Sunteți pe pagina 1din 162

State-of-Knowledge on Deposition, Part 2:

Assessment of Deposition Activity in Fossil


Plant Units

SED WARNING:
N
A L
LICE

Please read the License Agreement


on the back cover before removing
R I

M AT E
the Wrapping Material.
Technical Report

1.0

0.9

0.8

0.7
NORMALIZED RANK

SOURCES
0.6

SINKS
0.5

0.4

0.3

0.2

0.1

SUSPENDED SOLIDS
0.0 WATER SOLUBLE MATERIAL
TE SURFACE PHENOMENA ON SOLIDS
SA R
EN TE R
ND WA ILE
CO ED BO /RH
FE SH INE
RB
TU
R EVISED 07/06/2002
State-of-Knowledge on Deposition,
Part 2: Assessment of Deposition
Activity in Fossil Plant Units

1004930

Final Report, December 2003

EPRI Project Manager


K. Shields

EPRI 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10412, Palo Alto, California 94303 USA
800.313.3774 650.855.2121 askepri@epri.com www.epri.com
DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES
THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN
ACCOUNT OF WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH
INSTITUTE, INC. (EPRI). NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE
ORGANIZATION(S) BELOW, NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM:

(A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I)


WITH RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR
SIMILAR ITEM DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS
FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR
INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL
PROPERTY, OR (III) THAT THIS DOCUMENT IS SUITABLE TO ANY PARTICULAR USER'S
CIRCUMSTANCE; OR

(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER


(INCLUDING ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE
HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR
SELECTION OR USE OF THIS DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD,
PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT.

ORGANIZATION(S) THAT PREPARED THIS DOCUMENT

EPRI

Framatome ANP, Inc.

A. Aschoff

M. Ball

A. Bursik

ORDERING INFORMATION
Requests for copies of this report should be directed to EPRI Orders and Conferences, 1355 Willow
Way, Suite 278, Concord, CA 94520, (800) 313-3774, press 2 or internally x5379, (925) 609-9169,
(925) 609-1310 (fax).

Electric Power Research Institute and EPRI are registered service marks of the Electric Power
Research Institute, Inc. EPRI. ELECTRIFY THE WORLD is a service mark of the Electric Power
Research Institute, Inc.

Copyright 2003 Electric Power Research Institute, Inc. All rights reserved.
CITATIONS

This report was prepared by

EPRI
3412 Hillview Avenue
Palo Alto, California 94303

Principal Investigator
K. Shields

Framatome ANP, Inc.


155 Mill Ridge Road
Lynchburg, Virginia 24506-0935

Principal Investigators
W. Allmon
M. Pop

A. Aschoff
EPRI Consultant

M. Ball
EPRI Consultant

A. Bursik
EPRI Consultant

This report describes research sponsored by EPRI.

The report is a corporate document that should be cited in the literature in the following manner:

State-of-Knowledge on Deposition, Part 2: Assessment of Deposition Activity in Fossil Plant


Units, EPRI, Palo Alto, CA: 2003. 1004930.

iii
PRODUCT DESCRIPTION

Over the last 20 years, substantial advances have been made in the understanding and control of
fossil plant cycle chemistry. In spite of these advances, deposition activity, most notably in
boilers and steam turbines, remains an issue of concern to many organizations that operate fossil
units. The underlying science of deposition in fossil unit components has not, with the exception
of steam turbines, been studied extensively under the EPRI Boiler and Turbine Steam and Cycle
Chemistry Program. This report presents results of efforts to define the existing state-of-
knowledge and identifies information and data needs. EPRI intends to conduct research directed
towards improving the understanding of the science governing deposition in all parts of the
steam-water cycles of fossil units.

Results & Findings


The report reviews key parametersidentified in an earlier state-of-knowledge assessmentthat
need to be examined in future research. It discusses deposition experience in
condensate/feedwater systems, boiler waterside surfaces, steam tubing and piping, and steam
turbines of fossil plants. The approach followed has identified three basic deposition types and
analyzed related mechanisms and influencing factors applicable to each type. From over 40
parameters and factors that influence deposition, five key parameters were selected :(a) dynamics
of deposition and release (as it affects both boilers and turbines), (b) heat flux (primarily in boiler
waterwall tubes), (c) mass transport, (d) solubility in steam and water, and (e) surface finish.
Significant additional information was reviewed and summarized for later use in modeling the
deposition process. The findings substantiate EPRIs position that a better understanding of
deposition processes and the controllable factors influencing deposition is needed. Planned
research in deposition will be focused on boiler waterwalls and will include assessment and
possibly development of a computer model applicable to both drum-type and once-through
boilers.

Challenges & Objectives


Two factorsplant operating experience and prior research findingsmust be considered to
develop an understanding of the state-of-knowledge for deposition in fossil plants sufficient to
identify key influencing parameters. Both factors must be evaluated and assessed for their
applicability to modern fossil plants. Much of this work has general relevance but not direct
relevance. As an example, there has been considerable research directed towards understanding
deposition in nuclear steam generators. While portions of this work are relevant to deposition
processes in fossil boilers, it is clear that portions are not relevant and that data gaps, owing to
differences in design and operating environment, exist. In planning future research, it is
important to establish these areas of deficiency and include activities to resolve them.

v
Applications, Values & Use
Deposits and scales formed on heat transfer surfaces of fossil plant equipment are detrimental
because they reduce unit efficiency and increase heat rate. In boiler waterwall tubing, a buildup
of waterside solids is a prerequisite to activation of certain underdeposit corrosion mechanisms.
More extensive solids accumulation can lead to overheating damage and tube failures.
Deposition in steam turbines can lead to performance penaltiesreductions in capacity and/or
stage efficiencyor corrosion damage, depending on characteristics of the depositing material.
Deposit removal, usually by chemical cleaning and preferably done before deposits adversely
impact fossil unit operations, is costly and negatively impacts unit availability. With a better
understanding of basic deposition processes, it should be possible to improve the means by
which deposition activity is controlled and its effects are managed.

EPRI Perspective
A substantial portion of chemistry-related availability losses in fossil plants involve deposition of
solids on surfaces. Existing EPRI cycle chemistry guidelines are effective in minimizing
deposition but do so indirectly by controlling the chemistry to reduce impurity ingress and
corrosion product transport to acceptable levels. Continuing problems with deposition-related
availability losses in fossil plants underscore the need to focus on the science of deposition to
establish better criteria and tools for use by plant operating personnel. This state-of-knowledge
report for deposition in fossil plant units represents the first step in this process. Further
assessment of the state-of-knowledge for deposition will be made as needed to verify that the
scope of future research is appropriate.

Approach
A team of recognized fossil plant water chemistry experts evaluated available information on
deposition and identified relevant mechanisms, processes, and influencing factors. This
information was presented in an earlier report, State-of-Knowledge on Deposition: Part 1:
Parameters Influencing Deposition in Fossil Units (EPRI report 1004194). In the subject report,
the team considered deposition by breaking the steam-water cycle down into discrete sections.
Plant experience in each area of the cycle was evaluated, and data needs that would improve the
state-of-knowledge were identified.

Keywords
Deposition
Fossil plant availability
Boiler tube failures
Chemical cleaning
Turbine performance
Cycle chemistry

vi
ABSTRACT

Deposition in fossil plant cycles is probably the least understood aspect of operational cycle
chemistry applicable to these units. Previously, the primary emphasis has been to develop
working guidelines which, if properly selected and optimized, will serve to minimize impurity
ingress, corrosion and chemical transport. This approach has proven itself to be effective in
minimization of deposition activity, albeit by indirect means.

Development of a state-of-knowledge was conceived as the initial step in gaining an improved


understanding of deposition phenomena. A complete understanding of the subject represents an
ambitious undertaking indeed as it requires consideration of a large cross-section of the
chemical, physical science and engineering disciplines. Further, it requires that existing
information on deposition from the areas of industrial steam-water cycles, nuclear plant cycles,
and geochemistry, among others, be considered. Then it is necessary to overlay this body of
accumulated information on the deposition problems experienced in fossil plant cycles to
determine where the problems lay and if more research would be able to improve the situation.

In an earlier report on deposition in fossil plant cycles, the emphasis was placed on identification
of parameters that influence deposition, and a preliminary assessment of the significance of these
parameters in working plants. The resulting list of key parameters was compared to current
industry needs, using the latest economic and boiler tube failure statistics available and the short
and long range strategy of EPRI to make deposition activity more easily managed in operating
fossil units. In this second state-of-knowledge report, the deposition phenomena typically
observed in fossil power cycles is presented and research needs are identified. Key parameters
were selected because of their apparent, even if not fully understood, influence on one or more of
the three deposition process types. These key parameters are (a) dynamics of deposition and
release (as it affects both boilers and turbines), (b) heat flux (primarily in boiler waterwall
tubes),(c) mass transport,(d) solubility in steam and water and (e) surface finish (primarily in
turbines and steam generator water tubes). Additional parameters of interest were also identified
for possible further study.

The next step in improving the understanding of deposition should include deposition modeling
activity, starting with boiler waterwalls. The goal of the model development would be to create a
tool that would apply to new and existing fossil units to better manage any deposition activity
that cannot be avoided and, in so doing, eliminate or minimize the negative impacts of deposition
on performance and profitability. Optimization of fossil plant cycle chemistry in accordance with
the key cycle chemistry guidelines as it relates to the above processes should be a top priority
since doing so will serve to minimize the probability of significant deposition developing in a
power plant with a given material selection and stress distribution.

vii
ACKNOWLEDGMENTS

The authors gratefully acknowledge the advice and suggestions of the following personnel in
defining the initial compilations of deposition process types, mechanisms and parameters that
influence deposition activity in fossil plant units:
R. B. Dooley, EPRI
L. S. Lamanna, Framatome ANP, Inc.
D. A. Palmer, Oak Ridge National Laboratory
M. Gruszkiewicz, Oak Ridge National Laboratory
A. Petrov, Oak Ridge National Laboratory and Moscow Electric Power Institute

ix
CONTENTS

1 DEPOSITION IN FOSSIL PLANT CYCLES ............................................................. 1-1


1.1 Introduction ........................................................................................................ 1-1
1.2 Factors Influencing Deposition Activity............................................................... 1-3
1.3 EPRI Strategy to Address Deposition State-of-Knowledge Deficiencies............ 1-9
1.4 Report Overview .............................................................................................. 1-10
1.5 References....................................................................................................... 1-11

2 DEPOSITION IN CONDENSATE AND FEEDWATER SYSTEMS ........................... 2-1


2.1 Introduction ........................................................................................................ 2-1
2.2 Locations Subject to Corrosion and Deposition ................................................. 2-2
2.3 Mechanisms....................................................................................................... 2-5
2.4 Control ............................................................................................................... 2-6
2.5 Diagnostics ........................................................................................................ 2-9
2.6 Factors Linked to the Matrices ......................................................................... 2-10
2.7 State-of-Knowledge Deficiencies ..................................................................... 2-12
2.8 References....................................................................................................... 2-12

3 DEPOSITION IN BOILERS....................................................................................... 3-1


3.1 General Aspects of Waterside and Steamside Deposition in Boilers ................. 3-1
Contamination Originating in Makeup Water and Cooling Water......................... 3-1
Corrosion Products .............................................................................................. 3-3
Effects of Deposition............................................................................................ 3-4
3.2 Locations Subject to Deposition......................................................................... 3-4
Water-Cooled Tubing........................................................................................... 3-4
Steam-Cooled Tubing.......................................................................................... 3-8
3.3 Deposits in Waterwalls..................................................................................... 3-10
Rippled Deposits in Waterwalls of Once-through Steam Generators ................ 3-10

xi
Porous Deposits in Waterwalls of Drum Boilers ................................................ 3-11
Copper Deposition (10) ......................................................................................... 3-12
3.4 Deposits in Superheaters and Reheaters ........................................................ 3-14
Copper Deposition in Primary Superheaters ..................................................... 3-14
Sodium Sulfate Deposits in the Reheater .......................................................... 3-14
3.5 Corrective Actions ............................................................................................ 3-15
Selection of Optimum Chemical Treatment ....................................................... 3-15
Estimation of the Amount, Composition and Porosity of Boiler Tube
Deposits............................................................................................................. 3-16
Implementation of Chemical Cleaning ............................................................... 3-17
3.6 Factors Linked to the Matrices ......................................................................... 3-17
3.7 State-of-Knowledge Deficiencies ..................................................................... 3-23
3.8 References....................................................................................................... 3-24

4 DEPOSITION IN SUPERHEATERS AND REHEATERS ......................................... 4-1


4.1 Introduction ........................................................................................................ 4-1
4.2 Locations Subject to Deposition......................................................................... 4-1
4.3 Deposit Mechanisms.......................................................................................... 4-2
4.4 Methods to Control or Prevent Deposition ......................................................... 4-7
4.5 Diagnostic and Corrective Actions ..................................................................... 4-8
4.6 Factors Linked to the Matrices ........................................................................... 4-8
4.7 State-of-Knowledge Deficiencies ....................................................................... 4-9
4.8 References......................................................................................................... 4-9

5 TURBINES................................................................................................................ 5-1
5.1 Introduction ........................................................................................................ 5-1
History of Investigations into Turbine Deposition and Damage ........................... 5-2
Overview of Steam Path Damage Mechanisms and Impact................................ 5-2
Operational Consequences ................................................................................. 5-6
5.2 Source, Composition and Location of Deposits ................................................. 5-8
General Aspects of Deposition ............................................................................ 5-8
Salt Deposits ..................................................................................................... 5-16
Copper Deposits ................................................................................................ 5-19

xii
Deposition Processes and Mechanisms ............................................................ 5-24
Volatility of Materials in Steam........................................................................... 5-30
Contaminant Solubility in Steam and Its Deposition .......................................... 5-43
5.3 Methods of Control and Prevention of Deposits............................................... 5-49
General Aspects of Minimizing Deposition ........................................................ 5-49
Copper Deposition ............................................................................................. 5-52
Turbine Washing and Chemical Cleaning.......................................................... 5-54
5.4 Matrices for Deposition and Adsorption ........................................................... 5-56
Factors Linked to the Matrices........................................................................... 5-56
Findings for Reducing the Deposition on Turbines ............................................ 5-61
5.5 State-of-Knowledge Deficiencies ..................................................................... 5-62
5.6 References....................................................................................................... 5-63

6 CONCLUSIONS AND RECOMMENDATIONS......................................................... 6-1


6.1 Conclusions ....................................................................................................... 6-1
Condensate and Feedwater Systems.................................................................. 6-2
Sources ........................................................................................................... 6-2
Deposition ....................................................................................................... 6-2
Corrosion......................................................................................................... 6-2
Composition .................................................................................................... 6-2
Impact.............................................................................................................. 6-3
Mechanisms: ................................................................................................... 6-3
Control............................................................................................................. 6-3
Boilers.................................................................................................................. 6-3
Deposition ....................................................................................................... 6-4
Corrosion......................................................................................................... 6-4
Composition .................................................................................................... 6-4
Impact.............................................................................................................. 6-5
Superheaters and Reheaters............................................................................... 6-6
Sources ........................................................................................................... 6-6
Deposition ....................................................................................................... 6-6
Corrosion......................................................................................................... 6-6
Composition .................................................................................................... 6-7

xiii
Impact.............................................................................................................. 6-7
Mechanisms .................................................................................................... 6-7
Control............................................................................................................. 6-7
Deficiencies ..................................................................................................... 6-8
Steam Turbines ................................................................................................... 6-8
Sources ........................................................................................................... 6-8
Deposition ....................................................................................................... 6-8
Corrosion......................................................................................................... 6-8
Composition .................................................................................................... 6-9
Impact.............................................................................................................. 6-9
Mechanisms .................................................................................................... 6-9
Control............................................................................................................. 6-9
Deficiencies ..................................................................................................... 6-9
6.2 Recommendations ........................................................................................... 6-10
6.3 References....................................................................................................... 6-11

xiv
LIST OF FIGURES

Figure 1-1 Locations of Deposition, Corrosion and Impurity Ingress in Fossil Plant Cycles ......1-2
Figure 1-2 Influence Diagram for Deposition Parameters..........................................................1-9
Figure 2-1 High Pressure Heater Tube at Magnification of 200x (70-30 Copper-Nickel
after 29 Years of Service) ..................................................................................................2-4
Figure 2-2 High Pressure Heater Tube at Magnification of 500x (Alloy 400, Monel, after
26 Years of Service)...........................................................................................................2-4
Figure 2-3 Example of Increased Iron and Copper in Feedwater During Plant Startup.............2-7
Figure 3-1 Typical Boiler Waterwall Tube Deposit.....................................................................3-5
Figure 3-2 Waterwall Tube Heat Transfer Profile ......................................................................3-6
Figure 3-3 Waterwall Tube Hot Side Crown Region ..................................................................3-7
Figure 3-4 Steam Formation in Porous Waterside Deposits......................................................3-7
Figure 3-5 Representative Drum Boiler Mechanical Carryover..................................................3-9
Figure 3-6 Enthalpy-Entropy Diagram for Steam Turbines and Solubility of Sodium
Sulfate (Na2SO4)...............................................................................................................3-10
Figure 3-7 Multilayered Deposit with Metallic Copper in Pit ...................................................3-13
Figure 3-8 Multilayered Deposit with Metallic Copper Plated on Metal Surface ......................3-13
Figure 3-9 Summary of Boiler Water and Feedwater Treatments for Fossil Plants.................3-16
Figure 4-1 Copper Deposits (Visible as Bright Spheres) on Superheater Tube Magnetite
Scale at Magnification of 500x. Material is SA213-T11 with 16 Years of Service. .............4-3
Figure 4-2 Crystalline Metallic Copper Deposits on Primary Superheater Tube at 2000x
Magnification. Material and Service Conditions are the same as in Figure 4-1 ................4-4
Figure 4-3 Reheater Tubes with Scale ......................................................................................4-5
Figure 4-4 Damage to Turbine Nozzle Blocks by Exfoliated Scale............................................4-6
Figure 4-5 Erosion Damage to Turbine Blades by Exfoliated Scale ..........................................4-6
Figure 5-1 Turbine Deposits. (a)Blades in the zone of initial condensation in an LP
turbine. Only the blade root shows a thin deposit of iron oxides. (b) Failure due to
corrosion fatigue. The cracks occurred near the transition between oxide deposits
and bare metal. ................................................................................................................5-10
Figure 5-2 Typical Locations of Turbine Corrosion, Erosion and Deposits. (In figure 5-2
the symbols have the following signification: P, pitting; CF, corrosion fatigue; SCC,
stress-corrosion cracking; C, crevice corrosion; G, galvanic corrosion; E, erosion; E-
C erosion-corrosion; SPE, solid particle erosion; chemicals-deposits.) ...........................5-11
Figure 5-3 Cross-Section of LP Turbine Showing Locations of Chemistry and Corrosion
Processes ........................................................................................................................5-12

xv
Figure 5-4 Solubility and Deposition Turbines .........................................................................5-13
Figure 5-5 Relative Distribution of Components in Typical Turbine Deposits ..........................5-14
Figure 5-6 Summary of Deposit Compositions as a Function of Specific Volume of
Steam...............................................................................................................................5-14
Figure 5-7 Chemical Transfer in the Conventional Drum Boiler System..................................5-15
Figure 5-8 Chemistry and Corrosion Effects of Layup, Startup and Cycling for a Drum
Boiler Cycle ......................................................................................................................5-15
Figure 5-9 Mollier Diagram with LP Turbine Steam Expansion Line .......................................5-17
Figure 5-10 Turbine Copper Deposits......................................................................................5-22
Figure 5-11 Fossil Drum Cycle Illustrating Key Copper Processes .........................................5-23
Figure 5-12 Solubilities of Iron and Copper Passing through a Turbine in Supercritical
Cycle ................................................................................................................................5-25
Figure 5-13 Vaporous Carryover for Various Chemicals .........................................................5-26
Figure 5-14 Water and Steam Density-Pressure Relationship at Saturation ...........................5-28
Figure 5-15 Basic Steps in Deposit Accumulation on a Surface from Bulk Fluid.....................5-29
Figure 5-16 The Logarithm of the Partitioning Constants for 1: 1 Solutes as a Function of
Reciprocal Temperature in Kelvin ....................................................................................5-35
Figure 5-17 The Logarithm of the Partitioning Constants for Neutral Solutes as a
Function of Reciprocal Absolute Temperature .................................................................5-37
Figure 5-18 Partitioning Constants-1:2 Compounds ................................................................5-38
Figure 5-19 Solubility of Cuprous (Cu2O) and Cupric (CuO) Oxides in Superheated
Steam as a Function of Pressure (from data by Pocock and Stewart).............................5-41
Figure 5-20 An Example of Vaporous Carryover vs Temperature for All-Volatile
Treatment (Trends) ..........................................................................................................5-42
Figure 5-21 Mollier Diagram for a Fossil Plant.........................................................................5-44
Figure 5-22 H, S Diagram with Turbine Expansion Curves and Solubility Curves for NaCl
(Sodium Chloride Concentration Expressed as Sodium) .................................................5-48
Figure 5-23 H, S Diagram with Turbine Expansion Curves and Solubility Curves NaOH
(Sodium Hydroxide Concentration Expressed as Sodium) ..............................................5-48
Figure 5-24 H, S Diagram with Turbine Expansion Curves and Solubility Curves for
Na2SO4 (Sodium Sulfate Concentration Expressed as Sodium) ......................................5-49

xvi
LIST OF TABLES

Table 1-1 Deposition Processes and Mechanisms (Listed Alphabetically)................................1-7


Table 1-2 Deposition Processes and Mechanisms (Listed by Order of Merit) ...........................1-8
Table 2-1 Sources of Contaminant Ingress to Fossil Plant Cycles ............................................2-2
Table 2-2 Chemistry-Related Damage Mechanisms in Condensate/Feedwater Systems ........2-5
Table 2-3 High Pressure Heater Pressure Drop Improvements, psi (MPa) Following
Cleaning to Remove Copper Deposits ...............................................................................2-7
Table 2-4 EPRI Core Monitoring Parameters and/or Minimum Level of Continuous
(22)
Instruments for All Units Operating on AVT .................................................................2-10
Table 3-1 Crystalline Compounds Found in Water-formed Deposits.........................................3-2
Table 3-2 Copper Alloys Most Commonly used in Feedwater Heaters .....................................3-4
Table 3-3 Factors given High Scores in Tables 2-1 to 2-3 of Part 1 of Reference 1 for
Deposition in Boilers ........................................................................................................3-21
Table 5-1 Overview of Key Historical Research into Steam Impurities......................................5-3
Table 5-2 Turbine Steam Path Mechanisms Influenced by Cycle Chemistry ............................5-6
Table 5-3 Components in Turbine Deposits from W. H. Sammis Plant ...................................5-10
Table 5-4 Concentrations of Chemical Species in Turbine Deposits .......................................5-11
Table 5-5 Concentration of Impurities in Exfoliated Magnetite Filtered from Steam ................5-18
Table 5-6 Comparison of Turbine Deposit Composition (wt %) ...............................................5-18
Table 5-7 Relative Amounts of Water-soluble Turbine Deposits .............................................5-19
Table 5-8 List of Utilities/Units Losing Capacity from Copper Deposition ................................5-20
Table 5-9 Possible Sources of Copper and its Oxides in Steam .............................................5-21
Table 5-10 Summary of Steam Impurity Generation and Transport ........................................5-21
o o
Table 5-11 Boiling at 347 C (657 F), Early Condensate Infinitesimal Liquid Droplet* at
o o
100 C (212 F) ...................................................................................................................5-34
o o
Table 5-12 Example Calculation of Phosphate Levels in Steam at 350 C (662 F) ..................5-38
Table 5-13 Impurity Concentration Mechanisms in Steam Turbines .......................................5-46
o o
Table 5-14 Maximum Copper Solubilization at 4500 psi, 1150 F (30.1 MPa, 621 C) ..............5-47
Table 5-15 Concentrations of Impurities in Steam under Steady Load Conditions .................5-51
Table 5-16 Problems, R&D, Root Causes, and Solutions .......................................................5-53
Table 5-17 Possible Solutions for Copper Transport Problems ...............................................5-55
Table 5-18 Factors Given High Scores in Tables 2-1 to 2-3 of Reference 52 for
Deposition in Turbines .....................................................................................................5-58

xvii
Table 5-19 Areas where there is Insufficient Information about the Factors Affecting
Turbine Deposition or their Relative Importance ..............................................................5-63

xviii
1
DEPOSITION IN FOSSIL PLANT CYCLES

1.1 Introduction

The presence of deposits, within and on the heat transfer surfaces of equipment comprising the
steam/water cycles of fossil plant units, has been noted since the very beginning of the
commercial steam and power generation industry.

Areas of fossil plant cycles subject to deposition activity are identified in Figure 1-1. (1) Also
identified are sources of impurity ingress and areas where corrosion is problematic. (Impurity
ingress, corrosion and deposition are closely interrelated phenomena and must be carefully
considered when selecting and optimizing the water chemistry program, as discussed in Section
1.2) From casual inspection of the figure, it becomes clear that deposition activity can develop at
nearly all locations within the steam-water cycle. However, operating experience makes it clear
that the operational and resultant economic impacts of deposition vary significantly by location.

Advancements in unit design, specifically increased operating pressures, higher heat fluxes in
boilers and cycles with higher thermal efficiencies, have resulted in increased emphasis on
equipment cleanliness, requiring careful attention to operation and maintenance practices. The
influence of the cycle chemistry on deposition activity has long been recognized, as has the role
of deposition as an integral part of mechanisms that cause damage and produce failures in
pressure parts in contact with water and steam. The area of greatest concern to fossil plant
personnel is in boiler tubes, where failures can have substantial negative impacts on unit
(2)
availability and reliability. However, deposition occurs at other locations in the steam-water
cycle as well, and these deposition events also affect unit performance.

Deposition consists of three general processes: (3)


Physical attachment of suspended and colloidal solids to surfaces.
Formation of solids from low solubility impurities present in the water or steam on
component surfaces including reversible deposition phenomena, collectively referred to as
hideout and hideout return activity.
Deposition processes that are related to surface phenomena including adsorption, absorption
and ion exchange on surfaces.

1-1
Deposition in Fossil Plant Cycles

Figure 1-1
Locations of Deposition, Corrosion and Impurity Ingress in Fossil Plant Cycles

NOTE: Although a drum boiler is shown in this figure, the general locations are similar for once-
through boilers. Deposit locations within the boilers differ with boiler type (drum or once-through),
design, operation and other factors.

The first deposition process group includes deposition of metal oxides formed during corrosion
of cycle materials as well as any colloidal solids that may be present as a result of cycle
contamination. For example, colloidal iron, aluminum and silica are sometimes present in the
circulating water and/or the supply to the makeup water treatment system. Events such as
condenser leaks and improper operation of makeup water treatment plant can result in
introduction of colloids to the cycle.

The second deposition process group includes formation of solids from species of limited
solubility. Included here are many of the salts normally present in plant water supplies and
cooling waters. In general, solubility of salts decreases as the temperature of water increases and
decreases further in steam.

The second deposition process group includes situations where constituent solubility varies with
normal changes in unit operation. Hideout and return of phosphate in the boiler water of drum
units is the most common example of this, however, based on available research findings, it is
clear that other species exhibit similar tendencies, both within the boiler and at other locations
within the cycle.

1-2
Deposition in Fossil Plant Cycles

The third deposition process group includes transport of water soluble material on insoluble
particles. Mechanisms include absorption, adsorption and ion exchanges.

These processes, while generally well understood in nature, are influenced by a variety of factors
specific to design conditions and specific operating environments unique to fossil plant cycles.
An assessment of the state-of-knowledge, significance, controllability and deficiencies for each
(3)
parameter is extensively presented in an earlier study.

1.2 Factors Influencing Deposition Activity

Factors or parameters known or believed to be related to the basic deposition process categories
and mechanisms were evaluated earlier.(3) The list included factors that contribute to or
otherwise influence each mechanism and factors that result from each mechanism.

Parameters considered relevant to each process group were ranked according to approximate
influence on generation and deposition within condensate systems, feedwater systems, boiler
waterwalls, superheaters and reheaters and steam turbines. An influence diagram (Figure 1-2)
was produced to visualize the relative impact of each category on the major fossil plant
components comprising the steam-water cycle. Although generation and deposition of corrosion
products and impurities can occur throughout the fossil plant cycle, the source of deposits was
generally considered to be the condensate and feedwater systems, as indicated in Figure 1-2.
Likewise, deposition was assumed to be predominant in the boiler and steam portions of the
cycle. Suspended corrosion products and other particles generally deposit in single phase water
and two-phase water/steam regions. Water soluble material generally deposits in two-phase
water/steam and single phase steam regions. Although there are notable exceptions to these
general concepts, the influence diagram produces some useful insights into the overall
importance of various contributing parameters, considers their potential controllability in fossil
plant units, and points to areas that may need further investigation. The discussion in the study
ultimately leads to a statement of uncertainties in the state-of-knowledge and potential
opportunities for improved control of deposition by an enhanced understanding of the
(3)
mechanisms and approaches for direct control of deposition.

Over 40 parameters were identified and ranked as being known to influence deposition activity
(3)
in fossil plant steam-water cycles (Table 1-1). The ranking consisted of allotting grades from 0
to 5 to each deposition parameter as it relates to the behavior of (a) suspended material, (b)
soluble material and (c) surface phenomena within the main systems of the power plant
(condensate system, feedwater system, boiler, superheater/reheater and turbine). Most
parameters received a ranking between 0 and 4. One half of the ranking (2 points) was based on
the significance of the parameter to the system. If a parameter was insignificant with respect to
the system, the parameter was assigned a preliminary rank of zero. Parameters with moderate
and high significance were given a preliminary rank of 1 and 2 respectively. The second half (2
points) of the ranking reflects the degree to which that parameter can be controlled in existing
fossil power plants. An additional 1 or 2 points were added to the preliminary ranking to
indicate the moderate or high level of controllability, respectively.

The final ranking consists of the sum of the two values (significance and controllability) with an
additional point added for parameters that could be significantly improved by a moderate

1-3
Deposition in Fossil Plant Cycles

hardware change. The ranking could be different for new construction or very significant plant
modifications.

The results were presented in the Part 1 report (3) and are summarized differently in this report. In
this report, the results were obtained by summing individual contributions for all of the power
plant systems relative to the three deposition modes. A higher value defines a larger influence
on the deposition process for any of the 40 parameters. Table 1-1 lists each of the parameters
alphabetically and Table 1-2 lists each of the parameters by rank. The results are depicted
graphically in Figure 1-1.

The highest ranked parameters were found to be transient conditions (startup, shutdown, load
changes, etc.); concentrations of impurities (contaminants), oxidizing agents (oxygen) and
particles; pH at operating temperatures; mass transport; condensate polishing and filtration;
chemical treatment; temperature; boiling, condensing and flow regimes; agglomeration of
particles; stream quality (moisture); pressure; tube/surface and oxide/hydroxide compositions
and interactions; time; heat flux; steam/water properties; size and shape of particles; corrosion
rate; configuration of hardware and zeta potential.

The ranking was sorted by type of deposition (insoluble material, soluble material and surface
phenomena). The overall high ranking (score of 3 or higher) parameters with respect to
suspended particles were (in order of highest to medium rank):
1. pH at operating temperature,
2. Transient conditions and outages,
3. Concentrations of particles,
4. Chemical treatment,
5. Flow characteristics,
6. Concentrations of metal ions and impurities,
7. Velocity and flowrate,
8. Configuration of hardware (steam/water flow orientation),
9. Corrosion rate,
10. Zeta potential,
11. Condensate polishing and filtration,
12. Concentrations of oxygen and oxidizing agents,
13. Mass transport,
14. Particle size and shape,
15. Crystalline structure and
16. Tube/surface composition.

1-4
Deposition in Fossil Plant Cycles

The overall ranking high ranking (score of 3 or higher) parameters with respect to deposition of
soluble impurities and corrosion products (in order of highest to medium rank):
1. Concentration of metal ions and impurities,
2. Condensate polishing and filtration,
3. Transient conditions and outages,
4. Steam quality,
5. pH at operating temperature and
6. Concentrations of oxygen and oxidizing agents.

The overall ranking high ranking (score of 3 or higher) parameters with respect to surface
phenomena (sorption, hideout, etc) in order of highest to medium rank were:
1. Concentrations of metal and non-metal ions, impurities and particles and
2. Ion exchange properties of metal oxides and transient conditions and outages.

The ranking was used to select key parameters because of their apparent, even if not fully
understood, influence on one or more of the three deposition process types:
Dynamics of Deposition and Release (as it effects both boilers and turbines)
Heat Flux (primarily in boiler waterwall tubes)
Mass Transport
Solubility/Steam
Solubility/Water
Surface Finish (primarily in turbines and steam generator water tubes)

Additional parameters of interest, but for which more information is needed to establish their
role in fossil plant units and/or to determine their impact on possible solutions to deposition
problems, include the following:
Agglomeration of Particles
Catalysis
Electrical Charging of Water
Electrochemical Potential and Oxidation Potential (ECP/ORP)
Ion Exchange Properties of Metal Oxides
Magnetic Properties of Solids, Surfaces and Deposit/Oxide and Electromagnetic Fields
Molecular Cluster Collisions
Particle Size Distribution
Porosity (and density) of Deposits and Oxides

1-5
Deposition in Fossil Plant Cycles

Steam/Water Boiling and Condensing Regimes


Steam/Water Flow Regimes
Steam/Water Properties
Transport of Soluble Material and Impurities on Particles by Ion Exchange, Adsorption,
Absorption and Magnetic or Electromagnetic Attraction
Zeta Potential

Some of these parameters may require experimental work while others will need to be
considered in modeling investigation and development efforts.

Many of the remaining parameters not cited in the previous listings relate to cycle chemistry and
are adequately addressed by available EPRI cycle chemistry guidelines for fossil plants and
would, therefore, be regarded as control (fixed) parameters. Other parameters not listed
generally do not appear to require direct examination but many of them should be regarded as
important parameters for further investigative efforts.

