications pertinent to black powder formation [7].......... 21
Table 2: Typical H2S, CO2, O2, and H2O sales gas specifications and measured levels
within the Saudi Aramco sales gas network [8]................................................................ 21
Table 3: Common phases found in corrosion products formed in CO2, H2S, and O2
corrosion ........................................................................................................................... 29
Table 4: Sales gas composition from which experimental parameters are derived .......... 44
Table 6: Test matrix for studying the effect of H2S partial pressure on dewing corrosion.
........................................................................................................................................... 59
Table 7: Test matrix for analyzing the effect of time on sour dewing systems ................ 66
Table 8: Test matrix for the investigation of the effect of O2 on sweet dewing corrosion72
Table 9: Test matrix for the investigation of the effect of cyclic water condensation...... 77
LIST OF FIGURES
Page
Figure 1: Turbine blade eroded due to exposure to black powder. Reproduced with
permission from NACE International, Houston, TX. All rights reserved. Smart, Paper
11089 presented at CORROSION/2011, Houston, TX. © NACE International 2011 [6].
........................................................................................................................................... 18
Figure 2: (a) Measured dew point temperatures and (b) corresponding moisture contents
of sales gas measured during July and February at sampling locations in the Saudi
Aramco sales gas network [8]........................................................................................... 32
Figure 3: Ambient temperatures in the Eastern Province of Saudi Arabia. Points below
the red dotted line indicate times when dew may form in the pipeline. Reproduced with
permission from NACE International, Houston, TX. All rights reserved. A.M. Sherik,
S.R. Zaidi, E.V. Tuzan, J.P. Perez, Paper 08415 presented at CORROSION/2008, New
Orleans, LA. © NACE International 2008 [8].................................................................. 33
Figure 4: Flaky FeS corrosion product produced in marginally sour TLC conditions. (Test
Conditions: Duration: 7 days, pH2S: 0.15 H2S, pCO2: 0.93 bar, Water Condensation Rate:
0.25±0.04 ml/m2/s, Tgas: 40°C, Tsteel: 28°C Tsteel). Reproduced with permission from
NACE International, Houston, TX. All rights reserved. N. Yaakob, F. Farelas, M. Singer,
S. Nesic, D. Young, Paper 7695 presented at CORROSION/2016, Vancouver, BC. ©
NACE International 2016 [37].......................................................................................... 35
Figure 5: (a) Corrosion rates and (b) optical images taken after extraction of specimens
with NaCl deposits exposed to CO2/H2S gas mixtures with the indicated relative
humidity. Reproduced with permission from NACE International, Houston, TX. All
rights reserved. W. Litke, J. Bojes, P. Blais, J. Lerbscher, W. Wamburi, Paper 2339
presented at CORROSION/2013, Orlando, FL. © NACE International 2013. [42] ........ 38
Figure 6: (a) Corrosion rates and (b) optical images taken after extraction of specimens
with MgCl2 deposits exposed to CO2/H2S gas mixtures with the indicated relative
humidity Reproduced with permission from NACE International, Houston, TX. All rights
reserved. W. Litke, J. Bojes, P. Blais, J. Lerbscher, W. Wamburi, Paper 2339 presented at
CORROSION/2013, Orlando, FL. © NACE International 2013. [42] ............................ 39
Figure 7: Microstructure of the X65 steel specimens. Ferrite and cementite are found in
the dark and white locations, respectively. ....................................................................... 45
Figure 8: Dewing corrosion glass cell. Image courtesy of Cody Shafer. ......................... 48
11
Figure 9: Specimen holding and control stack including (A) specimen holder, (B) Peltier,
(C) Water cooled heat sink, and (D) Anchor bar and bolts. Image courtesy of Cody
Shafer. ............................................................................................................................... 49
Figure 10: XRD specimen holder. Image courtesy of Cody Shafer. ................................ 49
Figure 11: Hygroscopic corrosion glass cell. Image courtesy of Cody Shafer................. 53
Figure 12: Effect of H2S partial pressure and water condensation rate on corrosion
product morphology.......................................................................................................... 61
Figure 13: XRD data showing the effect of H2S partial pressure and WCR on corrosion
product composition.......................................................................................................... 62
Figure 14: Effect of H2S partial pressure and water condensation rate on corrosion rate.63
Figure 15: Effect of H2S partial pressure and water condensation rate (L = low, H = high)
on the measured and theoretical corrosion product masses. ............................................. 64
Figure 16: Effect of H2S partial pressure on maximum possible black powder production
rate in a 100 km, 42 inch ID pipeline................................................................................ 65
Figure 17: Effect of Time on Corrosion Product Morphology (0.3 mbar H2S, 0.96 bar
CO2, 0 mbar O2, 0.015 ml/m2/s WCR, 25 Tsteel, 30°C Tgas).............................................. 67
Figure 19: Effect of time on corrosion rate at 0.3 mbar pH2S and low WCR. ................. 69
Figure 20: Effect of time on (a) corrosion product mass and (b) maximum possible black
powder production rate in a 100 km long, 42 in. ID pipeline at 0.3 mbar pH2S and low
WCR. ................................................................................................................................ 70
Figure 21: Effect of O2 on (a) corrosion rate, (b) measured and theoretical corrosion
product mass, and (c) maximum possible black powder production rate in a 100 km, 42
in. ID pipeline. .................................................................................................................. 73
Figure 22: Surface analysis of corrosion products formed after a 3 day exposure to a 10
mbar O2 – 0.92 bar CO2 atmosphere at the low WCR. ..................................................... 75
Figure 23: (a) XRD and (b) Raman spectroscopy compositional analysis of corrosion
product formed after a 3-day exposure to a 10 mbar O2 – 0.92 bar CO2 atmosphere at the
low WCR. The red hematite spectrum is added as a reference [50]................................. 76
Figure 24: Effect of cyclic water condensation on corrosion product morphology. ........ 79
Figure 25: Effect of cyclic water condensation on the corrosion product composition. .. 80
12
Figure 26: Effect of temperature cycling on (a) corrosion rate, (b) measure and theoretical
corrosion product mass, and (c) maximum possible black powder production rate in a 100
km, 42 in. ID pipeline. ...................................................................................................... 81
Figure 27: SEM analysis of the NaCl layer formed before specimen insertion into test
environment. The specimen was coated with palladium to minimize electron beam
charging............................................................................................................................. 85
Figure 28: Slight localized corrosion on the outside boundary of a salt crystal formed by
drying NaCl....................................................................................................................... 86
Figure 29: (a) Surface microscopy after a 3 day exposures to CO2 atmosphere at 75% RH.
The region outlined in red is enlarged in (b)..................................................................... 87
Figure 30: EDS of corrosion product after a 3 day exposure to a CO2 environment at 75%
RH. NQ: Not Quantified. .................................................................................................. 88
Figure 31: XRD corrosion product composition analysis of specimens recovered after a 3
day exposure to CO2 at 75% RH. Peak labels are S: siderite (FeCO3), H: halite (NaCl),
and Fe: ferrite (Fe). ........................................................................................................... 88
Figure 32: Surface profilometry of specimen recovered after a 3 day exposure to CO2 at
75% RH. Scale bar units are in nm. .................................................................................. 90
Figure 33: Surface microscopy and XRD compositional analysis after a 3 day exposure to
CO2 at 58 and 33% RH. .................................................................................................... 91
Figure 34: Effect of relative humidity on (a) general corrosion rate, (b) measured and
theoretical corrosion product mass, and (c) maximum possible black powder production
rate in a 100 km, 42 in. ID pipeline under hygroscopic CO2 conditions. ......................... 92
Figure 36: Effect of H2S partial pressure on corrosion product composition. Phases are
M: mackinawite, S: siderite, H: halite, and Fe: ferrite. Peaks labeled with an asterisk (*)
correspond to unknown phases. ........................................................................................ 96
Figure 37: Effect of H2S partial pressure on surface morphology at 58 and 33% RH. .... 97
Figure 38: Effect of H2S partial pressure on corrosion product composition. Phases are
M: mackinawite, H: halite, and Fe: ferrite........................................................................ 99
Figure 39: Effect of H2S partial pressure on (a) general corrosion rate, (b) measured and
theoretical corrosion product mass, and (c) maximum possible black powder production
rate in a 100 km, 42 in. ID pipeline under hygroscopic conditions. ............................... 100
13
Figure 41: Cross-section view of the XRD specimen holder. Image courtesy of Cody
Shafer. ............................................................................................................................. 118
Figure 42: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, WCR: 0.05 ml/m2/s, Steel
Temperature: 15°C, Gas Temperature: 30°C, Duration: 3 days. .................................... 123
Figure 43: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 3 days. Images (a) and (b) were
from the same test as that shown in Chapter 5, whereas images (c)-(f) were from a test
with inconclusive gravimetric measurements. The corrosion product in the inconclusive
test was of a similar morphology to the CO2/O2 test. ..................................................... 124
Figure 44: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, WCR: 0.05 ml/m2/s, Steel
Temperature: 15°C, Gas Temperature: 30°C, Duration: 3 days. .................................... 125
Figure 45: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 3 days. .................................... 125
Figure 46: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.05 ml/m2/s, Steel
Temperature: 15°C, Gas Temperature: 30°C, Duration: 3 days. .................................... 126
Figure 47: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 6 hours. .................................. 126
Figure 48: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 1 day....................................... 127
Figure 49: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 3 days. .................................... 128
Figure 50: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 7 days. .................................... 129
Figure 51: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 10 days. .................................. 130
Figure 52: Conditions: pH2S: 1.0 mbar, pCO2: 0.96 bar, WCR: 0.05 ml/m2/s, Steel
Temperature: 15°C, Gas Temperature: 30°C, Duration: 3 days. .................................... 131
Figure 53: Conditions: pH2S: 1.0 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 3 days. .................................... 132
Figure 54: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 6 hours. .................................. 133
14
Figure 55: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015/0 ml/m2/s, Steel
Temperature: 25/40°C, Gas Temperature: 30°C, Duration: 3 days, Cycle Period: 1 day.
......................................................................................................................................... 134
Figure 56: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015/0 ml/m2/s, Steel
Temperature: 25/40°C, Gas Temperature: 30°C, Duration: 7 days, Cycle Period: 1 day.
......................................................................................................................................... 135
Figure 57: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, RH: 75%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. (d) and (e): after DI water rinse, (f):
after Clarke...................................................................................................................... 136
Figure 58: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, RH: 58%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. Micrograph (a): after extraction, (b):
after DI water rinse, and (c): after Clarke. ...................................................................... 137
Figure 59: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, RH: 33%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. Micrograph (a): after extraction, (b):
after DI water rinse, and (c): after Clarke. ...................................................................... 138
Figure 60: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, RH: 75%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl ........................................................... 139
Figure 61: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, RH: 58%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. .......................................................... 139
Figure 62: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, RH: 33%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. .......................................................... 140
Figure 63: Conditions: pH2S: 0.3 mbar, pCO2: 0.97 bar, RH: 75%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. .......................................................... 141
Figure 64: Conditions: pH2S: 0.3 mbar, pCO2: 0.97 bar, RH: 58%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. .......................................................... 142
Figure 65: Conditions: pH2S: 0.3 mbar, pCO2: 0.97 bar, RH: 33%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. .......................................................... 142
Figure 66: Conditions: pH2S: 1.0 mbar, pCO2: 0.97 bar, RH: 75%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. (c): after DI water rinse, (d): after
Clarke.............................................................................................................................. 143
Figure 67: Conditions: pH2S: 1.0 mbar, pCO2: 0.97 bar, RH: 58%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. Micrographs (c) and (d): after DI water
rinse................................................................................................................................. 144
15
Figure 68: Conditions: pH2S: 1.0 mbar, pCO2: 0.97 bar, RH: 33%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. Micrographs (c) and (d): after DI water
rinse................................................................................................................................. 145
16
CHAPTER 1: INTRODUCTION
Historically, energy demand in the industrialized world has largely been fulfilled through
the consumption of fossil fuels such as coal, oil, and natural gas. Investments into
alternative energy sources (e.g., wind, solar, and nuclear energy) have been made in an
attempt to diversify the global energy portfolio, but limitations including supply
intermittency, energy storage, and large up-front capital costs currently prevent the overall
transformation from a fossil fuel based economy into an alternative energy based economy.
of fossil fuels is likely to remain a primary source of energy. The United States Energy
48% between the years 2012 and 2040 to 815 quadrillion Btu, fossil fuels are expected to
supply 78% of this demand. The energy source with the largest increase in consumption is
projected to be natural gas. The abundance and low cost of natural gas makes it an
economically attractive energy source, particularly for the industrial and electrical power
also make natural gas attractive as it generates significantly less CO2 per kWh produced,
as well as combusting more cleanly, than both coal and liquid fossil fuels [2].
