Вы находитесь на странице: 1из 214

Revised Wet Stack Design Guide

2012 TECHNICAL REPORT

Revised Wet Stack Design


Guide

EPRI Project Manager


C. Dene

3420 Hillview Avenue


Palo Alto, CA 94304-1338
USA
PO Box 10412
Palo Alto, CA 94303-0813
USA
800.313.3774
650.855.2121
askepri@epri.com
www.epri.com

1026742
Final Report, December 2012

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.
REFERENCE HEREIN TO ANY SPECIFIC COMMERCIAL PRODUCT, PROCESS, OR SERVICE BY ITS TRADE
NAME, TRADEMARK, MANUFACTURER, OR OTHERWISE, DOES NOT NECESSARILY CONSTITUTE OR
IMPLY ITS ENDORSEMENT, RECOMMENDATION, OR FAVORING BY EPRI.
THE FOLLOWING ORGANIZATIONS, UNDER CONTRACT TO EPRI, PREPARED THIS REPORT:
Alden Research Laboratory, Inc.
Comit International des Chemines Industrielles

NOTE
For further information about EPRI, call the EPRI Customer Assistance Center at 800.313.3774 or
e-mail askepri@epri.com.
Electric Power Research Institute, EPRI, and TOGETHERSHAPING THE FUTURE OF ELECTRICITY are
registered service marks of the Electric Power Research Institute, Inc.
Copyright 2012 Electric Power Research Institute, Inc. All rights reserved.

Acknowledgments

The following organization, under contract to the Electric Power


Research Institute (EPRI), prepared this report:
Alden Research Laboratory, Inc.
30 Shrewsbury St.
Holden, MA 01520
Principal Investigators
D. Anderson
L. Maroti
CICIND
Comit International des Chemines Industrielles
(International Committee for Industrial Chimneys)
Talacker 50, CH-8001
Zurich, Switzerland
This report describes research sponsored by EPRI.
The authors would like to acknowledge the authors of the original
EPRI Wet Stacks Design Guide (EPRI TR-107099) upon which this
revised guide is based. The authors would also like to acknowledge
the program advisory committee for their guidance on the Revised
Wet Stack Design Guide project and for their comments on the draft
version of this guide. The committee members are:
Victor A. Bochicchio
Hamon Custodis, Inc.

Steven L. Reid, P. E.
Industrial Environmental Systems, Inc.

Prof. Gottfried Nonhoff


University of Aachen, Germany

John C. Sowizal, P. E.
Industrial Chimney Engineering Co., Inc.

This publication is a corporate


document that should be cited in the
literature in the following manner:
Revised Wet Stack Design Guide.
EPRI, Palo Alto, CA: 2012.
1026742.

iii

Product
Description

For the past 14 years, the design of wet stacks around the world has
been guided by the original EPRI Wet Stacks Design Guide (1996).
Since that time, the number of wet stack installations has grown
considerably, and a wealth of practical real-world operating and
maintenance experience has been obtained. The laws of physics have
not changed, and most of the information presented in 1996 is just as
valid today as it was when originally published. What has changed is
the power-generation industrys experience in using this information
and the day-to-day operation of wet stacks. Much had been learned
over the intervening years about the design and operation of wet
stack systems, and it had become clear that some updating of the
recommendations made in the original guide were needed. This
document, the Revised Wet Stack Design Guide, has been prepared to
present this updated information and to provide the powergeneration industry with the latest state-of-the-art information for
favorable wet stack design and operation. Much of this document
will be familiar to those who have read the original design guide.
Some sections of the original guide have been reused with only minor
changes; others have been significantly revised; and new sections
discussing the industrys experience with wet stack operation and
maintenance have been added. The outline of the guide has been
rearranged to be easier to use and follow, bringing the reader through
the entire wet stack design process, from the fundamentals of droplet
collection and liquid-film flow to the stacks final design,
commissioning, and operation. This new document strives to
thoughtfully update the original guide and to provide the industry
with the definitive reference needed by the engineers and designers
responsible for the specification, design, and implementation of
effective wet stacks.
Background
A wet stack is a chimney, stack, or flue that exhausts saturated flue
gas downstream from a wet-scrubbing process, such as a wet flue gas
desulfurization (WFGD) system. All recently designed and
constructed WFGD systems have installed wet stacks. Although the
technology is relatively mature, there are a number of technical issues
that utilities must address to achieve a successful installation. This
guide provides answers to these questions, whether the installation is
new or retrofit.

Objectives
To provide background information and updates of previously
published information

To summarize current state-of-the-art design

To list and discuss important parameters and options

To give specific recommendations for wet stack design

Approach
Investigators collected the information from a literature survey, the
in-house expertise of contractors, phone contacts with vendors, a
utility advisory committee, and a limited number of site visits. They
collated and summarized the information to produce the report,
which the advisory committee also reviewed.
Results
The information in the guide covers the design process and
operational issues for both new and retrofit wet stack installations.
Important issues addressed include system design for favorable wet
operation, stack liquid discharge, plume downwash, stack-liner
geometry, gas velocity in the liner, and liquid-collection devices and
drainage. In addition, the report also provides a guide to developing a
wet stack specification.
EPRI Perspective
Because most new FGD systems include wet stacks, it is imperative
that accurate, reliable information is available. This guide contains
the most up-to-date information, and it should be useful for
personnel responsible for wet stack design, specifications, or
operation. Care must be taken to use these recommendations with
good engineering judgment and consideration for site-specific
installations.
Keywords
Air-emissions control
Flue gas desulfurization (FGD)
Wet stacks
Wet scrubbers
SO2 control
Stack liquid discharge (SLD)

vi

Abstract
In 1996, the Electric Power Research Institute (EPRI), in a tailored
collaboration with New York State Electric & Gas (NYSEG),
retained Burns & McDonnell and DynaFlow Systems to prepare a
design guide for wet stacks. The purpose of this guide was to provide
the utility industry with information and recommendations
concerning the design and specification of wet stacks. Since that
time, the number of wet stack installations has grown considerably,
and a wealth of practical real-world operating and maintenance
experience has been obtained. This document, the Revised Wet Stack
Design Guide, has been prepared to present this updated information
and to provide the power-generation industry with the latest stateof-the-art information for favorable wet stack design and operation.
This new document strives to thoughtfully update the original guide
and to provide the industry with the definitive reference needed by
the engineers and designers responsible for the specification, design,
and implementation of effective wet stacks.

vii

Table of Contents
Section 1: Background and Objectives .....................1-1
1.1 Preface to the Revised Wet Stack Design Guide ............ 1-1
1.2 Introduction ............................................................... 1-1
1.3 Scenarios for Wet Stack Utilization .............................. 1-3
1.4 Important Considerations for Wet Stack Design and
Operation ....................................................................... 1-3
1.4.1 Mist-Eliminator Carryover ................................... 1-4
1.4.2 Deposition of Entrained Liquid Droplets ................ 1-4
1.4.3 Condensation .................................................... 1-5
1.4.4 Liquid Re-entrainment ......................................... 1-6
1.4.5 Washing of Wet Fans ........................................ 1-7
1.4.6 Stack Liquid Discharge (SLD) ............................... 1-7
1.4.7 Corrosion/Chemical Attack ................................ 1-8
1.4.8 Plume Downwash .............................................. 1-8
1.4.9 Icing Potential ................................................... 1-9
1.5 Contents of the Design Guide .................................... 1-11
1.5.1 Information About Key Issues ............................ 1-11
1.5.2 Knowledge of the State of the Art ...................... 1-11
1.5.3 Overview of Project Implementation ................... 1-12
Section 2: Wet Stack Design Fundamentals ..............2-1
2.1 Introduction ............................................................... 2-1
2.2 Droplet and Liquid-Flow Fundamentals .......................... 2-3
2.2.1 Flow of Droplets Suspended Within a Gas
Flow ........................................................................ 2-3
2.2.2 Liquid-Film Flow on Surfaces and Reentrainment Fundamentals ........................................... 2-4
2.2.3Gas-Flow in Wet Duct Systems ............................. 2-5
2.2.4Gas Flows in Wet Stack Systems .......................... 2-7
2.3 Sources of Liquid in a Duct/Wet Stack System ............ 2-12
2.3.1 Mist Eliminators and Absorbers ......................... 2-12
2.3.2 Liner Condensation .......................................... 2-20
2.4 Liquid-Film Behavior on the Liner Wall ........................ 2-22
2.5 Liner-Wall Discontinuities .......................................... 2-28
2.6 Re-entrainment ......................................................... 2-31
2.7 Recommended Liner-Gas Velocities ............................ 2-32
2.8 Stack Liquid Discharge ............................................. 2-33
ix

2.9 Plume Downwash and Icing ...................................... 2-35


2.9.1 Plume Downwash on Stack ............................ 2-35
2.9.2 Stack Top Icing ............................................ 2-40
Section 3: Practical Design Recommendations ..........3-1
3.1 Introduction ............................................................... 3-1
3.2 New Wet Stack Designs ............................................. 3-2
3.3 Retrofit Wet Stack Designs .......................................... 3-2
3.4 Wet Duct/Stack Operation ......................................... 3-3
3.5 Absorber Design and Placement .................................. 3-6
3.6 Wet Ductwork Design ................................................. 3-6
3.6.1 Guidelines for Selection of Geometry for Wet
Ducts ......................................................................... 3-8
3.7 Flow-Control Devices ................................................ 3-14
3.7.1 Flow-Control Liquid-collectors ............................ 3-14
3.8 Fans ....................................................................... 3-15
3.8.1 High-Gas-Temperature Fans .............................. 3-15
3.8.2 Wet Fans and Fan-Liquid Collectors ................... 3-15
3.9 Stack-Liner Design .................................................... 3-16
3.9.1 Liner Materials ................................................ 3-20
3.9.2 Recommended Liner-Gas Velocities .................... 3-33
3.9.3 Stack-Entrance Geometry .................................. 3-34
3.9.4 Liner-Diameter Changes.................................... 3-40
3.9.5 Internal Structures ............................................ 3-41
3.9.6 Stack-Exit Design ............................................. 3-43
3.9.7 Cyclonic Flow and Continuous-Emissions
Monitors .................................................................. 3-44
3.10 Miscellaneous Materials of Construction ................... 3-46
3.10.1 Exterior of Shell ............................................. 3-46
3.10.2 Exterior of Adjacent Stacks ............................. 3-46
3.10.3 Interior of Shell for Use with Brick Liners ........... 3-46
3.10.4 Liner Bands ................................................... 3-47
3.10.5 Rain Hoods and Roofs .................................... 3-47
3.10.6 Lightning-Protection System ............................. 3-47
3.10.7 Platforms and Ladders .................................... 3-48
3.10.8 Electrical Conduit .......................................... 3-48
3.11 Basic Liquid-Collector-System Design ........................ 3-48
3.11.1 Side-Wall Gutters .......................................... 3-50
3.11.2 Ceiling Collectors .......................................... 3-51
3.11.3 Floor Gutters ................................................. 3-53
3.11.4 Flow-Control Liquid Collectors ......................... 3-53
3.11.5 Internal Duct Supports and Expansion Joints...... 3-54
3.11.6 Stack-Entrance Collectors Side-Entry Breech.... 3-56

3.11.7 Stack-Entrance Collectors Bottom-Entry


Elbow ...................................................................... 3-58
3.11.8 Liner Collectors .............................................. 3-60
3.11.9 Sloped Liner Floor .......................................... 3-63
3.11.10 Stack-Outlet Collectors ................................. 3-63
3.11.11 Drains ........................................................ 3-65
3.11.12 Post-Installation Inspections ........................... 3-67
3.12 Laboratory Flow-Modeling....................................... 3-67
3.12.1 Computer Modeling. ...................................... 3-69
3.13 Plume Downwash and Icing .................................... 3-71
3.13.1 Downwash Modeling ..................................... 3-71
3.13.2 Windscreen Design ....................................... 3-72
3.13.3 Single versus Multiple Liners ............................ 3-72
3.13.4 Methods of Downwash Minimization ............... 3-73
3.13.5 Stack-Top Icing .............................................. 3-74
Section 4: Guide to Developing a Specification for
Wet Stacks ..............................................4-1
4.1 Specification Overview ............................................... 4-1
4.1.1 Phase I Feasibility Study ................................... 4-1
4.1.2 Phase II Design Process.................................... 4-3
4.2 Wet Stack Bid-Preparation Process ............................. 4-13
4.2.1 Establish Design Criteria ................................... 4-13
4.2.2 Define Absorber Outlet-Duct Geometry ............... 4-14
4.2.3 Define Chimney/Liner Geometry ....................... 4-14
4.2.4 Perform Model Study........................................ 4-14
4.2.5 Determine Liquid-Collection Devices ................... 4-14
4.2.6 Prepare Bid Specifications ................................ 4-15
4.2.7 Prepare Bid Drawings ...................................... 4-15
4.3 What to Specify in a Bid Document............................ 4-16
4.3.1 General Requirements ...................................... 4-16
4.3.2 Site Work ....................................................... 4-19
4.3.3 Concrete......................................................... 4-19
4.3.4 Liner ............................................................... 4-20
4.3.5 Ductwork and Expansion Joints ......................... 4-21
4.3.6 Miscellaneous Metals ....................................... 4-21
4.3.7 Platforms and Ladders ...................................... 4-21
4.3.8 Access Doors .................................................. 4-22
4.3.9 Protective Coatings .......................................... 4-22
4.3.10 Special Construction (Tuned Mass-Damping
System) .................................................................... 4-23
4.3.11 Personnel Elevator ......................................... 4-23
4.3.12 Mechanical ................................................... 4-24
4.3.13 Electrical....................................................... 4-24
xi

4.4 What to Ask For in a Bid Document ........................... 4-25


4.4.1 Warranties/Guarantees ................................... 4-25
4.4.2 Design Calculations ......................................... 4-25
4.4.3 Materials Testing ............................................. 4-27
4.4.4 Construction Procedures ................................... 4-28
4.4.5 On-Site Testing and Inspection .......................... 4-30
Section 5: Wet Stack Experience ..............................5-1
5.1 Experience with Wet Stacks ........................................ 5-1
5.1.1 General Operational Experience ......................... 5-1
5.1.2 International Experience with Wet Stacks ............. 5-3
5.2 Before Unit Startup ..................................................... 5-4
5.2.1 Review of Fabrication/Installation Drawings ......... 5-4
5.2.2 Field-Installation Inspection.................................. 5-5
5.3 After Unit Startup ....................................................... 5-5
5.3.1 Liquid-Drainage Monitoring ................................ 5-6
5.4 What to do if SLD Occurs ........................................... 5-6
5.4.1 Preliminary Stack Droplet Testing ......................... 5-7
Section 6: References ..............................................6-1
Appendix A: Glossary.......................................... A-1
A.1 Definitions................................................................ A-1
A.2 Units and Conversion Factors ..................................... A-8

xii

List of Figures
Figure 2-1 Liner Inlet Gas-flow Patterns Close-Coupled
Absorber......................................................................... 2-7
Figure 2-2 Typical Stack-Inlet Arrangements .............................. 2-8
Figure 2-3 Gas-flow Patterns Through a Standard 90 Elbow ...... 2-9
Figure 2-4 Liner-Inlet Gas-flow Patterns Side-entry Breech ....... 2-10
Figure 2-5 Liner-Inlet Gas-flow Patterns 3-Miter Cut BottomEntry Elbow ................................................................... 2-11
Figure 2-6 Common Mist-Eliminator Arrangements................... 2-14
Figure 2-7 Comparison of Gas-Path Lengths Through a Mist
Eliminator as a Function of Angle ..................................... 2-15
Figure 2-8 Typical Chevron-Type Mist-eliminator Pressure
Drop vs. Gas-Velocity Curve ............................................ 2-17
Figure 2-9 Typical Chevron-Type Mist-Eliminator Carryover
vs. Gas-Velocity Curve .................................................... 2-17
Figure 2-10 Typical Mist-Eliminator Collection Efficiency vs.
Droplet-Size Curve ......................................................... 2-18
Figure 2-11 Lower Liner Droplet- and Liquid-Film Flow
Patterns Side-Entry Breech............................................. 2-23
Figure 2-12 Liner-Floor Liquid-Film Flow Patterns Side-Entry
Breech .......................................................................... 2-24
Figure 2-13 Lower-Liner Droplet- and Liquid-Film Flow
Patterns Bottom-Entry Elbow .......................................... 2-25
Figure 2-14 Wetting vs. Nonwetting Surfaces ......................... 2-26
Figure 2-15 Liquid-Film Flow Patterns on a Smooth, Vertical
Wetting Surface vs. Vertical Gas Velocity ......................... 2-28
Figure 2-16 Liquid Flow over 1/8-in (3.1-mm) Weld Beads ...... 2-29
Figure 2-17 FRP Liner-Can Joints Internal-Taper
Requirement .................................................................. 2-30
Figure 2-18 Plume Downwash............................................... 2-36
xiii

Figure 3-1 Good and Poor Internal Duct-Truss Arrangements


for Wet Operation ......................................................... 3-12
Figure 3-2 Internal Duct-Truss Gusset Plate .............................. 3-13
Figure 3-3 Chimney with Constant-Diameter Brick Liner ............ 3-17
Figure 3-4 Chimney with Fiberglass-Reinforced Plastic Liner ...... 3-18
Figure 3-5 Chimney with Metal Liner Typical for Alloy,
Glass-Block, or Protective-Coating Liner System .................. 3-19
Figure 3-6 Borosilicate BlockLined Upper Wind Screen .......... 3-30
Figure 3-7 Sloped LinerFloor Arrangements ........................... 3-37
Figure 3-8 Miter-Cut Bottom-Entry Elbow Arrangements ............ 3-38
Figure 3-9 Side-Wall and Round-Duct Liquid-Collection
Gutters .......................................................................... 3-51
Figure 3-10 Ceiling Liquid-Collection Gutters .......................... 3-52
Figure 3-11 Horizontal Brace with Liquid-Re-entrainment
Prevention Disk .............................................................. 3-55
Figure 3-12 Upper and Lower Stack-liner Aerodynamic
Zones ........................................................................... 3-56
Figure 3-13 Typical Side-Entry Breech Liquid-Collection
System .......................................................................... 3-59
Figure 3-14 Side-Entry Breech Wing-Collector Design .............. 3-59
Figure 3-15 Bottom-Entry Elbow Liquid-Collection System
with Liner-Expansion Joint ................................................ 3-60
Figure 3-16 Liner-Expansion-Joint Placement and
Incorporation Within Ring Collectors ................................ 3-61
Figure 3-17 Liner-Expansion-Joint Liquid-Collector Concept ....... 3-61
Figure 3-18 Liner Rear-Wall V Diverter ................................ 3-62
Figure 3-19 Stack-Outlet Liquid Collector ................................ 3-64
Figure 3-20 Alternate Liner-Outlet Liquid-Collector Detail .......... 3-65
Figure 3-21 Typical Wet Stack Physical-Flow Models ............... 3-69
Figure 3-22 Typical Plume-Downwash Study CFD Model
Results .......................................................................... 3-71
Figure 3-23 Recommended Alignment of a Stack with Two
Flues ............................................................................. 3-73
Figure 5-1 Droplet Probe ........................................................ 5-8
xiv

Figure 5-2 Droplet-Probe Orientation ....................................... 5-8


Figure 5-3 Typical Droplet-Probe Test Results ............................. 5-9

xv

List of Tables
Table 2-1 Recommended Stack-Liner Velocities for Wet
Operation ..................................................................... 2-33
Table 3-1 Estimated Ranges of Liquid Flows in the Wet
Duct/Stack of a Typical 550-MW Unit ............................... 3-5
Table 3-2 Liner Material of Construction ................................. 3-21
Table 3-3 Physical Properties and Chemical Requirements
Of Acid-Resistant Brick1 .................................................. 3-22
Table 3-4 Recommended Stack Velocity Range for StacklinerDiameter Sizing ..................................................... 3-34
Table 4-1 Phase II Wet Stack Design Process ............................ 4-4
Table 4-2 Estimated Ranges of Flows in the Wet Duct/Stack
of a Typical 550-MW Plant ............................................... 4-8

xvii

Section 1: Background and Objectives


1.1 Preface to the Revised Wet Stack Design Guide
For the past 14 years, the design of wet stacks around the world has been guided
by the original EPRI Wet Stacks Design Guide (1996) [2]. Since that time, the
number of wet stack installations has grown considerably, and a wealth of
practical real-world operating and maintenance experience has been obtained.
The laws of physics have not changed, and most of the information presented in
1996 is just as valid today as it was when originally published. What has changed
is the power-generation industrys experience in using this information and the
day-to-day operation of wet stacks. Much had been learned over the intervening
years about the design and operation of wet stack systems, and it had become
clear that some updating of the recommendations made in the original guide
were needed. This document, the Revised Wet Stack Design Guide, has been
prepared to present this updated information and to provide the powergeneration industry with the latest state-of-the-art information for favorable wet
stack design and operation. Much of this document will be familiar to those who
have read the original design guide. Some sections of the original guide have
been reused with only minor changes; others have been significantly revised; and
new sections discussing the industrys experience with wet stack operation and
maintenance have been added. The outline of the guide has been rearranged to
be easier to use and follow, bringing the reader through the entire wet stack
design process, from the fundamentals of droplet collection and liquid-film flow
to the ducts and stacks final design, commissioning, and operation. This new
document strives to thoughtfully update the original guide and to provide the
industry with the definitive reference needed by the engineers and designers
responsible for the specification, design, and implementation of effective wet
stacks.
1.2 Introduction
A "wet stack" is a chimney, stack, or flue that exhausts saturated, scrubbed flue
gas. A wet stack is located downstream from a wet flue gas desulfurization
(WFGD) system. These systems spray slurry (typically limestone-based) into the
gas stream, which reduces the sulfur dioxide (SO2) content, saturates the flue gas
with water vapor, and reduces the temperature of the flue gas to 115130F (46
54.4C) for bituminous and hard coals and to 136145F (5763C) for lignite
and sub-bituminous coals.

1-1

The 1990 Clean Air Act Amendments (CAAA) required utilities to reduce
emissions of SO2. Many utilities have added WFGD systems to comply with the
Phase I requirements, whereas a lesser number have implemented dry-scrubber
technologies. Phase II of the program went into effect in 2000, further reducing
SO2 emissions and increasing the number of plants affected. As part of their
Phase II Compliance Plan, many plants with partial flue gas bypass systems have
decided to eliminate the bypass and scrub 100 percent of their flue gas. Many
other utilities have installed new or retrofit FGD systems, which typically use wet
stack operation because of reduced operating and maintenance costs.
The design of ducts and stacks for wet operation must address several issues that
were not present in unscrubbed or reheated gas stack designs. Some of the
important issues to consider in the design of a wet duct/stack system include

Stack liquid discharge (SLD)

Plume downwash and icing

Corrosion/chemical attack

Stack height

Absorber-outlet duct arrangement/geometry

Liner-breach geometry

Stack-liner geometry and material of construction

Gas velocity in the liner

Liquid-collection devices and drainage

The purpose of this revised guide is to provide the utility industry with
information and recommendations concerning the specification and design of
wet stacks. However, these recommendations should not be used without
applying good engineering judgment and consideration for site specifics.
Operating conditions, design conditions, and economics all play important roles.
The goals of the Revised Wet Stack Design Guide are to

Provide information for the fundamental understanding of wet stack


operation.

Update previously published information and recommendations.

Present the current state-of-the-art wet stack design.

Identify the parameters and options that will lead to favorable wet operation.

Give specific recommendations regarding wet stack design for new and
retrofit installations.

Present information related to wet stack startup and maintenance issues.

It is assumed that those who will use this guide have a general familiarity with
stack designs and FGD systems. This guide is intended for personnel who are
considering wet stacks and/or are responsible for designs, specifications, or
operation of wet stacks.
1-2

Section 2 of this guide provides fundamental information on the physical


processes and phenomena important to wet stacks and discusses the basic
objectives for favorable wet stack operation. Section 3 addresses new or retrofit
wet stack design and discusses the issues unique to wet stack operation. A guide
to developing specifications for wet stacks is presented in Section 4.
1.3 Scenarios for Wet Stack Utilization
There are two primary scenarios for wet stack utilization:

A new FGD system incorporating a new wet stack

A new FGD system that will convert an existing dry stack to wet operation

The first scenario involves adding a new FGD system that incorporates a wet
scrubber and a new wet stack. In such a scenario, the designer is given the most
freedom regarding plant layout, liner material, and stack-geometry selections. A
state-of-the-art wet stack design should be used that is based on the latest
analytical, experimental, and field data available. The goal is to produce an
effective wet stack design that minimizes both installation and operating costs.
The second scenario involves converting an existing dry stack to wet operation.
In such a scenario, the most important design consideration is whether or not the
existing liner and ductwork material can withstand the reduced temperatures and
wet conditions typical of wet stack operation and whether the liner-gas velocity
will result in droplet re-entrainment from the stack walls. If not, the existing
chimney liner must be modified or a new stack installed. Because of the outage
time required to modify an existing stack, it is often more economical to build a
new stack for scrubbed gas and use the existing stack for emergency gas bypass
around the scrubber as needed.
1.4 Important Considerations for Wet Stack Design and
Operation
The liquid inside a stack with a WFGD system exists as droplets from misteliminator carryover and as the moisture content of the gas flow entering the
stack liner. The vapor content is usually the maximum that can occur when the
flue gas is saturated with water vapor at the operating temperature. If there is an
induced-draft (ID) fan between the absorber and the stack, the fan-temperature
rise could result in a lower-than-saturation vapor content. What happens to the
vapor and liquid content before the gas reaches the top of the liner defines how
much liquid there is and in what form it is discharged from the stack. The major
gas/liquid flow processes that may lead to the SLD are described in the next four
subsectionsincluding the major features of the SLD. A description of the
liquid-flow balance in wet stacks is provided in [1], pp. 4-3 to 4-5. Liquid-flow
rates typical for wet ducts and stacks are given in Section 3.4, Wet Duct
Operation, Table 3-1.

1-3

1.4.1 Mist-Eliminator Carryover


The gas stream leaving the mist eliminator is saturated with water vapor, and
entrained fine liquid droplets are carried by the gas flow under normal design and
operating conditions. It is important that the mist eliminators operate properly at
the designed gas velocity and liquid load to minimize the size and rate of droplet
carryover. The mist eliminators are designed to reduce carryover of liquid and
slurry to the downstream ductwork. Mist-eliminator problems can result in
increased liquid carryover. Some larger droplets may also be present in the gas
stream, which have been re-entrained from the mist-eliminator blades, the misteliminator support structure, and areas of the mist eliminator containing solids
buildup. During the wash cycles, washing the mist eliminator can also increase
the amount of liquid carried with the gas flow downstream of the mist eliminator
because of the increased localized liquid loading associated with the wash sprays.
A field evaluation of the mist-eliminator performance can be made by
performing visual inspections and/or droplet sampling at the discharge face of the
mist eliminator. If carryover problems are noticeable, SLD will probably occur if
effective liquid collectors are not installed in the duct and the stack. Misteliminator selection, operation performance, and possible problems are described
in [3].
1.4.2 Deposition of Entrained Liquid Droplets
As droplets flow with the flue gas, they will impinge upon internal surfaces
because of droplet-trajectory paths controlled by the earth's gravity, centrifugal
forces on the droplets when the gas flow turns, and gas-drag forces. Large
droplets (3006000 m) from the gas-flow path deposit on duct and liner
surfaces most readily; medium droplets (100300 m) deposit only partially; and
small droplets (10100 m) deposit hardly at all. Fine droplets (<10 m) follow
the gas flow and do not deposit. The higher the gas velocity and the sharper the
turn of the flue gas path, the more likely it will be that significant numbers of
droplets will impinge upon walls, turning vanes, baffles, and internal duct
components (bracing, gusset plates, and expansion joints).
Small- to medium-sized droplets can also deposit on surfaces as a result of gasflow patterns in separation regions and turbulent deposition. The turbulent
deposition of fine droplets generated as a result of adiabatic condensation (caused
by pressure reduction with elevation change) along the height of the stack is
small. The gas velocity, duct geometry, stack geometry, and droplet size govern
the amount and location of liquid deposition. The major deposition area in the
stack is on the lower liner wall opposite the breeching opening. If the liner
incorporates a bottom-entry elbow, droplets will deposit along the elbows outer
radius.
Liquid deposition can be promoted, and the resulting liquid film collected, by
properly designed vanes, baffles, and liquid-collection devices installed in the
duct system. Maximum deposition and collection of the liquid on the duct walls
and stack liners must be achieved in order to control SLD. One of the main
1-4

objectives of physical-flow modeling is to design flow-control devices and to


develop liquid collectors such as ceiling gutters, side-wall channels, floor guides,
ring collectors, and drains to collect deposited liquid before re-entrainment can
take place. This process is discussed in Section 3.12, "Laboratory Flow
Modeling."
Another liquid-deposition area is on the choke surface in stacks that have chokes.
Because of the rapid change in flow direction created by the choke, some of the
fine droplets entrained in the gas flow will be deposited on the chokes inner
surface. The collected liquid will lead to SLD if the local gas velocities in the
choke are high enough to drag the liquid up to the stack outlet and no liquid
collector is installed.
1.4.3 Condensation
One of the sources of liquid in the wet duct/stack system is condensation of
water vapor from the saturated gas flow. Two types of condensation take place in
wet stacks.
Adiabatic Condensation (or bulk condensation) occurs in the bulk of the gas as
the pressure of the saturated gas flow decreases as a result of pressure loss and
elevation change along the height of the stack. The decrease in absolute pressure
of the saturated gas causes water vapor to condense on fine droplets or on small
solid particles in the bulk of the gas. Droplets formed by this type of
heterogeneous condensation will tend to be very small (<1 m) and to be located
throughout the volume of the gas flow. Only a very small fraction of this
adiabatic condensation reaches the liner surface by the turbulent-deposition
process. The rest is discharged from the top of the stack, which creates the white
plume typical of wet stacks, and which evaporates as the plume mixes with the
outside air.
Thermal Condensation (or wall condensation) caused by heat transfer is
important in wet stacks. Because the saturated flue gas temperature is usually
115145F (4663C), heat transfer through the liner, the annulus, and the outer
concrete shell to the cooler surrounding ambient air causes water vapor to
condense on the inside of the stack liner. The rate of condensation caused by heat
transfer is a function of the liner and shell construction, the type and thickness of
liner insulation, the gas-flow conditions, and the atmospheric conditions
(ambient temperature, wind velocity, and barometric pressure).
Annulus pressurization of brick-lined stacks using ambient air increases the rate
of condensation on the liner surfaces. Leakage of pressurizing air through the
liner in older brick stacks can significantly increase the condensation. Insulation
of alloy and FRP liners can significantly reduce the amount of thermal
condensation within the liner.

1-5

1.4.4 Liquid Re-entrainment


The liquid on the liner surface produced by deposition and condensation flows in
the form of film or rivulets is governed by gravitational, surface-tension, and gasshear forces. If high enough, the gas-shear force can shear the liquid off the
surface, causing re-entrainment. This is the most frequent source of SLD and the
fallout of liquid droplets in the vicinity of the stack.
Re-entrainment most frequently occurs because of one or more of the following:

High gas-velocity shear near rough surfaces with liquid deposits.

Instability of the liquid film (waves and ripples) on vertical stack-liner


surfaces. This instability occurs because, as gravity is pulling the liquid film
or droplet down, the gas-drag forces are pulling the liquid film or droplet
upward. This instability is a function of both the gas-shear force and the
local liquid-film thickness.

Surface discontinuities and protrusions, such as weld seams, fiberglassreinforced plastic (FRP) joints, and mortar joints in brick liners that disrupt
gas and liquid flow locally.

Vanes and baffles that cause gas-flow separation and recirculating liquid.

Gas-flow patterns that drag liquid along the surface to re-entrainment sites.

Strong vortex patterns that can pick water up in the core of the vortex from
horizontal or vertical surfaces.

Re-entrainment takes place at locations such as

Duct walls, roofs, floors, and stack-liner surfaces

Dampers, damper-blade guides, and trusses

Thermal-expansion joints

Trailing edges of turning vanes

Duct-junction corners and edges

Breeching-duct area changes

Stack floor

Stack-liner surfaces, particularly in the entrance region and at the


intersection with the breeching duct

Locations of internal solids buildup or scaling on surfaces

Stack-liner contraction or expansion sections

Stack choke surfaces

Liquid re-entrainment at these locations will be in the form of large droplets


(3006000 m); they will be discharged at the top of the stack, if not prevented
by a combination of favorable gas-flow path design and an optimized liquid 1-6

collector system. Droplets over 100 m do not evaporate outside of the stack, and
they will probably reach the ground level.
1.4.5 Washing of Wet Fans
Occasionally, ID fans will be located between the absorbers and the stack of a
utility power plant when retrofitted with an FGD system. These fans usually
require periodic or continuous washing to prevent solids buildup on the fan
impellers, which can cause fan-rotor imbalance. All of the liquid sprayed into the
fan inlet leaves the fan impeller as droplets. Most of the liquid will be propelled
by the high centrifugal-force field to the fan scroll, where it will deposit. Most of
this liquid will re-entrain back into the gas flow unless a properly designed highvelocity scroll-liquid collector is installed. Preferably, fans should be located
upstream of the absorber to avoid these problems. Alternatively, they should be
located such that there are a sufficient number of duct turns and/or duct lengths
between the fan outlet and the stack inlet to provide opportunities for liquid
collection.
1.4.6 Stack Liquid Discharge (SLD)
The stack liquid discharge (SLD) is also known as rainout or acid-mist fallout. It
is important to note that all wet stacks have some amount of SLD. However,
SLD is only a problem if the droplets are large enough to be detectable at ground
level near the stack.
The amount of liquid discharged at the top of the stack is a result of gas- and
liquid-flow processes that take place in the ductwork and stack system between
the discharge of the absorbers and the top of the stack liner. The source of the
liquid discharge can be categorized by the liquid-flow process and by the droplet
sizes as follows:

Liquid droplets carried from the absorber to the top of the stack liner by the
gas stream without depositing along the gas-flow path. Typically, these are
droplets less than 50 m in diameter, the majority of which can travel
through the ducts without significant deposition. Droplets of this size, when
ejected from the stack, will typically evaporate before reaching the ground.

Droplets re-entrained from liquid deposition and condensation on duct


surfaces that can be discharged. These tend to be larger dropsgreater than
300 m in diameter and are usually the major contributors to the SLD.

Droplets formed by adiabatic or bulk condensation. These represent a large


liquid-flow rate but are very fine droplets (<1 m), which have a negligible
effect on the SLD detectable at ground level. This condensation forms the
visible white plume typical of wet stack operation.

Liquid deposited and condensed on the liner, which flows either upward or
downward, depending on the local gas velocity. Liquid flows upward in those
areas of the stack where the gas velocity is higher than the flow-reversal
velocity for the specific liner material used. This phenomenon can be a
significant problem in tapered liners, in which the upper portion of the liner
1-7

can experience high gas velocities. The droplet sizes re-entrained at the top
of the liner, where velocities are highest, can be quite large (3006000 m),
depending on the liner-top geometry. The upper section of a stack choke is
usually in this mode of upward liquid flow.
The quantity and location of the fallout at a given plant is a function of the
droplet-size distribution of the discharged liquid and the atmospheric
conditionssuch as ambient temperature, wind, relative humidity, and
turbulence level. The very small droplets of the bulk condensation make the gas
plume white and visible. The plume is cooled as it mixes with the surrounding
air. This cooling results in condensation on the small droplets in the plume. The
increase in diameter is only 1 to 10 m; therefore, the droplets of the bulk
condensation can evaporate before reaching the ground and cannot be detected.
The mixing with the drier air away from the stack causes droplet evaporation.
The balance between the cooling and mixing processes defines the length of the
visible plume. Therefore, it is a function of the ambient air temperature and the
air relative humidity.
One of the major objectives in developing an effective wet stack design is to limit
SLD to a minimal acceptable level. The goal is to limit the droplet size that will
exit the stack. If the droplets are small enough, they will evaporate before hitting
the ground. However, if the droplets are large, they will land on nearby surfaces
such as plant structures, equipment, and cars. The wet stack survey indicates that
the fallout of liquid droplets, if noticeable, usually occurs on the downwind side
of the stack within a half-mile radius.
1.4.7 Corrosion/Chemical Attack
With wet stack operation, corrosion is less severe than with reheat, but it still
must be considered. Corrosion can be resisted by means of proper material
selection. The floors of the ductwork should be sloped to provide proper drainage
of the corrosive liquid after shutdown of the unit. The exposed materials of the
ductwork and the stack liner must be resistant to chemical attack. Several types of
material are available for both the scrubber-outlet duct and the stack liner. The
materials should be selected according to their corrosion resistance to the
chemical properties of the flue gas downstream of the absorber. Refer to Section
3.9.1, Materials of Construction, for more information.
1.4.8 Plume Downwash
Designing for a low gas velocity at the stack exit can decrease the amount of
SLD. However, this type of design increases the potential for plume downwash
at elevated wind velocities. During downwash episodes, the reduced staticpressure zone generated along the downwind side of the stack shell and liner
extensions can draw the flue gas into a downwash pattern along the stack shell.
In this process, saturated flue gas comes in contact with the liner extension, stack
hood, and stack shell. This contact can lead to deterioration of the stack
construction materials because of exposure to acid in the flue gas. It also increases
the potential for ice formation on the top of the stack. A cross wind at the top of
1-8

the stack will deflect the plume from its vertical path. As the ratio of vertical
plume momentum to horizontal wind momentum [(V2 ) FLUEGAS / (V2)WIND]
falls below a value of about 2.0 for a single-flue free-standing stack, the plume
may become partially entrained in the low-pressure zone that is formed on the
downwind side of the stack. This phenomenon is known as plume downwash.
To reduce the frequency of downwash, the height of the liner(s) can be increased
above the shell, or the momentum ratio can be increasedeither by reducing the
liner diameter or by installing a choke at the top of the liner.
For a multiple-flue stack, the equivalent momentum ratios for initiation of plume
downwash are higher, and they vary significantly with the wind direction relative
to the axis of the liners. The extent of downwash can also be more pronounced,
because of the larger size of the shell needed to enclose the multiple liners.
The interaction of a plume with a cross wind is a function of the following
parameters:

Liner-extension diameter

Shell diameter

Liner-extension height

Shape of hood/cap

Roof geometry

Number of liners

Wind direction, velocity, and air density at the top of the stack

The annual frequency of wind occurrence as a function of wind velocity

Liner exit-gas velocity and gas density

Plume buoyancy due to flue gas temperature

The most important fluid-dynamic parameter is the momentum ratio between


the stack-exit flow and the cross wind. This ratio defines the relative influence of
each flow on the other and is determined by the densities and velocities of the
flue gas and the wind. Plume downwash is most likely to occur during reduced
unit load operation, in which stack discharge velocities are reduced, and during
high wind conditions. Plume downwash is least likely to occur under the opposite
conditions of high unit load operation and low wind velocity. For a multiple-flue
stack, the liner-extension height above the shell is also very important for
minimizing plume downwash, as is the axis of the liners relative to the prevailing
wind direction. The selection of the stack-top geometry and stack-discharge
velocity for good plume escape must also consider the stack-discharge velocity for
plume dispersion and stack-liner velocity.
1.4.9 Icing Potential
Stack-top icing can occur at below freezing conditions all winter long, every
winter, and it creates a potential danger to people and property around the base
1-9

of the stack. Whether ice forms on the top of the stack depends on the
temperature of the stack surface, the temperature of the mixture of saturated flue
gas and cold ambient wind, and on whether water vapor will condense out of the
mixture. Ice formation is most likely at plants where below-freezing temperatures
are common and where they last for extended periods of time.
The locations for potential ice formation are discussed below.
The hood for a
single-flue stack

The stack hood is the first element exposed to plume downwash. A


hood for a single-flue stack is usually connected to the stack, liner,
and liner extension, which helps to keep it warm (particularly if
uninsulated). When ice forms on an outward-sloping hood, it
usually forms off the upwind edge and/or in the downwind half, at
two locations about 45 off the axis of the wind direction. These
ice buildups can be significant and can cause damage when they
fall to the ground. Stacks without a liner extension are most likely to
develop large ice buildups.

The liner extension


for single or multiple
liners

The liner extension is the second element exposed to plume


downwash. If the liner extension is uninsulated, the surface should
always be maintained above freezing temperature. If so, no ice
could form, but some condensation might occur that would then run
down to the stack roof. An insulated extension might have some ice
formation.

The roof of a
multiple-liner stack

The stack roof usually does not directly touch the liners, which
permits expansion and movement of the liners. The roof often
receives heat from the annulus of the stack, which prevents ice
buildup and melts any snow that deposits there. Therefore,
accommodations must be made for the drainage of liquid from the
roof area. Preferably, drains should be located on the inside of the
stack shell to keep them from freezing. Alternatively, they will need
to be heat-traced. Depending on the specific geometry and
weather conditions, some icing can occur on the roof, and
therefore may require heat-tracing to eliminate excessive ice
buildup.

Railings

The metal railings at the top of the shell will quickly cool to the
temperature of the mixture of flue gas and ambient air, and they
will be the first surfaces to become iced; in other words, these
railings are the most likely places for ice to form.

Platform near top of


stack

A metal platform on the outside of the stack near the top can also
be exposed to low concentrations of flue gas as a result of
downwash. Because it is metal and is exposed to the ambient and
flue gas mixture temperature, it may also experience occasional ice
buildup. Such platforms are the second most likely places for ice to
form.

Shell

If ice forms on the shell, it can be expected to be on the downwind


side of the stack. This condition occurs when visible downwash can
be seen on the downwind side of the stack shell.

1-10

Icing can often lead to serious buildups that can fall to the ground. However,
when the icing is occurring, the platform near the top of the stack, the railings,
and possibly the roof, may be slippery. If icing is observed at the top of the stack,
care should be taken in the area around the base of the stack because of the
possibility that the ice could become dislodged and fall to the ground.
The potential for icing can be reduced by employing the following steps:
1. Install a liner extension above the shell to minimize the potential for plume
downwash.
2. Select a stack-liner discharge velocity that minimizes plume downwash over
the expected operating range of the unit at the existing local wind conditions
and that is consistent with other design objectives.
3. During cold ambient temperature conditions with high winds, run the unit at
near full load. (Employing this operating procedure may be natural under
such conditions, because more power is consumed in below-freezing
weather.)
4. Use heated annulus air or heat-tracing to heat the stack hood, roof, or other
areas where ice forms on the top of the stack above the freezing point.
1.5 Contents of the Design Guide
The main contents of the design guide are intended to provide the industry with
information concerning the design and specification of wet stacks.
1.5.1 Information About Key Issues
A main objective of the Revised Wet Stack Design Guide is to provide the latest
information about key issues. For example, the design guide contains discussions
on gas velocity, materials for liner construction, liquid-collection devices and
drainage, continuous-emissions-monitoring systems, and flow modeling. This
guide is intended to provide background information and identify the parameters
that need to be considered in the development of an effective wet stack design.
The key to the successful design of a wet stack is an understanding of the choices
that lead to favorable wet stack operation at an early stage in the design process.
Often, simple low-cost design changes can result in significantly improved wet
operation. But these changes need to be identified early in the design process.
This design guide focuses on the options that should be considered, identifies
existing constraints, and describes the effects that variations in the parameters
have on the design of a wet stack.
1.5.2 Knowledge of the State of the Art
Another main objective of this Revised Wet Stack Design Guide is to summarize
the latest technology available regarding wet stacks. This design guide is a
follow-up to EPRI report CS-2520, Entrainment in Wet Stacks, published in
1982 [1]; EPRI report TR107099, Wet Stacks Design Guide, published in 1996
[2]; and EPRI report TR-109380, Guidelines for the Fluid Dynamic Design of
1-11

Power Plant Ducts, published in 1998 [4]. The Revised Wet Stack Design Guide
contains a presentation of the latest information available, updating what has
been learned in the laboratory and from field installations over the past 14 years.
1.5.3 Overview of Project Implementation
The Revised Wet Stack Design Guide is intended to provide the wet stack designer
with information and recommendations concerning the design and specification
of wet stacks. A systematic process is provided in Section 4 to assist the designer
in identifying the important issues that need to be addressed in the design
process.
The guide is presented in two main sections: 1) wet stack fundamentals and 2)
the design process. These sections provide the designer with the information and
background that lead to a bid specification and identify what to specify for work
scope, what information to provide in the specifications, and what to expect in
terms of state-of-the-art operation.