1-6
Deposition in Fossil Plant Cycles

Table 1-1
Deposition Processes and Mechanisms (Listed Alphabetically)

SOLIDS SOLUBLE SURFACE SOLIDS SOLUBLE SURFACE


Agglomeration of 9 0 10 Mass Transport 10 9 9
Particles
Catalysis 0 3 0 Molecular Cluster 5 0 0
Collisions
Chemical 16 8 0 Particle Shape 10 0 2
Treatment and Size
Coatings: 6 0 0 pH at Operating 19 10 0
Synthetic/Metallic Temperature
Concentration: 13 9 0 Porosity/Density 0 6 4
Metal Ion of Deposits/
Oxides
Concentration: 0 18 18 Pressure 7 5 5
Impurities and Ionic
Contaminants
Concentration: 11 10 0 Solubility and 0 8 0
Oxygen and Distribution in
Oxidizing Agents Water/Steam
Concentration 17 0 0 Steam Quality 6 11 0
Particles: (Moisture
Content)
Condensate 11 14 0 Steam/Water 14 5 0
Polishing and Boiling,
Filtration Condensing and
Flow Regimes
Configuration of 12 0 0 Steam/Water 6 6 0
Hardware Properties
(Steam/Water Flow
Orientation)
Corrosion Rate 12 0 0 Surface Finish 7 0 2
Crystalline 10 0 0 Temperature 7 6 6
Structure
Dynamics of 7 0 0 Time 9 6 0
Deposition/Release
Electrical Charging 5 0 0 Transient 18 11 10
of Water Conditions
(Including
Outages)
ECP/ORP 0 3 0 Transport of 0 0 10
Soluble Material
and Impurities on
Particles
Heat Flux 4 5 5 Tube/Surface 10 0 0
Composition
Interaction of Metal 9 7 0 Tube/Surface 9 7 0
Oxides and Oxide/Hydroxide
Hydroxides Composition
Ion Exchange 0 0 10 Velocity and 12 6 5
Properties of Metal Flowrate
Oxides
Magnetic and EMF 3 0 0 Zeta Potential 11 0 0
Properties of
Oxides, Surfaces
and Deposits

1-7
Deposition in Fossil Plant Cycles

Table 1-2
Deposition Processes and Mechanisms (Listed by Order of Merit)

SURFACE

SURFACE
SOLUBLE

SOLUBLE
EFFECT

EFFECT
SOLIDS

SOLIDS
TOTAL

TOTAL
Transient Conditions 18 11 10 39 Particle Shape and Size 10 0 2 12
(Including Outages)
Concentration: Impurities and 0 18 18 36 Corrosion Rate 12 0 0 12
Ionic Contaminants

pH at Operating Temperature 19 10 0 29 Configuration of Hardware 12 0 0 12


(Steam/Water Flow
Orientation)

Mass Transport 10 9 9 28 Steam/Water Properties 6 6 0 12


Condensate Polishing and 11 14 0 25 Zeta Potential 11 0 0 11
Filtration
Chemical Treatment 16 8 0 24 Tube/Surface Composition 10 0 0 10
Velocity and Flowrate 12 6 5 23 Transport of Soluble 0 0 10 10
Material and Impurities on
Particles
Concentration: Metal Ion 13 9 0 22 Porosity/Density of 0 6 4 10
Deposits/ Oxides
Concentration: Oxygen and 11 10 0 21 Crystalline Structure 10 0 0 10
Oxidizing Agents
Temperature 7 6 6 19 Ion Exchange Properties of 0 0 10 10
Metal Oxides
Steam/Water Boiling, 14 5 0 19 Surface Finish 7 0 2 9
Condensing and Flow Regimes
Agglomeration of Particles 9 0 10 19 Solubility and Distribution 0 8 0 8
in Water/Steam
Steam Quality (Moisture 6 11 0 17 Dynamics of 7 0 0 7
Content) Deposition/Release
Pressure 7 5 5 17 Coatings: 6 0 0 6
Synthetic/Metallic
Concentration: Particles 17 0 0 17 Molecular Cluster 5 0 0 5
Collisions
Tube/Surface Oxide/Hydroxide 9 7 0 16 Magnetic and EMF 3 0 0 3
Composition Properties of Oxides,
Surfaces and Deposits
Interaction of Metal Oxides and 9 7 0 16 Electrical Charging of 5 0 0 5
Hydroxides Water
Time 9 6 0 15 Catalysis 0 3 0 3

Heat Flux 4 5 5 14 ECP/ORP 0 3 0 3

1-8
Deposition in Fossil Plant Cycles

1.0

0.9

0.8

NORMALIZED RANK 0.7

SOURCES
0.6

SINKS
0.5

0.4

0.3

0.2

0.1

SUSPENDED SOLIDS
0.0
WATER SOLUBLE MATERIAL
TE SURFACE PHENOMENA ON SOLIDS
SA TE
R
EN R
ND WA I LE H
CO ED BO /R
FE SH INE
RB
TU
REVISED 07/06/2002

Figure 1-2
Influence Diagram for Deposition Parameters

1.3 EPRI Strategy to Address Deposition State-of-Knowledge Deficiencies

The relationships between impurity ingress, deposition and corrosion are well understood with
respect to the ways in which they can damage cycle components in contact with water and steam.
Deposits are integral to several damage mechanisms effecting the boiler and turbine. In addition,
deposits in these components and others can restrict flow and impede heat transfer, resulting in
reductions in capacity and efficiency and increases in heat rate.

While additional state-of-knowledge assessment efforts may be needed to determine long range
research plans and priorities, it would appear that the current assessment findings are sufficient
to define the short term needs.

Further, it is apparent that the short term strategy for deposition research is quite consistent with
the needs of the fossil industry. Key aspects of this include:
Experiments on reducing and oxidizing chemistry and on magnetite and copper chemistry).
Completion and reporting on results of boiler deposition studies at Moscow Electric Power
Institute in 2003; continuation of these studies will be considered upon completion of the
current work.
Initial feasibility assessments for development of a boiler water deposition model in 2004; if
the feasibility can be demonstrated, consideration will be given to its construction and testing
under future research initiatives.
Experiments to assess deposition in steam turbines, with the initial effort to be completed in
2005.

A more detailed accounting of the scientific aspects of the state-of-knowledge deficiencies will
be needed in order to proceed with long range research plans. This will be addressed by either

1-9
Deposition in Fossil Plant Cycles

issuing a follow-up report which compiles the existing sources of pertinent information and
overlays them on the main parts of the steam-water cycle or by addressing aspects of specific
interest within research report products that focus on specific deposition concerns. Long range
research needs will be dependent on results produced during these projects. However, several
aspects of deposition that appear likely candidates for inclusion in the long range plan are
already under consideration. These include:
Deposition model development, with initial interest in boiler deposits as needed to link to
work conducted under EPRI Program Boiler Corrosion.
Model development for deposition in turbines and possibly other fossil unit components.
Linkage of the models to an expert system, such as EPRI ChemExpert.
Assessment of the electrostatic and magnetic aspects of deposition including zeta potential,
point of zero charge, streaming potential, flow-induced charging and magnetizing of
droplets/particles, etc. with initial experimental efforts focusing on feedwater conditions and
subsequent efforts considering the boiler and turbine environments.
Evaluation of additives, surface conditioning or other techniques to inhibit deposition under
fossil plant conditions.
Assessment of the relationship between boiler deposition activity, initiation of under-deposit
corrosion activity, and the need for chemical cleaning.

The ultimate goal of the short and long range research activities is to identify new products and
develop tools that fossil plant personnel may apply to new and existing fossil units to better
manage any deposition activity that cannot be avoided and, in so doing, eliminate or minimize
the negative impacts of deposition on unit performance and profitability.

1.4 Report Overview

This report supplements work that was initiated in 2002 and document in the Part 1 report on
state-of-knowledge of deposition processes in the water/steam circuits at fossil-fuel fired power
(3)
plants. Section 1 of this report contains summary tables of deposition processes that extract
information from Section 2 of the Part 1 report. Sections 2, 3, 4 and 5 of this report focus on the
specifics of deposition in various parts of the fossil cycle, specifically:
Section 2: Condensate and feedwater systems,
Section 3: Boilers (once-through and drum type) including an introduction to the
superheaters and reheaters,
Section 4: Superheaters and reheaters and
Section 5: Turbines.

Finally, Section 6 provides conclusions and recommendations.

This report does not replace the Part 1 report since the Part 1 report contains information that is
not carried forward to this report. The Part 1 and Part 2 reports are complimentary.

1-10
Deposition in Fossil Plant Cycles

1.5 References

1. Cycle Chemistry Guidelines for Fossil Plants: All-Volatile Treatment: Revision 1, EPRI,
Palo Alto, CA: 2002. 1004187.

2. Guidelines for Chemical Cleaning of Conventional Fossil Plant Equipment, EPRI, Palo Alto,
CA: November 2001. 1003994.

3. State-of-Knowledge on Deposition, Part 1: Parameters Influencing Deposition in Fossil


Units, EPRI, Palo Alto, CA: December 2002. 1004194.

1-11
2
DEPOSITION IN CONDENSATE AND FEEDWATER
SYSTEMS

2.1 Introduction

This section of the report reviews corrosion and deposition in the condensate and feedwater
system. The condensate and feedwater systems have few areas of deposition. These systems
predominately function as generators of corrosion products which then migrate to other parts of
the heat cycle, namely the boiler, superheater, turbine and reheater. In these locations some
serious deposition problems can occur, if condensate and feedwater chemistry were not
maintained.

Areas in the condensate/feedwater systems where deposition has been reported are the high-
pressure feedwater heaters and, to a much lesser extent, on the steam side surfaces of the
condenser tubes. Feedwater deposits have been comprised of metallic copper and metallic
oxides of various types, depending on the materials of construction of the feedwater heaters.

Deposits on the steam side surfaces of the condenser tubes have been identified as iron or iron
oxides and, occasionally, oil.

The condensate system consists of the:


Condenser
Makeup water supply
Condensate discharge pumps
Condensate polisher and/or condensate filter (if used) and
Interconnecting piping.

The feedwater system consists of the:


Low pressure feedwater heaters and dearator (if present in the cycle)
Boiler feed pump
High pressure feedwater heaters and
Interconnecting piping.

The materials of construction of the condensate and feedwater systems have a direct bearing on
the corrosion and deposition in these systems as well as in the transfer of corrosion products to
other parts of the cycle, particularly the boiler, superheater, reheater and turbine.

2-1
Deposition in Condensate and Feedwater Systems

Figure 1-1 (presented in Section 1) depicts a typical heat cycle for a drum unit and indicates
locations where impurities can enter the cycle and where corrosion and deposition can occur.
While deposition can occur in the condensate and feedwater systems, as described later,
corrosion products formed in these systems are transferred to the boiler. There deposition and
additional corrosion can occur. From the boiler, impurities and corrosion products are
transferred to the superheater, turbine and reheater.
(1)
Table 2-1 lists potential sources of contaminant ingress. These contaminants can cause
corrosion, producing soluble and/or insoluble corrosion products that can deposit in downstream.
In this table, individual contaminant sources have been assigned an overall significance
designation (such as major or minor) to indicate the relative importance of each source in deposit
formation.

Table 2-1
Sources of Contaminant Ingress to Fossil Plant Cycles

Contaminant Source Overall Significance in Deposit References(1-9)


Formation*

Air Inleakage Often Major 4,5

Condenser Leaks Potentially Major 4-7

Control of Cycle Chemistry Normally Minimal To Minor 5


(Potentially Major)

Impurities in Treated Cycle Makeup Normally Minimal 5,8

Condensate Polishers Normally Minimal 5,8-9

Impurities in Water Treatment Chemicals Normally Minimal 5

Combustion Products Normally Minimal 5

Paints, Preservatives, Solvents, Etc. Normally Minimal 5

Ineffective Chemical Cleanings Minimal To Major 2-3


*
For individual plant units, all applicable contaminant sources should be considered.

2.2 Locations Subject to Corrosion and Deposition

The materials of construction in the condensate and feedwater systems vary considerably. In the
condenser, common tube materials include copper alloys, such as arsenical copper, aluminum
bronze, aluminum brass, Admiralty brass and copper-nickel, and stainless steel and titanium.
Low-pressure (LP) feedwater heater tubes may consist of copper alloys, such as Admiralty brass
or copper nickel, and stainless steels. High-pressure (HP) feedwater heater tubes can consist of
copper nickel, Monel, stainless steel and (more rarely today) carbon steel.

2-2
Deposition in Condensate and Feedwater Systems

All ferrous systems will have stainless steel or carbon steel feedwater heater tubes, but may have
copper alloys for condenser tubes. Mixed metallurgy systems will have copper alloys in the low
pressure and/or high-pressure heater tubes and perhaps in the condenser tubes.

Only minimal deposition occurs in the all-ferrous cycles, with the atypical exception of minor
iron deposits in the feedwater heaters (steam and water surfaces) and the steam side of the
condenser tubes. Occasionally, oil can be deposited on the steam side of condenser tubes as a
result of turbine bearing lube oil contaminating steam seals or other sources of oil leaking into
the steam/water cycle.

The situation is more complex in mixed metallurgy systems, where copper and metal oxide
(10-12)
deposits can form in the high-pressure heaters. Figure 2-1 shows copper that had deposited
on the waterside surfaces of a 70-30 copper- nickel high-pressure heater tube.(10-11) The deposits
increase heater pressure drop requiring additional power for the boiler feed pump to maintain the
required flowrates. In addition, deposit buildup impedes heat transfer, which may reduce final
feedwater temperature.

As copper deposits become thicker, they eventually come off the surface and are transported to
the boiler as copper foil or snake skins. Release of copper via exfoliation is also significant in
those units containing susceptible copper-nickel alloys, especially in peaking or cycling units.

Copper, if present in extraction steam from the turbine, can also deposit on high-pressure heater
(10-12)
steam-side surfaces, as shown in Figure 2-2. This copper may eventually re-enter the
feedwater and transport to the boiler and turbine.

De-alloying, which preferentially dissolves zinc from brass or nickel from copper-nickel alloys,
has also been cited as a means by which copper and other metals may be released from feedwater
heaters.

Certain areas of the condensate and feedwater parts of the cycle serve as reservoirs for loose
corrosion products and other particulate matter such as dust, dirt, and residues of abrasives and
other materials used during equipment maintenance. These include the condenser hotwell, the
deaerator heater and storage tank (when included in the cycle design) and the shell sides of
feedwater heaters. It is generally regarded as preferred practice to inspect these areas at each
suitable opportunity and arrange for removal of loose solids accumulations, this to avoid
transport of theses solids to locations where they could form adherent deposits during future unit
operations.

2-3
Deposition in Condensate and Feedwater Systems

Figure 2-1
High Pressure Heater Tube at Magnification of 200x (70-30 Copper-Nickel after 29 Years of
Service)

Source: References 10-12

Figure 2-2
High Pressure Heater Tube at Magnification of 500x (Alloy 400, Monel, after 26 Years of
Service)

Source: References 10-12

Deposits may form on either the waterside or steamside surfaces of condenser tubes. Deposition
on the waterside is the more serious problem. Deposition of iron oxides occurs on the steamside
surfaces of condenser tubes (and other surfaces within the condenser), producing generally
minimal impacts on efficiency.

2-4
Deposition in Condensate and Feedwater Systems

Damage mechanisms that occur in the condensate and feedwater portions of fossil plant cycles
are tabulated in Table 2-2. (13-15) None of the cited mechanisms requires deposition to be present to
initiate or perpetuate the damage. However, each of these mechanisms produces corrosion
products that can be transported to other locations in the cycle.
Table 2-2
Chemistry-Related Damage Mechanisms in Condensate/Feedwater Systems

Location Mechanism Chemistry Influence

Condensate and feedwater Carbon steel and copper alloy Low pH, acid constituents,
systems corrosion and corrosion product excess carbon dioxide and
transport oxygen present in condensate;
alternating oxidizing and
reducing conditions, excess
hydrazine dissolving magnetite

Heater drain and feedwater Flow accelerated corrosion Attack by reducing feedwater
piping (FAC) of carbon steel conditions and high velocities,
excessive hydrazine with zero
oxygen, low pH

Copper alloy condenser and Ammonia attack/condensate Simultaneous excess ammonia,


heater tubes corrosion oxygen and carbon dioxide in
steam synergistically oxidizing
and dissolving copper

Feedwater heater tubes Stress corrosion cracking Cu/Ni, Excessive corrodents in steam
Monel, stainless steel synergistic with tensile stress,
corrodent concentration in
crevices, dry-wet transition

Condenser and heater tubes Admiralty brass stress corrosion Excessive ammonia/chloride
cracking present in steam synergistic with
residual stress at tubesheets and
in u-bends

Feedwater heater tubes Copper/nickel exfoliation Excessive oxygen on shutdown


combined with thermal cycling
and thermal stresses

Source: References 13 (Adapted from References 14 and 15)

2.3 Mechanisms

The ingress of contaminants into the cycle is responsible for increased corrosion of the
condensate and feedwater circuits. This increased corrosion results in the generation of soluble
and insoluble metals and metal oxides from the alloys containing iron, copper, nickel, chromium,
zinc, titanium and other metals. Subsequently, these metals and metal oxides can be transported
to the boiler, superheater, turbine and reheater. Deposits are likely to accumulate in the high-
pressure heaters and in other parts of the heat cycle downstream.

2-5
Deposition in Condensate and Feedwater Systems

Obvious sources of severe contamination include condenser leaks, acid from chemical cleaning
operations, and acid and caustic from condensate polisher and makeup demineralizer
regenerations.

Air inleakage (typically into the condenser) can contribute to corrosion since air is a source of
both oxygen and carbon dioxide. The carbon dioxide suppresses the pH of the water toward the
corrosive, acidic region. Oxygen from air inleakage is especially important in mixed metallurgy
systems since these systems must be operated under reducing conditions (e.g. ORP of -300 to -
350 mV) to minimize copper transport. This is accomplished by controlling air inleakage and by
feeding a reducing agent. The effectiveness of this approach is evaluated by monitoring the ORP
at the deaerator inlet and copper in the feedwater.

The pH of mixed metallurgy systems is critical to maintaining minimal corrosion product


transport. A pH of 9.0 to 9.3 for mixed metallurgy systems contrasts with a pH of 9.2 to 9.6 in
all-ferrous systems is a necessity for minimizing corrosion in each of those systems.

The mode of power plant operations can contribute significantly to increased corrosion and
deposition. Units operated in cycling or peaking mode, with frequent startups and shutdowns
exhibit high corrosion product transport, especially during startups. This situation is caused by
inadequate layup procedures and chemical feed and oxygen (air inleakage) control prior to unit
restart and during restarting process. The layup procedure has to be tailored to the layup period.
For example, a layup procedure for a base loaded unit is different than one for a unit that
operates less frequently. An example of the increased iron and copper in boiler feedwater during
(16)
startup is shown in Figure 2-3.

In some instances, these deposits are sufficient to require temporary derating of the unit until
action can be taken to correct the problem. The literature cites one account where chemical
cleaning of high pressure heaters to remove copper based deposits resulted in significant
reductions in pressure differential following return of the unit to service, as summarized in Table
(17-18)
2-3.

2.4 Control

Proper condensate and feedwater chemistry limits, if conscientiously maintained, will minimize
corrosion of system materials, corrosion product transfer and deposition. EPRI recommended
chemistry limits for feedwater vary somewhat by the type of boiler water treatment used. The
feedwater (and boiler water) limits have been established in several operating guidelines for
(19-22)
phosphate treatment, OT, caustic treatment and AVT. There is also a guideline controlling
(10)
copper in fossil units. The use of condensate polishing and/or condensate filters have reduced
the amount of corrosion product transport and deposition in power plant cycles.(23)

2-6
Deposition in Condensate and Feedwater Systems

800 80

700 70

Copper Concentration (ppb)


Iron Concentration (ppb)
600 60

500 50

400 40

300 30

200 20

100 10

0 0
0 1 2 3 4 5
Time (Hours)
Figure 2-3
Example of Increased Iron and Copper in Feedwater During Plant Startup

Source: Adapted from Reference 16

Table 2-3
High Pressure Heater Pressure Drop Improvements, psi (MPa) Following Cleaning to
Remove Copper Deposits

No. 6 Heater No. 7 Heater

Before Cleaning 184 (1.29) 175 (1.23)

After Cleaning 120 (0.84) 84 (0.59)

Pressure Drop Reduction 64 (0.45) 91 (0.64)

Source: Reference 17 (Adapted from Reference 18)

As previously discussed, cycling and peaking units have higher corrosion product transport than
base loaded units, especially during frequent restarts. To minimize corrosion product transport
and deposition, proper lay-up provisions must be provided.(14) This includes the maintenance of
recommended layup chemicals and blanketing with nitrogen or dehumidified air.

Chemical feed is required to maintain pH and to control oxygen in condensate and feedwater
systems. The chemicals typically used are (a) aqueous ammonia for pH control and (b)
hydrazine for oxygen control and to maintain reducing conditions in mixed metallurgy systems.
Oxygen feed is used in place of hydrazine for units operating on oxidizing chemistry.

2-7
Deposition in Condensate and Feedwater Systems

Reducing chemical treatment applies to both all-ferrous and mixed metallurgy systems. For all-
ferrous systems, the pH is maintained in the range of 9.2 to 9.6. Hydrazine may be used in all-
ferrous systems, but with proper air inleakage control, many operators have found hydrazine to
be unnecessary. In fact, corrosion product transfer was found to be reduced by the elimination of
hydrazine when good control of air inleakage and water chemistry was practiced (ORP of 0 mV
(22,25)
or positive).

For mixed metallurgy systems, pH is controlled in the range of 9.0 to 9.3 and hydrazine must be
fed to provide a reducing atmosphere at the deaerator inlet (ORP of -300 to -350 mV) to
minimize copper transport and deposit formation in the high-pressure heaters.

Oxygenated treatment (OT) can only be used in all-ferrous systems which have rigid chemistry
control, usually accomplished through the use of condensate polishing. A controlled feed of
oxygen is maintained and hydrazine is not added (ORP of +100 to +150 mV). OT is applicable
to both once through and drum units, assuming the above requirements are satisfied.

OT, when properly applied and maintained, provides superior results relative to corrosion
product transfer compared to the traditional AVT approach. This can be explained as follows:
With AVT, the steel surface in contact with water is covered by a dense magnetite layer.
This dense magnetite layer is, in turn, covered by a thinner, more porous magnetite layer
which permits the underlying ferrous ions to be transported from the underlying steel to the
condensate and feedwater.
With OT, the pores in the upper, thinner layer are plugged with ferric oxide hydrate or ferric
oxide which mitigates the flow of ferrous ions, thus resulting in less iron transport.

Provided that the feedwater treatment is applied correctly, the following corrosion product levels
are achievable:
AVT (Reducing): <2 g kg1 (ppb) iron and/or copper
AVT (Oxidizing): Approximately 1 g kg1 (ppb) iron
OT: Approximately 0.5 g kg1 (ppb) iron

During the past 15 years, many new alternate amines and oxygen scavengers have been
introduced. Before applying these chemicals, their properties should be carefully determined,
and the experience with each should be verified. Then, within a few weeks of the first
application of the new chemical, the cycle chemistry should be analyzed in much more detail
(16)
than during the normal operation.

The selected treatment should protect all cycle components, and prevent generation of corrosion
products, and general and localized corrosion. These chemicals should be compatible with all
cycle component materials and lay-up practices, and should be compliant with all environmental
and health regulations. Most of the applications are for low- and medium-pressure industrial
units but some of the chemicals are also being used in utility cycles. The use of some of these
products can lead to corrosion, buildup of deposits, and other problems. It should be kept in
mind that the overall philosophy of the EPRI fossil plant cycle chemistry program is to keep the
cycle as pure as possible with as few chemical additions as possible. (16)

2-8
Deposition in Condensate and Feedwater Systems

These water treatment chemicals fall into the following two categories:
Neutralizing and filming amines for feedwater, steam and condensate pH control and
Reducing agents for oxygen removal.

However, to evaluate the effects of any water treatment chemical, data pertinent to its chemical
transport, decomposition products, cycle material corrosion, deposit and scale buildup and
(16)
analytical interferences must be known. These needed data include:
Hydrothermal stability in the cycle;
Kinetics of reactions;
Decomposition products and their effects;
Analytical interferences;
How to monitor/analyze;
Toxicity of the product, decomposition products, deposits, etc.;
Measured effects of pH, conductivity, cation conductivity and iron and copper
concentrations;
Stability in chemical addition tanks and storage containers;
Solubility and volatility of the chemical and its decomposition products;
Behavior of dried-out solutions (deposits in reheaters, superheaters, turbines and valves) and
Behavior under short-term and long-term lay-up conditions and during startup
(decomposition-acid formation, scale formation, disposal, etc.).

Very little is known about the impact of these alternate treatment chemicals on deposition. The
utility users of water treatment chemicals need to know the pressure and temperature range of
their application and the nature and behavior of the decomposition products.

2.5 Diagnostics

Control of condensate and feedwater chemistry is accomplished with the use of chemistry
monitoring of key parameters using both continuous instrumentation and grab samples.(19-22) Core
parameters for control of AVT and OT have been established by EPRI and are described in
(19-22)
guidelines documents. These core parameters are supplemented by additional monitoring
points and parameters to be used for commissioning and/or troubleshooting.(19-22) As an example,
(22)
the minimum guidelines for control of AVT (reducing and oxidizing) are given in Table 2-4.

The EPRI ChemExpert system processes readings from continuous instrumentation readings as
well as grab sample analysis data. (25) ChemExpert provides diagnoses of operating malfunctions
affecting water chemistry, such as condenser leaks, makeup demineralizer upsets, condensate
polisher breakthroughs, etc.

2-9
Deposition in Condensate and Feedwater Systems

Table 2-4
EPRI Core Monitoring Parameters and/or Minimum Level of Continuous Instruments for All
Units Operating on AVT (22)

Oxidizing AVT (AVT(O))

Drum or oncethrough units without a reducing agent and with an all-ferrous feedwater system.

Cation conductivity CPD, CPO or EI, RH (or MS), Downcomer (or blowdown)
Specific conductivity Makeup
pH (drum units) Downcomer (or blowdown)
Dissolved oxygen CPD, EI
Sodium CPD, CPO or EI, RH (or MS)
Air in-leakage
Carryover (drum units)

Reducing AVT (AVT(R))

Drum or once through units with a reducing agent and with a mixed-metallurgy or all-ferrous
feedwater system.

Cation conductivity CPD, CPO or EI, RH (or MS), Blowdown (or downcomer)
Specific conductivity Makeup
pH (drum units) Blowdown (or downcomer)
Dissolved oxygen CPD, EI
Sodium CPD, CPO or EI, RH (or MS)
ORP DAI
Air in-leakage
Carryover (drum units)

2.6 Factors Linked to the Matrices

The factors given higher scores (4 to 2) for deposition in condensate/feedwater systems


(originally presented in Tables 2-1 to 2-3 of Part 1 of this study (13)) have been abstracted from
these tables and are discussed as follows.

Condensate system high and medium ranked parameters with respect to suspended particles
were, in alphabetical order:
High (score of 4 or higher): Corrosion rate, chemical treatment, concentrations of oxygen and
oxidizing agents, concentrations of particles, pH at operating temperature, and transient
conditions and outages.
Medium (score of 3): Coatings, concentrations of metal ions and impurities, configuration of
hardware (steam/water flow orientation), flow characteristics, tube/surface composition,
velocity and flowrate and zeta potential.

2-10
Deposition in Condensate and Feedwater Systems

Feedwater system high and medium ranked parameters with respect to suspended particles were,
in alphabetical order:
High (score of 4 or higher): Corrosion rate, chemical treatment, concentrations of oxygen,
oxidizing agents and particles, condensate polishing and filtration and pH at operating
temperature, transient conditions and outages.
Medium (score of 3): Conditions and outages, concentrations of metal ions and impurities,
configuration of hardware (steam/water flow orientation), flow characteristics, tube/surface
composition, velocity and flowrate and zeta potential.

Condensate system high and medium ranked parameters with respect to deposition of soluble
impurities and corrosion products in alphabetical order were:
High (score of 4 or higher): Concentration of metal ions and impurities and concentrations of
oxygen and oxidizing agents.
Medium (score of 3): Electrochemical potential and oxidation potential (ECP/ORP).

Feedwater system high and medium ranked parameters with respect to deposition of soluble
impurities and corrosion products in alphabetical order were:
High (score of 4 or higher): Concentration of impurities, condensate polishing and filtration.
Medium (score of 3): Concentrations of metal ions, oxygen and oxidizing agents and pH at
operating temperature.

Condensate system high and medium ranked parameters with respect to surface phenomena
(sorption, hideout, etc) in alphabetical order were:
High (score of 4 or higher): Concentrations of metal and non-metal ions, impurities and
particles.
Medium (score of 2-3): Ion exchange properties of metal oxides, mass transport, and
transient conditions and outages.

Feedwater system high and medium ranked parameters with respect to surface phenomena
(sorption, hideout, etc) in alphabetical order were:
High (score of 4 or higher): Concentrations of metal and non-metal ions, impurities and
particles.
Medium (score of 2-3): Ion exchange properties of metal oxides, mass transport and transient
conditions and outages.

Damage mechanisms within condensate and feedwater systems are mainly carbon steel and
copper alloy corrosion, followed by corrosion product transport, flow accelerated corrosion
(FAC) of carbon steel, ammonia attack, stress corrosion cracking of copper/nickels, Monel,
Admiralty brass and stainless steel, and copper/nickel exfoliation. Possible root causes for
corrosion and deposition in the condensate and feedwater system are inadequate or inappropriate
chemistry control (for example, low pH, reducing feedwater conditions, excess ammonia,
oxygen and carbon dioxide in condensate, excess ammonia and chloride in steam ), inadequate

2-11
Deposition in Condensate and Feedwater Systems

purification of makeup water and condensate polishing, load cycling with high thermal stresses
and outages (including layup practices), contaminant ingress and air ingress or any combination
of the above. The corrosion mechanisms damage components in the condensate and feedwater
systems, sometimes resulting in failure and/or the need for major repairs and extended outages.
The corrosion generates soluble and insoluble metals and metal oxides from the alloys
containing iron, copper, nickel, chromium, zinc, titanium and other metals. Also, these corrosion
products may act as transportation and deposition sites for other impurities to the boiler,
superheater, reheater and turbines. Other related contributors are hardware configuration, mass
transport, flow characteristics, flow rate and velocity, electrochemical, oxidation and zeta
potential.

2.7 State-of-Knowledge Deficiencies

A large number of EPRI research projects are underway and planned for the future to investigate
knowledge deficiencies in the area of fossil plant deposition that may presently exist or that
might be uncovered as a result of these research projects. Regarding the condensate and
feedwater systems, no significant deficiencies are anticipated that have not been previously
explored or are planned to be explored.

2.8 References

1. Guidelines for Chemical Cleaning of Conventional Fossil Plant Equipment, EPRI, Palo Alto,
CA: 2001. 1003994.

2. Manual on Chemical Cleaning of Fossil-Fueled Steam Generation Equipment, EPRI, Palo


Alto, CA: 1984. CS-3289.

3. Guidelines for Chemical Cleaning of Fossil-Fueled Steam-Generating Equipment. Research,


EPRI, Palo Alto, CA: 1993. TR-102401.

4. Recommended Practices for Operating and Maintaining Steam Surface Condensers, EPRI,
Palo Alto, CA: 1987. CS-5235.

5. Interim Consensus Guidelines on Fossil Plant Water Chemistry, EPRI, Palo Alto, CA: 1986.
CS-4629.

6. Proceedings: Condenser Technology Conference, EPRI, Palo Alto, CA: 1996. TR-106781.

7. Condenser Technology Conference, EPRI, Palo Alto, CA: 1993.

8. Revised Guidelines for Makeup Water Treatment, EPRI, Palo Alto, CA: 1999. TR-113692.

9. Condensate Polishing Guidelines, EPRI, Palo Alto, CA: 1996. TR-104422.

10. Guidelines for Copper in Fossil Plants, EPRI, Palo Alto, CA: 2000. 1000457.

11. Courtesy A. G. Howell, New Century Energies, 2000.

2-12
Deposition in Condensate and Feedwater Systems

12. Mitigation of Copper Deposition in High Pressure Turbines of Utility Drum Boilers,
Power Plant Chemistry, 1(4), October 1999.

13. State-of-Knowledge on Deposition, Part I: Parameters Influencing Deposition in Fossil


Boilers, EPRI, Palo Alto, CA: 2002, 1004194.

14. Cycling, Startup, Shutdown and Lay-up Fossil Plant Cycle Chemistry Guidelines for
Operators and Chemists, EPRI, Palo Alto, CA: 1998. TR-107754.

15. J. Mathews, The Importance of Startup Chemistry to the Long-term Reliability of Power
Generating Equipment, Improvement of Chemistry Control During Startup of Fossil Units,
ASME Workshop, St. Louis, MO, April 9-10 1997.

16. State-of-Knowledge of Copper in Fossil Plant Cycles, EPRI, Palo Alto, CA: 1997, TR-
108460.

17. Cycle Chemistry Corrosion and Deposition: Correction, Prevention, and Control, EPRI,
Palo Alto, CA: 1993. TR-103038.

18. J. P. Engle, Chemical Cleaning of Feedwater Heaters, Corrosion/74, National Association


of Corrosion Engineers (NACE), March 4-8, 1974, Paper No. 11.

19. Cycle Chemistry Guidelines for Fossil Plants: Phosphate Treatment for Drum Units, EPRI,
Palo Alto, CA: 1994. TR-103665.

20. Cycle Chemistry for Fossil Plants: Oxygenated Treatment, EPRI, Palo Alto, CA: 1994, TR-
102285.

21. Sodium Hydroxide for Conditioning the Boiler Water of Drum-Type Boilers, EPRI, Palo
Alto, CA: 1995.

22. Cycle Chemistry Guidelines for Fossil Plants: All-Volatile Treatment: Revision 1, EPRI,
Palo Alto, CA: 2002. 1004187.

23. Condensate Polishing Guidelines, EPRI, Palo Alto, CA: 1996. TR-104422.

24. R. B. Dooley, J. Mathews, R. Pate and J. Taylor, Optimum Chemistry for All-Ferrous
Feedwater Systems: Why Use an Oxygen Scavenger, Proceedings of the55th International
Water Conference, Engineers Society of Western Pennsylvania, 1994.

25. Cycle Chemistry Advisor for Fossil Power Plants: ChemExpert Version 3.0, EPRI, Palo
Alto, CA: 2001. 1006402 (CD).

2-13
3
DEPOSITION IN BOILERS

This section of the report reviews deposition in boilers including the economizers, waterwalls
and drums. Deposition in superheaters and reheaters is covered in Sections 3.4 and 4. Two
additional EPRI reports supplement the information contained herein. The 2002 EPRI Part 1
report on deposition state-of-knowledge (1) and the 2003 EPRI report on Soviet and Russian
(2)
literature contain additional details and mathematical descriptions of boiler deposition
processes.

3.1 General Aspects of Waterside and Steamside Deposition in Boilers

Deposition of low-solubility solids and corrosion products in steam generators is as old as the
steam generating process itself. The low-solubility solids are formed either from impurities
introduced into the plant cycle via cooling water or makeup water or because of reactions
between the impurities and the boiler water treatment chemicals. Corrosion products
participating in the deposition process within the boiler originate in the low-temperature part of
the cycle (condenser, LP and HP feedwater heaters, deaerator and feedwater storage tank).

Contamination Originating in Makeup Water and Cooling Water

A lists of substances identified in deposits within boiler tubes and boiler drums of utility and
industrial boilers is given in Table 3-1.(3-4) Many of these contaminants enter the cycle with the
makeup water or as a result of ingress of cooling water.

Even very small in-leakage of cooling water causes cycle contamination and may result in
deposition in some cycle components with serious consequences. Cycle contamination by
cooling water may be caused by both shell-side and tube-side deficiencies in the condenser.
Shell-side deficiencies may result from mechanical damage, tube vibration, steam corrosion,
condensate corrosion and stress corrosion cracking; tube-side deficiencies may result from
biofouling, microbiologically-induced corrosion, erosion, flow-accelerated corrosion (FAC),
stress corrosion cracking, intergranular corrosion, crevice corrosion, pitting, de-alloying,
galvanic corrosion, uniform corrosion, tube joint leakage and leakage around the waterbox
(5)
flanges.