Natural gas is a byproduct of the decomposition of plant and animal matter trapped
underground at high pressures and temperatures over millions of years. The resulting
more valuable components, and remove the majority of undesired components. The
17
processed methane-rich gas, commonly known as sales gas, is transported to the consumer
through a network of pipelines. In 2008, the EIA estimated the network of natural gas
long, but the recent boom in natural gas production from unconventional sources (e.g.,
shale or tight gas reservoirs) means that this length will continue to increase [3].
A challenge to the operation of natural gas pipelines is corrosion, which can weaken
the integrity of the pipeline and ultimately may render the costly structure useless unless
mitigated. Internal corrosion in the oil and gas industry is mostly due to water wetting the
steel surface in the presence of acid gases including carbon dioxide (CO2) and hydrogen
sulfide (H2S). Processed natural gas can contain small concentrations of water, carbon
dioxide, hydrogen sulfide, and potentially oxygen that can be corrosive to the steel pipeline
transporting the gas. Products formed as a result of corrosion may spall off the steel surface
and become entrained in the gas becoming what is commonly known in the oil and gas
Black powder is a collective term used to generally describe particles that can be
entrained in the gas. Black powder is a problem for pipeline operators as it can erode
equipment, induce greater pressure drops, and clog instrumentation [4]. The erosive power
of entrained particles can be seen in Figure 1, which shows a power plant turbine blade
after exposure to black powder. Black powder particles vary in size and can range from
100 nm up 1 mm [5]. The composition of black powder can also vary and can include
corrosion products, salt, dirt, and other materials trapped in the pipeline during
pipelines. Transmission pipelines are large capacity pipelines that deliver gas from
processing plants to the consumer or distributor. These pipelines are designed to move
large quantities of natural gas over long distances so, therefore, have large diameters and
Figure 1: Turbine blade eroded due to exposure to black powder. Reproduced with
permission from NACE International, Houston, TX. All rights reserved. Smart, Paper
Although investigations into the impact of black powder have been performed, its
formation processes remain poorly understood. The present study investigates black
powder formation by examining the effect of CO2, H2S, and O2 on the corrosion of steel in
both water condensing conditions and humid conditions where hygroscopic material assists
wetting the steel surface. Research performed to investigate black powder formation is
powder formation.
and proposes future work to better understand black powder formation and ideas for
Black powder in gas pipelines is a costly problem for which effective mitigation
strategies are desired. Previous studies on black powder have tended to focus on its
superior mitigation strategies. A review of black powder literature was conducted by Khan
and Al-Shehhi [5] that examined black powder composition, formation, entrainment,
transport, and filtration. Due to the wide variety of topics covered, the review serves as an
excellent resource for engineers in the field with a black powder problem, but the review
of formation mechanisms was limited by the lack of available data in the scientific
literature. Black powder formation through corrosion processes is reviewed in more detail
in this chapter.
Compositions will vary, but sales gas is composed primarily of methane (CH4).
Larger hydrocarbon molecules such as ethane (C2H6) can be present, but their
concentrations are very low. Low concentrations of water (H2O), nitrogen (N2), carbon
dioxide (CO2), hydrogen sulfide (H2S), and sulfur-containing species such as methyl
mercaptan (CH3SH, added as a malodorant) are often present in sales gas, but a ceiling
content is generally specified to ensure the processed natural gas meets a gross heating
value and minimizes corrosion. Typical sales gas specifications pertinent to black powder
Table 1: Typical sales gas specifications pertinent to black powder formation [7]
Specification Value
Water Content 4-7 lbm H2O/MMscf
CO2 Content 3-4 mol%
N2 Content 4-5 mol%
Maximum O2 Content 0.01 mol%
H2S Content 0.25-1.0 grain/100 scf
Total Sulfur Content 0.5-20 grain/100 scf
Gross Heating Value 950-1200 Btu/scf
Solids and Free Liquid Commercially free
Typical specifications and measured levels of CO2, H2S, O2, and H2O were
published in a black powder study by Sherik, et al., [8] and are shown in Table 2. H2S,
CO2, and O2 levels were typically within limits, but the water content was consistently over
the 0.112 mg/L (7 lbs/MMscf) limit. The higher than specification water contents may
render the pipeline susceptible for liquid water to form and subsequently allow corrosion
to occur.
Table 2: Typical H2S, CO2, O2, and H2O sales gas specifications and measured levels
Both CO2 and H2S are considered acid gases due to their ability to form hydrogen
ions in water via dissociation processes. The H+ will react with iron at the steel surface in
an overall reaction given by Reaction 1. This electrochemical reaction can be broken into
Fe s 2H+aq → Fe aq
+
H g (1)
+
Fe s → Fe aq 2푒− (2)
If CO2 and H2S are included in the bulk gas, additional electrochemical reactions
need consideration. Carbonic acid has been postulated to directly react with the steel by the
reduction reaction given in Reaction 4, but this direct reaction mechanism is under dispute
in favor of a “buffering” mechanism [9]. Conversely, H2S is known to directly react with
Fe s H S aq → FeS s H g (5)
is dependent upon the charge transfer rate and the concentration of the oxidants at the steel
surface. The oxidant concentration at the surface is determined by both mass transfer and
the bulk concentration, which is governed by water chemistry. The effect of CO2 and H2S
The acidification of water caused by carbon dioxide takes place through several
steps. These steps include the dissolution of CO2 (Reaction 6), the hydration of CO2 to
form carbonic acid (Reaction 8), the first dissociation step of carbonic acid to form
bicarbonate ions (Reaction 10), and the dissociation of bicarbonate ions to form carbonate
ions (Reaction 12). The equilibrium constants are denoted in the reactions and are also
퐶
CO ⇌ CO (6) 퐾 (7)
g aq 푝
퐶
CO H O 푙 ⇌ H CO (8) 퐾hyd (9)
aq aq 퐶
퐶 퐶
H CO ⇌ HCO− aq H+aq (10) 퐾ca (11)
aq 퐶
퐶 퐶
HCO− aq ⇌ CO −aq H+aq (12) 퐾bi (13)
퐶
where:
With the exception of the lack of a hydration step, the acidification of water due to
H2S is similar to that of CO2. H2S does not require hydration to dissociate. To acidify water,
H2S first dissolves (Reaction 14) and then dissociates into bisulfide ions (Reaction 16),
which can dissociate further into sulfide ions (Reaction 18). Each dissociation step forms
hydrogen ions which can be reduced during the corrosion process [13]–[15].
24
퐶
H Sg ⇌ H S aq (14) 퐾 (15)
푝
퐶 퐶
H S aq ⇌ HS−aq H+aq (16) 퐾hs (17)
퐶
퐶 퐶
HS−aq ⇌ S aq
−
H+aq (18) 퐾bs (19)
퐶
Several correlations to calculate the equilibrium constants defined above are given
14.5 ∗
퐾 10− ( . + . × − . + . ) (22)
1.00258
.
퐾ca 387.6× 10− ( . − . 9 × + . × − . × − . + . ) (24)
.
퐾bi 10− ( . − .9 × + . × − . × − . +. ) (25)
퐾 10− . + . 9 − . × − 9× − .9 l og (26)
퐾 10 . 9 + . − . × − . − . ln (27)
퐾w a 10− ( 9. − . 9 + . × ) (29)
where:
푇
퐹= Temperature (°F)
푇= Temperature (K)
The ionic strength is calculated with Equation 30 where 푐, 푧, and 푁 are the
1
퐼 ∑ 푐푧 (30)
2
=
∑ 푐푧 0 (31)
=
If saturation is reached, both iron carbonate and iron sulfide can precipitate. The
precipitation reactions for iron carbonate and iron sulfide along with their attributed
, +
FeCO s ↔ Fe aq CO −aq (32)
퐾sp, 퐶 퐶 (33)
,
FeS s H+aq ↔ +
Fe aq HS−aq (34)
퐶 퐶
퐾sp, (35)
퐶
Considering temperature and ionic strength, the solubility products 퐾sp, and
퐾sp, are calculated with empirical correlations, which are given in Equations 34 and 35,
possible but do not give any indication on how fast precipitation can occur. Precipitation
nucleation. If the saturation with respect to a precipitate is greater than 1 then precipitation
may occur. The saturation (S) and precipitation (R) rates of iron carbonate and iron sulfide
퐶 퐶
푆 (38)
퐾sp,
. −
. 푆
푅 푒 퐾 푆 1 (39)
푉 sp,
퐶 퐶
푆 (40)
퐾sp, 퐶
−
, 푆
푅 푒 퐾 푆 1 (41)
푉 sp,
where:
FeS polymorph mackinawite has been postulated to form through both precipitation and a
Fe s H S aq → FeS s H g (42)
Mackinawite, FeS, is typically the first iron sulfide formed in the corrosion of steel
with iron has been justified by Sun and Nesic [20] due to the fast formation kinetics,
minimal effect of the bulk supersaturation on mackinawite formation, and the similarities
Two layers are typically observed in iron sulfide corrosion product layers: a thin,
inner layer and a porous, outer layer. Sun and Nesic [20] proposed a mechanism whereby
the thin, inner layer is formed by a reaction of H2S to iron. The outer layer is then formed
by the spallation of this thin inner layer due to internal compressive stresses that arise in
the inner mackinawite layer. The mechanism of mackinawite formation was later
reexamined by Zheng, et al., [21] who postulated that the inner layer forms through
chemisorption and the outer layer is formed by precipitation. The thin inner layer can act
natural gas transmission companies found that maximum allowable O2 concentrations were
between 0.01 and 0.10 mol% with a typical value of 0.02 mol% [24].
If oxygen is present in acidic environments such as those associated with CO2 and
H2S then the reduction of oxygen (Reaction 43) may occur [25] given the presence of a
The formation of iron oxides and iron oxyhydroxides proceeds through the
following reactions:
+
4Fe aq 4H+aq O aq
+
→ 4Fe aq 2H O l (44)
+
Fe aq 2H O l → Fe OH s 2H+aq (45)
+
Fe aq 3H O l → Fe OH s 3H+aq (46)
Fe OH s → FeO OH 푠 H Ol (47)
+
2FeO OH s Fe aq → Fe O s 2H+aq (48)
where FeO(OH) can be in the α, ß, or γ form depending on the formation conditions [26].