1-12

Section 2: Wet Stack Design Fundamentals


2.1 Introduction
Stack liquid discharge (SLD) is also known as rainout or acid-mist fallout. It is
found in the form of droplets deposited on the ground and on structures in the
vicinity of a wet stack. Typically, these droplets are acidic; they can lead to
various forms of damage in the vicinity of the stack, ranging from damaging
ground cover to etching spots on the painted surfaces of automobiles either
parked in or passing through the affected area. Damage to vehicle paint is the
most common complaint related to SLD. If not mitigated, long-term exposure
can lead to increased levels of metal corrosion on equipment and structures in the
vicinity of the stack.
It is important to note that all wet stacks will have some amount of SLD.
However, SLD is only a problem if the droplets are large enough to be detectable
at ground level near the stack.
The quantity of liquid discharged from the top of a stack is the result of gas- and
liquid-flow processes that take place in the ductwork and stack system between
the discharge of the absorber(s) and the top of the stack liner. The sources of
liquid discharge can be categorized by the type of liquid-flow process at the
location of their origin, and by the resulting droplet sizes, as follows:

Liquid droplets carried from the absorber to the top of the stack liner by the
gas stream without depositing along the gas-flow path. Only small droplets
(10100 m) travel through the ducts without complete deposition. This
means that most conventional scrubber designs, which have horizontal ducts
connecting the absorber to the stack, will result in the deposition of most of
their larger droplets on the internal duct and stack-liner surfaces.

Medium-sized drops (100300 m) are typically generated from a mist


eliminator that is experiencing performance issues. They may or may not
collect in the wet duct/stack system, depending on the geometry of the
system.

Droplets re-entrained from liquid deposition and condensation on duct and


liner surfaces that can be discharged if the local gas velocities are high and/or
if the local geometry is not favorable for wet operation. These droplets will
typically be large (3006000 m) and are usually the major contributors to
the SLD.
2-1

Droplets formed by adiabatic condensation. These represent a large liquidflow rate but are made up of very fine droplets (<10 m), which have a
negligible effect on the SLD detectable at ground level.

Liquid deposited and condensed on the liner, forming a liquid film that will
flow upward in those areas of the stack where the gas velocity is higher than
the flow-reversal velocity for the liner material used (see Section 2.2.2). If
this liquid film reaches the stack outlet, it will be re-entrained back into the
gas flow. The droplet sizes re-entrained at the top of the liner, where
velocities are highest, can be quite large (3006000 m), depending on the
liner-top geometry. Droplets of this size will impact ground-level surfaces in
the vicinity of the liner because they will not be able to evaporate before
reaching the ground. The upper section of the stack choke, if installed, is
usually operating in this mode of upward liquid flow.

All plants operating with a WFGD system will have a white plume. This plume
is a result of the very small droplets formed both by adiabatic condensation and
by additional condensation as the water vapor in the plume is cooled by mixing
with the surrounding air. This cooling process results in additional water-vapor
condensation on particulates and on small droplets within the plume. The
increase in diameter is small (only 110 m). At this size, these newly formed
droplets will follow the local gas-flow patterns; they should evaporate before
reaching the ground as a result of mixing with the drier air away from the stack.
The balance between the plume-cooling and turbulent-mixing processes defines
the length of the resulting visible plume. Therefore, the visible plume length is
primarily a function of the ambient air temperature and the air relative humidity.
The quantity and location of the fallout at a given plant is a function of the
droplet-size distribution of the discharged liquid and the local atmospheric
conditions, such as ambient temperature, wind, relative humidity, and turbulence
level. Large droplets (>300 m) discharged from the top of the stack will typically
reach the ground under most circumstances. Mid-sized droplets (100300 m)
can make it to the ground before evaporating, depending on the local weather
conditions. For example, droplet deposition is typically greater on hot humid
days than on dry days because of the suppressed droplet-evaporation rates. Small
droplets (10100 m) typically evaporate before hitting the ground.
As stated previously, all stacks will have occurrences of SLD. One of the major
objectives in developing an effective wet stack design is to limit SLD to a
minimal acceptable level through proper system design and the installation of an
effective liquid-collection/drainage system. The goal of this system is to collect
and remove deposited liquid from the system before it has an opportunity to reentrain back into the gas flow and to limit the size of droplets that will ultimately
reach the stack outlet. If the droplets are small enough, they will evaporate
before hitting the ground. However, if the droplets are large, they will land on
plant structures, equipment, and cars. A survey of plants with wet stacks found
that the fallout of liquid droplets, if noticeable, usually occurs within a half-mile
radius of the stack.

2-2

2.2 Droplet and Liquid-Flow Fundamentals


2.2.1 Flow of Droplets Suspended Within a Gas Flow
As droplets flow suspended within the flue gas, they will impinge upon and be
collected on internal surfaces as a result of droplet-trajectory paths controlled by
the earth's gravity, centrifugal forces on the droplets when the gas flows through
a turn, and gas-drag forces on the droplet. Because of their size and inertia, large
droplets (3006000 m) from the gas-flow path deposit most readily; mediumand small-sized droplets (10300 m) deposit only partially; and fine droplets
(<10 m) deposit hardly at all because these droplets follow the gas-flow patterns.
The higher the gas velocity and the sharper the turn of the flue gasflow path,
the more likely it will be that significant numbers of smaller droplets will impinge
upon and be collected on walls, turning vanes, baffles, and internal duct
components (bracing, gusset plates, and expansion joints).
Large drops are easily separated from the gas flow by both centrifugal and
gravitational forces. In addition to gravitational and centrifugal separation, smallto medium-sized droplets can also deposit on duct and liner surfaces in regions of
separated flow and as a result of turbulent deposition. However, because of the
generally uniform flow patterns within the upper liner, the turbulent deposition
of fine droplets formed by adiabatic condensation (as a result of the pressure
reduction with elevation change in the liner) along the height of the stack is
relatively small. Only 2.55% of the fine droplets formed as a result of adiabatic
condensation actually contact and collect on the liner wall.
Gas velocity, duct geometry, stack geometry, and droplet size govern the amount
and location of liquid deposition within the wet stack system. Significant
quantities of liquid collect along the outside radius of turns and on the pressure
side of turning vanes and baffles. The major droplet-deposition area in a stack is
on the liner opposite the breeching duct for stacks with side-entry breeches, and
along the outside radius of stacks with bottom-entry elbows.
Liquid deposition and removal from the system can be promoted by installing
properly designed collection devices, vanes, and baffles in the wet duct/stack
system. In order to minimize the potential for SLD, maximum deposition and
collection of the liquid on the duct walls and stack liners must be achieved.
Liquid-collection systems are typically designed by using a physical-flow model.
The main objectives of physical-flow modeling are 1) to understand the systems
internal deposition and liquid-film flow patterns and 2) to design liquidcollectorssuch as ceiling gutters, side-wall channels, and floor guidesto
collect and channel deposited liquid to a drain before re-entrainment can take
place. This process is discussed in Section 3.11.1, under "Laboratory Flow
Modeling."

2-3

2.2.2 Liquid-Film Flow on Surfaces and Re-entrainment


Fundamentals
The liquid on the duct and liner surfaces is produced by droplet deposition and
condensation flows in the form of a film, or as rivulets (small flow streams on the
liner surface), governed by the local gravitational, surface-tension, and gas-shear
forces. The liquid film naturally wants to flow both downward (because of
gravitational forces) and in the direction of the gas flow because of the gas-shear
forces). In ducting, these forces will result in liquid flowing toward the duct floor
as it moves in the direction of the gas flow near the surfaces.
The liquid-film flow in the vertical stack liner is a function of the gas-shear and
gravitational forces, which are acting in opposite directions to each other. For
most liner surfaces, in which gas velocities are below 70 ft/sec (21.3 m/s),
gravitational forces dominate, and the liquid film will flow downward. At
velocities between 70 and 90 ft/s (21.3 and 27.4 m/s), the gravitational and shear
forces have approximately the same magnitude, and the forces are balanced. In
this range, the liquid film in the liner will generally be stagnant on the wall and
will not move in either direction. At velocities above 90 ft/s (27.4 m/s), the gasshear forces dominate, and the liquid film will start to flow vertically toward the
stack outlet. This velocity point is called the flow-reversal velocity.
The changes in the liner liquid-film velocity are gradual. At low gas velocities,
gravity dominates, and the liquid film moves quickly downwardresulting in a
thin liquid film. As the flue gas velocity increases, the downward velocity of the
liquid film decreases because of increased vertical gas-shear forces. However, as
the liquid deposition and condensation rates are relatively independent of the gas
velocity, the total quantity of liquid accumulating on the liner wall remains
constant with time. This means that, as the downward velocity of the liquid-film
flow decreases, the drainage rate of collected liquid from the liner surface
decreases, and the thickness of the film on the liner wall must therefore increase.
The surface of thick liquid films is less stable than that of thin ones, and it is
easier for waves to form, which can lead to large quantities of re-entrained
droplets. Thick liquid films are the most frequent source of SLD and the
resulting fallout of liquid droplets in the vicinity of the stack.
Droplets can easily be re-entrained from the surfaces of thick liquid films because
of surface instabilities when they are exposed to a gas flow. Thick liquid films can
be generated in three different ways. First, as detailed above, thick liquid films
can be generated if the liner surface-drainage rate is reduced by exposing the film
to high gas velocities. Thick liquid films can also be generated if the liquid feed
rate to the wall is high as a result of high levels of droplet deposition or
condensation. Finally, they can be generated if the stack is too high and all the
collected liquid flows to the bottom of the liner without any intermediate
drainage locations.
Liquid re-entrainment most frequently occurs because of one or more of the
following:

High gas-velocity shear near surfaces covered by a liquid film.


2-4

Instability of the liquid film (waves and ripples) on vertical stack-liner


surfaces. This instability occurs because, as gravity is pulling the liquid film
downward, the gas-drag forces are pushing the liquid film or droplet upward.
This effect increases dramatically with liquid-film thickness.

Surface discontinuities and protrusions that disrupt gas and liquid flow
locally, allowing the gas flow to get under the liquid.

Vanes and baffles that cause gas-flow separation, recirculating liquid flow,
and re-entrainment.

Gas-flow patterns that drag liquid along the surface to re-entrainment sites.

Strong vortex patterns that can pick water up in the core of the vortex from
horizontal or vertical surfaces.

Re-entrainment takes place at locations such as

Duct walls, roofs, floors, and stack-liner surfaces

Dampers, damper-blade guides, and trusses

Thermal expansion joints

Trailing edges of turning vanes

Duct-junction corners

Breeching-duct area changes

Stack floor

Stack-liner surfaces, particularly in the entrance region and at the


intersection with the breeching duct

Solids buildup or scaling on internal surfaces

Stack-liner contraction sections

Stack choke surfaces, because of increasing gas velocity

Liquid re-entrainment at these locations will be in the form of large droplets


(3006000 m); they will be discharged at the top of the stack, if not prevented
by a combination of favorable gas-flow-path design and an optimized liquidcollector system. Droplets over 100 m do not completely evaporate outside of
the stack and will probably reach the ground level.
2.2.3 Gas-Flow in Wet Duct Systems
With the possible exception of materials, the design of wet ducts is the same as
that for dry ducts with respect to their overall design and pressure-loss
considerations. The reader is directed to EPRI Report TR-109380, Guidelines for
the Fluid Dynamic Design of Power Plant Ducts, for a detailed discussion of this
process [3]. Some basic considerations for wet operation are discussed in that
report.

2-5

The gasses exiting the absorber will be fully saturated and will contain suspended
droplets. These gasses will flow through the ducting, coming in contact with the
duct walls and with any internal structures within the duct. Liquid will be
continuously condensing from the gas onto these surfaces, forming a liquid film
that will flow downward under gravity and in the direction of the local gas flow.
Care needs to be taken to ensure that the gas-flow patterns moving this liquid
film are favorable for wet operation, minimizing conditions such as flow
separations or recirculation zonespotentially resulting in the re-entrainment of
liquid back into the gas-flow. Also, to the extent possible, the gas-flow patterns
should be managed in such a way as to move the resulting liquid films to drainage
points for removal from the system.
Turning vanes are typically used to optimize gas-flow patterns, reduce pressure
losses, and/or promote liquid collection. An economic assessment must be made
with respect to the value of turning vanes. Although desirable from a pressureloss standpoint, the corrosion-resistant materials required for installation in a wet
environment are very expensive, and the installation cost may not justify the
savings gained by the reduced operating cost. An economic study should be
performed to quantify the life-cycle cost or savings of such an installation.
Some turning vanes may be specifically required to reduce or eliminate SLD.
Such a situation might occur if multiple wet ducts combine immediately
upstream of the stack entrance. In such a situation, vane installation will be
required to ensure that the gas flow entering the stack is favorable for wet
operation and, possibly, to minimize swirling gas flow within the liner. Another
situation would be one in which the absorber is close-coupled to a stack, as in
Figure 2-1. Without vane installation, the gas flow would enter the stack with a
high vertical component, minimizing the ability of the stack to generate the
lower stack-liner flow profiles required for favorable wet operation and liquid
collection. Although this phenomenon is most prominent in close-coupled
arrangements utilizing a side-entry breech, it can also occur in close-coupled
units with bottom-entry elbows. In such a situation, a turning vane would be
required to ensure that the gas flow will enter the stack horizontally. Vanes that
are specifically required for the minimization of SLD should be installed as
recommended.
A more detailed discussion of wet duct design is presented in Section 3.6 of this
guide.

2-6

Droplet
Deposition
High on Liner
Wall
Droplet
Deposition
Low on Liner
Wall Liquid
Collection is
Enhanced

Absorber Outlet

Absorber Outlet

Figure 2-1
Liner Inlet Gas-flow Patterns Close-Coupled Absorber

2.2.4 Gas Flows in Wet Stack Systems


The gas-flow profiles within the stack liner are very important with respect to the
collection and drainage of liquid from the flow gas. It is important to ensure that
these flow profiles are as well defined and robust as possible, given the ductwork
arrangement leading to the liner. There are two major stack-entry design options:
a side-entry breech and a bottom-entry elbow (Figure 2-2). Both designs are
similar in terms of fluid dynamics, given that they both turn the gas flow 90
from horizontal to vertical. However, there are major differences between them
with respect to the intensity of the resulting flow patterns and to the way in
which these flow patterns interact with their respective physical geometries.

2-7

Liner

Absorber Outlet

Liner

Absorber Outlet

Side Entry Breach

Bottom Entry Elbow

Figure 2-2
Typical Stack-Inlet Arrangements

As a uniform gas profile flows through a 90 elbow, the gas flow will want to
continue in a straight line. This tendency will result in the formation of a region
of high gas velocity along the back wall of the liner for a side-entry breech
arrangement, and along the outside radius of the bend for a bottom-entry elbow
arrangement. The resulting gas velocity along the outer radius of the elbow can
be 50100% higher than the average gas velocity entering the elbow. When the
gas hits the back wall of the elbow, it splits in two, and each half flows
circumferentially around toward the inside radius of the bend. This flow
generates two counter-rotating secondary vortices, which form on the
downstream side of the turn. If the curvature radius of the elbow is small enough,
a flow-separation zone will also form along the downstream side of the inner
radius of the turn. These phenomena are shown in Figure 2-3. The strength of
the secondary vortices and the size of the separation zone increase with the
sharpness of the turn and with the velocity of the gas entering the elbow.

2-8

Higher than
Average Axial
Velocity
Along Outside
Radius of Bend

Secondary
Flows
Secondary
Flows

SECTION AA

Figure 2-3
Gas-flow Patterns Through a Standard 90 Elbow

For favorable wet operation, strong secondary vortices and a large region of flow
separation are desirable. For this reason, the side-entry breech, defined by a sharp
90 turn, is considered the most favorable design for wet operation. The flow
entering the liner through the breech generates strong robust secondary flows, in
the form of two large counter-rotating vortices, in the lower liner. A large gasrecirculation zone is formed just above the breech in the liner as the gas turns
from the vertical into the liner. These secondary flows, and the recirculation zone
detailed in Figure 2-4, are relied upon to move the collected liquid film on the
walls of the liner circumferentially out of the main gas flow and into the quiet
zones for drainage from the system. Experience has shown that the secondary
flows generated within a liner using a side-entry breech remain stable and welldefined even at reduced flows, ensuring effective wet operation over a wide range
of unit operating loads.

2-9

Uniform Vertical
Gas Flow

Flow
Reattaches

SECTION CC
High Gas
Velocity Along
Rear Wall

Flow
Separation

Upward Bias to
Flow Through
Breach

Gas Flow

B
Flow
Separation

High Gas
Velocity Along
Rear Wall

SECTION BB

Well
Centered
Stagnation
Point

A
SECTION AA

Chaotic Flow
Along Liner Floor

Secondary Flow
Vortices

Figure 2-4
Liner-Inlet Gas-flow Patterns Side-entry Breech

The use of bottom-entry elbows has become very popular with the increased use
of FRP liners. Bottom elbows are very similar to standard duct elbows and,
because of their size, are usually fabricated from liner-can sections spun and
assembled on site. Typically, bottom-entry elbows consist of three to four miter
cuts, forming a 90 elbow with the center-line turn radius to internal-diameter
ratio of between 0.8 to 1.5. This design is very clean in terms of fluid dynamics,
and it results in low system-pressure loss. When flue gas flows through an
unvaned bottom-entry elbow, two counter-rotating secondary-flow vortices are
generated, as detailed above. Because the change in direction from horizontal to
vertical is generally gradual, the intensity of the secondary-flow vortices is not as
great as that generated in the more abrupt side-entry breech arrangement.
Additionally, because the turn is more gradual, a region of flow separation may or
may not be generated along the downstream side of the inside radius of the
elbow. If a region of flow separation is generated, it will be smaller and less
intense than that generated in the side-entry breech arrangement. It has been
observed in flow-model studies of bottom-entry elbow arrangements that two
smallbut very intenseregions of flow recirculation are generated on both
sides of the elbow along the last miter cut (Figure 2-5). These recirculation zones
2-10

are formed because the change in direction from horizontal to vertical is abrupt
in these locations, and the incoming horizontal gas flow over shoots the inside
radius of the turn. Although small, these recirculation zones are problematic,
because they can be sites of significant liquid re-entrainment.
Uniform
Vertical Flow

Flow
Separation

Flow
Reattaches
Flow
Separation

SECTION CC
High Gas
Velocity
Along
Rear Wall

B
High Gas
Velocity Along
Rear Wall

SECTION BB

Gas Flow

SECTION AA
High Intensity
Gas Recirculation
Along Duct Wall
at Seam

Secondary Flow
Vortices

Figure 2-5
Liner-Inlet Gas-flow Patterns 3-Miter Cut Bottom-Entry Elbow

For both liner-entry arrangements, the counter-rotating secondary-flow vortices


generated in the entry region dissipate significantly within 34 liner diameters
above the roof of the inlet duct/breech. Above 56 liner diameters, the flow will
be essentially vertical and uniform. For this reason, a liner is typically divided into
two regions: The lower liner, where the secondary-flow vortices are generated
and dissipate, is defined as the region in the liner extending from the liner floor
to a point 34 liner diameters above the roof of the inlet/breech duct; the region
above this point is defined as the upper liner, where the gas-flow profile is
essentially uniform and near vertical.
It is clear that the secondary-flow vortices are desirable and necessary for
favorable wet stack operation. For this reason, flow controls in the liner are
2-11

discouraged unless absolutely necessary. Flow controls can reduce or eliminate


the secondary-flow vortices and can act as locations for liquid collection and reentrainment back into the gas flow with little chance for recollectionboth of
which can lead to increased levels of SLD. It should be noted that a division wall
separating two breech entries into a single liner would not normally be
considered to be a flow-control device as defined above. A properly designed
division wall will divide a liner into two separate aerodynamic regions, each of
which will be allowed to form its own secondary-flow vortices and flowseparation zones, resulting in favorable liquid-collection potential.
2.3 Sources of Liquid in a Duct/Wet Stack System
All of the water-vapor content in the stack-inlet flue gas flow comes from the
absorber outlet where the flue gas is fully saturated. The liquid content of the
stack-inlet flue gas flow comes from the following sources:

Mist-eliminator droplet carryover, which consists of fine drops (<10 m), a


portion of which never contact a mist-eliminator blade.

Medium-sized droplets (100300 m) will re-entrain from mist-eliminator


vanes at high velocities and from solids buildup during abnormal operation.

Condensation on the duct surfaces upstream of the stack inlet, which adds
liquid onto the duct surfaces.

Deposition of droplets on duct surfaces, vanes, baffles, and structure, which


removes droplets from the gas-flow stream.

Re-entrainment of droplets from duct surfaces, vanes, baffles, and structure,


which adds large droplets to the gas-flow stream.

Installation of liquid collectors on duct walls and ceilings, vanes, baffles, and
structure, which collect liquid from these surfaces and guide the liquid to the
duct floor of stack-bottom drains for removal from the duct/stack system.

The geometry of the ducts between the absorbers and the stack inlet; the
number of vanes, baffles, and liquid collectors in the ducts; and the
effectiveness of the mist eliminators and duct liquid collectors will determine
the amount of liquid, droplet sizes, and distribution of liquid in the stack
breeching duct at the stack inlet.

2.3.1 Mist Eliminators and Absorbers


The gas stream leaving the mist eliminator is saturated with water vapor, and
entrained fine liquid droplets are carried by the gas flow under ideal design and
operating conditions. Mist-eliminator carryover is a function of the gas velocity
through the blades and the liquid load in gpm/ft2 (l/min/m2) at its inlet face. It is
important that the mist eliminators operate properly to minimize the size and
rate of droplet carryover. The mist eliminators are designed to minimize
carryover of liquid and slurry from the WFGD absorber into the downstream
ductwork. Absorber problems resulting in a nonuniform gas-flow profile entering
the mist eliminator, or problems such as excessive mist-eliminator pluggage or
fouling, can result in increased liquid carryover. Some larger droplets may also be
2-12

present in the gas stream that were re-entrained from the mist-eliminator blades,
the mist-eliminator support structure, and areas of solids buildup on the mist
eliminator. During the wash cycles, washing the mist eliminator can also increase
the amount of liquid carried with the gas flow downstream of the mist eliminator
because of increased liquid load from the washed section.
A field evaluation of the mist-eliminator performance can be made by
performing visual inspections. If carryover problems are noticeable, SLD will
probably occur if effective liquid collectors are not installed in the duct and the
stack. The mist-eliminator selection, operation performance, and possible
problems are described in [3].
2.3.1.1 Mist-Eliminator Design and Arrangement
The mist-eliminator system is the first and most important element of the liquid
collectors in a wet duct/stack system. The selection of the best mist eliminators to
match the absorber and duct design will result in reduced liquid load in the
duct/stack system and less elaborate duct and stack liquid-collection systems. In
FGD service, the function of the mist eliminators is to remove entrained
scrubbing slurry from the flue gas before it travels out of the absorber into the
downstream ductwork and stack. The mist eliminators are located downstream of
the absorber spray zones, where the flue gas contacts the scrubbing slurry that
removes sulfur dioxide. The slurry content of the liquid carryover can cause solids
buildup in the outlet ductwork and stack, high particulate emissions, and
corrosion. The amount of liquid that the flue gas carries into the mist eliminator
will depend on many factors, including (but not limited to) the droplet size, the
gas velocity, the liquid-to-gas ratio, and the washing method.
There are many mist-eliminator designs, available from numerous manufacturers.
But the chevron type of collector is almost exclusively used by WFGD systems
on coal-fired units. Mist eliminators are manufactured from many different
corrosion-resistant materials, ranging from alloys to FRP and various extruded
plastics.
Mist eliminators can be oriented toward any number of gas-flow directions, from
horizontal to vertical (Figure 2-6). Most modern WFGD systems utilize a
vertical gas-flow orientation or a singularly supported roof top or diamond
design.
Older weir-type scrubbers utilize a horizontal gas-flow mist-eliminator design.
Depending on the separation efficiency required, a number of stages of mist
eliminators in series are used. In utility power-generation applications, two stages
are typically used. In situations in which high liquid loading are experienced, a
bulk entrainment separator may be placed in front of the first mist-eliminator
stage. This device is typically a series of baffles used to limit the liquid loading
entering the mist eliminator.

2-13

Gas Flow

Drainage through
Incoming Gas Flow

2nd ME Stage

Gas Flow

1st ME Stage

1st ME Stage

2nd ME Stage

Drain Box

Vertical Gas Flow

Horizontal Gas Flow

2nd ME Stage

Drain Box
Drainage through
Incoming Gas Flow

Gas Flow

1st ME Stage

Vertical Gas Flow


Diamond Arrangement
Figure 2-6
Common Mist-Eliminator Arrangements

The breakthrough velocity for a mist-eliminator design is a function of its ability


to drain away collected liquid, and it is therefore substantially a function of its
orientation relative to vertical. Vertical gas-flow mist eliminators require collected
liquid to drain back through and against the incoming gas flow. This process will
result in higher liquid loading entering the mist eliminator, because the smaller
droplets from the drainage flow will be immediately entrained back into the unit
after draining from the leading edge of the blades. A horizontal gas-flow mist
eliminator will drain its collected liquid perpendicular to the gas flow into a
collection box, which typically drains directly back to the absorbers reaction tank.
This design minimizes the liquid entrained within the mist eliminator and results
in better performance: either the ability to operate at a higher gas velocity for a
given liquid loading or the ability to handle a higher inlet-liquid loading for a
given inlet-gas velocity.
2-14

Roof-top, or diamond-type, arrangements fall between the two extremes. This


arrangement is becoming more popular because of its ability to support what is
essentially a two-stage mist-eliminator arrangement from a single support
elevation within the absorber. In this design, two stages of mist-eliminator blades
are oriented diagonally, forming a parallelogram which has been rotated so that it
is supported from two of its corners. The design enhances liquid drainage across
most of the collection surface. Higher-than-average liquid loadings will still be
experienced at the units drainage points because the collected liquid is still
required to drain against the incoming gas-flow at these locations. Another
concern with this arrangement is the need to use a mist-eliminator blade design
with a more torturous path (increased curvature) to maintain the ability for the
blades to inertially separate droplets from the gas flow. For a given blade design,
as the angle of the gas passing through it increases from 0 (perpendicular) to 45
or higher, the path length through the blades increases in the diamond
arrangement (Figure 2-7). If the curvature of the blade has not changed, the
side-to-side deflection will be the same for the increased path length through the
blades, so the resulting inertial separation will be lower.

Flow Path Through ME

Comparison of Flow
Paths Through ME
135

90

Perpendicular
137

Direction of
Gas Flow
Through Mist
Eliminator

95

ME Side Elevation

Figure 2-7
Comparison of Gas-Path Lengths Through a Mist Eliminator as a Function of Angle

2-15

The amount of liquid-droplet carryover into the wet duct/stack system from the
mist eliminators can be minimized by
1. Selecting the most suitable mist-eliminator elements for the operating
conditions of the installation
2. Using two-stage mist elimination
3. Using a vertical or angled mist-eliminator orientation
4. Using gas velocities with a safe margin under the critical re-entrainment
velocity for the mist-eliminator blades selected
5. Achieving a sufficiently uniform velocity profile into the second-stage misteliminator to keep all regions below the critical re-entrainment velocity
6. Selecting effective washing techniques and procedures to prevent solids
buildup on the mist eliminators, without entraining wash spray in the gas
flow
7. Establishing inspection and maintenance procedures for the mist eliminators
8. Maintaining good pH control to reduce solids precipitation in the mist
eliminators
EPRI Report GS-6984, FGD Mist Eliminator System Design and Specification
Guide [3] describes all essential aspects of mist-eliminator design, application,
operation, and maintenance necessary for an effective mist-elimination system.
The measured laboratory performance of several commercially available mist
eliminators is also furnished in this report.
Information regarding the performance of mist eliminators as a function of gas
velocity and inlet-liquid load is available from the mist-eliminator manufacturers,
and it should be obtained early in the system-design process. Performance is
given as a function of the gas velocity immediately upstream of the mist
eliminator. Figures 2-8 and 2-9 present typical vertical gas-flow chevron-type
mist-eliminator performance curves for pressure drop and carryover versus gas
velocity, respectively. Mist-eliminator collection efficiency versus droplet size can
also be supplied from the manufacturer if desired. This latter performance curve
is used to identify the maximum droplet size expected to pass through the mist
eliminator without being collected. (See Figure 2-10 for an example.)
For design and estimating purposes, a typical mist-eliminator manufacturers
carryover guarantee is 0.02 gr/acf (46 mg/m3) can be used.

2-16

Pressure Drop vs Gas Velocity

0.300

Pressure Drop (" W.C.)

0.250

0.200

0.150

0.100

0.050

0.000
0.00

2.00

4.00

6.00

8.00

10.00

12.00

Velocity (ft/s)

14.00

16.00

18.00

Figure 2-8
Typical Chevron-Type Mist-eliminator Pressure Drop vs. Gas-Velocity Curve

Figure 2-9
Typical Chevron-Type Mist-Eliminator Carryover vs. Gas-Velocity Curve

2-17

20.00

Figure 2-10
Typical Mist-Eliminator Collection Efficiency vs. Droplet-Size Curve

2.3.1.2 Mist-Eliminator Average Velocity


WFGD absorbers typically operate at an average gas velocity entering (but not
within) the mist-eliminator sections of 1014 ft/s (3.04.3 m/s) for vertical gasfloworiented mist eliminators and 1620 ft/s (4.96.1 m/s) for horizontal gasfloworiented mist eliminators. Under normal inlet-liquid loading conditions
typically 0.5 gpm/ft2 (0.0123 l/min/m2) carryover starts to become an issue at
velocities above 16 ft/s20 ft/s (4.96 m/s) for vertical gas-flow and horizontal
gas-mist eliminators, respectively.
Locally, higher-than-average gas velocities can occur through a mist eliminator
because of a number of reasons, including

Mist-eliminator fouling and/or plugging in localized areas

The location of the absorber outlet too close to the outlet of the mist
eliminators

A nonuniform gas-flow distribution into the mist eliminators

A nonuniform liquid loading from the absorber into the mist eliminators

Because local gas velocities in excess of the mist-eliminator breakthrough velocity


will result in increased droplet and slurry carryover into the downstream ducting,
every effort must be made to ensure that the mist-elimination system is clean,
well maintained, and operating correctly.
2-18

2.3.1.3 Mist-Eliminator Liquid Carryover


If designed and operated properly, a modern mist-elimination system should
experience the guaranteed minimum liquid carryover. Some carryover should be
expected during on-line mist-eliminator washing. Liquid carryover from the
mist-eliminator will occur if the local liquid loading and/or the local gas velocity
are too high. A sudden increase in SLD is often a key indicator of an issue
related to mist-eliminator performance and/or operation.
As stated previously, information regarding the performance of mist eliminators
as a function of gas velocity is available from the mist-eliminator manufacturers,
and it should be obtained early in the system-design process.
To the extent possible, all the liquid carryover from the mist eliminator that
deposits on the duct- and stack-liner surfaces must be collected. A significant
portion of this carryover must be collected in the downstream ducting before it
reaches the stack liner, making the downstream ducting an important part of the
overall liquid-collection system. For this reason, placing the stack liner directly
on top of an absorber should be discouraged. In this arrangement, liquid
carryover from the mist eliminator will have very limited deposition and little
opportunity for collection before being discharged from the top of the liner as
SLD.
2.3.1.4 Mist-Eliminator Wash

Mist eliminators have no moving parts. After they are properly installed, they
will operate at design conditions without any further attention. Because there
are no moving parts, the major potential operating problems are the plugging
of the mist eliminator with insoluble solids and/or the corrosion and
degradation of the materials of construction. The usual operating-pressure
decrease across the mist eliminator when clean is generally 0.20.5 in (5.1
12.5 mm) of water column per chevron stage.

If operating conditions remain unchanged, any significant increase in the


operating-pressure drop is undoubtedly attributable to partial plugging of the
mist eliminator. As a general rule, washing frequency is more important than
wash-water pressure and wash-water volume. However, all three factors must
be taken into consideration to ensure a clean, problem-free mist eliminator.
High-quality water should be used.

Mist eliminators are typically broken down into multiple smaller segments
for on-line washing. Each segment is washed sequentially in series. A
minimum wash frequency for each segment of 1.5 min/h is recommended.
More frequent washing is always better. The pressure of the wash-water
nozzles is generally 2545 psig (0.170.31 MPa). Mist-eliminator spray
nozzles are typically 90120 full-cone nozzles. Most manufacturers
recommend a 150200% overlap of the mist-eliminator spray-wash pattern.
A nozzle to mist-eliminator spacing of 23ft (0.60.9 m) should be used.

2-19

Generally, the lower mist eliminator in a two-stage mist-elimination system


is washed with more liquid than the upper mist eliminator. Most
manufacturers recommend washing the bottom of the lower mist eliminator
at 1.5 gpm/ft2 (61.1 l/min/m2) and the top of the lower mist eliminator at
0.50.6 gpm/ft2 (2125 l/min/m2). It is also recommended to wash the
bottom of the upper mist eliminator at 0.50.6 gpm/ft2 (2125 l/min/m2).
Typically, the top of the upper mist eliminator is not washed during absorber
operation.

During the wash cycle, the liquid within the mist-eliminator section being
washed will be significantly higher, and increased carryover into the downstream
ducting from this section will be expected. A properly designed wet duct/stack
system will be able to handle this additional carryover without increased potential
for SLD.
For a detailed discussion on the effectiveness of mist-eliminator systems, refer to
[3]. To maintain effectiveness and to prevent sludge buildup, it is very important
to keep mist eliminators clean by washing them according to the manufacturers
specifications.
2.3.2 Liner Condensation
The most significant source of liquid in a wet duct/stack system is condensation
of water vapor from the saturated gas flow exiting the WFGD absorber. This
condensation comes from two main sources: 1) thermal condensation on the wall
as a result of heat transfer from the flue gas to the outside air through the liner,
insulation, annulus air, and concrete shell and 2) adiabatic condensation in the
bulk of the gas flow because of the expansion of the saturated gases as a function
of the pressure decrease along the height of the stack.
Detailed analytical calculations must be completed to predict the quantity of
liquid produced by condensation in the absorber outlet duct and in the stack.
These computer models are based on the equations presented and discussed in
EPRI Report CS-2520, Entrainment in Wet Stacks [1]. The accuracy of the
condensation calculation is estimated to be 10%. The rate of liquid
condensation is a function of ambient air temperature, wind velocity, and linerdesign variables. Calculations are performed for a selected worst-case
atmospheric condition to estimate the maximum quantity of liquid condensation
expected in the liner of the subject unit. Determining the expected quantities of
liquid condensation in the duct and liner is necessary for the development of an
effective liquid-collection system, because it will affect the number, sizing, and
location of the various liquid-collection gutters and drains.
To perform the condensation calculations, the following information is required:

Detailed design drawings of the system ductwork from the absorber outlet to
the stack inlet, including material specifications and thicknesses

Information on duct insulation: location, type, thickness, and covering


2-20

Detailed design drawings of the stack shell and liner, including the number
and orientation of liners relative to true North, materials of construction,
material thickness versus elevation, annulus air-gap thicknesses versus
elevation, location and size of stack-shell vents to atmosphere

Information on the liner insulation: type, thickness, and covering

Expected flue gas conditions: temperature, flow rate, constituents, watervapor content

Plant elevation

Wind-rose data for a minimum of 1 year (5 years preferred), presenting air


temperatures, barometric pressure, wind speed, and direction

2.3.2.1 Thermal Condensation


Thermal condensation (or wall condensation) caused by heat transfer is
important in wet stacks. Because the saturated flue gas temperature exiting the
absorber is usually 120130F (4954C), heat transfer through the liner, the
annulus, and the concrete shell to the cooler surrounding ambient air causes
water vapor to condense on the inside surfaces of the ducts and stack liner. The
rate of condensation caused by heat transfer to the outside environment are
functions of the liner and shell construction and its thermal conductivity, the type
and quantity of liner insulation, the gas-flow conditions, and the local
atmospheric conditions (ambient temperature, wind velocity, and barometric
pressure).
Annulus pressurization of brick-lined stacks using outside air increases the rate of
condensation on the liner surfaces. Leakage of pressurizing air through the liner
in older brick stacks can significantly increase the condensation.
It should be noted that, with the exception of alloy liners, the use of insulation in
wet stacks is not very common. Insulation is typically used in dry stacks to keep
the metal temperature of the liner above the acid dew point. And because wet
stacks are always below the acid dew point, corrosion-resistant materials are used
that do not require insulation. FRP, borosilicate block, and brick liners have
some insulating value and rarely have needed to add insulation. However,
because a major source of the liquid generated on the liner surface is thermal
condensation, insulation can be used to significantly reduce this phenomenon.
Previous condensation studies have shown that the addition of two inches of liner
insulation can reduce the quantity of thermal condensation by approximately a
factor of four. Additional increases in the insulation thickness will reduce
condensation further, but at a diminishing rate of return.
2.3.2.2 Adiabatic Condensation and Stack Effect
Adiabatic condensation (or bulk condensation) occurs in the bulk of the gas as
the pressure of the saturated gas flow decreases as a result of pressure loss and
elevation change along the height of the stack. This type of condensation may
generate an appreciable quantity of liquid. The decrease in absolute pressure of
2-21

the saturated gas causes water vapor to condense on fine droplets or on small
solid particles within the bulk of the gas. Droplets formed by this type of
heterogeneous condensation will be very small (<10 m). Only a very small
fraction of this adiabatic condensation will reach and be deposited on the liner
surface by the turbulent-deposition process. Estimates of this fraction range from
2.5 to 5%. The expected percentage of droplets coming in contact with the liner
wall decreases with increasing liner diameter, because of the greater distance a
droplet would need to travel from the center of the liner to the liner wall.
Historically (as stated above), estimates of 2.55% have been used, and these
percentages have been anecdotally supported by observations from operating
units, which can monitor the liquid-drainage rate from their liquid-collection
systems. Condensed droplets that do not encounter the liner wall are discharged
from the top of the liner, which creates the familiar white plume associated with
wet stack operation. These droplets do not represent a problem, because their
size is very small, and they will evaporate before reaching the ground.
The stack effect is the hydrostatic-pressure difference between the inside and
outside of the stack liner. This difference is due to the lower density of the heated
gas in the stack compared to the density of the ambient air. Therefore, the stack
effect is calculated by multiplying the density difference of the gas in the stack
and the ambient air times the height of the stack. Net stack draft is the operating
pressure at the inlet of the stack liner, and it is calculated by subtracting the fluid
friction, the inlet losses, and the exit losses from the stack effect.
For wet stack design, the worst-case conditions are the maximum stack gas-flow
rate resulting in the highest friction loss, in combination with the hottest and
most humid ambient conditions. These conditions can normally be expected to
produce a positive gas pressure at the stack inlet. The stack effect at worst-case
conditions is generally only 0.20.3 in Wg (0.050.07 kPa) for wet stack
operationa small amount compared to the friction and exit losses at full load.
2.4 Liquid-Film Behavior on the Liner Wall
Because the flow characteristics are so different, the discussion of the liquid-film
behavior on the liner wall must be broken down into two sections: the lower liner
and the upper liner. The lower liner is defined as that portion of the liner located
below a point four liner diameters above the roof of the breech duct or bottomentry elbow. The upper liner is everything above that location.
As discussed previously, the gas-flow patterns in the lower liner are threedimensional and complex, and they are used for the majority of the liquid
collection in the stack. The flow entering the liner through the breech or bottomentry elbow generates secondary flows in the form of two large counter-rotating
vortices in the lower liner. These secondary flows are relied upon to drag the
liquid film on the walls of the liner circumferentially out of the high-velocity gas
flow into low-velocity or separated regions where it can be drained from the
system.

2-22

In a side-entry breech arrangement, the droplets entrained in the incoming gas


flow will typically impact on the rear wall of the liner. The droplets will impact in
a zone extending approximately 1 liner diameter from a point on the rear wall
corresponding to the roof of the breech opening (Figure 2-11). Finer droplets
will impact in the upper portion of this zone, whereas larger droplets will impact
lower in the lower portion. Secondary gas flows generated by the turn into the
liner will direct a large portion of the liquid collected on the rear wall of the liner
upward, then circumferentially around to the front of the liner above the breech,
where it will encounter the downward flow in the gas-recirculation zone located
above the breech opening on the front wall of the liner. When the liquid reaches
this recirculation zone, it will flow downward and be re-entrained from the top
edge of the breech opening. Liquid collected on the liner walls below the dropletimpact zone will either flow downward onto the liner floor or will be pulled
circumferentially around the liner to the vertical side edges of the breech, where
it will be re-entrained back into the gas flow. Effective liquid collectors can be
developed to work with these liquid-flow patterns to collect and direct these
liquid films to drain points for removal from the system.
Reentrainment

Uniform
Downward Liquid
Film Flow
From Upper Liner

Liquid Film Flow

Circumferential
Liquid Film Flow

Downward
Liquid Film
Flow in Flow
Separation

Upward Liquid
Flow

Droplet Impact
Zone

Reentrainment

Gas Flow

Downward Film
Flow Below
Impact Zone

Downward Flow
Along Breach
Opening

REAR WALL OF
LINER

Liquid Pooling with


Chaotic and Vortex
Flow On Liner Floor

Figure 2-11
Lower Liner Droplet- and Liquid-Film Flow Patterns Side-Entry Breech

2-23

Some liquid will make it to, and collect on, the liner floor. If the liner floor is flat,
the collected liquid will pool and slosh around until it finds a drain point.
Typically, two vortices can be observed on the flat floor of the liner, their centers
located approximately 1/3 of the way across the liner and even with the side walls
of the inlet breech duct (Figure 2-12). Under some conditions, these vortices can
be sufficiently strong to re-entrain liquid off of the floor and back into the gas
flow.
X

Liquid Pooling on
Liner Floor

X/3

Gas Flow
Vortex Flows on
Liner Floor
Reentrainment
Liquid Film Flow

Gas Flow

Liquid
Reentrainment
from Vortex Flow

Vortex Flows on
Liner Floor
Figure 2-12
Liner-Floor Liquid-Film Flow Patterns Side-Entry Breech

For a bottom-entry elbow, the droplets entrained in the incoming gas-flow will
typically impact in a zone starting on the outside radius of the elbow at an
elevation corresponding approximately to the roof of the inlet duct and extending
vertically approximately 11.5 liner diameters (Figure 2-13). Similar to the sideentry breech, finer droplets will impact along the top of the impact zone, and
larger droplets will impact along the bottom. Secondary gas flows generated by
the 90 elbow into the liner will drag a large portion of the liquid collected on the
2-24

rear wall of the liner upward. Then it will be dragged circumferentially around to
the inside radius of the inlet elbow, where it will encounter a lower vertical gas
velocity. It will then flow downward until it either re-entrains back into the gas
from the top of the duct, or it will flow circumferentially downward on either
side of the inlet duct toward the bottom of the elbow. In this latter case, there is a
high probability that the liquid will be shredded off of the wall by recirculation
zones located along the walls of the elbow near the inside radius. Liquid that
makes it to the floor of the inlet elbow will tend to pool there, because the
horizontal gas velocity along the floor is not high enough to drag the liquid back
up the outside radius of the elbow.
Reentrainment

Uniform
Downward
Liquid Film
Flow From
Upper Liner

Liquid Film Flow

Circumferential
Liquid Film Flow

Upward Film
Flow

Downward
Flow in Flow
Separation

Circumferential
Liquid Film Flow
Droplet
Deposition
Zone

Downward
Film Flow

Gas Flow

Liquid Pooling on Inlet


Duct Floor

Droplet Reentrainment
from High Intensity Gas
Recirculation

REAR WALL OF
LINER

Figure 2-13
Lower-Liner Droplet- and Liquid-Film Flow Patterns Bottom-Entry Elbow

Stacks are typically designed so that condensed liquid on the liner wall will flow
downward for collection. The stack-design velocity is the average gas velocity in
the liner defined as the volume flow rate divided by the area of liner. As discussed
above, for both inlet arrangements, the flue gas that enters the liner impacts the
wall opposite the inlet and forms a region of locally high gas velocity along the
rear wall of the liner or outer radius of a bottom-entry elbow. The gas velocity is
usually high enough to drag the deposited liquid vertically up the liner wall until
the secondary-flow vortices are strong enough to pull the liquid circumferentially
around and out of the high-velocity region. If the vertical gas velocity is not
sufficiently high, the deposited liquid will either move very slowly upward or not
2-25

move at all. Both of these situations are undesirable, because as more liquid is
deposited in this region, the liquid-film thickness will increase until the surface
becomes unstable, wavelets are formed, and droplets will begin re-entraining
back into the gas flow. There is little chance that many of these droplets can be
recollected before exiting the liner as SLD. If this situation occurs, liquidcollection devices in the form of V- shaped gutters and other types of liquid
collectors will need to be installed in the droplet-impact region to mechanically
help the liquid move circumferentially into the secondary-flow vortices. These
devices will be discussed in more detail in Section 3.11.8.
Liquid-film flows in the upper liner should be better behaved. In this region, the
gas-flow profile moving up the liner is more uniform, and the liquid film formed
as a result of thermal and adiabatic condensation should flow downward without
obstruction.
The motion of a vertical liquid film exposed to a gas flow is a function of many
variables. The most important of these are the surface tension of the liquid being
used, the gas velocity near the film, the roughness of the surface, and the liquidfilm thickness (which is itself a function of the condensation rate, the filmdrainage rate, and the height over which the condensation takes place).
To understand liquid-film flow on a liner surface, it is best to begin by looking at
the ideal case of a smooth wetting surface. A wetting surface is defined as a
surface upon which a liquid will form a film, as opposed to beading up after being
spread out (Figure 2-14). A wetting surface is referred to as being hydrophilic,
whereas a nonwetting surface is referred to as being hydrophobic. To be effective,
a stack-liner material must be hydrophilic. Some materials, such as FRP, can
transform from one surface characteristic to the other after exposure to the
corrosive conditions within the liner.