3-1
Deposition in Boilers

Table 3-1
Crystalline Compounds Found in Water-formed Deposits

Mineral Name Formula Mineral Name Formula


Fluorite CaF2 Pectolite NaCa2Si3O8OH
Sepiolite 2MgO3SiO22H2O Foshagite Ca4Si3O9(OH)2
Hopeite Zn3(PO4)24H2O Acmite NaFeSi2O6
Gypsum CaSO42H2O Dolomite CaMg(CO3)2
Kaolinite Al2Si2O5(OH)4 Whitlockite Ca3(PO4)2 (beta)
Viviante Fe3(PO4)28H2O Natrolite Na2Al2Si3O102H2O
Ferrous Hydrogen Carbonate Fe(HCO3)2 Malachite CuCO3Cu(OH)2
Iron Oxide Hydrate FeO(OH) Gehlenite Ca2Al2SiO7
Boehmite Al2(OOH)2 Olivine 2(Fe0.94Mg0.06)0SiO2
Mackinawite FeS Meta Thenardite Na2SO4
Gibbsite Al2(OH)3 Halite NaCl
Irons (II) Sulfate Hydrate FeSO44H2O Hydroxyapatite Ca5(PO4)3(OH)
Monohydrocalcite CaCO3H2O Covellite CuS
Goethite FeO(OH) Siderite FeCO3
Cristobalite (Low) SiO2 Thenardite Na2SO4
Witherite BaCO3 Burkeite Na6CO3(SO4)2
Nosean Na8Al6SO4(SiO4)6 Sodium Calcium CaNaPO4
Orthophosphate
Noselite Na8Al6(SiO4)6SO4 Thermonatrite Na2CO3H2O
Vaterite CaCO3 Hematite Fe2O3
Anhydrite CaSO4 Portlandite Ca(H)2
Irons (II) Sulfate Hydrate FeSO4H2O Magnetite Fe3O4
Barite BaSO4 Maghemite Fe2O3
Analcime (Analcite) NaAl(SiO3)2H2O Tenorite CuO
Aragonite CaCO3 Serpentine Mg3Si2O5(OH)4
Quartz (Alpha) SiO2 Zincite ZnO
Cancrinite 3NaAlSiO4CaCO3 Cuprite Cu2O
Xonotlite Ca6Si6O17(OH)2 Forsterite Mg2SiO4
Gyrolite Ca4(Si6O15)(OH)23H2O Willemite Zn2SiO4
Sphalerite ZnS Chalcocite Cu2S
Hemimorphite Zn4(OH)2Si2O7H2O Brucite Mg(OH)2
Sodium Silicate Na2SiO3 Wustite FeO
Calcite CaCO3 Periclase MgO
Calcium Sulfate Hydrate 2CaSO4 H2O Bunsenite NiO
Calcium Pyrophosphate Ca2P2O7 Troilite FeS
Wollastonite CaSiO3 Corundum Al2O3
Celestite SrSO4 Copper Cu
Calcium Hydrogen CaHPO4
Orthophosphate
Source: References 3 and 4

3-2
Deposition in Boilers

Inadequate design, poor condition or incorrect operation of makeup water treatment installations
and condensate polishers can contribute to cycle contamination and deposition. Many important
changes and developments in makeup water treatment and condensate polishing techniques have
been developed since the initial list of water-formed deposits was assembled, particularly with
(6-8)
respect to the application of new ion exchange technologies. Demineralized makeup has
become the industry standard even in industrial power generation. Numerous new technologies
are available for application in the treatment of makeup water for fossil plant cycles (e.g., reverse
osmosis, electromagnetic filtration, and continuous electrodeionization).

Nearly all once-through units and many drum units, in particular those with seawater or high-
TDS cooling water (cooling cycles with cooling towers), utilize condensate polishers. However,
many drum boilers are not equipped with condensate polishers. Condensate polishing offers the
(7)
following benefits:
Improvement in unit reliability and availability;
Reduction in boiler chemical cleaning frequency;
Reduction in turbine maintenance related to steam purity;
Faster startup (full load achieved more quickly);
Improved protection against condenser tube leaks;
Improvement in turbine efficiency (fewer turbine deposits);
Possibility of continued operation through peak demand periods during a minor condenser
leak;
Possibility of a controlled shutdown during a major condenser leak;
Reduction in replacement power costs caused by water chemistry-related forced outages;
Reduction in blowdown on drum units with a resultant decrease in makeup requirements and
energy losses; and
Reduction in fuel costs related to higher heat transfer efficiency associated with cleaner
boiler tubes.

Corrosion Products

Corrosion products from the pre-boiler areas are the source of most suspended solids that are
transported into the boiler. For all-ferrous feedwater systems (no copper alloys in the feedwater
and possibly copper-based condenser tubing), the generation of corrosion products (magnetite,
hematite and ferric oxide hydrate) occurs mainly due to corrosion and flow-accelerated corrosion
(FAC) of low-pressure and high-pressure feedwater heaters, deaerators, economizer inlet tubing
and piping, feedwater piping and drain lines.

For mixed metallurgy systems, the generation and transport of corrosion products (cupric and
cuprous oxide) is primarily a result of corrosion of copper-bearing tubes in the low-pressure
(8)
and/or high-pressure feedwater heaters. Table 3-2 lists copper alloys that are used for low-
pressure and high-pressure feedwater heater tubing.

3-3
Deposition in Boilers

Optimization of feedwater chemistry is important for the reduction of corrosion product transport
into the boiler. Target concentrations of <2 g kg1 (ppb) and approximately 0.5 g kg1 (ppb)
are achievable with AVT and OT, respectively. Guidance for optimizing the feedwater
treatments may be found in the respective Cycle Chemistry Guidelines and in the Guidelines for
(8-10)
Copper in Fossil Plants.
Table 3-2
Copper Alloys Most Commonly used in Feedwater Heaters

Low Pressure Feedwater Heaters High Pressure Feedwater Heaters

Admiralty Brass 80-20 Copper-Nickel

90-10 Copper-Nickel 70-30 Copper-Nickel

80-20 Copper-Nickel Monel (Alloy 400)

70-30 Copper-Nickel

Effects of Deposition

Deposition of chemicals, contaminants and corrosion products in the boiler may result in a
reduction in boiler efficiency and may reduce the service of boiler parts, and may finally lead to
boiler tube failures. A few typical examples:
Rippled deposits formed in economizers or waterwalls of once-through steam generators
operating on reducing all-volatile treatment cause a pressure drop increase. This increase
deteriorates cycle efficiency, requiring additional feed pump work, and full-load operation
may no longer be possible.
Buildup of excessive deposits in waterwalls of supercritical once-through steam generators
may cause or be one of the major root causes of waterwall cracking.
Thick and porous deposits in boiler tubes of drum boilers are the primary root cause of
under-deposit corrosion and may lead to boiler tube failures due to hydrogen damage, caustic
gouging and acid phosphate corrosion.
Mechanical carryover of sodium sulfate in steam during operation leading to the buildup of
sodium sulfate deposits in superheaters and reheaters may lead to or intensify pitting
corrosion during shutdown if inadequate or no layup methods are applied.

3.2 Locations Subject to Deposition

Water-Cooled Tubing

Water-side deposits in boilers typically consist of two layers a thin inner magnetite layer
adjacent to the metal (topotactic layer) and a more porous outer iron oxide deposit layer
(epitactic layer), Figure 3-1. The thin, hard, tenacious magnetite layer represents the normal
protective magnetite film formed from the reaction of water with the steel boiler tubing. Its

3-4
Deposition in Boilers

thickness depends on metal temperature and time of operation (parabolic law of formation). The
softer, more porous outer layer consists predominantly of corrosion products transported from
the pre-boiler system into the boiler by feedwater. Copper and suspended or other low-solubility
impurities are also found in the porous deposit layer, if present in the feedwater. This layer may
be essentially non-adherent.

Soft Transported
Iron Oxide Deposit
Mounting Material

Hard In-place Iron


Oxide Deposit

Base Material

Figure 3-1
Typical Boiler Waterwall Tube Deposit

Source: Reference 11 (Adapted from Reference 32)

Boiler tube deposits cause an increase in the boiler tube temperature. Heat flux is established by
the furnace flame temperature spectrum and the bulk fluid temperature, Tf, in the tube, which is
fixed by boiler pressure (Figure 3-2). The temperature profile is established by the total
temperature difference (T = T1 + T2 + T3) required to drive radiant heat from the furnace
through each segment of thermal resistance into the bulk fluid in the tube. With nucleate boiling
in a clean tube, T3 across the fluid film is usually small and the overall temperature differential
T primarily results from the thermal conductivity (t) of the tube metal, i.e., from the
temperature difference T1. With internal deposit on the tube, an additional temperature
difference T2 is required to drive the heat flux through the deposit (thermal conductivity, s) so
(11, 12, 32)
that inside (T2) and average tube metal temperatures raise accordingly.

3-5
Deposition in Boilers

Internal
Deposit Heat Flux, q
2 2
ls lt W/m (Btu/hr/ft )

T1

DT1
T2

DT2
T3

Tf DT3

ds
Figure 3-2
Waterwall Tube Heat Transfer Profile

Source: References 11 and 12 (Adapted from Reference 32)

Increases in the boiler tube temperature can play a significant role in tube failures caused by
under-deposit corrosion, as well as in boiler tube failures involving locally accelerated creeping
phenomena due to overheating.

Independent of the boiler type (drum or subcritical or once-through boilers), almost all boiler
parts that come into contact with water may be the subject to deposition. Particularly endangered
are all locations where (a) the water flow adjacent to the tube wall is disrupted or almost stagnant
(b) locations with a high heat flux and (c) locations where the nucleate boiling first initiates.
Geometrical features may also contribute to the formation of deposits. Very often, the influence
of fire-side conditions (e.g., burner misalignment) on deposit buildup is underestimated. All fire-
side conditions leading to localized overheating cause increased deposit formation.

Whereas the deposits on heating surfaces outside of the combustion chamber (e.g., economizer
tubes) are uniformly allocated around the tube circumference, the waterwall tubes experience the
(11, 32)
thickest deposits at the hot side crown area of the tube (Figure 3-3).

3-6
Deposition in Boilers

Figure 3-3
Waterwall Tube Hot Side Crown Region

Source: Reference 11

Porous deposits of corrosion products transported from the pre-boiler sections into the
waterwalls are particularly dangerous. Within the deposits, many individual microsystems may
develop that generate steam ("wick boiling," Figure 3-4). Within the deposits, circulation and
evaporation lead to an additional concentration of boiler water. Thus, a chemistry condition is
established within or underneath the deposits which is totally different from that in the bulk
boiler water.

Chemical interactions between the concentrated boiler water and magnetite or steel are the
cause of under-deposit corrosion processes that may lead to boiler tube failures.

Figure 3-4
Steam Formation in Porous Waterside Deposits

Source: Reference 11 (Adapted from Reference 32)

In addition to corrosion products transported into the boiler from the pre-boiler section, hideout
of salts dissolved in the boiler water may be a very serious source of a temporary deposit
buildup. In drum boilers operated with a phosphate treatment at low sodium-to-phosphate molar

3-7
Deposition in Boilers

ratios, severe hideout problems can occur. Hideout is a function of chemical composition,
pressure, temperature and design.

Phosphate species hidden during full-load periods can re-dissolve during low-load operation or
during standby (hideout return). Concomitantly, the sodium-to-phosphate molar ratio and the
boiler water pH decrease. This pH reduction is a possible environment-dependent component of
the synergistic system "stress and environment" which may lead to corrosion fatigue failures;
reactions of phosphate species with magnetite may be a cause of acid phosphate corrosion.

Steam-Cooled Tubing

Deposits in steam-cooled tubing are caused by mechanical carryover, chemical (volatile)


carryover and water used for attemperation. Steam-cooled tubing is also discussed in Section 4
of this report.

Mechanical carryover is strongly pressure dependent (Figure 3-5). Furthermore, the correct
function of secondary scrubbers (drum internals in natural and forced circulation boilers) for
separating water droplets from steam exiting the boiler drum depends on the mechanical
conditions of the internals. Determination of mechanical carryover at regular intervals and
appropriate maintenance measures during scheduled outages are important for assuring low
levels of mechanical carryover.

Chemical carryover is particularly noticeable at drum pressures higher than 15 MPa (about 2200
psi). A comprehensive monitoring program should be established to ensure that the normal
target values for boiler water are maintained and, thus, chemical carryover remains in a tolerable
range. (8-10, 13)

All non-volatile salts and suspended corrosion products in attemperation water may be deposited
in superheaters and reheaters. For this reason, the quality of the attemperation water with respect
to cation conductivity and the content of iron and copper should meet the requirements on
(8-9)
feedwater for once-through steam generators.

Some material that hides out in the boiler may be flushed to the superheater during load changes
and startup and shutdown. Also, once-through units may be operated in a recirculation mode
during startup, changing the amount and location of boiling surfaces and providing a means to
deposit deleterious chemicals into the superheater.

Two types of deposits in steam tubing are particularly noteworthy:


Copper deposits (copper compounds and metallic copper) and
Sodium sulfate deposits.

Copper compounds are introduced into the steam by mechanical and chemical carryover and
with attemperation water. The limited solubility of copper compounds in steam results in
deposition in the primary superheater at pressures below about 16 MPa (2300 psi). The
(10)
precipitate is typically metallic copper in primary superheater deposits.

3-8
Deposition in Boilers

Drum Pressure (MPa)


4.8 6.2 7.6 9.0 10.3 11.7 13.1 14.5 15.8 17.2 18.6 19.6
0.3

0.2
Mechanical carryover (%)

0.1

0.09

0.08

(Note: This curve includes a safety factor of 2)


0.07

600 700 900 1100 1300 1500 1700 1900 2100 2300 2500 2700 2850
Drum Pressure (psig)

Figure 3-5
Representative Drum Boiler Mechanical Carryover

Source: Reference 8

Sodium sulfate deposition is predominantly caused by mechanical carryover into the superheater
and by chemical carryover into the reheater.(14-17) Figure 3-6 demonstrates the low solubility of
sodium sulfate in the reheater.(17) Also, attemperation water containing sodium sulfate may be
another cause of sodium sulfate deposition.

3-9
Deposition in Boilers

Figure 3-6
Enthalpy-Entropy Diagram for Steam Turbines and Solubility of Sodium Sulfate (Na2SO4)

Source: Adapted from Reference 17


Note: Expansion lines correspond to a subcritical unit (540C [1004F]/540C [1004F] and
17.8 MPa [2582 psi]/4.4 MPa[638 psi]) and solubility is expressed in parts per billion, ppb, where
1 ppb = 1 g kg1

3.3 Deposits in Waterwalls

The three most significant types of deposition in waterwalls are:


Formation of rippled deposits in once-through steam generators,
Possible effects of porous deposits in drum boilers in connection with the most frequent
under-deposit corrosion (UDC) failures and
Copper deposition and behavior in drum boilers.
Each type of deposition is described in the following text discussions.

Rippled Deposits in Waterwalls of Once-through Steam Generators

The formation of rippled deposits in waterwalls (and in economizers of some units) were
investigated in the 1960s and 1970s, particularly in Germany.(1822) Fluid vortices and turbulent
bursts were found to influence the structure of deposits.(23-24)

The formation of rippled deposits in boiler tubes requires two conditions: (25)
Physical and chemical reactions of the operating fluid with the base metal and its oxide(s)
must lead to formation of a layer which adheres with an adequate strength to the metal or
metal oxide surface.
The shear stress of flow velocity on the surface (within the laminar sub-layer) must be within
certain limits such as not to wash away the adherent deposition.

3-10
Deposition in Boilers

The structure of deposits in an existing boiler can be influenced only when the first condition is
not fulfilled, i.e., when the dissolution and re-deposition of metal oxide (magnetite) is markedly
reduced. This is the case when applying oxidizing chemistry, such as OT.

Porous Deposits in Waterwalls of Drum Boilers

Three different under-deposit corrosion (UDC) failures may occur in the presence of porous
deposits in waterwalls of drum boilers:
Hydrogen damage,
Caustic gouging and
Acid phosphate corrosion.

These mechanisms all require the presence of thick, porous deposits. These failure mechanisms
differ distinctly from each other.

Hydrogen damage requires breakdown of the protective magnetite layer. This breakdown is
caused by concentration of acidic contamination within and underneath the deposits. The
mechanism of the oxide growth changes; magnetite formation does not follow the parabolic law:
and the growth rate is linear. Oxide is formed on the interface of the boiler tube metal and the
epitactic oxide. The oxide layer delaminates, and the formation of the next oxide layer on the
tube surface occurs (multi-lamination process).

Hydrogen forms on the metal surface according to the Schikorr reaction and diffuses into the
steel:
3 Fe + 4 H2O Fe3O4 + 4 H2 Equation 3-1

Hydrogen atoms react with iron carbide in the base metal and form methane:
Fe3C + 4 H 3 Fe + CH4 Equation 3-2

The microstructure is irreversibly affected by decarburization, and microfissures form in the base
metal. The final failure is thick-edged with a brittle appearance.(26)

Caustic gouging is another under-deposit corrosion mechanism. As a result of "wick boiling"


within the porous deposits, very high concentration factors (up to several thousands) may be
reached. With this mechanism, the protective magnetite layer is dissolved (fluxed) in
concentrated alkaline solutions. The typical reactions of sodium hydroxide with magnetite or
steel are:
Fe3O4 + 4 NaOH 2 NaFeO2 + Na2FeO2 + 2 H2O Equation 3-3

or
2 NaOH + Fe + Na2FeO2 + H2 Equation 3-4

3-11
Deposition in Boilers

(26)
The layered deposits contain crystals of sodium ferroate (NaFeO2) and/or sodium ferroite.

Acid phosphate corrosion is the third of the under-deposit corrosion mechanisms. A breakdown
of the magnetite protective layer is required with this mechanism, also. This breakdown occurs
as a fluxing reaction by the acid phosphate solution concentrated in the porous deposits.
According to Dooley and Paterson, the presence of maricite (NaFePO4) is a typical characteristic
of this corrosion mechanism.(27) The possible reactions leading to maricite formation are:
2 Na2HPO4 + Fe + O2 NaFePO4 + Na3PO4 + H2O Equation 3-5

2 Na2HPO4 + Fe3O4 NaFePO4 + Na3PO4 + Fe2O3 + H2O Equation 3-6

3 NaH2PO4 + Fe3O4 3 NaFePO4 + O2 + 3 H2O Equation 3-7

Acid phosphate corrosion is most likely to occur in units with significant phosphate hideout
problems originating in monosodium or disodium phosphate (MSP/DSP) additions to the boiler
water. (26)

Copper Deposition (10)

The feedwater corrosion products in units with mixed metallurgy, consisting of particulate Cu2O
(cuprous oxide) and CuO (cupric oxide), are transported through the economizer into the boiler
drum and waterwalls (Figures 3-7 and 3-8). It is rare to find copper in significant amounts in
economizer tube deposits. In waterwalls, extensive amounts of copper can be found interspersed
in the magnetite layers. In high heat flux locations, the portion of copper in deposits may range
from 20 to 50 % or even more.

Because of the reducing conditions applied in units with mixed metallurgy, metallic copper
(rather than copper oxides) is present in the deposits. During operation, most of the copper
remains on tube walls. During shutdown, if correct layup is realized (nitrogen blanketing and/or
reducing agent added to the boiler water), the dissolution of copper is minimized. Further
oxidation of metallic copper and cuprous oxide is avoided.

Without a nitrogen blanket and/or reducing agent, the boiler water may become air saturated.
Under this condition, the concentration of soluble copper species increases. An increase in pH
leads to increased copper concentration in the boiler water, too. During startup following
unprotected shutdown, the concentration of copper species in the saturated steam is relatively
high. The highest level of partitioning occurs across the temperature range from 100 to 350C
(212 to 662F). For this reason, EPRI has developed a boiler water copper concentration vs.
1 (10)
drum pressure diagram based on the copper concentration limit of 2 g kg (ppb) in steam.

3-12
Deposition in Boilers

Figure 3-7
Multilayered Deposit with Metallic Copper in Pit

Source: Reference 11 (Adapted from Reference 32)

Figure 3-8
Multilayered Deposit with Metallic Copper Plated on Metal Surface

Source: Reference 11 (Adapted from Reference 32)

3-13
Deposition in Boilers

3.4 Deposits in Superheaters and Reheaters

The two most important types of deposition in superheaters and reheaters are:
Deposition of copper species in primary superheaters of units with mixed metallurgy and
Deposition of sodium sulfate in reheaters.

Deposition in superheaters and reheaters is introduced in the following text and further discussed
in Section 4.

Copper Deposition in Primary Superheaters

EPRI-supported research work performed at the Oak Ridge National Laboratory has helped to
markedly improve the knowledge about solubility and partitioning of copper species in the plant
(28-29)
cycle. The solubility of both cuprous and cupric oxide in steam was found to be independent
1
of temperature and pH in the range of 200 to 350C (212 to 662F) and to be several g kg .
Only the neutral hydroxide forms-CuOH and Cu(OH)2-exist in steam, and their concentrations
are independent of the pH in the water phase. However, partitioning constants for both species
show strong pH dependence because the solubility of both oxides in water is strongly pH
dependent.

Since the high level of partitioning occurs across the temperature range of 100 to 350C (392 to
662F) and the solubility of cuprous and cupric oxides in the steam is very low over a wide
temperature range, any copper species in saturated steam precipitate on primary superheater
surfaces when the drum pressure is less than about 16 MPa (2300 psi). For this reason, the
startup period, even the early startup phase, is important for the copper deposition in the primary
superheaters. Mechanical carryover as a result of poor drum level control and/or poor function
of devices (drum internals) that should separate water droplets from steam may worsen (i.e., may
increase) the copper deposition in the superheater.

Boilers operating at pressures higher than the critical pressure of about 16 MPa (2300 psi) may
also be subject to copper deposition in the superheater. During startup (when the drum pressure
is lower than 16 MPa (2300 psi) or due to sliding pressure operation or to the pressure drop
between the drum and the turbine, copper species start to precipitate at intermediate points in the
superheater.

Sodium Sulfate Deposits in the Reheater

Deposition of sodium sulfate in the reheater is very often underestimated. Almost microscopic
deposits formed during operation can hardly be recognized since they do not cause any pressure
drop. Sodium sulfate solubility at pressure and temperature conditions typical for reheaters is
1
less than 0.1g - kg as Na (Figure 3-6). Thus, deposition of sodium sulfate may occur at sodium
1
concentrations of greater than 0.1g - kg and coincident trace concentrations of sulfate (> 0.2 g -
kg1).

3-14
Deposition in Boilers

This deposition is harmless during operation. However, during idle periods when the steam
condensate forms within the reheater, concentrated sodium sulfate solutions may be formed. If
incorrect layup procedures (or no layup) are applied during the unit shutdown, oxygen saturation
of sodium sulfate solution is possible and conditions for pitting damage are created.

3.5 Corrective Actions

The most important corrective actions in units suffering depositions on boiler tubes include:
Selection of optimum feedwater treatment in once-through and drum boiler cycles and
optimum boiler water treatment in units with drum boilers,
Estimation of the amount, composition and porosity of boiler tube deposits,
Implementation of chemical cleaning if the amount of deposits exceeds a tolerable deposit
weight limit and
Corrections in feedwater treatment or boiler water treatment based on analysis of boiler tube
deposit composition.

Selection of Optimum Chemical Treatment

When selecting the optimum cycle chemistry and the optimum boiler water treatment, the cycle
metallurgy and site-specific conditions (e.g., cooling water, equipment capability and start-up
and layout practices) must be considered. The continuum of treatments was first introduced by
Dooley, in 2002 (30); the most recent version of the continuum is shown in Figure 3-9. (35)

3-15
Deposition in Boilers

Figure 3-9
Summary of Boiler Water and Feedwater Treatments for Fossil Plants

Source: Reference 35

Detailed road-maps have been developed for optimizing feedwater treatment for all-ferrous
once-through and mixed-metallurgy systems and for boiler water treatment for drum boilers.(31)
Also, guidelines have been developed for optimum phosphate treatment for a drum boilers
including equilibrium phosphate treatment (EPT) and phosphate treatment (PT) and conversion
(13)
to AVT. The step-by-step optimization approach is summarized in a road-maps with detailed
explanations of the individual approach steps.

Updated road-maps covering optimization of feedwater for all-ferrous and mixed-metallurgy


feedwater systems and of boiler water treatment for drum boilers are an important part of the
(8)
revised EPRI AVT guideline.

Estimation of the Amount, Composition and Porosity of Boiler Tube Deposits

Waterside cleanliness of boiler tubes is influenced by many factors including cycle chemistry,
tube metal temperature, heat flux, layup conditions and number of operating hours. To avoid
boiler tube failures caused by excessive deposits (such as overheating and under-deposit
corrosion failures), appraisal of boiler tube cleanliness at regular intervals is required.

3-16
Deposition in Boilers

Long-term operation (more than 10 or 15 years) of boiler operation without operational chemical
cleaning is possible based on experience with units that consistently meet EPRI feedwater and
boiler water guidelines. This is true particularly for once-through boilers operated with OT. As
a rule, the deposit weight on the tubes in these boilers does not exceed 20 mg/cm2 or about 18.5
2
g/ft , even after two decades of operation. In contrast, boilers operated under conditions located
in the continuum of treatments nearer to the open end of the funnel (right hand side in Figure 3-
9) experience high deposition rates (e.g., higher than 50 mg/cm2 per year or about 46.3 g/ft2) and
require frequent chemical cleaning.

Knowledge of the amount, composition and porosity of boiler tube deposits helps in determining
the need to chemically clean a boiler. A detailed road-map for methodology used to determine
the need to chemically clean fossil plant boilers and explanations of the individual roadmap steps
(11, 32)
are provided in EPRI guidelines.

Implementation of Chemical Cleaning

If the amount of deposits exceeds an acceptable deposit weight, chemical cleaning is required to
avoid possible boiler tube failures. Numerous chemical cleaning issues need to be addressed to
assure a successful cleaning. This includes determination of the need to clean, solvent selection
and application, planning for chemical cleaning, chemical cleaning procedures for boilers, post-
cleaning inspections, cleaning of steam-cooled tubing and parts, cleaning of other cycle
components and cleaning wastes. (32) One significant difference between European cleaning
practices and those in the United States is the common use of inhibited hydrofluoric acid for
routine and pre-operational chemical cleaning of European units.

3.6 Factors Linked to the Matrices

The factors given higher scores (4 to 2) for deposition in boilers (originally presented in Tables
2-1 to 2-3 of Part 1 of this study (1)) have been abstracted from these tables and are listed in Table
3-3.

High and medium ranked parameters with respect to suspended particles in alphabetical order
were:
High (score of 4 or higher): Chemical treatment, concentrations of particles, condensate
polishing and filtration, pH at operating temperature and transient conditions and outages.
Medium (score of 3): Agglomeration of particles, concentrations of metal ions, impurities,
oxygen and oxidizing agents, configuration of hardware (steam/water flow orientation), flow
characteristics, oxide/hydroxide composition on tubes/surfaces, velocity and flow rate and
zeta potential.

High and medium ranked parameters with respect to deposition of soluble impurities and
corrosion products in alphabetical order were:
High (score of 4 or higher): Concentration of impurities and concentrations of metal ions.

3-17
Deposition in Boilers

Medium (score of 3): Condensate polishing and filtration, concentrations of oxygen and
oxidizing agents, pH at operating temperature, solubility in steam and water, solubility in
steam and water, steam quality and transient conditions and outages.

High and medium ranked parameters with respect to surface phenomena (sorption, hideout, etc)
in alphabetical order were:
High (score of 4 or higher): Concentrations of metal and non-metal ions, impurities and
particles.
Medium (score of 2-3): Heat flux, ion exchange properties of metal oxides, mass transport,
porosity, temperature, and transient conditions and outages.

The factors affecting boilers listed in Table 3-3 are reviewed in alphabetical order.
Chemical treatment can reduce the generation of soluble and insoluble corrosion products in
boiler. Chemical treatment includes the addition of pH control agents, reducing agents and
oxidizing agents. Chemical additives may alter the nature and extend of deposits in boilers.
The concentration of metal ions is a measure of the soluble corrosion products generated,
which contribute to the deposition of suspended particles and water soluble material. The
concentration in the bulk liquid depends on corrosion rate of the materials exposed, chemical
treatment, pH, concentration of oxidizing/reducing agents, electrochemical potential (ECP)
and/or oxidation reduction potential (ORP), and the solubility in water/steam. Metallic ions can
be converted to suspended oxide particles as a result of corrosion. Transport within the boiler
(e.g. transition zone within once-through boilers) may occur as a result of solubility in steam
and/or transport of moisture droplets (steam quality).
The concentration of non-metal ions and impurities (e.g. chloride, fluoride, sulfate, carbonate,
organics, etc.) contributes to corrosion and the generation of soluble and insoluble corrosion
products. Non-metal ions and impurities can also participate indirectly by changing solution pH.
Impurities can dissolve in boiler recirculation water or steam and precipitate in the superheated
area.
The concentration of particles is a measure of the insoluble corrosion products generated,
which can deposit on the heat transfer surfaces of the boiler. The concentration of particulate
corrosion products is a key factor in influencing deposition rates.
Condensate polishing reduces corrosion of the boiler system by removing soluble, ionic
impurities by ion exchange. It also removes suspended material by filtration.
The configuration of the water/steam circuit hardware can contribute to the deposition of
suspended particles, including corrosion products. Water and steam flow can generate soluble
and insoluble corrosion products by erosion and/or flow assisted corrosion (FAC). Other factors
depending on hardware include high or low velocity, susceptible materials, pH, chemical
treatment, electrochemical potential (ECP) and oxidation/reduction potential (ORP). Hardware
configuration can alter deposition of suspended particles by changes in direction and velocity of
the fluid.
The rate of general corrosion, erosion and FAC affects the generation of soluble and insoluble
corrosion products and suspended solids function of the materials selected in design and other
boiler design parameters (elbow or collector positioning etc.). Carbon steel and low alloy steels

3-18
Deposition in Boilers

are the primary sources. Correct design material selection and chemical treatment (pH, oxygen
control, etc.) are the primary methods of controlling corrosion. Copper deposition can be a
particularly difficult problem for boilers in units with copper alloys in the feedwater heaters.
Crystalline structure of deposit can affect the nature of the soluble and insoluble corrosion
products from corrosion. Some deposit can be non-crystalline (amorphous). The
crystalline/amorphous structure of deposits may affect the generation, deposition, release and
movement of suspended particles.
Heat flux can affect the nature of the deposition and the concentration of impurities and ionic
contaminants (Cl, SO4 etc.) in the deposit as well as the concentration of metal ions in the
deposit. Higher heat fluxes result in higher deposition and corrosion rates in steam generators
and influence the DNB location and value.
The ion exchange properties of metal oxides, suspended solids and deposits can affect the
deposition in boiler water system or superheater. Adsorbed ions on these materials can exchange
with the ions in the water.
Electrochemical potential (ECP) and oxidation reduction potential (ORP) of deposits are
increased by oxidizing agents. Reducing agents decrease the ECP and ORP.
The pH at operating temperature affects the generation of soluble and insoluble corrosion
products, the volatility of contaminants in steam carryover and the deposition of suspended
particles in superheaters. The neutral point of water changes with temperature. The pH of the
chemical solution must be evaluated relative to the neutral pH of water.
Porosity (density) of oxides and deposits of suspended metal oxide particles provide sites that
can enhance deposition and concentrating of water soluble material inside the deposits on heat
exchanger surfaces of boilers (especially the deposits on the fireside of the boiler since the
fireside deposits can contribute to elevated temperatures inside the tubes and, finally to creep
failures).
Steam quality (carryover) can affect the transport and deposition of soluble and insoluble
material in the boiler superheater section. Carryover consists of mechanical and vaporous
carryover. The steam separation equipment in drum boilers is designed to separate most of the
moisture from the steam. The mechanical carryover of moisture droplets results in the transport
of additional water soluble material to the superheater. Since many water soluble chemicals are
much more soluble in the moisture than in steam, moisture carryover significantly enhances the
transport of water soluble chemicals with the steam. Transport of suspended particles with the
water can also occur.
Steam and water properties, including density, viscosity, entropy and enthalpy, contribute to
the deposition of soluble and insoluble material from water and steam. Steam and water
properties are a function of temperature and pressure.
Surface roughness contributes to the deposition of suspended particles and water soluble
material present on the surface of particles. Deposits tend to accumulate in the troughs formed
by machining and polishing marks. Surface roughness increases the deposition of particles on
the surface and also reduces the rate of release of deposited particles from the surface. The
impact of surface finish is strongly dependent on bulk fluid velocity and particle size.
Temperature is a contributor to the generation of soluble and insoluble corrosion products and
deposition of water soluble material from steam. The solubility of water soluble material in

3-19
Deposition in Boilers

steam is a function of the steam density, which is itself a function of temperature and pressure.
Increasing temperature through the feedwater system can enhance the generation of corrosion
products. Some of the material deposited in the boiler can be transported to the superheater as a
result of solubility in steam and/or transport with moisture droplets.
Time is a factor in the generation of soluble and insoluble corrosion products, deposition of
suspended particles and water soluble material. The rate of generation and deposition normally
decreases with time.
Transient operations, including load changes, cycling and load follow operation, shutdown,
startup and outages, contribute to the generation of soluble and insoluble corrosion products and
deposition, release and re-deposition of suspended particles and water soluble material. Existing
deposits may be released and redeposit at other locations. The impact is generally proportional
to the size of the transient. Transients can alter the rate of deposition and deposit location
throughout the cycle. Layup practices (use of nitrogen, reducing agents and blanketing) during
outages are also a factor in formation and release of deposits. Modeling of plant chemistry data
from nuclear power plants has confirmed field data that show significant transportation and re-
deposition of soluble chemical impurities from steam generators during relatively small load
changes.
Transport of soluble material and impurities on or within suspended particles contributes to
the deposition of water soluble contaminants and possibly to underdeposit corrosion of the
surface onto which the particles deposit. Adsorption and ion exchange may be involved. Other
factors include condensate purification, chemical treatment, the concentration and composition
of impurities, corrosion rate, oxidizing agents, ECP/ORP, particle size, pH, porosity and
solubility in steam.
Tube surface composition (carbon steel, alloying materials, etc.) affects the generation of
soluble and insoluble corrosion products and deposition of suspended particles. The effective
tube surface composition changes as oxides and deposits accumulate, obscuring or partially
obscuring the tube metal from the bulk fluid. Porous deposits can entrap and concentrate
corrosive impurities.
Velocity of the fluid contributes to the generation of soluble and insoluble corrosion products,
deposition of suspended particles and deposition/sorption of water soluble material. Velocity is a
function of flow rate, hardware configuration and steam/water properties and contributes to
generation of corrosion products by erosion and FAC.

3-20
Deposition in Boilers

Table 3-3
Factors given High Scores in Tables 2-1 to 2-3 of Part 1 of Reference 1 for Deposition in
Boilers

Deposition of Suspended Deposition of Water Soluble Adsorption and Ion


Solids Material Exchange on Surfaces
pH at operating temperature Concentration of impurities and Concentration of impurities
ionic contaminants (CO2, Cl, SO4, and ionic contaminants
etc) (CO2, Cl, SO4, etc)
Transient conditions during Transient conditions during various
various plant modes (base plant modes (base load, cycling,
load, cycling, etc) including etc) including use of nitrogen,
use of nitrogen, reducing reducing agents and blanketing
agents and blanketing during during layup
layup
Condensate polishing pH at operating temperature
(including filtration capacity
for particle removal, shift in
soluble/insoluble fractions
and impact of resin leakage
on deposits
Oxygen and oxidizing agents Solubility in water and precipitation
including copper of water soluble compounds
Zeta potential Condensate polishing
Agglomeration of particles Steam quality (fraction of steam and
water)
Tube surface oxide/hydroxide Concentration of oxygen and
composition, thickness, oxidizing agents including copper
density and porosity
Tube surface composition Solubility in steam and steam/water
(carbon steel, alloying distribution
materials, etc.)
Concentration of particles
Steam/water flow
characteristics
Flow regimes of steam and Attraction of soluble
water including T/H transients materials to suspended
and oscillations solids, including suspended
corrosion products by ion
exchange, absorption and
adsorption
Corrosion rate of materials of Tube surface oxide/hydroxide Transient conditions during
construction in prevailing composition, thickness, density and various plant modes (base
chemistry porosity load, cycling, etc) including
use of nitrogen, reducing
agents and blanketing
during layup
Chemical treatment (pH Chemical treatment (pH control,
control, reducing or oxidizing reducing or oxidizing agents, etc)
agents, etc)
Concentration of metal ions
and impurities

3-21
Deposition in Boilers

Deposition of Suspended Deposition of Water Soluble Adsorption and Ion


Solids Material Exchange on Surfaces
Hardware configuration and Porosity (density) of oxides and Porosity (density) of oxides
steam/water flow deposits and deposits
Surface finish
Velocity and flow rate Heat Flux Heat Flux
Crystalline structure and
amorphous material
Particle size distribution
Mass transport (local and Mass transport Mass transport
bulk)
Time, including deposition Time, including deposition rate
rate changes changes
Dynamics of deposition and
release (including exfoliation)
Pressure Pressure
Temperature Temperature Temperature
Steam/water properties Steam/water properties (density,
(density, viscosity, etc) viscosity, etc)
Steam quality(fraction of
steam and water)
Steam/water boiling and Steam/water boiling regimes
condensing regimes including DNB
Ion exchange properties of
metal oxides
Note: The factors in the upper, center and lower parts of the table were given scores of 4, 3 and 2,
respectively, in Tables 2-1 to 2-3 of Reference 1

3-22
Deposition in Boilers

3.7 State-of-Knowledge Deficiencies

The chemical aspects of deposition in utility boiler tubes (steam and water cooled) have been
researched extensively in boiler tubes over the last 10 years. These units are operated with AVT
or OT. This research has been incorporated into EPRI guideline documents. Unfortunately,
these guidelines are not always applied to plant operation and layup. The costs of plant cycle
chemistry operation outside of the optimum chemistry target values and missing or inadequate
layup procedures are underestimated by many utilities. In many cases, the root cause analysis of
forced outages concentrates on non-chemical phenomena as a result of the prominence of other
operative aspects and the expense and relative slowness of deposit chemistry analyses.