Corrosion products that form in sales gas conditions may include iron carbonate,
iron sulfide, iron oxide, and iron oxyhydroxide. The phases formed as a product of
corrosion will ultimately depend on both thermodynamics and kinetics, which are, in turn,
determined by pipeline conditions including CO2 partial pressure, O2 partial pressure, H2S
partial pressure, and temperature. The names, formulae, unit cell types, and the
products formed due to the internal corrosion of pipeline steel are given in Table 3.
29
Table 3: Common phases found in corrosion products formed in CO2, H2S, and O2
corrosion
Sherik, et al., [8] collected black powder samples from Saudi Aramco sour gas and
sales gas pipelines for analysis by X-ray diffraction (XRD) and X-ray fluorescence (XRF).
The sour gas pipeline black powder samples were predominantly composed of FeS and
30
FeS2 with minor amounts of Fe3O4, FeO(OH), and FeCO3. XRD of black powder samples
from the sales gas pipelines detected Fe3O4, α-FeO(OH), and γ-FeO(OH) with small
amounts of FeCO3. However, sulfur was detected in all collected samples by XRF, which
suggests the black powder may contain an amorphous sulfur-containing species. Elemental
sulfur was detected with XRD in some samples, but it is unclear if it was detected in the
Japanese sales gas network. The black powder was determined to be predominantly
magnetite, but small amounts of hematite, goethite, lepidocrocite, and siderite were also
found. Iron sulfide and elemental sulfur was detected in trace amounts in pipelines that
historically transported H2S-containing sales gas. Both mill scale and the corrosion of
pipeline steel due to oxygen and water vapor ingress is attributed to be a major source of
black powder due to the high quantity of magnetite collected. Ingress of O2 and H2O during
The identification of iron oxide and iron oxyhydroxides as the major components
of black powder in sales gas pipelines containing O2 and liquid water were used by Sherik,
et al., and Yamada, et al., to justify the corrosion of steel due to oxygen as the primary
source of black powder. However, the presence of lepidocrocite and elemental sulfur
suggests the formation of iron oxide and iron oxyhydroxide from an iron sulfide precursor
likely occurred. The formation of lepidocrocite in sales gas conditions is largely attributed
to the oxidation of mackinawite, which is shown in Reaction 49 [31], [32]. Magnetite can
then form through the reaction of lepidocrocite and ferrous ions (Reaction 50) [33].
31
+
2훾-FeO OH s Fe aq → Fe O s
2H+aq (50)
The oxidation of iron sulfide is not the only pathway to elemental sulfur. Elemental
sulfur may be obtained through the oxidation of H2S as shown in Reaction 51.
8H S aq 4O aq → 8H O l S s (51)
The oxidation of iron carbonate may also occur and is reported to follow Reaction
52.
Measured dew points of water in the sales gas network were reported by Sherik, et
al., [8] and are given in Figure 2. The sales gas moisture content consistently exceeded the
maximum moisture level of 7 lbs/MMscf (0.112 mg/L) set by sales gas specifications. The
dew point temperature of water was generally lower than 10°C, but several dew points
exceeded 15°C, which may lead to water condensation depending on the pipeline steel
temperature.
32
(a) (b)
Figure 2: (a) Measured dew point temperatures and (b) corresponding moisture contents
of sales gas measured during July and February at sampling locations in the Saudi
Sherik compared the measured water dew points to meteorological data to examine
the potential for water condensation on the steel pipeline. As indicated by Figure 3, winter
ambient temperatures in Saudi Arabia were often below the water dew point temperatures
measured; therefore, water condensation was likely. If the ambient temperature serves as
an adequate proxy for the steel temperature, then water condensation is likely to be cyclical
where water condensation can be expected to occur during cool time periods, such as
overnight, and not occur during warm time periods such as during the day. Cyclical water
wetting may affect the corrosion product layer morphology due to fast precipitation during
unknown, but corrosion in dewing conditions has been studied in top-of-the-line corrosion
studies.
Figure 3: Ambient temperatures in the Eastern Province of Saudi Arabia. Points below
the red dotted line indicate times when dew may form in the pipeline. Reproduced with
permission from NACE International, Houston, TX. All rights reserved. A.M. Sherik,
S.R. Zaidi, E.V. Tuzan, J.P. Perez, Paper 08415 presented at CORROSION/2008, New
hydrocarbons, and organic acids. If the gas is saturated with water and the steel temperature
is cooler than the gas temperature, then the water can condense on the steel and cause
corrosion. Corrosion in these conditions can be severe near the top of the pipeline where
34
nonvolatile corrosion inhibitors cannot reach, therefore, this type of corrosion has been
Water condensation plays a critical role in TLC. A high water condensation rate
(WCR) would limit the ability of a corrosion product to precipitate due to dilution and thus
leave the steel unprotected. Conversely, a low WCR would allow for higher ferrous ion
product layer [35]. A model for determining the WCR based on heat transfer was developed
Studies have been performed that investigated the formation of corrosion products
in TLC conditions since corrosion products can offer some protection against corrosion.
Yaakob, et al., examined the effect of iron sulfide on localized corrosion in marginally sour
conditions similar to sales gas conditions without O2. Steel specimens mounted in the lid
of a glass cell with the polished side facing down were exposed to water-saturated CO2/H2S
mixtures with H2S partial pressures of 0, 0.015, 0.03, 0.08, and 0.15 mbar. To force water
condensation on the specimens, their surfaces were cooled to below the 40 or 60°C gas
temperature. Severe localized corrosion was found at 0.015 and 0.03 mbar H2S and was
attributed to failures within the FeS corrosion product layer. The corrosion product layers
obtained at 0.08 and 0.15 mbar H2S were more protective as no localized attack was
observed, but the FeS layers were flaky and appeared to readily detach from the steel
surface. Figure 4 shows the corrosion product morphology from the 40°C Tgas, 0.15 mbar
H2S partial pressure experiment. Corrosion product spallation appears to have occurred
during either test exposure or the isopropanol rinse after specimen extraction, as evidenced
35
by the visibility of the inner FeS layer seen in the right half of Figure 4 [37]. The role of
unknown.
Figure 4: Flaky FeS corrosion product produced in marginally sour TLC conditions. (Test
Conditions: Duration: 7 days, pH2S: 0.15 H2S, pCO2: 0.93 bar, Water Condensation Rate:
0.25±0.04 ml/m2/s, Tgas: 40°C, Tsteel: 25°C). Reproduced with permission from NACE
with water and subsequent pressurization to check for leaks. Due to volume requirements
and water costs, a pipeline operator may elect to use seawater, which can leave a
36
hygroscopic salt residue on the pipe wall after the seawater is drained. If not properly
rinsed, the salt residue can accumulate water from a humid sales gas stream.
hygroscopic material is dependent primarily on its affinity toward water, relative humidity
(RH), and temperature. Soluble materials with a high affinity toward water can dissolve if
the availability of water is sufficiently high through a process called deliquescence. The
crystallizes due to drying. The DRH and the efflorescence relative humidity (ERH) are not
necessarily the same as hysteresis between the DRH and the ERH exists due to
crystallization kinetics [38]. Hygroscopic uptake of water by chloride and perchlorate salts
in Martian soil has been hypothesized to be the reason for recurring slope lineae, dark
streaks on Martian slopes only present during warmer seasons [39]. The potential for
Hygroscopic salts and their potential to lead to corrosion have been examined in
atmospheric corrosion studies where the deposition of salt aerosols on metals can lead to
corrosion in oxic conditions. The corrosive effect of hygroscopic NaCl particles on mild
steel in atmospheric environments was investigated by Schindelholz, et al. [40]. NaCl was
deposited onto mild steel specimens and subjected to air with relative humidities ranging
from 1.5% to 90% RH for up to 300 days, demonstrating that water adsorbed on the NaCl
particles and the steel can initiate corrosion at relative humidities as low as 33%. The
37
initiation of corrosion was attributed to water adsorbed to the NaCl crystals and the steel
surface with capillary condensation, which accumulates water at the salt/steel interface.
The effect of CO2 and H2S on hygroscopic corrosion is even less understood.
Kolts [41] investigated the corrosion of sales gas pipelines due to salt deposits left behind
by a seawater hydrotest. Steel samples with and without dried synthetic seawater deposits
were exposed to RH-controlled CO2. Kolts found salts can induce corrosion in the
conditions tested but oxygen ingress is suspected to occur which will skew results.
Litke, et al., [42] investigated the corrosion of steel under hygroscopic salts in
salt dried under vacuum onto UNS G10180 steel coupons and then inserted the specimens
into autoclaves with CO2 and H2S. The humidity of the autoclaves was controlled with
before loading the samples. Figure 5 and Figure 6 show optical images of the exposed
surface after extraction from a 14 day exposure to a gas with 15 psig H2S and 57 psig CO2.
General and localized corrosion rates for the NaCl and MgCl2 deposit tests were measured
and are shown in Figure 5 and Figure 6, respectively. Both corrosion and corrosion product
formation were found to occur in the conditions tested, however, reservations exist due to
the measurement of similar general corrosion rates in the NaCl experiments at 86% and
0.7% RH and the appearance of corrosion product at 0.7% RH in the NaCl experiments.
The surface of the NaCl/0.7% RH specimen is expected to look more like the surface of
the MgCl2/0.7% RH specimen with the white crystalline salt still on the surface. Corrosion
underneath the NaCl crystals at 0.7% RH should be negligible due to the minute amount
38
of water in the vapor phase. The unexpected behavior is likely due to the ethylene glycol
(a)
(b)
Figure 5: (a) Corrosion rates and (b) optical images taken after extraction of specimens
with NaCl deposits exposed to CO2/H2S gas mixtures with the indicated relative
humidity. Reproduced with permission from NACE International, Houston, TX. All
(a)
(b)
Figure 6: (a) Corrosion rates and (b) optical images taken after extraction of specimens
with MgCl2 deposits exposed to CO2/H2S gas mixtures with the indicated relative
humidity Reproduced with permission from NACE International, Houston, TX. All rights
Both Kolts and Litke, et al., found corrosion to occur due to salt deposits in humid
hygroscopic salts can contribute to black powder production as neither Kolts nor Litke
corrosion product can grow in a pipeline, but if that corrosion product cannot spall then it
cannot become entrained in the gas and thus form black powder. However, to understand
how corrosion product spallation leads to black powder, the processes that lead to corrosion
Water availability has been identified as the primary unknown for corrosion in sales
gas conditions since the concentrations of CO2, H2S, and sometimes O2 are known to be
adequate for signification corrosion to occur. If the water content of the sales gas is high,
then water may condense onto the steel surface. However, if the water content is below the
thermodynamic dew point temperature, then a hygroscopic material like salt must be
present for corrosion to occur. Cyclical temperature fluctuations were also identified as
potentially changing the corrosion product morphology. Spalled corrosion products were
obtained in dewing conditions with H2S, but the effect of acid gas concentration and water
condensation rates on the production of friable corrosion products is still unknown. While
is known about the potential for hygroscopic corrosion in CO2 and H2S conditions to
3.1 Objective
products, associated with black powder when spalled, developed through dewing and
hygroscopic processes. The overall goal is to gain a better understanding of the role of
water availability and H2S concentration on the corrosion of sales gas pipelines.
This study is broken into two parts. Firstly, corrosion under dewing conditions is
examined where water is forcibly condensed by cooling steel specimens in simulated sales
gas conditions. The objective is to examine the corrosion products produced in a worst-
case scenario; the effect of cyclic water condensation on corrosion product formation is
determine if salts on the surface of the steel may lead to black powder production in
3.2 Hypotheses
form on the steel surface although kinetically favored compounds will also be
encountered.