Wetting Surface
Figure 2-14
Wetting vs. Nonwetting Surfaces

2-26

Non-wetting Surface

If liquid if deposited on a vertical, smooth wetting surface at a uniform rate


without being exposed to a vertical gas flow, the resulting liquid film will flow
easily downward under the influence of gravity. If a vertical gas flow is applied to
this wetting surface, a new forcegas-shearis added to the downward-flowing
liquid film. This force produces an upward drag on the liquid film, slowing its
downward movement. As the gas velocity is increased, the gas-shear force
increases, and more drag is applied to the liquid filmfurther slowing it down.
Because the liquid is being deposited at a constant rate, the reduction in the
downward flow of the film means that the drainage rate is reduced and the
inventory of liquid held on the plate is increased. This process results in an
increase in the thickness of the liquid film. The thickness is not uniform; it
occurs in waves and ripples on the film surface.
At gas velocities below 50 ft/s (15.2 m/s), the liquid-film motion will be
downward, and the surface of the film will be smooth. Between 50 and 60 ft/s
(15.218.3 m/s) the liquid-film motion will still be downward, but waves and
ripples will form on the surface of the film.
If the gas velocity is increased furtherto 70 ft/s (21.4 m/s)the downward
flow of the liquid film will be reduced, causing further increases in the film and
surface-wave thickness. The influence of the surface-tension forces will begin to
diminish relative to the gas-drag forces to the point at which small waves or
wavelets will begin to form on the surface of the film. At first these wavelets will
be sporadic, and it will look like the surface of the film is standing still or even
moving vertically.
At 7085 ft/s (21.425.9 m/s), the magnitude of the gas-shear force equals that
of the gravitational force, and the liquid film stops moving and is stagnant on the
wall. At this point, the thickness of the liquid film continues to increase, until the
surface of the film becomes unstable and starts to break down. Waves are formed
of sufficient size to allow the gas flow to get under them, forming white caps
not unlike waves on the ocean. Large quantities of liquid are shed back into the
gas flow. If left unchanged, a steady-state condition will be reached, in which all
of the liquid deposited on the plate must leave the plate as re-entrained droplets.
Operating under these conditions, all of the liquid condensing within a stack
liner would leave the liner as SLD.
If the gas velocity is increased to 8590 ft/s (25.927.4 m/s), the gas-shear forces
will begin to dominate the gravitational forces, and the liquid film will begin to
move againbut now vertically up the surface. Significant quantities of droplets
will still be re-entrained off of the surface of the liquid film. Liquid will begin to
move toward the top edge of the plate for drainage. If uncollected, this liquid will
flow over the top of the plate or be re-entrained back into the gas flow.
Additional increases in gas velocity will increase the vertical speed of the liquid
film.

2-27

At velocities above 90 ft/s (24.7 m/s), re-entrainment off of the liquid-film


surface is significantly reduced (although not eliminated), and the vertical motion
of the liquid film is well-defined. This condition can often exist at the outlet of a
wet stack incorporating an outlet choke.

0-50 ft/s
(0-15.2 m/s)

55-65 ft/s
(16.7-19.8 m/s)

Upward Liquid Film Flow


Droplet Reentrainment Reduces with
Increased Gas Velocity

Little to no Downward Liquid Flow


Droplets Reentrained from Entire Surface

Downward Liquid Flow with Waves/Rivulets Forming on


Surface, Occasional Droplets Reentrained from Surface

Smooth Downward Liquid Film Flow


No Droplet Reentrainment

Gas Flow

The observations described above are presented visually in Figure 2-15. A


number of factors can impact these ideal observations, particularly with regard to
the liquid-film thickness. The surface of a thick liquid film will become unstable
at lower gas velocities than that of a thin liquid film, so film-thickness
management must be considered during the design of a wet stack system.
Specifically tall stackstypically greater than 500 ft (152.5 m)without
intermediate liquid collectors can experience thick liquid films. Stacks in colder
climates can experience thick liquid films because of high condensation rates
within the liner. In these situations, the addition of insulation (or more
insulation) is the easiest manner in which to control the resulting liquid-film
thickness.

70-85 ft/s
(21.3-25.9 m/s)

>90 ft/s
(>27.4 m/s)

Figure 2-15
Liquid-Film Flow Patterns on a Smooth, Vertical Wetting Surface vs. Vertical Gas
Velocity

2.5 Liner-Wall Discontinuities


The observations described above apply to the ideal case of a smooth wetting
surface. In reality, the surfaces are anything but smooth. On alloy liners, there are
horizontal weld beads; on FRP liners, there are joints between adjacent liner
cans; and on brick liners, there are horizontal mortar joints every 24 in (50100
mm) up the entire height of the liner. These disturbances are referred to as linerwall discontinuities.
2-28

When the liquid film flows over a horizontal discontinuity, there is a potential for
the upward- flowing flue gas to get under the liquid, resulting in the formation of
droplets. A portion of these droplets will be re-entrained back into the gas flow
and will exit the liner as SLD.
The major areas of concern in an alloy liner are the horizontal weld beads joining
the plates together. Vertical weld beads will have little to no impact on droplet
re-entrainment. But horizontal weld beads on alloy liners can cause liquid reentrainment in the stack at gas-flow velocities as low as 40 ft/s (12 m/s) if not
properly made. The height of the horizontal weld beads is critical, and their
height should be the minimum possible that does not exceed the standard for
good weld quality: 1/8 in (3.1 mm). Grinding all horizontal beads flush to the
base material is ideal. However, laboratory studies have demonstrated that, for
gas velocities up to 50 ft/s (15.2 m/s), a 1/8-in (3.1-mm) bead height on a butt
joint will result in minimal levels of droplet formation (assuming the bead is
smoothly rounded). Slight increases above this velocity5560 ft/s (16.718.3
m/s)resulted in liquid holdup along the joint and the start of droplet reentrainment back into the gas flow. At velocities above 65 ft/s (19.8 m/s), droplet
re-entrainment becomes more pronounced. Figure 2-16 details liquid-flow
patterns over a 1/8-in (3.1-mm) weld bead as a function of gas velocity.
Droplets
Reentrained From
Surface

Liquid Flows
Over Weld Bead

Wave/Rivulets
Forming on Surface

Gas Flow

Weld Bead
Liquid Starts to
Holdup Over
Weld Bead
With some
Reentrainment

Gas Velocity
<50 ft/s (15.2 m/s)

Film Thickening
Due to Liquid
Holdup
Droplets
Reentrained
From Weld Bead
Gas gets Under
Liquid at Weld
Bead and Waves

Down Flow of
Liquid Reduced or
Stopped
Gas Velocity
Gas Velocity >65
55 - 60 ft/s (16.8-18.3 m/s)
ft/s (19.8 m/s)

Figure 2-16
Liquid Flow over 1/8-in (3.1-mm) Weld Beads

The rings of an FRP liner should be joined together with material mostly on the
outside of the liner. Only a minimal inward protrusion with smooth transitions is
recommended. FRP liner-can joints are typically lap-type joints with interior and
exterior overlays (Figure 2-17). A 6:1 to 10:1 taper should be used from the
stiffener ringtop and bottomback to the base-liner material, with smooth
transitions.
2-29

Liner Wall
Detail 1

Inside Flue

Interior
Reinforcing Band

(Minimum Thickness
Possible)

Minimum Smooth
1:6 Taper Back To
Liner Wall
(1:10 Preferred)

Outside
Flue

Exterior
Reinforcing
Band

Detail 1
Figure 2-17
FRP Liner-Can Joints Internal-Taper Requirement

Special attention must be paid to brick liners. Every effort must be made to
ensure the proper radial alignment of the bricks and to strike the mortar flush
with the inside surface of the liner. If the maximum projection or offset between
bricks on the interior surface of the liner exceeds 1/8 in (3.1 mm), re-entrainment
may occur from the liner surface. Extruded or depressed mortar joints can be
significant sources of liquid re-entrainment back into the gas flow. The
construction should be clean, without mortar smeared on the inside surface of the
liner, because this could also create a location for droplet re-entrainment.
For borosilicate-block lining systems, such as the PennGuard, every effort
must be made to ensure the proper radial alignment of the blocks and to strike
the adhesive/mastic flush with the inside surface of the liner. The surface of the
block should be clean and free from smeared adhesive/mastic material. If the
maximum projection or offset between bricks on the interior surface of the liner
exceeds 1/8 in (3.1 mm), re-entrainment may occur from the liner surface.
Extruded or depressed adhesive/mastic material can be a significant source of
liquid re-entrainment back into the gas flow.

2-30

2.6 Re-entrainment
The liquid on the duct and liner surfaces produced by deposition and
condensation flows in the form of film or rivulets is governed by gravitational,
surface-tension, and gas-shear forces. The gas-shear force can shear the liquid off
of the surface, causing re-entrainment. This phenomenon is the most frequent
source of SLD and fallout of liquid droplets in the vicinity of the stack.
Re-entrainment most frequently occurs because of one or more of the following:

High gas-velocity shear near rough surfaces with liquid deposits.

Instability of the liquid film (waves and ripples) on vertical stack-liner


surfaces. This instability occurs because, as gravity is pulling the liquid film
or droplet down, the gas-drag forces are pulling the liquid film or droplet
upward.

Surface discontinuities and protrusions that disrupt gas and liquid flow
locally.

Vanes and baffles that cause gas-flow separation, recirculating liquid flow,
and re-entrainment.

Gas-flow patterns that drag liquid along the surface to re-entrainment sites.

Strong vortex patterns that can pick water up in the core of the vortex from
horizontal or vertical surfaces.

Re-entrainment takes place at locations such as

Duct walls, roofs, floors, and stack-liner surfaces

Dampers, damper-blade guides, and trusses

Thermal expansion joints

Trailing edges of turning vanes

Duct-junction corners

Breeching-duct area changes

Stack floor

Stack-liner surfaces, particularly in the entrance region and at the


intersection with the breeching duct

Internal solids buildup or scaling on surfaces

Stack-liner contraction sections

Stack choke surfaces

Liquid re-entrainment at these locations will be in the form of large droplets


(3006000m); most will be discharged at the top of the stack, if not prevented
by a combination of favorable gas-flow-path design and an optimized liquidcollector system. Droplets over 100 m typically do not completely evaporate
outside of the stack before they reach the ground level.
2-31

2.7 Recommended Liner-Gas Velocities


Wet stack designers should select the liner diameter based on liquid-collection
considerations. Selecting the liner diameter so that the gas velocity is below the
critical re-entrainment velocity increases the opportunity for liquid to be collected
within the stack rather than being emitted with the gas stream. The critical reentrainment velocity varies depending on the liners surface roughness, material,
and quality of construction. Surfaces with high levels of discontinuities and
roughness, such as those of a brick liner, will be more likely to re-entrain liquid,
in comparison to smoother surfaces, such as those of an alloy or an FRP liner.
Therefore, the liner diameter will depend upon the liner material selected. For
instance, brick liners will require a larger-diameter liner than an alloy or FRP
liner with the same gas-flow rate. Although there are many economic drivers for
minimization of the stack-liner diameter, it must be clearly understood that the
primary controlling parameter for effective wet stack operation is that of
minimizing the potential for SLD.
The stack-liner velocity recommendations in the original EPRI Wet Stack Design
Guide were developed on the basis of laboratory testing of different liner-material
surfaces. These materials were tested in a vertical wind tunnel under ideal
conditions, and their surface behaviors were characterized without weld beads or
other surface discontinuities. Since this testing was performed, a large amount of
valuable field experience has been obtained. On the basis of this experience, the
recommended velocities have been slightly reduced to accommodate practical
variabilities in the quality of the field installation, such as weld-bead heights on
alloy liners and joints in FRP liners. The currently recommended liner-gas
velocities are presented in Table 2-1. These slightly reduced values also provide
the plant some margin to account for increases in the flue gas flow rate as a result
of changes in fuel source, increases in plant efficiency, and/or future increases in
plant output. On the basis of this experience, it is now recommended that alloy
and FRP liners should be designed for a maximum velocity of 55 ft/s (16.8 m/s),
whereas brick liners (which inherently have a less smooth surface) should be
designed for 45 ft/s (13.7 m/s). The maximum recommended liner velocity for
borosilicate block is 60 ft/s (18.3 m/s). This recommendation takes into account
the significant increase in the effective surface area afforded by the closed-cell
surface structure of the material and the resulting increased surface-tension forces
holding the liquid to the material.

2-32

Table 2-1
Recommended Stack-Liner Velocities for Wet Operation

*Laboratory testing of coating material is recommended to finalize liner velocity for favorable wet
operation.

A number of different liner-coating materials have been suggested for wet stack
operation, ranging from gunite to spray-on polymers. Most of these coatings
have not been evaluated for determination of their recommended liner velocity
for favorable wet operation. The favorable velocity for coatings is highly
dependent upon the smoothness and wetting nature of the material after it has
been applied within the liner. Laboratory testing of the coating system should be
performed before a liner velocity is finalized for a specific material and
application. A velocity of 55 ft/s (16.8 m/s) is recommended for materials that
result in a smooth, discontinuity-free surface within the liner.
2.8 Stack Liquid Discharge
Stack liquid discharge (SLD) can occur at any time, even from a well-operating
stack. Disturbances in the operating conditions of the plant and/or absorber as
well as ambient weather conditions can initiate the start of an SLD event.
However, SLD is only a problem if the droplets are large enough (>100 m) to
be detectable at ground level near the stack.
The amount of liquid discharged at the top of the stack is a result of gas- and
liquid-flow processes that take place in the ductwork and stack system between
the discharge of the absorbers and the top of the stack liner. The source of the
liquid discharge can be categorized by the liquid-flow process and by the droplet
sizes as follows:

2-33

Liquid droplets carried from the absorber to the top of the stack liner by the
gas stream without depositing along the gas-flow path. Only small droplets
(10100 m) travel through the ducts without complete deposition. This
means that most conventional scrubber designs, with horizontal ducts
connecting the absorber to the stack, will result in deposition of most large
droplets on the duct surfaces.

Droplets re-entrained from liquid deposition and condensation on duct and


liner surfaces that can be discharged. These are large drops, and they are
usually the major contributors to the SLD.

Droplets formed by bulk condensation. These represent a large liquid-flow


rate but are very fine droplets (<10 m), which have a negligible effect on the
SLD detectable at ground level.

Liquid deposited and condensed on the liner, which flows upward in those
areas of the stack where the gas velocity is higher than the flow-reversal
velocity for the liner material. The droplet sizes re-entrained at the top of the
liner, where velocities are highest, can be quite large (3006000 m),
depending on the liner-top geometry. The upper section of a stack choke is
usually in this mode of upward liquid flow.

The quantity and location of the fallout at a given plant is a function of the
droplet-size distribution of the discharged liquid and the local atmospheric
conditionssuch as ambient temperature, wind, relative humidity, and
turbulence level. The very small droplets of the bulk condensation make the gas
plume white and visible. The plume is cooled as it mixes with the surrounding
air. This cooling results in condensation on the small droplets in the plume.
These droplets are typically 110 m in diameter and therefore do not fall to the
ground and cannot be detected. The mixing with the drier air away from the
stack causes droplet evaporation. The balance between the cooling and mixing
processes defines the length of the visible plume. Therefore, it is a function of the
ambient air temperature and the air relative humidity.
One of the major objectives in developing an effective wet stack design is to limit
SLD to a minimum acceptable level. The goal is to limit the droplet size that will
exit the stack. If the droplets are small enough, they will evaporate before hitting
the ground. However, if the droplets are large, they will land on nearby surfaces
such as plant structures, equipment, and cars. The wet stack survey indicated that
the fallout of liquid droplets, if noticeable, usually occurs within a half-mile
radius of the stack.

The most common cause of an unexpected SLD is a problem with the


WFGD mist-eliminator system, such as blade pluggage, fouling, and
washing issues. Liquid-collection system pluggage could also be occurring.
It is for this reason that a method for monitoring the liquid-collection
drainage flows is often recommended.
It is also common for newly commissioned liners to experience some level of
SLD for the first few weeks of operation, until the inside surfaces of the liner
have been seasoned and become fully wetted.
2-34

If SLD is experienced, note the time of day, location of the fallout relative to the
stack liner, and local weather conditions. If the location of the fallout is known,
rolling aluminum foil out on the ground in 26-ft (0.61.8-m) squares for a
period of time sufficient to collect numerous drops will be extremely helpful.
Both the droplet size and the quantities of droplets can be obtained from the foil.
This information, along with information on plant operations at the time of the
event, will be critical in evaluating the source and possible cause of the rainout
event. If SLD is a continuous problem, some stack sampling may be required.
Knowing the sizes of the droplets entrained within the gas flow inside of the liner
can be very helpful in identifying the source (or sources) of the discharge.
Additionally, an inspection of the liquid-collection system should be performed
as early as possible. The inspection should be performed and photo-documented
before anything inside of the wet ducts and stack are disturbed in any way. Often,
clues to the source of the discharge problem can be as simple as a deposit on a
wall or marks on the liner surface.
2.9 Plume Downwash and Icing
2.9.1 Plume Downwash on Stack
A cross wind at the top of the stack will deflect the plume from its vertical path.
As the cross wind increases, the plume may become partially entrained into the
low-pressure region behind the stack windscreen. This phenomenon is known as
plume downwash.
Designing a liner for favorable wet operation will decrease the potential for SLD.
However, such a design increases the potential for plume downwash at elevated
wind velocities (Figure 2-18). During downwash episodes, saturated flue gas
comes in contact with the liner extension, stack hood, and stack shell. This
contact can lead to deterioration of the stack construction materials because of
exposure to acid in the flue gas. It also increases the potential for ice formation
on the top of the stack. The potential deterioration problem is a long-term
situation caused by continual exposure to SO2, which has been evidenced in
recent years by the deteriorating conditions of tombstones, stone buildings, and
the top of stack shells. The icing problem can occur at below-freezing conditions
all winter long, every winter, and creates a potential danger to people and
property.
The most important fluid-dynamic parameter is the momentum ratio between
the stack exit flow and the cross wind. This ratio defines the relative influence of
each flow on the other and is determined by the densities and velocities of the
flue gas and the wind. This momentum ratio (MR) is defined as:

As the ratio of vertical plume momentum to horizontal wind momentum falls


below a value of about 2.0 for a free-standing single-flue stack, the plume may
become partially entrained in the vortex patterns that are formed on the
2-35

downwind side of the stack. To reduce the frequency of downwash, the


momentum ratio must be increasedeither by reducing the liner diameter or by
installing a choke at the top of the liner.

Figure 2-18
Plume Downwash

The second most important ratio is the shell outer diameter to the liner inside
diameter at the top of the shell. A larger shell- to liner-diameter ratio results in
increased downwash potential at the same momentum ratio.
Plume downwash is most likely to occur during reduced unit load operation, in
which stack-discharge velocities are reduced, and during high wind conditions.
Plume downwash is least likely to occur under the opposite conditions of high
unit load operation and low wind velocity.
For a multiple-flue stack, the equivalent momentum ratios for initiation of plume
downwash are higher, and they vary with the relative wind direction. This
difference in momentum ratios occurs because the shell- to liner-diameter ratio is
larger for multiple-flue stacks. The reduced static pressure in the wind vortices
generated off the stack shell and liner extensions can draw the flue gas into a
downwash pattern along the stack shell. In the process, the saturated flue gases
that are drawn into the vortices come into contact with the roof and sides of the
stack liner and shell.
The interaction of a plume with the cross wind is a function of the following
parameters:

Liner-extension diameter

Shell diameter

Liner-extension length

Roof shape

Number of liners
2-36

Wind direction, velocity, and air density at the top of the stack

The annual frequency of wind occurrence as a function of wind velocity

Liner exit-gas velocity and gas density

Plume buoyancy due to flue gas temperature

Ambient air temperature

The momentum ratio (MR) is the most important fluid-dynamic force ratio that
controls plume downwash. The individual parameters within this ratio have the
following importance for wet stack plume downwash:
1. Flue gas density (flue gas) is almost constant for saturated flue gas at about
0.066 lb/ft3 (1.06 kg/m3) at sea-level ambient pressure.
2. Wind density (wind) varies with ambient temperature at sea-level ambient
pressure from 0.090 lb/ft3 (1.44 kg/m3) at -20F (-29C) to 0.071 lb/ft3
(1.14 kg/m3) at 100F (38C) (an approximately 25% spread).
3. Reduced ambient pressure for plants at high elevations above sea level
reduces density by 1020%, but it changes both density values by the same
amount.
4. The flue gas discharge velocity (Vflue gas) is at the highest stack elevation and is
set by the liner or choke discharge diameter and the low-load to high-load
flue gas volume flow rate. This value can range from 20 to 120 ft/s (637
m/s) for a wet stack plume discharge, including partial-load operation.
The wind-velocity magnitude, direction, and frequency of occurrence varies
significantly at different geographical locations, but this information is usually
available at weather stations and airports located near existing or new powerplant sites. The information needed is usually referred to as the "wind-rose" data
in graphical or tabular form, and it includes wind velocity and frequency as a
function of direction. The quantity of cumulative data is usually supplied on a
yearly basis, but the data is also available on a monthly (or more frequent) basis if
time of year is important.
When reviewing the wind data, a judgment must be made when selecting a
worst-case wind magnitude and cumulative frequency of occurrence (to eliminate
unusually high windstorms of short duration that seldom occur).
Items 1, 2, 3, and 5 above can usually be set for a specific plant site. Item 4, the
flue gas discharge velocity, must then be selected to satisfy the requirements for

Liquid re-entrainment from the inside of the liner

Liquid collection and drainage

Plume downwash

For plume-downwash considerations, design decisions must be made regarding


the extent of allowable downwash and the cumulative amount of time that
downwash occurs over the period of a typical year. Factors to be considered are:
2-37

Unit load schedule versus time

Wind-rose data

Geometry of the top of the stack (liner inner diameter, liner-extension


height, shell outer diameter, roof or cap geometry)

Extent of downwash allowed:

None (may not be achievable)

On liner extension

On roof or hood*

On stack shell

Cumulative time of downwash (per year):


-

Zero (may not be achievable)

Less than 2 days (0.5%)

24 days (0.51%)*

48 days (12%)

*The extent of downwash and the cumulative time of downwash marked are suggested
values at which to start the evaluation of downwash and its effect on stack-discharge
velocity.

If the resulting stack-discharge velocity is unreasonably high, the downwash


criteria may have to be relaxed to allow more downwash and longer cumulative
time duration per year.
A computational fluid dynamic (CFD) flow-model study of the stack top for a
specific plant and geometry is required for an accurate evaluation of the extent of
plume downwash and the cumulative time of downwash. Presented below are
some approximate guidelines that can be used for the preliminary evaluation of
the expected degree of plume downwash and for the selection of starting
geometries of the stack top for CFD model selection.
The MR ranges presented below represent a variation from intermittent
downwash, which is nearly continuous for the low MR and intermittent for the
larger MR. Below the lower MR for each range, the downwash is continuous.
Above the upper MR for each range, the downwash is almost unnoticeable. The
following two MR ranges are discussed below:

MR1 MR range for intermittent downwash to roof or hood

MR2 MR range for intermittent downwash to the shell

For all stack-top geometries discussed below, the stack-shell diameter should be
as small as practicable to contain the number of liners.

2-38

Single-Flue Stacks
1. For an angled hood and no liner extension:
-

MR1 = 35

MR2 = 23

2. For a flat or angled roof and a short liner extension of 0.5 liner diameter:
-

MR1 and MR2 expected to be lower (1020%)

Two-Flue Stacks
1. For a flat roof plus a liner extension of about one liner diameter:
Best wind direction (prevailing-wind direction in line with the axis of the
two flues)
-

MR1 = 1.52.0

MR2 = 1.11.5

Worst wind direction (prevailing-wind direction perpendicular to the axis of


the two flues)
-

MR1 = 3.55

MR2 = 34.5

2. For a minimum liner extension of 0.75 liner diameters, the above MR values
will remain the same for the best wind direction, but the above MR values
for the worst wind direction will rise by 1020%.
3. For a taller liner extension of 1.5 extension diameters, the above MR values
for the best wind direction will remain the same, but the above MR values
for the worst wind direction will drop by 1020%.
4. A chamfered stack top (2 sides) will slightly increase (by1020%) the best
wind direction MR values, but will reduce (by 3050%) the worst wind
direction MR values.
Three- and Four-Flue Stacks

The best wind direction MR values are expected to be similar to those for
single-flue stacks.

The worst wind direction MR values are expected to be similar to those for
two-flue stacks.

In most locations, there is a prevailing-wind direction at which the highest wind


velocities occur for longer periods of time. For single-flue stacks, this prevailingwind direction has no effect on the extent or duration of plume downwash.
However, for multiple-flue stacks, it will be advantageous to orient the multipleflue arrangement so that the prevailing wind is aligned with the best wind
direction, which gives the lowest MR values. This type of alignment will result in
2-39

the lowest cumulative time for plume downwash for a specific combination of
flue gasdischarge velocity and stack-top geometry.
These plume-downwash guidelines can be used to obtain preliminary estimates
of the stack-top design to limit the extent and duration of plume downwash:
1. Select plume downwash criteria at a specific unit load.
2. Select a basic stack-top design except for flue gasdischarge velocity.
3. Obtain the wind-rose data and select reasonable design wind velocities.
4. Select the appropriate MR-guideline value that is compatible with steps 1
and 2.
5. Calculate the corresponding flue gasdischarge velocity, flue gasdischarge
area, and the preliminary stack-top geometry.
6. Compare the results in Step 5 to values compatible with good plume
dispersion, stack liquid collection and drainage, and construction cost.
7. Repeat steps 1 through 6 until satisfactory results are obtained for all
important criteria by making appropriate compromises.
The final stack-top geometry should be optimized by CFD flow-model studies,
particularly for multiple-flue stacks for which downwash-minimization is
important.
2.9.2 Stack Top Icing
Whether ice forms on the top of the stack depends on the temperature of the
stack surface, the temperature of the mixture of saturated flue gas and cold
ambient wind, and on whether water vapor will condense out of the mixture. Ice
formation is most likely at plants where below-freezing temperatures are
common and where they last for extended periods of time.
The locations for potential ice formation are discussed below:
The hood for a
single-flue stack

The hood is the first element exposed to plume downwash. A hood


for a single-flue stack is usually connected to the stack, liner, and
liner extension, which helps to keep it warm. When ice forms on
an outward-sloping hood, it usually forms off the upwind edge
and/or in the downwind half at two locations about 45off the axis
of the wind direction. The ice deposits formed are similar to those
on the roof valley of a house and can be quite large.

The liner extension


for single or multiple
liners

The liner extension is the second element exposed to plume


downwash. If the liner extension is uninsulated, the surface should
always be maintained above freezing temperature. If so, no ice
could form, but some condensation might occur that would then run
down to the stack roof. An insulated extension might have some
ice formation.

2-40

The roof of a
multiple-liner stack

The stack roof usually does not directly touch the liners, which
permits expansion and movement of the liners. The roof can
receive heat from the stack annulus. Liquid must be drained from
the roof area. Preferably, drains should be located on the inside of
the annulus to keep them from freezing.

Railings

The metal railings at the top of the shell will quickly cool to the
temperature of the mixture of the flue gas and the ambient air, and
they will be the first surfaces to become iced. In other words, these
railings are the most likely places for ice to form.

Platform near top of


stack

A metal platform on the outside of the stack near the top can also
be exposed to low concentrations of flue gas. Because it is metal
and is exposed to the ambient- and flue gasmixture temperature, it
may also experience occasional ice buildup. Such platforms are
the second most likely places for ice to form.

Shell

If ice forms on the shell, it can be expected to be on the downwind


side of the stack. This condition occurs when visible downwash can
be seen on the downwind side of the stack shell.

Icing can result in significant ice buildups along the top of the stack. These ice
buildups can dislodge and fall, potentially causing serious damage to people and
equipment on the ground.
The potential for icing can be reduced by employing the following steps:
1. Select a stack-liner discharge velocity that minimizes plume downwash over
the expected operating range of the unit at the existing local wind conditions
and that is consistent with other design objectives.
2. During cold ambient temperature conditions with high winds, run the unit at
near full load. Employing this operating procedure may be natural under
such conditions, because more power is consumed in below-freezing weather.
3. Use heated annulus air or electrical heating elements to heat the stack hood,
roof, or other areas where ice forms on the top of the stack.
4. Do not use insulation on sloped rain hoods constructed of a corrosionresistant material.
5. Use a parapet wall instead of metal railings around the top of the liner shell.
A parapet wall is usually an extension of the concrete shell above the roof.
There is normally less icing on a concrete parapet than on a metal railing,
and the parapet wall will contain ice that could fall from the liner or liner
extension.

2-41

Section 3: Practical Design


Recommendations
3.1 Introduction
Having presented the basic understanding of the fundamentals of gas and liquid
flows within wet duct/stack systems in Section 2, we can now turn our attention
to the use of these fundamentals for practical applications. Essentially, all new
and retrofit WFGD units are typically designed for wet stack operation. During
the initial design phases of the project, a feasibility study is normally conducted
to determine such factors as layout, materials of construction, economics, and
construction time. The Revised Wet Stack Design Guide addresses what needs to
be considered during the feasibility-study phase of a new or retrofit wet stack
project. For example: Will the existing stack-liner exit velocity be acceptable for
wet stack operation? Can the existing stack be modified, or will a new stack be
required?
Wet stacks have been designed and operated effectively at new WFGD
installations for many years, thanks largely to the guidance provided to the
industry by the original EPRI Wet Stacks Design Guide. Although most of the
design criteria presented in this original guide is still valid, experience with its
practical implementations has shown that some revisions and updates are
warranted. In each situation, the rules affecting a systems favorability for wet
operation are the same. However, there are application-specific issues that need
to be addressed. These issues are discussed in detail in the following sections of
this guide.
The recommendations presented in this revised design guide will both help the
designer make decisions that will affect wet operation and provide the necessary
background information to understand why the recommendation is being made.
Typically, the wet duct/stack liquid-collection and drainage-system arrangement
is prepared by a flow-modeling laboratory skilled in the art of wet stack design. It
is imperative that these modelers be brought into the design process early enough
so that their input on the overall favorability for wet operation can be
implemented before the system arrangement has been finalized.

3-1

3.2 New Wet Stack Designs


WFGD installations incorporating a new wet duct/stack system have the
advantage of being designed for favorable wet operation from the start. Using the
recommendations presented in the following sections of this revised design guide
should result in a wet stack system that operates well, with a minimal potential
for SLD. New wet stacks can be designed using materials that are acceptable for
the wet corrosive environment; can incorporate design characteristics that
minimize the potential for droplet re-entrainment back into the gas flow; and can
be designed to operate within the range of gas velocities recommended for the
stack-liner material used.
Although the processes controlling the design of an effective wet duct/stack
system are well understood, much of the re-entrainment potential for a particular
unit is site-specific, and a physical-flow-model study for each installation is
highly recommended. Such a study will evaluate liquid re-entrainment and liquid
collection in the ductwork and liner, based on the design of the actual unit and
anticipated gas velocities.
3.3 Retrofit Wet Stack Designs
Stacks that operate under reheat, partial bypass, or full bypass generally have been
designed for a higher gas velocity in the liner than new wet stacks. When
operating the stack wet, by eliminating reheat, the gas velocity in the liner
decreases and the density increases. Wet operation results in system-pressure-loss
decreases of 510%, compared to the system-pressure loss with a reheat
temperature rise of 3060F (1733C).
Operating an existing liner that is converted to wet stack operation can often
result in a gas velocity in excess of the recommended liner velocities previously
presented in Table 2-1 for wet liners; this higher gas velocity makes liquid
collection within the liner more difficult. Dry stacks were typically designed for
higher gas velocities. Existing liners that are converted to wet operation and have
velocities that exceed these recommendations may require that the majority of
the liquid be collected in the absorber top and the absorber-outlet ductwork. The
use of additional and more-extensive liquid collectors in the absorber top and the
absorber-outlet ductwork will need to be evaluated. Droplets entrained within
the gas flow will impinge upon the liner, forming a liquid film. Because of the
unfavorably high gas velocity in the liner, a large fraction of the liquid may
become re-entrained back into the gas flow. If the wet-gas velocity is higher than
the flow-reversal velocity for the liner, all deposited and condensed liquid will
flow upward to the top of the liner and be re-entrained there.
If gas velocities in the ductwork are also high, liquid collection in the ductwork
may also be difficult; therefore, the gas velocities in the absorber-outlet ductwork
should be evaluated as well. Ductwork for reheated or bypass gas is typically sized
for 60 ft/s (18 m/s) or higher. Ductwork for wet stacks should be sized for 50 ft/s
(15 m/s) or lower to minimize the potential for liquid re-entrainment back into
3-2

the gas flow. Critical re-entrainment velocities in the ductwork are difficult to
determine and are not available.
If the liner velocity is lower than the range of recommended maximum values,
the liquid re-entrainment rate is lower and the need for additional liquid
collectors is reduced.
A single-stack liner that serves multiple FGD modules may produce large
volumes of liquid for drainage. Drainage-pipe diameters need to be sized
accordingly. If the liquid collected from multiple FGD modules is returned to
the FGD process, the volume of liquid may be too large to be returned to a single
FGD module without significant changes to its operational characteristics.
A flow-model study should be performed to evaluate liquid re-entrainment and
liquid collection in the ductwork and liner, based on the anticipated gas
velocities. If adequate liquid collection cannot be accomplished at the anticipated
velocities, a new liner and/or ductwork sized for a lower gas velocity should be
considered.
3.4 Wet Duct/Stack Operation
Initially, water enters the duct and the stack by droplet and water carryover, with
the saturated flue gas passing through the mist eliminators of the scrubbers. The
liquid accumulates on surfaces by deposition of liquid droplets and by vapor
condensation from the saturated flue gas.
Mist-eliminator carryover is usually characterized by fine droplets that can pass
through the mist eliminator without being collected and by larger droplets reentrained from deposits and during washing. The finer droplets will follow the
gas flow and will generally not be collected within the system before exiting the
stack. However, because these droplets are very small, they will evaporate on the
outside of the liner before reaching the ground. The larger droplets (both in size
and quantity) re-entrained from partially plugged mist eliminators and during
mist-eliminator washing, will generally be deposited on the surface within the
ductwork and the lower liner.
Most of the liquid suspended in the gas will be deposited on the walls and
internal structures (turning vanes, dampers, trusses) by impingement, caused by
the departure of the trajectory of a liquid droplet from the streamline flow in a
curved gas-flow field. Some deposition also occurs because of turbulent
deposition, caused by the turbulence-induced motion of small droplets in a
direction perpendicular to the main gas-flow direction. The liquid on the walls
and internal structures will be in the form of films, rivulets, attached droplets,
and (in some cases) large pools on horizontal surfaces. The gas-shear force can
drag the liquid through the system (even upwards against gravity and wall
friction) if the gas velocities are high enough; it can also shear off and re-entrain
droplets from the liquid flowing on the walls and internals. Some small droplets
may negotiate the gas-flow path from the scrubber outlet to the top of the stack
without contacting a solid surface.
3-3

During normal operation, the largest contributor to the liquid flow on a liner
surface is condensation, rather than deposition of the droplet carryover from the
mist eliminators. The condensation on the liner cannot be measured in a
laboratory-scale model of the stack. Therefore, analytical calculations are required
to define the amount of condensed liquid-flow rate in the stack.

There are two types of condensation mechanisms: 1) condensation on the


duct walls and liner surface due to heat transfer in the boundary layer and
2) condensation in the bulk of the gas due to pressure changes and
adiabatic expansion from elevation changes. Both of these mechanisms
can produce significant liquid by condensation. A basic description of
these mechanisms is given in Section 2.3, "Sources of Liquid in Wet
Stacks."
Liquid condensation occurs continuously on all wet duct and stack-liner surfaces
because the flow gas is saturated and the liner's inside surface temperatures are
lower than the gas dew-point temperature. But the rate of condensation per unit
surface area can vary significantly, depending on geometry, materials,
temperatures, and wind speed.
The duct and stack condensation rate cannot be measured in experimental flow
models. A 2-D condensation analysis is satisfactory for calculating the
condensation rate that accounts for stack geometry, mass flow, heat loss, gas
psychometric conditions, and the ambient atmospheric-pressure variation along
the height of the stack. Assuming axial symmetry for the stack, all the variables
can be described or calculated as a function of vertical height with computer
programs developed for this purpose as design tools.
The major conclusions drawn from the results of these calculations are as follows:

The bulk condensation in the liner is a function of stack height.

For brick liners, the pressurizing-air leakage through the liner increases the
bulk condensation by additional cooling. Thermal condensation also
increases because of the colder annulus air used for pressurization.

The wall condensation is directly proportional to the temperature difference


between the gas and the ambient air.

With the exception of borosilicate block (which has an inherently high


insulating value), the choice of liner material only marginally affects the wallcondensation rate, because the thermal resistance of the liner is a small
fraction of the total thermal resistance between the flue gas and the ambient
air.

The wall condensation may be reduced by a factor of about four by insulating


the outside of alloy or FRP liners with a 2-in (0.05-m) mineral wool layer.
Increasing the insulation thickness from 2 in (0.05 m) to 4 in (0.1 m) will
result in a small reduction in condensation rate. Therefore, more than 2 in
(0.05 m) of insulation is seldom required. Insulation does not eliminate the
liner condensation, but it can reduce it significantly.
3-4

A wind-speed increase of 2040 mph (918 m/s) increases the condensation


rate by only a few percentage points, because the convective-heat transfer
levels off above 20 mph.

Having more than two liners in a shell increases the total wall-condensation
rate by about 5%. This increase is mainly attributable to a higher
condensation rate on the liner extension above the shell.

The condensation in the ductwork is primarily thermal wall condensation;


adiabatic condensation (due to pressure reduction with elevation change) is
negligible. Therefore, the rate of condensation here is not as important as in the
stack, because most of the condensed liquid and the deposited droplets can be
collected and drained out before they enter the stack.
The results of the condensation calculation are used to design the liquid
collectors and to plan for the liquid-disposal system. The estimated ranges of the
liquid-flow rates in the wet duct and stack of a typical 550-MW unit are listed in
Table 3-1. These numerical values help the stack designer estimate the amount
of liquid flow to be controlled and accounted for in a wet duct/stack system.
Table 3-1
Estimated Ranges of Liquid Flows in the Wet Duct/Stack of a Typical 550-MW
Unit
Source of Liquid at the Stack

Estimated Flows

Mist-eliminator carryover:

The vapor mass-flow rate in the saturated gas


3
11001800 gpm (250410 m /h)
stream is rather large. The total vapor content at
saturation temperatures115130F (46
54C)is equivalent to a liquid-flow rate of

The fine liquid mist (diameters <10 m) with


high-efficiency mist eliminators

0.31.5 gpm (0.10.3 m /h)

Duct-wall condensation

15 gpm (0.21.1 m /h)

Stack-wall condensation

16 gpm (0.21.4 m /h)

Bulk condensation in the stack

68.4 gpm (1.41.9 m /h)

Deposition of liquid droplets

112 gpm (0.22.7 m3/h)

Droplets re-entrained from the walls (diameters


202,000 m)

010 gpm (02.3 m /h)

Liquid from stack drains

212 gpm (0.42.7 m /h)

SLD (a function of the duct/stack design and the


3
015 gpm (03.4 m /h)
effectiveness of the liquid-collectors installed)

3
3

All the droplet-producing sources are controlled primarily by the local gas
velocity. The two basic mechanisms for producing liquid droplets are the
shear force caused by the local gas velocity and the dynamic pressure of the
gas flow at local flow disturbances on the solid surfaces. The particle-size
range for re-entrained droplets is primarily 3006,000 m.
3-5

3.5 Absorber Design and Placement


There are many types of WFGD absorber designs in use around the world.
These include, but are not limited to, open-spray towers (with and without liquid
hold-up trays), dual-contact absorbers, jet-bubbling reactors, venturi scrubbers,
and horizontal gas-flow weir or duct-type absorbers. All absorbers incorporate
mist eliminators at their outlet to ensure that the gas flow exiting into the ducts
toward the stack or other downstream equipment has a minimum quantity of
suspended droplets.
As a matter of economics, most new WFGD installations incorporating a new
stack will locate the absorber and stack close together to minimize the length of
connecting ductwork. Some length of ducting is required for effective liquid
collection. For example, placing the liner directly on top of the WFGD absorber
module is not recommended, because droplets generated off the mist eliminator
during upset conditions will not be collected before exiting from the top of the
stack. However, in some situationsmost often in those in which a new WFGD
absorber is going to use an existing stackthere may not be sufficient real estate
in the vicinity of the existing stack in which to place the new absorber. In such
situations, the absorber will be located at a significant distance from the stack,
and a long duct run will be required to connect them.
After the gasses pass through an absorbers mist-eliminator section, they are
directed to the absorber outlet, which connects to the ductwork leading to the
stack. In the outlet hood, the flue gas is accelerated from the mist-eliminator
outlet velocity to the duct velocitytypically 50 ft/s (15.2 m/s) for wet operation.
Absorber outlets can be either horizontal or vertical, and either rectangular or
round in shape, depending on the design of the outlet ducting. Absorbers with
horizontal outlets tend to be rectangular, whereas absorbers with vertical outlets
tend to be round.
3.6 Wet Ductwork Design
The proper design of power-plant ducting is detailed in [4]. The design of a duct
that is to operate in a wet environment is basically the same as that for a dry duct.
However, some considerations specific to wet operation must be addressed.
Properly designed ducting with effective mist-elimination and liquid-collection
systems can result in a system that is very favorable for wet operation.