Deposition in boiler tubes containing water and/or steam is the result of transport processes
within the plant cycle. Contaminants and corrosion products entering or generated in the pre-
boiler area and may deposit in the boiler or be involved in reactions resulting in deposition. Any
measures leading to a reduction in the corrosion product concentration in the feedwater and to
the avoidance of contaminant ingress markedly contribute to a reduction in deposition.

Application of state-of-the-art feedwater and boiler water treatments, as outlined in current EPRI
guidelines, creates conditions that minimize deposition and generally provide for long-term
operation of boilers without frequent chemical cleaning to remove those deposits. Although the
current guidelines are supported by research and operating experience, comprehensive model of
the deposition processes in critical areas is needed to further minimize deposition and to evaluate
changes in operating conditions. For example, the quantitative effect of altering operational
conditions on deposition is not sufficiently known. Deposition in water tubes is highly
dependent on numerous factors, including:
Heat flux
Circulation,
Boiler tube temperatures,
Feedwater and boiler water purity and
Type of feedwater and boiler water treatment.

Heat flux, circulation and boiler tube temperatures are design parameters that may be altered
when the operating conditions are changed substantially from the original design. These changes
may not be recognized by plant management as having potentially significant impact on
deposition and boiler tube failures. Such changes may occur under one or more of the following
(11)
conditions:
Coal firing is changed to oil or gas firing,
Coal with higher caloric value is used,
A different type burner is installed,
The burner angle is altered,
The boiler is fired harder because a feedwater heater is out of service or

3-23
Deposition in Boilers

A unit is changed to overpressure operation when the unit was not designed for such
operation.

The application of a comprehensive model would provide insight into current operating practices
and proposed changes in those practices. Also, some changes in current practices may be
suggested as a result of the modeling effort.

Additionally, an objective scientific-based evaluation of the influence of organic chemical


additives on corrosion and deposition processes is still missing. These organic additives are
often used in boilers that are operated by refineries, paper mills, sugar refineries and chemical
plants. Polyamines, volatile amines, dispersants and various oxygen scavengers are utilized.
Catalysts and stabilizers are added to many of these chemicals. Some of these additives are
intended to control deposition within boiler tubes. The behavior of dispersants has been studied
(33)
in a trial in a nuclear power plant, but not under the conditions found in a fossil cycle. Product
purity requirements need to be established and the products need to be shown to have no
unintended deleterious effects on cycle chemistry and materials. Then, the efficiency of the
products to perform the intended function needs to be evaluated and application guidelines need
(34)
to be established.

3.8 References

1. State-of-Knowledge on Deposition, Part I: Parameters Influencing Deposition in Fossil


Boilers, EPRI, Palo Alto, CA: 2002. 1004194.

2. Deposition in Boilers: Review of Soviet and Russian Literature, EPRI, Palo Alto, CA: 2003.
104193.

3. H. Kirsch and S. Pollmann, Chem.-Ing. Technik, 24(18), 1968, pp 897903.

4. American Society for Testing and Materials. Annual Book of ASTM Standards: Water and
Environmental Technology, Volume 11.02, Water (11), ASTM, Philadelphia, PA, 1989.

5. Cycle Chemistry Corrosion and Deposition: Correction, Prevention, and Control, EPRI,
Palo Alto, CA: 1993. TR-103038.

6. Revised Guidelines for Makeup Water Treatment, EPRI, Palo Alto, CA: 1999. TR-113692.

7. Condensate Polishing Guidelines, EPRI, Palo Alto, CA: 1996. TR-104422.

8. Cycle Chemistry Guidelines for Fossil Plants: All-Volatile Treatment: Revision 1, EPRI, Palo
Alto, CA: 2002. 1004187.

9. Cycle Chemistry Guidelines for Fossil Plants: Oxygenated Treatment, EPRI, Palo Alto, CA:
1994. TR-102285.

10. Guidelines for Copper in Fossil Plants, EPRI, Palo Alto, CA: 2000. 1000457.

3-24
Deposition in Boilers

11. K. J. Shields and R. B. Dooley, Proc. of Conference on Boiler Tube Failures, HRSG Tube
Failures, and Inspections (Phoenix, Arizona), EPRI, Palo Alto, CA: 2001.

12. K. J. Shields and R. B. Dooley, Power Plant Chemistry, 2002, 4(11), p 671.

13. Cycle Chemistry Guidelines for Fossil Plants: Phosphate Treatment, EPRI, Palo Alto, CA:
1994. TR-103665.

14. M. Ball and M. A. Jenkons, Proc. of Second Fossil Plant Cycle Chemistry Conference, EPRI,
Palo Alto, CA: 1998-99. GS-6166.

15. K. Daucik, Power Plant Chemistry, 2001, 3(5), p 280.

16. K. Daucik, Proc. of Fourth International Conference on Fossil Plant Cycle Chemistry, EPRI,
Palo Alto, CA: 1995. TR-104502.

17. J. P. Jensen and A. Bursik, Proc. of Steam Chemistry: Interaction of Chemical Species with
Water, Steam, and Materials during Evaporation, Superheating, and Condensation, (Ed: B.
Dooley and A. Bursik), EPRI, Palo Alto, CA: 2000. TR-114837.

18. W. Schoch, VGB-Mitteilungen, 48(4), 1968, p 239.

19. W. Schoch, R. Richter and H. Khle, VGB-Mitteilungen, 49(3), 1969, p 202.

20. W. Schoch, R. Richter and P. H. Effertz, Der Maschinenschaden, 43(12), 1970, p65.

21. H. Schuster, Allianz-Berichte, 16(4), 1971, p 28.

22. W. Schoch, H. Wiehn, R. Richter and H. Schuster, VGB-Mitteilungen, 52(3), 1972, p 228.

23. J. S. Kline, W. C. Reynolds, F. A. Schraub, P. W. Rundstadler, J. Fluid Mech., 30(4), 1967, p


741.

24. W. R. B. Morrison, K. J. Bullock and R. E. Kronauer, J. Fluid Mech., 47(4), 1971, p 639.

25. B. Pfau, Technik der Wrmekraftwerke, VCH Verlagsgesellschaft mbH, Weinheim,


Germany, 1987, 73.

26. R. B. Dooley and W. McNaughton, Boiler Tube Failures: Theory and Practice, EPRI, Palo
Alto, CA: 1966. TR-105261-V1V3.

27. R. B. Dooley and S. Paterson, Proc. of International Water Conference, Engineering Society
of Western Pennsylvania, Pittsburgh, PA, 55, 1994, p 420.

28. Behavior of Aqueous Electrolytes in Steam Cycles: The Solubility and Volatility of Cupric
Oxide, EPRI, Palo Alto, CA: 2000. 1000455.

29. Behavior of Aqueous Electrolytes in Steam Cycles. The Solubility and Volatility of Copper (I)
and Copper (II) Oxides, EPRI, Palo Alto, CA: 2001. 1003993.

3-25
Deposition in Boilers

30. R. B. Dooley, Power Plant Chemistry, 4(6), 2002, p 320.

31. Selection and Optimization of Boiler Water and Feedwater Treatments for Fossil Plants,
EPRI, Palo Alto, CA: 1997. TR-105040.

32. Guidelines for Chemical Cleaning of Conventional Fossil Plant Equipment, EPRI, Palo
Alto, CA: 2001. 1003994.

33. K. Fruzzetti, P. Frattini, P. Robbins, A. D. Miller, R. D. Varrin and M. Kreider, Power Plant
Chemistry, 4(9), 2002, p 513.

34. A. Bursik, Power Plant Chemistry, 3(8), 2001, p 459.

35. B. Dooley and K. Shields, Cycle Chemistry for Conventional Fossil Plants and Combined
Cycle/HRSGs, 7th International Conference on Fossil Plant Cycle Chemistry, EPRI, Palo
Alto CA. 1009194. 2003.

3-26
4
DEPOSITION IN SUPERHEATERS AND REHEATERS

4.1 Introduction

This section of the report reviews deposition in superheaters, reheaters and steam piping.
Additional discussion is provided in Section 3.4. The 2002 EPRI Part 1 report on deposition
state-of-knowledge (1) and the 2003 EPRI report on Soviet and Russian literature contain
additional details and mathematical descriptions of boiler deposition processes. (2)

Deposit formation is possible in superheaters, reheaters and steam piping. Deposition in these
areas can be caused by:
Contamination in feedwater and steam,
Contamination in attemperation water,
Boiler mechanical and vaporous carryover,
Contamination from layup practices and/or
Contamination from chemical cleaning practices.

The superheater and reheater are fabricated from ferrous alloys, such as 1.25% chrome-0.5%
molybdenum and 2.25% chrome-1.0% molybdenum, and austenitic stainless steels, depending
on pressure and temperature requirements.

Deposits in superheaters, reheaters and steam piping can be controlled by:


Maintenance of proper water and steam chemistry,
Maintenance of designed boiler scrubber efficiency (drum units),
Elimination of high boiler water levels and foaming (drum units) and/or
Adoption of proper layup and chemical cleaning procedures.

Chemical cleaning is the only option for removal of deposits in these tubes.

4.2 Locations Subject to Deposition

Normally, deposition from vaporous carryover during plant operation is not expected if steam
purity is maintained as needed to protect the steam turbine. Conversely, if the unit is operated
with impure steam, the development of deposits on surfaces comprising the steam path becomes
more likely. This is particularly true in units with drum boilers operated on solids based

4-1
Deposition in Superheaters and Reheaters

treatments (phosphate or caustic) and subject to high rates of mechanical carryover, a situation
which can lead to formation of sodium salt deposits.

Copper deposition can occur in primary superheaters of drum units in cycles with mixed
metallurgy in the condensate and feedwater components. Copper deposition in the primary
superheater primarily results from volatile carryover during startups. This phenomenon is
presented, along with other findings resulting from investigations of the science of copper in
(3)
fossil cycles, in an EPRI guideline on copper.

Oxides can form in superheater and reheater tubes by a reaction between steam and iron, a
phenomenon which progresses with time. Oxidation of superheater and reheater internal
surfaces can result in exfoliation of oxides, after a number of operating years, and severe damage
(solid particle erosion or SPE) to steam turbine components.

4.3 Deposit Mechanisms

The composition of deposits formed by mechanical and vaporous carryover and attemperation
water will depend on the composition of contaminants entering the superheater and reheater and
their solubility in steam at superheater and reheater operating conditions.

In drum boiler cycles, a major source of deposits in the superheater, and to a lesser extent, in the
reheater, results from both mechanical and vaporous carryover from the boiler. Mechanical
carryover will be present in all drum boilers. Typical values of mechanical carryover were
shown earlier in Figure 3-5, which is based on a consensus of boiler manufacturers. The values
shown in Figure 3-5 contain a factor of two to provide conservatism.

In once-through units, all contaminants entering the boiler either (a) deposit in the boiler or (b)
transport with the steam (vaporous carryover) and possibly deposit in the superheater, turbine
and/or reheater.

Vaporous carryover depends on the chemical species involved, the feedwater and boiler water
treatment and the boiler pressure and temperature. Vaporous carryover was the subject of an
EPRI research project that was conducted at Oak Ridge National Laboratory over the last
decade. Additional discussion is provided in Section 5.

Mixed metallurgy feedwater systems have feedwater heaters comprised of various copper alloys,
such as Admiralty brass, copper-nickel and Monel. All-ferrous systems generally have stainless
steel tubes, and, much less commonly today, carbon steel tubes. All once-through units and
many drum type units feature all-ferrous systems. However, mixed metallurgy systems are still
common in drum type units. Either mixed metallurgy or all-ferrous systems may have copper
alloy condenser tubes.

Copper and copper compounds from mixed metallurgy systems are introduced into the steam
from a drum boiler by both mechanical and vaporous carryover and with attemperation water.
The limited solubility of copper compounds in steam results in deposition in the primary
(3)
superheater. Surfaces where saturated steam is converted to superheated steam are highly
susceptible to deposition at pressures below 2300 psi (about 16 MPa).

4-2
Deposition in Superheaters and Reheaters

Copper deposits in primary superheaters are usually present as crystalline, metallic copper, as
(3-5)
shown in Figures 4-1and 4-2 , indicating that:
Much of the copper was transported in vaporous rather than particulate form and
Deposition in superheaters occurs as a result of exceeding a solubility limit.

Steam solubility considerations indicate that copper deposited in the superheater at low pressure
can reenter the steam at higher pressures. Thus, the primary superheater appears to represent an
important deposition/release point along the path by which copper compounds are transported to
the high-pressure turbine.

Figure 4-1
Copper Deposits (Visible as Bright Spheres) on Superheater Tube Magnetite Scale at
Magnification of 500x. Material is SA213-T11 with 16 Years of Service.

Source: References 3-5

4-3
Deposition in Superheaters and Reheaters

Figure 4-2
Crystalline Metallic Copper Deposits on Primary Superheater Tube at 2000x Magnification.
Material and Service Conditions are the same as in Figure 4-1

Source: References 3-5

The attemperation water can be a source of deposition in both superheaters and reheaters. This
source of deposition is normally minimal, but can become important if the source (usually
feedwater) is contaminated or if an excessive quantity of attemperation water is required to
maintain final superheater and/or reheater temperature.
(6)
Superheater and reheater problems are frequently caused by contamination or improper layup.
Introduction of abnormal quantities of salts or other solids (in addition to carryover and
attemperation water) can be caused by:
Poor chemical cleaning procedures which allow the introduction of chemicals to the
superheater and reheater,
Poor layup procedures prior to initial operation of the boiler or after long outages,
Poor layup procedures associated with forced outages and
Contaminants introduced during startup when steam tubing may be used as boiling surfaces.

Pendant superheater and reheater tubes cannot be drained and, thus, contaminants that are
introduced during startup, operation, shutdown and layup may corrode superheater and reheater
tubing if condensate forms during shutdown, and especially if oxygen is subsequently introduced
during the outage.

4-4
Deposition in Superheaters and Reheaters

These sources of contamination can lead to general corrosive attack, overheating (impaired heat
transfer or flow blockage by internal deposits) and stress corrosion cracking of austenitic
stainless steels.

The inner surfaces of the superheater and reheaters oxidize in the presence of steam to form
(6)
oxide layers. In many cases, this steamside oxide periodically spalls away from the metal
surface in the form of flakes. This phenomenon is termed exfoliation. The exfoliated scale
can be transported at very high velocities into the steam turbines causing extensive solid particle
erosion (SPE) of both stationary and rotating blades in the front-end stages of the high- and
intermediate-pressure sections. This form of material damage has been recognized as a severe
and expensive problem for many utilities. Some examples are shown in Figures 4-3, 4-4 and
(7)
4-5.

The reheater tube in Figure 4-3 shows a considerable amount of high temperature oxidation.
When this iron oxide scale exfoliates, it causes serious damage to turbines, diaphragms and
valves as it can be seen in Figures 4-4 and 4-5. In Figure 4-4, the extensive damage done to
turbine nozzle blocks by exfoliated scale is clearly visible. Figure 4-5 presents erosion damage
to turbine blades by exfoliated oxide.

Figure 4-3
Reheater Tubes with Scale

Source: Reference 7

4-5
Deposition in Superheaters and Reheaters

Figure 4-4
Damage to Turbine Nozzle Blocks by Exfoliated Scale

Source: Reference 7

Figure 4-5
Erosion Damage to Turbine Blades by Exfoliated Scale

Source: Reference 7

4-6
Deposition in Superheaters and Reheaters

4.4 Methods to Control or Prevent Deposition

In drum units, carryover from the boiler drum is an important source of contamination and
deposition in superheaters and reheaters. Total carryover on drum units must be monitored with
regularity, and compared to the mechanical carryover obtained from Figure 4-1. Values that are
slightly above those in Figure 4-1, perhaps, can be attributed to vaporous carryover. Values that
significantly exceed those in Figure 4-1 are indicative of (a) malfunction of boiler internals
(scrubbers), (b) high boiler water levels and/or (c) foaming. Such conditions should be
immediately corrected.

The deposition of copper in the superheater tubes can be prevented by replacing all copper-
bearing heater tubes with stainless steel, obviously a very costly solution. Replacement of
copper-bearing components may not completely eliminate copper transport if significant
amounts of copper has been transported and deposited in other components. Other less drastic
solutions for minimizing the deleterious effects of copper in the heat cycle include sort term
(3)
approaches, long term approaches and chemistry monitoring approaches.

Short term approaches include:


Cleaning the turbine,
Cleaning the feedwater heaters and
Cleaning the superheater.

Long term approaches include:


Changing the cycle design and piping where necessary,
Retrofitting condensate polishers,
Minimizing mechanical carryover,
Minimizing cycle air inleakage,
Mimimizing oxygen in cycle makeup,
Nitrogen blanketing boiler and heaters,
Optimizing feedwater heater options, and
Minimizing the extent of overpressure operations.

Chemistry monitoring approaches include:


Upgrading sampling system,
Installing an ORP monitor and
Monitoring copper transport.

For both drum type and once-through units, the feedwater limits (relative to attemperation water
and carryover) and steam limits shown on the cycle diagrams of the applicable EPRI guidelines
will minimize deposits on the superheater and reheater surfaces. (3, 8-11) Careful monitoring of

4-7
Deposition in Superheaters and Reheaters

feedwater and steam samples, as outlined in these guidelines, is required to determine the
comparison between actual and recommended maximum contaminant concentrations. The
prevention of contamination from layup and chemical cleanings is addressed in the EPRI layup
(4, 12)
guidelines and the chemical cleaning guidelines, respectively.

The formation of oxide scale in the superheaters and reheaters is, unfortunately, a natural
phenomenon, which progresses with age and operational aspects. No practical preventative
measures for this phenomenon are known at this time. However, it is possible to extend
component life by chemical cleaning and perform remaining life assessments to minimize the
impact of creep damage on unit availability. (13)

4.5 Diagnostic and Corrective Actions

Oxide formation in superheater and reheater tubes can be diagnosed based on an examination of
tube samples, increased pressure drop, and the existence of SPE in the turbine. Excessive oxide
formation may collect in bends causing steam starvation, overheating and tube failures.

The presence of copper in superheater tubes can be expected if there are high concentrations of
copper in the feedwater and decreased turbine efficiency caused by copper deposits in the
turbine. The presence of copper and other contaminants can be assessed by (a) analysis of
condensate, feedwater and boiler water samples, (b) analysis of representative tube samples and
(c) inspections of the feedwater heaters and turbine.

Chemical cleaning is the only known method of removing excessive iron oxides from
superheater and reheater steam surfaces. Chemical cleaning for removal of copper is a relatively
new development, but it is an effective choice when this problem occurs. Procedures for
chemical cleaning of superheaters and reheaters are given in the EPRI Chemical Cleaning
(5)
Guidelines. It is very important to realize that chemical cleaning of the primary superheater
(3)
will be needed in any effective utility copper program.

4.6 Factors Linked to the Matrices

The factors given higher scores (4 to 2) for deposition in superheaters and reheaters (originally
presented in Tables 2-1 to 2-3 of Part 1 of this study (1)) have been abstracted from these tables
and are discussed as follows.

Superheater/Reheater high and medium ranked parameters with respect to suspended particles in
alphabetical order were:
High (score of 4 or higher): None.
Medium (score of 3): pH at operating temperature and transient conditions and outages.

Superheater/Reheater high and medium ranked parameters with respect to deposition of soluble
impurities and corrosion products in alphabetical order were:
High (score of 4 or higher): None.

4-8
Deposition in Superheaters and Reheaters

Medium (score of 3): Steam quality.

Superheater/Reheater high and medium ranked parameters with respect to surface phenomena
(sorption, hideout, etc) in alphabetical order were:
High (score of 4 or higher): None.
Medium (score of 2-3): Concentrations of metal and non-metal ions, impurities and particles,
ion exchange properties of metal oxides, and transient conditions and outages.

Principal mechanisms for damage in superheaters and reheaters are:


Deposition of the corrosion products in the superheater and reheater,
Corrosion damage (including SCC) in the superheater/reheater areas,
Concentration and dilution of acidic species, alkaline species or hydrogen,
Crystallization and adsorption of hydrous metal oxides due to mechanical and vaporous
carryover from the boiler drum and
Excessive creep under load swings.

Possible root causes of superheater or reheater damages are inadequate or inappropriate


chemistry control in the feedwater and condensate systems and/or the boiler Other important
factors are improper startup, shutdown or cycling operations, improper drum level and
contaminated attemperation water.

Mechanical carryover will be present in all drum boilers. Mechanical carryover depends on the
design, operation and maintenance of the boiler. Vaporous carryover depends on the chemical
species involved, the feedwater and boiler water treatment and the boiler pressure and
temperature. Chemicals that are soluble in high pressure steam may deposit in the superheater or
reheater as the solubility changes with changes in temperature and pressure.

4.7 State-of-Knowledge Deficiencies

No specific state-of-knowledge deficiencies exist with respect to deposition in the superheaters


and reheaters. Deposit minimization requires reducing the input of copper and sulfates.

4.8 References

1. State-of-Knowledge on Deposition, Part I: Parameters Influencing Deposition in Fossil


Boilers, EPRI, Palo Alto, CA: 2002. 1004194.

2. Deposition in Boilers: Review of Soviet and Russian Literature, EPRI, Palo Alto, CA: 2003.
104193.

3. Guidelines for Copper in Fossil Plants, EPRI, Palo Alto, CA: November 2000. 10000457.

4-9
Deposition in Superheaters and Reheaters

4. Courtesy A. G. Howell, New Century Energies, 2000.

5. A. G. Howell, Mitigation of Copper Deposition in High Pressure Turbines of Utility Drum


Boilers, Power Plant Chemistry, 1(4), October 1999.

6. The ASME Handbook on Water Technology for Thermal Systems, Paul Cohen-Editor-in-
Chief, ASME and EPRI Research Project RP 1958-1, Copyright 1989.

7. Guidelines for Chemical Cleaning of Conventional Fossil Plant Equipment, EPRI, Palo Alto,
CA: November 2001. 1003994.

8. Cycle Chemistry Guidelines for Fossil Plants: Phosphate Treatment for Drum Units, EPRI,
Palo Alto, CA: December 1994. TR-103665.

9. Cycle Chemistry Guidelines for Fossil Plants: Oxygenated Treatment, EPRI, Palo Alto, CA:
December 1994. TR-102285.

10. Sodium Hydroxide for Conditioning the Boiler Water of Drum Type Boilers, EPRI, Palo Alto,
CA: January 1995. TR-104007.

11. Cycle Chemistry Guidelines for Fossil Plants: All-Volatile Treatment: Revision 1, EPRI, Palo
Alto, CA: 2002. 1004187.

12. Cycling, Startup, Shutdown, and Layup Fossil Plant Cycle Chemistry Guidelines for
Operators and Chemists, EPRI, Palo Alto, CA: August 1998. TR-107754.

13. R. B. Dooley and W. McNaughton, Boiler Tube Failures: Theory and Practice, EPRI, Palo
Alto, CA: 1966. TR-105261-V1V3.

4-10
5
TURBINES

5.1 Introduction

The practical way to minimize damage within existing steam turbines that have a set material
and stress distribution is to control the environment in which those materials are operated.
Environmental control is accomplished by controlling the chemistry of the steam that enters the
turbine. This section of the report reviews the status of knowledge of turbine steam path damage
mechanisms, highlighting the role of cycle water-steam chemistry on deposition, corrosion and
erosion.

The most serious turbine deposit problems are (a) failures from corrosion fatigue and stress
corrosion cracking since these failures can lead to long and often unplanned outages and (b)
significant reductions in turbine performance leading to as much as 4 to 8% reduction in power
(as discussed in Section 5.2). Due to the utmost importance of chemistry processes in the above
performance changes and corrosion damage, it is appropriate to recap the several steps involved
in the deposition deleterious materials within steam turbines:
Formation of corrosion products in the feedwater heaters and boiler,
Formation of solids from low solubility impurities in the water,
Transport of the deposits to the steam circuit and
Transport of water soluble impurities to the steam circuit.

Corrosion products, as well as any impurities present in the feedwater, may be transported
through the boiler to the turbine. Deposits can form in the superheater and turbine, especially if
the unit is being operated with impure steam. Deposition of sodium salts in the turbine can occur
in all units, including drum boilers operated on non-volatile alkalis (phosphate or caustic) with
high rates of mechanical carryover.

For units with mixed metallurgy in the feedwater systems, copper can deposit in the primary
(1)
superheater and the high pressure (HP) turbine as a result of carryover during unit startups.
Copper deposition in the HP turbine is one of several turbine deposition and damage
mechanisms.

Deposition in turbines usually results from the presence of chemicals in the steam at
concentrations in excess of the solubility in steam. A chemical may be soluble in higher pressure
steam at the turbine inlet and become insoluble as the pressure and temperature decreases
through the turbine. Purity criteria for steam are based on the understanding of solubility in
steam under all operating conditions.

5-1
Turbines

Deposits may be formed either in situ, as a result of chemical reaction (usually oxidation) with
the base metal or due to transport of corrosion products from another part of the cycle. The
presence of impurities in the cycle presents opportunities for various chemical reactions and
transport processes to take place. Potential sources of contaminant ingress include makeup
water, condenser leaks, combustion products, poor maintenance and plant operation. Deposit
formation can be affected also by the feedwater and boiler water treatments. Particularly
troublesome are unit startups and shutdowns and periods of load cycling operation. Chemistry
control is most difficult at these times, in part because air inleakage rates tend to be higher
compared to periods of sustained operation.

History of Investigations into Turbine Deposition and Damage

Turbines have always been affected by deposits. In the early days, the main damage concern
was often silica, which deposited in the low pressure (LP) turbines. As early as 1962, Pocock
(2)
and Stewart reported on the research undertaken into investigating the loss of capacity of the
supercritical units at the Philo Power Plant, due to the rapid development of copper deposits on
the HP turbine.
(3)
Martynova reported that in the 1960s, her co-workers in Russia were able to establish the
solubilities of the main impurities in boiler water, i.e. salts, silica, iron and copper oxides, in
superheated steam and established that the solubility of these substances decreases through the
turbine. In addition, they were able to predict the location of deposits, for example that of copper
oxides at the first stages of the HP turbine, and estimate how much deposits could be tolerated in
the first stages of the turbine. They reported that the deposits crystallize from supersaturated
steam and strongly adhere at the blades surface. These deposits are sometimes readily "washed
away" by the steam flow at very high velocities.

Lindsay, Lee and Gould reported on investigations into the behavior of impurities in steam
(4-6)
turbines. Jonas reported that 154 chemical compounds in steam and deposits had been
identified and that potentially corrosive sodium hydroxide, sodium and ammonium chlorides and
(7)
sulfates, and organic and inorganic acids were frequently present.

An overview of the key historical progress is given in Table 5-1. (8) Reference to original work of
each of the cited investigators can be found in Reference 8.

Overview of Steam Path Damage Mechanisms and Impact

The impurities in turbine steam and subsequent deposition and corrosion can occur when
impurities are transferred to the turbine by mechanical and chemical (vaporous) carryover from
the boiler. This includes the mechanical entrainment of minute droplets of water in the steam. (9)
Contributing factors can include incorrect drum level, inadequate blowdown to maintain proper
boiler water chemistry and inadequate monitoring of steam quality. These turbine deposits have
a composition similar to that of the dissolved solids in the boiler.

A further source of turbine deposits even with good steam separation can be the use of impure
water for desuperheating (attemperation). The effects of attemperator spray water on steam
purity can be assessed by comparing the impurity levels in saturated and main/reheat steam

5-2
Turbines

samples or by calculation from the feedwater composition, spray rate, saturated steam
composition and boiler steaming rate.(10) Excessive use for attemperator spray can be a
contributing factor.

Cycling, startup and shut down affect the production and transport of deposits to the turbine and
increase the risk of damage. However, such units are much more likely to be subjected to
condensate washing, which can remove some of the deposited chemicals.

When deposition is substantial, it may be necessary to de-rate the unit until the deposits can be
removed mechanically or chemically during an outage.

Jonas explained that the precipitation of chemical compounds from superheated steam;
deposition and evaporation from hot surfaces; and change of concentration in oxides in
(7)
depositions are three mechanisms for concentrating impurities in steam turbines. Chemicals
precipitate when concentration exceeds solubility in superheated steam.

Allmon, et. al. demonstrated that a deposit may form a solid, crystalline or a liquid (aqueous)
film on the internal surfaces of the turbine.(11) The solubility of the chemical in steam and the
concentration of the liquid deposit depends on the chemical nature of the impurity and the
temperature and pressure at each location on the turbine surface. In some conditions both states
can coexist for different chemicals.

McCloskey et. al. summarized the role of chemistry, mechanisms and areas of turbine affected in
(8)
Table 5-2. The aqueous deposit can attack the metallic surfaces of stressed or susceptible
materials. Experience has shown that a concentration of corrosive impurities (chloride and
sulfate ions) in turbine deposits below 0.25% usually do not cause pitting and corrosion fatigue
(12) (9)
of blades. Pitting corrosion is most often associated with chloride deposits. This
contaminant is most commonly found in systems employing brackish or salt water cooling.
Pitting corrosion is rarely the direct failure mechanism of a turbine component. However, it can
initiate a site for corrosion fatigue and has also been implicated in stress corrosion cracking
failures. Stress corrosion cracking is probably the most serious deposit related problem, because
it can lead to long outages.

Jonas, et. al. showed that the impact of deposits on turbine performance depends on their
(13)
thickness, their location and the resulting surface roughness.

Detailed monitoring of plant conditions revealed discrepancies between field data and the
historical representation (ray diagram) of the expected carryover of impurities into steam.(14)
Thus, laboratory investigations were undertaken on the following compounds and related
species: NH4Cl, NH3Cl, HCl, NH4HSO4, Na2SO4, NaHSO4, NaOH, H2SO4, organic acids (acetic
and formic), Cu(OH)2, H3PO4, NaH2PO4, and Na2HPO4. As a result, an improved method of
predicting vaporous carryover was developed by EPRI, as described later in this report.
Table 5-1
Overview of Key Historical Research into Steam Impurities

Decade Researcher(s) or Topic(s)


Agency
1930s Straub Steam turbine deposits, boiler carryover, field survey and

5-3
Turbines

Decade Researcher(s) or Topic(s)


Agency
laboratory data.
Faulk and Ulmer Steam separation.
Baker Steam separation.
Powell Steam contamination.
1940s Fuchs and Rudoff Carryover and deposition of silica by its volatility.
GE Effects of deposits on turbine capacity and efficiency.
Straub Measured boiler carryover and solubilities in superheated
steam.
1950s Morey, Coulter, Measurements and discussion of steam solubilities of
Kennedy and others silica.
Styrikovich, Martynova Experimental and theoretical work.
and others
Various researchers Extensive studies of condensation, separation, and mist
formation.
USSR Guidelines for cycle chemistry.
1960s USSR researchers Solubility in steam and steam-water distribution.
Various researchers Boiling, condensation and impurity behavior in boiler tubes.
Heitmann Silica solubility diagram for a wide range of steam and
water conditions.
Kirsch Deposits.
Pocock Study of solubility of copper in steam.
1970s ASME Research Intensified work in steam chemistry.
Committee on Water in
Thermal Power
Systems (Steam Purity
Task Group)
IAPS (now IAPWS
International Assn for
Properties of Steam
and Water)
Turbine and boiler First U.S. steam chemistry limits.
manufacturers.
Many field measurements of carryover, steam composition,
and turbine deposits.
Lindsay, Martynova, Improved theoretical understanding of steam chemistry.
and others.
EPRI Corrosion fatigue of turbine blades.
Stress corrosion of disks.
CRIEPI Guidelines for cycle chemistry.

5-4
Turbines

Decade Researcher(s) or Topic(s)


Agency
1980s EPRI, CEGB, VGB Comprehensive plant-wide guidelines for cycle chemistry.
EPRI Guidelines for monitoring, and international practice.
Effect of phosphate treatment on steam and deposits.
Steam impurity interactions with magnetite and metal
surfaces.
Concern of high carryover of chlorides and sulfates.
Research on solubilities (NaOH, NaCI, NH4C1).
Modeling volatility.
Experimental work on volatility of boric acid and amines.
Pitzer Thermodynamics of NaCI solutions in steam.
Gallagher and Sengers Modeling of steam chemistry near critical region.
First attempt to verify concentrated impurities in "salt zone"
of running turbine.
Lindsay and Lee Thermodynamic and kinetic aspects of precipitation/
condensation of low volatility impurities in turbines.
ABB/Svoboda Composition of "first condensate" and steam moisture.
First laboratory study of deposition under dynamic and
expanding steam conditions.
ASME and EPRI Water Technology Handbook published in 1989.
Development of practical multi-stage through-flow
analyses.
Computer prediction of condensate phenomena (without
effect of impurities).
Hill; Moses and Stein; Influence of impurities in condensing steam.
Kantola
1990s EPRI Volatility of impurities, salts, organics and oxides in steam
Source: Reference 8

5-5
Turbines

Table 5-2
Turbine Steam Path Mechanisms Influenced by Cycle Chemistry

Mechanism Role of Chemistry/Moisture Affected


Turbine
Sections

Localized corrosion Impurity transport from the condensate, transport & volatilization of LP
(pitting and crevice the impurity into steam and its condensation and concentration in
corrosion) liquid films on the blade and disc surfaces which during "moist"
shutdown conditions lead to localized corrosion and pitting.

Stress corrosion Impurity transport from the condensate, transport & volatilization of LP
cracking of disc the impurity into steam, and its condensation and concentration in
attachments liquid films on the blade and disc surfaces. Pitting initiated during
shutdown.

Stress corrosion Impurity transport from the condensate, transport & volatilization of LP
cracking/ corrosion the impurity into steam and its condensation and concentration in
fatigue of blades liquid films on the blade and disc surfaces. Pitting initiated during
shutdown.

Deposition on Impurity transport from the condensate, transport, and LP


LP blades volatilization of the impurity into steam and its condensation and
concentration in liquid films on the blade and disc surfaces

Copper deposition Feedwater corrosion of copper alloys, transport and volatility of HP; IP
copper and its oxides into steam.

Condensation Moisture nucleation is affected by impurity transport and the LP


(moisture nucleation) condensation process heavily affects the formation of liquid films
on blades needed for various mechanisms.

Liquid droplet erosion Moisture formation and shedding of liquid films off blades & discs. LP

Moisture-related Moisture formation and shedding of liquid films oft blades & discs. LP
(except liquid droplet
impact)

Flow-accelerated Acidification of moisture increases flow-accelerated corrosion. Exhaust


corrosion hoods

Source: Reference 8

Operational Consequences

Turbine operating problems and failures at several nuclear power plants have been attributed to
deposition of chemical impurities from the steam.(11) When a deposit forms in a turbine, it can
cause a loss in thermal efficiency, changes in pressure drop, erosion, localized corrosion (e.g.,
pitting), corrosion fatigue or stress corrosion cracking. Damage in a turbine usually results from
a combination of material susceptibility, design stresses and the environment. Unfortunately, all
materials can be susceptible to damage in various environments. The most practical way to
minimize damage within existing units is to control the environment.