2. The formation of corrosion products follows a cyclical pattern that correlates with
3. The cyclic behavior of water condensation and evaporation in sales gas pipelines
4. Corrosion products will form on the steel surface in non-water saturated sales gas
3.3 Motivation
The Abu Dhabi based natural gas company Gasco is reporting black powder
formation in their sales gas pipelines. Filtration and pigging are currently employed to
minimize the impact to downstream operations, but are costly. A joint research partnership
between the Abu Dhabi Petroleum Institute and the Ohio University Institute for Corrosion
and Multiphase Technology has been tasked with examining the role of corrosion products
on black powder formation in the Gasco pipelines as well as to examine potential methods
Under dewing conditions, water condensation was forced by cooling the steel specimens
hygroscopic conditions, NaCl was deposited onto steel specimens that were then placed in
NaCl or maintains a solid salt layer on the steel. Upon specimen extraction, corrosion
product morphology was examined with scanning electron microscopy, and corrosion
product composition was obtained with energy dispersive X-ray spectroscopy, Raman
profilometry.
experimental parameters are shown in Table 4. Oxygen and water contents were not
measured. The 40 inch (1 m) GASCO pipeline transported the gas at a flow rate of 550
From the specified gas concentration and the highest operating pressure, the partial
pressures of CO2 and H2S were calculated to be approximately 0.55 bar and 0.3 mbar,
respectively.
44
The experimental gas composition was primarily based on the partial pressure of
H2S measured in the pipeline. The experimental H2S partial pressures were 0, 0.1, 0.3, and
1.0 mbar to examine corrosion in simulated GASCO conditions, and the effect of increased
and deceased H2S contents. The H2S was diluted with CO2 for a CO2 partial pressure of
0.96 to 0.98 bar. Corrosion at 0.1 – 1 mbar H2S is known to be H2S dominant, so dilution
over the GASCO conditions should not affect results. Tests with O2 were conducted at 10
mbar O2, the O2 partial pressure expected if the O2 content was approximately 0.02 mol%.
Table 4: Sales gas composition from which experimental parameters are derived
The steel used for all experiments is API 5L X65 mild steel, a grade of steel
typically used for gas transmission pipelines. The steel composition is listed in Table 5.
The steel is a low carbon steel with a ferritic-pearlitic microstructure. The microstructure
is shown in Figure 7, which was determined after etching with a 2% Nital solution (HNO3
45
in CH3CH2OH). Cementite (Fe3C) and ferrite are found in the white and dark regions,
respectively.
C Al As Co Cu Mn Mo Nb Ni P
0.05 0.033 0.015 0.012 0.14 1.21 0.16 0.03 0.38 0.004
S Sb Si Sn Ti V Zn Fe
Figure 7: Microstructure of the X65 steel specimens. Ferrite and cementite are found in
CO2/H2S or CO2/O2 mixtures. Water condensation was stimulated onto the steel specimens
by cooling their surfaces to 15°C or 25°C for high or low water condensation rates (WCR),
through deionized (DI) water heated at a temperature to maintain the gas temperature at
30°C. A 30°C gas temperature was selected by rounding down the 33°C GASCO sales gas
operating temperature; the temperature most likely to condense water. The apparatus and
procedures for testing in dewing conditions are based largely on those used by Yaakob, et
al., [37] to test marginally sour TLC conditions. Modifications, described below, were
A glass cell setup, as shown in Figure 8, was used for testing black powder
H2S, and O2 concentrations, was bubbled through 1 liter of heated deionized (DI) water
located in the bottom of the glass cell to heat and saturate the gas with water. The
temperature of the water, which is controlled by the hotplate, is set such that the gas above
(31.7 mm × 12.7 mm) cylindrical specimen for surface and corrosion rate analysis, and a
0.5×0.5×0.08 inch³ (12.7×12.7×2 mm³) specimen for XRD compositional analyses. The
cylindrical specimens were inserted directly into PEEK specimen holders designed to
47
provide a gas-tight seal to the lid while exposing a specimen surface for temperature
control. The smaller XRD specimens were suspended from the lid with a specially designed
holder, shown in Figure 10, where the steel specimens were held by a magnet within an
aluminum body, which aids heat transfer to the Peltier thermoelectric cooler. The steel
specimens were coated with Xylan on surfaces in contact with the holder to prevent
C
B
A
Figure 9: Specimen holding and control stack including (A) specimen holder, (B) Peltier,
(C) Water cooled heat sink, and (D) Anchor bar and bolts. Image courtesy of Cody
Shafer.
mounted to the side of the cylindrical specimens or within the aluminum body of the XRD
50
feedback control loop, and supplying the computed power to a thermoelectric cooler in
contact with the specimen. The thermoelectric cooler, more commonly known as a Peltier
after the French physicist Jean Charles Athanase Peltier, acts as a heat pump when
electrical power is applied. By controlling the amount of power supplied to the Peltier, the
over the Peltier to prevent it from overheating. Temperature control was primarily done
with a TC-48-20 Peltier controller by TE Technology®, but very similar results were
obtained using a controller developed by the author of this thesis based on an Arduino®
microcontroller.
steel specimen must be periodically changed. To control the temperature changes with a
Peltier, the polarity of the Peltier must be switched. Changing the polarity of the Peltier
changes the direction heat is pumped and, therefore, switches the hot and cold sides of the
Peltier. The TC-48-20 Peltier controller cannot change the polarity internally, so a double-
pole-double-throw mechanical relay was connected between the controller and the relay.
MOSFETs in an H-bridge configuration were used to control both the power and the
polarity in the Arduino controller. Additional information regarding the controllers can be
found in Appendix A.
51
The glass cell shown in Figure 8 was cleaned then 1 liter of deionized water
introduced. The glass cell was then placed on the hot plate and the DI water heated to 35°C
to maintain the gas temperature at setpoint. The glass cell was purged with CO2 for at least
procedures, the specimens were then ground with silicon carbide abrasive papers to a 600
grit finish with water as the coolant. The water was rinsed from the surface with
were cleaned in isopropanol in an ultrasonic bath, dried with cool air, and then the initial
mass was recorded. Specimens were stored in nitrogen until insertion, which was ordinarily
The specimens were inserted into the specially designed holders outlined above and
mounted to the glass cell lid. The Peltiers and heatsinks were placed over the specimens,
and the resultant stack fastened to the lid. The thermistors and Peltiers were connected to
the controllers, and specimen temperature control was started. The glass cell was allowed
to purge for an additional 30 minutes with CO2 before H2S was then added into the system.
The H2S concentration was set by mixing a CO2 and a CO2/H2S stream metered by a
rotameter upstream of the glass cell. The H2S concentration was measured with
colorimetric gas detector tubes (GASTEC 4HM) and adjusted as necessary. Gas was
continuously purged through the glass cell throughout the experiments. Effluent gas
52
containing H2S was passed through a sodium hydroxide solution and finally an activated
carbon scrubber. Test conditions were monitored and adjusted as needed to maintain
experimental parameters.
After the allotted exposure time, the specimens were removed and immediately
rinsed with isopropanol to remove water and prevent unwanted oxidation. The rinse was
performed by targeting the top of the specimen with a stream of isopropanol from a squirt
bottle and allowing the isopropanol to fall over the rest of the specimen. The wall shear
stress of the isopropanol rinse was calculated to be about 1 Pa, roughly the same wall shear
stress in the GASCO pipeline. Whether or not corrosion product was found in the waste
isopropanol was recorded. The specimens were dried under nitrogen, optical images were
taken, and the specimen with corrosion product was weighed. The specimens were then
The effect of salts on the corrosion of steel in unsaturated conditions was examined
by depositing salts onto steel specimens and exposing the specimens to relative humidity-
controlled CO2, CO2/H2S and CO2/O2 mixtures. The three relative humidities tested were
75%, 58%, and 33% controlled through the use of saturated aqueous solutions of NaCl,
Hygroscopic corrosion tests were performed in a glass cell with the specimens
located on a platform above the saturated salt solutions in the bottom of the glass cell, as
shown in Figure 11. The apparatus consisted of a glass cell, TeflonTM lid, lid clamp, gas
53
inlet, gas outlet, specimen platform, and thermocouple. A large hole was made in the lid
for fast specimen insertion and extraction. The large hole was sealed with a rubber stopper,
Figure 11: Hygroscopic corrosion glass cell. Image courtesy of Cody Shafer.
Prior to testing, the glass cell, lid and stage are thoroughly cleaned with DI water
and isopropanol then dried. In the bottom of the glass cell, 100 ml of DI water was mixed
with salt at 125% of the solubility limit in water at 25°C. The stage is then placed into the
glass cell and the lid is fastened. The glass cell is purged with N2 for a minimum of 4 hours
to remove oxygen.
54
X65 steel specimens, the same geometry as those used in dewing conditions, were
coated with Xylan on all surfaces where corrosion is undesired. The steel surface was
ground with carbide abrasive paper and then polished to a mirror finish by polishing with
a 0.25 µm diamond suspension. The specimens were then thoroughly rinsed with DI water
and ultrasonically cleaned in isopropanol. Residual isopropanol was evaporated with cool
air, and the specimen mass was recorded. The specimens were stored in N2 until salt was
deposited onto the surface or the clean specimen was inserted into the glass cell.
Salt layers were generated by drying a NaCl/DI water solution placed on the steel
specimens with N2. A 3.5 wt.% NaCl solution was prepared prior to experimentation and
used for all experiments; this is a similar salinity to seawater. The salt solution was purged
with N2, and 250 µl of solution was deposited onto the steel surface with a pipette. The
specimen was immediately placed in a N2 environment, and the water was then spread over
the entire surface by tilting the specimens. The salt solution was dried by passing dry N2
over the wetted steel surface. Once the salt crystallized, the specimen mass with salt was
measured, photographed, and then placed in the N2-purged glass cell on the platform salt-
side up.
The glass cell was allowed to purge with N2 for another 30 minutes before
replacement with the test gas, which was continuously sparged into the system throughout
each experiment. The dry test gas entered the glass cell through a tube with the outlet
approximately 1 cm above the saturated salt solution. The gas was not bubbled through the
salt solution to prevent aerosol formation, which can contaminate the steel surface; this
phenomenon was observed in preliminary experiments with saturated MgCl2. A small gas
55
flow rate was maintained throughout the experiment to prevent a decrease in the gas RH.
In sweet conditions, a hygrometer (GE Panametrics Moisture Monitor Series 35) was
placed downstream of the glass cell to verify the relative humidity, but the hygrometer
could not be used in H2S testing due to chemical incompatibility. Specimen extraction was
examine the corrosion product morphology. SEM was performed with a JEOL JSM-
minimize charging in the SEM. Cross-sectional SEM was performed after first mounting
the specimen in epoxy to stabilize the corrosion product and cutting the specimen normal
to the steel surface with a diamond-tipped saw blade. The cross-section was polished to a
mirror finish through the use of consecutively finer grit silicon carbide abrasive paper and
minimize charging.
dispersive X-ray spectroscopy (EDS), Raman spectroscopy, and, more generally, by X-ray
diffraction (XRD). EDS was performed with an EDAX Genesis attachment on the SEM.
Raman spectroscopy was conducted with a Bruker Senterra with a 785 nm wavelength
laser. XRD was performed on a Rigaku Ultima IV with a CuKα source over 10-70° 2θ at
a scan rate of 1° per minute. The corrosion product was removed in accordance with ASTM
56
G1-03 [43], and the corrosion rate and corrosion product mass was calculated by
comparing the after removal mass to the initial mass and the after extraction mass,
InfiniteFocus microscope.