The seven primary droplet-producing sources in a wet duct are:

1. Mist-eliminators: The primary source of liquid in stacks is the misteliminator carryover from normal operation, partially plugged operation, and
washing cycles. If a WFGD retrofit to an existing unit is made with nonideal
duct/stack conditions and a wet stack is used, one needs to specify the best
possible mist eliminator. The quoted efficiency and carryover from mist
eliminators is typically much lower (0.020.04 gr/acf [4692 mg/m3]) than
the limited field data, which may go as high as 0.172.00 gr/acf (3904600
g/m3). The wash cycle of the mist eliminator can produce short periods of
increased liquid-droplet discharge. EPRI has conducted field tests on mist 3-6

eliminator carryover drop sizes and loading levels. EPRI has also conducted
tests on many mist-eliminator modules.
2. Surface discontinuities in duct walls: The liquid films of rivulets flowing in a
horizontal, inclined, or vertical passage will be partially entrained at surface
discontinuities, such as a step-up, a step-down, and sharp bends. The
amount of liquid entrained is a strong function of the local geometry and the
gas-flow disturbance caused by the discontinuity. Liquid-droplet production
at wall discontinuities has been observed in both laboratory models and field
wet stack systems, but it has not been measured quantitatively.
3. Thermal-expansion joints: These duct components represent a special
discontinuity because they usually have a capacity to store liquid. Thermalexpansion joints represent a potential liquid re-entrainment area in the ducts
and in the stack, although the expansion joint is often covered. Expansion
joints can act as liquid collectors on the walls, and they should have drains
installed to remove collected liquid to reduce the possibility of reentrainment. Expansion-joint drains are only practical at duct pressures near
ambient pressure downstream of ID fans. At locations where negative
pressure exists in the duct, traps are needed to prevent air leakage into the
duct. Other re-entrainment control methods may also be needed.
4. Internal duct components: Necessary internal components such as mounting
plates, louver-damper blades, and turning vanes collect liquid by inertial
deposition, and re-entrainment takes place on or downstream of these
components.
5. Special flow patterns: Gas-flow passages of power plants are designed to be
functional and low-cost, but they are usually aerodynamically crude. Flow
separations, reattachments, and secondary flows are encountered at several
locations in most power-plant duct systems. The liquid droplets and films
collect in separated zones, forming a large pool or a heavy stream that
entrains more easily than a thin film at lower gas velocity. An increase of gas
velocity from a load change, or the addition of an FGD module on-line, can
produce re-entrainment from unprotected pools.
6. Roughness of surfaces: Liquid entrainment by gas streams from rough
surfaces is proportional to the gas velocity and the roughness of the surface,
starting at some critical threshold velocity that varies with the surface
characteristics. Because large surfaces with thousands of square feet are
involved, the integrated effect of surface roughness is very important
particularly in the vertical stack flow. Discontinuities and other sources of
surface roughness can be controlled or eliminated by design and careful
installation, but practical acid-resistant liner materials have to be accepted
and installed as produced.
7. Fan washing: If ID fans are included downstream of wet scrubbers, they may
have an intermittent on-line wash system for cleaning deposits off of fan
impellers. When in use, this system will produce entrained droplets.

3-7

3.6.1 Guidelines for Selection of Geometry for Wet Ducts


Because expensive corrosion-resistant materials must be used to construct or line
wet ducts downstream of wet scrubbers, duct sizes will tend to be smaller than in
dry ducts, resulting in higher velocities. Therefore, special care should be given to
reducing duct-component pressure-loss coefficients by the use of lower-loss ductcomponent types and more aerodynamically designed shapes.
The following three objectives need to be achieved by the design and the
operation of a wet duct and stack system without reheat:

Minimum re-entrainment of the deposited liquid by the gas shear and


turbulence

Maximum deposition and separation of the liquid on the duct walls, baffles,
turning vanes, and stack liners

Good collection and protected drainage of the liquid from the duct and stack
system

In addition, special care is needed if an ID fan located in the wet portion of the
system is used. See Section 3.8.2 for guidelines on these components.

3.6.1.1 Duct and Stack Design Velocity Levels


The selection of wet duct design-area average-velocity levels are based on the
following considerations:

The effect of velocity and velocity head on pressure loss and the costs
associated with the fan-energy requirement

The upper limit of the velocity that starts to entrain liquid droplets from the
walls and wall-related discontinuities

The movement of liquid films, rivulets, and droplets along the wall because
of gravity and gas-drag forces

The incorporation of liquid collectors and drains to prevent entrainment of


liquid from surfaces and remove it from the duct system

The addition of liquid collectors will add a small amount of additional pressure
loss as a result of the contraction and expansion of the gas flow to get past the
liquid collectors.
For wet ducts with saturated gas flow downstream of WFGD modules, the ductdesign velocity levels are limited by high gas-velocity droplet re-entrainment
values. Suggested maximum design-area average-velocity values for different wet
duct design situations are presented below. The design-area average velocities
used can be any value less than these. It is necessary that liquid collectors and
drains be designed, optimized, and evaluated by an experimental two-phase
laboratory-model test for each installation.

3-8

1. 65 fps (20 m/s): This maximum velocity level should be used if the following
duct features are present:
-

Any duct with an upward slant of less than 30, and in which most of the
duct is horizontal or sloped downward. Some sections with vertical,
upward, smooth duct slants (6090) could also be included if the
velocity for the stack design is compatible.

Duct components are less aerodynamic than those of Item 1 above,


but are still reasonably good.

A few internal trusses, beams, and surface discontinuities are allowed


near the stack, but no louver dampers can be near the stack.

Ducts have some restrictions on vane extensions to minimize cost,


but no restrictions on liquid-collector design.

2. 50 fps (15 m/s): This maximum velocity level should be used if the
following duct features are present:
-

Ducts have several or numerous sections that slant upward by 30


60angles from the horizontal.

Duct components are aerodynamically crude and include sharp


corners.

Ducts have no restrictions on internal trusses, gussets beams, surface


discontinuities, expansion joints, or louver dampers.

Ducts have some restrictions on vanes, baffles, and liquid-collector


installation.

If higher velocities must be used, then the re-entrained liquid could still be
collected by the installation of more extensive baffles and liquid collectors in the
stack bottom.

3.6.1.2 Duct-Component Selection


In the wet sections of a duct system, duct components should be selected to
promote droplet impingement, liquid-film collection, and drainage out of the
system. In addition, the duct system should be oriented to encourage liquid films
to flow naturally with the gravitational and gas-drag forces. If these forces are
opposed to each other, resulting in liquid flow back against the gas-drag forces,
droplet re-entrainment may occur. Liquid collectors and drains must be used to
intercept the flowing liquid films and guide the liquid to drain locations where
liquid is removed from the duct system. These liquid collectors must be
optimized by two-phase (gas-liquid) flow experimental-model tests for each
installation to insure satisfactory liquid-collection performance

3-9

The following duct components have desirable features for inclusion in wet duct
sections:

Horizontal large-angle bends for elbows, junctions, and manifolds with


several large vanes or baffles that can act as impingement and collection
surfaces

Vertical to horizontal elbows with large vanes that can act as impingement
and collection surfaces

Ducts that are horizontal or sloped downward, in which water will easily flow
on surfaces in the direction of the gas flow

Rounded inner corners on all bends

The use of a few large vanes in bends that can have vertical or inclined liquid
collectors on the outlet edges of the vanes. The vanes may also have extended
trailing edges with inclined trailing-edge collectors for improved collection
and drainage.

Expansion joints are natural collection regions and must have drains installed
to prevent overflow and re-entrainment. Drains must be designed to prevent
flow of air into the flue gas duct. The flexible portion of the expansion joint
must not protrude into the duct during plant operation.

The following duct components have undesirable features for inclusion in wet
duct sections:

Ducts that slope upwards on the floor or roof, because liquid can flow back
against the gas flow and be re-entrained.

Horizontal to vertical upward-vaned and unvaned bends, because the liquid


wants to flow downward as a result of gravity against the gas flow. This
situation cannot be avoided in the stack bottom. Special liquid collectors are
needed in these bends.

Internal pipe trusses, gussets perpendicular to the gas flow, louver dampers,
and structural beams located near the stack entrance where liquid can
impinge and be re-entrained as larger droplets.

Steps or discontinuities in the duct walls at flanges, dampers, and expansion


jointsparticularly near the stack.

Sharp corners on elbows, junctions, or manifold entrances where liquid can


be re-entrained.

Unprotected natural liquid-pool areas on the floors where gas flow can reentrain liquid from the pool surfaces. These areas can be protected to prevent
re-entrainment.

Location of dampers close to vaned bends can result in undesirable flow


interactions.

A large number of small vanes in bends, because they would require liquid
collectors on each vane.
3-10

In a field installation, liquid collectors and drains will include some of the
following types of devices, fabricated out of corrosion-resistant materials:

Gutters on walls and vanes constructed from angles and channels

Vane extensions and baffles constructed from plate material

Protection covers for water films or pools constructed from grids or


perforated plates

Drains made from pipes and properly located

To optimize the selection and design of these liquid collectors and drains
for a specific installation, experimental two-phase flow-model tests are
recommended. These tests will provide the designer with a high level of
confidence for successful operation without significant SLD.
3.6.1.3 Internal Structures
Ideally, the inside of the ducting used for wet operation should be as smooth as
possible, with a minimum of discontinuities to act as sites for droplet reentrainment back into the gas flow. However, some structures are often required,
and their negative impact on wet operation must be understood andto the
extent possibleminimized.
To the extent possible, all structural supports should be located on the exterior of
ducts used for wet operation. Internal bracing is often installed in rectangular
ducts on retrofit units that originally operated dry. Horizontal bracing is also
commonly used to support the tall side-walls of breech openings on FRP liners
with side-entry breeches. These internal support structures are not favorable for
wet operation because they provide additional impingement surfaces and reentrainment sites for liquid droplets. Liquid collectors should be designed for
these braces, which reduce the re-entrainment of large droplets.
Internal trusses, if necessary, should be designed to direct liquid collected on
them to the side-walls and duct floor (Figure 3-1). Trusses that direct collected
liquid to the center of the duct should be avoided. Internal truss work is typically
fabricated from 36-in (75150-mm) piping attached to the duct surfaces using
gusset plates. Liquid deposited or condensed on the trusses will flow around to
the back or downwind side of the truss and along the length of the trusseither
downward because of gravity, or across the duct if the support is horizontal.

3-11

LIQUID FILM
FLOW TO WALL
DUCT WALL

LIQUID FILM
FLOW TO
FLOOR

Good Truss
Arrangement

LIQUID FILM
FLOW TO
CENTER OF
DUCT

DROPLET
REENTRAINMENT
FROM CENTER OF
DUCT

Poor Truss
Arrangement

LIQUID FILM
FLOW TO
FLOOR

Figure 3-1
Good and Poor Internal Duct-Truss Arrangements for Wet Operation

3.6.1.4 Expansion Joints


Expansion joints are typically required in ducting to allow for thermal expansion
of the ducting as well as to isolate duct sections either structurally and/or
seismically. These duct components represent a special discontinuity, because
they usually have a capacity to store liquid. Thermal-expansion joints represent a
potential liquid re-entrainment area in the ducts. Expansion joints can act as
liquid collectors on the walls, and they should have drains installed along the
floor to remove collected liquid and reduce the possibility of re-entrainment.
Expansion-joint drains are only practical at duct pressures near ambient pressure
downstream of ID fans. At locations where negative pressure exists in the duct,
traps are needed to prevent air leakage into the duct. Other re-entrainmentcontrol methods may also be needed, such as covering the joint with a sliding
plate.
There are a number of different types of expansion joints in use in wet powerplant ducting. Duct-expansion joints can range from 824 in (0.20.6 m) in
width and are fabricated from a corrosion-resistant flexible rubber composite or
elastomeric material. The most typical joint utilizes a flexible boot, which forms a
cavity running around the perimeter of the duct. Some designs incorporate a
reverse boot designed to be nearly flush with the duct-liner surface. Experience
has shown that after a few months of operation, many of these boots have twisted
and inverted, with portions protruding into the gas flowresulting in increased
potential for liquid re-entrainment back into the gas flow.

3-12

Internal duct-support trusses are often located at the end of duct sections, and as
such, can be located immediately upstream and downstream of an expansion
joint. If the truss members are attached to the floor using a gusset plate, care
must be taken to ensure that the expansion joint is protected from a gas-flow
separation zone formed downwind of the gusset plate(s) (Figure 3.2). This
recirculation zone can be very intense, and it can easily rip water out of the
expansion-joint cavity if the joint is not drained. If such an arrangement exists, it
is highly recommended that a 34-ft (0.91.2-m)-wide metal plate be placed
over the joint. The plate should be centered immediately behind the gusset plate
for protection from the gas-recirculation zone. The plate will need to be
fabricated from a corrosion-resistant material and will need to be secured in such
a manner that it will not hinder the relative motion between the adjacent duct
sections.
3-4 DIA PIPE

GUSSET PLATE

DUCT FLOOR

RECIRCULATION ZONE
FORMED BEHIND GUSSET
PLATE

GAS FLOW

GUSSET PLATE

EXPANSION
JOINT

Figure 3-2
Internal Duct-Truss Gusset Plate

3-13

3.6.1.5 Man-Ways and Test Ports


Man-ways and test ports can act as sites for liquid re-entrainment back into the
gas flow if they are encountered by the moving liquid films within the wet duct
systems. To minimize the potential for liquid re-entrainment, liquid-collection
gutters may be required to direct the moving liquid films around these structures.
Because the size and location of these devices is substantially a function of the
local gas- and liquid-film flow patterns, the need for these devices can only be
identifiedand their design finalizedas part of the wet stack physical-flowmodel study required for the overall liquid-collection-system design.
3.7 Flow-Control Devices
Flow-control devices are often required to optimize the gas-flow patterns and
minimize the pressure losses within duct systems. This requirement is true for
wet as well as dry ducts. However, because of the high cost of corrosion-resistant
materials required for use in this environment, the need for flow controls must be
weighed against the economics of their installation. For example, the fabrication
and installation cost of a turning vane desired for pressure-loss reduction may be
higher than the expected lifetime operating-cost savings attributable to the
reduced system-pressure loss.
Modern WFGD installations with a simple straight run of ductwork between
absorbers and stack liner typically do not require flow-control devices. However,
flow controls required for ensuring favorable wet operation should be installed as
recommended. For example, a WFGD absorber-outlet turning vane designed to
ensure that the gas flow enters the liner in a way that will maximize liquid
collection and the generation of strong and robust secondary flows, which are
required for effective in-stack liquid collection, should be installed if
recommended.
Flow-control devices such as small vanes, baffles, and diverters are often
recommended for the elimination of recirculation and/or dead zones where liquid
films will collect and possibly re-entrain droplets back into the gas flow.
3.7.1 Flow-Control Liquid-collectors
All turning vanes and other flow-control devices should be evaluated regarding
their need for trailing-edge liquid-collectors. These devices are often directly
exposed to, and therefore collect the droplets entrained within, the gas flow.
These fine droplets collect as a film. They can then re-entrain back into the gas
flow as larger droplets from the downwind or trailing edges of the device. It this
potential exists, trailing-edge collection gutters will need to be designed and
installed to direct the collected liquid film off of the device and onto the duct
walls or floor for subsequent collection and drainage.

3-14

3.8 Fans
Occasionally, ID fans will be located between the absorbers and the stack of a
utility power plant. In this type of application, the ID fans will operate wet. Only
a few existing FGD systems were designed for wet fans, primarily because of the
high cost associated with the requirements for corrosion-resistant materials and
the inherent problem of rotor imbalance caused by scale buildup. For most
installations, wet fans should be considered only if they can be constructed of
corrosion-resistant alloy and if a spare fan can be installed to permit a regular
maintenance program of cleaning and rebalancing.
Wet ID fans will require periodic or continuous washing to prevent solids
buildup on the fan impellers, which can cause fan-rotor imbalance. All of the
wash liquid sprayed into the fan inlet leaves the fan impeller as droplets. Most of
the liquid will be propelled by the high centrifugal-force field to the fan scroll,
where it will deposit but immediately re-entrain as a result of the high gas-stream
velocity. Most of the washing liquid will escape the fan discharge as droplets
entrained in the gas stream, causing additional liquid load in the ducts and in the
stack. The fan-wash method can be optimized for effective washing with the
minimum wash-flow rate. Two-dimensional computer calculations are used to
optimize the fan washing. The duct and stack liquid-collection system must be
designed to account for the extra liquid carryover from the wet fan-wash system.
Fans are used with WFGDs for system pressurelevel control and for the
absorber-pressure loss. The fans are either ID fans or booster fans added to the
ID fan to produce the required pressure rise. Large centrifugal fans are used most
frequently, but axial fans are also used. These fans are mostly located upstream of
the absorber, but sometimes the design places them downstream of the absorbers.
3.8.1 High-Gas-Temperature Fans
The fans located upstream of the absorber are operating with high gas
temperatures and a lower gas density, which means that a higher-head-rise fan
must be used to produce the pressure loss of the FGD system. They are
downstream of the electrostatic precipitators or bag-houses and encounter only
limited fly-ash erosion. Axial fans are more sensitive to erosion than centrifugal
fans. No fan-liquid collectors are needed.
3.8.2 Wet Fans and Fan-Liquid Collectors
Fans located downstream of WFGDs operate with saturated gas at adiabatic
temperature level (wet fan). The head rise of the wet fan produces a higher
pressure rise with the lower temperature and a higher-density gas. The gastemperature rise produced by the fan reduces the relative humidity below
saturation in the ducts and the stack. The rate of liquid condensation in the stack
is reduced or eliminated completelyan important beneficial effect.
The carryover from the mist eliminators usually contains water droplets, fly ash,
and scrubber sludge, which cause solids buildup on the blades of the fan
3-15

impellers. Wet fans have to be washed to keep the impellers in balance. The fanwash spray has to be designed to work effectively with the specific fan-blade
shape and speeds, spraying periodically or continuously. The fan wash is less
sensitive to the fan-blade design of the axial fans.
To minimize the SLD, liquid collectors are needed. Fans are good liquid
separators, but the liquid has to be captured in the high-velocity gas flow in the
scroll, and in the ducts and the stack downstream of the fan. The centrifugal fans
can be equipped with scroll-liquid collectors especially designed for the high gas
velocity in the scroll. The ducts and the stack may require the usual liquidcollection system with wet fans, depending on the temperature riseabove
saturation remaining in the gas entering the stack.

If an ID fan is located in the wet portion of the duct system, the following
steps should be taken:

As much liquid should be collected and drained ahead of the fan as possible.

An on-line periodic fan-wash system should be installed and used to keep the
fan impeller clean and in balance. This system needs to be custom-designed
for the particular fan design.

Liquid collectors should be incorporated into the scroll and the downstream
ducts to collect and drain the fan-wash water.

The power input to the fan may provide a 520F (2.811.2C) reheat of the flue
gas if the liquid-droplet load and its total latent heat of evaporation is minimized.
This reheating will help reduce condensation in the downstream ducts and stack
liner.
3.9 Stack-Liner Design
A variety of factors should be considered in determining the overall geometry
configuration of a new wet stack. The stack-liner geometry should provide the
most economical configuration to accommodate the physical and design
constraints required. The stack-liner diameter should be based on the
recommended gas velocities to accommodate the liquid downflow on the liners.
Platforms, access system, and means of inspection must be considered in the
stack-diameter selection. Stack diameters can be plumb, single-tapered, multipletapered, or incorporate stepped diameter changes. Stack-liner geometry should
also be evaluated to minimize the design implications associated with wind and
earthquake loads. The major items for the preliminary stack design are discussed
in the following sectionsfrom the stack floor to the top of the stack. For
examples of stack geometry for a brick, FRP, and metal-lined stack, refer to
Figures 3-3, 3-4, and 3-5, respectively.
Steel, alloy, and FRP liners are typically, but not necessarily, limited to plumb or
constant diameters. Several different support methods are available for these
types of liners. For steel and alloy liners, a common arrangement is to support the
liner at or near the base or breeching entry. The liner is supported in
compression, and bumpers or stay rods are used to provide lateral support.
3-16

RAINHOOD

LINER
EXTENSION

PLATFORM

OBSTRUCTION
LIGHTS

REINFORCED
CONCRETE SHELL
BRICK LINER
OBSTRUCTION
LIGHTS

TEST & MONITORING


PORTS
PLATFORM

LINER BANDS (TYP.)

BREECHING DUCT

LIQUID COLLECTION RING

EXPANSION JOINT

LIQUID COLLECTION
GUTTERS
ABSORBER
OUTLET DUCT

LINER FLOOR

DRAINS

CONCRETE
PEDESTAL
PRESSURIZATION
FAN

FOUNDATION

Figure 3-3
Chimney with Constant-Diameter Brick Liner

3-17

LINER
EXTENSION

RAINHOOD
OBSTRUCTION
LIGHTS

PLATFORM

STAY RODS
REINFORCED
CONCRETE SHELL

LINER SUPPORT

STAY RODS

FRP LINER

OBSTRUCTION
LIGHTS

TEST & MONITORING


PORTS
PLATFORM

LIQUID COLLECTION
RING

LINER CAN JOINT (TYP.)

LINER SUPPORT

LINER EXPANSION
JOINT

EXPANSION JOINT

LINER SUPPORT
LIQUID COLLECTION
GUTTERS

ABSORBER
OUTLET DUCT

LINER BOTTOM
ENTRY ELBOW
DRAINS

LINER EXPANSION
JOINT

FOUNDATION

Figure 3-4
Chimney with Fiberglass-Reinforced Plastic Liner

3-18

RAINHOOD

LINER
EXTENSION

PLATFORM

OBSTRUCTION
LIGHTS
STAY RODS

REINFORCED
CONCRETE SHELL

LINER SUPPORT
ALLOY LINER

STAY RODS

TEST & MONITORING


PORTS

OBSTRUCTION
LIGHTS

PLATFORM

STIFFENERS (TYP.)

BREECHING DUCT

EXPANSION JOINT
EXPANSION JOINT
LIQUID COLLECTION RING
LINER SUPPORT
LIQUID COLLECTION
GUTTERS
ABSORBER
OUTLET DUCT

SLOPED LINER
FLOOR

DRAINS

FOUNDATION

Figure 3-5
Chimney with Metal Liner Typical for Alloy, Glass-Block, or Protective-Coating
Liner System

Tall steel and alloy liners supported in compression would need to be evaluated
for the cost of increased wall-plate thicknesses, versus providing a second support
level near the top of the chimney to support most of the liner in tension. An
expansion joint would be required above the lower support to accommodate
thermal movements.

3-19

FRP liners are generally supported with the major part of the liner in tension.
Supports are provided at two locationsone near the top and the second near
the breeching levelwith an expansion joint located above the breeching level.
Expansion joints in wet stack liners are a problem because they tend to collect
liquid, resulting in complete liquid re-entrainment at the expansion-joint cavities.
Therefore, expansion joints should be avoided or provided with liquid collectors.
If required, expansion joints should be located 1 to 1.5 liner diameters above the
roof of the breeching duct or bottom-entry elbow to ensure that the lower liner is
unobstructed for liquid collection and that liquid film flows to the liner-drainage
points.
Because of structural considerations to accommodate breeching openings and
design loads, brick liners can be plumb or tapered. Depending upon the
breeching-opening elevation and other factors, brick liners can be supported from
the stack foundation, from a concrete pedestal, or from a corbel support in the
concrete column. Chimneys with breeching openings that are less than 50 ft (15
m) above the foundation would probably use a brick liner that is supported
directly off the foundation. Breeching openings that are more than 50 ft (15 m)
above the foundation should consider utilizing a concrete pedestal to support the
lining. For tall brick liners with thick walls at the base, it is generally more
economical to use a concrete pedestal for high breeching entries.
Brick liners in high-risk seismic locations sometimes use a corbel-supported liner
to reduce the seismic loads. The liner is constructed in short segments, which are
supported from concrete ledges in the chimney, called corbels. This method of
construction is not recommended for wet stack applications. Liquid and flue gas
leaking through the brick liner can attack the concrete support system.
Additionally, it is difficult to pressurize the annular-space areas at each of the
corbel sections.
3.9.1 Liner Materials
Several types of liner materials and coating systems are available and have been
used successfully in WFGD environments. The advantages and disadvantages of
each system must be considered prior to the selection of the liner material or
coating. The following sections describe the most common materials that are
currently considered appropriate for wet stack applications. These materials have
previously been utilized with some degree of success. Other materials and
suppliers may be available and acceptable, and they should be evaluated on a
case-by-case basis. Operating conditions, design conditions, and economics all
play important roles in this decision. The advantages and disadvantages of several
different liner and coating-system options, along with estimated installation
costs, are presented in Table 3-2.

3-20

Table 3-2
Liner Material of Construction
Liner Material or
Coating

Advantages

Disadvantages
o

FiberglassReinforced Plastic

Good corrosion resistance


Easy to add liquidcollection devices

Maximum 300 F (149 C) (approx.)


gas-temperature exposure
Quality control during fabrication
Compressive strength limitations
usually requires two support levels
and expansion joint.

Good corrosion resistance

Borosilicate-Glass
Block

Good insulator (ductwork


and liner should not be
insulated)
Ability to retrofit to
existing steel liner systems

Cannot tolerate abrasion or


physical and mechanical abuse
Care in installing

Good surface for liquid


flow
Good corrosion resistance
Acid-Resistant Brick

Cost-effective
Liquid adheres to the
porous surface.

Surface discontinuities re-entrain


liquid.
Not recommended in high seismic
areas
Maintenance of liner accessories
Annulus pressurization

Protective Coating
on Carbon Steel

Fair corrosion resistance


Ability to retrofit to
existing steel liner systems

Surface preparation prior to


placement
Frequency of repair and
maintenance
Limited acceptable selections
Welding quality control
High material costs

Alloy C276

Excellent corrosion
resistance

Welding seams
Iron Contamination
Acid cleaning
Condensation

3.9.1.1 Brick
Acid-resistant brick is a solid kiln-fired brick made of clay, shale, or mixtures of
clay and shale, which conforms to the requirements of ASTM C980, Standard
Specification for Industrial Chimney Lining Brick. ASTM C980 defines three types
of brick, which are categorized by compressive strength, water absorption, and
acid solubility. Of the three types, Type III has the most stringent requirements,
allowing the least amount of water absorption and acid solubility, whereas Type I
3-21

allows the highest levels. Because of the stringent requirements, Type III brick is
the most difficult and expensive to manufacture. Because of availability
limitations for Type III brick, most brick liners for FGD applications are
constructed using Type II bricks.
The physical properties and chemical requirements for Type I, Type II, and Type
III brick, as defined by ASTM C980, Standard Specification for Industrial
Chimney Lining Brick, are presented in Table 3-3.
Table 3-3
Physical Properties and Chemical Requirements Of Acid-Resistant Brick1
Minimum Compressive Strength,
Gross Area (psi)

Designation
Type I

Maximum Water Absorption


by 2-h Boiling
(%)

Average
of 10

Individual

Average
of 10

Individual

Maximum
Average
Weight Loss
by H2SO4
Boil Test
(%)

8,500

8,000

6.0

7.0

20

9,000

4.0

5.0

12

1.0

1.5

Type II

10,000

Type III

12,000

10,000

Per ASTM C980, Standard Specification for Industrial Chimney Lining Brick 1 psi = 6.895 kPa

As previously stated, chimney brick can be manufactured from clay, shale, or a


mixture of clay and shale. Bricks manufactured from clay are called fireclay bricks
and are typically buff in color. Bricks manufactured from shale are called redshale bricks and are red in color. Red shale usually has a higher iron oxide
content than fireclays. The iron oxide acts as a flux, which permits a lower firing
temperature for shale than for fireclay. Fireclay and red-shale brick provide
comparable physical and chemical characteristics when fired at their respective
temperatures. If fireclay and red-shale brick are fired at the same temperature,
red-shale brick will typically be denser and have lower absorption characteristics
than fireclay brick. When properly manufactured, both will meet the
requirements of ASTM C980.
Brick liners use potassium silicate mortar that conforms to the requirements of
ASTM C466, Standard Specification for Chemically Setting Silicate and Silica
Chemical-Resistant Mortars. These inorganic mortars are resistant to most of the
strong acids present in the scrubbed flue gas from coal-fired power plants (except
for significant concentrations of acid fluorides or hydrofluoric acid). Potassium
silicate mortars are more resistant to sulfation than sodium silicate mortars or
other types of silicate-based cements.
Brick liners have been widely used by the power-plant industry. These liners
generally require the lowest initial capital-cost expenditure in comparison to
3-22

other types of liners. However, maintenance for a brick liner can generally be
higher than that for an alloy or FRP liner. Typical maintenance for a brick liner
would include repair of brick cracks, repair or replacement of bands, maintenance
of the pressurization fans, and repair of expansion-joint seals at the rain hood and
breeching.
Brick liners should be designed and constructed in accordance with the
recommendations of ASTM C1298, Standard Guide for the Design and
Construction of Brick Liners for Industrial Chimneys.
3.9.1.1.1 Mortar Joints
Careful attention must be paid to the quality of mortar joints in a brick-lined wet
stack. All joints must be struck smooth with the brick surface in order to
eliminate (to the extent possible) horizontal discontinuities on the liner surface.
Mortar protuberances and depressions will act as sites for liquid collection and
re-entrainment back into the gas flow.
3.9.1.1.2 Annular Space and Pressurization
Sufficient space should be provided between the inside of the concrete column
and the outside of the liners to allow for inspection and maintenance over the full
height of the stack. A minimum annular space of 2 ft 6 in (0.76 m) is
recommended. If platforms, ladders, and an elevator are located on the interior of
the stack, sufficient annular space needs to be provided for these items. The size
of the annular space should account for differential movements that may occur
between the concrete column and the liner as a result of wind, seismic, or thermal
expansion. Damage to platforms or the liner may occur if sufficient space is not
provided. Access and adequate clearance for clean-out doors and continuousemissions-monitoring equipment also needs to be considered. Ports should be
accessible and provide sufficient clearance between the column and liner to insert
EPA test probes. To minimize annular-space requirements, doors can be
provided in the column. Monorails can be used to assist in installation of the
probes. In order to ensure personnel safety, the annular space should be wellventilated, especially when an interior access system is used.
Annulus pressurization of the brick-liner stacks increases the rate of condensation
on the liner surfaces. Leakage of pressurizing air through the liner in older brick
stacks can significantly increase the condensation.
Because of the porous characteristics of brick and mortar, brick liners operating
under WFGD conditions should use an annulus-pressurization system. Annulus
pressurization is not needed for other liner materials. An annulus-pressurization
system consists of fans that provide ambient air under positive pressure into the
annular space of the stack, ensuring that permeation of the flue gas and liquids
through the liner and liner cracks is minimized. The pressure created by these
fans should exceed the maximum anticipated positive pressure inside the liner by
a minimum of 1 in Wg (249 Pa). Flue gas and liquids that permeate the liner
3-23

create a highly corrosive environment and can cause damage to the liner bands
and concrete column.
Airtight seals are required at all openings in the stack and liner to minimize air
leakage from the pressurization system. Air-locking chambers should be provided
at door locations to provide safe passage into and out of the chimney.
Pressurizing fans typically have a 30,00070,000-cfm (1433-m3/s) flow rate and
a 34-in Wg (7471245-Pa) rise for a 500650-MW unit.
An adjustable damper can be provided at the top of the chimney for use when
reduced-pressurization pressure and flow rate are acceptable. Excessive annulus
pressurization could lead to air leakage through cracks in the brick into the inside
of the liner and cause increased liquid condensation in the flue gas.
3.9.1.1.3 Target Walls and Linings
Some brick liners operating under WFGD applications have experienced
differential liner growth, resulting in a permanent deflection of the liner.
Although the cause of this permanent deflection has not been determined, wet
saturated conditions and nonuniform temperature appear to be contributing
factors. EPRI has previously funded laboratory and field studies directed at
preventing this problem. The results of these studies are presented in [21].
Precautions such as installing a target wall in the bottom of the liner should be
considered. Mixing of gases with a nonuniform temperature profile should also
take place prior to entering the stack.
A target wall or target lining provides protection to the brick liner from the wet
flue gas. Target walls or target linings can be constructed partially or fully around
the inside circumference of the brick liner, and they should extend from the
bottom of the liner to 1 or 2 liner diameters above the breeching entry.
A target wall is an acid-resistant brick wall, constructed independently of the
brick liner. An air space should be provided between the target wall and the brick
liner.
A target lining is constructed directly against the inside surface of the brick liner.
The lining can be constructed using an elastomeric coating or a borosilicate-glass
block with an adhesive membrane, which provides a good moisture- and
chemical-resistant barrier.
Another method that has been considered to reduce the direct impingement of
wet gases exiting the breeching onto the interior surface of the liner is providing a
bottom entry into the brick liner. This can be accomplished by providing an alloy
breeching-thimble section that extends below the bottom of the brick liner.
Target walls may be needed for good liquid collection and reduced erosion of the
brick liners in the stack-entrance region.

3-24

3.9.1.1.4 Liner Bands


Circumferential liner bands should be provided around the liner to minimize
vertical cracking and to add stability to the liner if cracking occurs. The
minimum recommended band size is 3/8 x 3 in (10 x 75 mm) plate spaced at 5-ft
centers. Materials of construction will depend upon the operating conditions and
effectiveness of the annulus-pressurization system. For applications in which the
pressurization system is working effectively and the exterior of the liner remains
dry, bands constructed of carbon steel with a protective-coating system are
adequate. For applications in which liquid permeates the brick and is in contact
with the bands, a nickel-alloy material should be used for this corrosive
environment. Use of carbon-steel bands in a wet environment may require future
maintenance, and possibly future replacement. Several utilities replaced their
carbon-steel liner bands because of corrosion.
3.9.1.2 FRP/GRP
The most common material currently used for wet stack liners is fiberglassreinforced plastic (FRP). This material is also referred to as glass-reinforced
plastic (GRP). FRP is a composite material made of glass reinforcement and a
thermosetting resin. FRP is a lightweight laminated product that provides good
service in cool, wet, dilute acid environments. Because of size and shipping
restraints, these liners are usually filament-wound in sections on-site.
The FRP liners inner surface is a corrosion-resistant barrier that consists of a
glass mat and resin. Vinyl-ester resins are recommended instead of the polyester
resins that have been used in many past projects. Vinyl-ester resins provide
higher temperature and corrosion resistance and have a lower coefficient of
thermal expansion. The structural wall of the FRP liner is composed of
continuous glass strands that are helically wound around a mandrel.
FRP is suitable for the low temperatures that are characteristic of wet-scrubbed
conditions and is not recommended for bypass conditions or temperature
excursions. Exposure to bypass conditions will result in loss of strength,
shortening the life of an FRP liner. An inlet-water spray-quench system is
recommended to reduce the effects of bypass temperature excursions. FRP liners
downstream of FGD systems that do not have bypass capability do not need a
quench system. FRP liners should be designed, fabricated, and erected in
accordance with the recommendations of ASTM D5364, Guide for Design,
Fabrication, and Erection of Fiberglass Reinforced Plastic Chimney Liners With CoalFired Units [7].
3.9.1.2.1 FRP Liner Joints
FRP liners are typically fabricated on-site in 3040-ft (9.112.2-m) sections, or
cans, which are lifted into place within the windscreen and joined together.
The cans themselves are butted together with additional glass and resin, applied
to both the interior and exterior of the joint. Most of the joint material is
recommended to be placed on the exterior of the liner. The amount of joint
3-25

material on the inside of the liner should be minimized to reduce the potential
for creating horizontal discontinuities on the liner surface, from which droplets
may be re-entrained back into the gas flow. The surface of the joint material
must be tapered back to the base-liner material with a minimum 6:1 taper (10:1
preferred).
3.9.1.3 Alloy
Stainless steel covers a wide variety of corrosion-resistant materials. Operating
parameters and environmental factors greatly affect alloy performance, especially
pH, temperature, and chloride and oxygen levels. The corrosion resistance of
stainless steel is improved by increasing its molybdenum, nitrogen, chromium,
and nickel contents. Molybdenum and, to a lesser extent, nitrogen improve
resistance to pitting. Chromium aids in the development of protective, passive
surface films and increases corrosion resistance in oxidizing environments. Nickel
assists in the renewal of damaged passive films and stabilizes the austenitic
microstructure with its improved fabricability and weldability.
On average, chloride levels in the new FGD systems will likely be much higher
than earlier FGD systems because of the requirement for zero-discharge water
systems. Designs specifying chlorides above 15,000 ppm have been fairly
common, and some designs have anticipated sustained operation above 30,000
ppm. Closed-loop operation causes chloride levels to increase, thereby increasing
the possibility of pitting and/or crevice corrosion of stainless steel. Stainless steels
that have been used successfully in scrubbers may not perform well in FGD
outlet ducts or stack liners. These locations are wetted by essentially unbuffered
mist and/or condensate. Absorption of residual SO2 can cause the pH to drop to
below 2. Because of the potential for high chlorides and low pH in the absorber
outlet duct and stack, stainless steels are not recommended. For these areas, a
nickel-alloy material such as Alloy C276 or Alloy C22 is recommended.
High-nickel alloys are an effective corrosion-resistant material in wet FGD
applications. Nickel alloys containing high percentages of chromium and
molybdenum provide good resistance to pitting and crevice corrosion in a highchloride, low-pH environment.
Although several different nickel alloys may be adequate for FGD applications,
Alloy C276 (ASTM B575, UNS N10276) and Alloy C22 (ASTM B575, UNS
N06022) have been shown to provide the best overall corrosion-resistance
characteristics. Alloy C276 contains 16% molybdenum and 15.5% chromium.
Alloy C22 contains approximately 13% molybdenum and 22% chromium. These
alloys have performed well in numerous WFGD environments and offer
approximately the same level of corrosion resistance. The two products have
successfully withstood a wide range of acid concentrations and extremely high
chloride levels, in conditions ranging from 130F (54.4C) saturated flue gas to
full system bypass.

3-26

3.9.1.3.1 Solid Alloy


Alloy C276 and Alloy C22 is produced in bar stock, plate, and sheet. Liners
constructed using this material could be fabricated from solid-alloy plate, from
carbon-steel plate with an alloy-sheet lining, from roll-clad plate, or from
Resista-Clad plate. Although liners fabricated from all of these applications are
highly corrosion resistant, liners fabricated from solid Alloy C276 plate offer the
most corrosion resistance. Because welding between the alloy material and carbon
steel is a potential source of dilution and contamination, the use of solid-alloy
plate eliminates this concern. However, the use of solid-alloy plate also requires
the highest initial capital cost.
Stiffeners on the exterior of the liner can be fabricated from carbon steel because
this material is not in contact with the flue gas. Any internal stiffening, such as
bracing in the ductwork, should be fabricated from alloy material.
3.9.1.3.2 Wallpaper Lining
Alloy-sheet lining involves the welding of thin alloy sheets, usually 1/16-in (1.6mm) thick, to a carbon-steel liner. This approach has commonly been referred to
in the industry as a "wallpaper lining." The sheet lining is installed with lapped
joints. The joints exposed to flue gas must be fully seal-welded to ensure
protection of the carbon-steel liner surface. The sheets are attached to the liner
with plug welds and intermittent fillet welds. The plug welds should be covered
with cap plates to eliminate any concerns about dilution. From a design
standpoint, only the carbon-steel portion of the liner should be considered
structurally effective in supporting the liner.
This lining process can be installed on a new or existing carbon-steel liner. If the
lining is to be installed on an existing liner, the liner-support steel should be
evaluated to accommodate the additional load of the sheet lining. Alloy C276
sheet lining has been utilized in the stacks at a number of plants. It has been
reported that after several years of operation, the material is performing well.
However, buckles in the lining sheets have developed throughout the liner as a
result of thermal stress.
3.9.1.3.3 Roll-Clad Plate
Alloy-clad plate consists of alloy sheet, usually 1/16-in (1.6-mm) thick, that is
roll-bonded to a thicker carbon-steel backing plate. The two metals are millrolled under heat and pressure until they are integrally bonded over the entire
interface. Clad plates are readily joined by welding. However, on the clad side, it
is important to minimize dilution to maintain adequate corrosion resistance of
the weld metal.
3.9.1.3.4 Resista-Clad
Another type of lining system that utilizes highly corrosion-resistant alloy as a
lining material is a process called Resista-Clad. Resista-Clad plate is a metal 3-27

fabrication technique that uses a patented process in which a relatively expensive,


thin sheet of corrosion-resistant metal cladding is attached to an inexpensive
substrate. The Resista-Clad plate process is capable of cladding carbon steel to
both weld-compatible metals, such as nickel-based alloys and stainless steels, and
nonweld compatible metals, such as titanium. This technique allows the costeffective use of both relatively high-cost alloy metals, for corrosion resistance, and
relatively low-cost carbon steel, for structural support. In this technique, 1/16-in
(1.6-mm) alloy sheet is bonded to the carbon-steel backing by the use of
resistance welding. During liner fabrication, the carbon-steel plates are welded
together, and then a batten strip of alloy is laid over the joints and seal-welded to
the alloy lining.
3.9.1.3.5 Weld Seams
The horizontal weld beads on alloy liners can cause a localized liquid reentrainment in the stack above a velocity of approximately 55 fps (16.8 m/s). The
height of the weld bead should be the minimum possible and not exceed the
standard of 1/8 in (3 mm) for good weld quality Vertical weld seams in a wet
stack, including the choke, are also typically limited to 3/16 in (5 mm). Limiting
the size of the horizontal weld seams is more critical than that of the vertical weld
seams with regard to minimizing liquid re-entrainment. If the shape of the weld
bead is gradually rising and falling, liquid re-entrainment is reduced, and higher
liner-gas velocities can be used. To achieve liner maximum velocities for smooth
metal, horizontal weld beads must be ground flat and smooth.
3.9.1.4 Borosilicate Block
Borosilicate foamed-glass block is an inorganic manufactured product that is
chemically resistant to most acids, solvents, and weak bases. Borosilicate-glass
block is supplied by Hadek, Inc., under the trade name Pennguard and has
been successfully used in numerous wet stacks all over the world. Its closed-cell
composition is impervious to liquids and gases and provides a low thermal
conductivity. This light-weight product is manufactured in block form and is
installed directly against the surface that it is to protect. An adhesive membrane
is used behind and between the blocks and serves as a moisture- and chemicalresistant barrier. These characteristics make it ideal for installation on liner
materials that are not generally suitable for corrosive environments such as carbon
steel or concrete. Also, because the mastic substrate provides a moisture-resistant
barrier, its installation in a brick stack will eliminate liquid seepage and bleeding
through to the exterior of the liner and the need for annulus pressurization.
Borosilicate block can operate at temperatures up to 960F (516C); however, the
adhesive membrane is limited to 200F (93C). Because the glass block is an
excellent insulator, the ductwork and liner should not be insulated if glass block is
used. If insulated, the temperature of the mastic may exceed the maximum
recommended temperature of 200F (93C) if the duct/stack system is operated
dry.

3-28

Because of its high insulative properties and low thermal mass, borosilicate block
will minimize the quantity of thermal condensation within the liner during unit
startup. The block also provides some level of protection to the base-liner
material in the event of a stack fire. In such a situation, the lining system will
most likely need to be replaced, but the structural portion of the liner may be
undamaged.
Borosilicate-glass block is well suited for lining ductwork. However, because the
block can be easily damaged, areas that are susceptible to physical or mechanical
abuse should be protected with an abrasion-resistant coating. Internal bracing is
difficult to line with block. Membrane coatings or alloy materials are
recommended for internal braces.
3.9.1.4.1 Mastic Joints
Borosilicate block is attached to the substrate material using an
adhesive/membrane material. It is important that the liner substrate be fully
covered with the membrane material to ensure that a continuous chemical- and
moisture-resistant barrier is formed between the inside and outside of the flue. A
nominal adhesive/membrane thickness of 1/8 in (3.2 mm) is recommended
behind and between each block. A 1/16-in (1.6-mm) thick layer should be
toweled onto the liner, as well as to the sides and back of each brick, before
installation. This double-buttering technique ensures a full bond of each block to
the liner and to each other. Excess adhesive material squeezed out during
placement of the block should be struck clean. Every effort should be made to
minimize the quantity of adhesive material smeared on the exposed surface of the
block, because these locations can lead to liquid re-entrainment back into the gas
flow during wet operation.
3.9.1.4.2 Lining Inside of Concrete Shell
Borosilicate-glass block can be attached directly to the inside surface of a
concrete shell. This technique has been used to reduce the cost of stacks and also
as a method for reusing older stacks whose liners are not favorable for wet
operation. The most common application is in existing stacks with tapered brick
liners previously operating with flue gas reheat. After conversion to wet
operation, the upper portions of the liner would be operating at gas velocities in
excess of the recommended velocity for wet operation for brick-liner material. In
such a situation, the upper portion of the brick liner can be removed down to a
point at which the gas velocities are acceptable within the brick portion of the
liner, and the inside of the existing larger-diameter reinforced-concrete shell can
be used as the upper portion of the liner. In such a situation, the inside of the
upper shell would be lined with borosilicate block to protect the reinforced
concrete. This arrangement is presented in Figure 3-6, where it can be seen that
a ring collector will be required at the interface between the lower and upper liner
sections of the flue to collect condensate generated within the upper portion of
the liner. This arrangement allows the entire liner to operate at acceptable liner
velocities.
3-29

BOROSILICATE
BLOCK
ATTACHED TO
INSIDE OF
SHELL

REINFORCED
CONCRETE SHELL

LIQUID
COLLECTION
GUTTER BETWEEN
LINER SECTIONS

BRICK LINER
REMOVED TO
THIS
ELEVATION
LIQUID COLLECTION DRAIN
LINER BANDS (TYP.)