5-6
Turbines

Deposits not only foul the turbine blades and cause losses in efficiency, but as Svoboda et. al.
noted, deposits may block the movement of the turbine valves, leaving the turbine unprotected
against potential overspeed if there is a turbine trip. (15)

The most serious failures are LP turbine blade and disk cracking and corrosion fatigue.
Martynova in 1993 reported that one of the most difficult areas to protect is the region of steam
condensation, specifically the region where condensate first forms (early condensate) as steam
expands through the turbine.(3) This is the region where the most serious failures (turbine blade
rupture, disc cracking etc.) have taken place. It is referred to as the phase transition zone (PTZ).
The high stresses on high strength materials combined with a corrosive media leads to
difficulties in managing intergranular stress corrosion cracking (IGSCC) and corrosion fatigue of
the last LP turbine blades.

According to a 1993 survey of the steam turbine LP blade failures, the majority of the failures in
(9)
the period 1965 to 1977, gave fairly consistent results, as follows:
Steam path deposits cause excessive corrosion,
Chlorides are the most harmful contaminant with respect to blade life,
Pitting is usually associated with blade cracking,
Blade failures most commonly occur in the wet regions of LP turbines,
Failures are most frequently attributed to a corrosion fatigue mechanism and
Low pH and high oxygen, especially in conjunction with chloride ions are very deleterious to
blade life.

McCloskey et. al. in 1999 stated that steam path damage, particularly of blades, has long been
recognized as a leading cause of steam turbine unavailability for large fossil fuel plants. The
(8)
conclusions of the EPRI analysis indicated that:
Corrosion fatigue of blades in low pressure (LP) turbines continues to be the most significant
form of steam path damage;
Corrosion fatigue is often found at the dry-to-wet transition zone (PTZ);
The majority of LP blade damage occurs in the last two rows in fossil fuel units:
Significant damage occurs by stress corrosion cracking in rotors (disc rim blade attachment
region) in fossil units:
Damage in high pressure (HP) and intermediate pressure (IP) turbines of fossil units caused
by solid particle erosion continues to be significant:
Deposition of copper in the HP and IP turbines has seen recent increases in occurrence,
although this phenomenon originally occurred in the late 1950s; and
Steam path damage is often concurrent with other problems in the plant in many cases, more
than one problem is evident in units reporting blade failures.

5-7
Turbines

Historically, steam turbines in the United States from all vendors experienced deposition and
corrosion. An estimated 40% of utility and 60% of industrial turbines in the United States
operate with high concentrations of impurities in steam.

Steam turbines have been the fifth leading cause of forced outages.(16) Most of that outage time is
attributed to corrosion of LP turbine blades. Even without corrosion problems, a buildup of
deposits on turbine blades results in costly heat rate increases. Only 0.1 mm thick deposit can
reduce stage efficiency by 3%. The estimated cost of the turbine problems to the utilities in the
United States is extremely high and maybe as much as $1 billion/year. At the same time, there
are hundreds of turbines with up to 200,000 hours of service without any corrosion problems.

Jonas and Dooley stated in 1999 that whilst improvements have been made in turbine design
(reduction of stresses and material yield strength) and in steam purity control, there are still
several major turbine problems, including: (17)
Stress corrosion of LP turbine disk blade attachment,
Pitting, corrosion fatigue, and stress corrosion of blades in the dry-to-wet transition zone and
Loss of generating capacity and efficiency due to deposition of copper, iron, sodium
phosphate, and other impurities.

Jonas et. al. stated in 2001 that even in new turbines, the efficiency can be improved by better
blade surface finish and prevention of foreign object damage.(13) In today's competitive business
environment, there are strong incentives for improving efficiency. With high replacement power
costs typically at $100/MWh and costing as much as $7,000 per MWh under tight market
conditions, the savings can be very substantial. It has been demonstrated by field experience that
up to 15% of MW generating capacity loss can be caused by buildup of blade deposits and
several percent of turbine efficiency can be lost due to a deterioration of blade surface finish and
the deposits. For a copper deposition problem with a 30 MW loss of capacity and the difference
between the sale price and cost of generation of $30 per MWh the yearly loss is over $5 million.

5.2 Source, Composition and Location of Deposits

Most deposits are salts and oxides, principally iron and copper oxides. Copper deposits on
turbines can affect the output capability of the turbine. The deposits can reduce the efficiency of
the turbine, but more seriously, they can lead to corrosion and mechanical failure of the blades or
disks. The otherwise innocuous iron oxides that deposit on the turbine can absorb salts,
increasing the risk of corrosion. A particular concern with salts is in the phase transition zone of
the LP turbine, where high concentrations and potentially more acidic or alkaline conditions can
form. The location that impurities deposit in the turbine depends on their nature, solubility in
steam and operating conditions, including temperature and pressure.

General Aspects of Deposition


(7)
Jonas listed 154 chemical compounds identified in steam and deposits using X-ray diffraction.
The list included sodium hydroxide, sodium and ammonium chloride and sulfate and organic and
inorganic acids. Also, copper oxides, which can aggravate corrosion, are frequently present.

5-8
Turbines

Major corrosive deposits, identified in a literature survey given in EPRI report NP-3002, were
sodium hydroxide (NaOH), sodium chloride (NaCl), hydrochloric acid (HCl), sodium silicate
(Na2SiO3), sodium carbonate (Na2CO2), sodium bicarbonate (NaHCO3), ammonium chloride
(NH4Cl) and potassium chloride (KCl).(11) Sodium sulfate (Na2SO4), ammonium sulfate
(NH4)2SO4 and ammonium hydrogen sulfate (NH4HSO4) are compounds that may be present in
turbine deposits. Sodium ions are common contaminants and ammonium ions are present from
water treatment chemicals. Sodium hydroxide, sodium chloride and hydrochloric acid were
considered the most undesirable because these chemicals can promote stress corrosion cracking.

Salts undergo hydrolysis to produce hydroxides and since sodium salts are most prevalent,
NaOH may be present in the water. Caustic deposits are very aggressive and have been found in
turbines, especially in the last few "dry" stages of the turbine.

Silica, a common constituent in surface waters, can enter the steam cycle via makeup water or a
condenser leak. Silica has a relatively high volatility and can be carried into the turbine in
significant quantities. Vaporous silica precipitates mainly as insoluble deposits in the
low-pressure turbine, affecting capacity and efficiency.

Svoboda, et. al. observed that deposits were present above the phase boundary whereas the blade
(15)
surfaces were clean below the phase boundary (Figure 5-1). In a few cases, some attack was
present near the phase boundary, corresponding to the sporadic formation of a corrosive phase.

In a later report, Svoboda, et. al. noted that on one turbine investigated, the deposit zone" was a
uniform oxide layer about 0.1 mm thick.(17) It consisted of iron-oxides (alpha-Fe2O3, Fe3O4, alpha
and gamma-FeOOH) and silicates (alpha SiO2 and amorphous silica), and contained ionic species
2
(mostly chlorides, sulfates and sodium in the range of 1 to100 mg/m ). These substances were
not found in the "clean zone."

Leonard et. al. reported on the deposits and resulting losses on a turbine at the W. H. Sammis
(17)
plant. The deposits contained magnetite and hematite, with traces of -Fe and silica (Table 5-
3). The concentrations of components in turbine deposits are given in Table 5-4. (18) Figures 5-2
(19,20)
and 5-3 show the location of various types of deposits and damage.

The current understanding of the physical-chemical processes occurring in low pressure turbines
are summarized in Figure 5-3. The inlet steam composition is a result of impurity transport from
the boiler and is influenced by the chemical hideout on superheater and reheater surfaces. In the
superheated steam region, as the steam expands and the solubility of steam impurities decreases,
chemical substances, the concentration of which exceeds the solubility, precipitate and some of
them deposit. There is co-deposition of solid particles and sorption on oxidized surfaces.

5-9
Turbines

(a) (b)

Figure 5-1
Turbine Deposits. (a)Blades in the zone of initial condensation in an LP turbine. Only the
blade root shows a thin deposit of iron oxides. (b) Failure due to corrosion fatigue. The
cracks occurred near the transition between oxide deposits and bare metal.

Source: Reference 15

Table 5-3
Components in Turbine Deposits from W. H. Sammis Plant

Location (under shroud) Major Medium Minor/Trace

LP-1 Turbine L-2 Fe3O4 Fe2O3 -Fe, SiO2

LP-1 Turbine L-6 Fe3O4, Fe2O3 -Fe

HP/IP Turbine L-9 Fe3O4, Fe2O3 -Fe

HP/IP Turbine L-12 Fe3O4 Fe2O3 -Fe

Turbine 1990 Losses (KW) 1996 Losses (KW)

HP/IP 805 603

LP 216 198

Source: Reference 17

5-10
Turbines

Table 5-4
Concentrations of Chemical Species in Turbine Deposits

Boiler Si Cu Pb Fe Ca Na Cl SO4 PO4 CO3 pH Thickness


Water Treatment (Wt. %) (Mils)
Drum, > AVE. 17.0 11.5 0.70 29.3 0.66 8.38 0.51 3.13 6.82 0.78 9.3 42
100.MW.
(P04) MAX 96.2 96.9 40.0 96.6 28.0 40.9 16.1 63.5 62.0 18.0 12.3 500
OT- Subcrit. AVE 31.5 2.1 2.84 27.0 0.97 6.67 5.98 0.81 0.22 0.00 9.5 38.0
(AVT) MAX. 98.9 28.5 60.6 97.2 10.0 38.0 42.7 13.9 2.03 0.01 9.8 62
OT- Supercrit. AVE. 11.4 4.68 0.29 43.7 0.50 4.86 1.22 2.05 0.38 0.25 6.8 42.4
(AVT) MAX. 94.6 75.0 25.0 98.5 30.0 67.0 24.0 43.8 7.04 3.00 10.7 300
All Units AVE. 16.3 7.20 0.65 34.7 0.70 6.80 1.20 3.20 4.00 0.70 8.45 41.4
(P04 + AVT) MAX. 98.9 96.9 60.6 98.5 30.0 67.0 42.7 63.5 62.0 57.0 12.3 500
Units with AVE. 17.6 5.90 0.15 25.6 0.40 4.50 2.20 4.30 1.70 0.10 8.0 22.2
Blade Failures MAX. 98.9 68.0 4.50 96.1 25.0 33.8 42.7 54.5 25.3 3.00 10.8 125

Source: Reference 18

Figure 5-2
Typical Locations of Turbine Corrosion, Erosion and Deposits. (In figure 5-2 the symbols
have the following signification: P, pitting; CF, corrosion fatigue; SCC, stress-corrosion
cracking; C, crevice corrosion; G, galvanic corrosion; E, erosion; E-C erosion-corrosion;
SPE, solid particle erosion; chemicals-deposits.)

Source: Reference 16

5-11
Turbines

Figure 5-3
Cross-Section of LP Turbine Showing Locations of Chemistry and Corrosion Processes

Source: Reference 20

Martynova linked the solubility of iron, copper, silica and sodium chloride to deposition on HP
(3)
and IP turbines (Figure 5-4). Figures 5-5 and 5-6 show that different materials tend to deposit
on different sections of the turbine.(12,21,22) This is attributed to a decline in solubility as the
specific volume of steam increases. Figure 5-7 shows the chemical transport processes for drum
(22)
boilers. Most of those processes are also applicable to cycles with once-through boilers.

Layup, startup, and cycling can exacerbate corrosion and transport of impurities. In particular,
air inleakage, corrosion and exfoliation products, condenser leaks, aerated makeup water,
condensate polishers and, sometimes, the combustion products entering leaking reheater tubes
during initial firing, when the reheater is under vacuum (Figure 5-8).(23)

5-12
Turbines

(a) HP Turbine

(b) LP Turbine
Figure 5-4
Solubility and Deposition Turbines

Source: Reference 3

5-13
Turbines

IP
HP Turbine LP Turbine
Turbine
100
90 Fe

80 Cu
Turbine Deposits

70

60

50

40

30
Si
20

10 SO4
Na
0
0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30
3
Specific Volume, ft /lb
Figure 5-5
Relative Distribution of Components in Typical Turbine Deposits
Source: Reference 12

Figure 5-6
Summary of Deposit Compositions as a Function of Specific Volume of Steam
Source: Reference 22

5-14
Turbines

Figure 5-7
Chemical Transfer in the Conventional Drum Boiler System
Source: Reference 22

Figure 5-8
Chemistry and Corrosion Effects of Layup, Startup and Cycling for a Drum Boiler Cycle
Source: Reference 23

5-15
Turbines

Salt Deposits

Mechanical carry-over in drum boilers increases with pressure, since the density difference
between water and steam decreases with increasing pressure and temperature.(24) Vaporous
carry-over occurs in drum boilers due to the inherent volatility of the compounds present in the
boiler and is also a function of pressure and temperature.

The influence of water and steam chemistry on condensation is illustrated in Figure 5-9.(16) This
is a Mollier diagram showing an LP turbine steam expansion line, which includes key chemical
processes influencing condensation. From the superheated to the wet steam, Figure 5-9
illustrates the following points:
Salts, such as NaCl, are present as single molecules or dry clusters and crystals until they
reach the top of the salt zone at slight superheat temperatures. There they start retaining
water molecules, become droplets of concentrated solutions, and as the steam expands,
become more and more diluted by water.
Within the blade path flow, there are several regions of superheat and salt zones. Some
regions, below the salt region on the graph, are associated with the condensation shock and
the transonic flow shock. The salt zone is a result of boiling point elevation of the
contaminant and could be either solid or liquid. In these zones, both homogeneously and
heterogeneously formed droplets partially or totally evaporate.
Particles of metal oxides, such as magnetite, will adsorb some moisture, but the larger
particles may stay dry throughout the steam expansion process because of their heat capacity
and temperature above the saturation temperature.

Jonas and Dooley noted that in the dry-wet phase transition zone (PTZ), moisture droplets first
nucleate in the salt zone by a heterogeneous process on impurity molecules and dusters, before
(16)
the theoretical saturation line. These droplets, which start as a concentrated aqueous solution,
are diluted by addition of water molecules as the steam expands. Liquid films of these
concentrated salt solutions are formed on blade surfaces. Nevertheless, early condensate
contains increased concentrations (up to 150 to 200 times) of solutes. When the droplets hit a
hot surface and evaporate, low volatility impurities such as NaCl and NaOH can concentrate to
% levels (salt zone) leaving liquid films which can be very corrosive to all turbine materials.

5-16
Turbines

Figure 5-9
Mollier Diagram with LP Turbine Steam Expansion Line

Source: Reference 16

Jonas et. al. collected exfoliated oxides and oxides on filters installed on the superheated steam
and analyzed for water leachable and total sodium, chloride, sulfate, phosphate and silica.(25)
Results of analysis of magnetite from the main steam pipe of a supercritical unit are given in
Table 5-5. Ammoniated water leached more impurities then distilled water, but there were still
some impurities retained on the magnetite.

There is significant carry-over of boiler water chemicals and impurities into the turbine, as
(26)
shown in Table 5-6. High concentrations of chloride, sulfate, and silica in some turbines are
of particular concern.

Jensen and Bursik found that sodium sulfate deposits were predominant in reheaters and HP
turbines, and sodium hydroxide and sodium chloride deposits were predominant in IP and LP
(27)
turbines (Table 5-7).

A computer model was developed to incorporate the partitioning constants and the hydrolysis
reactions at high temperature to predict pH and conductivity of the early condensate from the
boiler water composition. (14) These initial qualitative simulations have shown that the
concentrations of impurities in steam can reach ppb levels and those in the early condensate, ppm
levels. Chloride is much more volatile than sulfate, as seen in early condensate measurements.
Chloride is transported as hydrochloric acid, and sulfate as sulfuric acid, although ammonium
and sodium bisulfate can also play a role.

5-17
Turbines

As noted previously, discrepancies were found between data obtained from operating plants and
the values predicted from the ray diagram. (28) In particular, the plant measurements of
distribution coefficients for chloride and sulfate were higher by orders of magnitude than those
given by the ray diagram for NaCl and Na2SO4.
Table 5-5
Concentration of Impurities in Exfoliated Magnetite Filtered from Steam

Na Cl SiO2

1 Leached by water, ug/g* 32 18 24

2 Leached again by 10 ppm NH4 solution, ug/g* 28 12 12

3 Concentration in magnetite after leaching by solutions 38 --- 27


1 and 2 (determined by digestion in acid), ug/g*

* The concentration on oxides depends on the size of magnetite particles or the surface area, up to 6
ppb of sodium, 0.6 ppb of sulfate, and 0.3 ppb of chloride can adsorb on 30 ppm of magnetite in
superheated steam; 3.4 g/cm2 of sodium, 0.4 g/cm2 of sulfate, and 0.1 g/cm2 of chloride was found
on oxidized surfaces.

Source: Reference: 25

Table 5-6
Comparison of Turbine Deposit Composition (wt %)

Na Cl PO4 SO4 Fe2O3 SiO2

RP1886 A* 7.5 1.3 1.9 1.1 28.9 40.3

RP1886 SD* 10.3 2.9 4.3 2.4 25.2 23.7

RP1886 M* 33.0 10.9 17.0 10.0 92.4 79.4

RP912 A* 8.4 0.5 6.9 3.1 29.3 17.0

RP912 M* 40.9 16.1 62.0 63.5 96.6 96.2

A - Average, SD - Standard Deviation, M Maximum

Source: Reference 26

5-18
Turbines

Table 5-7
Relative Amounts of Water-soluble Turbine Deposits

Turbine High Pressure Intermediate Pressure Low Pressure


Pressure Reduction 220-45/3190-653 45-5/653-73 5-0/73-0
(bar/psi)
Na2SO4 XXXXXXX XX X
Na2HPO4 XXXX XXXX XX
NaHCO3 XX XXXX XXXX
NaOH XXXXXX XXXX
NaCl XX XXXXXX
Source: Reference 27

Silica is very volatile in steam at high pressures, but only slightly soluble under the conditions in
the low pressure turbine. Thus, silica tends to deposit in the low pressure turbine and can affect
turbine efficiency.

Copper Deposits

Hundreds of high pressure drum boilers are experiencing output (MW) and efficiency losses due
to deposition of copper in the high pressure turbines. (22) Experience with seven high pressure
drum boiler units is summarized Table 5-8. These very significant losses range from 18 MW to
60 MW or 4% to 8% of the nominal unit power.

The problem of copper deposition was first recognized in the late 1950s with the introduction of
high pressures in supercritical units in the United States. One of the first experiences of the
deposition of copper oxides in the HP turbine was at Ohio Power Companys Philo 6
supercritical unit. During the first year of operation, serious deposition occurred in the HP
turbine. Within a short time, the unit was shut down. Heavy deposits of copper oxides were
found on stage wheels and diaphragms, as well as lighter deposits on the first and second reheat
turbines. This incident was followed shortly by a similar one in Cleveland Electrics Avon
Unit 8.

5-19
Turbines

Table 5-8
List of Utilities/Units Losing Capacity from Copper Deposition

Unit Max Boiler Type Copper in Boiler Condensate Operation MW


MW Pressure* Treatment Polisher Loss

Carolina P&L 690 Forced LPFWHs AVT No Load Follow 60


Roxboro 2 Circulation
2850 psi

Tri-State Craig 2 447 Natural Condenser AVT Powdex Coordinated 18


Circulation LPFWHs Load Follow
2820 psi HP-Monel

Wisc. P&L 549 Forced Condenser EPT No Load Follow 29


Columbia 1, 2 Circulation LPFWHs
2800 psi HPFWHs

Montana Power 776 Forced Condenser PO4 Startup Base Load 30


Colstrip 3, 4 Circulation All FWHs Powdex
2800 psi

CPS 405 Forced Condenser PO4 No Base Load 30


Circulation All FWHs
2800 psi

* 1 kPa = 0.145 psi

Source: Reference 22

Sources of copper are cited in Table 5-9. (12) A summary of the transport processes is given in
Table 5-10. Changes in feedwater heater tube material accepted for once-through boilers have
not been applied as extensively to drum boilers, although recognition of the benefits of avoiding
copper alloys has improved over the past 20 years. Replacement of copper alloy feedwater
heater tube bundles is not normally initiated until justified by the costs associated with reductions
in efficiency or reliability due to poor heater performance. A variety of copper-bearing materials
is employed in tubing used in feedwater heaters, condensers and other heat exchangers (such as
(1)
gland steam condensers) in the pre-boiler circuits of fossil plant drum boilers .

5-20
Turbines

Table 5-9
Possible Sources of Copper and its Oxides in Steam

Source Mechanism Possible Causes

Boiler Water (Drum Mechanical Carryover High drum level


Boilers)
Drum liner leakage (controlled circulation boilers only)

Damaged or faulty steam separation equipment

Vaporous Carryover High copper in boiler water. Operation above design


pressure

Spray Direct Injection Copper in feedwater

Attemperation Extraction point for attemperation water too close to


Water economizer inlet header (controlled circulation boilers
only)

Feedwater Direct Injection Copper in feedwater

(Once-Through
Boilers)

Source: Reference 12

Table 5-10
Summary of Steam Impurity Generation and Transport

Steam Impurity Generation Main Transport When


Mechanism

Copper and its oxides Corrosion of feedwater heaters Boiler carry-over, Startups
during layup and operation attemperation

Iron oxides Exfoliation in SH, RH, erosion- With steam, boiler Startups
corrosion in FW carry-over

PO4, Na Used for boiler water treatment Boiler carry-over Transients

Cl, SiO2,SO4 Condenser leak Boiler carry-over, Transients


attemperation

Cupric oxide, (tenorite, CuO) and cuprous oxide (cuprite, Cu2O) are predominant under HP
turbine conditions. Copper in the steam is likely to deposit at any location subject to pressure
losses. Specific volume increases resulting from the steam passing through the high pressure
turbine makes this area a likely deposition site. Pocock and Stewart found that the supercritical
(2)
turbine at Philo contained mixtures of cupric and cuprous oxides. Figure 5-10 shows the
(1)
typical location of copper deposits in high pressure turbines.

5-21
Turbines

(29)
Locations subject to deposition were documented by Shields, et. al. Transport depends on unit
operating practices. Locations subject to copper deposition during normal operation may serve
as copper release sites during startup. Copper deposits can form at several locations including
the high pressure section of the turbine. Solubility plays an important role in deposition, with
much of the copper present under reducing conditions in suspended rather than dissolved form.
Volatile carryover is, in most cases, the primary transport path. The volatility of cupric oxide
can be as high as 30% in high pressure drum boiler units, and is significantly greater than that of
cuprous oxide. This relationship clearly illustrates the benefits of reducing chemistry with
(1)
respect to volatile carryover activity (Figure 5-12).

Axley et. al. described significant efficiency losses in the high pressure turbine of a 900 MW
supercritical unit primarily as a result of heavy deposit buildup on the turbine blades.(30) The
copper concentrations were normally greater than the EPRI recommended limit of 2 ppb, usually
ranging from 3 to 15 ppb. A thorough investigation of metals transport confirmed that the
primary source of copper to the HP turbine was the Monel 400 tubes in the HP feedwater heaters.
The highest concentrations were contributed by the shell side of the heaters.

Figure 5-10
Turbine Copper Deposits

Source: Reference 1

5-22
Turbines

Copper oxide
deposition
HP turbine
Chemical
feed
Copper IP LP
volatility turbine turbine
and
solubility Air and
CO2
Condenser ingress
Attemperation Makeup
Copper Copper oxide
solubility transport

Copper
deposition
Boiler Deaerator
Copper
oxide
transport
HP heaters LP heaters

Condensate
polisher

Corrosion of Corrosion of Feed


HP Alloys LP Alloys NH4OH
Imuprity ingress N2H4
Corrosion
Deposition

Figure 5-11
Fossil Drum Cycle Illustrating Key Copper Processes

Source: Reference 1

Tenorite (CuO) has been identified in the turbine deposits but not in the superheater deposits,
which may reflect the higher solubility in steam. However, reduced forms of copper (metallic
copper and cuprite) are also believed to be less volatile than tenorite. Reduced forms of copper,
by virtue of their low volatility and solubility relative to tenorite, may enter the superheater by
mechanical carryover. Reduction of tenorite to cuprite within the superheater could also account
for the deposit characteristics.
Palmer et. al. identified the chemistry that occurs when copper and its oxides are exposed to
(31)
conditions in fossil-fired boiler from start-up to normal operation. A combination of solubility
and volatility experiments were used to determine the concentrations of copper in steam under
varying conditions of temperature and pH, where the oxidation state was controlled either by
trace oxygen levels or by the copper metal/cupric oxide oxidation/reduction couple.

For over 40 years copper has been a major problem in fossil plants including:

Corrosion of feedwater heater and condenser tubing,


Deposition of copper on boiler waterwalls and
Deposition on the first few blade stages in the HP turbine.

5-23
Turbines

Of particular concern was the frequency of turbine deposition problems, which increased
markedly towards the end of the1990s (32) The EPRI Program Copper" included the following
elements to overcome these deficiencies:
The corrosion of copper materials in LP and HP feedwater heater environments
The high temperature electrochemistry,
The volatility, partitioning and solubility of copper and its oxides in water and steam and
The deposition process in the turbine.

Recent research has shown that oxidation-reduction potential (ORP) must be kept reducing
during all periods of operation and shutdown.

Deposition Processes and Mechanisms

Correlations have been found between chemical environment, the composition of the steam and
the formation of deposits, and turbine blade and disc cracking in fossil drum units. (33) Harmful
concentrations of potentially corrosive impurities were found in deposits of about 50% of the
turbines. The pitting and corrosion fatigue susceptibility of low pressure turbine blades
correlated well with the corrosiveness of the turbine deposits (concentration of chloride). When
the concentrations of chloride in the deposits was low (<0.1% Cl), the turbines did not
experience blade failures. Little sodium was found in deposits in turbines when the sodium in
steam was below 2 ppb, but higher concentrations occurred in deposits, when the sodium in
steam was above 10 ppb.

In the salt zone, salts and acids can be present as solids and/or concentrated aqueous solutions.
The relatively benign concentrations of a few parts per billion (ppb) in the steam can be
concentrated into deleterious concentrations on metal surfaces by several mechanisms, such as
evaporation, drying, concentrating in oxides and concentrating in liquid as a result of a
temperature gradient. Heat transfer down the rotor can elevate local surface temperatures above
that of the surrounding steam/liquid.

The deposition from steam, as a dynamic process, is a balance between three simultaneous
processes:
Precipitation from expanding steam,
Deposition on a surface and
Re-dissolution of the deposit.

Deposition is expected when the concentration of an impurity in the steam exceeds the local
solubility at any location. Conversely, when the concentration of an impurity in the steam
decreases below the solubility, a deposited salt can start dissolving and be carried away with the
steam. A deposit can be washed away by moisture during load changes, shutdown and start-up.

A good correlation has been observed between the solubility and turbine deposition patterns for
molecular impurities. Figure 5-12 shows the solubilities of several impurities plotted for

5-24
Turbines

conditions through a fossil turbine. (16) The highest rate of deposition is observed where the
gradient of solubility is high, such as at the high pressure turbine inlet for copper oxides and at
the exit from the IP turbine for sodium chloride, and in the LP turbine for silica.
2 3
10 10
Temper
a ture
Fe(FeO
3 4)

Temperature & Pressure,atm


1 2
10 10
Cu
(Cu
O)
Copper & I ron, ppb

-0 1
10 Pres 10
sure

10-1 10-0

Reheater
-2 -1
10 10

-3 -2
10 10
0 2 4 6 8 10 12 14 16 18 20 22 24

HP Turbine IP Turbine LP
Turbine

Figure 5-12
Solubilities of Iron and Copper Passing through a Turbine in Supercritical Cycle

Source: Reference 16

Gabrielli and Goodstine noted that the carryover of solids into the steam occurs by three
mechanisms. (34) The total carry-over is the sum of the mechanical and the vaporous carry-over.
The theoretical values of mechanical carryover used by some boiler designers and utilities for
various boiler pressures was shown earlier as Figure 3-5. The vaporous carry-over (distribution
coefficient) for a number of chemical species is shown in Figure 5-13 as a ray diagram. The
total carryover of low volatility species, such as sodium phosphate, sodium chloride and sodium
hydroxide is similar, because mechanical carryover is almost independent of the chemistry. The
total carry-over is reasonably well defined by the ratio of sodium in the drum steam to the
sodium in boiler water. The contribution of attemperation sprays can be evaluated by a
comparing the drum (saturated) steam composition with the main steam, LP steam, and
condensate compositions. The volume of attemperating water varies widely from unit to unit
and with load, and varies from zero to over 15% of the feedwater flow.

The rate of mechanical carryover is influenced by the design and integrity of boiler steam drum
(10)
internals, specifically the moisture separator devices and baffles. Mechanical carryover
increases with boiler pressure; the typical values used for deriving the EPRI limits for properly
designed boilers with the internals in good condition are indicated in Figure 5-14. A factor of 2

5-25
Turbines

is included in the figure to provide a conservative estimate of mechanical carryover. Actual


mechanical carryover rates should be measured.

The ray diagram, shown in Figure 5-15, provides an estimate of vaporous carryover for several
chemical species which may be present in the boiler water of drum units. Inspection of the
diagram indicates that predicted volatile carryover of cuprous oxide is very close to that of silica.
Cupric oxide exhibits a greater tendency for volatilization than either silica or cuprous oxide and
at 2600 to 2800 psi (about 17.9-19.3 MPa) the volatile carryover of cupric oxide is between 30 to
40% of the concentration in the boiler drum.

Drum Pressure (psia)


3200

3000
2800
2600
2200
2000

1600

1200
1000

600

400

200
10-0

10-1 Fe3O4

Al2O3
Concentration In Vapor/Concentration in Liquid

10-2 NH B2O3
4 Cl
, (N NiO
H
4)
2 SO
4 CuO
10-3

SiO2
10-4
Cu2O
CaSO4
MgO BaO
CaCl2
10-5
Na

LiCl
Cl
Na 3

10-6
Na

PO
2S
O4

NaOH

10-7
22 20 16 12 8 6 4 3 2 1
Drum Pressure (MPa)

1 2 6 10 20 60 100
Density of Liquid/Density of Steam

Figure 5-13
Vaporous Carryover for Various Chemicals

Source: Reference 51

5-26
Turbines

As shown in Figure 5-14, the densities of steam and water in equilibrium approach the same
limiting value as temperature (saturation pressure) is increased to the critical point and the
physico-chemical properties of steam and water converge. The two phases have increasingly
similar solvating characteristics as saturation pressure increases. In addition, some solid
materials may exhibit significant vapor pressures at elevated temperatures. The net result is that
solids are no longer completely confined to the liquid phase. Instead, the materials exhibit a
distribution coefficient between the steam and water phases. The distribution coefficient is
simply the ratio of the concentration of a material in the steam phase divided by the
concentration of the material in the water phase. The equilibrium distribution of material
between the water and steam phases, which results in the presence of solids in the steam, is
termed vaporous carryover.

Silica was the first material identified as exhibiting significant vaporous carryover and it has long
been recognized by boiler operators that the steam silica concentration is predominantly
governed by vaporous carryover and not by mechanical carryover. Many other solid materials,
including treatment chemicals, are subject to significant vaporous carryover at high pressure.

The ray diagram presented in Figure 5-13, as originally proposed by Styrikovich in the late
1950's, illustrates the distribution coefficient for many compounds which might be encountered
in boiler water. This diagram is based on the relationship that the distribution coefficient is equal
to the density ratio of steam to water raised to a power, n. This exponent is the slope of the line
for that material on the ray diagram. However, the actual extent of vaporous carryover of a
particular solid species is a complex function of the properties of the chemicals present in
(14)
combination with the properties of steam and water (Figure 5-14).

The cycle chemistry guidelines published by the American Boiler Manufacturers Association
(ABMA 3rd ed, 1982) included a calculation of the equivalent total carryover represented by the
ratio of steam to boiler water solids. (These excluded silica, because of the high degree of
vaporous carryover of this material.) High pressure boilers operating at over 1800 psig were
assumed to have a mechanical carryover of no more than 0.2%. Vaporous carryover of other
solid materials was considered significant only at the highest pressure range listed; an additional
allowance of 0.13% vaporous carryover was provided in this situation.

Ball and Jenkins reported that the carry-over from 500, 350 and 200 MW drum boilers ranged
(33)
from 0.02 to 0.7%. Carry-over increased when factors, such as load changes, affected the
drum water level or stability. The chemical purity of the steam was normally within the limit of
5 ppb sodium. However, a certain amount of carryover appeared to be inevitable and, when the
concentrations of sodium, chloride and sulfate exceeded 1 ppb, there were measurable losses
across the superheaters, high pressure turbines and reheaters, indicating deposition. Some of the
salts may become re-entrained, but sodium sulfate tends to be locked in the reheaters. Some
light deposits were found on the turbines but no damage.

5-27
Turbines

Pressure (psig)
725 1450 2175 2900 3277
1.0 1.0

0.8 0.8
Wa
ter den
sity
(d )
Density of Steam and Water (kgl-1)

Vapor/Liquid Density Ratio (dV /dL)


L

Critical Pressure
0.6 0.6

Differential Density
0.4 0.4

0.2 0.2

eam
Density ratio
a t e d st )
ur (d V
Sat ensity
d
0 0
0 5 10 15 20 22.6
Pressure (MPa)
Figure 5-14
Water and Steam Density-Pressure Relationship at Saturation

Source: Reference 14

Warwood et. al. reported that deposition processes are dependent on several transport and
reaction mechanisms (Figure 5-15). (35, 36) The mechanisms can be described as: 1) generation, 2)
convective bulk transport in the system, 3) transport from the bulk fluid to a surface, 4)
interaction with the surface and substrate through adsorption, 5) surface relocation to a reaction
site (chemisorption or crystallization processes), 6) multilayer sorption onto underlying deposit
material (amorphous deposition or crystallization, aging processes, etc.). In addition, processes
of desorption of 7) molecular or 8) particulate deposit material or reactant products also occurs.

Each of the processes can lead to deposition of a solid or liquid phase onto an underlying base
material, usually iron oxide:
Transport processes include traditional mechanisms that are rate limited by transport of
material into the system or from the bulk solution to the surfaces (diffusion and convection).
Interfacial attachment phenomena are mechanisms that are surface driven such as adsorption,
precipitation, and crystallization processes, i.e. surface attachment processes, and aspects
which include surface component rearrangement and relocation during crystal formation.
Surface reactions are mechanisms that include chemisorption, surface acidic properties,
surface catalysis, and dissolution phenomena, i.e. reactions affecting solubility at the surface.

5-28
Turbines

Colloidal properties are mechanisms involving colloidal materials which overlap in many
ways with each of the above categories, but can provide unique properties (for example, a
mobile interface).
Magnetic and electrostatic interactions can occur between solid materials and stationary
surfaces.

Figure 5-15
Basic Steps in Deposit Accumulation on a Surface from Bulk Fluid

Source: Reference 35

Adsorption of ionic, molecular or particulate materials to the process surfaces can provide the
initiation mechanism for deposit formation. The adsorption processes that occur in the steam
phase can be driven by concentration and solubility gradients, electrochemical interactions,
electrostatic or magnetic interactions, or a combination of these mechanisms.
Colloidal particles provide mobile surfaces for adsorption of highly insoluble materials.
Adsorption of ionic and molecular species onto colloidal particles in the steam phase will
increase the transport rate of materials that otherwise may be present as volatile compounds
only very sparingly.
The chemical transformation of an adsorbed ionic or molecular species to surface
immobilized material occurs via surface precipitation reactions. Further surface
rearrangement into organized structures can occur during crystallization. Deposition on a
surface introduces a concentration gradient within the overlying layer as surface reactivity
consumes the suspended material and provides a transport driving force.