4.6 Safety
The primary risk during the work conducted for this thesis is exposure to hydrogen
sulfide through inhalation. H2S atmospheres above 100 ppm are considered immediately
dangerous to life and health. To protect against H2S inhalation, experiments with H2S
concentrations above 0.1 mbar (100 ppm) H2S were conducted in a segregated laboratory
at the Ohio University Institute for Corrosion and Multiphase Technology, which has a
The safety system contained multiple levels of defense. The first level is, of course,
maintaining a well-sealed glass cell. If the gas inlet flow rate did not match the gas outlet
flow rate, then H2S was not added into the system. The second level was the vent hood in
which the glass cell was placed. Any leakage of H2S from the glass cell would be largely
contained by the vent hood and sent to an activated carbon scrubber. Then, if necessary,
the effluent gas is passed through a combustion system and released. The third level of
defense was personal protective equipment (PPE). PPE worn included both an H2S badge
sensor and a self-contained breather apparatus (SCBA). The H2S badge was worn during
any entry into the H2S room, regardless of whether or not experiments were being
conducted. The badges have two audible and tactile alarms with the low alarm at 10 ppm
57
and a high alarm at 15 ppm. The badges were checked daily with a 25 ppm H2S/air mixture
to ensure they functioned properly. The SCBA was worn during any entry in the H2S room
while an H2S experiment was conducted. Air was supplied to the SCBA wearer from
cylinders located outside the H2S room through hoses to prevent H2S inhalation in the event
of a release. Lastly, in case all previous safety layers failed, the buddy system was
employed where trained personnel were stationed outside the H2S room in visual contact
with the person inside and ready to enter the laboratory at a moment’s notice in the event
of incapacitation. H2S experiments below the 100 ppm H2S threshold were conducted in a
vent hood outside the H2S laboratory. A badge was still worn when working in the vent
acid is a primary component of the solution, commonly known as Clarke solution, used to
remove corrosion product. Any work with hydrochloric acid was done in a vent hood to
minimize exposure to HCl fumes. Nitrile gloves were also worn to prevent skin contact.
58
If pipeline conditions are favorable, water may condense on the steel wall and allow
CO2, H2S, and O2 to corrode the steel and develop corrosion products, which may spall and
corrosion products in dewing sales gas conditions. The results of the experiments are
To examine the effect of H2S on the corrosion of pipeline steel in water condensing
conditions, experiments were conducted at varying H2S partial pressures. H2S partial
pressures tested were 0, 0.1, 0.3, and 1 mbar with calculated water condensation rates
(WCR) of 0.015 and 0.05 ml/m2/s. Experimental conditions are summarized in the test
Table 6: Test matrix for studying the effect of H2S partial pressure on dewing corrosion.
After extraction from the glass cell, the specimens were analyzed using SEM.
Micrographs of the corrosion product on the steel surface after the extraction of specimens
are shown in Figure 12. In CO2 conditions, minimal corrosion product other than residual
iron carbide was present. Some iron carbonate was present at the low WCR, whereas iron
carbide was only present at the high WCR. At 0.1 mbar H2S, a friable corrosion product
was present at both WCRs, but buckling was present in regions presumed to be the edges
of water droplets at the low WCR. At higher H2S partial pressures, FeS spallation was
much more pronounced. A higher degree of corrosion product layer buckling was observed
at 0.3 mbar H2S than at 0.1 mbar, but the buckles were confined to specific regions.
Buckling was also present at the high WCR at 0.3 mbar H2S. At 1.0 mbar H2S, severe
buckling was observed over the specimen surface at the low WCR. At the high WCR,
60
buckling still occurred, but was not nearly as severe. Buckle-driven spallation is examined
the effect of pH2S are shown in Figure 13. At the low WCR, XRD did not pick up any
corrosion product on the specimens analyzed from the CO2 experiment, but mackinawite
(FeS) was detected in tests with H2S. Increasing the H2S partial pressure saw an increase
in mackinawite peak intensity, but the relative peak intensities did not match the
mackinawite XRD reference data [28] indicating preferred orientation of the mackinawite
layer on the steel substrate. This has been postulated to be due to a growth mechanism
favoring parallel [100] and [001] iron and mackinawite planes, respectively, at the steel
surface. At the high WCR, mackinawite was detected at 0.3 and 1.0 mbar H2S, but not at
0.1 mbar.
61
Figure 12: Effect of H2S partial pressure and water condensation rate on corrosion
product morphology.
62
Figure 13: XRD data showing the effect of H2S partial pressure and WCR on corrosion
product composition
63
The effect of H2S concentration on corrosion rate is shown in Figure 14. For both
WCRs the addition of 0.1 and 0.3 mbar H2S generally decreased the corrosion rate due to
the corrosion product offering more protection. However, increasing the H2S partial
pressure to 1.0 mbar saw an uptick in the corrosion rate, particularly for the low WCR. The
dramatic increase in corrosion rate at 1.0 mbar H2S and low WCR is presumed to be due
to the much higher degree of spallation observed at 1.0 mbar H2S than at 0.3 and 0.1 mbar
H2S, since a spalling layer cannot offer as much protection to the steel as a more stable
layer. The spallation of a corrosion product would expose the steel surface to further attack.
The smaller uptick in corrosion rate observed in the 1.0 mbar and high WCR experiment
relative to its low WCR counterpart is likely due to the lesser degree of corrosion product
spallation.
Figure 14: Effect of H2S partial pressure and water condensation rate on corrosion rate.
64
Figure 15: Effect of H2S partial pressure and water condensation rate (L = low, H = high)
The measured corrosion product mass and theoretical corrosion product mass are
shown in Figure 15. The measured corrosion product mass is the mass of corrosion product
removed with the inhibited hydrochloric acid (Clarke solution) in the procedure described
by ASTM G1. The theoretical corrosion product mass is determined by calculating the
maximum amount of corrosion product possible derived from iron lost due to corrosion.
The theoretical corrosion product mass calculation assumes all iron lost due to corrosion,
not in the form of Fe3C, is converted to FeCO3 under sweet conditions and FeS under sour
comparison against the measured corrosion product mass as it gives some insight into the
amount of iron that was lost in removed condensed water or spalled corrosion product. As
65
evidenced by the large difference in the measured and theoretical corrosion product mass,
most of the iron available for corrosion product formation on the steel surface was lost.
The difference in the theoretical and measured corrosion product masses generally
decreased with increasing H2S concentration, but this is expected since corrosion product
layers are easier to produce in sour systems than sweet systems at the temperatures tested.
extrapolating the theoretical corrosion product mass averaged by the experimental duration
to a hypothetical 100 km length of a 42 inch (106.7 cm) ID pipeline. The maximum black
powder production rates varied from 900 to 2000 kg/day for a 100 km, 42-inch ID pipeline.
These numbers are high, but they represent a worst-case scenario where the entire
Figure 16: Effect of H2S partial pressure on maximum possible black powder production
The effect of time was examined by exposing steel specimens to 0.3 mbar H2S and
0.96 bar CO2 in an O2-free environment for up to 10 days at the low WCR. Experimental
Table 7: Test matrix for analyzing the effect of time on sour dewing systems
The effect of time on the corrosion product layer morphology is shown in Figure
17. After 6 hours, the only visible corrosion product was confined to the structures shown.
Flakes were already beginning to form, but no buckling was observed. FeS layer buckling
was evident after 1 day where the buckled corrosion product was evident only surrounding
the flakey structures seen after 6 hours. The severity of buckling intensified with time
developing large buckled regions after 7 days. Similar sized features were not observed
after 10 days as they likely had been removed from the surface due to spallation.
67
3 Days 7 Days
10 Days
Figure 17: Effect of Time on Corrosion Product Morphology (0.3 mbar H2S, 0.96 bar
The effect of time on the corrosion product composition is shown in Figure 18. No
change in phase identity associated with corrosion product was observed, however,
68
absolute peak intensities generally increased as the corrosion product developed. No XRD
Time dependency relating to corrosion rate is shown in Figure 19. The corrosion
rate decreased with time from 0.28 mm/yr to a steady corrosion rate of 0.10 mm/yr after
corrosion product, which can offer some protection, had time to develop on the steel
surface. The error in the measurement decreased significantly over time since mass
measurement uncertainties had less impact on the error-propagated result as time increased.
69
Figure 19: Effect of time on corrosion rate at 0.3 mbar pH2S and low WCR.
As shown in Figure 20, the measured corrosion product mass on the steel surface
is approaching a steady value near 10 g/m2 while the theoretical corrosion product mass
becomes approximately linear with time. This combined with the changes in corrosion
product morphology over time indicate the corrosion product formation rate and spallation
rate are approaching equilibrium and may ultimately lead to a black powder production
rate of ca.1000 kg/day for a 100 km, 42 in. ID pipeline. Note, the errors in the measured
and theoretical corrosion product mass measurements were relatively constant since they
are based soley on gravimetric measurements and did not consider time, whereas the
maximum possible black powder production rate is based on the corrosion rate, which had
(a) (b)
Figure 20: Effect of time on (a) corrosion product mass and (b) maximum possible black
powder production rate in a 100 km long, 42 in. ID pipeline at 0.3 mbar pH2S and low
WCR.
The increased degree of buckling with increasing H2S partial pressure and time in
the FeS layers is related to the accumulation of compressive stress accompanying FeS layer
growth. Spontaneous buckling and layer delamination occurs to release stored energy
model [20] of iron sulfide layer formation, stress accumulation, which leads to cracking
and spallation, is postulated to be due to volume differences between the ferrite and the
mackinawite. The ratio of the scale volume to the substrate volume, more commonly
known as the Pilling-Bedworth ratio (PBR) [44], for FeS relative to ferrite is about 2.6.
Since the PBR is greater than 1, compressive stress within the layer is expected to
71
accumulate and lead to layer failure. However, the PBR is not especially useful for layers
grown through outward cation diffusion, which is the case for most sulfides. Iron cations
may diffuse faster than the much larger sulfide anions. Layers predominantly grown via
outward cation diffusion can grow outward from the scale surface and, therefore, are not
Sun and Nesic suggest epitaxial stresses, which arise due to unit cell size
differences may lead to stress accumulation [20]. However, the adherence of the sulfide to
the metal is expected to be fairly weak due to void formation as the steel substrate corrodes,
which may limit epitaxial stresses [38]. A perhaps more plausible method of stress
accumulation may be due to crystal growth at cracks, pores, and grain boundaries. Crystal
growth due to the addition of ferrous and sulfide ions to an already formed crystal may lead
to crystals pushing against one another therefore compressively stressing the layer. Crystal
growth at grain boundaries and cracks has been used to explain the buckling of chromium
In all H2S experiments, black particles were present in collected fluid after the
isopropanol rinse of extracted specimens. The wall shear stress imposed by the isopropanol
was estimated to be approximately 1 Pa based on the flow of a falling laminar film [49].
This shear stress is similar to the wall shear stress in the GASCO pipeline. It is unlikely
this small shear stress would cause damage to a corrosion product layer experiencing little
predominantly caused by intrinsic stresses within the layer generated by layer growth rather
72
than the extrinsic stresses caused by fluid flow. Further study is needed to better quantify
The effect of oxygen was examined by exposing steel specimens at the low WCR
of 0.015 ml/m2/s to a CO2/air mixture with a 10 mbar O2 partial pressure. Test conditions
Table 8: Test matrix for the investigation of the effect of O2 on sweet dewing corrosion
(a)
(b) (c)
Figure 21: Effect of O2 on (a) corrosion rate, (b) measured and theoretical corrosion
product mass, and (c) maximum possible black powder production rate in a 100 km, 42
in. ID pipeline.
in corrosion rate and corrosion product mass, as shown in Figure 21(a) and Figure 21(b).