FLEXIBLE
CONNECTION
BREECHING DUCT

LIQUID COLLECTION RING

EXPANSION JOINT

LIQUID COLLECTION
GUTTERS
EXISTING
BRICK LINER
DRAIN
PIPES

CONCRETE
PEDESTAL

FOUNDATION

PRESSURIZATION
FAN

Figure 3-6
Borosilicate BlockLined Upper Wind Screen

There are a number of disadvantages to this approach, which must be addressed


when considering such a retrofit. First is the loss of the annular space between
the old liner and the stack shell. All utilities, electrical conduits, ladders, drains,
and test ports located in this area will need to be moved to the exterior of the
stack shell.
The second disadvantage is potential damage to the concrete column if leakage
through the block and membrane develops. The location of the leakage and the
extent of the damage to the concrete column would be difficult to detect and
3-30

evaluate in such a system. The site-specific situation and risks involved should be
evaluated when considering this system for a utility wet stack.
3.9.1.5 Liner-Protective Coatings
Many existing units utilizing reheat and bypass liners have been constructed
using bare carbon-steel plate. This liner material is not acceptable for wet
applications. Carbon steel will corrode rapidly under wet acidic conditions.
Steel liners with an existing gunite lining are also not considered acceptable for
wet operation. Gunite is porous and tends to develop cracks through which liquid
can penetrate and attack the substrate material. Conversion to wet operation
requires removal of the gunite and relining with a corrosion-resistant material.
If liquid collection can be effectively collected without modifying the existing
liner geometry, applying corrosion protection to the carbon-steel liner should be
considered. Because the available options involve extensive field construction, an
extended unit outage is required to perform this work. The expense of
performing these modifications and the lost revenue associated with the unit
outage needs to be evaluated against the construction costs for a new stack that
would require a minimal amount of unit outage.
The following modification options can provide corrosion protection to the
carbon-steel liner:

Alloy wallpaper lining

Borosilicate-glass block lining.

Protective coating

Surface preparation is required before applying a lining or coating. The liner's


interior surface must be blast-cleaned. The degree of cleanliness and the surface
profile for the blast-cleaning depends upon the lining or coating being installed.
A coating system on an existing liner typically requires weld preparation, surface
preparation, testing for the pH of the surface, and if necessary, neutralizing acids.
Coating also requires careful control of humidity and temperature, which limits
installation. If alloy wallpapering is used, field measurements should be taken to
verify existing dimensions before cutting and fitting the lining sheets. Existing
welds must be ground flush for alloy linings and smoothly transitioned for
coatings.
Existing carbon-steel liners can be lined with borosilicate-glass block to provide
corrosion protection. Installation of the block is similar to installing a new liner,
except that additional surface preparation may be required for cleaning the
interior surface of the liner.
A wide variety of coatings are available, several of which are applicable to
WFGD operations. One such coating is the flake-glass-reinforced, vinyl-ester
system. This system is a glass-filled, thermosetting, resin-based, corrosionresistant lining system. The vinyl-ester resins are more durable and corrosion 3-31

resistant than the polyester resins that have been used in older liner-coating
applications. This coating system should be applied in three layers. The first layer
is a primer coat, which is used to prevent abrasive blasted steel from developing
rust bloom. The lining is then trowel-applied in two 3040 mil layers. Toweling
and subsequent rolling allows the glass-flake filler in each layer to be properly
oriented to the substrate and to achieve maximum resistance to water-vapor
permeation. In order to ensure adequate coverage, the two layers of resin can be
applied using different color pigments. Historical experience indicates that this
coating system may develop permeability, and liquid may eventually seep through
and reach the carbon-steel substrate. Coating systems other than the flakeline
resin systems have been utilized, but with only a limited degree of success.
The performance of any coating is dependent upon surface preparation and
application. Strict quality-control measures are required. The coatings need to be
applied under good weather conditions or under a controlled environment. The
coating should not be applied at temperatures less than 50F (10C) or greater
than 120F (49C). In order to get optimum performance out of a coating,
application and curing conditions are critical. The steel-substrate surface requires
significant surface preparation prior to application of the coating. All sharp edges
and imperfections must be ground smooth. Degreasing the surface and then
sand-blasting to a white metal blast (SSPC SP-5) is required. Blasting should
take place only over an area that can be primed within the same day. Surface
preparation and installation should carefully follow the manufacturer's written
instructions.
Coatings have been used successfully in some wet stack applications and have
failed in others. There is some risk involved in using a coating system, and the
potential for failure should be realized. Maintenance to repair chips and blisters
should be anticipated. Total lining replacement is typically required after a life of
about 810 years. However, coatings can represent substantial savings, and they
have been included as an alternative here for cost-comparative purposes.
3.9.1.6 Liner Insulation
A wet stack with a flue gas temperature of 130F (54C) typically will not require
insulation on the exterior surface of an alloy or FRP liner, because the liner is
already operating in a wet condition. Insulation is installed on coated steel liners
to reduce the temperature gradient across the coating, which may help in
extending the life of the coating. Borosilicate-lined stacks are not insulated to
ensure that the temperature of the mastic/adhesive material between the block
and liner wall does not exceed the manufacturers recommendations.
Insulation should be considered if the condensation calculations predict a high
liquid-film thickness within the liner, which could lead to an increased potential
for SLD. A thick liquid film could occur because of a high condensation rate or
could be the result of a tall stack. Insulation should also be considered as a means
of minimizing the potential for SLD if its minimization is a high priority; for
example, such might be the case if the stack were located in a densely populated
area.
3-32

At 130F (54C), the addition of 2 in (50 mm) of mineral-wool insulation


reduces the rate of liner-wall condensation by a factor of approximately four. The
addition of more insulation will reduce the condensation rate further, but with a
diminishing rate of return.
Uninsulated liners should have a protective coating on any carbon-steel plate or
stiffeners on the exterior of the liner that are exposed to the ambient
environment.
3.9.2 Recommended Liner-Gas Velocities
The gas velocity within the liner is a significant factor in minimizing liquid reentrainment into the gas stream. Liquid droplets, either carried over from the
mist eliminator or condensed on cold walls, are deposited on the ductwork and
stack-liner surfaces. As the droplets accumulate, they are pulled downward by
gravity, whereas the gas drags the liquid in the same direction as the flow of the
gas. When the force from the gas reaches or exceeds the forces of gravity and
surface tension, the liquid is sheared from the ductwork or liner walls. Liquid
then re-enters or is re-entrained back into the gas stream and is carried out of the
stack. When this occurs, the gas velocity is referred to as the critical reentrainment velocity.
Wet stack designers should select the liner diameter based on liquid-collection
considerations. Selecting the liner diameter so that the gas velocity is less than
the critical re-entrainment velocity (with a desirable margin for fabrication
inconsistencies and potential future plant up-rates) increases the opportunity for
liquid to be collected within the stack rather than being emitted with the gas
stream as SLD. The critical re-entrainment velocity varies depending upon the
liner's material, surface roughness, and quality of construction. Surfaces with a
high level of discontinuities and roughness (such as those of a brick liner) will be
more likely to re-entrain liquid, in comparison to smoother surfaces (such as alloy
or FRP liners). Therefore, the liner diameter will depend upon the liner material
selected. For example, brick liners will require a larger diameter than an alloy or
FRP liner for the same gas-volume-flow rate.
The recommended liner-velocity range for sizing the liner diameter when using
various liner materials was previously presented in Table 2-1 and is represented
in Table 3-4 below.

3-33

Table 3-4
Recommended Stack Velocity Range for Stack-linerDiameter Sizing

LINER MATERIAL

Liner
Velocity

Liner
Velocity

45

13.7

55

16.8

55

16.8

60

18.3

55*

16.8*

(ft/s)

Acid Brick
Radial tolerance of construction is 1/8 (3mm)

Alloy
Weld bead height <= 1/8 (3mm)

Fiberglass Reinforced Plastic


Minimum 6:1 taper on interior joints

Borosilicate Block
Radial tolerance of construction is 1/8 (3mm)

Coatings

(m/s)

* Laboratory testing of coating material is recommended to finalize liner velocity for favorable wet
operation.

3.9.3 Stack-Entrance Geometry


As discussed previously, there are two major stack-entrance geometries in use
today: the side-entry breech and the bottom-entry elbow. Each approach has its
own unique advantages and disadvantages, and each has specific design
requirements for both structural application and favorable wet operation.
The stack entrance is the last good place between the mist eliminator and the
stack outlet to collect and drain any liquid that enters the stack. The primary
techniques are intended to promote droplet impingement onto the surfaces by 1)
directing droplet trajectory to the far wall or a center baffle and 2) creating
a single or double gas-swirl pattern that will help deposit droplets on the liner or
baffle wall.
Once the liquid is on a solid surface in the lower stack, it must be

Allowed to drain naturally where upward gas velocities are low

Collected and drained in a gutter or shielded area

Prevented from re-entering the breeching area by side and top gutters on
each entrance opening

Drained to a protected sump in the bottom of the stack liner and out of the
system

Although single entrances are easiest to work with to obtain good liquid
collection and drainage, a double-entrance stack with a special center baffle for
liquid collection and drainage can also be made to work satisfactorily.
3-34

Variations can be made in all of these parameters and still achieve a satisfactorily
low level of SLD. However, more extensive liquid-collector and drain systems
will be needed, and further steps may be needed to reduce condensation in the
stack if velocities are in excess of stack-liner design values.
3.9.3.1 Side-Entry Breech
Side-entry breeches can be used on liners of any material. Because the flow
gasses entering the liner must make a sharp 90 turn to vertical after passing
through the breech opening, strong secondary-flow vortices are generated, which
are required for effective liquid collection. These secondary flows are stronger in
a liner with a side-entry breech than in one with a bottom-entry elbow.
Additionally, the liquid-collection system in the lower liner of a unit with a sideentry breech is less complexand generally performs betterthan that of a unit
with a bottom-entry elbow. For this reason, from the perspective of effective
liquid collection, a side-entry breech is preferred to a bottom-entry elbow for wet
operation.
For tall breeching openings, the vertical edges of the opening should be designed
to possess adequate stability as a column element. Multiple and single openings
under axial compression should be reinforced or proportioned to fully restore the
structural capacity of the entire cross section of the liner. Brick liners utilize brick
pilasters, and alloy liners utilize vertical jamb stiffeners, to provide vertical-edge
stability.
For brick liners, the width of the breeching opening may dictate the diameter of
the liner. The ASTM C1298 brick-liner design guidelines recommend that the
opening width not exceed one- half the internal diameter of the liner at the
opening elevation.
3.9.3.2 Breech D/W, H/W Ratios
Side-entry breeching-duct dimensions are typically rectangular, with the long
dimension in the vertical direction to minimize the structural effects on the
column and liner. It is beneficial to increase the height-to-width ratio of the
breeching by transitioning the width narrower and the height taller as the
breeching approaches the liner and by keeping the floor level. These
precautionary steps minimize the amount of liner cross-sectional area that is
interrupted at the breeching openings. Tall, thin breech openings also result in
the enhanced generation of strong secondary-flow vortices in the lower liner.
These vortices are used to push the collected liquid films to their respective drain
points. Experience has shown that the breech height (H) to width (W) ratio
(H/W) of the breeching-duct opening should be in the 23.5 range for favorable
wet operation.
Another dimensionless ratio affecting a breechs ability to generate strong
secondary-flow patterns is the liner-diameter (D) to breech-width (W) ratio
(D/W). This value should be between 1.5 and 2.5 for favorable wet operation.
3-35

3.9.3.3 Liner-Floor Geometry


Stack-liner floors must be designed to promote the drainage of liquid that will
collect in the bottom of a wet stack. Steel and alloy liners can utilize conical
hoppers or floors sloped toward the breech side of the liner. FRP liners
traditionally utilize a 90 mitered elbow, and liquid is collected in a drain at the
base of the elbow. Brick liners use a built-up sloped brick floor with a drainage
system. Floors or conical hoppers should be located a sufficient distance below
the breeching sill to prevent the liquid that has been collected on the stack
bottom from being re-entrained. EPRI Report CS-2520, Entrainment in Wet
Stacks, recommends that the hopper for liners with side-entry breeches be located
a minimum of 1/2 liner diameter below the sill of the breechingboth to prevent
flue gas from sweeping liquid from the floor up the liner wall and to reduce the
potential for generation of flow vortices on the liner floor that could re-entrain
liquid back into the gas flow [1]. This distance can be reduced to about 4 ft for
cost reduction, but the draining liquid may have to be protected from the gas
flow by a floor grating or by covers over the drain locations. Flow modeling can
define whether protective grating or baffles are needed and what their geometry
should be.
For alloy or FRP liners, the conical hopper, sloped-floor, or elbow material
should be the same material as the liner. Sloped floors in brick liners are
supported from an elevated support structure or by the concrete foundation.
Brick-liner floors consist of multiple layers of materials that provide corrosion
and thermal protection to their support structures. These floors should be
constructed of the following layers of materials: lead pan over asphalticimpregnated felt, acid-resistant mortar fill sloped to the drains, and a minimum
of two layers of acid-resistant brick and mortar. The lead pan should be returned
up the side of the brick liner to provide protection to the wall.
3.9.3.4 Sloped Liner Floor
Most recent alloy and FRP liners with side-entry breeches have incorporated a
sloped liner floor. This floor can is commonly incorporated into the liner wall, or
it can be bottom-supported within the liner (Figure 3-7). Flow-model studies
performed on numerous installations have shown the advantage of a sloped liner
floor, both for a reduction in system-pressure losses as well and for the generation
of robust secondary-flow vortices in the lower liner for enhanced liquid
collection. With a sloped liner floor, the secondary flows were found to be
stronger and more robust over a wider range of plant operations than those
generated in liners without a sloped liner floor. The design of the sloped liner
floor also results in the formation of a small, quiet sump area below the breech
entry for the collection and drainage of liquid from the liner. The final design of
the sloped liner floor should be optimized as part of the liquid-collection
physical-flow-model study.

3-36

SLOPED LINER WALL


SUPPOERTED OFF OF
LINER FLOOR
STACK LINER

STACK LINER

6-8 GAP
TO WALL

DRAIN

BREECH

BREECH

~1/3 BREECH
HEIGHT

SLOPED LINER
FLOOR INTEGRATED
INTO LINER WALL
DRAIN

~1/4 LINER
DIA.

QUIET SUMP
AREA

Figure 3-7
Sloped LinerFloor Arrangements

A good starting point for material-estimating purposes is to assume that the top
point of the sloped floor impacts the rear wall of the liner at a point 2/3 of the
way up the height of the breech opposite the opening, and that its bottom point
terminates 1 ft (0.3 m) below the floor of the breech at a point approximately 1/4
of the way across the diameter of the liner. Allowing the bottom edge of the
sloped floor to extend below the floor of the breech opening ensures that the
bottom edge of the incoming gas stream will not be peeled off and directed into
the sump area, thus ensuring that the sump remains aerodynamically quiet for
effective drainage. The sloped liner floor can be fabricated as part of the liner
wall, or it can be fabricated separately and supported off of the flat liner floor. In
the latter case, it is recommended that a 68 in (150100 mm) gap be provided
between the sloped floor and the liner wall, both to allow for differential
expansion and to allow liquid flowing down the liner wall to pass into the quiet
area behind the sloped portion of the floor.
Incorporating the sloped liner floor into the liner wall will reduce the quantity of
liner material required. This consideration can be important if the liner is being
fabricated from an expensive material such as an alloy.
3.9.3.5 Bottom-Entry Elbow
Bottom-entry elbows are fabricated from cylindrical sections, typically of the
same material and diameter as the liner, which have been cut into mitered
sections and joined together to form a 90 elbow. The number of miter cuts
defines the elbow. For example, a three-miter-cut elbow has three miter cuts; a
four-miter-cut elbow has four miter cuts (Figure 3-8). Three-miter-cut elbows
are recommended for favorable wet operation.
3-37

3 MITER CUT ELBOW

4 MITER CUT ELBOW

Figure 3-8
Miter-Cut Bottom-Entry Elbow Arrangements

It is well understood that, for FRP liners, bottom-entry elbows are a less
expensive approach. Although side-entry breeches are preferred for wet
operation, bottom-entry elbows with a properly designed liquid-collection system
can be just as effective. Their main disadvantage is that, because of the curvature
in the elbow, the liquid-collection system will require more complex shapes and
the existence of strong gas-recirculation zones exactly in the place where a key
liquid-collection ring must be located. If the ring is not properly designed, these
recirculation zones can strip liquid out of the collector back into the gas flow.
Bottom-entry elbows are commonly used on FRP liners because the elbow
segments can be fabricated from additional liner cans fabricated on site. Care
must be taken when fabricating elbows to ensure that the sections remain round.
As with liner joints, every effort should be made to minimize the joint material
on the inside of the elbow and to taper this material back to the base-liner/elbow
material with a minimum 1:6 taper (10:1 preferred) . Discontinuities in the
diameters of adjacent sections can create sites for significant levels of liquid reentrainment back into the gas flow.
3.9.3.6 Liners with Multiple Entries
Stacks can be designed to enclose a single or multiple number of liners. Each
liner can also be designed to accommodate a single or multiple number of
operating units. Economics, plant layout, ductwork layout, and the need to
perform maintenance during unit outages should be considered when evaluating
these options. The capital cost for a stack with multiple liners is typically less
than the combined costs of several stacks with individual liners. However, this
savings is partially offset by the increased costs associated with a larger
foundation, longer ductwork, and (possibly) larger pressure-rise ID fans.
Building a single large-diameter liner to serve multiple units is not
recommended, because if maintenance needs to be done on one unit, all the units
3-38

feeding into the liner will need to be taken out of service. Multiple units
operating within a common liner will also have additional design considerations
that will need to be addressed, such as gas-temperature differentials, variableflow rates, chaotic flow patterns within the liner, and continuous emissionsmonitoring-system certification. In order to monitor readings from individual
units, the continuous-emissions-monitoring system for a liner serving multiple
units would need to locate its equipment in the upstream ductwork rather than in
the stack. Another concern would be that, with one unit off-line, low-load
operations would result in a low exit velocity, causing potential downwash
problems. For good liquid collection, a one-breeching-duct-per-liner design
configuration is recommended. However, if multiple entries into a single liner
exist (or are required) consideration should be given to 1) placing one breech
opening directly above the other or 2) using a division wall within the lower liner
as a means of separating the lower liner into two separate aerodynamic zones that
can be optimized for liquid collection.
3.9.3.7 Division Wall
A division wall is a wall within the lower liner designed to separate the lower
liner into two (or more) separate aerodynamic zones. The objective is to decouple
the operation of the individual units feeding into the liner from each other so
that the flow patterns and liquid collection in each zone can be optimized
independently from the others.
Division walls typically extend vertically to a point at least one liner diameter
above the roof of the highest breech opening. The sides of the wall are typically
attached to the liner wall, although some designs provide a gap between the
division wall and the liner wall to allow for pressure equalization between the two
zones. The division wall must be able to withstand the full force of the gas jet
entering the liner through the breech openings. As such, it needs to be designed
to withstand the full pressure of the jet, even if the opposite equalizing breech is
out of service.
A significant quantity of liquid will be deposited on the division wall opposite the
breech opening, and provisions must be made to collect the resulting liquid film
and direct it to the liner wall and floor for drainage from the system. Liquid
collectors will also be required along the top edge of the division wall to prevent
the re-entrainment of liquid back into the gas flow. The wall can be placed
perpendicular to the incoming gas flow, or at an angle to it, to promote the
motion of any collected liquid films from the division wall to the liner wall for
drainage.
The design, placement, and orientation of a division wall and its associated
liquid-collection system must be optimized in a physical-flow model of the
subject unit.

3-39

3.9.3.8 Stack-Entry Pressure Losses


The total and static pressure of flue gas inside the stack is a function of the
following:

Ambient temperature and barometric pressure

Total pressure losses in the stack entrance

Total pressure frictional losses through the liner

Total pressure losses across expansion joints, corbel joints, and other stackliner breeching openings, expansions, or contractions

Total pressure loss associated with the liner-inlet geometry

Pressure variation with height (stack effect)

Static-pressure changes across area expansions or conical contractions and


chokes

Total pressure loss from the stack breeching duct to the stack discharge plume is
the value needed for determining the ID-fan requirements for the stack. The
maximum differential static pressure between the flue gas and the stack annulus
at the same elevation is the pressure needed to design the fans for the annuluspressurization system with brick stack liners. For a wet stackconversion project,
flue gas pressure data should be collected and, after correction to saturated
nonreheat conditions, compared to model study and calculated values.
Details of the pressure-loss coefficients for various liner-entrance geometries can
be found in [4].
3.9.4 Liner-Diameter Changes
Some stacks incorporate changes in the diameter over the height of the liner.
These can be gradual, such as in the case of a tapered brick liner. Or they can be
abrupt, such as in the case of a brick liner with multiple corbels, or in a liner in
which the top sections have been removed so that the upper section of the shell
can be used as the flue. A number of brick stacks have recently been
commissioned in which the lower breech-entry sections were fabricated from
alloy to allow for effective liquid collection, which then discharged into a largerdiameter brick liner. The transition between these two sections incorporated a
ring collector to prevent any liquid flowing down the brick portion of the liner
from entering into the lower alloy portion of the liner.
Gradual changes in the liner diameter are usually characterized by a reducing
diameter in the direction of the gas flow. This characteristic results in an increase
in the liner-gas velocity with height. Attention needs to be given to ensure that
the local liner-gas velocity does not exceed the recommended velocity for the
liner material in use.
Abrupt changes in the liner diameters are typically associated with a sudden
increase in the diameter of the flue. The transition from one section to another is
3-40

an ideal location for liquid re-entrainment back into the gas flow. In these
situations, the transition should incorporate a liquid-collection gutter to drain
away the liquid from the upper section before it flows down into the highervelocity lower section. One notable exception to this rule of thumb is the stackoutlet choke, which incorporates a rapid decrease in the liner diameter in the
direction of the gas flow. Stack chokes are unique in purpose and are discussed in
detail in Section 3.9.6.2.
3.9.5 Internal Structures
Ideally, stacks should be smooth and clear over their entire height. However,
reaching this ideal is not possible, because of the need to incorporate the many
structures necessary for the proper operation of both the liner and the plant.
Access into the liner in the form of ports for such purposes as testing, continuous
emissions-monitoring, and expansion joints must be incorporated into the liner.
Every effort must be made to minimize these structures, and their placement
should take into account their potential impact on the fluid-dynamic and liquidcollection performance of the liner.
3.9.5.1 Liner-Expansion Joints
Alloy and FRP liners sometimes incorporate expansion joints within the liner if
required by the liner-suspension method. These joints are often located at a point
less than one liner diameter above the roof of a side-entry breech or are placed at
the outlet of a bottom-entry elbow. These locations are not favorable for wet
operation because they lie in the region where most of the liner liquid-collection
occurs. Liner-expansion joints are typically longer than their ductwork
counterpartsranging from 25 ft (0.61.5 m) in length and 612 in (150300
mm) in radial depth.
The liquid condensing along the stack liner flows down on the liner surface when
gas velocities are below the liquid flow-reversal velocity. This liquid will flow
downward until it encounters a liner-expansion joint. If an expansion joint is
encountered, this surface discontinuity will cause all of the downward-flowing
liquid to re-entrain back into the gas flow by the gas recirculation in the
expansion-joint cavity. Most of these re-entrained droplets will be discharged
from the stack, and many will be large enough to reach the ground before
evaporating. For this reason, the liner should ideally be designed so that linerexpansion joints are not needed. If an expansion joint is required, it should be
located at least 1 to 2 liner diameters above the roof of the breech duct or the
inlet roof of the bottom-entry elbow. This arrangement will ensure that the
surface of the lower liner, where a significant amount of the liner liquiddeposition occurs, is free from discontinuities that would lead to liquid reentrainment. The expansion joint should be protected from the downwardflowing liquid through the installation of a liner-ring collector above the joint, or
the joint should be equipped with liquid collectors especially designed for linerexpansion joints. Laboratory flow modeling should be used to design and develop
liner expansion-joint collectors that are effective at all operating loads for a given
unit.
3-41

3.9.5.2 Observation Ports, Test Ports, Man-ways


Observation ports, test ports, and continuous-emissions-monitoring (CEM)
ports are required on all stacks. These ports are typically 610 in (150250 mm)
in diameter and are flush with the inside surface of the liner. Because of their
limited number, no special attention with respect to their impact on the liners
liquid-film flow is usually given. Some plants have incorporated inverted Vshaped gutters over the ports to direct the downward-flowing liquid film around
the openings on the liner wall.
A few plants have incorporated large 2 x 2-ft (0.6 x 0.6-m) access doors in the
liner, typically above the breech opening. The doors to these man-ways must be
flush with the inside of the liner wall. Both because of their size and of their
potential for disturbing a significant quantity of the liners downward liquid-film
flow, these man-ways should incorporate liquid-diversion gutters to direct the
downward-flowing liquid film around the opening on the liner wall.
3.9.5.3 Trusses, Stiffeners, Gusset Plates
To the extent possible, structures extending into the liner-gas flow should be
discouraged; liquid collected on these structures will be re-entrained back into the
gas flow, with little opportunity to be recollected before discharging from the
stack as SLD. The liners should be designed so that all structural supports are on
the outside of the liner.
Some side-entry breech openings incorporate internal horizontal bracing to
stabilize the tall side- walls of the opening. This bracing is common on FRP
liners with side-entry breeches, and it usually consists of three to four pipes, 10
12 in (250300 mm) in diameter, spaced evenly up the height of the breech
opening. Liquid will collect on these supports and will flow around to their
downwind side, from which droplets will be re-entrained back into the gas flow.
Liquid flowing on the breech side-walls can also be drawn into the low-pressure
region on the downwind side of these supports, further increasing the quantity of
liquid that will be re-entrained from this location. To prevent liquid from being
drawn into the region of flow separation on the downwind side of these
horizontal supports, it is recommended that liquid-flowprevention baffles, in
the form of 18 in (450 mm) in diameter disks, should be placed on the
supportslocated 18 in (450 mm) from, and parallel to, the breech side-walls.
Additional details of these disks are presented in Section 3.11.5.
Unless absolutely necessary, turning vanes, vortex breakers, and other flowcontrol devices in the entrance region of the liner should be avoided, because
these devices will act as locations for liquid collection and re-entrainment back
into the gas flow. Droplets generated from this location will have little
opportunity to be recollected before exiting the top of the stack as SLD.

3-42

3.9.6 Stack-Exit Design


For most modern installations, the stack exit is simply the end of the liner. If the
liner velocity is within the recommended range for the liner material used, no
special attention needs to be given to this area. If the liner velocity is near or
above the liquid flow-reversal gas velocity, the outlet of the liner should
incorporate a method for collecting the upward-moving liquid film. Details of
stack-outlet liquid collectors are presented in Section 3.11.10.
3.9.6.1 Liner Extensions
Stack-liner extensions are a common method of minimizing the potential for
plume downwash. Extension heights are determined as a result of a
computational flow-model study. Such a study should be performed to
investigate a stacks potential for plume downwash, using site-specific stack
design and weather conditions. Typical liner-extension heights range from 0.5 to
1 liner diameters above the top of the stack shell.
Liner extensions are typically not insulated so that they will remain above
freezing during the winter months, and ice will not form on them. This situation
will only apply if the liner is operating at or below the recommended velocity for
the liner material used because, during the winter, the condensation rate in the
uninsulated extension will increase significantly as it is exposed to the cold
environment. A liner operating at or below the recommended gas velocity should
be able to easily handle the increased condensate load on the liner wall.
If the liner is operating above the recommended gas velocity, the extension
should be insulated to minimize the increase in condensed liquid on, and
potential droplet re-entrainment from, the liner wall. Heat-tracing may be
required to control icing on the exterior of the extension and stack roof.
3.9.6.2 Chokes
Stacks that operate with a low discharge velocity sometimes have problems with
downwash or insufficient plume rise. To reduce these problems, a choke can be
incorporated at the top of the liner. Chokes are designed to increase the gas-exit
velocity by decreasing the discharge diameter. It must be noted that the use of a
choke, which constitutes a "dispersion technique" under EPA stack-height
regulations, cannot result in a relaxation of the emission limitation for the
facility. Chokes will increase draft losses, and the fan-pressure rise should be
selected accordingly. Chokes will also cause a positive pressure inside the liner
and increase the leaching problems for a brick liner. Chokes can be fabricated
from corrosion-resistant alloys and FRP. If a choke is installed on a wet stack
liner, a choke liquid collector will be needed at the top of the choke, because the
choke discharge velocity usually exceeds the liquid-reversal velocity for the choke
surface material. The choke surface material for recent choke designs is nickel
alloy or FRP, even for brick liners.

3-43

In stacks that have chokes, some of the fine droplets entrained in the gas flow
will be deposited on the choke surface, and the liquid collected on the choke will
lead to SLD if the local gas velocities in the choke are high enough to exceed the
flow-reversal velocity. The gas velocity on the choke increases from the liner
velocity, up to an 80120 fps (24.436.6 m/s) exit velocity, which is the usual
design range for chokes. The choke exit velocity normally exceeds the flowreversal velocity. Flow-reversal velocity is the flue gas velocity at which the flow
of the liquid on the stack walls is reversed from down- to up-flow.
The magnitude of the liquid discharge is a function of the choke geometry, the
gas-velocity variation through the choke, the droplet-size distribution, and the
spatial variation of the different-sized droplets. Flow-reversal velocity is in the
range of 7090 fps for common liner materials [1]. Therefore, liquid film will be
flowing upwards over a large percentage of the choke surface for most choke
geometries. The liquid must be collected at the top of the choke and drained out
to reduce the SLD.
3.9.6.3 Rain-Hood and Roof Design
The stack rain-hood shape is designed to cover the gap between the shell and the
single liner. A hood is typically installed on a stack in which the shell to linerdiameter ratio is small. Personnel access to this area will usually be limited. A
stack cap should be fabricated from a corrosion-resistant material such as FRP
and be angled downward away from the liner to promote drainage. Stack hoods
or caps are engulfed by wet stack gas all along their perimeter, and icing will
occur in cold climates. The amount of icing can be both minimal or large and
dangerous.
Hoods are typically left over from the dry stack design and are not favorable for
wet operation. A liner extension above a shell with a roof is favorable.
A stack roof is used on stacks in which the shell to liner-diameter ratio is large
and the liner extends above the roof. In such a case, the annulus is larger, and
personnel access is allowed to the roof. For this reason, a stack roof is structurally
stronger, and it incorporates a guard rail or parapet around its perimeter. Roofs
should be provided with a slope toward drains, leading the collected liquid to the
ground.
Typically, roofs and caps are not insulated, so that heat from the annulus can be
used to eliminate icing from this area.
3.9.7 Cyclonic Flow and Continuous-Emissions Monitors
EPA 40 CFR Part 60, Appendix A, Method 1 states that the flow condition is
unacceptable if the average yaw angle is greater than 20. If potential flowmeasurement problems are significant, it is recommended that a 3-D traverse be
performed to more accurately determine the flow characteristics of each unit.
These characteristics must be discussed with the regional EPA to determine
acceptability. If unacceptable, an evaluation of flow-straightening must be done.
3-44

Although typically undesirable from a wet-operation point of view, such an


evaluation could include the addition of straightening vanes near the monitor
location and/or a modification of the deflection plates at the base of the stack.
Any modifications would probably require a three-dimensional flow-model
study.
The CEM-system requirements for wet stack applications are generally the same
as for unscrubbed applications, with the exception of monitoring opacity. Opacity
monitors will not function properly in a wet stack because water droplets will
result in high opacity readings. The EPA has exempted wet stack units that are
otherwise affected by the acid-rain program from monitoring opacity. However,
no such exemption has been granted for new source performance standard units.
In addition, there may be local requirements for monitoring opacity.
EPA regulations specify that a quality-control program must include calibrations
and calibration checks for the CEM equipment installed. Provisions for field
testing need to be madenot only to meet regulatory requirements, but to
provide preventive maintenance and to ensure proper functioning of the
monitoring equipment. It may be required to cut holes in existing chimney liners
and concrete shells in order to get the probes through. Access into the annulus
space must also be considered. For units converting to wet stack operation, flow
monitors need to be recalibrated for the revised flow conditions.
CEM-system considerations include provisions for field testing, installation
and/or modification of platforms and access systems, and compliance to EPA
testing requirements.
According to EPA test methods for gas sampling (EPA 40 CFR Part 60
Appendix A), flow monitors should be located a minimum of eight stack or duct
diameters downstream, and two diameters upstream, from any flow disturbance.
If necessary, location of two stack or duct diameters downstream, and one-half
diameter upstream, from a flow disturbance may be employed. The purpose is to
make sure that the flue gassample data obtained are precise enough that the
total emissions can be accurately calculated.
Minimum testing requirements are satisfied if the EPA criteria for port locations
are adhered to and if no cyclonic or stratified-flow conditions exist.
Extensions to existing platforms may be required to provide access to new port
locations. Existing ports may be reused wherever possible. If ultrasonic flow
monitoring is used, an additional platform landing will be required, located
approximately one liner diameter above the main testing platform. Modification
to existing handrails may be required at certain locations to provide clearance for
the monitoring-equipment enclosure boxes.
To provide continuous access to the monitoring location and ensure personnel
safety, an elevator is advisable for accessing equipment on tall stacks. Platform
enclosures may be needed to provide personnel and equipment with weather
3-45

protection during the winter months. For personnel working in enclosed


conditions, adequate ventilation should be ensured.
3.10 Miscellaneous Materials of Construction
Because of the corrosive environment associated with FGD operations, materials
on the stack other than the liner also need to be evaluated.
3.10.1 Exterior of Shell
Gas-exit velocities on WFGD stacks should be designed for a lower velocity than
for reheat or bypass stacks. Consequently, plume rise on the gas exiting a wet
stack is not as high. With the gas velocity low and heavily saturated with liquid,
it is common for the flue gas to come in contact with the upper portion of the
stack. This condition is called downwash. As the gas comes in contact with the
concrete shell, liquid condenses on the concrete. Over time, this condensate can
cause the deterioration of unprotected concrete on the upper portion of the
stack.
To protect the upper portion of the concrete stack against downwash, a
protective-coating system should be applied to the top of the chimney. The
actual height that should be coated is typically determined as part of a
computational plume-downwash flow-model study. For estimating purposes, a
coated height of 50100 ft (1530 m) can be used. Applying two coats of an
acrylic-emulsion coating is an acceptable protection system for the concrete
surface. A high-build (46 mils dry-film thickness) epoxy coating has also been
used for this application. The painting system selected should be checked for
compatibility with the concrete curing compound and with any FAA painting
requirements.
3.10.2 Exterior of Adjacent Stacks
A stack that is located adjacent to and downwind of another stack may be subject
to the corrosive effects of flue gas impinging on its concrete surface. If the
upwind stack is significantly shorter than the downwind stack, the potential for
concrete corrosion is increased. This is a common occurrence in retrofit WFGD
installations, which incorporate new shorter stacks close to the old original stack.
The potential for plume impingement on a taller adjacent stack can easily be
estimated using a computational flow-model study and can usually be
incorporated into the plume-downwash study being performed for the new stack.
If the expected duration of plume impingement on the adjacent stack is high, it
can be protected by applying two coats of an acrylic-emulsion coating system.
3.10.3 Interior of Shell for Use with Brick Liners
For a stack with a brick liner, flue gas may enter the annular space of the stack if
the liner is leaking or if the annulus-pressurization system is not performing
adequately. This condition can be severe enough to result in corrosion of the
3-46

interior surface of the concrete shell. Additional protection should be provided by


applying two coats of an acrylic-emulsion coating system (or its equivalent) to the
interior concrete surfaces.
3.10.4 Liner Bands
Liner bands can be a potential maintenance concern if the band material or
coating system is not properly selected to resist the corrosive environment.
Uncoated carbon-steel bands have failed on several stacks. A coating system
using a first coat of high-build epoxy, with a second coat of polyurethane, can be
used for this application. However, in severe environments with excessive leakage
through the liner, this coating system may not be adequate, and alloy bands will
be required.
3.10.5 Rain Hoods and Roofs
The rain hood at the top of the stack is exposed to flue gas, so the rain-hood
material must be able to withstand this corrosive environment. Liquid reentrained in the flue gas is deposited on the rain-hood surface. In addition, the
continuous cycle of evaporation and deposition increases the acid content of
liquids deposited on the rain-hood surface.
Accordingly, the rain-hood material must resist higher acid concentrations than
the liner. FRP is an excellent choice, because this material is lightweight and
corrosion-resistant.
Stainless steels have been used on occasion for rain hoods, but most stainless
steels cannot resist the high acid concentrations that may occur at these locations.
Use of coated carbon-steel rain hoods is not recommended for rain-hood
applications because of maintenance concerns for the coating system.
Stacks with multiple liners use a roof instead of a rain-hood system at the top of
the stack. Roofs should be provided with a minimum slope of 1/4 in/ft to permit
adequate drainage. Roof materials consistent with those referenced for rain hoods
should be used. In addition, concrete roofs can be utilized. Type V sulfateresistant cement is recommended for the concrete. A coating system can be
applied to the top of the concrete to provide additional protection.
3.10.6 Lightning-Protection System
Because most of the items associated with the lightning-protection system are
located at the top of the stack, materials of construction for these items must
provide adequate corrosion resistance. The air terminals should be fabricated
from a solid rod made of Alloy C276. Conductors, attachments, and fasteners
located within the top 25 ft (8 m) of the stack should be lead-covered.