5-29
Turbines

Precipitation is believed to be the most prevalent driving force for deposition within
superheated steam environments. When the concentration of a steam impurity exceeds its
solubility under the existing pressure and temperature conditions, the impurity starts forming
molecular clusters and precipitates. This event occurs either on a surface or as part of a
colloidal formation reaction.
Crystal formation occurs from a saturated solution or from aging of an amorphous solid.
When formation is from a saturated solution, the equilibrium kinetics drive the precipitation
of the saturated component. Adsorption of the component occurs on the surface of a growing
crystal, with subsequent rearrangement on the surface to the final location of the ion in the
lattice. The adsorbed ions become part of the stabile crystal lattice.
Crystal formation also occurs through rearrangement of amorphous precipitates through
aging processes. The aging process is actually the rearrangement of the ions within the
amorphous solid into crystalline structures. The growth of crystals within an amorphous
precipitate due to the aging process is much slower than crystal growth at a surface due to
the intervening atoms within the solid interacting with the rolling ions as they relocate.

Volatility of Materials in Steam

An important factor in the volatility of materials in steam is the distribution ratio, i.e. the
concentration of a salt in the vapor phase versus the concentration of the salt in the liquid phase.
This indicator of vaporous carryover was developed into the ray diagram (Figure 5-15), which
shows curves for a range of salts (impurities and conditioning chemicals) that may be present in
boiler water. The ray diagram is empirical and has a number of shortcomings, as will be
explained later. Improved data on volatility has been obtained in EPRI sponsored work
undertaken by Oak Ridge National Laboratory (ORNL) and a thermodynamically rigorous
treatment of volatility has been developed. This covers the volatility of materials in steam and
vaporous carryover and has been extended to cover condensation in the phase transition zone of
the LP turbine. Since these two aspects are very closely linked, this aspect of condensation will
be considered in this sub-section with volatility, rather than the final sub-section on the solubility
in steam and deposition.
(6)
Gould reported that the solubility of impurities in steam was recognized and studied
extensively in the 1950's, following the observation of silica carryover from high pressure
boilers. Straub and Coulter worked on measuring the high pressure, high temperature
equilibrium distribution of salts between the liquid and vapor phases of water. The expression for
the distribution coefficient, K, given by Styrikovich is:
n
K = (CV/CL) = (V/L) Equation 5-1

Where C is the impurity concentration of the liquid or vapor phase, is the water or steam
density and n is the coordination number of hydration.

The experimental work was done by Styrikovich, and Martynova on many chemical compounds
at high pressures (~3,000 psi or nearly 20.7 MPa), and the distribution coefficients showed a
strong dependence on pressure (steam density), becoming smaller as the pressure decreased. This
is shown on the ray diagram in Figure 5-15.

5-30
Turbines

(11)
The theoretical considerations were considered in some detail in an EPRI report . The
distribution ratio (KD), defined as the concentration of a contaminant in the vapor phase (CV)
divided by the concentration of the contaminant in the liquid phase (CL) is function of
temperature and pressure and can be used to determine the amount of vaporous carryover from a
boiler and to predict whether deposition will occur in a steam turbine. According to this theory,
the contaminant should remain soluble in the steam and not deposit, if the vapor phase
concentration is less than KD x CL.

The transportation of ammonium salts, such as ammonium chloride (NH4Cl), may be an


important factor in steam turbine chemistry. The solubility of ammonium chloride is several
orders of magnitude higher than the solubility of sodium hydroxide or sodium chloride in steam.
Therefore, the presence of ammonium ion (NH4+) from the feedwater may lead to the transport of
additional chloride ions (Cl-) from the steam supply to the turbine. However, the increased
solubility of ammonium chloride may also prevent the deposition of chloride ion on the internal
surfaces of the turbine. Based on the above, the steam phase in the low-pressure turbine may
contain ammonium chloride, ammonium hydroxide, and hydrogen chloride. Also, associated to
NH3, NH4Cl and HCl, compounds such as (NH4)2SO4, (NH4)HSO4, (NH4)2CO3 and (NH4)HCO3
might be expected to exist in the water/steam cycle.

Vaporous contaminants deposit selectively in the turbine because of differences in the


thermodynamic properties. Each impurity has its own particular solubility in the liquid and vapor
phases of water. The solubility determines the amount of impurity leaving the boiler with the
steam and the region in the turbine where the impurity precipitates.

As steam expands through the turbine, the solubility of an impurity reaches saturation in the
steam. Deposits form in those sections of the turbine where the concentration of a specific
impurity exceeds the saturation point, resulting in precipitation of some of the impurity. If the
impurity has a small distribution coefficient, the impurity concentration in the precipitate will be
much greater than its concentration in the steam.
(37)
Simonson and Palmer described the experimental study of the volatility of ammonium
chloride from aqueous solutions to high temperatures undertaken at ORNL for EPRI. The fossil
plant cycle chemistry monitoring survey had shown that at four US power plants operating on
all-volatile treatment (AVT), the chloride levels observed in the high-temperature steam were
significantly higher than predicted on from the ray diagram. As part of that project, the
partitioning of chloride from aqueous solutions under AVT conditions and the liquid-vapor
partitioning ratios at lower temperatures found in LP turbines near the Wilson line was
determined.

The low concentrations of solutes present under AVT, made it impractical to perform laboratory
experiments at the concentrations found in steam generators. Application of the results required a
detailed knowledge of the chemistry at the temperatures and pressures encountered in steam
generator. The species responsible for transport of solutes to the vapor phase needed to be
identified and their concentrations in the aqueous phase known. This information could only be
obtained through the study of high temperature chemistry and thermodynamics, combined with
measurements of the partitioning of solutes, under conditions where the compositions of both
liquid and vapor phases could be accurately determined.

5-31
Turbines

During measurements of the volatility of ammonium chloride, ammonia and hydrochloric acid
were present from the hydrolysis of ammonium ion. Ammonium chloride volatility was
indicated by the excess of ammonia or chloride in the vapor phase, over that due purely to the
volatility of ammonia or hydrochloric acid.

The solute transports to the vapor phase as combinations of NH4Cl, HCl and NH3, with the
relative proportions of each strongly dependent on temperature, total chloride, solution pH, and
the presence of other cations (e.g. Na+).

At high temperatures and pH levels commonly encountered in steam generator systems, under
AVT, the transport of chloride to the vapor phase is predominantly as HCl, with additional
chloride transported as NH4Cl.

Although it has become customary to compare field and laboratory carryover and volatility data
on the ray diagram (Figure 5-15), Jonas found that the measured distribution coefficients for
chloride and sulfate were orders of magnitude higher than those for NaCl and Na2SO4, in the ray
diagram. This indicated that ammonia may contribute to the volatility of Cl- and SO42-. The
volatilities of Na+ and silica were similar to those expected from the ray diagram.

EPRI report TR-102377 (38) gave a more detailed description of the behavior of ammonium salts
in steam cycles and concluded that the ray diagram for NH4Cl and other species was unsuitable
for predicting the levels of solutes expected in steam, or for assigning the compound primarily
responsible for liquid-vapor partitioning.

The equilibrium constants of thermodynamic partitioning analyses represent ratios of the


thermodynamic activities of products and reactants for specific chemical reactions, and are
independent of the concentrations of any of the species. The vapor-phase molality of the species
of interest may be calculated at any conditions of interest within the range of the data, regardless
of the total molalities and identities of the species in the liquid phase.

The early condensate is likely to have enhanced concentrations of ammonium and chloride ions,
and due to the partitioning of HCl, the condensate is likely to be acidic.
(39)
Palmer et al. described the measurement of partitioning constants, with particular emphasis on
NaOH, H2SO4 and the intermediate salts NaHSO4 and Na2SO4. Much lower partitioning constants
were found for Na2SO4, than for NaHSO4.

A computerized model was developed incorporating the partitioning constants for HCI, NH3,
NH4Cl, H2SO4, NaHSO4, Na2SO4, NaOH, and NaCl; as well as the hydrolysis constants in the
- +
liquid phase for HSO4 and NH4 ; and the dissociation constant, Kw for water, with the goal of
predicting the composition of steam and condensate formed in the water/steam cycle of power
plants with drum boilers. These calculations were based on equilibrium conditions, but
demonstrated the impact of the degree of condensate, particularly regarding the pH and the
build-up of high concentrations of solutes in the condensate.

5-32
Turbines

o o
The computer code was based on chosen boiler water compositions and pH at 25 C (77 F). In
order to determine the speciation through the water/steam cycle with a drum boiler, the following
sequence of five stepwise calculations was carried out:
1. Boiler water composition at 25oC (77oF). This included the concentrations and pH or
ammonia could be adjusted to achieve charge balance. The actual speciated concentrations
were then computed from the hydrolysis equilibrium constants for HSO4- and NH4+, and Kw of
water.
2. Boiler water composition at operating temperature. The compositions were recalculated as in
Step 1 at the prevailing temperature.
3. Steam composition at operating temperature. The compositions from step 2 were combined
with the partitioning constants to yield the equilibrium concentrations of the corresponding
neutrally-charged, species in the steam.
4. Condensate composition at same temperature. This procedure involved consideration of mass
and charge balance constraints, steam- and liquid-phase composition, partitioning and
hydrolysis constants, and the fraction of steam condensed. The pH and liquid phase
concentrations were then calculated.
5. Condensate composition at 25oC (77 oF). This calculation was similar to Step 2.

In order to simulate the formation of an early condensate in the LP, a factor was introduced into
the code to take into account the fraction of the steam passing through the turbine which does not
condense. The concentrations of solutes in a micro-volume of early condensate are extreme as
can be seen from Table 5-11. Note that in practice a higher degree of condensation may occur,
e.g. 0.1%, diluting high the concentrations.
(40)
The EPRI report TR-106017 reviewed the ray diagram with particular emphasis on developing
a more quantitative method for predicting the composition of coexisting liquid and steam under
plant operating conditions. Unfortunately, the intuitive format of the ray diagram does not lend
itself to quantitative representation of precise experimental data over wide ranges of solute
concentrations, and is particularly poorly suited to multicomponent solutions. Styrikovich et. al.
noted the dependence of the apparent distribution ratio Kapp = cs/cw on the concentration of NaCl
in either phase. Their analysis showed a change of as much as two orders of magnitude in Kapp
with changing liquid-phase NaCl concentration cw, depending on the experimental pressure. This
is clearly at variance with the assignment of a single value for the distribution ratio at each
temperature (or for each value of the density ratio w/ s) as is characteristic of the ray diagram. In
principle it is possible to include mixed electrolytes within the framework of the ray diagram by
considering the partitioning of all species present in solution.

The ray diagram incorporates a linear dependence of solute concentration in steam on


liquid-phase concentration and will give incorrect extrapolations to low solute concentrations,
unless measurements have been made at extreme dilution. If such measurements are made,
adequate control or knowledge of the liquid-phase pH is essential as products of hydrolysis
reactions (e.g., HCl and NaOH in NaCI solutions) may contribute significantly to the observed
partitioning of solutes. Measurements at higher concentrations may appear to follow the ray
diagram due to the effect of activity coefficients in the liquid phase, but overall, the ray diagram
describes incompletely the thermodynamics of liquid-vapor partitioning of impurities and the
partitioning ratio represented by the ray diagram and is not independent of solute concentrations.
5-33
Turbines

Table 5-11
Boiling at 347oC (657 oF), Early Condensate Infinitesimal Liquid Droplet* at 100oC (212oF)

Species 25oC/77oF 3470C/657oF 3470C/657oF 100oC(212oF) 1000C/212oF


liquid liquid vapor vapor liquid

Na+ 75 ppb 75 ppb 0.11 ppt, (NaCl) 50000 ppm

NH4+ 206 ppb 4 ppb 0.76 ppt (NH4Cl) 0.97 ppt 1100 ppm
(NH4Cl)

NH3 120 ppb 311 ppb 740 ppb 740 ppb 53 ppb

Cl- 72 ppb 72 ppb 29 ppt 29 ppt 3.8 ppm

(HCl) (HCl)

HSO4- 71 ppb 19000 ppm

SO42- 73 ppb 3 ppb 0.14 ppt (H2SO4) 0.02 ppt 98000 ppm
(H2SO4)

OH- 190 ppb 20 ppb 0.02 ppt (NaOH) 25 ppt

pH (-log 9.04 6.27 3.10


[H+]

*Concentrations at 347 and 100oC (657 and 212oF) are at equilibrium (values <0.01 ppt are not shown)

The partitioning constants for the salts, acids and bases are shown in Figure 5-16, where the
partitioning constant, KD, is defined for a simple 1:1 electrolyte as the ratio of the concentration
of the neutral molecule in the vapor phase to the activities of the component ions in the liquid
(41)
phase .

5-34
Turbines

Tc 300 200 150C


20

Partitioning Constants

10

HCI (1:1)

Log KD 0
NH4CI (1:1)

NaCI (1:1)

-10 H2SO4 (1:1)

NaOH (1:1)
CH3COONa (1:1)
NaH2PO4 (1:1)
NH4HSO4 (1:1)
-20

NaHSO4 (1:1)

1.6 1.8 2.0 2.2 2.4 2.6


1000 K/T

Figure 5-16
The Logarithm of the Partitioning Constants for 1: 1 Solutes as a Function of Reciprocal
Temperature in Kelvin

Source: Reference 14

Reference 42 described the measurement of vapor - partitioning of sulfuric acid and ammonium
sulfate. Using the published partitioning constant for ammonia, estimates were made of the
contributions of sulfuric acid and ammonia to the volatility of ammonium sulfate and bisulfate.
Sulfuric acid is the dominant form of sulfur (VI) species in steam under most operating
conditions. The dominant sulfur vaporous carryover under mildly reducing conditions is sulfur
dioxide (SO2), which may lead to sulfate deposits due to reoxidation. Ammonium bisulfate was
the other form of sulfur (VI) present in steam at high pH under AVT conditions. Owing to the
effect of hydrolysis (and the intrinsic low volatility of 1:2 electrolytes, this salt was not detected
in any of the vapor phase samples.

A key conclusion from these studies was that the chemistry is significantly more complex than
can be predicted from the simple ray diagram, particularly with the addition of more potentially
volatile species. With AVT, carryover of chloride to steam is most probably by hydrochloric

5-35
Turbines

acid, and not ammonium chloride. The carryover of sulfate is probably predominantly as sulfuric
acid, although ammonium and sodium bisulfate are also likely to contribute to carryover. Thus,
the transport of chloride occurs mainly as HCl. Sulfuric acid and NH4HSO4 both contribute to the
steam inventory of total sulfate, with NaHSO4 also being significant at high sodium levels in the
boiler water. Ammonium sulfate and sodium sulfate were shown to be insignificant in the steam.

The partitioning constants that are available from the EPRI program are summarized graphically
in Figure 5-17 for neutral compounds and Figures 5-16 and 5-18 for 1:1 and 1:2 compounds,
respectively. Note that since these compounds have different stoichiometries (neutral, 1:1, 1:2),
the relative positions of the curves do not indicate the relative volatilities.

The volatilities of acetic acid, formic acid and sodium acetate were reported in EPRI report
TR-113089 (43). The partitioning was measured experimentally at temperatures between 30oC
o o o
(86 F) and 250 C (482 F). The partitioning constant for sodium acetate was found to be
comparable to the partitioning constants of other simple 1:1 electrolytes, so it is much less
volatile than acetic acid. The relative volatilities of acetic acid and sodium acetate favor the
former by several orders of magnitude at high temperatures and low concentrations. The
contribution of sodium acetate to partitioning is then negligible. As a result of the relatively weak
acidity of acetic and formic acids, their partitioning behavior can be expected to depend
significantly on pH. Startup and shutdown may give rise to conditions where the partitioning of
acetic acid is very different from under normal operating conditions.

Further information on the volatility of acetic and formic acids, and phosphoric acid was given
(44)
by Palmer et al. . The partitioning constants for these neutrally-charged solutes are shown in
Figure 5-20 as a function of reciprocal temperature (Kelvin). The volatilities (all neutral species
in steam) may be compared directly and at high temperatures the order of decreasing volatility is
NH3 > SO2 >CH3COOH> HCOOH >>H3PO4. Formic acid is less volatile in aqueous solutions
than acetic acid, although its molecules are smaller and its boiling point is lower. This is due to
the strong interactions between formic acid and water. It should be remembered that the actual
concentration in the vapor phase also depends on the pH of the liquid, as acids tend to dissociate
with increasing pH.

It is clear from Figure 5-16 that the volatilities of these 1:1 solutes vary over many orders of
magnitude. This is particularly true at lower temperatures and this dictates their relative
concentrations in the early condensate.

The order of volatilities of the various compounds investigated is as follows: carbon dioxide >
ammonia > sulfur dioxide > acetic acid > formic acid > > phosphoric acid hydrochloric acid >
ammonium chloride > sodium dihydrogen phosphate = sulfuric acid > sodium chloride > sodium
hydroxide > sodium hydrogen sulfate.

5-36
Turbines

Tc 300 200 150C


20

NH3 (N)
Partitioning Constants
SO2 (N)

10 CH3COOH (N)

HCOOH (N)

Cu(OH)2 (N)
Log KD 0

H3PO4 (N)

-10

-20

1.6 1.8 2.0 2.2 2.4 2.6


1000 K/T
Figure 5-17
The Logarithm of the Partitioning Constants for Neutral Solutes as a Function of
Reciprocal Absolute Temperature

Source: Reference 14

5-37
Turbines

Tc 300 200 150C


20

Partitioning Constants

10

Log KD 0

-10
Na2HPO4 (1:2)

-20 Na2SO4 (1:2)

1.6 1.8 2.0 2.2 2.4 2.6


1000 K/T
Figure 5-18
Partitioning Constants-1:2 Compounds

Source: Reference 14

Gruszkiewicz et al. (45) measured the concentrations of phosphate and sodium in liquid and vapor
phases at 250, 300 and 350oC (482, 572, and 662 oF). The species considered in the liquid were
phosphoric acid (H3PO4), dihydrogen phosphate (H2PO4-), monohydrogen phosphate (HPO42-) and
3-
phosphate (PO4 ), and those in the vapor were phosphoric acid, sodium dihydrogen phosphate
(NaH2PO4) and sodium monohydrogen phosphate (Na2HPO4). The volatility of Na3PO4 was
neglected, since attempts to detect it in the vapor phase at 350oC (662 oF) yielded nothing. Most
of the phosphate exists as HPO4 and H2PO4, either anions or neutral species formed with the
sodium cation. The volatilities of the phosphate species decrease in the order phosphoric acid >
sodium dihydrogen phosphate > sodium monohydrogen phosphate > sodium phosphate. Table 5-
o o
12 gives examples of the concentrations of phosphate species in steam at 350 C (662 F).

Table 5-12
Example Calculation of Phosphate Levels in Steam at 350oC (662oF)

5-38
Turbines

Concentration of phosphate in steam (ppm)

Conc. in water = 0.1 ppm Conc. in water = 10 ppm

a) H+ + H2PO4- 0.0013 13

b) Na+ + H2PO4- 9.0 x 10-8 9 x 10-4

c) 2Na+ + HPO4 2- 7.2 x 10-10 7.2 x 10-4

Phosphoric acid in power plant water/steam cycles originates from hydrolysis of the phosphate
treatment chemicals added to drum boilers. Phosphoric acid is a relatively weak inorganic acid,
which like all neutrally-charged acids, becomes weaker with increasing temperature, so that
H3PO4 may represent the dominant phosphate specie in steam. Phosphoric acid is surprisingly
volatile at high temperatures.

The volatility of copper was assessed based on available data on the thermodynamics of
copper/copper oxide solubility, hydrolysis and complexation, solubility in steam and the
(10)
influence of complexing species on copper solubility in steam. Some of the important
findings were summarized, as follows:
Cupric oxide appears to be the most stable solid phase present in plant cycles.
Cupric oxide is more soluble in steam than cuprous oxide or metallic copper.
Solubility of copper oxides increases with temperature and steam density.
Copper carbonate (CuCO3) may be able to enhance the volatility of cupric oxide.
Cuprous chloride (CuCl) may enhance the solubility of cuprous oxide.
Cuprous chloride complexes could exhibit high volatility. `
Copper-ammonia complexes are less volatile than copper oxides but may enhance the
solubility of cupric oxide.

The research described in EPRI report 1000457 (1) was designed to determine the vaporous
carryover of copper, cuprous and cupric oxides from drum boilers. The following key results
were derived:
The stable copper specie that partitions to steam from (boiler) water in contact with CuO is
Cu(OH) and with Cu2O is CuOH. It should be noted that Cu(OH), has a very high
o o
partitioning constant that is relatively insensitive to temperature from below 150 C (302 F) to
o o
above 350 C (662 F).
The solubility of cupric oxide in (boiler) water is minimal over the normal working pH range
of a boiler, but increases at higher pH. The solubility of CuO increases slightly with
temperature in neutral to basic pH solutions.
Solubility data for cuprous oxide indicate a strong dependence on pH exists at (boiler) water
pH conditions. The solubility of Cu2O in water, at pH values from neutral to basic, increases
o o
dramatically with temperature, appearing to reach a maximum at 482 F (250 C).

5-39
Turbines

The solubility of Cu2O in water appears to decrease slightly with increasing temperature,
which agrees with previous estimates.
The effect of ammonia is to increase the solubilities of CuO and Cu2O at lowtemperatures
and at low to near neutral pH. The effect decreases with temperature, but is stronger for
Cu2O, which persists to at least to 250oC (482oF).
The solubility of cupric oxide in steam is very low (about 2 ppb) over a wide temperature
range (150-350oC) (302-662oF). The solubility of cuprous oxide appears similar.
The solubilities of CuO in steam measured in this study are consistent with the early results
of Pocock and Stewart (Figure 5-19). However, their findings that CuO is more soluble than
Cu2O in supercritical steam appears inconsistent.

Copper must be in the oxide form to volatilize from the boiler. Reducing environments favor the
formation of cuprous oxide, which is less volatile than cupric oxide, as shown earlier in Figure 5-
13 and in 5-19 (2, 10). This figure illustrates that cuprous oxide is less soluble in steam than cupric
oxide. Recent Program Copper research indicates that the two oxides have similar solubilities in
steam. Below a pressure of about 2300 psi (~15.8 MPa), both forms of copper oxide have
minimal solubility in steam.
(14)
EPRI report 1001042 also contains a good summary of the model and predictions based on the
volatility of impurities in water/steam cycles. The model predicts only the vaporous carryover,
but with power plant boilers, vaporous carryover is always accompanied by mechanical
carryover and further impurities can reach the steam via attemperation sprays. Using the
program, it is possible to make predictions for various methods of conditioning.

With AVT/OT, chloride is much more volatile than sulfate and sodium (Figure 5-20). Chloride is
transported largely as hydrochloric acid, with ammonium chloride playing only a minor role, and
sodium chloride hardly any role at all. Sulfate is normally transported as sulfuric acid, with
ammonium bisulfate and sodium bisulfate playing significant roles under certain conditions.

5-40
Turbines

Figure 5-19
Solubility of Cuprous (Cu2O) and Cupric (CuO) Oxides in Superheated Steam as a Function
of Pressure (from data by Pocock and Stewart)

Source: References 2 and 10

Sodium sulfate plays no measurable role. The chloride contents of the steam and early
condensate from vaporous carryover are doubled by relatively small amounts of mechanical
carryover (e.g. 0.05%). The volatilities of sodium and sulfate are much lower than chloride and
the concentrations of these constituents can be doubled by smaller amounts of mechanical
carryover.

With caustic treatment, the addition of sodium hydroxide to the boiler water significantly reduces
the amount of chloride carried over into the steam by vaporous carryover, compared with AVT.
It also counteracts the effects of sulfuric acid ingress better than AVT. However, there is a need
to strictly limit the amount of mechanical carryover (e.g. <0.1 %) to reduce the risk of exceeding
the limits for sodium in steam.

With equilibrium phosphate treatment, the vaporous carryover of chloride is significantly lower
than for AVT. Phosphate treatment deals much better with sulfuric acid ingress than AVT,
particularly in suppressing chloride vaporous carryover. The vaporous carryover of sodium (and
phosphate) is higher than with AVT. With oxygenated treatment, there is a need to carefully
restrict the amount of mechanical carryover to limit the concentration of sodium in steam.

5-41
Turbines

Saturation Pressure (MPa)


9.8 11.3 12.9 14.6 16.5 18.7 21.0
100 1

Cl

10-1 0.1
Concentration of Ion in Steam (ppb)

10-2 Na 0.01

SO4

10-3 0.001

10-4 0.0001

10-5 0.00001

310 320 330 340 350 360 370


(590) (608) (626) (644) (662) (680) (698)
Temperature C (F)

Figure 5-20
An Example of Vaporous Carryover vs Temperature for All-Volatile Treatment (Trends)

Source: Reference 14

Sulfuric acid ingress considerably increases the vaporous carryover of sulfate and, to a lesser
degree, chloride. Sodium hydroxide ingress increases the carryover of sodium and, to a lesser
extent, chloride. Because the concentrations in the boiler water are much higher for phosphate
treatment than for equilibrium phosphate treatment, the concentrations of sodium and phosphate
in steam and early condensate are higher due to both vaporous and mechanical carryover. Even
a small amount of mechanical carryover (e.g. 0.02%) can more than double the amount of
chloride already present in the steam due to vaporous carryover. Strict limits for mechanical
carryover need to be applied for sodium (and phosphate) to prevent exceeding the limits for
steam.

In summary, the concentrations of impurities in the steam due to carryover can reach ppb levels
and those in early condensate (0.1% condensation), ppm levels. Unless the boiler water is very
pure, even small amounts of mechanical carryover can produce more impurities in the steam and
early condensate than vaporous carryover.

5-42
Turbines

Contaminant Solubility in Steam and Its Deposition

Lindsay (4) used a Mollier diagram to describe the behavior of sodium hydroxide and sodium
chloride in steam as it passes through the turbines of a fossil unit. After evaporation, the steam
passes through the superheater, expands through the HP turbine, is reheated and expands again
through the IP turbine, and passes through most of the LP turbine before it enters the wet region.

The curves for NaOH and NaCl described by Lindsay, are included in the Mollier diagram
(46)
Figure 5-21. Solutions of sodium hydroxide can become very concentrated; these concentrated
solutions have elevated boiling points so they can exist in the superheated steam. These
concentrated solutions cover occur for almost all of the steam expansion line. It is only in the
wet region, near the outlet of the LP turbine, where the sodium hydroxide concentration becomes
so dilute as to be non-corrosive. Caustic contamination could put a quantity (say 100 ppb) of
sodium hydroxide in the steam, which could produce a 90% concentrate of sodium hydroxide as
a liquid somewhere in the IP turbine.

Another contaminant of importance is sodium chloride. The thermodynamic properties of


sodium chloride are very different from those of sodium hydroxide, because it has a more limited
solubility. The shaded area in the Mollier diagram is the only region where concentrated salt
solutions are stable. Above the shaded band, only pure dry sodium chloride is stable in the dry
superheated steam. Below the band, in the wet region, salt contamination becomes so dilute as
to be without much corrosive significance. Along the upper boundary of the shaded region there
are solutions that are saturated with solid salt at the point where the LP expansion line crosses
into this region, the concentration is 28% sodium chloride at the upper boundary and close to 0%
at the lower boundary. The important feature is that corrosive salt solutions are
thermodynamically stable, only within one stage of the LP turbine. This location can shift
slightly during load changes.
(6)
Gould noted from the work of Allmon and Mravich on sodium hydroxide in a flowing
autoclave through which steam was passed over solid sodium hydroxide, that there is a tendency
for condensation of high concentration of sodium hydroxide solutions even at concentrations of 1
ppb sodium hydroxide in the steam in the later stages of the low pressure rotor (see Figure 5-9 in
Section 5.2). It is likely that the actual onset of condensation of sodium hydroxide solutions is at
temperatures and pressures below those predicted by this equilibrium analysis just as the Wilson
line (wet-dry transition for pure water) is depressed below the equilibrium saturation line.
(7)
Jonas noted that deposition from steam, as a dynamic process, is a balance between three
simultaneous processes: precipitation from expanding steam, deposition on a surface, and
erosion of the deposit. This is under steady conditions in a dry steam region. With load changes
and under shutdown and start-up conditions, a deposit can be washed away by moisture. With
steam composition changes from high to low concentrations a deposited salt can start dissolving
back in dry steam and be carried away. Chemical compounds precipitate and deposit when their
concentration in superheated steam exceeds solubility. Since the solubility in dry steam
decreases as the steam expands, this type of deposition mechanism is most likely to occur on
components operating at lowest pressures near the PTZ. Evaporation occurs in wet steam
regions where the temperature of metal surfaces is higher than the steam temperature. On such a
surface, the moisture evaporates and non-volatile impurities are left.

5-43
Turbines

Solubility, mmol/kg: Corrosion Rate, mm/year

Corrosion Linear

Magnetite Fe(OH)42-
101
solubility

100
Linear Fe2+

10-1

10-2 Fe(OH)3-
Fe(OH)+
Cubic
Parabolic

10-3

10-4
Fe(OH)2

2 4 6 8 10 12
pH 300C

100 10-2 10-4 10-4 10-2 100


Conc. HCl, mol/kg Conc. NaOH, mol/kg
Neutral

Figure 5-21
Mollier Diagram for a Fossil Plant

Source: Reference 46

EPRI report TR-108184-Vl (24) summarized the concentration of impurities on turbine surfaces by
deposition from superheated steam as:
Precipitation of salts, acids, hydroxides and oxides from superheated steam after the
solubility limit is reached and deposition of the precipitates.
Evaporation of moisture on the surfaces hotter than the saturation temperature and retention
of the mineral residue. This process can be partially controlled by design and by control of
steam purity.
Deposition of oxide particles (with or without adsorbed impurities) and salt particles formed
elsewhere in the steam cycle (mechanical boiler carry-over, exfoliation in the superheater and
reheater, etc.).

5-44
Turbines

Adsorption of dissolved and volatile impurities on oxidized surfaces. This process can be
slowed down by control of steam purity. Early deposits formed on clean surfaces are patchy,
irregular submicron areas.

Sampling showed that steam contained particulates of magnetite. Depending on the size of
magnetite particles or the surface area, up to 6 ppb of sodium, 0.6 ppb of sulfate, and 0.3 ppb of
2
chloride could be adsorbed on 30 ppm of magnetite in superheated steam and 3.4 g/cm sodium,
2 2
0.4 g/cm of sulfate, and 0.1 g/cm of chloride was found on oxidized surfaces.

McCloskey et. al. (8) noted that in the guidelines for fossil units, the levels of impurities in inlet
steam should be small, in the order of ppb. Because of various concentration mechanisms, even
such low levels of particularly harmful impurities, such as chloride and sulfate, can result in
harmful deposits and serious turbine damage. Table 5-13 summarizes the concentration
mechanisms, typical locations where they occur in the turbine and means of control. In the "wet"
zone, after the saturation line, deposition in the steam flow path is mostly due to precipitation
from the steam. In the "dry" zone, prior to the saturation line, deposition of salts, oxides and
other contaminants can occur as a result of the salt zone or by impaction mechanisms.

Reference 1 considered steam in the superheater and steam in the HP turbine in some detail. In the
superheater, the partitioning of copper oxides into saturated steam is in the form of dissolved
CuO/Cu2O/Cu(OH)2/CuOH and particulates. While CuO (and Cu2O) has increasing solubility in
(boiler) water with increasing pH, the solubility in steam is very low and a function of pressure,
with a possible maximum in steam of around 2-4 ppb. Above a steam pressure of 2300-2400 psi
(15.8-16.5 MPa) the data indicate that CuO is more soluble in steam than Cu2O. Below this
pressure, both copper oxides have very low solubility in steam. This pressure effect controls the
transport of copper from the drum (saturated steam) to the primary superheater and then to the
turbine.

During a unit startup, when the drum pressure is less than this "critical pressure," any
CuO/Cu2O/Cu(OH)2/CuOH in the saturated steam will precipitate onto primary superheater
surfaces as crystalline growth. If any moisture is present in the saturated steam, the copper
oxides will remain soluble until superheated steam predominates (no moisture remains), and then
these oxides will also precipitate out on the tube surfaces. As the startup continues and the drum
pressure becomes greater than the threshold pressure of about 2300-2400 psi (15.8-16.5 MPa),
the copper oxides that have partitioned from the boiler water will remain soluble in steam, but
only up to 2-4 ppb limit. The superheated steam now can transport the copper oxides towards
the turbine. It should be remembered that the flowing superheated steam has the ability to
re-dissolve the copper oxides from the tube surfaces up to the solubility limit for that pressure. In
units operating markedly above the threshold pressure (overpressure operation), copper oxides
can reach the HP turbine directly without precipitation from the steam at intermediate points.

5-45
Turbines

Table 5-13
Impurity Concentration Mechanisms in Steam Turbines

Description of Locations Likely to be Can Problem be Control by


Mechanism Affected Avoided?

Precipitation and Depends on impurity, can Yes Steam purity-the higher


deposition from occur in HP, IP, or LP the steam purity the fewer
superheated steam although most common in problems that will be
phase transition zone of LP. encountered.

Formation of droplets LP starting in phase Yes As above.


transition zone.

Evaporation and drying Surfaces above saturation No Design


of wet steam on hot because of heat transfer
surfaces through metal.

Locations of flow stagnation. Design and operation


(moving of the phase
transition zone).

Concentration in surface Depends on impurity, can No, Keeping surfaces clean


oxides (processes such occur in HP, IP, or LP by reduced deposition of
as ion exchange, although most common in but can be oxides, control of water
chemisorption, capillary phase transition zone of LP. controlled. chemistry, and design.
condensation, capillary
boiling, etc.)

Distribution of solubles As above. Yes Steam purity and design


from steam to moisture. so as to avoid stagnant
liquid.

Source: Adapted from Reference 16

In the HP turbine, this involves the solubility of the CuO/Cu(OH)2/Cu2O/CuOH mixture, their
precipitation onto superheater tube surfaces, and re-dissolution into the steam. The largest
pressure drop occurs across the first controlling stage in the HP turbine (Curtiss stage), and
within the first few stages the pressure drops below the threshold pressure for the solubility of
copper oxides and hydroxides. The CuO/Cu(OH)2/Cu2O/CuOH individually and together, can
then precipitate onto the blade surfaces as crystalline growth.

The predominant mechanism for copper carryover into steam appears to be related to the
(2)
solubility of cupric and cuprous oxides in steam. The solubilities found by Pocock and Stewart
were strongly influenced by pressure, (Figure 5-20) and the specific volume of steam, but are
relatively independent of temperature, at least over the range 900-1150oF (482-621oC). In
particular, the solubility for CuO increases sharply above 15.8 MPa (2300 psi) and for Cu2O
above 17.2 MPa (2500 psi), with the CuO being more soluble. Cupric oxide (CuO) has about
twice the solubility of cuprous oxide (Cu2O) at 17.2 MPa (2500 psi), which in turn, is more
soluble than Cu in steam. Table 5-14 shows the maximum solubilities of copper in steam.

5-46
Turbines

Table 5-14
Maximum Copper Solubilization at 4500 psi, 1150oF (30.1 MPa, 621oC)

Species Solubility ppb

pH 7.5 pH 9.5

Metallic Copper 6.3 6.3

Cuprous Oxide 8.9 10.2

Cupric Oxide 15.4 17.2

Jensen and Bursik(27) considered the solubility of sodium chloride, sodium hydroxide and sodium
sulfate within the context of a density model. The entropy/enthalpy diagrams developed can be
used for identifying the parts of the turbine likely to be affected by deposition and the critical
steam contaminant concentrations. The wet region in the low-pressure turbine is especially
sensitive to the presence of impurities in steam. Many impurities have a tendency to concentrate
in the first droplets or in the liquid film formed in the low-pressure turbine where there is a
potential risk of corrosion initiated by the steam impurities.