The measured corrosion product mass was much closer to the theoretical corrosion product
74
mass than the CO2-only experiment indicating the water chemistry was more favorable for
precipitation.
The dramatic increase in the corrosion rate and corrosion product mass resulted in
22(a). The inner layer shown in Figure 22(b) was well adhered to the steel, but the outer
layer, Figure 22(c), was more loosely bound. Red flakes of corrosion product were
observed in the collected isopropanol after rinsing, which suggests black powder
The origin of the outer layer is unclear. The outer layer is presumed to have grown
on the gas/liquid interface since a flat surface is present on the outer layer, whereas the
inner layer is more rough. It is postulated that the precipitation of iron carbonate in the bulk
occurred away from the steel surface whereupon it collected at the gas/water interface due
to gravity.
75
(a)
(b) (c)
Figure 22: Surface analysis of corrosion products formed after a 3 day exposure to a 10
analysis with XRD and Raman spectroscopy as shown in Figure 23. Siderite (FeCO3) was
detected with XRD, but hematite (Fe2O3) was detected by Raman spectroscopy. It is
corrosion product surface, which cannot be detected by XRD but is detected by Raman
spectroscopy. Indeed, the high XRD baseline is also indicative of the presence of an
76
amorphous corrosion product. As Raman is more suface specific, the iron carbonate
Figure 23: (a) XRD and (b) Raman spectroscopy compositional analysis of corrosion
product formed after a 3-day exposure to a 10 mbar O2 – 0.92 bar CO2 atmosphere at the
Under pure CO2 environments, corrosion products were minimal and consisted
primarily of iron carbide. However, the predominance of iron carbonate over iron
oxyhydroxides and iron oxides even in the presence of dissolved O2 suggests black powder
77
formation may still be possible under pure CO2 conditions if conditions are favorable for
precipitation to occur into a morphology, such as the bilayered morphology seen above,
Pipeline conditions are not static. Temperatures are expected to change throughout
the year and even throughout the day due to meteorological conditions. To examine the
specimens was changed every 12 hours between 25°C and 40°C. These temperature
changes are supposed to mimic temperature changes that occur due to the day/night cycle.
Table 9: Test matrix for the investigation of the effect of cyclic water condensation
Par ameter Value
Test Mater ial API 5L X65
Total Pr essur e (bar ) 1
H 2S Par tial Pr essur e (mbar ) 0.3
O 2 Par tial Pr essur e (mbar ) 0
CO 2 Par tial Pr essur e (bar ) 0.96
N2 + Ar Par tial Pr essur e (bar ) 0
Test Dur ation (days) 3, 7
Water Condensation Rate (ml/m 2/s) 0.015, 0
Steel Temper atur e (oC) 25, 40
Gas Temper atur e (oC) 30
Temper atur e Cycle Per iod (days) 1 (change every 12 hours)
78
SEM surface analysis was performed after each experiment to examine the effect
that experienced cyclic water condensation are compared to their constant WCR
counterparts in Figure 24. Overall, FeS spalling was more apparent in the constant WCR
experiments than the cycling experiments. The less sheet-like characteristics of the cycled
corrosion product layers, as compared to the layers formed under constant water
condensation rates, is likely a result of high FeS precipitation rates as water was quickly
removed from the steel surface. It is postulated that precipitated corrosion product would
have settled on layers similar to those observed from the constant WCR experiments,
thereby reinforcing the structural integrity of the layer making spalling less pronounced.
79
determined by X-ray diffraction is shown in Figure 25. Mackinawite was the primary
corrosion product detected, but siderite was also detected in the corrosion product
developed after 7 days under cycling conditions. Peak broadening is apparent in the
temperature cycled experiments. It is postulated that the fast precipitation rates caused by
the drying process results in growth of small mackinawite crystals or induces lattice strain
80
broadening.
Figure 25: Effect of cyclic water condensation on the corrosion product composition.
81
(a)
(b) (c)
Figure 26: Effect of temperature cycling on (a) corrosion rate, (b) measure and theoretical
corrosion product mass, and (c) maximum possible black powder production rate in a 100
The effect of temperature cycling on corrosion rate, corrosion product mass, and
maximum possible black powder production rate is illustrated in Figure 26. As would be
expected for the case of a steel surface water wetted for about half the experimental
duration, the corrosion rates of the cycled specimens were less than those from the constant
WCR conditions. However, while the corrosion rate from the 3 day-cycled experiment was
about half of its constant WCR counterpart, the corrosion rate from the 7 day – cycled
experiment was 36 ± 4 % of the 7 day – constant WCR corrosion rate suggesting that the
solid corrosion product produced during the drying phase is more protective than the
conditions favorable to corrosion product precipitation and iron sulfide film forming
conditions. The precipitated corrosion product formed in oxic conditions consisted of two
layers. The outer layer is the more likely layer to detach and form black powder as the inner
layer is attached to the steel with no evidence of inner layer delamination. The outer layer
appeared only loosely held to the steel surface with portions projected away from the
surface, which would make the layer more susceptible to removal by a flowing gas stream.
The bilayered morphology is postulated to only occur at the top of the pipeline (i.e., 12
o’clock position) where iron carbonate, which nucleated away from the steel surface, can
collect at the gas/liquid interface. Away from the top of the line, gravity will likely draw
precipitated products back to the steel surface to merge with the inner layer.
83
under constant water condensation rates. These thin films were likely a results of the direct
reaction of H2S to the steel, which forms a framework for iron sulfide precipitation as per
Zheng’s model of sour corrosion [21]. The spallation of these thin films occurs as intrinsic
stresses, which accumulate during layer growth, then relax during the buckling process. A
gas pipeline experiencing water condensation would likely have to contend with this
method of black powder formation over the entire pipe wall as there is no bilayered
morphology. It appears the only requirement to form these types of layers is a water-wetted
and hematite. Of the three, mackinawite is metastable and its transformation to more
appears corrosion product formed through precipitation served to structurally reinforce the
iron sulfide layer. No major change in corrosion product composition was obtained due to
temperature cycling.
To mitigate the formation of black powder due to dew formation within the
pipeline, proper gas dehydration is absolutely necessary. Dehydrating the gas to below
100% RH, however, may not be completely effective. The potential for corrosion below
corrosion in water unsaturated conditions, which can lead to black powder formation. The
100% was investigated by depositing NaCl on polished steel specimens and exposing the
specimens to differing H2S partial pressures and relative humidities. The test matrix that
Surface microscopy of the salt layer, generated by drying a 3.5 wt% NaCl/deionized
water solution onto a steel surface polished to a mirror finish, is shown in Figure 27. The
85
solution was purged with N2 prior to deposition, and drying was performed by passing dry
N2 over the steel specimen and allowing the water to evaporate, slowly.
Figure 27: SEM analysis of the NaCl layer formed before specimen insertion into test
environment. The specimen was coated with palladium to minimize electron beam
charging.
As seen in Figure 27, the salt was predominantly grouped into roughly cubic
structures approximately 100 µm in width. A slight salt residue is apparent on the steel
The salt layer generation method produced some pitting on the outside of the border
of crystals, as shown in Figure 28. The micrographs shown in Figure 28 were taken after
the rinsing the salt layer shown in Figure 27 with N2-purged DI water. The methodology
employed during salt generation, sputter coating, SEM, and salt removal of the salt layer
86
in the micrographs is the same as during experimentation, therefore the attack morphology
would propagate to the experiments, whose results are shown in subsequent sections.
Attack beyond that shown in Figure 28 is a result of exposure to the atmosphere dictated
(a) (b)
Figure 28: Slight localized corrosion on the outside boundary of a salt crystal formed by
drying NaCl.
conditions (75% RH) is shown in Figure 29. The occurrence of deliquescence is evident at
75% due to the small size of the cubic salt crystals relative to those seen in Figure 27.
Distinct regions were found under deliquescing conditions, which are the darker and the
lighter regions seen in Figure 29(a) and better detailed in the enlarged micrograph in Figure
29(b). More corrosion product is found in the darker shaded regions, whereas coverage was
87
sparser in the circular region where islands of corrosion product were obtained amidst a
material structurally consistent with iron carbide. EDS analysis of a similar region shown
carbonate, which was also more generally detected by XRD, as shown in Figure 31.
(a) (b)
Figure 29: (a) Surface microscopy after a 3 day exposures to CO2 atmosphere at 75% RH.
Figure 30: EDS of corrosion product after a 3 day exposure to a CO2 environment at 75%
Figure 31: XRD corrosion product composition analysis of specimens recovered after a 3
day exposure to CO2 at 75% RH. Peak labels are S: siderite (FeCO3), H: halite (NaCl),
The reason for the non-uniform corrosion product is unclear but is attributed to the
presence of a large salt crystal that once covered the circular area. The circular shape
compared to the typical cubic shape is due to the rounding of the NaCl crystal edges during
Significant corrosion took place beneath the salt crystal as evident by the
approximately 1 µm difference between regions beneath salt crystals and the surrounding
regions in the surface profilometry shown in Figure 32. Deeper localized corrosion was
also observed typically beneath the smaller crystals than the larger crystals.
90
Figure 32: Surface profilometry of specimen recovered after a 3 day exposure to CO2 at
Corrosion product was not directly observable by SEM after a 3-day exposure to
CO2 at 58% and 33% RH, as shown in Figure 33. The large NaCl crystals were still present
at 58% and 33% RH indicating deliquescence did not occur, however, a ring surrounding
the salt crystal is apparent at 58% RH. Also, no corrosion product was detected by XRD
Figure 33: Surface microscopy and XRD compositional analysis after a 3 day exposure to
The effect of relative humidity on the general corrosion rate due to hygroscopic
corrosion in CO2 atmospheres is shown in Figure 34(a). As expected, the corrosion rate
increased with the availability of water in the atmosphere; however, the corrosion rates
were all very small. The corrosion rate at 33% RH cannot be distinguished from no
corrosion due to its small value and gravimetric measurement uncertainty. The small
corrosion rates are partially owed to the nonuniform coverage of the salt layer making
92
corrosion rate is still useful for comparison purposes, so the value is still provided.
Figure 34: Effect of relative humidity on (a) general corrosion rate, (b) measured and
theoretical corrosion product mass, and (c) maximum possible black powder production
The effect of relative humidity on the corrosion product mass and black powder
production rate is shown in Figure 34(b) and Figure 34(c), respectively. Below the DRH,
the measured corrosion product mass was higher than the theoretical corrosion product
mass. This is believed to be due to the presence of NaCl in the corrosion product that could
shown in Figure 35. Flakey corrosion products potentially susceptible to spallation were
only obtained at 0.3 and 1.0 mbar H2S where the corrosion product appeared to create a
shell around the location once occupied by a salt crystal. Corrosion product formation atop
of the salt crystal is postulated to be due to ferrous ion diffusion through the water layer
surrounding the deliquescing salt crystal. FeS precipitation then occurred with sulfides
already present. The shell of corrosion product would then be all that remains once the salt
crystal fully deliquesces. No evidence of a friable corrosion product was found at 0.1 mbar
H2S.
94
The effect of H2S partial pressure on the composition of corrosion products formed
under deliquescing conditions is shown in Figure 36. With the increase of H2S partial
pressure to 0.1 mbar, mackinawite was the dominant phase, however the [001] peak at
17.6° 2θ was quite broad. At 0.3 mbar and 1.0 mbar H2S, both siderite and mackinawite
were present with an additional unknown phase(s). The baseline of the diffraction patterns
95
collected at 0.3 and 1.0 mbar H2S were high indicating the presence of amorphous or
nanocrystalline material.