3-47

3.10.7 Platforms and Ladders


To provide adequate protection against downwash and liquid fallout, exterior
platforms and ladders located within the top 100 ft (30 m) of the stack should be
fabricated from a corrosion-resistant material. These include all platform steel,
grating, handrails, ladders, ladder fall-prevention devices, and attachments.
Typically, Type 317LMN stainless steel provides sufficient corrosion protection
at these locations. Items fabricated from Type 317LMN are not readily available,
and material substitutions may need to be evaluated.
3.10.8 Electrical Conduit
All rigid steel conduits and accessories within the top 100 ft (30 m) on the
exterior of the stack should be installed with a bonded PVC jacket. Conduit
located within the top 100 ft (30 m) on the exterior of the stack should be
supported with Type 316 stainless-steel unistruts, clamps, and fasteners. Because
of corrosion resistance and weather-tight requirements, all electrical boxes
located within this area should be NEMA 4X.
3.11 Basic Liquid-Collector-System Design
Saturated gas flow and condensation on the surfaces are normal operating
conditions for ducts and wet stacks. Some mist-eliminator liquid carryover, and
the liquid condensed from the large amount of water vapor present in the gas,
will also increase the liquid load in the ducting and in the stack liner. Liquidcollection devices and drains are necessary to collect and drain the liquid in the
absorber-outlet ductwork and in the stack before it exits the chimney. Liquid
collectors and drains are therefore an important design consideration in a
duct/stack system used for wet operation. The basic behavior of gas and liquid
flows are described in Section 2, "Wet Stack Design Fundamentals," whereas the
guidelines for selecting the duct and stack geometries suitable for wet operation
are discussed in Section 3.6, Ductwork Design, and Section 3.7, Stack-liner
Design, respectively.
The following sections of the Revised Wet Stack Design Guide provide an update
as well as additional detail for the basic liquid-collection designs previously
presented in [1] and [2].
The successful design and operation of a wet stack liquid-collection system starts
with the mist eliminators and includes all elements of the absorber duct,
breeching duct, stack entrance, stack liner, and stack outlet. Each element along
the gas-flow path is important.
The liquid collectors collect all the deposited and condensed liquid from the duct
and stack surfaces and direct it to drains for removal from the system. The type,
size, geometry, and location of liquid collectors are selected to utilize the systems
gas-shear and gravitational forces to move and collect the liquid on the duct and
liner surfaces and drain it out of the system before re-entrainment can take place.
3-48

The actual design and optimization of the liquid collectors and drains is sitespecific work, and it depends on the absorber type, the duct/stack geometry, the
planned operating-system operating conditions, and the gas-velocity levels.
Because the resulting three-dimensional gas- and liquid-flow patterns are unique
for each unit, the liquid-collection system must be optimized for each specific
unit. A flow-model study should be performed by a flow-modeling company
experienced in state-of-the-art wet stack design to determine the optimum
location and configuration of flow controls, liquid-collection devices, and drains.
Liquid collectors and drains will be required on any unit operating with wet ducts
and a wet stack. Because the collected liquid is acidic, the material selection for
liquid collectors and drains must be evaluated. The most common material for
liquid collectors are corrosion-resistant alloy materials and FRP.
For new installations, a general system arrangement and liquid-collection system
can be designed from the start for favorable wet operation. For units converting
from dry to wet operation, a number of unique considerations must be addressed.
The amount of vapor and liquid carryover from the mist eliminator is the same
for both new and retrofit applications. However, in many retrofit applications,
the gas velocities will be 510% lower than before conversion. Nevertheless, these
velocities are still usually higher than the gas velocities that would be selected for
the design of a new wet stack for the same gas-flow rate.
Because of the amount of liquid condensation and deposition on the duct and
stack surfaces and the relatively high gas velocities, design of liquid collectors and
drains is an important part of the retrofit process.
Mist-eliminator operation and performance need to be reviewed before a
conversion to wet operation. Mist-eliminator washing procedures will need to be
optimized to limit the amount of liquid-droplet carryover. In some instances, it
may be necessary to modify or replace the mist eliminators or mist-eliminator
wash system. Mist eliminators need to be monitored, inspected, and exchanged if
damaged or if solid scaling is excessive. The pH control needs to be good to
ensure that solid precipitation in the mist eliminators and the downstream duct
work is minimized.
Some units undergoing conversion to wet operation may have turning vanes or
other flow-control devices in the gas path between the absorber outlet and the
stack. These devices are common on units with multiple absorption modules and
interconnecting outlet ducting. If this is the case, liquid-collection devices will
probably need to be installed on them to minimize the potential for re-entraining
deposited liquid back into the gas flow.
The following sections discuss the basic liquid collectors typically found in most
effective wet stack installations. These discussions are designed to provide
engineers who are considering wet stack operation a basic understanding of the
design, location, and installation details of these collectors, as well as a basic
understanding of their purpose and operation. Although exhaustive, this list is by
3-49

no means complete, because many units often require custom-designed collectors


to address local gas- and/or liquid-film flow issues unique to the unit under
consideration. With the information provided in the following sections in hand,
engineers should be able to understand the operation of the overall liquidcollection system. They should also have enough information to estimate, with
some degree of accuracy, the quantity and types of materials required for a liquidcollection system, along with their fabrication and installation costs.
3.11.1 Side-Wall Gutters
Good liquid collection in the ducts means reduced liquid load in the stack and a
reduced potential for SLD. In horizontal rectangular-duct sections, simple
slanted side-wall collectors can be used to direct the liquid film flowing on the
duct wall to the floor. These collectors are usually L-shaped in cross section, and
their sizes range from 2 x 2 in to 6 x 6 in (0.05 x 0.05 m to 0.15 x 0.15 m),
depending on duct geometry, gas velocity, and expected level of liquid and solid
carryover from the absorber. The L-shape faces into the gas flow to form a
collection gutter. Side-wall gutters typically start at the roof of the duct and are
sloped 1015 in the direction of the gas flow (high side into the flow) to
promote drainage toward the floor. These gutters can also be placed vertically if
they are being used to keep the liquid film on the wall from encountering an area
where re-entrainment is likely to occursuch as immediately upstream of a ductexpansion joint or at the edges of the breech where they meet the stack liner. The
side-wall gutters typically terminate 68 in (15200 mm) above the floor of the
duct to promote drainage and prevent pluggage.
Depending on the level of mist-eliminator carryover, duct length, and duct
geometry, multiple side-wall gutters may be required down the length of the
duct. Also, short sections of wall gutters may be necessary at specific locations to
direct the liquid film on the duct wall away from an area where re-entrainment
might occur as a result of flow separation. These locations would be identified as
part of the recommended flow-model study.
Round ducts also need to be equipped with liquid collectors. Round ducts usually
require fewer collector elements, but they are more difficult to fabricate and
install because they need to follow the curved surface contour inside the duct.
These collectors are in the form of a ring that can go completely around the
perimeter of the duct with a drain at the base. Alternatively, a 1020 section at
the bottom of the ring can be removed to allow the collected liquid to flow along
the floor of the duct toward a drain point. Similar to the basic side-wall gutter, a
round-duct wall-ring collector will be sloped 1015 in the direction of the gas
flow (high side into the wind) to promote collection and drainage to the bottom
of the duct. Because the ring will be at an angle, and not perpendicular to the axis
of the circular duct, the resulting ring will be in the form of an ellipse. Circular
duct liquid-collection rings are usually L-shaped in cross section, and their sizes
range from 2 x 2 in to 6 x 6 in (50 x 50 mm to 150 x 150 mm), depending on
duct geometry, gas velocity, and expected level of carryover from the absorber.
The L-shape faces into the wind to form a collection gutter.
3-50

To be effective, the side-wall and ring gutters must be sealed to the duct wall
along their entire length. Details of basic side-wall and round-duct ring collectors
are presented in Figure 3-9.
LOCAL GAS FLOW
DIRECTION

SEALED TO
DUCT WALL
2-6 (50-150mm)
ANGLE
DUCT WALL

LIQUID FILM
FLOW
PATTERNS

SECTION AA

A
A

LOCAL GAS
FLOW
DIRECTION

10-20

6-8" (150-200mm)

GAP TO FLOOR

20

GAP OPENING

RECTANGULAR
DUCT

CIRCULAR
DUCT

Figure 3-9
Side-Wall and Round-Duct Liquid-Collection Gutters

3.11.2 Ceiling Collectors


Liquid will also condense and deposit on the roof of the duct. This liquid film
will flow in the direction of the local gas flow. The resulting liquid-film thickness
will increase as the film moves toward the liner until the droplets fall from the
roof. Many of these droplets may re-entrain back into the gas flow. To prevent
this from happening, gutters are attached to the roof of the duct to direct the
moving liquid film to the side-wall, where it will then flow onto the wall and
then down to the floor. Ceiling gutters are very similar to side-wall gutters in
size, shape, and orientation with respect to the local gas-flow direction. To be
effective, the gutter must be sealed to the duct roof along its entire length.
Depending on the local gas-flow patterns, a ceiling gutter either can be a straight
gutter running at a 1045 angle from one side-wall to the opposite wall, or it can
be configured in the shape of a V, with the point of the V upwind. This
arrangement directs the collected liquid film to both side-walls simultaneously
(Figure 3-10).
As detailed in Figure 3-10, the ends of the ceiling collectors are trimmed back at
a 30 angle where they attach to the duct side-walls. This cutback provides a
means for any gas collected in the gutter to leak out of the gutter in a controlled
3-51

manner as the liquid is directed onto the side-wall, minimizing the potential for
the scouring of the collected liquid back into the gas flow.
10-45
10-20

DUCT ROOF

DIRECTION OF
LOCAL LIQUID
FILM FLOW

DIRECTION OF
LOCAL LIQUID
FILM FLOW

SEALED TO
DUCT ROOF

SECTION AA
DIRECTION OF
LOCAL LIQUID
FILM FLOW

DIRECTION OF
LOCAL LIQUID
FILM FLOW

2-6 (50-150mm)
ANGLE

DETAIL 1

SECTION BB
DIRECTION OF
LOCAL LIQUID
FILM FLOW

30

DIRECTION OF
LOCAL LIQUID
FILM FLOW

DETAIL 1
Figure 3-10
Ceiling Liquid-Collection Gutters

In some instances, the local gas velocity may be opposite to the direction of the
bulk gas flow. This phenomenon is a common occurrence along the roof of
vertical transitions, particularly in the transition connecting the absorber-outlet
ducting to the stack breech. If the vertical expansion is abrupt, a region of flow
separation will be formed along the roof of the transition, and liquid collected
and/or condensed on the roof of the transition will flow downward back toward
the absorber. In such a situation, a ceiling V collector would need to be
installed, with the point of the V aiming up the slope of the roof. This V
collector will prevent the collected liquid film from reaching the bottom of the
transition roof, where it would be completely re-entrained back into the gas flow.
By orientating the V uphill, the downward-flowing liquid film will be directed
to the duct side-walls, where it will ultimately be directed to the duct floor for
collection
The size, shape, and location of ceiling collectors can only be accurately
determined through a flow-model study of the unit.
3-52

3.11.3 Floor Gutters


In most modern WFGD installations, the absorber is located close to the stack,
with a short duct connecting them. This ducting is typically sloped 11.5 either
toward the liner or back toward the absorber to enhance liquid drainagemainly
during unit shutdown. For these arrangements, floor gutters are typically not
required. For older units and/or units with long and/or complicated duct runs
between the absorber and the stack liner, floor gutters may be required to control
the flow of collected liquid along the duct floorusually toward a drain. A
trench or scupper installed across the duct in the floor with a drain is also
effective. Floor gutters are also used to prevent liquid from flowing into ducts
that are connected to the floor of the absorber-outlet duct, as is common for
units converting from reheat operation.
Floor gutters are typically angled to the gas flow to promote liquid flow to a
specific location such as a drain. The gutters are usually L-shaped in cross
section, and their sizes range from 2 x 2 in to 6 x 6 in (50 x 50 mm to 150 x 150
mm), depending upon duct geometry, gas velocity, and expected level of
carryover from the absorber. The L-shape faces into the wind to form a
collection gutter. To be effective, the gutter must be sealed to the duct floor
along its entire length.
3.11.4 Flow-Control Liquid Collectors
Many new and retrofit WFGD installations will have, or will require, turning
vanes located in the wet gas path between the mist eliminator and stack outlet.
Liquid condensed and deposited on these devices will be dragged to the
downwind or trailing edge of the device, at which point the collected liquid film
will detach, re-entraining liquid back into the gas flow.
In horizontal ducts, turning vanes and baffles are very effective for promoting
deposition of suspended droplets of all sizes and are frequently recommended if
additional droplet deposition is needed. In many instances, the additional cost of
these devices may be justified by the combined effect of reduced pressure loss and
the additional liquid collection generated. On the other hand, horizontal to
vertical turning vanes in the stack entrance should be avoided, because droplets
re-entrained from these devices will have little opportunity to be collected before
exiting from the top of the liner.
If turning vanes or other types of flow-control devices are present in the system
(or are required), liquid collectors will be necessary to minimize the potential for
liquid re-entrainment back into the gas flow. These collectors are typically
located on the trailing edge of the pressure side of the vanes. However,
depending upon the orientation of the vanes, collectors may be needed on the
back or low-pressure side of the vanes as well.
Like the side-wall gutters discussed previously, trailing-edge liquid collectors are
L shaped and range in size from 2 x 2 in to 3 x 3 in (50 x 50 mm to 75 x 75 mm),
depending on duct geometry, gas velocity, and expected level of carryover from
3-53

the absorber. The L-shape faces into the wind to form a collection gutter. The
gutters are typically placed on the vane or a vane extension in a slanted
orientation, angled 530 relative to the perpendicular of the gas flow. This
arrangement will direct the collected liquid film to the side-wall or the floor of
the duct, where it can drain onto the wall or floor and, ultimately, to the duct
floor. To be effective, the gutter must be sealed to the turning vane along its
entire length. Similar to the ceiling gutters, the ends of the flow-control
collectors are trimmed back at a 30 angle where they attach to the duct sidewalls. This orientation provides a means for any gas collected in the gutter to leak
out of the gutter in a controlled manner as the liquid is directed onto the sidewall. This arrangement minimizes the potential for the scouring of any collected
liquid back into the gas flow.
The gas- and liquid-film flow patterns in and around turning vanes can be very
complicated, and the development of effective liquid collectors is difficult. It is
highly recommended that a flow-model study be performed to develop and
optimize these collectors.
3.11.5 Internal Duct Supports and Expansion Joints
To the extent possible, all structural supports should be located on the exterior of
ducts used for wet operation. Internal bracing is often installed in rectangular
ducts on retrofit units that originally operated dry. Horizontal bracing is also
commonly found supporting the tall side-walls of breech openings on FRP liners
with side-entry breeches. These internal support structures are generally not
favorable for wet operation because they provide additional impingement surfaces
and re-entrainment sites for liquid droplets. Details of internal truss-support
designs for wet operation were previously discussed in Section 3.6.1.3.
Horizontal braces in side-entry breeches are a unique problem because of their
large diameter and because they are located very close to the stack liner. When
the liquid film flowing on the duct wall encounters a horizontal brace, a portion
of the film will flow onto the support and be pulled toward the center of the duct
in the recirculation zone on the back or downwind side of the truss. Liquid
pulled into this zone will flow along the back side of the brace, re-entraining back
into the gas flow along its entire length. Because of their close proximity to the
liner, many of these droplets will pass through without being recollected. To
prevent this from happening, a circular disk approximately 1012 in (250300
mm) in diameter can be added to the brace from 612 in (150300 mm) off of
the duct wall (Figure 3-11). These disks have been proven to be effective in
keeping the liquid film on the duct wall from spreading onto the support and
being pulled into the center of the duct, where it will be re-entrained back into
the gas flow.

3-54

DUCT WALL

10-12
(250-300mm)
DIA DISK

6-12
(150-300mm)
OFF WALL

GAS FLOW

Figure 3-11
Horizontal Brace with Liquid-Re-entrainmentPrevention Disk

Expansion joints (although required for duct design) are not effective for wet
operation, because they form discontinuities on the duct surfaces (wall, ceiling,
floor), from which the flowing liquid film can be re-entrained back into the gas
flow. The liquid films approaching an expansion joint should be collected and
directed to the duct floor by appropriately designed liquid-collection gutters
upstream of the joint. It is recommended that expansion joints in horizontal
ducting be drained. Liquid will naturally collect in the bottom cavity of these
joints, and they are ideal locations for eliminating collected liquid from the
system. Depending upon the design of the joint, the drain connection can be
made through the rubber boot or through a side structural member of the joint.
A 23-in (5075-mm) drain pipe should be sufficient. The drain piping will
need to pass through a seal pot or run directly back to the absorber reaction tank
to a point below the lowest liquid level to prevent gas from back-flowing up the
drain pipe, potentially resulting in liquid re-entrainment back into the gas flow.
If the expansion joint is not drained, it will fill with liquid. And any liquid
flowing along the duct floor will flow over the joint as it is pushed by the gas flow
toward the liner. Duct misalignments, or a rubber boot protruding into the gas
path, may be sufficient to generate regions of flow separation that could scour
liquid from an undrained joint back into the gas flow.
As described previously, in Section 3.6.1.4, units with internal duct supports will
typically have a brace located on either side of the expansion joint. If the truss
support attaches to the duct floor upstream of an expansion joint filled with
liquid, the recirculation zone formed on the downwind side of the attachment
gusset will scour the collected liquid out of the joint back into the gas flow. To
prevent this, a cover plate should be placed over the joint, extending
approximately 2 ft (0.6 m) on either side of the attachment gusset.

3-55

3.11.6 Stack-Entrance Collectors Side-Entry Breech


The liquid-film flow patterns in a wet stack can be defined by two basic
aerodynamic zones: the lower zone, typically extending from the liner floor to a
point approximately 1.5 to 2 liner diameters above the roof of the breech inlet,
and the upper zone above this point (Figure 3-12). The lower region of the liner
is defined by complex multidimensional gas- and liquid-film flow patterns,
whereas the upper zone is defined by generally well-behaved flow patterns. Most
liner-liquid collectors are located in the lower zone.

Arrows Show
Liquid Film Flow
Patterns
Upper Aerodynamic
Zone
Lower Aerodynamic
Zone

1-2 Liner
Diameters Above
Roof of Breech

Breech

Main Droplet
Impact Zone

Figure 3-12
Upper and Lower Stack-liner Aerodynamic Zones

The stack liner opposite the breech opening on a side-entry liner and along the
outer radius of stacks with a bottom-entry elbow generally receives more liquid
deposition than any other area downstream of the absorber. Effective liquid
collection in the stack-entrance zone is therefore critical for both inlet
arrangements. As described previously, when gas flows though a 90 turn, higher
gas velocities are experienced along the outer radius of the turn, and two counterrotating secondary vortices are generated at the outlet of the turn. Similar gasflow patterns are generated in the entrance region of liners as the gas flow turns
vertically up the stack. These gas-flow patterns both drag the deposited liquid
vertically up the liner and also push them circumferentially around the inside of
the liner back toward the breech opening.

3-56

The liquid-collection systems in the stack-entrance region of liners with a sideentry breech or a bottom-entry elbow are basically similar, but they have a
number of geometry-specific differences.
Figure 3-13 presents a generic liquid-collectionsystem arrangement for the
lower liner of a stack with a side-entry breech. This arrangement consists of a
T-shaped sloped-ring collector, typically 6 x 6 x 6 in (150 x 150 x 150 mm),
used to separate the upper and lower aerodynamic zones of the liner. The
upward-facing portion of this collector is used to collect liquid flowing downward
from the upper liner, whereas the downward-facing portion of the gutter is used
to help direct the vertically moving liquid film along the rear wall of the liner
circumferentially around toward the region of lower gas velocity located along the
front wall of the liner above the breech opening. The liquid film in this region
generally flows downward toward the top of the breech opening. To collect this
liquid, a gutter is located across, and extending slightly beyond, the top of the
breech opening. This gutter is typically deep, 1624 in (0.40.6 m), and
incorporates a number of strategically located vertical or angled tabs to prevent
scouring of liquid back into the gas flow. The floor of this gutter is sloped 12
from its center point to its outer ends to promote the movement of collected
liquid to drain pipes located at each end. These 68-in (150200-mm) drain
pipes run down either side of the breech opening, terminating 812 in (200300
mm) above the liner floor.
Liquid collected in the upward-facing portion of the liner-ring collector is
drained into the breech-top gutter through a drain box running down the front
wall of the liner. This drain box is sealed to the liner wall and incorporates wing
extensions to form two vertical gutters running along the sides of the box. These
gutters provide stability to the liquid film flowing circumferentially around
toward the front wall of the liner, directing it into the breech-top gutter.
To prevent the circumferentially moving liquid film from re-entraining back into
the gas flow along the sides of the breech opening, two large wing collectors are
located on either side of the breech, running from the breech-top gutter to the
bottom of the drain pipes (Figure 3-14). These wing collectors are generally L
shaped, but have sides of 8 in (200 mm) and 1824 in (450-600 mm). The
collectors are positioned such that the 8-in (200-mm) leg is attached and sealed
to the liner wall adjacent to the breech opening, and the 1824-in (450600mm) wing is roughly parallel to the liner wallangled slightly so that the
resulting opening or gap to the liner wall at the opposite edge of the wing is 46
in (100150 mm) wide. The gap to the liner wall is maintained by support rods
or straps located approximately every 4 ft (1.2 m) up the height of the gutter.
This gutter is designed to collect the liquid film flowing circumferentially around
the inside of the liner from the main droplet-impact zone on the rear wall of the
liner back around toward the breech opening. In many wet stack installations,
the breech-top gutter-drain pipes will be located inside the channel formed by
this gutter.

3-57

3.11.7 Stack-Entrance Collectors Bottom-Entry Elbow


In units incorporating a bottom-entry elbow, the lower-liner liquid-collection
system is basically similar to the side-entry breech arrangement. However,
accommodations are required to account for the curvature of the ducting in the
elbow (Figure 3-15). Similar to the side-entry breech arrangement, the gas
velocity opposite the stack inlet (in this case, along the outer radius of the elbow)
will be higher than average, and two counter-rotating secondary vortices will be
formed. Liquid deposited on the outer radius of the elbow will be pushed
vertically by the high-velocity gas flow and pulled circumferentially around
toward the front wall of the liner by these secondary-flow vortices.
A T-shaped ring collector, shown in Figure 3-15 as the lower ring collector, is
used to separate the upper and lower regions of the liner and to help direct the
circumferentially flowing liquid film in the lower section around to the front wall
of the liner. When a liner-expansion joint is used, a second vertically facing Lshaped gutter, referred to as the upper ring collector, will be necessary
immediately above the joint to prevent any down-flowing liquid from
encountering the expansion joint. When the liquid reaches the region of low gas
velocity above the elbow inlet, it will naturally flow downward because of the
reduced vertical gas velocity in this area. Because a breech-top gutter cannot be
used, a second (or lower) ring collector is located along the uppermost angled
miter cut of the elbow. This upward-facing lower ring, typically 6 x 6 in (150 x
150 mm), directs the collected liquid to a location on the outer radius of the
elbowwhere the vertical gas velocities are not high enough to push the liquid
vertically. Discharged at this point, the collected liquid will drain downward to
the floor of the inlet elbow. Similar to the side-entry breech arrangement, the
upper ring is drained into the lower ring through a drain box connecting the two
rings.
Because of strong gas vortices formed along the inner radius of the elbow in the
vicinity of the lower ring, numerous vertical tabs may be required in the lower
ring collector to prevent these vortices from scouring liquid out of the ring back
into the gas flow. If the vortices are strong enough, the ring may need to be
partially covered in these regions. The extent of scour protection required can
only be determined by a flow-model study of the unit.

3-58

UPPER RING
COLLECTOR
UPPER RING
COLLECTOR

EXPANSION
JOINT
LOWER RING
COLLECTOR

BREECH TOP
GUTTER

DRAIN BOX

SECTION AA
DRAIN BOX

DRAIN BOX
BREECH TOP
GUTTER

B
BREECH

SECTION BB
WING
COLLECTOR

WING
COLLECTOR

LINER SUMP

BREECH TOP
GUTTER DRAIN
PIPE

LINER DRAIN

SECTION CC

Figure 3-13
Typical Side-Entry Breech Liquid-Collection System

Breech Top
Gutter Drain
Pipe
Breech Top
Gutter

Breech Wall

4-6 (100-150mm)
Open Gap to Wall
Wing

Horizontal Section Through Wing Collector


and Breach Top Gutter

Figure 3-14
Side-Entry Breech Wing-Collector Design

3-59

Horizontal Section Through Wing Collector


Below Breach Top Gutter

UPPER RING COLLECTOR


(ONLY WITH EJ IF NEEDED)
DRAIN BOX
W/SLIDING
CONNECTION

EXPANSION
JOINT

A
ELBOW RING
COLLECTOR ANTISCOUR TABS

DRAIN BOX
LOWER RING
COLLECTOR
ELBOW RING
COLLECTOR

SECTION CC

LINER DRAIN

ELBOW RING
COLLECTOR ANTISCOUR TABS

BREECH TOP
GUTTER DRAIN
PIPE

SECTION AA

ELBOW RING
COLLECTOR

ELBOW RING
COLLECTOR
DRAIN BOX

SECTION DD

SECTION BB

Figure 3-15
Bottom-Entry Elbow Liquid-Collection System with Liner-Expansion Joint

3.11.8 Liner Collectors


The gas- and liquid-flow patterns in the stack-entrance region of a liner are
complex. In the upper portion of a liner, the gas-flow patterns are typically well
behaved, and liquid condensing within the stack liner in this zone will flow
downward if liner-gas velocities are below the liquid-reversal velocity. The lower
ring collector described in the previous sections is designed to separate these two
zones. The upper portion of the gutter collects the downward-flowing liquid
film, whereas the downward-facing portion of the gutter is used to direct the
moving liquid film in the lower liner to the quiet region above the liner-breech
opening. This ring collector is typically angled 510 from horizontal to promote
drainage, with the low point of the ring located on the front wall of the liner.
Liquid collected in this ring is directed to the breech-top gutter on units with
side-entry breeches, and to the lower ring collector on units with bottom-entry
elbows, through a drain box attached to the liner wall. The entrance to this drain
box should incorporate a debris screen to prevent pluggage.
Many stacks incorporate a liner-expansion joint. Liquid encountering this joint
will re-entrain back into the gas flow, a significant portion of which will be
discharged from the stack as SLD. Ideally, from a liquid-collection point of
view, expansion joints should not be used. If an expansion joint is necessary,
every effort should be made to place the joint at a location 1.5 to 2 liner
diameters above the roof of the breech or inlet duct. At this location, the joint
will be situated approximately at the division point between the upper and lower
aerodynamic zones of the liner, and it can be incorporated into the upper ring
collector (Figure 3-16). This arrangement will minimize the potential for SLD
caused by the expansion joint.

3-60

ELBOW RING
COLLECTOR
DRAIN BOX

Upward Facing
Ring Collector
10

Liner Expansion
Joint

1.5-2 Liner
Diameters Above
Roof of Breech

Breech

Downward Facing
Ring Collector

Figure 3-16
Liner-Expansion-Joint Placement and Incorporation Within Ring Collectors

A second alternative would be to incorporate liquid collectors into the expansion


joint itself (Figure 3-17). Such collector concepts have been used with some
success, but they require a special design and additional preventative maintenance
to ensure their proper long-term operation.
Downward
Liquid Film
Flow

Flue
Gas
Flow

Expansion
Joint Boot

Drain
Figure 3-17
Liner-Expansion-Joint Liquid-Collector Concept

3-61

Only laboratory flow-modeling can be used to design and develop linerexpansion-joint collectors that are effective at all operating loads of a given unit.
In some units, the secondary gas vortices in the lower liner may not be
sufficiently strong to push the liquid collected on the rear wall of the liner
circumferentially around to the front wall of the liner. This phenomenon can
occur in units with either side-entry breeches or bottom-entry elbows and is
typically associated with units in which the gas flow entering the liner has a
significant vertical-velocity component. (In other words, it is not horizontal when
entering the liner.) If this situation occurs, the deposited liquid film on the rear
wall of the liner will not drain properly. And as more liquid is deposited, the
resulting liquid-film thickness will increase until droplet re-entrainment occurs
off of the films surface. To prevent this, one or more large V collectors are
placed either in and/or slightly above the droplet-impact zone on the rear wall of
the liner (Figure 3-18). These collectors are made of L-shaped material and are
typically 6 x 6 in (150 x 150 mm), with the gutter opening facing downwards.
These V collectors (sometimes referred to as V diverters) use the vertical
motion of the liquid film to mechanically push it circumferentially away from the
rear wall of the liner to a point where the secondary gas flows are strong enough
to continue pushing the film around to the front wall of the liner. A gap opening
and cover plate are provided at the bottom of the V to prevent re-entrainment
of any liquid collected on the inside of the V back into the gas flow.
All liquid collectors in the stack liner will need to be sealed to the liner wall along
their full length to prevent leakage and possible re-entrainment of liquid back
into the gas flow.

V Diverter
Main
Droplet
Impact Zone
Liquid Film
Flow Pattern

Figure 3-18
Liner Rear-Wall V Diverter

3-62

Bottom
Opening
Cover Plate

3.11.9 Sloped Liner Floor


Stacks with side-entry breeches can accommodate a number of different linerfloor options, ranging from flat to conical to sloped. Of the many options
available, the sloped liner floor is most favorable for wet operation. Flat and
conical liner floors create a large open volume in the bottom of the liner, which
can lead to unstable and chaotic gas- and liquid-flow patterns as a function of
boiler load. These instabilities can result in the formation of vortices that are
strong enough to pull liquid off of the liner floor back into the gas flow, as well as
make drainage difficult because of liquid sloshing.
A sloped liner floor mitigates these issues, creating a small, quiet sump area for
effective drainage, eliminating the formation of floor vortices, and creating stable
well-defined lower liner gas- and liquid-film flow patterns across a wide range of
boiler loads. A sloped liner floor also reduces the stack-entrance pressure losses.
These benefits associated with the sloped liner floor justifies the additional cost
in most installations. A typical sloped liner floor was previously presented in
Figure 3-7. Although the optimal design is unit-specific, a good starting point
for estimating purposes would locate the high point of the floor on the back wall
of the liner at a point 2/3 of the way up the height of the breech opening. The
bottom point of the sloped floor would be located approximately 1/4 of the way
across the liner from the breech opening, with a sump depth of 610 ft (1.83
m). If the sump depth is 4 ft (1.2 m) or less, the top of the sump opening should
be covered with a grating to help prevent the flue gas from flowing into the sump
and potentially scouring collected liquid back into the gas flow.
The sloped floor either can be an internally supported structure or can be
incorporated directly into the liner wall. If the sloped floor is internally
supported, a 410-in (100250-mm) gap between the sloped floor and the liner
wall is recommended to allow liquid flowing down the wall to flow into the quiet
area behind the sloped floor.
3.11.10 Stack-Outlet Collectors
A well-designed stack operating below the liquid-film flow-reversal gas velocity
will have no need for stack-outlet liquid collectors. Units with high gas velocities
at the top of the liner or outlet chokes may require stack-outlet liquid collectors if
the discharge velocity of the choke exceeds the liquid-film flow-reversal velocity
for the choke material used. The recommended choke material for alloy and brick
stack liners is solid alloy. FRP choke material should be used with FRP liners.
Liquid reaching the top of the choke is discharged and falls to the ground within
the near vicinity of the stack.
Chokes for a wet stack need to be equipped with an effective liquid collector to
prevent this liquid discharge. Some current stack and choke-outlet liquid
collectors incorporate a downward-facing gutter at the liner outlet to collect the
upward-moving liquid film (Figure 3-19). This design is only marginally effective
because the pressure drop through the collection system will be higher than if it
was just passing around the collector. Also, the gas flow will always follow the
3-63

path of least resistance. Because of these factors, a large quantity of liquid will
collect at the point where the upward-flowing gas separates from the liner to go
around the gutter. The quantity of liquid in this area will increase until it starts
re-entraining back into the gas flow. Some gas and liquid will be collected in the
downward-facing gutter as a result of the momentum of the gas, but some
bypassing of the gutter and the resulting SLD should be expected.
Reduced Gas
Velocity Within
Collector
Liquid Film
Build-up

DETAIL 1
STACK LINER

GAS FLOW

DRAIN

Detail 1
Expected

DRAIN

Detail 1
Actual

Figure 3-19
Stack-Outlet Liquid Collector

An alternative design is presented in Figure 3-20, in which it can be seen that the
outer top edge of the liner has been sharpened to a knife edge. This arrangement
allows the liquid film to flip over the top edge of the liner outlet before it has a
chance to re-entrain back into the gas flow. Although effective, this collector
design is susceptible to fouling, and such a collector needs to be cleaned on a
periodic basis to ensure that deposits are not formed from which liquid could be
re-entrained.
A flow-model study of any proposed stack-outlet collector should be evaluated
and optimized in a laboratory flow model before installation at a plant.

3-64

COLLECTION
GUTTER

SHARPENED
DETAIL
2
EDGE

DETAIL 1

GAS FLOW

INSIDE LINER

STACK LINER
DRAIN

DETAIL 1
Figure 3-20
Alternate Liner-Outlet Liquid-Collector Detail

3.11.11 Drains
Liquid collected within a wet duct/stack must ultimately be drained from the
system. In general, drains can be installed in the horizontal absorber-outlet ducts,
duct-expansion joints, stack floor, liner-expansion-joint collectors, and choke
collectors. The preliminary wet stack design should plan an adequate drainage
system by considering the drainage needs discussed in this section. The final
number of drains needed and their optimum location are specified as one of the
results of the laboratory flow modeling.
The absorber-duct drains are most effective if a trench or scupper is installed
across the duct in the floor with a drain. The drains, if installed in ductexpansion joints, are selected on the basis of local conditions. Sloping the floor of
the absorber-outlet ducts either back toward the absorber or toward the liner
(preferred) eliminates water puddles after shutdown, when all the liquid from the
walls drains to the floor. The sloping floor does not alter the liquid flow on the
duct floor during operation because the gas shear dominates the liquid motion on
the floor.
Duct-floor drains may be required at locations where liquid pools can build up
and water re-entrainment from the pool may take place. The best locations of the
floor drains are specified by the wet stack flow model.
Drains located on a flat liner floor may need to have a solid cover positioned at
812 in (200300 mm) above the drain opening to provide a quiet area
underneath for effective liquid drainage. These covers are typically 3 x 3 ft (1 x
3-65

1m) and are supported over the drain opening by legs (not unlike a table). All
drainage points should incorporate a wire cage (or equivalent) over their inlet to
prevent plugging by flaking solid scale and other debris.
Drains connect points of different static-pressure levels inside the ducts and liner.
To prevent the possibility of gas back-flowing through the drain pipespossibly
resulting in droplet re-entrainment back into the gas flowthey must all be
individually pressure-balanced through loop seals or drain pots with good on-line
cleaning capability.
Because the liquid is acidic, the material selection for the drain line and the
means of final disposal must be evaluated. FRP pipe is commonly used for wet
FGD drain-line applications. FRP pipe provides good chemical resistance to
these acids, and is more economical and more readily available than stainless steel
or nickel-alloy pipes. Drain lines are sized to prevent pluggage and for ease of
cleanout. To accommodate the potential for solids plugging, drains lines should
be sized to handle a larger liquid-flow capacity than necessary. Main liner-floor
drains should be 1012 in (250300 mm) in diameter, because these are the final
drain points in the system. As such, they are points of possible single-point
failures. Seal pots or siphons should be provided as needed. These devices will
prevent air inflow under negative pressure, which would prohibit liquid discharge
through the drain line or cause droplet re-entrainment back into the gas flow.
Different drain lines can be connected only if the static pressure at the flue gas
end is the same; in other cases, pressure-loop seals or drain pots are required.
Chimney-drain lines should use pipe crosses, tees, or Ys with bolted blind
flanges for cleanout. To reduce pluggage, the drain lines should slope without
horizontal sections. Access to cleanout connections should be provided.
Outdoor FRP drain lines should have a steep-enough slope to prevent freezing.
For long, relatively flat horizontal runs of drain line, the pipe should be heattraced and insulated.
Stacks with a roof instead of a rain hood are relatively flat so as to serve as a
walking surface. Because of the tendency for liquid-fallout accumulation in this
area, discharged liquid and rainwater sometimes freeze on the roof during cold
weather. Ice formation at drain locations may prevent liquid from entering the
drains. In order to prevent this occurrence, roof slopes should be sufficient to
prevent liquid from pooling, and multiple drains are recommended. The roof
drain should be separate from the liner or choke liquid-collector drains because of
pressure differences and different disposal methods.
Liquid collected from within the ductwork or stack can be routed to the waste
sump or returned to the FGD-system process. Liquid routed to the waste sump
does not re-enter the process and is eliminated. Liquid returned to the process
can be routed to the absorber reaction tank. Liquid returned to the process may
be of sufficient volume and pH level that it could change the operational
characteristics of the FGD process. The volume and pH of the liquid returned to
the process should be considered to compensate for any operational changes that
may be needed.
3-66

A single stack liner that serves multiple FGD modules may produce large
volumes of liquid for disposal. Drainage-pipe diameters need to be sized
accordingly. If the liquid collected from multiple FGD modules is returned to
the FGD process, the volume of liquid may be too large to be returned to a single
FGD module without significant changes to its operational characteristics.
3.11.12 Post-Installation Inspections
An often-overlooked part of the liquid-collection-system installation process is
the post-installation inspection. It is important to ensure that the designed flow
controls, liquid collectors, and drains are installed properly. The objectives of this
inspection are to ensure that the liquid-collection system has been fabricated and
installed correctly; to ensure that it is free from leaks; and to identify any
deviations to the recommended design as a result of field modifications made to
ease installation. Inspection of the installation is usually scheduled for a day when
installation of the liquid-collection system is 8090% complete. Such timing
allows on-the-spot modifications to be defined, if necessary, and corrections to
be made while the construction/installation crew is still mobilized on-site.
Inspections of the liquid-collection system and drains should be performed
during the first scheduled or unscheduled plant shutdown. If any stack-emission
incidence occurs during normal operation, the need for inspection is obvious. If
SLD is not experienced, it is still important to inspect the liquid collectors to
ensure satisfactory long-term operation. Problems with solid depositions and
liquid drainage can be detected and corrected during a scheduled or unscheduled
outage. This inspection includes the stack-inlet duct, stack liner, and stack outlet,
if accessible.
3.12 Laboratory Flow-Modeling
An experimental gas-flow-model study is a valuable engineering tool used to
design and develop liquid-collection systems for wet operation.
The primary objectives of the wet stack flow model include
1. Minimizing the discharge of droplets from the top of the stack that are large
enough to reach the ground before evaporation (SLD)
2. Investigating the potential for plume downwash and for icing on the stack
liner or shell at low ambient temperatures
3. Providing velocity profile and swirl patterns at the CEM elevation that
satisfy EPA requirements for flow uniformity
Laboratory flow-modeling provides the following benefits for wet stack design:
1. Estimation of the total liquid load on the liner surface as a result of thermal
and adiabatic condensation. This estimate is based on the liner and shell
geometry, thickness, materials of construction and the flue gasflow rate and
properties over a range of expected plant-operating conditions and worst-case
ambient weather conditions.
3-67

2. Design and optimization of liquid collectors in the absorber-outlet hood,


absorber-outlet ducts, stack-liner entrance region, and if necessary, stackliner outlet to prevent SLD. The collectors collect liquid from duct and stack
surfaces, prevent re-entrainment, and guide the liquid to locations where it
can be drained out of the system. This development work is conducted on a
scale model of the wet duct and stack system, from the outlet of each
absorbers mist eliminator to approximately four stack-liner diameters above
the top of the breech or inlet duct. Model scales typically range between 1:8
and 1:16. Even though the absorber-outlet-duct region is not a part of the
chimney design, the liquid-collector designs should be developed here and
provided to the owner so that an integrated complete system of liquid
collectors can be developed for the unit. Typical wet stack physical-flow
models are shown in Figure 3-21.
3. Evaluation of the liner-material and liner-surface discontinuities that will be
produced by the construction technique. A laboratory test determines liquid
re-entrainment and drainage behavior over the expected range of unit
operating conditions and confirms or selects the stack-liner design velocity.
This step can be eliminated if sufficient reliable information already exists for
the selected liner material.
4. Evaluation of re-entrainment at liner-expansion joints in a laboratory test rig
and development of a unit-specific liquid-collector geometry to collect and
drain the liquid from the liner-expansion joints.
5. Development of an optimum choke geometry (if required), including liquid
collectors and drains to prevent re-entrainment and discharge of large
droplets from the top of the choke.
6. Performance of a computational 3-D fluid-dynamic-model study of the
upper third of the stack liner and shell to evaluate downwash of the wet
plume. Downwash could lead to stack-surface deterioration, unacceptable
ground-level concentrations of SO2, or icing problems at the top of the
stack. The final choices of the liner-extension height and the geometry of the
shape of the chimney cover will be specified as a result of this work.

3-68

Figure 3-21
Typical Wet Stack Physical-Flow Models

Laboratory flow-modeling has a number of limitations:


1. Physical cold-flow models cannot accurately simulate thermodynamic
processes such as thermal and adiabatic condensation. These must be
evaluated using mathematical or computer-based models.
2. The physics controlling the processes are different; droplet trajectories and
deposition patterns cannot be modeled simultaneously with the liquid-film
motion in a physical-flow model. Droplet trajectories are controlled by the
gravitational and centrifugal forces acting on the droplets as they pass
through the system, whereas the motion of the liquid film on the duct and
liner surfaces is controlled by gravitational and gas-shear forces. Because
these two phenomena cannot be modeled simultaneously, they must be
modeled separately and the results combined to get a full understanding of
droplet deposition and liquid-film flow processes within the wet duct/stack
system.
3. Because droplet trajectories and deposition patterns cannot be modeled
simultaneously with the liquid-film motion, physical-flow models cannot be
used to accurately determine the collection efficiency of the final
recommended liquid-collection system.
3.12.1 Computer Modeling.
A number of powerful computer codes are available for the evaluation of threedimensional fluid flow and thermodynamic processes. The applicable programs
can be separated into two categories: 3-D and 2-D computer programs. How and
where these computer codes can provide help and information for the wet stack
designers is described briefly in the following sections.

3-69

3.12.1.1 3-D Computer Programs


These are a number of commercially available programs designed for general
three-dimensional thermal and fluid-dynamic calculations. These codes can be
used to evaluate the overall gas-flow patterns and droplet trajectories, but they are
incapable of predicting liquid-film formation, liquid-film flow, and droplet reentrainment at a scale useful for the prediction of actual field performance. These
processes must still be evaluated in a physical-flow model of the subject unit. A
complete wet stack analysis with all the necessary detailsfrom the absorber
outlet to the top of the stackwill not be possible using 3-D computer code in
the near future.
The 3-D computer models are very useful for calculating the steady gas-flow
patterns in and around the ducts and stack. They are also very good at describing
the trajectories of liquid droplets of all sizes and at identifying their deposition
points in the wet duct/stack system when combined with the gas-flow
calculations and droplet-evaporation rates. Three-dimensional computer models
can also be used to predict condensation rates within the liner. However, 2-D
models can do this effectively without needing to model the entire geometry of
the wet stack system.
Three-dimensional computer models are often used to evaluate stack-plume
downwash. Details of this type of modeling are presented in Section 3.13.1.
The accuracy of a 3-D computer model is substantially a function of the data
used to set up the model; the quality of the computational grid used; and the
settings used within the code for the modeling of such factors as flue gas,
material properties, turbulence, and droplet-evaporation rates. Computer
modeling should only be performed by engineers experienced in the use of the
code and familiar with the modeling of the unique issues associated with the
modeling of utility power-generation systems.
3.12.1.2 2-D Computer Programs
Two-dimensional computer programs and spreadsheets are specifically written to
evaluate selected flow and thermodynamic processes. Therefore, their use is
important and the results are good, but they are generally limited to one detail of
the complete wet stack design process.
Special-purpose 2-D computer programs can be very useful for providing
quantitative results for cases in which experimental models are not applicable and
3-D codes are too time-consuming for the objectives. Their use often yields
satisfactory results for 2-D processes such as the prediction of thermal and
adiabatic condensation within a wet stack liner.

3-70

3.13 Plume Downwash and Icing


Designing liners for a low-velocity plume can decrease the amount of SLD.
However, having a low-velocity plume increases the potential for plume
downwash at elevated wind velocities. During downwash episodes, saturated flue
gas comes in contact with the liner extension, stack hood and stack shell. This
contact can lead to deterioration of the stack construction materials as a result of
exposure to acid in the flue gas. It also increases the potential for ice formation
on the top of the stack, which creates a potential danger to people and property
in the vicinity of the stack.
Details of the processes controlling the onset and extent of plume downwash
were previously presented in Section 2.9.1.
3.13.1 Downwash Modeling
The extent of downwash under various weather and plant-operating conditions is
typically evaluated using a computational fluid-dynamic (CFD) model of the
subject unit. A three-dimensional model of the upper 1/3 of the stack and its
surrounding environment is constructed for the study. The model will then be
used to study the flow patterns at the top of the stack and the likelihood of
downwash for different combinations of boiler-operating load and wind
conditions. Typical results from a downwash model of a single flue liner are
presented in Figure 3-22.

Prevailing Wind Direction

Liner Extension
Stack Shell

No Downwash onto Shell

Figure 3-22
Typical Plume-Downwash Study CFD Model Results

3-71

A typical plume-downwash study starts with an evaluation of the "as-designed"


stack (baseline geometry) at plant-operating conditions corresponding to full and
minimum load and at atmospheric conditions corresponding to the worst-case
velocities and prevailing wind directions expected during a typical full year of
operation. Simulations are then performed to determine whether, and to what
extent, the chosen ambient wind conditions create plume downwash. The results
of the simulations performed for ambient temperatures below freezing will be
used to determine the potential for ice formation at the top of the stack.
If the results of initial downwash simulation are unsatisfactory (for example,
downwash occurs at unacceptably low wind velocities or at a velocity for which
the cumulative number of hours at or above that velocity are too high), stack-top
modifications and/or liner extensions will be evaluated as a means of reducing
downwash and minimizing the likelihood of ice formation at the top of the stack.
If liner extensions prove to be ineffective, the addition of a choke or other stacktop modifications can be evaluated.
For plants with multiple liners within a single stack shell, additional simulations
will need to be performed as a function of the prevailing wind direction.
To perform a plume-downwash study, the following information will be
required:
1. Detailed drawings of the stack shell and liner.
2. Hourly wind-rose data (temperature, wind speed, and direction) from the
plant or nearest weather station for at least one full year (35 years preferred).
3. Stack base elevation.
4. Expected plant-operating conditions (gas temperature, flow, density) at
minimum and maximum load. For stacks with multiple flues, this
information will be required for each flue.
3.13.2 Windscreen Design
Because the potential for plume downwash is substantially a function of the
windscreen to liner-diameter ratio, the outer diameter of the stack windscreen
should be kept as small as possible at the top of the stack.
3.13.3 Single versus Multiple Liners
Plume-downwash potential for a traditional single-liner stack is independent of
the prevailing wind direction. The potential for downwash is only a function of
the stack geometry and the plume to wind-momentum ratio.
For units with multiple liners, the potential for and extent of plume downwash is
substantially a function of the prevailing wind direction and the number and
arrangement of the liners within the windscreen.

3-72

If the prevailing wind direction is in line with the axis of a stack with two liners,
the extent of downwash will be significantly less than if the wind direction is
perpendicular to the axis of the liners (Figure 3-23). In other words, when the
wind is aligned with the axis of the liners, it is more difficult to push the
combined plumes over than when the wind is perpendicular to the axis and the
wind is acting on both plumes simultaneously. Therefore , it is important to
know the prevailing wind direction when developing the initial system design
and setting the relative locations of the liners.

Prevailing Wind Direction


Stack Shell

Liners

Favorable
Liner Orientation

Unfavorable
Liner Orientation

Figure 3-23
Recommended Alignment of a Stack with Two Flues

Stacks with three or more liners will require multiple downwash calculations with
respect to the relative wind direction and individual liner-gas velocities to
establish the units downwash potential.
3.13.4 Methods of Downwash Minimization
If the actual or predicted cumulative duration and/or extent of downwash are not
acceptable, there are a limited number of methods available to reduce them:
3.13.4.1 Liner extensions
The most common method of minimizing the potential for plume downwash is
to increase the distance between the top of the windscreen and the top of the
liner in the form of liner extensions. These extensions minimize the potential for
the plume to be pulled into the low-pressure region formed on the downwind
side of the windscreen. Typical liner-extension heights range from 0.5 to 1.5
liner diameters above the top of the stack shell. Within these heights, the
extensions should be self-supporting.