The density model" for the solubility of salts in superheated steam was based on:
log S = F+ G logw Equation 5-2

where S is the solubility of the salt and w the steam density. F is a function of temperature and
G may be constant or a function of temperature and steam density.

Solubility relationships were established for NaCl, NaOH and Na2SO4 as a function of density at
temperatures of 380, 390, 400, 410, 420, and 430oC (716, 734, 752, 770, 778, and 796oF). These
were confirmed by reference to results published by other workers. Solubility diagrams were
then produced for these compounds.

Figure 5-22 gives the solid-liquid-vapor (SLV) line for NaCl-H2O indicates where concentrated
sodium chloride solutions may form. It should only be possible to form concentrated sodium
chloride in the LP turbine or in low and high-pressure heaters on the steam side. If the steam
content of sodium is less than 1 g/kg (1 ppb), deposits of sodium chloride may be found in the
last part of the IP and in the LP. Figure 5-23 shows the solubility of sodium hydroxide (given as
sodium) in steam. It shows that if the sodium content in steam reaches a concentration of about
1 g/kg sodium, then concentrated sodium hydroxide solution may form in IP and LP turbine.
Figure 5-24 shows solubility curves for sodium sulfate, shows that there is a potential risk of
deposition of Na2SO4 in the reheater, as well as turbines, if the sodium content is about 1 g/kg in
steam.

5-47
Turbines

Figure 5-22
H, S Diagram with Turbine Expansion Curves and Solubility Curves for NaCl (Sodium
Chloride Concentration Expressed as Sodium)
Source: Reference 27

Figure 5-23
H, S Diagram with Turbine Expansion Curves and Solubility Curves NaOH (Sodium
Hydroxide Concentration Expressed as Sodium)
Source: Reference 27

5-48
Turbines

Figure 5-24
H, S Diagram with Turbine Expansion Curves and Solubility Curves for Na2SO4 (Sodium
Sulfate Concentration Expressed as Sodium)
Source: Reference 27

5.3 Methods of Control and Prevention of Deposits

Methods of diagnosis, control, prevention and corrective actions are considered in this part of
Section 5 on turbine deposition. First, the general aspects of minimizing deposition, monitoring
and control are considered. There has been a lot of concern recently about copper deposition and
this is considered in the second part. The removal of turbine deposits by washing or chemical
cleaning is considered in the final part of Section 5.4.

General Aspects of Minimizing Deposition


(7)
Jonas recognized in 1982 that the allowable concentrations of impurities in steam were in the
low ppb range and that the equilibrium solubility was not applicable to steam chemistry control.
Information was required on dynamic solubility. Sodium concentration and cation
(11)
conductivity were considered to be the best indicators of steam purity. Silica (SiO2)
deposition could be almost eliminated, if the concentration in steam was kept below 15 ppb SiO2.
(47)
EPRI report NP-3487 reported that the equilibrium solubility data for sodium hydroxide and
sodium chloride indicated that corrosive deposits would be expected in most turbines. Sodium
hydroxide in the steam at concentrations more than the equivalent of approximately 1.5 ppb
sodium, at the inlet of a typical LP turbine, would be expected to deposit on the surfaces within
the turbine upstream of the wet-dry zone. Fortunately, most turbines can be operated without
damage, when the actual vapor-phase sodium concentration is 1 ppb to 2 ppb.

5-49
Turbines

(48)
Ball and Jenkins conducted extensive surveys of steam circuit chemical conditions of 165 bar
(2450 psi) drum boilers under startup and steady load conditions. Under steady-load conditions,
the steam from the 500 MWe units operating with caustic treatment typically contained 0.25
g/kg sodium and less than 0.2 g/kg chloride and sulfate throughout the steam circuit. This was
equivalent to 0.015% carryover of boiler water impurities into steam. However, rapid load
changes and abnormally high boiler drum water levels increased the carryover. The saturated
steam from the 350 MWe units also operating with caustic treatment contained typically 3 g/kg
sodium, 1.2 g/kg chloride and 1.7 g/kg sulfate under steady-load conditions, corresponds to
0.2% carryover. However, although the 200 MWe units produced acceptable steam purity when
operating on all volatile treatment (the normal method of operation), typically 1.0 g/kg sodium,
1.2 g/kg chloride and 5 g/kg sulfate in the saturated steam during steady-load conditions and
short duration tests using dilute caustic treatment, carryover was typically 0.8%, but could be up
to 1.7%.

Significant reductions of sodium, chloride and sulfate were observed for the 350 and 200 MWe
units, indicating deposition, particularly in the superheaters (Table 5-15).

Some carryover appears to be inevitable and when concentrations of sodium, chloride and sulfate
reach 1 g/kg, there are measurable losses through the steam circuit indicating deposition. Some
salts may re-entrain in the steam but sodium sulfate tends to accumulate in the reheaters.
(34)
Gabrielli and Goodstine noted that the guidelines published by the American Boiler
Manufacturers Association (3rd Ed 1982) included a calculation of the equivalent total
carryover as the ratio of steam to boiler water solids. The values excluded silica because of the
high degree of vaporous carryover for this material. High pressure utility boilers operating at a
drum pressure over 1800 psig (12.4 MPa) were assumed to have a mechanical carryover of no
more than 0.2%. Vaporous carryover was considered significant only at the highest pressure
range listed; an additional allowance of 0.13% vaporous carryover is provided in this situation.
They noted that mechanical carryover depended on the design and condition of the drum
internals, and was affected by load and drum water level. Tests indicated that mechanical
carryover remained at very low levels with wide variations in these parameters. However, when
critical values were exceeded, carryover increased dramatically and as little as a 5% increase in
load, could cause an order of magnitude increase in carryover. Testing the carryover
performance of the steam drum is done for warranty purposes, but tests should also be carried
out periodically to monitor the amount of impurities introduced into the steam and carried on to
the turbine
(15)
Svoboda et al. recommended that acceptance tests should always include the measurement of
carryover. Tracer measurement techniques are most useful. A worsening of the carryover of a
boiler during operation, if not caused by the chemistry, could be caused by failure of the
moisture separators.
(49)
EPRI report TR-103665 describes the rationale for target levels of sodium, chloride, sulfate
and silica in steam for drum units operating with phosphate treatment (PT):

Monitoring of sodium is necessary because sodium hydroxide and sodium chloride are two
major corrodents of the turbine, especially in relation to stress corrosion cracking of turbine
blades and stainless steel superheater tubes. The sodium target value in reheat steam recognized

5-50
Turbines

that some of the sodium present would be in a non-corrosive form (sodium phosphate
compounds) and a limit value of 5 ppb was selected.

Monitoring of chloride is necessary because chloride contributes to corrosion fatigue, stress


corrosion cracking, and pitting in LP turbines. The chloride target in steam was based on the
solubility of sodium chloride in superheated steam, the limit was 3 ppb.

Monitoring of sulfate is necessary because sodium sulfate in combination with chloride


contributes to turbine corrosion, sulfate causes off-load corrosion of steam-side components and
acid sulfate can cause turbine corrosion. The sulfate target value in steam was based on the
solubility of sodium sulfate in superheated steam (considered to be similar to that of sodium
chloride) and that acid sulfate is as corrosive as acid chloride. Thus, the same limit was chosen
for sulfate in steam as for chloride.
Table 5-15
Concentrations of Impurities in Steam under Steady Load Conditions

a) 350 MWe Units - Boilers Conditioned with Caustic Treatment (CT)


Parameter Saturated Superheater Reheater Reheater
Steam Outlet Inlet Outlet
(g/kg) (g/kg) (g/kg) (g/kg)
Sodium 3.0 1.3 0.5 0.5

Chloride 1.2 0.6 0.6 0.6

Sulfate 1.7 1.2 0.6 0.6

b) 200 MWe Units - Boilers Conditioned with All Volatile Treatment (AVT)
Parameter Saturated Superheater Reheater Reheater
Steam Outlet Inlet Outlet
(g/kg) (g/kg) (g/kg) (g/kg)
Sodium 1.0 0.6 0.4 0.4

Chloride 1.2 0.5 0.5 0.4

Sulfate 5 2.5 1.5 1.4

c) 200MWe Units Boilers Conditioned with Caustic Treatment (CT)


Parameter Saturated Superheater Reheater Reheater
Steam Outlet Inlet Outlet,
(g/kg) (g/kg) (g/kg) (g/kg)
Sodium 18 5 0.5 0.3

Chloride 0.7 0.6 0.3 0.3

Sulfate 8 2.5 1.3 1.1

Monitoring of silica is necessary because the precipitation of silica forms silicate deposits on the
turbine that are not soluble in water and are very difficult to remove. Silica deposits cause losses
in turbine capacity and efficiency. The silica target value in steam was based on the solubility of

5-51
Turbines

silica in superheated steam, a limit of 10 ppb was chosen for units with reheat and 20 ppb was
chosen for non-reheat units. These target values should limit deposition of silica in the turbine.

EPRI report TR-105040 (41) on the selection and optimization of boiler water and feedwater
treatments gave the limits for steam for other forms of chemical conditioning. For AVT and
equilibrium phosphate treatment, the limits were <3 ppb sodium, <0.15S/cm cation
conductivity, <10 ppb silica, <3 ppb chloride and <3 ppb sulfate.
(50)
Jonas et. al. described a turbine deposit collector, developed several years ago, that can be
used for troubleshooting to allow collection and analysis of deposits without turbine
disassembly.

The EPRI report on steam, chemistry, and corrosion in the phase transition zone (PTZ) of steam
turbines (24) concluded that there seemed to be a limit for concentration of common impurities in
steam, below which there was no measurable deposition. The limits for sodium, chloride, and
sulfate are between 0.3 and 1 ppb. Findings of this work was considered before tightening of
steam purity guideline values, as described in Revision 1 of the guidelines for all-volatile
(51)
treatment. It appeared that more reliable operation in the steam turbine PTZ might be
accomplished by reducing the steam purity guideline limits for Na, Cl, and SO4 to below 1 ppb
each (for units with reheat). However, there was still transport and deposition of steam
impurities as particles, even in units with concentration of sodium and chloride in steam less than
0.1 ppb. Table 5-16 summarizes the problems, causes, and solutions of transport and deposition
of impurities in fossil fuel cycles.

Copper Deposition

Particular concern has been expressed about copper deposition on turbines, because it can
quickly lead to a reduction in output capacity and efficiency of HP turbines. EPRI report
TR-107754 (22) covers cycling, startup, shutdown and layup of fossil plant. After years of
corrosion of copper alloys, the corrosion products are transported and deposited throughout the
steam cycle, including heaters, boiler, superheater and the turbine. The deposited copper is often
in the form of metallic copper and cuprous oxide (Cu2O), because during operation there is a
reducing environment. During unprotected layup, some of these deposited species oxidize to
cupric oxide (CuO), which can then be dissolved and transported downstream. For mixed
metallurgy systems, it is clear that reducing conditions (established with use of a reducing agent
such as hydrazine) are required. The EPRI guidelines for AVT suggest that oxygen levels should
be kept below 5 ppb at the economizer inlet (oxidation/reduction potential, ORP <0mV).

5-52
Turbines

Table 5-16
Problems, R&D, Root Causes, and Solutions

Problem Under- R&D* Root Causes* Solutions


standing

LP blade corrosion fatigue 70% CH, D D, CH Design - reduce vibration, improve


steam chemistry, inject buffers

SCC of blade attachment 60% CH D, (CH) Redesign - lower stress, improve


steam chemistry, inject buffers,
(shot peening)

LP turbine efficiency loss 30% CH, D Thermodynamic Modification of condensation by


vs. moisture design chemicals and by application of
high voltage, removal of moisture

MW loss due to Cu deposits 70% C, V, S Cycle design with Replace Cu in HP heaters,


in the HP turbine Cu, high boiler possibly LP heaters, optimize
pressure, sensitive chemistry, reduce boiler pressure
turbines when Cu is high, clean SH, turbine
chemical wash

MW loss due to HP + IP 70% V, S Too much Reduce phosphate concentration


turbine phosphate deposits phosphate at high in boiler water, go to EPT or AVT
boiler pressure (need cond. polishers), turbine
wash with wet steam

Flow assisted corrosion in 80% CH Acidified moisture Minimize or eliminate oxygen


wet steam extraction piping, drops, scavenger, increase pH of
LP rotors ox-scavenger, flow moisture eliminate acids, use alloy
velocity and steels, reduce flow velocity and
turbulence turbulence

Water droplet erosion of 90% D, (CH) High % moisture Erosion-resistant shields, reduction
blades large droplets of moisture, velocity and mass flow

*CH - Chemistry, M - Material, D - Design, C - Corrosion, V - Volatility, S Solubility

The following actions should lead to reduction in copper problems around the cycle: (1)
Keep feedwater copper levels at guideline values (<2 ppb at the economizer inlet) during
normal operation.
Establish conditions in the feedwater which favor metallic copper or cuprous oxide (Cu2O)
rather than cupric oxide (CuO) under all operating conditions.
Maintain reducing chemistry (ORP <0 mV) at all times, including shutdown and startup.
This necessitates using a reducing agent.
Control feedwater pH in the range 9.0 to 9.3.
Implement shutdown procedures and layup programs which effectively minimize copper
transport during return to service.

5-53
Turbines

At startup, as the drum pressure exceeds the threshold pressure, copper oxides that have
partitioned from the boiler water will become soluble in steam, but only up to 2-4 ppb. It is very
important to recognize that chemical cleaning of the primary superheater will be needed to
remove deposited copper. It is also clear that copper deposits in the primary superheater are
unavoidable in any unit with mixed metallurgy. But turbine deposition problems will only occur
in units that operate above the threshold pressure.

During normal unit operations, full compliance with EPRI chemistry guidelines should be
practiced. The guidelines are intended to minimize copper corrosion and transport in the
condensate and feedwater systems and to control volatile carryover from the boiler drum, and
need to be integrated into plant chemistry programs for units with copper alloys in feedwater
heaters. Copper levels in saturated and main steam should consistently be less than 2 ppb.

Actions which should be considered during the maintenance outages include:


Insure proper lay-out of all components of steam-feedwater cycle. For once-through steam
generators insure that the superheater and reheater are protected from contact with oxygen.
Remove tube samples from the primary superheater to appraise the extent of copper
deposition and, if needed, efficacy of deposit removal methods.
If the turbine will not be open, plans should be made to examine the inlet and high pressure
section using fiber optics via any available access ports.
If signs of deposition are present near the turbine inlet, it may be possible to collect a small
sample for analysis of copper and other constituents.

Table 5-17 lists some of the solutions to copper control problems. Short term solutions 1-4
emphasize removal of copper deposits formed during unit operations. Long-term solutions 5-13
focus on equipment changes which either eliminate existing sources of copper in the cycle or
reduce the impact of copper transport. Chemistry surveillance solutions 14-16 give criteria for
surveillance, both routine operation and monitoring campaigns intended to evaluate copper
transport activity.

Turbine Washing and Chemical Cleaning

If deposition on the turbines can not be prevented or controlled adequately, it may be necessary
to clean the turbine. Water soluble deposits can be removed by water or steam washing. It may
be necessary to remove non-water soluble deposits by grit blasting and, in the case of copper
deposits by chemical cleaning.

5-54
Turbines

Table 5-17
Possible Solutions for Copper Transport Problems

Deposit Removal (Short Term Actions)


Solution 1. Cleaning turbine
Solution 2. Cleaning feedwater heaters
Solution 3. Cleaning boiler waterwalls
Solution 4. Cleaning superheater
Equipment and Operational Changes (Long-Term Actions)
Solution 5. Change feedwater heater materials
Solution 6. Change cycle design/piping changes
Solution 7. Retrofit condensate polisher
Solution 8. Minimize mechanical carryover
Solution 9. Minimize cycle air in-leakage
Solution 10. Minimize oxygen in cycle makeup
Solution 11. Nitrogen blanket boiler and heaters
Solution 12. Optimize feedwater heater operations
Solution 13. Evaluate overpressure operation
Chemistry Monitoring Approaches
Solution 14. Upgrade sample system
Solution 15. Install ORP monitor
Solution 16. Monitor copper transport

The EPRI guidelines for chemical cleaning (20) noted that a survey conducted in the 1980s of
turbine manufacturers and utilities found the average interval between major turbine
maintenance outages was five to six years. More recently, it has increased to seven to eight
years. Such outages require disassembly of the turbine for inspection and maintenance. If
deposits are apparent on a turbine, these are usually effectively removed by grit-blasting, with
either aluminum oxide or other suitable material. If grit-blasting at each major turbine outage is
insufficient, the turbine may have to be chemically cleaned, as is often the case with copper
deposits.

Good steam purity virtually eliminates the need for cleanings more frequent than the average
maintenance outage, when the turbine is disassembled. Elimination of copper-based alloys in
feedwater heaters, installation of condensate polishers, and chemically cleaning steam/water
cycle components that contain residual copper and copper oxide deposits, and operation in
agreement with EPRI copper guidelines, usually eliminates the need for chemically cleaning
copper from turbines.

For water-soluble, non-corrosive turbine deposits, such as sodium phosphate, water-washing can
(1)
restore lost efficiency. There are several different procedures which have been called water-

5-55
Turbines

washing. One involves injecting high purity water in the steam upstream of the turbine as the
unit is shutting down to scour deposits from turbine blading. Such steam/water-washing is
performed occasionally, but the potential for damage makes it an undesirable approach. Yet
another water washing procedure that avoids the use of high purity water injection in steam
consists of sliding the parameters of the steam boiler during the shut-down process all the way
above the deaerator pressure (approximately at 160 C and at the saturation pressure) while
turning the turbine with this steam at zero power generation. Another water-washing procedure
for removing non-corrosive water-soluble deposits, involves allowing the turbine to cool and
then filling it to the turbine shaft (through a shell drain) with high purity water while the turbine
is on turning gear. For large, modern turbines, steam/water-washing should be undertaken only
with great caution and under special circumstances. Water-soluble deposits in turbines also have
o o
been removed with foamed water rinses at 82-88 C (180-190 F). This is essentially the same
procedure used for copper removal, minus the ammonia-based compounds or oxidants.

Grit-blasting produces a soft, grey satin finish and slightly increases the fatigue strength of the
material. Mild sand blasting has been used by at least one utility routinely to remove copper
deposits. Provided no loss in turbine efficiency due to deposition is detected during unit
operation, this practice of inspecting and grit-blasting every five to six years will suffice and is
preferable to chemically cleaning the turbine.

HP turbines have been successfully cleaned to remove copper deposits for many years. The
advantage of chemical cleaning over grit-blasting is that chemical cleaning does not require
disassembly of the unit. The need to chemically clean is usually based on losses in turbine
efficiency or capacity due to deposition. It should be stressed that a turbine chemical cleaning
should be viewed only as a temporary measure to regain capacity or efficiency. If performed in
isolation without addressing the other areas, such as superheater, boiler, turbine and the operating
and shutdown chemical environments, then the problem of deposition in the turbine will return.

The standard method of chemical cleaning uses a foam cleaning operation. Choice of the correct
solvent is critical and will depend on the composition of the deposit and the compatibility of the
solvent with turbine components. A typical solvent consists of about 5 to 6% ammonium
hydroxide, 12 to 16% ammonium bicarbonate, and sufficient foaming agent. Proprietary
solvents for copper and copper oxide removal are also available. The addition of hydrogen
peroxide to the foamed mixture described above is used when copper base metal is present. The
hydrogen peroxide converts the copper to copper oxide and enables dissolution by the ammonia.

5.4 Matrices for Deposition and Adsorption

Factors Linked to the Matrices

The factors given higher scores (4 to 2) for deposition in turbines (originally presented in Tables
2-1 to 2-3 of Part 1 of this study (52)) have been abstracted from these tables and the results are
given in Table 5-18 and discussed below.

High and medium ranked parameters with respect to suspended particles in the turbine in
alphabetical order were:

5-56
Turbines

High (score of 4 or higher): pH at operating temperature.


Medium (score of 3): Concentrations of particles, flow characteristics, and transient
conditions and outages.

High and medium ranked parameters with respect to deposition of soluble impurities and
corrosion products in the turbine in alphabetical order were:
High (score of 4 or higher): Concentration of metal ions and impurities, condensate polishing
and filtrations and solubility in steam and water.
Medium (score of 2-3): Steam quality, transient conditions and outages, chemical treatment,
concentrations of metal ions, mass transport, pH at operating temperature, pressure,
steam/water boiling/condensing regimes, steam/water properties, temperature, time, velocity
and flow rate.

High and medium ranked parameters with respect to surface phenomena (sorption, hideout, etc)
in the turbine in alphabetical order were:
High (score of 4 or higher): Concentrations of metal and non-metal ions, impurities and
particles.
Medium (score of 2-3): Ion exchange properties of metal oxides, mass transport, pressure,
temperature, and transient conditions and outages.

5-57
Turbines

Table 5-18
Factors Given High Scores in Tables 2-1 to 2-3 of Reference 52 for Deposition in Turbines

Deposition of Suspended Deposition of Water Soluble Adsorption and Ion


Solids Material Exchange on Surfaces
pH at operating temperature
Concentration of impurities
Non-metal ion composition and
concentration
Solubility in water and precipitation
of water soluble compounds
Condensate polishing
Conditions during various Conditions during various plant
plant modes (base load, modes (base load, cycling, etc)
cycling, etc)
Moisture carryover
Solubility in steam and steam/water
distribution
Concentration of particles
Steam/water flow
characteristics
pH at operating Temperature
Plant modes (base load,
cycling, etc)
Chemical treatment (pH Chemical treatment (pH control,
control, reducing or oxidizing reducing or oxidizing agents, etc)
agents, etc)
Concentration of metal ions Concentration - metal ion
and impurities composition and concentration
Hardware configuration and
steam/water flow
Velocity and flowrate Velocity and flowrate
Crystalline structure and
amorphous material
Particle size distribution
Mass transport Mass transport Mass transport
Time, including deposition Time, including deposition rate
rate changes changes
Pressure Pressure Pressure
Temperature Temperature Temperature
Steam/water properties Steam/water properties (density,
(density, viscosity, etc) viscosity, etc)
Steam/water boiling and Steam/water boiling and
condensing regimes condensing regimes
Ion exchange properties of
metal oxides
Note: The factors in the upper, center and lower parts of the table were given scores of 4, 3 and 2,
respectively, in Tables 2-1 to 2-3 of Reference 52.

5-58
Turbines

The factors affecting turbines, listed in Table 5-18 are reviewed below:

Chemical treatment can reduce the generation of soluble and insoluble corrosion products in
the feedwater heaters and boiler, and the carryover and deposition on the turbine. Chemical
treatment includes the addition of pH control agents, reducing agents and oxidizing agents.
Chemical additives may alter deposition from steam.

The concentration of metal ions is a measure of the soluble corrosion products generated,
which contribute to the deposition of suspended particles and water soluble material. The
concentration in the bulk liquid depends on corrosion rate of the materials exposed, chemical
treatment, pH, concentration of oxidizing/reducing agents, electrochemical potential (ECP)
and/or oxidation reduction potential (ORP), and the solubility in water/steam. Metallic ions can
be converted to suspended oxide particles as a result of corrosion. Transport to the turbine may
occur as a result of solubility in steam and/or transport of moisture droplets (steam quality).

The concentration of non-metal ions and impurities (e.g. chloride, fluoride, sulfate, carbonate,
organics, etc.) contributes to corrosion and the generation of soluble and insoluble corrosion
products. Non-metal ions and impurities can also participate indirectly by changing solution pH.
Impurities can dissolve in steam and precipitate as the pressure and temperature reduces through
the turbine.

The concentration of particles is a measure of the insoluble corrosion products generated,


which can deposit on the turbine. The concentration of particulate corrosion products is a key
factor in influencing deposition rates. Some particles may be transported out of the boiler as
solid aerosols or with moisture droplets.

Condensate polishing reduces corrosion of the boiler and feed system by removing soluble,
ionic impurities by ion exchange. It also removes suspended material by filtration.

The configuration of the water/steam circuit hardware can contribute to the deposition of
suspended particles, including corrosion products. Steam and water flow can generate soluble
and insoluble corrosion products by erosion and/or flow assisted corrosion (FAC). Other factors
depending on hardware include high velocity, susceptible materials, pH, chemical treatment,
electrochemical potential (ECP) and oxidation/reduction potential (ORP). Hardware
configuration can alter deposition of suspended particles by changes in direction and velocity of
the fluid.

The rate of general corrosion, erosion and FAC affects the generation of soluble and insoluble
corrosion products and suspended. Carbon steel, low alloy steels, copper and copper alloys in
the feedwater systems are the primary sources. Correct material selection and chemical
treatment (pH, oxygen control, etc.) are the primary methods of controlling corrosion. Copper
deposition can be a particular problem for turbines in units with copper alloys in the feedwater
heaters.

Crystalline structure of deposit can affect the nature of the soluble and insoluble corrosion
products from corrosion. Some deposit can be non-crystalline (amorphous). The

5-59
Turbines

crystalline/amorphous structure of deposits may affect the generation, deposition, release and of
suspended particles.

The ion exchange properties of metal oxides, suspended solids and deposits can affect the
deposition of water soluble material. Adsorbed ions on these materials can exchange with the
ions in the water.

Electrochemical potential (ECP) and oxidation reduction potential (ORP) of deposits are
increased by oxidizing agents. Reducing agents decrease the ECP and ORP.

The pH at operating temperature affects the generation of soluble and insoluble corrosion
products, the volatility of contaminants and the deposition of suspended particles. The neutral
point of water changes with temperature. The pH of the chemical solution must be evaluated
relative to the neutral pH of water.

Porosity of the deposition of suspended metal oxide particles provide sites that can enhance
deposition and concentrating of water soluble material on the surfaces of turbines blades.

Cycle pressure affects the volatility of impurities and deposition of material from the steam,
since solubility in steam is a function of steam density, which itself is a function of pressure and
temperature. The contribution of pressure to fluid density and other steam/water properties
means that pressure makes a contribution of to deposition of suspended material and surface
phenomena (adsorption and ion exchange on surfaces).

Steam quality (carryover) can affect the transport and deposition of soluble and insoluble
material in the turbine. Carryover consists of mechanical and vaporous carryover. The steam
separation equipment in drum boilers is designed to separate most of the moisture from the
steam. The mechanical carryover of moisture droplets results in the transport of additional water
soluble material to the turbine. Since many water soluble chemicals are much more soluble in
the moisture than in steam, moisture carryover significantly enhances the transport of water
soluble chemicals with the steam. Transport of suspended particles with the water may also
occur.

Steam and water properties, including density, viscosity, entropy and enthalpy, contribute to
the deposition of soluble and insoluble material from water and steam. Steam and water
properties are a function of temperature and pressure.

Surface roughness contributes to the deposition of suspended particles and water soluble
material present on the surface of particles. Deposits tend to accumulate in the troughs formed
by machining and polishing marks. Surface roughness increases the deposition of particles on
the surface and also reduces the rate of release of deposited particles from the surface. The
impact of surface finish is strongly dependent on bulk fluid velocity and particle size.
Mechanical or electropolishing can be used to reduce surface roughness of turbines.

Temperature is a contributor to the generation of soluble and insoluble corrosion products and
deposition of water soluble material from steam. The solubility of water soluble material in dry
steam is proportional to the steam density, which is itself a function of temperature and pressure.

5-60
Turbines

Increasing temperature through the feedwater system can enhance the generation of corrosion
products. Some of the material deposited in the boiler can be transported to the turbine as a
result of solubility in steam and/or transport with moisture droplets. Deposition of material,
transported with water droplets, occurs when the water droplet impinges and evaporates on hot
surfaces of the turbine. Deposition of steam soluble material occurs with decreasing temperature
within the turbine, but very high flow velocities within the turbine can delay the onset of
deposition.

Time is a factor in the generation of soluble and insoluble corrosion products, deposition of
suspended particles and water soluble material. The rate of generation and deposition normally
decreases with time.

Transient operations, including load changes, cycling and load follow operation, shutdown,
startup and outages, contribute to the generation of soluble and insoluble corrosion products and
deposition, release and re-deposition of suspended particles and water soluble material. Existing
deposits may be released and redeposit at other locations. The impact is generally proportional
to the size of the transient. Transients can alter the rate of deposition and deposit location
throughout the cycle.

Transport of soluble material and impurities on or within suspended particles contributes to


the deposition of water soluble contaminants and possibly to underdeposit corrosion of the
surface onto which the particles deposit. Adsorption and ion exchange may be involved. Other
factors include condensate purification, chemical treatment, the concentration and composition
of impurities, corrosion rate, oxidizing agents, ECP/ORP, particle size, pH, porosity and
solubility in steam.

The composition of porosity of oxides, their thickness and density affect the generation of
soluble and insoluble corrosion products and deposition of suspended particles. Porous deposits
can entrap and concentrate corrosive impurities.

Velocity of the fluid contributes to the generation of soluble and insoluble corrosion products,
deposition of suspended particles and deposition/sorption of water soluble material. Velocity is a
function of flowrate, hardware configuration and steam/water properties and contributes to
generation of corrosion products by erosion and FAC.

Findings for Reducing the Deposition on Turbines

Table 5-18 suggests the following methods of reducing deposition on turbines and the associated
problems due to corrosion, reduced efficiency, output and other failures:
Reduced corrosion in the feedwater heaters, to minimize corrosion and the production of
corrosion products, by ensuring sufficiently pure water, appropriate chemical conditions, i.e.
chemical dosage, pH, mildly oxidizing, except for copper, where reducing conditions must be
maintained.
Maintain the boiler in a clean condition, with the minimum of deposits.

5-61
Turbines

Minimize vaporous carryover of impurities and non-volatile conditioning chemicals in drum


boilers by ensuring correct boiler water pH, chemical purity and appropriate conditions
during off-load, startup and normal operation.
Minimize mechanical carryover of impurities and particulate matter in drum boilers by
maintaining and periodically monitoring the steam separation equipment to ensure that it is in
good condition.
Minimize the transport of material from the superheaters and reheaters during startup.
Minimize deposition in turbines.
If necessary, remove water soluble materials from the turbines by water washing under
carefully controlled condition. It even may be necessary to removed deposits by chemical
cleaning.

5.5 State-of-Knowledge Deficiencies

While the science of these phenomena are now understood and guidelines are available which, if
followed, greatly reduce the risk of turbine deposition and corrosion, the continued occurrence of
such problems within the industry and their associated costs indicate a clear need to better
understand the deposition processes involved to determine if there are other solutions that could
be practically implemented. Some of these factors are listed in Table 5-19.

Two of the most important aspects where more knowledge is required are on mechanical
carryover and the dynamics of deposition.

The curve used in the EPRI guidelines for mechanical carryover (Figure 5-14) was based on
information supplied by boiler manufacturers many years ago. The relationship used is probably
still a reasonable one for most plants, as it incorporates a safety factor of two, and operators are
advised to conduct measurements on their own plants. It is most important to know how
(48)
mechanical carryover can be controlled and reduced. Ball and Jenkins showed that in well
designed and operated plant total carryover could be much lower than given in the Figure 5-14.

Information on the factors that affect the dynamics of deposition and the proportions of particles
depositing and adhesion is needed.

5-62
Turbines

Table 5-19
Areas where there is Insufficient Information about the Factors Affecting Turbine
Deposition or their Relative Importance

Deposition of Suspended Solids Deposition of Water Soluble Adsorption and Ion Exchange
Material on Surfaces
Mechanical carryover Mechanical carryover
Steam quality
Conditions during various plant Conditions during various plant Conditions during various plant
modes (base load, cycling, etc) modes (base load, cycling, etc) modes (base load, cycling, etc)
Hardware configuration and
steam/water flow
Chemical treatment (pH control, Chemical treatment (pH control,
reducing or oxidizing agents, etc) reducing or oxidizing agents, etc)
pH at operating temperature
Concentration of metal ions and Concentration - metal ion
impurities composition and concentration
Porosity (density) of oxides and Porosity (density) of oxides and
deposits deposits
Agglomeration of particles
Crystalline structure and
amorphous material
Surface finish
Condensate polishing
Zeta potential Attraction of soluble material to
deposited solids
Ion exchange properties of metal
oxides

5.6 References

1. Guidelines for Copper in Fossil Plants, EPRI, Palo Alto, CA: 2000. 1000457.

2. F. J. Pocock and J. F. Stewart, The Solubility of Copper and Its Oxides in Supercritical
Steam, American Society of Mechanical Engineers, New York, ASME Publication
61-WA-140, 1963, pp 33-44.

3. O. I. Martynova, Mechanism of "Early Condensate" Formation in the LP Turbine.


Interaction of Iron-Based Materials with Water and Steam, EPRI, Palo Alto, CA: 1993.
TR-102101.

4. W. T. Lindsay, Behavior of Impurities in Steam Turbines, Power Engineering, 83(5),


1979.

5. W. T. Lindsay and P. K. Lee, Condensation of Low-Volatility Impurities in Steam


Turbines, EPRI Workshop on Corrosion Fatigue of Steam Turbine Blade, R.I. Jaffee, Ed.
Pergamon Press, 1983.

5-63
Turbines

6. G. C. Gould, Vapor Phase Carryover and Corrosion of Turbine Materials, EPRI Workshop
on Corrosion Fatigue of Steam Turbine Blade, R.I. Jaffee, Ed. Pergamon Press, 1983.

7. O. Jonas, Characterization of Steam Turbine Environment and Selection of Test


Environments, EPRI Workshop on Corrosion Fatigue of Steam Turbine Blade Materials,
R.I. Jaffee, Ed. Pergamon Press, 1983.

8. Turbine Steam Path Damage. Theory and Practice Volume 1: Turbine Fundamentals, EPRI,
Palo Alto, CA: 1999. TR-108943.

9. Cycle Chemistry Corrosion and Deposition: Correction, Prevention, and Control, EPRI,
Palo Alto, CA: 1993. TR-103038.

10. State-of-Knowledge of Copper in Fossil Plant Cycles, EPRI, Palo Alto, CA: 1997.
TR-108460.

11. Deposition of Corrosive Salts from Steam, EPRI, Palo Alto, CA: 1983. NP-3002.

12. Guidelines for Chemical Cleaning of Fossil-Fueled Steam-Generating Equipment, EPRI,


Palo Alto, CA: 1993. TR-102401.

13. O. Jonas, W. Steltz and R. B. Dooley, Steam Turbine Efficiency and Corrosion: Effects of
Surface Finish, Deposits, and Moisture, Sixth International Conference on Fossil Plant
Cycle Chemistry EPRI, Palo Alto, CA: 2001. 1001363.

14. The Volatility of Impurities in Water/Steam Cycles, EPRI, Palo Alto, CA: 2001. 1001042.

15. R. Svoboda, M. Bodmer and H. Sandmann, Chemistry Requirements for Turbines Working
with Medium and Low Pressure Boilers, Second Cycle Chemistry Conference, EPRI, Palo
Alto, CA: 1989. GS-6166.

16. Turbine Steam, Chemistry, and Corrosion, EPRI, Palo Alto, CA: 1994. TR-103738.

17. R. Svoboda, R. Sandmann, S. Romanelli, and M. Bodmer, M., Early Condensate in Steam
Turbines, Interaction of Iron-Based Materials with Water and Steam, EPRI, Palo Alto, CA:
1993. TR-102101.

18. M. F. Leonard, J. E. Neidhardt, G. J. and Verib, Turbine Deposits on Once-through Units on


Oxygenated Water Treatment, Fifth International Conference on Fossil Plant Cycle
Chemistry, EPRI, Palo Alto, CA: 1997. TR-108459.

19. O. Jonas, and R. B. Dooley, Impurity Concentration Processes in Steam Turbines, Steam
Chemistry: Interaction of Chemical Species with Water, Steam, and Materials during
Evaporation, Superheating, and Condensation, EPRI, Palo Alto, CA: 2000. TR-114837.