96
Figure 36: Effect of H2S partial pressure on corrosion product composition. Phases are
M: mackinawite, S: siderite, H: halite, and Fe: ferrite. Peaks labeled with an asterisk (*)
Figure 37: Effect of H2S partial pressure on surface morphology at 58 and 33% RH.
98
The effect of H2S partial pressure at relative humidities below the DRH on surface
micrographs, however XRD detected the presence of mackinawite at 0.3 and 1.0 mbar H2S
at 58% RH as shown in Figure 38. As was the case in their 75% RH analogous experiments,
the mackinawite peaks were very broad. No corrosion product was detected by XRD at
Figure 38: Effect of H2S partial pressure on corrosion product composition. Phases are
Figure 39: Effect of H2S partial pressure on (a) general corrosion rate, (b) measured and
theoretical corrosion product mass, and (c) maximum possible black powder production
The effect of H2S partial pressure on the general corrosion rate, corrosion product
mass, and maximum possible black powder production rate is shown in Figure 39. At 0.3
and 1.0 mbar H2S, the corrosion rates were close to those seen in dewing conditions,
however, the corrosion rates saw a large drop as the RH fell below the deliquescence
relative humidity (DRH) of NaCl (75% RH). Corrosion was measurable at 0.3 mbar H2S
and 33% RH, but the corrosion rates at 0, 0.1 and 1.0 mbar and 33% RH measured could
not be distinguished from a corrosion rate of zero due to measurement uncertainties. The
measured corrosion product mass was consistently higher than the theoretical mass, but
this is likely due to the presence of NaCl in the measured corrosion product that was not
quantified.
Based on the results above, it is apparent that corrosion occurred below the DRH,
Schindelholz, et al., [40] examined the effect of relative humidity on the corrosion of steel
beneath NaCl particles in oxic environments. Corrosion below the DRH was attributed by
Schindelholz, et al., to adsorbed water on both the salt and the steel, deliquescence-
The initiation of corrosion due to the hygroscopic salt particles was attributed to
the presence of adsorbed water on both the salt and the steel surface and capillary
condensation at the gas/salt/metal interface. Both the salt and the steel surface can adsorb
water. NaCl has been shown to produce adsorbed water layers with properties similar to
bulk water at a relative humidity of 40-50% [52]. Similarly, water adsorption on iron was
102
found to lead to a potentially corrosive water layer at about 50% RH [53]. The occurrence
of corrosion at 33% RH in the oxic environments tested was attributed to the interaction of
the two water layers on the salt and the steel through capillary condensation at the confined
crystallization kinetics. In short, the driving force for crystallization at the DRH is
frequently insufficient for nucleation and crystal growth, therefore a lower RH for
efflorescence is necessary. However, the presence of corrosion products and metal surfaces
In the limited water available for corrosion to occur, a small amount of corrosion
can significantly impact the water chemistry. The addition of ferrous ions to water can
depress the effective DRH allowing corrosion to occur at even lower RH values than those
FeCl2·4H2O, which has a DRH of approximately 55% [55], may indeed allow for
corrosion of archaeological iron artifacts down to 20% RH and induce the “weeping” of
the artifacts, whereby a highly acidic solution with high Fe2+ and Cl– concentrations is
additional species leading to locally decreased RH values is potentially important for black
powder production through hygroscopic corrosion processes as pipeline conditions are not
static. Corrosion beneath a salt crystal present on the steel surface may initiate while gas
conditions are relatively humid, but the change in chemistry and deliquescence-
efflorescence hysteresis could sustain deliquesced water even if bulk gas RH is decreased.
For black powder to form via corrosion at RH values below 100%, the species on
the surface must assist in the accumulation of moisture. NaCl deposits on the steel surface
were found to lead to corrosion and corrosion product formation at relative humidities as
low as 33%, however, corrosion products that can potentially lead to black powder were
deliquescence makes black powder prediction computationally simpler. Rather than trying
to predict the corrosion due to adsorbed water layers on salt particles of variable size, one
can assume full deliquescence is needed thereby allowing the thermodynamics of salt
deliquescence in the particular system to be the deciding factor in the prediction of black
powder formation. However, this still requires some knowledge of the composition of
materials present on the steel surface. A sea salt residue will have a different DRH than a
pure NaCl residue. Luckily, the thermodynamics of salt deliquescence is relatively well
The experiments discussed above do not necessarily negate the potential for black
powder to form without the presence of a hygroscopic salt. Other materials such as glycols
104
or corrosion products may assist in the wetting of the steel surface and sustaining corrosion.
present upon a steel surface. A porous corrosion product may lead to water accumulation
RH values much less than 100% due to increased intermolecular interactions within the
Glycols, especially triethylene glycol (TEG), are commonly used to dry natural gas
to specification. Glycol carryover from gas processing plants can lead to corrosion and
possibly black powder production through co-condensation of glycol and water and/or
through the hygroscopic uptake of water by glycol. Deposited TEG/H2O mixtures of >95%
TEG in a 0.4 mbar H2S/CO2 (1 bar total pressure) atmosphere were found to sustain
corrosion rates of 0.6 – 30 µm/year [58]. For reference, a 1 µm/yr general corrosion rate
over a 100 km, 42 in. ID pipeline would yield a maximum possible black powder
production rate of approximately 10 kg/day. Mixtures of TEG with black powder and other
residual components will result in a black sludge, which requires pigging for it to be
removed [59]. If this sludge is present within the pipeline or glycol carryover is suspected,
then a thorough cleaning with methanol or another solvent is recommended before inhibitor
application.
Great care was taken to ensure consistency in laboratory practices so that repeatable
results were obtained, but the results herein are generally from one experiment per set of
difficulties with H2S environments, limited the number of experiments that could be
conducted for corrosion rate and corrosion product mass measurements. However, the
features observed in the above microscopy were consistent with those seen in experiments
with inconclusive gravimetric measurements though otherwise fine. SEM images from
those tests and additional locations in the specimens shown above are shown in Appendix
C.
106
Under dewing conditions, the two corrosion product morphologies were identified
which can lead to black powder production. The first is formed due to corrosion product
mackinawite films that buckle and spall due to compressive stress generation due to layer
growth.
Potential was found for black powder to form through hygroscopic corrosion if
deliquescence occurs. Without deliquescence the corrosion product was minimal due to a
The hypotheses formulated at the start of this study are revisited here in light of the
form on the steel surface although kinetically favored compounds will also be
encountered.
The composition of the corrosion product did indeed change with the differing
conditions tested. Mackinawite was the predominant phase detected under sour conditions,
iron carbonate was found under sweet conditions, and a mix of iron carbonate and hematite
phase as compared to the more thermodynamically favored pyrite. The preference for
107
mackinawite is likely due to the temperatures being too low for meaningful phase
With regards to black powder formation, the corrosion product composition seemed
to have some effect on the propensity of the layer to spall. Buckling was readily observed
in sour conditions, and loose layers were found in the oxic environments. However,
corrosion products which form through a precipitation-only mechanism did not seem to be
direct reaction with the steel and precipitation. Precipitation directly onto the steel surface
appeared to have developed a much more adherent corrosion product layer than the
The formation of corrosion products follows a cyclical pattern that correlates with
The cyclic behavior of water condensation and evaporation in sales gas pipelines
Corrosion and subsequent corrosion product formation did indeed seem to have
correlated with the wet/dry cycle based on the 50% reduction in corrosion rate after 3 days
in a temperature cycled condition compared to the 3 days, constant WCR experiment, but
the cyclic water condensation seemed to have an opposite effect than hypothesized on the
development of friable corrosion product layers. Less buckling was observed in the cycled
108
conditions than the constant temperature due to corrosion product precipitation upon
Corrosion products will form on the steel surface in non-water saturated sales gas
The development of corrosion products did indeed occur under non-water saturated
conditions due to the presence of hygroscopic salts on the steel surface. Corrosion products
Black powder mitigation efforts should emphasize proper gas dehydration as the
primary means of black powder mitigation. However, periodic upsets may allow water to
condense despite best efforts in dehydration. As a means of insurance against black powder
formation if liquid water was to condense, the use of corrosion inhibitors, particularity
volatile corrosion inhibitors (VCIs), should minimize the effects of corrosion and therefore
limit the production of resulting black powder. Finding an appropriate VCI to mitigate
dewing conditions.
Determining the role of iron sulfide growth mechanisms (i.e., epitaxial growth,
REFERENCES
[2] U.S. Energy Information Administration, How much carbon dioxide is produced per
kilowatthour when generating electricity with fossil fuels? [Online]. Available:
http://www.eia.gov/tools/faqs/faq.cfm?id=74&t=11. [Accessed: 13-Nov-2016].
[4] R. Baldwin, “Black Powder in the Gas Industry - Sources, Characteristics and
Treatment,” Mechanical and Fluids Engineering Division Southwest Research
Institute, San Antonio, TA 97-4, 1997.
[9] T. Tran, B. Brown, and S. Nesic, “Corrosion of Mild Steel in an Aqueous CO2
Environment–Basic Electrochemical Mechanisms Revisited,” in
CORROSION/2015, Houston, TX, 2015, Paper No. 5671.
[12] D. A. Palmer and R. Van Eldik, “The chemistry of metal carbonate and carbon
dioxide complexes,” Chem. Rev., vol. 83, no. 6, pp. 651–731, 1983.
[16] W. Sun, S. Nešić, and R. C. Woollam, “The effect of temperature and ionic strength
on iron carbonate (FeCO3) solubility limit,” Corros. Sci., vol. 51, no. 6, pp. 1273–
1276, Jun. 2009.
[18] W. Sun and S. Nešić, “Kinetics of Corrosion Layer Formation: Part 1—Iron
Carbonate Layers in Carbon Dioxide Corrosion,” CORROSION, vol. 64, no. 4, pp.
334–346, Apr. 2008.
[21] Y. Zheng, J. Ning, B. Brown, D. Young, and S. Nesic, “Mechanistic Study of the
Effect of Iron Sulfide Layers on Hydrogen Sulfide Corrosion of Carbon Steel,” in
CORROSION/2015, Houston, TX, 2015, Paper No. 5933.
[24] A. Sherik, “Black Powder in Gas Transmission Pipelines,” in Oil and Gas Pipelines,
R. W. Revie, Ed. John Wiley & Sons, Inc., 2015, pp. 423–436.
[25] M. G. Fontana, Corrosion Engineering, 3rd ed. New York, NY: McGraw-Hill
Education, 1986.
[27] B. Lavina et al., “Structure of siderite FeCO3 to 56 GPa and hysteresis of its spin-
pairing transition,” Phys. Rev. B, vol. 82, no. 6, Aug. 2010.
[29] D. Rickard and G. W. Luther, “Chemistry of Iron Sulfides,” Chem. Rev., vol. 107,
no. 2, pp. 514–562, Feb. 2007.
[30] J. Yamada, H. Kaneta, and K. Nakayama, “Analysis of Black Powder in Natural Gas
Pipelines,” in CORROSION/2011, Houston, TX, 2011, Paper No. 11088.
[32] B. Craig, “Corrosion Product Analysis - A Road Map to Corrosion in Oil and Gas
Production,” Mater. Perform., vol. 41, no. 8, pp. 56–58, 2002.
[37] N. Yaakob, F. Farelas, M. Singer, S. Nesic, and D. Young, “Localized Top of the
Line Corrosion in Marginally Sour Environments,” in CORROSION/2016,
Houston, TX, 2016, Paper No. 7695.