3-73

3.13.4.2 Chokes
Because downwash is controlled by the plume to wind-momentum ratio,
increasing the liner-exhaust velocity will have an immediate impact on reducing
downwash. Although this goal can be accomplished by decreasing the liner
diameter, doing so would most probably result in the liner operating at a velocity
above the recommended value for wet operation, leading to a significant increase
in SLD.
Another method of increasing the liner-outlet gas velocity is the addition of a
liner-outlet choke. This option is typically considered as a last resort; it is an
expensive option with respect to both installation and operation because of the
increased fan pressure required to overcome the increased pressure drop for the
life of the plant.
3.13.4.3 Other Stack-Top Modifications
A limited number of nontraditional stack-top design modifications have been
successfully used to minimize the potential for plume downwash.
Chamfering the top of the windscreen (on 2 sides) can reduce both the onset and
extent of downwash, particularly if the chamfering is aligned perpendicular to the
worst-case prevailing wind direction.
For smaller-diameter stacks without a windscreen, a disk placed around the stack
near the outlet has been shown to reduce the onset of plume downwash. This
disk can either be solid or in the form of a walkway made of grating. The
diameter of the disk and its position relative to the stack outlet must be
optimized by means of a plume-downwash study.
Potential approaches are highly site-specific and they should be evaluated as part
of the plume-downwash CFD evaluation.
3.13.4.4 Protective Shell Coatings
If the expected cumulative duration of plume downwash events is acceptable, or
the extent of plume downwash cannot be reduced, one option is simply to live
with it and ensure that the top of the stack is adequately protected with an
effective shell-coating system.
3.13.5 Stack-Top Icing
Whether ice forms on the top of the stack depends on the temperature of the
surface, the temperature of the mixture of saturated flue gas and cold ambient
wind, and on whether water vapor will condense out of the mixture. Ice
formation is most likely at plants where below-freezing temperatures are
common and where they last for extended periods of time.

3-74

Icing usually does not cause serious ice buildups that can fall to the ground.
However, when the icing conditions are occurring, the platform near the top of
the stack, the railings, and possibly the roof, may be slippery.
The potential for icing can be reduced by employing the following steps:
1. Select a stack-liner discharge velocity that minimizes plume downwash over
the expected operating range of the unit at the existing local wind conditions
and that is consistent with other design objectives.
2. During cold ambient temperature conditions with high winds, run the unit at
near full load. Employing this operating procedure under such conditions
may be natural, because more power is consumed in below-freezing weather.
3. Use heated annulus air or electrical heating elements to heat the stack hood,
roof, or other areas where ice forms on the top of the stack.
4. Insulation is not required on sloped rain hoods constructed of a corrosionresistant material.

3-75

Section 4: Guide to Developing a


Specification for Wet Stacks
The designer of a wet stack system must follow a logical process to develop a bid
specification for a new or retrofit wet stack. This process involves evaluating
specific design issues that are important to wet stack design. The engineering
process that leads to a bid specification and the completion of a wet stack design
is discussed in this section.
4.1 Specification Overview
The design of a new or retrofit wet stack may be broken down into two distinct
phases of work:

Phase I Feasibility Study

Phase II Design Process

4.1.1 Phase I Feasibility Study


Essentially all modern wet flue gas desulfurization (WFGD) systems will operate
with a wet stack. Therefore, the two major questions to be answered during the
Phase 1 feasibility study are:
1) Can the existing stack be reused in a retrofit application?
2) If not, what will be the requirements for the new stack installation?
The Phase I feasibility study includes the following steps:

Address regulatory issues.

Define existing plant considerations.

Define operating parameters.

Determine if the existing stack can be converted to wet operation.

Determine the required stack height.

Determine liner alternatives.

Establish liner-material options.

Determine the impact of liner modifications on existing construction.


4-1

Establish liner-inlet options.

Determine liner-insulation requirements.

Perform operating conditions versus liner-design compatibility analysis.

Select design velocity limit to prevent liquid re-entrainment for each


candidate liner material.

Perform a preliminary economic analysis for liner material and inletgeometry options for wet stack operation.

Select liner material(s) for wet stack operation.

The feasibility study should address all considerations necessary to ensure that a
wet stack is allowable and that the proposed design will work. The study should
include permit considerations and plume-dispersion modeling as well as
preliminary design choices such as the absorber outlet-duct arrangement, choice
of inlet geometry, velocity limits for liner materials, and preliminary economics.
4.1.1.1 Identification of Issues (Phase I)
A number of important issues must be addressed during the Phase I feasibility
study for the design of a new or retrofit wet stack. These include the following:

Choice of liner materials

Gas velocity in the liner

Allowable gas velocity in the liner for favorable wet operation

Stack height

Foundation requirements as affected by liner choice

Absorber outlet-duct geometry

Liner-inlet geometry

Liquid-collection devices and drainage

Miscellaneous materials of construction

Operating conditions

Model-study testing to minimize system-pressure losses and optimize liquidcollection and drainage

Continuous-emissions-monitoring (CEM) system considerations

Outage time

Cost

A detailed discussion of each issue that is important for new or retrofit wet stacks
is included in this section of the design guide.

4-2

4.1.1.2 Evaluation of Issues (Phase I)


During the wet stack Phase I feasibility study, several critical issues must be
evaluated. For example, what material will be used for construction of the stack
liner? The liner diameter is established on the basis of the liner material. Each of
the specific design issues identified in this section needs to be evaluated from a
technical and an economic standpoint to enable the wet stack designer to decide
how to proceed.
4.1.1.3 Economics (Phase I)
Economic evaluation of a new or retrofit wet stack should include the following
factors:

Capital costs

Operating costs

Maintenance costs

Outage costs

Capital costs for a chimney include the cost of the concrete shell, liner,
foundation, and of miscellaneous items such as access platforms, doors, elevator,
rain hood, pressurization system, drain system, protective coatings, and electrical
system.
Chimney-maintenance costs will vary depending upon choice of liner material.
Coated carbon-steel liners need to be recoated periodically, whereas Alloy C276
clad liners and FRP liners require little or no maintenance. Maintenance on brick
liners can include repair of cracks, replacement of liner bands, and replacement of
expansion joints. All chimneys require periodic inspection.
Outage time does not apply to a new wet stack, because the unit has not yet been
placed into operation. For a retrofit wet stack, outage time depends upon how
long it will take to tie the new absorber-inlet duct to the existing ductwork and, if
applicable, reline or replace the liner within the existing stack shell. Outage costs
can be minimized by scheduling the ductwork tie-in to take place during a
normal maintenance outage. Outage costs per day will vary depending upon time
of year (energy demand), loss of revenue, and cost of replacement power.
The number of years considered in the economic analysis should encompass the
design life of the FGD system. Initial capital costs, periodic or annual
maintenance costs, and annual energy costs need to be converted to present value
to compare the options on an equal basis.
4.1.2 Phase II Design Process
Table 4-1 describes Phase II, which is the process for the design of a wet stack.
This table provides a brief description of each step of the design process, a
designation of each step of the design process, a designation of responsibility, and
4-3

a reference to the section of this guide that provides the relevant background
information.
The design-process phase essentially takes the feasibility study and turns it into a
bid specification. During this phase, it is necessary to establish the design criteria,
define the stack and inlet-duct geometry, perform the flow-model study to
determine liquid-collection devices and drains, and prepare the bid specifications
and drawings.
Table 4-1
Phase II Wet Stack Design Process
Step

Description

Responsibility

Reference
Section 2: "Specific
Design Issues

1. Preliminary
design

Perform component-bycomponent design.

2. Preliminary
design review

Adjust design for


Engineer and
Section 2: "Evaluation
favorable wet operation. modeling company for Wet Operation

Engineer

3. Fluid-dynamic Perform flow-model


design
study.

Modeling company

Section 2: "Evaluation
for Wet Operation

Prepare bid
4. Preparation of
specifications for
bid specification
chimney contract.

Engineer

Section 4: "Guide to
Develop Specifications
for Wet Stacks

5. Final design

Perform detailed design


of chimney for
construction.

Chimney contractor

----

Note: The utility must be an active participant in all major decisions in all steps of the design
process.

Several items considered during the feasibility-study phase should be revisited or


finalized during the design-process phase. These include such items as the
subsurface investigation and foundation requirements, stack and inlet-duct
geometry, and liquid-collection devices.
Detailed designs of the chimney foundation, concrete column, liner, and access
platforms are normally provided by the chimney contractor. However, the
foundation is often designed by the owners or general contractors engineer. The
electrical power-feed design is usually performed by the owners or general
contractors engineerbased on information from the chimney electrical
engineerbecause the electrical raceway system must be designed ahead of time
in order to be embedded in the chimney foundation. The chimney-related scope
of work is typically included in a separate contract. However, if schedule and site
constraints warrant it, the chimney may be included in the FGD system contract.
The preliminary duct and stack design is completed using the component-bycomponent guidance outlined in Section 3, "Specific Wet Stack Design Issues."
Next, the complete wet duct and stack geometry is evaluated for wet operation as
4-4

an integrated system, in which the individual components operate together as a


unit.
The first objective of this evaluation is to review the preliminary duct and stack
design as a complete system for wet operation. How well will all the flow
passages work together as a system, from the absorber outlet to the top of the
stack? How suitable are they to be outfitted with liquid collectors to achieve
minimum SLD? If improvements are possible by means of modifications to the
duct and stack geometry, are the gains in reduced SLD worth the cost of the
geometry changes?
The second objective is to design an effective liquid-collection system that is
optimized for the integrated geometry and for the specified range of operating
conditions and absorber combinations.
The combined wet duct and stack systems use a variety of design methods. The
best methods that have been developed are described in the following
subsections. The design methods and tools used to achieve these objectives are:

Preliminary design review based on experience

Laboratory flow-modeling 3-D

Computer flow-modeling 3-D

Computer program calculations 2-D

These design methods are summarized in the first four subsections below.
The subsection on "Comparison of Experimental and Computer Modeling" is an
overview of the basic gas/liquid-flow processes that need to be evaluated in a wet
system using the design tools, and it summarizes the usefulness of the
experimental and computer tools to carry out the needed evaluation.
The last two subsections are a discussion of the system tests and inspections
conducted after the unit is in operation to help identify how well the final
duct/stack geometry and the installed liquid-collection system are working. If
there are wet-operation problems, the tests and inspections can identify the
causes and practical solutions.
4.1.2.1 Preliminary Design Review Based on Experience
The first two steps in Phase II are to perform a component-by-component
design and then adjust the design for liquid-collection requirements. The results
of these first two steps should be evaluated by a preliminary design review.
The preliminary duct and stack-liner design geometry for new construction must
be reviewed to assess how well all individual components will work together as an
integrated system. The second goal of the design modifications implemented at
this phase of wet stack design is to make the integrated geometry more suitable
for liquid collection. This review should be based on actual field experience with
liquid collectors designed for different power plants. This modified wet duct and
4-5

stack design is the starting geometry for the design of the liquid-collection
devices. It also optimizes gas-flow passages by experimental flow-modeling. It is
important to get the engineer and modeling company together to review the
preliminary design early in the design process, because simple changes made at
this point in the process can have a big impact on the liquid-collection
performance of the entire system.
4.1.2.2 Condensation Calculations
A larger portion of the liquid flow on the liner surfaces is caused by condensation
than by deposition of the droplet carryover from the mist eliminators. The
condensation on the liner cannot be measured in a laboratory-scale model of the
stack. Therefore, analytical calculations are required to define the amount of
condensed-liquid flow rate in the stack.
The basic description of the adiabatic bulk condensation and the thermal-wall
condensation processes is given in Section 2.3, "Sources of Liquid in a Duct/Wet
Stack System." This section discusses the methods used to predict condensation
rates in the duct and the stack and also presents calculated condensation data for
typical sample stacks.
Liquid condensation occurs continuously on all wet duct and stack-liner surfaces
because the flow gas is saturated and the liner's inside surface temperatures are
lower than the gas dew-point temperature. But the rate of condensation per unit
surface area can vary significantly, depending on geometry, materials,
temperatures, and wind speed.
The duct- and stack-condensation rate cannot be measured in experimental flow
models. Three-dimensional computer modeling can be adapted for condensation
calculations; however, such models can be complex and are unnecessary for this
essentially two-dimensional process.
A 2-D condensation analysis is satisfactory for calculating the condensation rate
that accounts for stack geometry, mass flow, heat loss, gas-psychrometric
conditions, and the ambient atmospheric-pressure variation along the height of
the stack. Assuming axial symmetry for the stack, all the variables can be
described or calculated as a function of vertical height with computer programs
developed for this purpose as design tools.
Numerous studies performed using these programs have drawn the following
conclusions:

The bulk condensation in the liner is a function of stack height.

For brick liners, the pressurizing-air leakage through the liner increases the
bulk condensation by additional cooling.

The wall condensation is directly proportional to the temperature difference


between the gas and the ambient air.

4-6

The liner material does not significantly affect the wall-condensation rate
because the thermal resistance of the liner is a small fraction of the total
thermal resistance between the flue gas and the ambient air.

The wall condensation may be reduced by a factor of about 4 to 5 by


insulating the outside of the alloy or FRP liners with a 2-inch (0.05-m)
fiberglass layer. Increasing the insulation thickness from 2 in (0.05 m) to 4 in
(0.1 m) will result in a small reduction in condensation rate. Therefore, more
than 2 in (0.05 m) of insulation is seldom required. Insulation does not
eliminate the liner condensation but can reduce it significantly.

A wind-speed increase from 20 to 40 mph (9 to 18 m/s) increases the


condensation rate by only a few percentage points because the convectiveheat transfer levels off above 20 mph.

Having more than two liners in a shell increases the total wall-condensation
rate by about 5%. This increase is mainly the result of a higher condensation
rate on the liner extension above the shell.

The condensation in the ductwork primarily consists of thermal-wall


condensation; adiabatic condensation (due to pressure reduction with elevation
change) is negligible. Therefore, the rate of condensation here is not as important
as in the stack, because most of the condensed liquid and the deposited droplets
can be collected and drained out before they enter the stack. The ductcondensation rate is calculated either by a 2-D computer program or by handcalculation for the various ductwork surfaces.
The results of the condensation calculation are used to design the liquid
collectors and to plan for the liquid-disposal system.
The estimated ranges of the liquid-flow rates in the wet duct and stack of the
typical 550-MW unit are listed in Table 4-2. These numerical values help the
stack designer to estimate the amount of liquid flow to be controlled and
accounted for in a wet duct/stack system.

4-7

Table 4-2
Estimated Ranges of Flows in the Wet Duct/Stack of a Typical 550-MW Plant
Source of Liquid at the Stack

Estimated Flows

Mist-eliminator carryover
-

The vapor mass-flow rate in the


saturated gas stream is rather large.
The total vapor content at saturation
o
temperatures of 115130 F (46
o
54 C) is equivalent to a liquid-flow
rate of

11001800 gpm (250410 m /h)

The fine liquid mist of diameters <10


m with high-efficiency mist
3
eliminators is
0.31.5 gpm (0.10.3 m /h)
3

Duct-wall condensation

15 gpm (0.21.1 m /h)

Stack-wall condensation

16 gpm (0.21.4 m /h)

Bulk (adiabatic) condensation in the


stack

68.4 gpm (1.41.9 m /h)

Deposition of liquid droplets

112 gpm (0.22.7 m3/h)

Droplets re-entrained from the walls


(diameters 202000 m)

010 gpm (02.3 m /h)

Liquid in drains

212 gpm (0.42.7 m /h

SLD (a function of the duct/stack design


and the effectiveness of the liquid
collectors installed)

015 gpm (03.4 m /h)

4.1.2.3 Laboratory Flow-Modeling


An experimental gas-flow-model study is a valuable engineering tool used to
evaluate wet operation of the absorber-outlet duct and stack system from each
FGD module mist eliminator to the top of the stack.
The primary objectives of flow-modeling include:

Preventing the discharge from the top of the stack of droplets that are large
enough to reach the ground before evaporation

Investigating the potential for plume downwash and for icing on the stack
liner or shell at low ambient temperatures

Providing velocity-profile and swirl patterns at the CEM elevation that


satisfy EPA requirements

Laboratory flow-modeling provides the following benefits for wet stack


design:

Design and development of liquid collectors (experimentally) where needed


in the absorber-outlet ducts and stack-liner entrance region to prevent large
droplets from reaching the stack liner. The collectors collect liquid from duct
4-8

and stack surfaces, prevent re-entrainment, and guide the liquid to locations
where it can be drained out of the system. This development work is
conducted on a scale model of the wet duct and stack system, from the outlet
of each absorber to approximately two stack-liner diameters above the top of
the breeching duct. Even though the absorber-outlet-duct region is not a
part of the chimney design, the duct liquid-collector designs should be
developed here and provided to the owner so that an integrated complete
system of liquid collectors can be developed for the unit.

Evaluation of the liner-material and liner-surface discontinuities that will be


produced by the construction technique. A laboratory test determines liquid
re-entrainment and drainage behavior over the range of operating conditions
represented by the range of boiler operation and confirms or selects the
stack-liner design velocity. This step can be eliminated if sufficient reliable
information already exists for the selected liner material.

Evaluation of re-entrainment at liner-expansion joints in a laboratory test rig


and development of a liquid-collector design to collect and drain the liquid
from the liner-expansion joints.

Development of the optimum choke geometry, liquid collectors, and drains


to prevent re-entrainment and discharge of large droplets from the top of the
choke.

Wind-tunnel tests that evaluate downwash of the wet plume, which could
lead to stack surface deterioration, unacceptable ground-level concentrations
of SO2, or icing problems at the top of the stack. The final geometry of
chimney-cover shape and the extension of the liner will be specified as a
result of this work.

Modeling of the absorber-outlet ductwork, stack-liner breeching duct, and


the stack liner up to two liner diameters above the CEM elevation. This
modeling measures the gas-velocity profile and swirl angles for comparison to
EPA requirements at the CEM station.

4.1.2.3.1 Laboratory Flow-Modeling for Wet Stack Conversion


The geometry of the ductwork and the stack is inherited from the existing system
and cannot be altered in most conversion cases. For this reason, gas-flow passages
are less suitable for gas/liquid flow and for installation of liquid collectors than
optimized flow passages of new or retrofit units.
Experimental flow-modeling is the best tool to evaluate existing duct/stack
geometry and to design liquid-collection devices needed for a successful
conversion. Because of the geometric limitations, the modeling work is more
critical and more difficult. Laboratory flow-modeling provides the following
results, specifically pertaining to the conversion of an existing system to wet
operation:

Evaluation of existing duct and stack gas-flow passages for gas/liquid flow or
wet operation. Existing surface discontinuities, duct bends, and stack-liner
4-9

surface and geometry are evaluated, with the air and water flow simulating
the field conditions.

Design of duct and stack modifications to make the geometry more suitable
for wet operation. These modifications are limited to internal changes only.
Internal changes are inserts, target walls, turning vanes, and baffles that make
the flow more favorable for liquid collection and promote the deposition of
liquid droplets entrained in the gas flow.

Flow-modeling to decide if reheat-duct openings can stay open to the


isolation damper or whether they must be blanked off at the duct interfaces.

Evaluation of wet operation with existing internal trusses. Decide if they are
acceptable or if they need to be equipped with liquid-collection devices or
replaced with a more suitable truss design.

Design and development of liquid collectors. These collectors are


experimentally used where needed in the absorber-outlet ducts and stackliner entrance region to prevent large droplets from reaching the stack liner,
to collect liquid from duct and stack surfaces, to prevent re-entrainment, and
to guide liquid to locations where it can be drained out of the system.

Definition of the best location for drains, selection of possible drain locations
within the geometric limitations, and experimental evaluation of the final
drain system.

Evaluation of re-entrainment at liner-expansion joints in a laboratory test rig


and development of a liquid-collector design to collect and drain liquid from
liner-expansion joints.

Development of the optimum choke-liquid collectors and drains to prevent


re-entrainment and discharge of large droplets from the top of the choke (if
there is a choke in the existing liner).

Measurement of pressure losses in the model, scaling them to field-operating


conditions, and comparing them to the available pressure rise of the existing
ID fan. Reduction of pressure losses if needed.

Wind-tunnel tests to evaluate downwash of the wet plume that could lead to
stack-surface deterioration, unacceptable ground-level concentrations of
SO2, or icing problems at the top of the stack. The final geometry of the
chimney-cover shape and the extension of the liner will be specified as a
result of this work.

Use of a model of the absorber-outlet ductwork, stack-liner breeching duct,


and stack liner up to two liner diameters above the CEM elevation.
Measurement of the gas-velocity profile and swirl angles for comparison to
EPA requirements at the CEM station.

4.1.2.4 Computer Modeling


The use of computer programs for computational fluid mechanics has been
increasing during the last 20 years. How and when computer programs can
provide help and information for wet stack designers is briefly described here.
4-10

The applicable programs can be separated into two categories: 3-D and 2-D
computer programs.
In the following subsections, these categories of computer modeling are
compared to laboratory experimental flow-modeling from the wet stack design
point of view only.
4.1.2.4.1 3-D Computer Programs
These are commercially available programs designed for general threedimensional fluid-mechanics calculations. These codes have the potential to
handle all aspects of the gas/liquid-flow in wet stacks. But at the present stage of
development they cannot describe some of the most important wet stack flow
processes, such as re-entrainment and liquid flow on surfaces with gas flow.
Therefore, a complete wet stack analysis, with all the necessary details from the
absorber to the top of the stack, will not be possible by using 3-D computer code
in the near future.
The 3-D computer models are useful to calculate steady gas-flow patterns in the
duct and stack. They are good for major flow patterns, but they become complex
and slow-running programs when small details are included that require larger
computers (for example, the flow in the expansion-joint cavity, the flow around
trusses, and the flow near liquid collectors). The 3-D computer models are very
good at describing the trajectories of liquid droplets of all sizes and their impact
on deposition points in the duct system when combined with the gas-flow
calculations. These codes can be used to calculate deposition on the choke-cone
surface with assumed points of droplet origin. Some of the codes come with
trajectory subroutines that yield very detailed quantitative results.
The condensation on the liner of a wet stack can be described by a 3-D computer
code. But because the process is basically two-dimensional, 2-D analytical codes
are effective and are easier to use.
Several available 3-D computer codes are written for dispersion-modeling only.
The degree of complexity and accuracy of these codes varies a great deal as they
account for the terrain and atmospheric conditions over large distances from the
stack. Computerized modeling, verified by field-dispersion measurements, is the
best tool available at this time to predict the plume-dispersion process. Plume
downwash at the top of the stack could be predicted by an elaborate 3-D
computer model with a high degree of accuracy.
4.1.2.5 Full-Scale Field Tests and Inspections
After the wet stack is placed into operation, the effectiveness of mist eliminators
and downstream liquid-collection devices should be tested to verify that design
objectives have been met. EPRI Report CS-2520, Entrainment in Wet Stacks,
evaluates various methods for making carryover measurements and presents
recommendations for obtaining accurate results [1].
4-11

Effective operation of the liquid-collection devices and drainage systems should


also be verified during an off-line inspection of the liner system. Gas and liquid
flow on the liner and around the liquid collectors can be determined by observing
the pattern of residue on the internal surfaces. If heavy solids buildup or deposits
are found and are causing SLD, necessary improvements can be defined on the
basis of this inspection.
4.1.2.6 Flow Measurements at CEM Level
Flow measurements can be taken to confirm the operational characteristics of the
FGD system and the liquid-collection devices. Several of these measurements are
required by the EPA and are performed as part of the CEM and testing.
Information that can be provided from the CEM system measurements includes:

Stack gas-flow rates at different operating loads

Gas-velocity levels and yaw angles of the cyclonic flow

Additional in-place testing at the CEM level can measure the effectiveness of the
liquid-collection system. These tests can be performed as needed to establish the
liquid-flow in the stack. The following liquid-flow measurements are not
mandatory tests performed as part of the CEM system; however, they would be
beneficial in furnishing some or all of the following information:

Measurements of entrained droplet-size distribution using video droplet


analyzer (VDA) or similar methods

Qualitative amount of liquid near the wall by impingement methods

Liquid-flow direction on wall (up, down, or slanted) by observations at the


ports covered by a Plexiglas cover

Qualitative droplet sizes, using a suitable impingement probe exposed to the


flow for a few seconds

Wherever there is a large amount of entrained liquid, the droplets may interfere
with the CEM system velocity and angle measurements. This situation shows
that liquid collection and drainage must be improved, even though the SLD is
acceptable.
4.1.2.7 Outage Time
Outage time is the amount of time necessary for an existing unit to be out of
service because of repairs, maintenance, or modifications. The outage may be
caused by an emergency (unscheduled), or it may be routine (scheduled). Outage
time applies only to an existing, operating unit. Therefore, the designer does not
need to consider outage time for a wet stack on a new unit. For a retrofit FGD
project, outage time is the time it takes to tie the new FGD-system ductwork
into the existing ductwork.
Outage time for retrofit wet stacks should be minimized for economic reasons.
Depending on unit size, energy demand, loss of revenue, and cost of replacement
4-12

power, outage time can be very expensive. To minimize the outage time required
for a retrofit wet stack project, the ductwork connections or tie-ins should be
made during a scheduled maintenance outage.
4.2 Wet Stack Bid-Preparation Process
Once the decision has been made to utilize wet stack operation, it is necessary to
prepare a bid specification for the wet stack. For a new or retrofit wet stack, the
chimney can either be covered in a separate contract or be part of the FGDsystem contract. This section of the guide addresses key issues that should be
included in a wet stack bid specification. These guidelines should not be used
without application of good engineering judgment and consideration of site
specifics.
Preparing a bid specification for a wet stack is a multi-step process. The design
criteria for a wet stack are determined on the basis of plant considerations,
applicable codes, and system-performance requirements. General arrangement
drawings that define duct and chimney geometry should also be included. A
flow-model study should be performed in order to optimize the absorber outletduct arrangement and to design the necessary liquid-collection devices and
drains. At this point, the bid specifications can be prepared. Exhibit drawings
showing the wet stack arrangement and typical design details are included with
the specifications to define the scope of work and obtain competitive bids.
4.2.1 Establish Design Criteria
Minimum standards for design materials and construction are established in the
bid specifications. The chimney contractor is normally responsible for the design
of the wet stack to meet the specified operating conditions and any applicable
codes, such as the latest edition of the American Concrete Institute (ACI 307),
Code Requirements for Reinforced Concrete Chimneys, or ASTM D5364, Standard
Guide for Design, Fabrication, and Erection of Fiberglass Reinforced (FRP) Chimney
Liners with Coal Fired Units.
New stacks are typically designed in accordance with the latest wind and seismic
sections in the applicable steel-stack or concrete-chimney codes. Dynamic wind
loads should be evaluated in relationship to the chimney's critical wind velocity
and natural frequency. Wind effects from adjacent structures and nearby
chimneys should be considered. Closely spaced chimneys can cause an
amplification of the vortex-shedding wind loads for both the new and existing
chimneys. Spacing between adjacent chimneys can be increased in order to
decrease the amplification factor for the vortex-shedding wind loads.
In addition, a subsurface investigation is usually performed to determine the most
appropriate foundation type. For example, the subsurface investigation may
require that a deep, pile-supported foundation (versus a shallow, soil-supported
foundation) be implemented.

4-13

4.2.2 Define Absorber Outlet-Duct Geometry


General arrangement drawings that indicate chimney inlet-duct size, elevation,
and configuration are typically included with the bid specifications. The duct
arrangement is influenced by the number of absorbers, the size of absorber, and
the operating conditions of the unit.
4.2.3 Define Chimney/Liner Geometry
To ensure favorable wet operation, the chimney shell and liner-inlet geometry
and the liner diameter should be defined as early as possible in the initial study
phase of a wet stack project. Some flexibility is required to accommodate the
different scrubber arrangements and sizes available. Refer to Sections 2 and 3 of
this design guide for more information on chimney-geometry requirements.
Special consideration should be paid to choosing a liner diameter that results in
liner-gas velocities favorable for the wet operation of the liner material under
consideration.
4.2.4 Perform Model Study
Laboratory flow-modeling should be conducted before the final design phase to
design and optimize liquid-collection devices and drains for the absorber outlet
duct and stack. The objectives of a flow-model study are as follows:

Experimental evaluation of the absorber outlet duct and stack liner in order
to assure proper geometry and favorable wet operation

Design of flow controls to minimize system-pressure losses

Design of liquid-collection devices

Determination of drainage locations

Performance of a plume-downwash computational fluid dynamics (CFD)


study to determine the required liner-extension height above the top of the
stack shell and/or to determine how much of the top of the stack shell should
be painted with an acid-resistant coating

Laboratory flow-modeling is a valuable tool for the design of a successful wet


stack system and should be conducted by an experienced model-testing firm.
4.2.5 Determine Liquid-Collection Devices
Liquid-collection devices are critical to the successful operation of a wet stack
system. After the flow-model study is performed, liquid-collection devices can be
located and incorporated into the bid drawings. If the model study is to be
performed after the chimney contract is awarded, then bid quantities need to be
provided with the specifications to cover any potential liquid-collection devices.
Basic system designs from previous studies of similar units can be provided for
estimating purposes.

4-14

4.2.6 Prepare Bid Specifications


The bid specifications can be generated once the feasibility study, preliminary
layouts, and material selections have been completed. The bid specifications
should be as detailed and complete as possible to ensure competitive bids and a
successful wet stack system. Major items included in the bid specifications are
discussed below in the subsection, "What to Specify in a Bid Document."
4.2.7 Prepare Bid Drawings
Drawings are prepared and included with the bid specifications in order to clearly
define the scope of work. At a minimum, the bid documents include a site plan,
chimney general arrangement, typical details, and electrical drawings.
4.2.7.1 Site Plan
A site plan included with the bid specifications conveys the arrangement of the
power plant in relation to the area for the proposed construction. These drawings
should include lay-down areas for material storage and prefabrication/assembly
activities. These drawings will allow the contractor to plan his construction
operations. It also indicates restricted or congested areas on the plant site with
which the contractor will have to contend.
4.2.7.2 Chimney General Arrangement
A general arrangement of the chimney included in the bid documents gives the
bidders overall guidelines about the scope of work. Major features of the wet
stack are typically included on the chimney general-arrangement drawing. The
chimney general arrangement is only meant to serve as a guideline and is not
intended to provide final design information. The chimney inlet-duct elevation
may need to be adjusted, depending upon the absorber height.
4.2.7.3 Typical Details
Typical details that the owner or specifier wishes to incorporate into the final
design are included in the bid specifications. Typical details may include
platforms and handrails, rain hoods, expansion joints, duct supports, test ports
and probe-support jibs, insulation and lagging, lining systems, and liquidcollection devices such as gutters.
4.2.7.4 Electrical
Electrical drawings include items such as an electrical-grounding grid,
obstruction lighting, test-platform power and lighting, personnel-elevator power
wiring, conduit, annulus-pressurization wiring (if applicable), and electricalequipment layout.

4-15

4.3 What to Specify in a Bid Document


This section discusses items included in a bid specification for a wet stack.
4.3.1 General Requirements
In the general-requirements section of the specifications, the scope of work must
be clearly defined. Construction schedules, submittal schedules, temporary
construction power, and chimney design and construction requirements should
also be specified. The general requirements must be specific either to a new
chimney or to an existing chimney that is to be modified.
4.3.1.1 Scope of Work
The scope of work for the wet stack contract must be clearly defined in the bid
specifications. A complete description of the entire project may also be included
in the specifications. A new or retrofit wet stack, which involves constructing a
new stack, is usually designed, furnished, and constructed by a chimney
contractor who must comply with a performance specification that is usually
prepared by an architectural/engineering firm. A conversion to a wet stack, which
involves converting an existing stack to wet operation, is usually designed by an
architectural/engineering firm that produces a construction-contract
specification.
All work to be performed under a certain contract is clearly defined to avoid any
change orders or schedule delays later. The specification should also clearly state
that the contractor is responsible for furnishing and erecting all materials (if
applicable), and for providing the supervision, labor, tools, and transportation
required for construction of the work specified. Work by others is included in
multiple-contract projects, to assist the contractor with on-site coordination and
scheduling issues. A general work sequence may be helpful in avoiding
unnecessary construction delays. A list of contract drawings, reference drawings,
and reporting requirements is also helpful to the contractor performing the work.
4.3.1.2 Site Requirements
Consideration is given to site requirements such as construction parking, access
roads, lay-down areas, and any plant restrictions that might impact construction
efforts.
4.3.1.2.1 Construction Parking
Construction-parking areas may be identified to provide sufficient parking for all
construction personnel. Construction-parking areas may need to be maintained
throughout the construction process and restored to the original condition after
construction (if so desired).

4-16

4.3.1.2.2 Access Roads


The bid specifications may include a section to provide and maintain vehicular
access, both to the site and within the site during construction. New access roads
may need to be constructed to perform the work. Restrictions should be imposed
on existing access roads which cannot support construction activities.
4.3.1.2.3 Lay-Down Areas
Construction lay-down areas may need to be considered in the bid specifications.
The following items should be addressed:

Adequate area for construction-material storage

Provisions for substantial, weather-tight enclosures (if necessary for storage


of materials)

Control of temperature and humidity according to manufacturers'


recommendations

Platforms, blocks, or skids to protect materials from soiling or staining

Provisions for fabrication areas and/or enclosures if substantial on-site


prefabrication/assembly activities are planned, such as the on-site winding of
FRP liner cans

4.3.1.2.4 Safety Barrier


A provision addressing the need for a safety barrier around the base of the
chimney, which would protect personnel working below from falling objects, is
typically included with the specification. Chimney contractors usually require a
50-ft-wide safety zone around the base of the chimney during construction of the
concrete shell, exterior platforms, and ladders. A debris net on the outside of the
chimney shell is typically required to catch falling objects during construction.
4.3.1.2.5 Coinciding Construction Activities
This section alerts the contractor of any concurrent construction activities in the
chimney area. It allows the chimney contractor to make necessary provisions
regarding coordination of construction activities between different contractors
and/or the owner. For example, other construction activities within the safetybarrier zone cannot take place while the concrete chimney shell is being
constructed.
4.3.1.2.6 Plant Restrictions
Restrictions on construction imposed by the owner or by plant operations are
noted in this section. Restricted-access or lay-down areas may be addressed.
Plant-outage schedules may also be included in this section.

4-17

4.3.1.3 Project Meetings and Schedules


This section of the bid specifications includes requirements for the following:

Project meetings

Preconstruction conference

Finalizing schedules

Progress meetings and methods of submittal

Coordination conferences

Safety meetings

Schedules and reports

Initial coordination submittals

Work-progress schedule

Work-progress reports

4.3.1.4 Submittals
Submittal requirements specify what items must be reviewed by the engineer
before fabrication or construction. Submittals include both compliance submittals
and miscellaneous submittals. Compliance submittals include shop fabrication
drawings, product data, and samples that are submitted by the contractor,
subcontractor, manufacturer, or supplier. Miscellaneous submittals include
technical reports, administrative submittals, certificates, and guarantees.
4.3.1.5 Temporary Utilities and Facilities
Temporary utilities and facilities are specified to address water distribution,
drainage, dewatering equipment, enclosure of work, heat, ventilation, electricalpower distribution, lighting, hoisting facilities, stairs, ladders, and access roads.
In general, the specifications usually require the contractor to furnish, install, and
maintain temporary utilities required for construction, safety, and security.
4.3.1.6 Chimney Design and Construction Requirements
This section of the bid specifications includes minimum standards for design,
materials, and construction. The contractor is normally responsible for the
detailed design of the wet stack, whether it be new or a retrofit construction.
Specific design parameters for the piling, foundation, chimney shell, and liner
and/or lining material should be included in this section so the contractor can
perform the final design. The wet stack geometry can be included in this section
of the specifications unless indicated on the drawings. The operating conditions
for the unit, such as maximum gas-flow rate, maximum temperature, and gas exit
velocity at maximum gas flow, should be specified in this section. Specific site
conditions, such as the location of nearby structures and chimneys, are considered
for vortex-shedding wind loads on the new or existing chimneys.
4-18

4.3.2 Site Work


A bid specification for a new or retrofit wet stack that involves constructing a
new stack includes excavation, backfill, and foundation piling (if necessary).
Modification of an existing stack (if converted to wet operation) may require the
designer to analyze the existing stack and foundation, including piling, for new
wind and seismic loads (depending on applicable code requirements). The
existing stack and foundation also need to be analyzed for additional dead load, if
the existing liner is revised or replaced.
4.3.2.1 Excavation and Backfilling
A new stack will require excavation and backfill for the chimney foundation.
Dewatering requirements for the excavation during construction should be
addressed. Proper fill material and backfilling procedures are also typically
addressed.
4.3.2.2 Piling or Drilled Shafts
This section addresses furnishing all the labor, equipment, and materials
necessary to install foundation piling or drilled shafts for the chimney structure
(if applicable).
4.3.3 Concrete
The concrete section of the bid specifications for a new wet stack includes such
items as formwork, reinforcement, concrete materials, and quality control. The
contractor submits the concrete-mix design, which should preferably be based on
field experience. However, when sufficient or suitable strength-test data is not
available, concrete shall be proportioned on the basis of a laboratory-trial mix
design.
4.3.3.1 Materials
Materials of construction relating to the concrete work are usually specified.
Concrete materials typically comply with applicable building codes, ACI
standards, and ASTM requirements that are referenced in the concrete section of
the specifications. The following items should be specified:

Type of Portland cement (and fly ash if applicable)

Fine aggregate

Coarse aggregate

Mixing water

Admixtures (such as water-reducing type, air-entraining type, retarding type,


high-range water-reducing type)

Mix proportions (based on field experience or laboratory-trial batches)

Compressive strength
4-19

Slump

Air Content

4.3.3.2 Foundation
Foundation requirements for a new wet stack are included in the concrete section
of the bid specifications. In some cases involving a new stack, the chimney
contractor is responsible for designing the chimney foundation, including the
number and layout of piling (if required). However, the foundation is often
designed and supplied by the general contractor. Pile type and capacity is
generally determined by the architect/engineer.
4.3.3.3 Chimney Column, Construction Tolerances
The chimney column or concrete chimney shell may require special cement or
concrete additives in the mix design to facilitate rapid construction. Coatings are
usually used to withstand the potentially corrosive environment at the top of the
column. Construction tolerances for vertical alignment of center point, diameter,
and wall thickness for a new chimney column conform to ACI 307.
4.3.3.4 Brick-Liner Support Pedestal
The liner support pedestals usually are specified to meet the same concrete
requirements as the chimney column.
4.3.3.5 Roof Slab
Specifications for a concrete roof slab should consider either a corrosion-resistant
sealant, a coating applied on the concrete surface, or a sulfate-resistant cement
such as ASTM C150 Type V cement.
4.3.4 Liner
Liner materials of construction, strength requirements, construction
requirements, and ambient considerations are specified in this section.
4.3.4.1 Materials
Liner materials of construction for a new chimney liner are specified in the bid
documents. The material selected should conform to the respective ASTM
standards, unless specifically noted otherwise. Refer to Sections 2 and 3 of this
guide for liner-material discussions.
4.3.4.2 Strength Requirements
Strength requirements for various liner materials are dictated by the structural
design of the liner. For example, the yield strength of a steel or FRP liner and the
compressive strength of a brick liner should be specified.
4-20

4.3.4.3 Construction Requirements


Construction requirements for the selected liner materials are typically included
in the bid specifications. These requirements are primarily based on good
construction practice, recognized and accepted throughout the industry. Unique
construction requirements are also included in the bid specifications.
4.3.4.4 Construction Tolerances
Construction tolerances depend on the liner materials of construction selected.
Construction tolerances are either specified or referenced to a specific code or
standard.
4.3.4.5 Ambient Conditions
Ambient conditions may adversely affect the quality of liner construction,
depending on the liner material selected. Certain materials, such as coatings,
adhesive membranes, or mortars, may need to be stored and installed in a
controlled environment to provide satisfactory performance.
4.3.5 Ductwork and Expansion Joints
The chimney inlet-duct and expansion-joint section includes provisions for the
contractor to design, fabricate, furnish, deliver, and erect these items. The
ductwork and expansion-joint section specifies the materials of construction,
design criteria, and fabrication and erection requirements.
4.3.6 Miscellaneous Metals
Structural and miscellaneous metals are included in the metals section of the bid
specifications. Miscellaneous metals include such items as platforms and ladders,
liner or roof-support framing, test ports, and rain hood (if applicable).
4.3.7 Platforms and Ladders
The materials of construction for wet stack platforms and ladders are specified.
Carbon steel is typically specified for the entire ladder, except for the upper 50
100 ft, which should be fabricated from a more corrosion-resistant material. The
ladder should also include a Saf-T-Climb device for personnel safety. Wet stack
platforms may be fabricated from carbon steel, with the exception of the upper
portion of the stack, in which case the material should be similar to the upper
portion of the ladder. The extent and type of corrosion-resistant materials
specified will depend on the environment to which the materials will be exposed
and the extent of downwash anticipated.

4-21

4.3.7.1 Liner or Roof-Support Framing


Material selection should be considered when specifying the liner or roof-support
framing. The environment to which the materials will be exposed will dictate the
material selection.
4.3.7.2 Test Ports
Test-port materials are typically specified and conform to applicable ASTM
standards, depending on the materials selected. The materials selected will
depend upon such factors as the stack-liner material, their location on the stack,
corrosive environment, intended purpose of the test port, and material costs.
4.3.7.3 Insulation and Lagging
It is unnecessary to maintain the temperature of the saturated, absorber-outlet
flue gas above a certain minimum temperature. However, insulation may reduce
the amount of condensation in a metal or FRP liner. If insulation is required, the
insulation and lagging materials and design are specified in the bid documents.
4.3.8 Access Doors
Access doors are usually specified for the chimney liner at floor level or in the
chimney-inlet ductwork for cleanout and maintenance. Access doors may be
required in the chimney shell for annular platform and test-port access. Access
doors may also be specified for a brick-liner support pedestal (if applicable).
Additional liner-access doors can be provided adjacent to liquid collectors to
visually inspect the condition of the collectors and determine if the drain pipes
need to be cleaned. These access doors should be at the following locations:

Liner-entrance area above breeching duct

Near expansion-joint drains

At the choke

The size and location of the access doors may need to be evaluated and
determined by the flow-modeling company. Liquid collectors and/or liquid-film
diverters may be required around access doors to prevent droplet re-entrainment
from these areas.
4.3.9 Protective Coatings
Protective-coating systems for each particular component of the wet stack are
included in the bid specifications, if needed. Interior and exterior coating systems
may also be considered for the stack shell and liner. An exterior protective
coating may also be considered for the chimney inlet duct, if it is fabricated using
wallpapered construction or a clad-plate material.