20. Steam, Chemistry, and Corrosion in the Phase Transition Zone of Steam Turbines Volume 1:
Key Results, Summary, and Interpretation, EPRI, Palo Alto, CA: 1999. TR-108184-Vl.

5-64
Turbines

21. Guidelines for Chemical Cleaning of Conventional Fossil Plant Equipment, EPRI, Palo
Alto, CA: 2001. 1003994.

22. Steam Turbine Efficiency and Corrosion: Effects of Surface Finish, Deposits, and Moisture,
EPRI, Palo Alto, CA: 2001. 1003997.

23. Cycling, Startup, Shutdown, and Layup Fossil Plant Cycle Chemistry Guidelines for
Operators and Chemists, EPRI, Palo Alto, CA: 1998. TR-107754.

24. M. Ball, A. Bursik, R. B. Dooley, M. Gruszkiewicz, D. A. Palmer, K. J. Shields and J. M.


Simonson, The New Approach to the Volatility of Impurities in Water/Steam Cycles, Sixth
International Conference on Fossil Plant Cycle Chemistry, EPRI, Palo Alto, CA: 2001.
1001363.

25. O. Jonas, R, K, Mathur, J. K. Rice and E. E. Coulter, Development of Steam Sampling,


International Conference on Cycle Chemistry, EPRI, Palo Alto, CA: 1991. TR-100195.

26. O. Jonas and B. Syrett, Chemical Transport and Turbine Corrosion in Phosphate Treated
Drum Boilers, Proceedings of the 48th International Water Conference, Engineers Society
of Western Pennsylvania, 1994.

27. J. P. Jensen and A. Bursik, Solubility of Sodium Salts in Superheated Steam and Related
Deposition Processes, Steam Chemistry: Interaction of Chemical Species with Water,
Steam, and Materials during Evaporation, Superheating, and Condensation, EPRI, Palo
Alto, CA: 2000. TR-114837.

28. O. Jonas, Transport of Ionic Impurities in Fossil and PWR Cycles New Observations,
Proceedings of the42nd International Water Conference, Engineers Society of Western
Pennsylvania, 1994.

29. K. J. Shields, R. B. Dooley, T. H. McCloskey, B. C. Syrett and J. Tsou, Copper Transport in


Fossil Plant Units Fifth International Conference on Fossil Plant Cycle Chemistry, EPRI,
Palo Alto, CA: 1997. TR-108459.

30. R. G. Axley, D. W. Beaver, S. O. Hilton, and M. G. Sexton, Turbine Copper Deposits in a


Supercritical Once-through Unit International Conference on Cycle Chemistry, EPRI, Palo
Alto, CA: 1991. TR-100195.

31. D. A. Palmer, P. Bnzeth, J. M. Simonson, A. Y. and Petrov, The Transport and


Chemistry of Copper in Power Plants as Determined by Laboratory Experiments. Sixth
International Conference on Fossil Plant Cycle Chemistry, EPRI, Palo Alto, CA: 2001.
1001363.

32. R. B. Dooley, A. Aschoff, K. J. Shields and B. Syrett, Copper in the Fossil Plant Cycle.
Sixth International Conference on Fossil Plant Cycle Chemistry, EPRI, Palo Alto, CA: 2001.
1001363.

33. Water, Steam, and Turbine Deposit Chemistries in Phosphate Treated Drum Boiler Units,
EPRI, Palo Alto, CA: 1987. CS-5275.

5-65
Turbines

34. F. Gabrielli and S. L. Goodstine, Utility Boiler Steam Purity Considerations International
Conference on Cycle Chemistry, EPRI, Palo Alto, CA: 1991. TR-100195.

35. B. K. Warwood, F. Roe, J. Sears and R. B. Dooley, Fundamental Mechanisms of Deposition


in Power Plants, Fourth International Conference on Fossil Plant Cycle Chemistry, EPRI,
Palo Alto, CA: 1995. TR-104502.

36. Deposition in Fossil Plants: Mechanisms and Impacts, EPRI, Palo Alto, CA: 1998.
WO9002-04. (Unpublished)

37. J. M. Simonson and D. A. Palmer, An Experimental Study of the Volatility of Ammonium


Chloride from Aqueous Solutions to High Temperatures, Proceedings of the52nd
International Water Conference, Engineers Society of Western Pennsylvania, 1994.

38. Behavior of Ammonium Salts in Steam Cycles. EPRI, Palo Alto, CA: 1993. TR-102377.

39. D. A. Palmer, J. M. Simonson and J. P. Jensen, Measurement of the Volatilities of


Electrolytes: Application to Water/Steam Cycles, Fourth International Conference on
Fossil Plant Cycle Chemistry, EPRI, Palo Alto, CA: 1995. TR-104502.

40. Assessment of the Ray Diagram, EPRI, Palo Alto, CA: 1996. TR-106017.

41. Selection and Optimization of Boiler Water and Feedwater Treatments for Fossil Plants,
EPRI, Palo Alto, CA: 1997. TR-105040.

42. Vapor-Liquid Partitioning of Sulfuric Acid and Ammonium Sulfate, EPRI, Palo Alto, CA:
1999. TR-112359.

43. Volatility of Aqueous Acetic Acid, Formic Acid and Sodium Acetate, EPRI, Palo Alto, CA:
2000. TR-113089.

44. D. A. Palmer, S. L. Marshall, J. M. Simonson and M. S. Gruszkiewicz, The Partitioning of


Acetic, Formic, and Phosphoric Acids between Liquid Water and Steam, Steam Chemistry:
Interaction of Chemical Species with Water, Steam, and Materials during Evaporation,
Superheating, and Condensation, EPRI, Palo Alto, CA: 2000. TR-114837.

45. M. S. Gruszkiewicz, D. B. Joyce, S. L. Marshall, D. A. Palmer and J. M. Simonson, The


Partitioning of Acetate, Formate, and Phosphate around the Water/Steam Cycle, Sixth
International Conference on Fossil Plant Cycle Chemistry, EPRI, Palo Alto, CA: 2001.
1001363.

46. Cycling, Startup, Shutdown, and Layup Fossil Plant Cycle Chemistry Guidelines for
Operators and Chemists, EPRI, Palo Alto, CA: 1998. TR-107754.

47. Role of Chemical Additives in Transporting and Depositing Corrosive Impurities in Steam,
EPRI, Palo Alto, CA: 1984. NP-3487.

48. M. Ball and M. A. Jenkins, Steam Chemical Purity from CEGB Drum Boilers, Second
Cycle Chemistry Conference, EPRI, Palo Alto, CA: 1989. GS-6166.

5-66
Turbines

49. Cycle Chemistry Guidelines for Fossil Plants: Phosphate Treatment for Drum Units, EPRI,
Palo Alto, CA: 1994. TR-103665.

50. O. Jonas, R. Mathur and R. B. Dooley, EPRI and International Projects on Turbine Steam
Chemistry and Corrosion, Fourth International Conference on Fossil Plant Cycle
Chemistry, EPRI, Palo Alto, CA: 1995. TR-104502.

51. Cycle Chemistry Guidelines for Fossil Plants: All-Volatile Treatment: Revision 1, EPRI, Palo
Alto, CA: 2002. 1004187.

52. State-of-Knowledge on Deposition, Part 1: Parameters Influencing Deposition in Fossil


Units, EPRI, Palo Alto, CA: December 2002. 1004194.

5-67
6
CONCLUSIONS AND RECOMMENDATIONS

6.1 Conclusions

This report provides an update on deposition in the steam-water cycle at fossil plants and
supplements work that was initiated in 2002 and continued in 2003.(1,2) Three basic deposition
types and the related mechanisms and influencing factors applicable to each type were identified
in Reference 1 (See Section 1). (1)

Optimization of fossil plant cycle chemistry in accordance with the key cycle chemistry
guidelines indicated throughout this document should be a top priority since doing so will serve
to minimize the probability of significant deposition developing. Additional consideration
should be given, where needed and feasible, to upgrading the chemistry to either oxygenated
treatment (OT) or all volatile treatment (AVT), this based on the EPRI benchmarking activities
of the last few years that clearly indicate these treatments minimize corrosion and corrosion
product transport which in turn leads to minimized deposition activity. The units with the top
benchmarking scores (rated above average to excellent or World Class) almost without
exception employ these treatments in accordance with EPRI Guidelines and report high levels of
(3, 4)
unit availability and reliability. However, it is also recognized that many fossil plant units
will need to operate with chemistries that are not fully optimized and will not be able to upgrade
the chemistry in the near term. Further, cycle chemistry optimization in accordance with existing
guidelines would not entirely eliminate deposition activity in all areas of the cycles of every
fossil unit.

While the initial assessment demonstrated that deposition can and will occur in any cycle
component in contact with water and steam, it also confirmed the relative importance of
deposition in condensate and feedwater systems, boiler waterwalls, steam-cooled tubing and
steam turbines. Important considerations of industry needs for each area of the cycle are
delineated in the following report subsections.

Over 40 parameters and factors that influence deposition were identified in Part 1 of the state-of-
(1)
knowledge assessment. Although no additional parameters were identified under the Part 2
assessment presented herein, significant additional information and mathematical relationships
were reviewed and summarized for later use in modeling the deposition process. The highest
ranked parameters were found to be transient conditions (startup, shutdown, load changes, etc.);
concentrations of impurities (contaminants), oxidizing agents (oxygen) and particles; pH at
operating temperatures; mass transport; condensate polishing and filtration; chemical
treatment; temperature; boiling, condensing and flow regimes; agglomeration of particles;
stream quality (moisture); pressure; tube/surface and oxide/hydroxide compositions and

6-1
Conclusions and Recommendations

interactions; time; steam/water properties; heat flux; size and shape of particles; corrosion
rate; configuration of hardware and zeta potential.

Key parameters were selected because of their apparent, even if not fully understood, influence
on one or more of the three deposition process types. These key parameters are (a) dynamics of
deposition and release (as it affects both boilers and turbines), (b) heat flux (primarily in boiler
waterwall tubes), (c) mass transport, (d) solubility in steam and water and (e) surface finish
(primarily in turbines and steam generator water tubes). Additional parameters of interest were
also identified for possible further study.

Condensate and Feedwater Systems

The components considered in this review included the condenser, makeup water supply,
Condensate pumps, condensate polisher and/or condensate filter (if used), low pressure feedwater
heaters, boiler feed pump, high pressure feedwater heaters and interconnecting piping.

Sources

Potential sources of contamination that leads to deposition/corrosion air inleakage, condenser


leaks, improper chemistry control, impurities in makeup water, condensate polishers leakage,
impurities in treatment chemicals, combustion products, paints, preservatives, solvents and
ineffective chemical cleaning.

Deposition

The condensate and feedwater systems have few areas of deposition. Deposition can form in the
high-pressure feedwater heaters and, to a much lesser extent, on the steam side surfaces of the
condenser tubes. Deposits can form on either the waterside or steamside surfaces of condenser
tubes. Deposition on the waterside is the more serious problem.

Corrosion

These systems predominately function as generators of corrosion products which then migrate to
other parts of the heat cycle, namely the boiler, superheater, turbine and reheater, resulting in
serious deposition problems in those components.

Composition

Feedwater deposits have been comprised of metallic copper and metallic oxides of various types,
depending on the materials of construction of the condenser and feedwater heater tubes.
Deposits on the steam side surfaces of the condenser tubes have been identified as iron or iron
oxides and, occasionally, oil.

6-2
Conclusions and Recommendations

Impact

Corrosion in the condensate and feedwater systems represents the source of most of the material
deposited in other components. Reductions in corrosion in the former reduce deposit inventories
in the latter. Deposition on the waterside of the condenser can result in significant loss of
condenser efficiency and possible derating of the plant. Deposition of iron oxides on the
steamside surfaces of condenser tubes generally produces minimal impacts on efficiency.
Deposits in the condensate and feedwater systems can increase heater pressure drop requiring
additional power for the boiler feed pump to maintain the required flowrates. In addition,
deposit buildup impedes heat transfer, which may reduce final feedwater temperature.

Mechanisms:

Inadequate or inappropriate chemistry control, load cycling, contaminant ingress and air ingress
can contribute to several corrosion mechanisms that damage components in the condensate and
feedwater systems, sometime resulting in failure and/or the need for major repairs and extended
outages. The corrosion generates soluble and insoluble metals and metal oxides from the alloys
containing iron, copper, nickel, chromium, zinc, titanium and other metals. Also, these corrosion
products may as transportation and deposition sites for other impurities to the boiler, superheater,
reheater and turbines.

Control

Corrosion damage and transport of corrosion products to the boiler, superheater, reheater and
turbine can be minimized with adequate chemistry control, high quality water
purification/treatment equipment, tight condensers and proper material selection.

Deficiencies

While industry experience indicates some issues and concerns with deposition in this part of the
cycle, it is apparent that these incidents almost always relate to poor design, deficient operating
and maintenance practices or, most frequently, improper choices with respect to cycle chemistry.
Organizations with deposition problems in this part of the cycle are therefore strongly
encouraged to implement improvements based on existing guidelines publications.

Boilers

Deposit formation on boiler waterwall tube surfaces remains the most widespread issue of
concern to the industry and is likely to remain so until either chemistry is optimized in more
units or new approaches to controlling it are identified, developed, evaluated, verified and
implemented. Recently issued statistics on cost of corrosion in fossil plants in the United States
compiled and reported by EPRI indicate that boiler tube failures involving waterside/steamside
corrosion remain the number one corrosion item with annual costs in excess of $1.1 billion
(5)
according to 1998 figures. The total cost of corrosion damage involving deposition in the
steam-water cycle was nearly $1.9 billion, about 40% of the total cost of corrosion. While not all
tube failures and associated costs are due to failure mechanisms involving waterside deposits,

6-3
Conclusions and Recommendations

other EPRI statistics indicate that these mechanisms are still contributing to availability losses in
far too many fossil plant units.(6)

Existing EPRI guidelines define the role of waterwall deposits in initiation and propagation of
several boiler tube failure mechanisms and outline the role of chemical cleaning in their
prevention or correction. However, in view of recent industry experiences as embodied in the
latest statistics, it is clear that an improved understanding of boiler waterwall deposition that
would allow either better control or improved utilization of plant resources to either predict or
more readily identify deposition activity that endangers short term unit availability would
represent a significant practical improvement in the deposition state-of-knowledge.

Sources

Corrosion products from the pre-boiler areas are the source of most suspended solids that are
transported into the boiler. For all-ferrous feedwater systems (no copper alloys in the feedwater
and possibly copper-based condenser tubing), the generation of corrosion products (magnetite,
hematite and ferric oxide hydrate) occurs mainly due to corrosion and flow-accelerated corrosion
(FAC) of low-pressure and high-pressure feedwater heaters, deaerators, economizer inlet tubing
and piping, feedwater piping and drain lines. For mixed-metallurgy systems, the generation and
transport of corrosion products (cupric and cuprous oxide) is primarily a result of corrosion of
copper-bearing tubes in the low-pressure and/or high-pressure feedwater heaters.

Deposition

Deposits form in the areas of highest heat flux and temperature. Subsequent chemical reactions
with the base metal, other deposits and chemical treatment additives may result in tube damage.

Corrosion

In addition to corrosion products transported into the boiler from the pre-boiler section, hideout
of chemical contaminants dissolved in the boiler water can result in underdeposit corrosion
(UDC) and corrosion fatigue failures. In drum boilers operated with a phosphate treatment at
low sodium-to-phosphate molar ratios, severe hideout problems can occur. Hideout is a function
of chemical composition, pressure, temperature and design.

Composition

Over seventy-five different chemical compounds have been identified in boiler deposits. Many
of these contaminants enter the cycle with the makeup water or as a result of ingress of cooling
water. Deposits are comprised of metallic metal and metal oxides (primarily iron and copper
oxides) plus a combination of compounds that contain metals and constituents from makeup
water, cooling water and, in some cases, the water treatment chemical (for example, phosphate).

6-4
Conclusions and Recommendations

Impact

Deposition of chemicals, contaminants and corrosion products in the boiler may result in a
reduction in boiler efficiency and may reduce the service of boiler parts, and may finally lead to
boiler tube failures. Rippled deposits in economizers or waterwalls of once-through steam
generators cause a pressure drop increase. This increase deteriorates cycle efficiency, requiring
additional feed pump work, and full-load operation may no longer be possible. Buildup of
excessive deposits in waterwalls of supercritical once-through steam generators may cause or be
one of the major root causes of waterwall cracking. Thick and porous deposits in boiler tubes of
drum boilers are the primary root cause of under-deposit corrosion and may lead to boiler tube
failures due to hydrogen damage, caustic gouging and acid phosphate corrosion. Mechanical
carryover of sodium sulfate in steam during operation leading to the buildup of sodium sulfate
deposits in superheaters and reheaters may lead to or intensify pitting corrosion during shutdown
if inadequate or no layup methods are applied. Porous deposits of corrosion products can
develop and generate steam ("wick boiling"). Circulation and evaporation within the porous
deposit can lead to an additional concentration of boiler water .

Mechanisms

Deposition is closely related to heat flux and can significantly alter the local tube temperatures.
Heat flux is established by the furnace flame temperature and the bulk fluid temperature, in the
tube, which is fixed by boiler pressure. With internal deposit on the tube, an additional
temperature difference is required to drive the heat flux through the deposit so that inside and
average tube metal temperatures rise accordingly. Increases in the boiler tube temperature can
play a significant role in tube failures caused by under-deposit corrosion, as well as in boiler tube
failures involving locally accelerated creeping phenomena due to overheating. Deposition in
water tubes is highly dependent on numerous factors, including heat flux, circulation, boiler tube
temperatures, feedwater and boiler water purity and type of feedwater and boiler water treatment.
Heat flux, circulation and boiler tube temperatures are design parameters that may be altered
when the operating conditions are changed substantially from the original design. Such changes
may occur under one or more of the following conditions:
Coal firing is changed to oil or gas firing,
Coal with higher caloric value is used,
A different type burner is installed,
The burner angle is altered,
The boiler is fired harder because a feedwater heater is out of service or
A unit is changed to overpressure operation when the unit was not designed for such
operation.

The application of a comprehensive model would provide insight into current operating practices
and proposed changes in those practices. Also, some changes in current practices may be
suggested as a result of the modeling effort.

Control

6-5
Conclusions and Recommendations

The most important corrective actions in units suffering depositions on boiler tubes include (a)
selection of optimum feedwater treatment in once-through and drum boiler cycles and optimum
boiler water treatment in units with drum boilers, (b) estimation of the amount, composition and
porosity of boiler tube deposits, (c) implementation of chemical cleaning if the amount of
deposits exceeds a tolerable deposit weight limit and (d) corrections in feedwater treatment or
boiler water treatment based on analysis of boiler tube deposit composition.

Deficiencies

Application of state-of-the-art feedwater and boiler water treatments, as outlined in current EPRI
guidelines, creates conditions that minimize deposition and generally provide for long-term
operation of boilers without frequent chemical cleaning to remove those deposits. Although the
current guidelines are supported by research and operating experience, comprehensive model of
the deposition processes in critical areas is needed to further minimize deposition and to evaluate
changes in operating conditions. Additionally, an objective scientific-based evaluation of the
influence of organic chemical additives on corrosion and deposition processes is needed. Some
of these additives are organic and some are used to control deposition.

Superheaters and Reheaters

Sources
Potential sources of contamination that leads to deposition/corrosion contamination in feedwater
and steam, contamination in attemperation water, boiler mechanical and vaporous carryover,
contamination from layup practices and contamination from chemical cleaning practices.

Deposition

Gross deposition is not generally a concern in steam tubing. However, certain failure
mechanisms are influenced by chemistry and appear to involve deposition of constituents which
are soluble in superheated steam and also, in drum boiler units of constituents that are partitioned
into the vapor phase. Copper readily partitions into the vapor phase in drum boilers,
subsequently forms deposits in the primary superheater.

Corrosion

Contamination can lead to general corrosive attack, overheating (impaired heat transfer or flow
blockage by internal deposits) and stress corrosion cracking of austenitic stainless steels.
Pendant superheater and reheater tubes cannot be drained and, thus, contaminants that are
introduced during startup, operation, shutdown and layup may corrode superheater and reheater
tubing if condensate forms during shutdown, and especially if oxygen is subsequently introduced
during the outage. Additionally, the inner surfaces of the superheater and reheaters can oxidize
in the presence of steam to form oxide layers. In many cases, this steamside oxide periodically
spalls away from the metal surface in the form of flakes. This phenomenon is termed
exfoliation and is not directly related to aforementioned mechanical and vaporous carryover of
contaminants.

6-6
Conclusions and Recommendations

Composition

The composition of deposits formed by mechanical and vaporous carryover and attemperation
water will depend on the composition of contaminants entering the superheater and reheater and
their solubility in steam at superheater and reheater operating conditions. Contaminants can
include copper and copper oxides, sodium salts (such as sodium sulfate), silica and many others.
Copper is dissolved by the superheated steam only in those units which operate above the critical
pressure required to transport copper to the superheater, reheater and turbine. Deposits can be
comprised of metallic copper and metallic oxides of various types, depending on the materials of
construction of the condenser and feedwater heater tubes.

Impact

General corrosive attack, overheating (impaired heat transfer or flow blockage by internal
deposits) and stress corrosion cracking of austenitic stainless steels can lead to tube failures and
exfoliation of oxides can result in solid particle erosion of turbine. The exfoliated scale can be
transported at very high velocities into the steam turbines causing extensive solid particle erosion
(SPE) of both stationary and rotating blades in the front-end stages of the high- and intermediate-
pressure sections. This form of material damage has been recognized as a severe and expensive
problem for many utilities.

Mechanisms

Inadequate or inappropriate chemistry control in the feedwater and condensate systems and/or
the boiler can contribute to corrosion in those areas and deposition of the resultant corrosion
products in the boiler, superheater and reheater. Also, startup, shutdown, cycling operations,
improper drum level operating, contaminated attemperation water and contamination from
boiling on steam surfaces during startup can lead to corrosion damage in the superheater/reheater
areas. In drum boiler cycles, a major source of deposits in the superheater, and to a lesser extent,
in the reheater, results from both mechanical and vaporous carryover from the boiler.
Mechanical carryover will be present in all drum boilers. Vaporous carryover depends on the
chemical species involved, the feedwater and boiler water treatment and the boiler pressure and
temperature. Chemicals that are soluble in high pressure steam may deposit in the superheater,
reheater or turbines as the solubility changes with changes in temperature and pressure.
Mechanical carryover depends on the design, operation and maintenance of the boiler.

Control

Deposit minimization requires reducing the input of copper and sulfates. Deposits in
superheaters, reheaters and steam piping can be controlled by maintenance of proper water and
steam chemistry, maintenance of designed boiler scrubber efficiency (drum units), elimination of
high boiler water levels and foaming (drum units) and/or adoption of proper layup and chemical
cleaning procedures. Chemical cleaning is the only option for removal of deposits in these tubes.

6-7
Conclusions and Recommendations

Deficiencies

No specific state-of-knowledge deficiencies exist with respect to deposition in the superheaters


and reheaters. As previously discussed for condensate and feedwater, control of the chemistry in
accordance with existing EPRI Guidelines effectively minimizes the risks of steam
contamination, and subsequent deposition and corrosion in steam tubing. Overall experience in
the fossil plant industry is that availability losses due to mechanisms involving steam side
deposition are relatively minor. However, the potential for transport of oxides, possibly with
contaminants attached is very high and a better understanding of the parameters governing
attachment and release of these impurities is needed.

Steam Turbines

Deposition in turbines is a more serious problem, which can lead to corrosion damage and failure
(primarily in the low pressure section) or capacity losses and performance degradation (primarily
in the high and intermediate pressure sections). The costs of stress corrosion cracking and
corrosion fatigue corrosion damage were around $600 million in 1998, while costs of turbine
(6)
copper deposits were nearly $150 million.

Sources

The sources of contamination are generally the same as those for superheaters and reheaters.
Chemical contaminants are transported to the turbine as a result of mechanical carryover,
vaporous carryover and impure attemperation water.

Deposition

Deposits may form solid, crystalline or liquid (aqueous) film on the internal surfaces of the
turbine. The solubility of the chemical in steam and the concentration of the liquid deposit
depend on the chemical nature of the impurity and the temperature and pressure at each location
on the turbine surface. In some conditions both states can coexist for different chemicals.

Corrosion

The aqueous deposit can attack the metallic surfaces of stressed or susceptible materials.
Experience has shown that a concentration of corrosive impurities (chloride and sulfate ions) in
turbine deposits below 0.25% usually do not cause pitting and corrosion fatigue of blades.
Pitting corrosion is most often associated with chloride deposits. This contaminant is most
commonly found in systems employing brackish or salt water cooling. Stress corrosion cracking
is probably the most serious deposit related problem, because it can lead to long outages. The
otherwise innocuous iron oxides that deposit on the turbine can absorb salts, increasing the risk
of corrosion. A particular concern with salts is in the phase transition zone of the LP turbine,
where high concentrations and potentially more acidic or alkaline conditions can form.

6-8
Conclusions and Recommendations

Composition

Most deposits are salts and oxides, principally iron and copper oxides. Copper deposits on
turbines can affect the output capability of the turbine. Major corrosive deposits, identified in a
literature survey given in EPRI report NP-3002, were sodium hydroxide (NaOH), sodium
chloride (NaCl), hydrochloric acid (HCl), sodium silicate (Na2SiO3), sodium carbonate (Na2CO2),
sodium bicarbonate (NaHCO3), ammonium chloride (NH4Cl) and potassium chloride (KCl).
Sodium sulfate (Na2SO4), ammonium sulfate (NH4)2SO4, and ammonium hydrogen sulfate
(NH4HSO4) are sulfate compounds that may be present in turbine deposits. Sodium ions are
common contaminants and ammonium ions are present from water treatment chemicals. Sodium
hydroxide, sodium chloride and hydrochloric acid were considered the most undesirable because
these chemicals can promote stress corrosion cracking. Silica is a common component of turbine
deposits.

Impact

When a deposit forms in a turbine, it can cause a loss in thermal efficiency, changes in pressure
drop, erosion, and localized corrosion. Also, deposits can block the movement of the turbine
valves, leaving the turbine unprotected against potential overspeed.

Mechanisms

Deposition in turbines usually results from the presence of chemicals in the steam at
concentrations in excess of the solubility in steam. A chemical may be soluble in higher pressure
steam at the turbine inlet and become insoluble as the pressure and temperature decreases
through the turbine. The location that impurities deposit in the turbine and the nature of the
deposit (solid or liquid), depends on their nature, solubility in steam and operating conditions,
including temperature and pressure.

Control

Purity criteria for steam are based on the understanding of partitioning to saturated steam in
drum boilers and solubility in steam under all operating conditions.

Deficiencies

While the science of these phenomena are now understood and guidelines are available which, if
followed, greatly reduce the risk of turbine deposition and corrosion, the continued occurrence of
such problems within the industry and their associated costs indicate a clear need to better
understand the deposition processes involved to determine if there are other solutions that could
be practically implemented. Understanding of the factors that affect the dynamics of deposition
and the proportions of particles depositing and adhesion is required.

6-9
Conclusions and Recommendations

6.2 Recommendations

The state-of-knowledge improvement process should include deposition modeling activity in


addition to research into understanding the multitude of thermal, hydrodynamic, electrostatic and
other parameters that influence deposition. The need for modeling activity is greatest in boilers
due to the impact of boiler deposits on fossil unit availability and reliability. Experimental results
and model predictions must subsequently be verified through demonstration projects that permit
observation of conditions in working fossil plant units.

The goal of boiler deposition model development would be to create a tool that would apply to
new and existing fossil units to better manage any deposition activity that cannot be avoided and,
in so doing, eliminate or minimize the negative impacts of deposition on performance and
profitability.

The application of a comprehensive model would provide insight into current operating and
chemical practices and proposed changes in those practices. Also, some changes in current
practices may be suggested as a result of the modeling effort. In view of the needs in this area,
EPRI has begun to assess the feasibility of developing such a model.

The first step of the feasibility assessment is to examine the possibility of developing a
comprehensive model for the deposition processes in fossil plant cycles combining known
models relating deposition to heat transfer, mass transfer and steam solubility. The following
four general areas will be examined:
Heat transfer models,
Mass transport models,
Steam solubility models and
Synthetic corrosion products models.

Examination will consist of the following activities:


Identifying existing equations and models that address various aspects of the deposition
process,
Documenting input and output parameters and major assumptions and limitations of existing
models,
Identifying overlapping input/output parameters in existing models,
Identifying missing data that is basic to the success of the model and missing data that may
enhance the model (if no data exists for use in developing equations) and
Identifying portions of the selected models (theoretical or empirical relationships) for future
development.

While the science of these phenomena are now understood and guidelines are available which, if
followed, greatly reduce the risk of turbine deposition and corrosion, the continued occurrence of
such problems within the industry and their associated costs indicate a clear need to better
understand the deposition processes involved to determine if there are other solutions that could

6-10
Conclusions and Recommendations

be practically implemented. Specifically, additional knowledge is needed on (a) copper


deposition, (b) impact of chemical additives on deposition, (c) mechanical carryover, (d)
dynamics of deposition, and (e) transient conditions.

Although no significant knowledge deficiencies were identified with respect to deposition in


superheaters and reheaters, minimizing copper and sulfate transport is important to these
components as well as turbines.

Further minimization of corrosion product generation and transportation to the boiler would
reduce deposition in the boiler. Electrochemical experiments and assessment of commercially
available alternative chemical additives and/or development of alternative chemicals for deposit
minimization via reduction in corrosion product generation, transportation and adherence would
be beneficial.

Finally, the conditions that exist during plant transients including load changes, cycling
operation, startup, shutdown and outages can contribute significantly to the transportation and
deposition of soluble and insoluble chemicals throughout the plant. Modeling of data from
nuclear power plants has confirmed field data that show significant transportation and re-
deposition of soluble chemical impurities from steam generators during relatively small load
changes.

6.3 References

1. State-of-Knowledge on Deposition, Part I: Parameters Influencing Deposition in Fossil


Boilers, EPRI, Palo Alto, CA: December 2002. 1004194.

2. Deposition in Boilers: Review of Soviet and Russian Literature, EPRI, Palo Alto, CA:
August 2003. 104193.

3. R. B. Dooley. The Relationship Between Cycle Chemistry and Performance of Fossil


Plants, Power Plant Chemistry, 4(6), 2002.

4. Strategic Decision Analysis for Cycle Chemistry Improvements. EPRI, Palo Alto, CA. EPRI
1004641. December 2002.

5. Cost of Corrosion in Fossil Plants. EPRI, Palo Alto, CA. EPRI 1007274. September 2002.

6. Survey Results. International Conference on Boiler Tube Failures and HRSG Tube Failures
and Inspections, EPRI, Palo Alto, CA: 2002. 1007347.

6-11
Program: SINGLE USER LICENSE AGREEMENT
THIS IS A LEGALLY BINDING AGREEMENT BETWEEN YOU AND THE ELECTRIC POWER RESEARCH INSTI-
Boiler and Turbine Steam and Cycle Chemistry TUTE, INC. (EPRI). PLEASE READ IT CAREFULLY BEFORE REMOVING THE WRAPPING MATERIAL.
BY OPENING THIS SEALED PACKAGE YOU ARE AGREEING TO THE TERMS OF THIS AGREEMENT. IF YOU DO NOT AGREE TO
THE TERMS OF THIS AGREEMENT,PROMPTLY RETURN THE UNOPENED PACKAGE TO EPRI AND THE PURCHASE PRICE WILL
BE REFUNDED.
1. GRANT OF LICENSE
EPRI grants you the nonexclusive and nontransferable right during the term of this agreement to use this package only for your own
benefit and the benefit of your organization.This means that the following may use this package: (I) your company (at any site owned
or operated by your company); (II) its subsidiaries or other related entities; and (III) a consultant to your company or related entities,
if the consultant has entered into a contract agreeing not to disclose the package outside of its organization or to use the package for
its own benefit or the benefit of any party other than your company.
This shrink-wrap license agreement is subordinate to the terms of the Master Utility License Agreement between most U.S.EPRI mem-
ber utilities and EPRI.Any EPRI member utility that does not have a Master Utility License Agreement may get one on request.
2. COPYRIGHT
About EPRI This package, including the information contained in it, is either licensed to EPRI or owned by EPRI and is protected by United States
and international copyright laws.You may not, without the prior written permission of EPRI, reproduce, translate or modify this pack-
EPRI creates science and technology solutions for age, in any form, in whole or in part, or prepare any derivative work based on this package.
the global energy and energy services industry. U.S. 3. RESTRICTIONS
electric utilities established the Electric Power You may not rent, lease, license, disclose or give this package to any person or organization, or use the information contained in this
package, for the benefit of any third party or for any purpose other than as specified above unless such use is with the prior written
Research Institute in 1973 as a nonprofit research permission of EPRI.You agree to take all reasonable steps to prevent unauthorized disclosure or use of this package. Except as speci-
consortium for the benefit of utility members, their fied above, this agreement does not grant you any right to patents, copyrights, trade secrets, trade names, trademarks or any other
intellectual property, rights or licenses in respect of this package.
customers, and society. Now known simply as EPRI,
4.TERM AND TERMINATION
the company provides a wide range of innovative This license and this agreement are effective until terminated.You may terminate them at any time by destroying this package. EPRI has
products and services to more than 1000 energy- the right to terminate the license and this agreement immediately if you fail to comply with any term or condition of this agreement.
Upon any termination you may destroy this package, but all obligations of nondisclosure will remain in effect.
related organizations in 40 countries. EPRIs
5. DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES
multidisciplinary team of scientists and engineers
NEITHER EPRI,ANY MEMBER OF EPRI,ANY COSPONSOR, NOR ANY PERSON OR ORGANIZATION ACTING ON BEHALF
draws on a worldwide network of technical and OF ANY OF THEM:
business expertise to help solve todays toughest (A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I) WITH RESPECT TO THE USE
OF ANY INFORMATION,APPARATUS, METHOD, PROCESS OR SIMILAR ITEM DISCLOSED IN THIS PACKAGE, INCLUDING
energy and environmental problems.
MERCHANTABILITY AND FITNESS FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR
EPRI. Electrify the World INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTYS INTELLECTUAL PROPERTY, OR (III) THAT THIS
PACKAGE IS SUITABLE TO ANY PARTICULAR USERS CIRCUMSTANCE; OR
(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER (INCLUDING ANY CONSE-
QUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH
DAMAGES) RESULTING FROM YOUR SELECTION OR USE OF THIS PACKAGE OR ANY INFORMATION, APPARATUS,
METHOD, PROCESS OR SIMILAR ITEM DISCLOSED IN THIS PACKAGE.
6. EXPORT
The laws and regulations of the United States restrict the export and re-export of any portion of this package, and you agree not to
export or re-export this package or any related technical data in any form without the appropriate United States and foreign gov-
ernment approvals.
7. CHOICE OF LAW
This agreement will be governed by the laws of the State of California as applied to transactions taking place entirely in California
between California residents.
8. INTEGRATION
You have read and understand this agreement, and acknowledge that it is the final, complete and exclusive agreement between you
and EPRI concerning its subject matter, superseding any prior related understanding or agreement. No waiver, variation or different
terms of this agreement will be enforceable against EPRI unless EPRI gives its prior written consent, signed by an officer of EPRI.

2003 Electric Power Research Institute (EPRI), Inc. All rights


reserved. Electric Power Research Institute and EPRI are registered
service marks of the Electric Power Research Institute, Inc.
EPRI. ELECTRIFY THE WORLD is a service mark of the Electric
Power Research Institute, Inc.

Printed on recycled paper in the United States of America

1004930

EPRI 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10412, Palo Alto, California 94303 USA
800.313.3774 650.855.2121 askepri@epri.com www.epri.com

S-ar putea să vă placă și