[39] L. Ojha et al., “Spectral evidence for hydrated salts in recurring slope lineae on
Mars,” Nat. Geosci., vol. 8, no. 11, pp. 829–832, Sep. 2015.
[41] J. Kolts, “Design for Internal Corrosion Resistance of Sales Gas Pipelines,”
presented at EUROCORR, Nice, France, 2004.
[43] “Standard Practice for Preparing, Cleaning and Evaluating Corrosion Test
Specimens,” ASTM Standard G1-03, 2011.
[52] M. C. Foster and G. E. Ewing, “Adsorption of water on the NaCl(001) surface. II.
An infrared study at ambient temperatures,” J. Chem. Phys., vol. 112, no. 15, pp.
6817–6826, 2000.
[53] S. Lee and R. W. Staehle, “Adsorption of Water on Copper, Nickel, and Iron,”
CORROSION, vol. 53, no. 1, pp. 33–42, Jan. 1997.
[58] A. Dugstad and H. Sirnes, “Experimental Study of Black Powder in field TEG
solutions,” in CORROSION/2011, Houston, TX, 2011, Paper No. 11083.
[59] A. M. Sherik, A. Rasheed, and A. Jabran, “Effect Of Sales Gas Pipelines Black
Powder Sludge on the Corrosion of Carbon Steel,” in CORROSION/2011, Houston,
TX, 2011, Paper No. 11087.
control system. The system was built around a TC-48-20 PID (Preportional, Integral,
heating and cooling control modes, 2 thermistor inputs, a pulse-width modulated output, 2
temperature alarms, and a computer interface through serial RS-232 communications. The
two alarms included in the controller monitor their respective thermistor input and activate
activates and allows electrical current to pass to ground completing a circuit. The alarms
were designed for controlling fans or other loads. For instance, if a high temperature is
measured on the Peltier heatsink by the secondary thermistor, then a fan on the heat sink
Figure 40. Specimen temperature is measured with the primary thermistor (10 kΩ at 25°C)
held in contact with the specimen. The thermistor bulb is mounted through a hole in the
PEEK specimen holder and held in place with hot glue. The hot glue acts as both support
for the thermistor and as thermal insulation isolating the sensor from external influences.
The PID controller computes the pulse-width modulation (PWM) duty cycle required to
maintain the temperature setpoint which modulates the 12 VDC power supply voltage to
A limitation of the TC-48-20 Peltier controller is that the positive and negative
Peltier terminals are fixed preventing the polarity change required for temperature cycling
installed between the controller and the Peltier in the configuration shown in Figure 40,
above. Break-before-make contacts are required in the relay to prevent short circuits. A
diode was placed in a reverse bias orientation to minimize the effects of voltage spikes due
to a collapsing magnetic field in the relay’s inductor. The relay is activated by triggering
the secondary alarm which allows current to pass through the relay inductor thereby
activating the relay. Since measurements are required to activate the alarm, the secondary
117
thermistor was replaced with a 10kΩ resistor to measure a constant 25°C secondary
temperature. By changing the secondary alarm temperature to above or below 25°C, then
Since the Peltier controller was to be used in an H2S laboratory, the controller was
housed in a sealed steel box which isolates the controller and potential sparks within the
box from a potentially flammable environment. Panel mount connectors were installed in
the box door to provide a connection from the controller to the Peltier, primary thermistor,
and the computer kept outside the H2S laboratory. The serial RS-232 connection and
software included with the controller allowed for specimen temperature monitoring and
microcontroller. As the microcontroller cannot supply the electrical power to the Peltier, a
MOSFET H-bridge motor driver (Pololu High-Power Motor Driver 18-25) was installed
between the Peltier and the power supply for power modulation and Peltier polarity control.
The Arduino controller acts in a similar manner as the TC-48-20 controller in that a PID
algorithm computes the duty cycle for the PWM output. The AC output is passed to the
MOSFET gates in the motor driver along with a binary signal which determines the polarity
Specimen size limitations in the X-ray diffractometer prevented the use of the large
cylindrical specimens the dewing corrosion apparatus was designed around, so a new
holder was developed to hold smaller specimens within a body close to the same size as
118
the cylindrical specimens. A cross-sectional view of the apparatus is shown in Figure 41.
The specimen holder holds a 0.5×0.5×0.08 inch³ (12.7×12.7×2 mm³) specimen within a
thick (31.7 mm × 15.9 mm) cylindrical specimen. Aluminum was chosen for its thermal
conductivity and corrosion resistance. The specimen was coated with Xylan to prevent
galvanic effects between the aluminum and the steel. A magnet within a stainless steel part
embedded within the aluminum holds the specimen to the holder. Threads in the aluminum
body and the stainless steel part allow for specimen height adjustment. Lubrication was
applied to the threads to prevent galling. A hole slightly offset from center was drilled to
place a thermistor near the steel specimen for temperature control. The aluminum body
was fitted into a PEEK sleeve designed to provide a seal with the stainless steel glass cell
lid. O-rings were embedded within the body and PEEK sleeve to prevent gas leakage.
Figure 41: Cross-section view of the XRD specimen holder. Image courtesy of Cody
Shafer.
119
APPENDIX B: CALCULATIONS
Three primary specimen mass measurements were taken for each specimen: (1)
initial mass, 푚 (g); (2) post-extraction mass, 푚 (g); and (3) mass after corrosion product
removal, 푚푓 (g). The procedures for calculating the general corrosion rate, measured
specimen mass, and theoretical specimen mass from those measurements are detailed in
Corrosion Rate
푚 푚푓
CR 365 (53)
푡
휌st l 퐴s
where:
The measured corrosion product mass is the mass of corrosion product removed
per unit area. The measured corrosion product mass, in g/m2, is calculated with Equation
54:
푚 푚푓
푚m as (54)
퐴푠
120
The theoretical corrosion product mass is less straightforward to calculate than the
measured corrosion product mass. The theoretical corrosion product mass is the mass of
corrosion product that would be produced based on the mass of steel lost due to corrosion.
The X65 steel includes alloying elements (i.e., C, Mn, Cu, Ni, etc.), so the measured
corrosion product mass should account for these species. The steel components are grouped
into three categories: iron, carbon, and other (which includes all other alloying elements).
The mass lost due to corrosion in these three categories is calculated by multiplying the
total mass lost due to corrosion by the mass fraction (푥) of that category:
∆푚 푚 푚푓 (55)
푚 푥 ∆푚 (56)
푚 푥∆푚 (57)
Carbon within the steel is assumed to all be present as iron carbide (Fe3C).
Realistically, some carbon will be interstitially intercalated through the ferrite, however
this mass is assumed to be small and therefore negligible. The mass of iron carbide is
therefore determined by Equation 59. The mass of iron available for iron carbonate, iron
sulfide, or iron oxide production is thereby the difference in the total mass of Fe and the
mass of Fe within the iron carbide (Equation 60). 푀 is the molar mass of species 푖
.
푚푀
푚 (59)
푀
푚 푀
푚 , P 푚 3 (60)
푀
121
which the corrosion product is developed. The corrosion product with identity 푗is assumed
to be Fe2O3 in oxic environments, FeS in sour conditions, and FeCO3 in sweet conditions.
The mass of the iron corrosion product with 푛moles of Fe per mole 푗produced is calculated
as follows:
푚 , P푀
푚 (61)
푛푀
The total theoretical mass of corrosion product per unit area is calculated by
summing the individual corrosion product components and dividing by the specimen area
as in Equation 62.
푚 푚 푚ot h r
푚t h or (62)
퐴s
(1.07 m) ID, 100 km pipeline with area 퐴pip for the experimental duration to determine
푚t h or 퐴pip
푚̇ BP 10− (63)
푡
Error Analysis
Corrosion rate and corrosion product mass measurements are highly dependent upon
AB204-S, which has a readability and repeatability (standard deviation) of 0.1 mg. The
uncertainty in the mass measurement was propagated throughout the corrosion rate and
corrosion product mass measurements with the assistance of the Python package
122
Uncertainties [60]. The Uncertainties package uses linear error propagation theory to
calculate the standard deviation of a mathematical expression with inputs that include a
probability distribution function. The calculated results were reported using a 95%
confidence interval.
123
Many more SEM images were taken than those shown in Chapter 5. The images
selected for viewing in the main body of the text were selected based on how representative
the micrograph is of the specimen at large and their content in relation to understanding
black powder formation. Addition SEM surface microscopy is shown in this appendix for
context. The micrographs are grouped by experiment condition and, unless specified
Figure 42: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, WCR: 0.05 ml/m2/s, Steel
(a) (b)
(c) (d)
(e) (f)
Figure 43: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Temperature: 25°C, Gas Temperature: 30°C, Duration: 3 days. Images (a) and (b) were
from the same test as that shown in Chapter 5, whereas images (c)-(f) were from a test
125
Figure 44: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, WCR: 0.05 ml/m2/s, Steel
Figure 45: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Figure 46: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.05 ml/m2/s, Steel
Figure 47: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Figure 48: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Figure 49: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Figure 50: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Figure 51: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Figure 52: Conditions: pH2S: 1.0 mbar, pCO2: 0.96 bar, WCR: 0.05 ml/m2/s, Steel
Figure 53: Conditions: pH2S: 1.0 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Figure 54: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015 ml/m2/s, Steel
Figure 55: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015/0 ml/m2/s, Steel
Temperature: 25/40°C, Gas Temperature: 30°C, Duration: 3 days, Cycle Period: 1 day.
135
Figure 56: Conditions: pH2S: 0.3 mbar, pCO2: 0.96 bar, WCR: 0.015/0 ml/m2/s, Steel
Temperature: 25/40°C, Gas Temperature: 30°C, Duration: 7 days, Cycle Period: 1 day.
136
(a) (b)
(c) (d)
(e) (f)
Figure 57: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, RH: 75%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. (d) and (e): after DI water rinse, (f):
after Clarke.
137
(a) (b)
(c)
Figure 58: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, RH: 58%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. Micrograph (a): after extraction, (b):
Figure 59: Conditions: pH2S: 0 mbar, pCO2: 0.96 bar, RH: 33%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. Micrograph (a): after extraction, (b):
Figure 60: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, RH: 75%, Temperature: 25°C,
Figure 61: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, RH: 58%, Temperature: 25°C,
Figure 62: Conditions: pH2S: 0.1 mbar, pCO2: 0.96 bar, RH: 33%, Temperature: 25°C,
Figure 63: Conditions: pH2S: 0.3 mbar, pCO2: 0.97 bar, RH: 75%, Temperature: 25°C,
Figure 64: Conditions: pH2S: 0.3 mbar, pCO2: 0.97 bar, RH: 58%, Temperature: 25°C,
Figure 65: Conditions: pH2S: 0.3 mbar, pCO2: 0.97 bar, RH: 33%, Temperature: 25°C,
(a) (b)
(c) (d)
Figure 66: Conditions: pH2S: 1.0 mbar, pCO2: 0.97 bar, RH: 75%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. (c): after DI water rinse, (d): after
Clarke.
144
(a) (b)
(c) (d)
Figure 67: Conditions: pH2S: 1.0 mbar, pCO2: 0.97 bar, RH: 58%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. Micrographs (c) and (d): after DI water
rinse.
145
(a) (b)
(c) (d)
Figure 68: Conditions: pH2S: 1.0 mbar, pCO2: 0.97 bar, RH: 33%, Temperature: 25°C,
Duration: 3 days, Salt Layer Composition: NaCl. Micrographs (c) and (d): after DI water
rinse.
Thesis and Dissertation Services