4-22

4.3.9.1 Shell Exterior/Interior


Protective coatings may be considered for the upper portion of the concrete shell
of a wet stack to protect it from the corrosive effects of saturated flue gas during a
plume-downwash event. The extent of coverage can be determined from a
plume-downwash study. Additional consideration is given to coating the exterior
of adjacent existing stacks, which may be subject to the corrosive effects of
saturated flue gas from the wet stack.
4.3.9.2 Liner Exterior/Interior
If a liner is selected with a carbon-steel exterior such as an alloy-clad or
wallpapered liner, or one utilizing a borosilicate-glass or elastomeric interior
lining system, then an exterior coating system may be specified. If a coated
carbon-steel liner is to be used, then the interior coating must also be specified.
Refer to Section 2 of this guide for additional information on recommended
lining systems.
4.3.9.3 Structural and Miscellaneous Steel
The coating system selected for structural and miscellaneous steel is evaluated on
the basis of the environment to which the steel is exposed, access requirements
for maintenance, the cost of the coating system, and performance history.
4.3.10 Special Construction (Tuned Mass-Damping System)
Special construction may be necessary when converting an existing stack to wet
operation. For example, structural modifications may be necessary when
converting a steel stack to wet operation. If an existing liner is replaced with a
new liner, or an existing liner is relined with a glass-block or coating-lining
system, the change in the overall mass of the stack could affect the natural
frequency of the structure. A dynamic analysis is typically performed for vortexshedding wind loads. If resonance at critical wind speeds is a problem, then
structural modificationssuch as helical stakes, stiffeners, or the addition of a
tuned mass dampenermay be necessary to ensure the structural integrity of the
existing steel stack.
4.3.11 Personnel Elevator
The bid specifications for a new wet stack may require furnishing and installing
an electric rack-and-piniondrive personnel elevator. Design requirements, such
as conformance to applicable sections of ANSI, access to established platform
elevations, minimum lifting capacity, and minimum-design wind loads (for
exterior elevators) are typically included. If fabricated from carbon steel, the
exposed metalwork of the elevator and supporting structure should be galvanized
or protected with a coating system. For the upper 50100 ft of the chimney, the
mast and rack is usually specified to be Type 316 stainless-steel material. The
electrical power supply and wiring also needs be specified.
4-23

4.3.12 Mechanical
The following mechanical sections are typically included in the wet stack bid
specifications.
4.3.12.1 Drain Piping
The specifications provide for all pipes, fittings, hangers, supports, and
accessories required to complete the chimney corrosive drain-piping systems
associated with the liquid-collection system. Specific piping material should be
specified for each application. FRP or alloy piping is normally specified for drainpipe material. Minimum piping-installation standards are usually also addressed.
Stack-liner liquid-collector drains should incorporate some easily accessible
method for monitoring or observing the drainage flow. This capability will be
useful for determining the performance (or nonperformance) of the liquidcollection system.
4.3.12.2 Pressurization-System Fans/Dampers
This section of the bid specifications applies to a concrete chimney with an acidresistant brick liner. An annulus-pressurization system is required to keep flue
gases contained within the liner. The pressurization-system section should
include furnishing and installing the required number of fans, including drive
motors, controls, and accessories.
4.3.12.3 Louvers
The bid specifications include a section requiring the contractor to furnish and
install louvers to provide ventilation between the chimney liner and shell.
4.3.13 Electrical
The electrical section of the bid specifications includes provisions to furnish and
install all electric equipment, wiring (including plant-interface wiring),
grounding, and lighting necessary for a wet stack system.
4.3.13.1 Rounding
Grounding for all new electrical equipment is included in the bid specifications.
The grounding requirements for a new wet stack are similar to the requirements
for a dry stack. Any new electrical equipment used in a wet stackconversion
project needs to be properly grounded.
4.3.13.2 Obstruction Lighting
The bid specifications include furnishing and installing obstruction lighting,
including temporary obstruction lighting during construction. Obstruction
lighting must comply with the governing standards of the U.S. Department of
4-24

Transportation, Federal Aviation Administration (FAA) Advisory Circulars and


the National Electrical Manufacturers Association (NEMA).
4.3.13.3 Lightning-Protection System (Master Label)
The bid specifications include provisions to furnish and install a complete
chimney lightning-protection system, including a below-grade, chimney-ground
system. The Contractor may need to provide a Master Label C plate issued by
Underwriters' Laboratories, Inc., and attach the plate to the base of the chimney.
4.3.13.4 Plant Interface Wiring, Power Panel, and Disconnect Switch
The electrical section of the bid specifications address plant-interface wiring, the
power panel, and the disconnect/lockout switch. The contents of this section are
usually similar to standard power-plant chimney electrical requirements.
4.3.13.5 Pressurization-System Fan Wiring and Controls
The pressurization-system fan wiring and controls section of the specifications
usually follow the same general requirements as those for a dry chimney.
4.4 What to Ask For in a Bid Document
This section describes the Contractor's responsibilities for information that is
typically submitted, and inspection and testing that is usually performed, for the
design and construction of a wet stack or for wet stack modifications.
4.4.1 Warranties/Guarantees
The Contractor is typically required to furnish a performance and payment bond.
A performance bond is a guarantee that all work will be carried out in accordance
with the Contract Documents. A payment bond guarantees that the Contractor
will pay all lawful claims for payment to subcontractors and material suppliers
and for labor supplied in performing the work under the Contract.
Normally, a one-year warranty is specified for material and labor (one year from
initial operation). If an unproven technology is being proposed by the
Contractor, an extended warranty would be needed.
4.4.2 Design Calculations
The Contractor may be requested to submit calculations that confirm the
adequacy of the design. The calculations confirm that the design was performed
in accordance with the codes, standards, and specification requirements specified.
Design calculations are sealed by a professional engineer registered in the state
where the stack is being constructed.

4-25

4.4.2.1 Chimney Foundation


Design calculations for the chimney foundation may also be required for
submittal. The following is a list of items to be included with this submittal:

Foundation layout and size

Foundation loads

Soil-bearing pressure load or pile loading

Factor of safety against overturning

Radial and tangential design moments

Radial and tangential reinforcing-steel requirements

Shearing-stress calculations

4.4.2.2 Concrete Shell


The concrete-shell calculations specify all design loads and operating conditions
used. Calculations are provided for representative elevations throughout the
height of the chimney. The following information is typically provided with these
calculations:

Wind analysis and resulting loads

Dynamic seismic analysis and resulting loads

Column geometry, including diameters and wall thicknesses

Column deflections

Thermal stresses

Reinforcing-steel requirements

4.4.2.3 Liner
Calculations are also typically submitted for the liner design. The design loads,
operating conditions, and material properties are specified. Load combinations
applicable to the liner material are addressed per the governing code
requirements.
Information that is typically submitted with the liner design is as follows:

Loads and stresses for wind, seismic, thermal, and dead-load conditions

Liner geometry, including diameters and wall thicknesses

Liner deflections

Factor of safety against overturning (brick liners)

Factor of safety against buckling (steel or alloy liners)

Joint/weld specifications
4-26

4.4.2.4 Pressurization-Fan Size


The Contractor is usually required to size the pressurization fans if a brick liner is
to be constructed or converted to wet operation. The information that the
Contractor is required to furnish is as follows:

Fan manufacturer and model number

Fan static pressure, including accessory losses at the design point

Fan-volume flow rate at the design point

Fan-motor manufacturer and type

Fan-motor horsepower rating

Fan-performance curves, including capacity versus static pressure, efficiency,


and shaft horsepower

Design basis for determining fan size and leakage values

Drawings of the fan housing and base frame

4.4.3 Materials Testing


To confirm compliance with specification requirements, the Contractor is
required to submit materials-testing results and certifications for materials of
construction.
4.4.3.1 Acid-Resistant Brick
Certified laboratory-test reports may be submitted for each lot of brick
manufactured. Brick should not be shipped until it is confirmed that the certified
laboratory-test reports comply with ASTM C980 and the specified requirements.
Random-sample testing of the brick delivered to the job site may also be
performed in the field, as needed.
4.4.3.2 Brick-Mortar Certification
The brick-mortar manufacturer typically submits certification that the mortar
conforms to ASTM C466. Mortar-data submittals usually include the following
information:

Mix proportions and mix methods

Application temperature information

Curing and curing protection required

4.4.3.3 Fiberglass-Reinforced Plastic (FRP)


Material components of the FRP liner, such as resin and glass, are typically
certified for compliance with the appropriate ASTM standards and specification
requirements. All materials, prior to and during fabrication, as well as the
completed sections of the liner, may be subject to laboratory tests. Resin-batch
4-27

numbers and shelf life are recorded for future reference. For each batch of
material, resin gel-time tests are performed in accordance with ASTM D2471,
and resin-viscosity tests are performed in accordance with ASTM D2393.
4.4.3.4 Alloy-Material Certifications
Material-certification reports are usually furnished for all alloy materials and for
the alloy-welding materials. Certification reports should verify that the chemical
composition of the material is in conformance with the ASTM or specification
requirements.
4.4.3.5 Concrete-Material Certifications/Mix Design.
Mill certificates may be submitted for the cement being utilized to confirm its
conformance with ASTM C150 and to ensure that the correct type is being
furnished. Certifications are also required for the coarse and fine aggregate
(ASTM C33) and for any admixtures (ASTM C494) that will be utilized.
Concrete-mix design proportions and compressive-strength-test results
submittals are used to substantiate the proposed mix design.
4.4.4 Construction Procedures
The Contractor is responsible for his construction means, methods, techniques,
sequences, or procedures. The Engineer or Owner usually do not dictate and
approve the Contractor's construction practice. Submittal of construction
procedures are only required to confirm that the work can be performed within
schedule and meet the intent of the Contract Documents.
4.4.4.1 Foundation Construction
Chimney foundations are large-mass concrete placements that require substantial
planning and coordination. Before placement, the Contractor, Engineer, and
Owner review and coordinate the following issues:

Concrete-placement scheme (terracing or full lifts)

Method of concrete placement (pump trucks, buckets, or chutes)

Concrete-delivery schedule

On-site concrete-testing requirements

Method of concrete curing

Adequacy of formwork

Adequate personnel and equipment available

4.4.4.2 Hot- or Cold-Weather Concrete Procedures


Precautions need to be taken to ensure that hot- or cold-weather conditions do
not adversely affect the quality of the concrete. ACI 305R provides the guidelines
and recommendations for hot-weather concreting. Although several options are
4-28

available, the most commonly used hot-weather concreting construction


techniques are:

Using ice or cold water in the concrete mix

Applying cold-water spray to aggregates

Placement of concrete during the coolest time of the day or evening

ACI 306R provides the guidelines and recommendations for cold-weather


concreting. Because of the difficulty and expense of providing cold-weather
protection, construction of the chimney shell during the winter is not
recommended. In some instances, cold-weather situations cannot be avoided, and
precautions need to be taken to ensure that the concrete is not damaged.
Insulating blankets are commonly used to protect the concrete while it cures, and
they are often used during occasional cold-weather periods. An adequate supply
of blankets or other means of protection should be readily available on-site in
cold weather.
The Contractor may be requested to submit their hot- and cold-weather concrete
procedures for approval and be prepared to implement these procedures on short
notice, if needed.
4.4.4.3 Liner Construction
Liner-construction procedures are important for scheduling and coordination.
Liner materials of construction can take up a considerable amount of lay-down
and/or prefabrication area. Access to and installation of these materials need to
be coordinated with other construction activities in the area. Scheduling and laydown areas associated with liner materials may be submitted and incorporated
within the overall project schedule.
4.4.4.4 Welding Procedures
Submittal and implementation of welding procedures is essential in maintaining
quality control of welding for a wet stack application. Because the quality of the
weld affects the corrosion resistance of an alloy liner, a high level of quality
control needs to be enforced during welding. Coated carbon-steel liners also
require high-quality welds to ensure proper performance of the coating system.
For weld-fabrication requirements for lining applications, refer to ASTM
D4618, to NACE Standard RP0178-89, Standard Recommended Practice
Fabrication Details, Surface Finish Requirements, and Proper Design Considerations
for Tanks and Vessels to be Lined for Immersion Service, and to NACE Standard
RP0292-92, Standard Recommended Practice Installation of Thin Metallic
Wallpaper Lining in Air Pollution Control and Other Process Equipment.
Horizontal weld seams in a wet stack are typically limited to 1/8-in (0.003-m)
reinforcement or projection on the inside of the liner. Horizontal welds on the
inside of a choke are usually ground flush to prevent liquid re-entrainment.
Vertical weld-seam reinforcement in a wet stack, including the choke, is also
typically limited to 1/8 in (0.003 m). Limiting the size of the horizontal weld
4-29

seams is more critical than the vertical weld seams with regard to minimizing
liquid re-entrainment.
For all welding processes, the Contractor prepares and submits for approval the
welding-procedure specification (WPS) and the procedure-qualification test
results (PQR). For carbon-steel welding, the welding-procedure specification
includes the nonmandatory information included in Appendix E of AWS D1.1,
in addition to the mandatory information listed in Appendix IV, Table IV-1 of
AWS D1.1. For alloy welding, the welding procedures typically address all
essential and nonessential variables of Section IX of the ASME Boiler and
Pressure Vessel Code. Only approved welding procedures are typically used.
4.4.4.5 Coating Procedures
Proper preparation and application of a protective coating is important to ensure
its corrosion resistance and longevity. The following submittals may be required
to ensure that the coating is properly installed:

Surface-preparation procedures

Application procedures, including thickness tolerances

Curing procedures

Test procedures and test results

Coating formulation with number and thickness of coats required

Weld and seam repair procedures

Manufacturer's standard data and recommendations for surface preparation,


application, curing, and testing

Quality-control reports and certifications

Temperature, humidity, and ventilation conditions to be maintained during


application and curing

4.4.5 On-Site Testing and Inspection


Ongoing on-site testing and inspection are required throughout the project's
construction phase. Testing confirms that the materials of construction meet
contract requirements. Inspections verify that the Contractor is performing the
work in accordance with specification requirements.
4.4.5.1 Piling
Piling is often used to support the chimney foundation. Several types of piling
systems, such as steel H-piles or auger-cast concrete piles, are commonly used.
Tests and inspection applications for piling are as follows:

4-30

4.4.5.1.1 Load Test


The Contractor is typically required to install test piling to confirm design
conditions and to verify that the installed piling will provide an adequate factor of
safety versus the design loads calculated. The piling is tested for compression,
uplift, and lateral loads, as applicable.
4.4.5.1.2 Grout Cube Tests
A qualified testing laboratory may be retained by the Owner to perform grout
testing for pressure-grouted, auger-cast piling systems. Testing is performed in
accordance with CRD-C 620 and ASTM C109. The results of the cube tests are
used to substantiate evidence of adequate strength and uniform consistency of the
grout.
4.4.5.1.3 Visual Inspection
Piling may be inspected to ensure correct installation and location. Auger-cast
pile reinforcing steel should be inspected before placement. Steel H-piles should
be inspected to confirm that the flanges are oriented properly with respect to the
strong and weak axis direction. Recommended piling-installation tolerances are
as follows:

Piles should not exceed a variation from the vertical of more than 1/4 in
(0.006 m) per foot of pile length.

The center of the pile head should not vary from plan location at the cutoff
by more than 3 inches.

4.4.5.2 Foundation
Inspection of the chimney foundation is required before and during concrete
placement.
4.4.5.2.1 Embedded Items
Placement and positioning of all embedded items should be verified. Examples of
items that are typically embedded in a chimney foundation are:

Anchor bolts

Electrical conduits

Grounding

Drains

4.4.5.2.2 Reinforcing-Bar Placement


Reinforcing steel embedded in or projecting out of the foundation may be
inspected to confirm correct number, size, and spacing. Splice lengths and
dowel-projection lengths may also be verified. If mechanical splice connectors are
4-31

used, the connectors are capped off to provide protection from the wet concrete.
Mud or debris is removed from the rebar before placement of the concrete.
4.4.5.2.3 Concrete Placement
Inspection during concrete placement is performed to ensure that the concrete is
being properly placed. Inspection also confirms that the concrete is vibrated after
its placement. Concrete should be placed in lift heights that permit the concrete
to be vibrated into the lower section. The Contractor should select a placement
sequence so that a minimal amount of concrete surface area is exposed at a given
time. This reduces the opportunity for the concrete to develop its initial set and
maintains a working surface that is easy to vibrate.
4.4.5.2.4 Concrete Testing
Concrete testing is usually performed by a qualified laboratory-testing company.
Before placement, the concrete is tested for temperature, slump, and air content.
Concrete cylinders are taken to perform compression-strength tests.
4.4.5.3 Concrete Column
Inspection of the concrete column is performed during construction to confirm
that the chimney is being constructed in accordance with the design drawings.
4.4.5.3.1 Dimensional Tolerances
Dimensional tolerances for the concrete column are established by ACI 307.
Vertical alignment of the columns center point is taken by plumbing down to a
reference point on the chimney foundation. The column's diameter is taped, and
measurements are taken for out-of-roundness tolerances.
4.4.5.3.2 Location of Openings
The orientation, elevation, and size of openings in the column should be
confirmed. The breeching opening should be accurately located in order to avoid
misalignment problems with the absorber outlet-duct tie-in. Because the
breeching is typically installed before the outlet ductwork, a survey crew should
position the breeching from an established benchmark that can also be used by
the absorber contractor.
4.4.5.3.3 Reinforcing-Bar Placement
Vertical reinforcing bars should be counted and bar sizes checked for each
concrete-lift height to confirm that the bar quantities match the design
requirements. Horizontal bars are checked for size and location. Splice lengths
and concrete cover should also be checked.

4-32

4.4.5.3.4 Wall Thickness


Wall thicknesses are typically measured to verify conformance with the design
requirements.
4.4.5.3.5 Concrete Testing
Concrete testing is typically performed for each lift section (jump form) or for
each 8-hour shift (slip form) by a qualified laboratory-testing company. Testing
is normally performed in the bottom of the chimney at the point of discharge
from the concrete truck. The temperature, slump, and air content of the concrete
are measured and recorded. Concrete cylinders are taken to verify compressive
strength.
4.4.5.4 Acid-Resistant Brick and Mortar
Materials of construction and construction tolerances are typically monitored for
liners constructed of acid-resistant brick and mortar.
4.4.5.4.1 Mortar-Cube Strength Tests
During construction of the brick liner, mortar cubes are taken daily for
compressive-strength tests. This testing is usually performed on-site by a
qualified testing laboratory. It is recommended that two sets of mortar-cube
samples should be tested: one set that is cured in the bottom of the stack to
represent actual in situ conditions during construction, and another set that is
cured under laboratory conditions in accordance with ASTM guidelines. Because
the mortar-curing process is heat-sensitive, compression-strength test results
from mortar cured in the bottom of the stack may indicate whether adequate heat
is provided during construction in cold weather.
4.4.5.4.2 Visual Inspection of the Liner
Inspection during construction of the brick liner may be used to confirm that the
following recommended construction practices are maintained:

Projections on the inside face of the liner should not exceed 1/8 in (0.003 m).

The liner's vertical axis should neither be off the theoretical axis by more
than 0.1 percent of its height nor by 1 in (0.03 m), whichever is greater. Also,
the center of the liner should not vary by more than 1 in (0.03 m) in 10 ft
(3 m).

Neither the diameter nor the radius of the liner should differ by more than
2% from those specified.

During the liner's construction and curing period, a minimum temperature of


50F (10C) should be maintained in the shell's interior.

4-33

4.4.5.5 Metal Liners


On-site inspection of metal liners is primarily related to welding-quality control.
Welding inspection for alloy materials is critical to ensure the corrosion resistance
of the weld in a wet environment.
4.4.5.5.1 Welder Qualifications
For carbon-steel welding, all welders must typically be qualified by passing tests
prescribed in the AWS D1.1 Standard Qualification Procedure. Welders should
have been tested within the past 12 months, and their qualifications may be
considered in effect unless the welder is not engaged in a given process of
welding for more than six months.
For alloy welding, all welders are typically being qualified in accordance with
Section IX of the ASME Boiler and Pressure Vessel Code. In addition, at a
minimum, the tests should meet the radiographic acceptance criteria of QW-191
of Section IX.
4.4.5.5.2 Welding Inspection
Verification inspection and testing of welds for carbon-steel liners may be
performed in accordance with AWS D1.1 and ASCE's Design and Construction of
Steel Chimney Liners. Inspection of material and workmanship may verify
conformance with the AISCs Manual of Steel Construction Specifications and Codes
and the AISCs Quality Criteria and Inspection Standards.
For alloy welding, all welds should be visually inspected by an AWS Certified
Welding Inspector (CWI), qualified per AWS Standard QC1. All alloy welds
that will be exposed to flue gas are tested using liquid-penetrant examination
according to Article 6 of ASME Section V. All liquid-penetrant testing is
performed by an ASNT Level II inspector. Certification verifying the
qualifications of the welding inspector should be submitted before welding
operations start.
4.4.5.5.3 Weld-Quality Workmanship Standards
The Contractor usually prepares and submits workmanship samples for approval
prior to the start of work. The workmanship samples are visually inspected and
liquid penetrant tested prior to submittal. The approved workmanship samples
are visibly displayed at the shop and job site for the benefit of the welders and
inspectors, as examples of acceptable weld surfaces.
4.4.5.6 FRP Liners
Inspection and testing of an FRP liner is typically performed during the liner's
fabrication and installation.

4-34

4.4.5.6.1 Physical Testing


Physical testing may be performed and submitted for representative wall samples.
The following tests may be performed as appropriate:

Resin/glass ratio for interior and exterior layers per ASTM D2584

Barcol hardness for surface cure per ASTM D2583

Tensile modulus per ASTM D638

Flexural modulus per ASTM D790

Compressive modulus per ASTM D695

Coefficient of thermal expansion per ASTM E228

Coefficient of thermal conductivity per ASTM C117 or C518

Specific gravity per ASTM D792

Chemical resistance per ASTM C581 or D4398

Heat-deflection temperature per ASTM D648

Flame retardancy per ASTM E84

4.4.5.6.2 Visual Inspection


Wall samples may be provided by the Contractor for use as representative
samples of visual standards for the liner. The liner appearance is typically free
from defects, including delaminations, cracks, bubbles, pinholes, and other
defects affecting strength or serviceability.
For curing considerations, the air temperature in the work area is maintained at a
minimum of 50F (10C). All resins and FRP components are stored in
accordance with the manufacturer's recommendations. The resin temperature at
its point of use should typically be between 60F (15C) and 75F (24C).
4.4.5.7 Borosilicate-Glass Block
All glass-block lining installations may be subject to inspection. Quality-control
procedures and requirements for installing a borosilicate-glass-block lining
system may be performed in accordance with the manufacturers specifications.
Specific items that are typically inspected or tested are presented below:
4.4.5.7.1 Surface Preparation
For installation on steel liners, any existing fly ash buildup or coating should be
removed, and any abrupt contours or edges on welds should be rounded off
smoothly by grinding. The interior steel-substrate surface should be blast-cleaned
to white metal in accordance with SSPC-SP5.
For installation on brick liners, the condition of the inside surface of the liner
should be inspected. Fly ash buildup may require that the surface be cleaned by
4-35

hydroblast or sandblast. Mortar joints should be inspected. Protruding mortar


should be chipped off.
4.4.5.7.2 Visual for Oxidation
After sandblasting the steel substrate, an inspection should be performed to
confirm that rust bloom on the steel has not developed. A primer should be
applied on the sandblasted steel substrate to protect the steel from the formation
of flash rust during the period of time between completion of sandblasting and
application of the block-lining system.
4.4.5.7.3 Surface-Contamination Tests
The inside surface of the liner should be tested to confirm the removal of any
existing acids. Any existing acid on the steel should be neutralized with an
ammonia-water rinse followed by a light brush-blast cleaning in accordance with
SSPCSP7.
The pH of the blast-cleaned surface should be 6 to 7. The Contractor should
take pH readings in all four quadrants at 20-ft (6-m) intervals throughout the
height of the stack. The steel-substrate surface should also be inspected for
chloride contamination. If the chloride concentration is greater than 10 g/cm2,
the surface is contaminated, and it should be recleaned.
4.4.5.7.4 Workmanship Standards
The installation procedure and workmanship standards of the lining system
should be verified by installing the system on a 4 x 4-ft (1.2 x 1.2-m) transparent
Plexiglas sheet. The test panel should be visually inspected, mallet-tested, and
tested for adhesion. Testing procedures and acceptance criteria should be in
accordance with the manufacturer's requirements.
4.4.5.7.5 Temperature and Humidity
Inspection should be performed to ensure that the lining system is being stored
and installed under the proper temperature and humidity conditions. During the
installation of the lining, the Contractor should measure and record air
temperature, temperature of the surface being lined, moisture dew point, and
relative humidity. Inspection guidelines for these issues are:

Relative humidity in the working area where the lining system is being
installed should not be greater than 90%.

The substrate temperature in which the block lining is installed must be at


least 5F (3C) above the moisture dew point temperature.

Temperature of the substrate, ambient air temperature in the work area, and
curing temperature should be between 50F (10C) and 90F (32C).

4-36

4.4.5.7.6 Visual
Visual inspection of the lining system should include the following items:

All joints must be full and of a 1/8-in (0.003-m) minimum in thickness.

No air voids should exist between the adhesive-coated block and the
adhesive-coated substrate.

The adhesive membrane is trowel-applied.

4.4.5.8 Protective Coatings


All protective-coating installation work is typically inspected. The services of a
quality-control coating inspector certified by NACE are usually required. Items
that are typically inspected or tested are presented in this section.
4.4.5.8.1 Surface Preparation
The interior surface of the liner should be inspected to ensure that surface
preparation is being adequately performed. The surface of the metal to be lined
should be blasted to obtain a white metal surface as defined in SSPC-SP5. The
sand or grit used for blasting should be clean and dry and produce a minimum
surface profile of 2.0 mils when measured with a Keane-Tator comparator,
Model No. 372.
4.4.5.8.2 Visual for Oxidation
After blasting, inspection should confirm that surface rust has not developed. All
surfaces to be coated, as well as coated surfaces requiring additional thickness,
should be primed. All sand, dust, or grit should be removed by brushing, air
blasting, or vacuuming before priming. The blasted surface must be primed
within the same work shift (812 h) and before any visible surface rusting
develops. In the event of surface rusting, the area must be reblasted.
4.4.5.8.3 Surface-Contamination Tests
Surface-contamination tests measuring pH, chloride concentrations, and ferrous
sulfate concentrations should be performed before applying primer. Tests should
be performed using a "kTA Surface Contamination Kit" from kTA-Tator (or
approved equal). If the pH is less than 5.0, the chloride concentration is greater
than 10 g/cm2, or the ferrous sulfate concentration is greater than 20 g/cm2,
the surface is contaminated, and it should be recleaned. High-pressure water or
sodium bicarbonate solution should be used to remove the contamination.
4.4.5.8.4 Thickness
Random wet-film thickness readings should be taken during the coating. A
minimum of four readings should be taken per 100 sq ft (9 m2). After the
coatings are fully cured, dry-film thickness measurements should be taken, based
4-37

on a 6-ft (1.8-m) square grid pattern. Each measurement should consist of the
average of three random readings taken near the center.
4.4.5.8.5 Temperature and Humidity
Inspection should be performed to ensure that the coating system is being stored
and installed under the proper temperature and humidity conditions. During the
installation of the coating, the Contractor should measure and record air
temperature, substrate temperature, moisture dew point, and relative humidity.
Inspection guidelines for these issues are:

Relative humidity in the work area should not be greater than 90%.

Surface temperature must be at least 5F (3C) above the dew point of the air
in the work area.

Temperature in the work area should be 50120F (1049C), or it should be


within the manufacturer's recommendations.

4.4.5.8.6 Spark Test


The cured coating should be tested for discontinuities in accordance with NACE
RP0188-90, Standard Recommended Practice for Discontinuity (Holiday) Testing of
Protective Coatings. Pinholes should be repaired.
4.4.5.9 Elevator-Load Test
The Contractor usually conducts a load test of the personnel elevator, witnessed
by the Owner and Engineer. The Contractor should furnish the Engineer and
Owner a copy of the load-test procedure. The Contractor usually certifies in
writing that the elevator meets all the applicable local, state, and federal codes for
safety and performance.
4.4.5.10 Pressurization Fans
Pressurization fans for brick liners will require on-site testing, as well as start-up
upon delivery to the job site. The Contractor is typically required to perform the
following procedures during the installation of the fans:

Level the fan to the foundation using shim stocks; grout and anchor.

Align the rotating assembly.

Check for proper rotation of motor.

Verify balance of fan.

Perform functional test of pressurization system.

4.4.5.11 Electrical
Completion of the electrical and grounding system is typically one of the last
items to be completed on a new chimney installation. Before demobilization of
4-38

the electrical subcontractor, all electrical and grounding components associated


with the chimney should be checked.
4.4.5.11.1 Obstruction Lights
All obstruction lights should be inspected to verify that they are operational. If
strobe lights are used, the photocell should be checked to verify the correct
daytime and evening intensity levels. The alarm contacts should also be checked
to verify that the lights are not malfunctioning.
4.4.5.11.2 Grounding
The grounding system for the chimney should be checked for continuity. The
ground resistance should be measured and recorded at the ground-cable risers.
All physical connections between the ground cable and the air terminals should
be checked.
4.4.5.11.3 Electrical-System Checkout
The electrical system for the chimney should be inspected. Examples of items
that should be inspected are as follows:

Confirm correct voltage level at the power panel.

Reset transformer taps, if required.

Test each ground-faultinterrupter receptacle to ensure proper operation

Check that all ambient lights are operational.

Verify proper position of disconnect switches and power-panel breakers.

4.4.5.11.4 Lead- and PVC-Coated Components


Grounding and conduit components at the top of the chimney require additional
corrosion protection. A visual inspection should be performed to verify that all
grounding cables and fittings in this area are lead-covered. All conduits located at
the top of the chimney should also be inspected to verify that polyvinyl chloride
(PVC)-coated material was used.

4-39

Section 5: Wet Stack Experience


The original Wet Stacks Design Guide was issued in 1996. Since that time, this
guide has been used as the de facto standard for wet stack design and operation;
hundreds of wet stacks around the world have been designed using its
recommendations. Most of these units operate well, with little if any SLD
experienced under normal operating conditions.
This section of the Revised Wet Stack Design Guide reviews some of the
experience gained in the industry since the issuance of the original guide. To the
extent possible, experience from locations around the world has been
incorporated. Also included is a discussion of the steps that can be taken inhouse if a unit experiences an SLD issueboth regarding finding the source of
the issue and determining the information that should be collected before outside
help is engaged.
5.1 Experience with Wet Stacks
5.1.1 General Operational Experience
A survey of units utilizing wet stacks was performed, and for the most part, it was
found that the majority of them were running well. On occasion, some units
would experience SLD, but typically these events could be traced back to upsets
in the WFGD absorber, maintenance issues such as fouled mist eliminators
and/or plugged drains, or correlated with local weather conditions (for example,
high humidity/low wind).
Units experiencing SLD from the start were biased toward those that had been
retrofitted with WFGD systems but kept the original stacks. For these units, the
stack-liner velocities were not favorable for wet operation from the start, and
some level of SLD was not unexpected. Retrofit installations for units with
multiple absorber towers are also more likely to experience SLD issues, because
these units often have complex absorber-outlet-ducting arrangements and gasrouting schemes requiring complex liquid-collection systems. To minimize costs,
it is not unheard of for utilities to limit the number of liquid collectors actually
installed to those nearest the stack. This practice is not recommended.
A limited number of units, although designed within the recommendations of
the original Wet Stack Design Guide, experienced SLD from the start. It was
found that most of these units had been designed for the higher end of the
recommended liner-gas-velocity range and had some issues with the correct
5-1

installation of the liquid-collection system or the overall quality of the liner


construction. In many instances, simple corrections to the errors in the liquidcollection system installation were sufficient to correct the problem. For alloy
liners, construction-quality issues leading to an increased potential for SLD were
primarily related to the height of horizontal weld seams; for FRP liners, the
problems were related to the height and taper on the internal joints between liner
segments or misalignment/roundness between adjacent segments.
A more recent cause for the onset of SLD issues is related to increases in the
stack-liner velocity brought about by increases in the unit heat rate and, to a
lesser extent, by changes in the fuel source/type/ mix used at the plant. Similar to
the other issues described above, these latter issues occur at plants whose liners
were originally designed for the upper end of the recommended liner-gas
velocities for favorable wet operation.
A limited number of units had also underestimated the total volumetric gas-flow
rate used for the stack-liner design basis. This error resulted in liner diameters
that were too small, with the resulting liner-gas velocities above the
recommended design range for favorable operation.
As can be seen, there is a common thread connected to the maximum
recommended liner flue gas velocity. Units initially designed to operate at the
higher end of the recommended liner-gas velocity are more prone to SLD issues
than those designed for the mid-point or lower end of the range. For this reason,
the recommended velocities for favorable wet operation have been slightly
lowered from those provided in the original Wet Stack Design Guide. It should
also be noted that the new recommendations, previously detailed in Table 2-1,
have been provided as a single design value, not a range. By lowering the
recommended velocities by a few feet per second, the wet stack system will be
significantly more robust and capable of handling the inevitable increases in gas
velocity that will turn an otherwise well-performing stack into one experiencing
some level of SLD.
Although they are only now being codified in this document, these slightly lower
liner-gas-velocity recommendations have effectively been in use over the past few
years. The result has been that significantly fewer units have been experiencing
SLD issues related to the system design and normal unit-operating conditions.
Issues at new plants are generally related to liquid-collectionsystem installation
errors or drain pluggage caused by absorber carryover and/or the buildup of
debris.
An increasing number of plants are utilizing borosilicate-block lining systems. In
this system, the blocks are adhered directly onto a substratetypically, but not
limited to, a carbon-steel flue or the inside of the reinforced-concrete
windscreen, using a flexible membrane material. This material is also used
between the blocks, not unlike the mortar between bricks. Units utilizing this
system appear to be operating well, but they experience many of the same generic
issues as alloy, FRP, and acid-brick liners. It is very important that the blocks are
installed so that they form a smooth surface, without any block edges sticking out
5-2

or recessed away from the main surface. These edges could act as sites for liquid
re-entrainment back into the gas flow. This is also true for the membrane
between the bricks, which must be struck flush with the liner surface. No issues
related to block misalignment have been reported, indicating the existence of
good quality control during the installation process. Very few issues have been
reported when borosilicate blocks have been used in new plants designed from
the start to use the system. Some retrofit applications have had SLD events
related to specific unit-geometry issues, liquid-collection-system drainage issues,
and of course, higher than recommended liner-gas velocities.
Based on discussions with plant operators, most wet stack liquid-collection
systems appear to be operating well. A common theme linking plants without
problems is that they perform liquid-collection-system maintenance on a regular
basisincluding inspections of the absorber mist eliminators, liquid-collection
gutters and baffles, and ring collectors and drains for buildup, pluggage, and
debris. They also clean and flush the system drains. Photographs should be taken
before any cleaning has been performed to document the location and extent of
any buildup within the system. A documented history of the liquid-collectionsystem performance will be very useful should an SLD issue occur.
5.1.2 International Experience with Wet Stacks
The United States was an early adopter of wet stack operation and, for the most
part, wet stack operation has only recently been accepted for use in other areas of
the world. These areas include the EU, countries hoping to join the EU, and
countries in which the plants are being funded by international agencies. As
experienced in the United States, the conversion of older, previously dry, stacks
to wet operation has resulted in a number of units experiencing SLD, primarily
because of unfavorably high liner-gas velocities. Because of this experience, most
new wet-scrubber installations also include a new properly designed wet stack.
These units are using the latest absorber and wet stack technologies, and if
operated according to prescribed procedures, they appear to be operating well.
SLD is not a regulated emission in the United States. As part of the Wet Stack
Design Guide revision process, efforts were made to determine if SLD was
regulated in other parts of the world. To the best of our knowledge, it is not.
However, as in the United States, it is considered a nuisance, and many
countries/regions recommend the installation of a liquid-collection system to
minimize the potential for SLD.
5.1.1.1 European Standard EN 13084
The European Standard for free-standing chimney design, EN 13084 Parts 1-8,
makes no mention of the need for liquid collection.
5.1.1.2 VGB PowerTech
VGB PowerTech Instruction Sheet VGB-M 643 Ue, Chimneys for Operation
without Flue Gas Reheating after FGD, discusses the formation of a liquid film
5-3

on the liner wall and the need for an effective liquid-collection system, although
no specific design details or guidance are provided. As in the EPRI/CICIND
Revised Wet Stack Design Guide, recommendations for the maximum flue gas
velocities inside the liner as a function of the liner material are presented. These
velocities are similar to those recommended in this guide for FRP, alloys, and
coated surfaces; they are higher for acid-resistant brick and slightly lower for
borosilicate block. The need to minimize horizontal discontinuities, as well as to
control plume downwash, are also discussed in this document.
5.1.1.3 China
In 2006 and 2007, more than 100 GWe power-generation capacity was installed
in China. In 2008 through 2010, an additional 130 GWe of capacity was
installed. Many of these plants incorporated FGD systems, and the dominant
technology is wet scrubbers. Most of these units have wet stacks; however there is
no evidence that these plants have incorporated liquid-collection systems.
Discussions with the China Electric Council have revealed that at present there
are no regulations in China controlling SLD from their plants. It is known that
SLD is an issue at many plants in China, and at present there is significant
interestboth at the governmental and the plant levelin the introduction of
liquid-collection technology to these plants.
5.1.1.4 India
No standards or guidelines related to wet stack design and/or operation in India
could be found.
5.2 Before Unit Startup
Before a wet stack is brought on line, a number of reviews and inspections should
be performed to ensure that the liquid-collection and drainage system operates as
expected, with minimal potential for SLD when the unit comes on-line.
5.2.1 Review of Fabrication/Installation Drawings
The first inspection that should be performed is a review of the liquid-collection
system installation drawings. The liquid-collection system design provided by the
laboratory/flow-model vendor is typically limited to providing the size, shape,
and location of the required collection gutters, ring collectors, flow diverters, flow
controls, and drains. These recommendations are then provided to the stack
vendor or architecture engineering (AE) firm, who develops the actual
fabrication and field-installation drawings. These drawings should be reviewed by
the liquid-collection system designer to ensure that their recommendations have
been properly interpreted and that any modifications required for ease of
fabrication or installation do not adversely impact the efficacy of the liquidcollection system.

5-4

5.2.2 Field-Installation Inspection


It is important to ensure that the liquid-collection system has been installed
properly. A second inspection should be performed in the field after the liquidcollection system has been installed, but before the access scaffolding has been
removed from the liner. This way the entire extent of the liquid-collection system
can be closely inspected, and on-the-spot modifications and corrections can be
made while the construction crew and scaffolding is still available.
One of the major sources of SLD from an otherwise well-designed liquidcollection system consists of leaks from the components (which should be sealed
to the duct and liner walls). During the pre-startup inspection, a leak check of
the various ring collectors, gutters, and drainage channels should be performed.
This check can be accomplished by introducing water at the high point of the
liquid-collection system (typically the liner-ring collector) and following the
waters path through the entire drainage system. If leaks are identified, they need
to be sealed before the unit goes on-line.
During the pre-startup inspection, the gutters, ring collectors, and drains should
be inspected and cleared of any debris that might have collected in them during
the construction process. Debris can range from snack wrappers to plastic
sheeting to tools.
5.3 After Unit Startup
Some SLD may be observed during the first few weeks of operation, particularly
from new stacks. Such SLD is a normal occurrence and should be expected. After
fabrication, a new liners interior surface is often coated with a material that is
nonwetting. In alloy liners, this coating may be an oily film or scale from the
mill-rolling process; in FRP liners, it may be a separating medium or gloss
coating. Because these surfaces are initially nonwetting, the liquid film
condensing and depositing on the liner wall will form beads that are easily reentrained back into the gas flow. Over time, the weak acidity of the condensed
liquid will etch or season the liner surface, and it will become wettingallowing
the formation of thin liquid films that bond to the surface via surface tension.
This seasoning process may take a number of weeks to occur, and during this
time some SLD may be experienced.
If any stack-emission incidence occurs during normal operation, the need for
inspection is obvious. If SLD is not experienced, it is still important to inspect
the liquid collectors to ensure satisfactory long-term operation.
A third inspection should be performed at the first scheduled or unscheduled
outage after approximately six months of continuous operation. It is important
that this inspection be performed before the unit is cleaned in any way, because
the solids-deposition patterns, buildup, and water stains on the walls, roof, and
duct floor are all indicators of the systems liquid-collection performance. Special
attention should be paid to the system drains. Different problemssuch as areas
of solids buildup and liquid drainage or re-entrainment problemscan be
5-5

detected. Modifications to the collection system can be designed for installation


during a scheduled or unscheduled outage. This post-startup inspection should
include the absorber outlet, outlet duct, stack liner, and choke/exhaust plane, if
accessible.
The liquid-collection system should be inspected during each outage, and all
drains and liquid-seal pots or loop seals should be cleaned. It is recommended
that a photographic record be kept of the inspections; this record will be
invaluable should an SLD issue arise.
5.3.1 Liquid-Drainage Monitoring
The liquid-flow rate at each drain location should be monitored for the first six
months of the operation after startup. These records should show the quantity of
liquid collected from each of the drains, the unit load during the measurement,
and the ambient temperature at the time the measurements were taken. This
information will be very valuable should an SLD issue be encountered. If this
information is monitored on a continuous basis, trends can be detected and
corrective measures taken before an SLD issue is experienced.
5.4 What to do if SLD Occurs
If SLD remains a persistent problem, the liquid-collection system designer (or
someone skilled in the art) should be contacted to help assess the problem. To
make their initial analysis effective, the following information should be collected
prior to contacting them:

How long has the SLD problem been occurring?

What is the velocity of the flue gas in the stack when SLD occurs?

Was the onset of the SLD issue related to a change in operation or to


procedures at the plant, such as fuel type or absorber operation?

Have there been any recent problems with the absorber chemistry?

Have the mist eliminators been inspected since the SLD issue started, and if
so, was any plugging or fouling observed?

Can the occurrence of SLD be correlated to the mist-eliminator wash cycle?

Has the mist-eliminator wash system been inspected for damage?

Can the occurrence of SLD be correlated to the local weather conditions?

Can the occurrence of SLD be correlated to the time of day?

How far from the stack do the droplets fall?

What is the estimated droplet size?

Have any droplet-size measurements been made in the stack, and if so, what
are the results?

When was the last time the liquid-collection system was inspected? Were
there any issues such as plugging or fouling?
5-6

Has the liquid-collection system drainage-flow rate changed?

This information will provide a basis for evaluation of the SLD issue. Hopefully,
it will provide the investigators with sufficient information to begin their
assessment of the cause of the issue and its elimination.
Specific actions can also be taken. The first should be to measure the liquid-flow
rate exiting the stack through the liquid-collection system drains. If these flow
rates are significantly lower than the baseline or calculated values, it is likely that
the drain system is plugged. An inspection of the liquid-collection system and
drains should be made at the earliest opportunity. The entire drainage path
should be inspected, but particular attention should be paid to areas where the
collected liquid is constricted. Such areas include drain ports; locations where
liquid enters a drainage pathway, such as found on internally drained ring
collectors; and (on units with side-entry breaches) the drains located on either
end of the breach-top gutter. Past experience has shown that debris tends to
collect in these areas. At some units experiencing unexpected SLD, these areas
have been found to be plugged with plastic sheeting; pieces of the plastic film
used as a separating media during FRP liner fabrication that was not fully
removed prior to startup; and sometimes even dead birds. These locations are the
first places that should be inspected if SLD is encountered.
During this inspection, the mist eliminators and the mist-eliminator wash system
should also be inspected for pluggage and damage. Plugged mist eliminators can
result in higher than design velocity values through the unit, which could lead to
liquid breakthrough and increased levels of droplets re-entrained in the flue gas
flow. Some droplets could also be re-entrained directly from the edges of the
plugged areas of the mist eliminator. Some of these droplets will not be collected
in the absorber-outlet ducting or liner and will pass through the unit, exiting
from the top of the stack.
5.4.1 Preliminary Stack Droplet Testing
The location of the droplets within the liner can often be related to the source of
their generation. Approximate droplet-size measurements and their spatial
distribution within a few feet of the liner wall can be estimated using a simple
droplet-impact probe made from a 1012-ft long 2 x 4 board painted flat black,
the smoother the surface the better (Figure 5-1).
The probe should be oriented with the test side facing horizontally, then rapidly
inserted through a test port between 2 and 8 ft (0.6 and 2.5 m). The probe is
then quickly rotated 90 to place the test face directly into the upward-rising gas
flow, and it is left in this orientation for 34 seconds. The probe is then rotated
back to its original orientation and rapidly removed from the liner (Figure 5-2).

5-7

~2
~24

12

4 Hard Plastic Tube

Impact
Probe

Stack Linier

Figure 5-1
Droplet Probe

Insertion
Orientation

Test
Orientation

Withdrawal
Orientation

Test Surface

Gas Flow
Figure 5-2
Droplet-Probe Orientation

Droplet spots will be observed on the surface of the probe (Figure 5-3). The
probe exposure time may need to be increased or decreased, depending on the
number of droplets captured. The objective is to get enough droplets so that the
droplet spots do not overlap and that enough spots are obtained to have some
statistical significance. Tests should be repeated multiple times. The diameter of
the spots should be measured and their distribution on the probe noted. Large
droplets occurring in the area adjacent to the liner surface indicate that the issue
could be related to droplets being re-entrained from the liner surface. Large
droplets located farther from the liner wall are indicative of droplet carryover
5-8

from the mist eliminators or of droplet re-entrainment from somewhere in the


absorber-outlet ducting.
The probe testing should be performed at multiple locations around the
perimeter of the liner to determine if the droplets are evenly distributed across
the liner or are biased toward one side. A biased distribution can provide further
evidence as to the actual source of the droplets. If accessible, this technique can
also be performed at the stack outlet.
After each insertion, the probe should be wiped down and allowed to dry before
being reinserted into the liner.

Figure 5-3
Typical Droplet-Probe Test Results

The use of this probe technique is not intended to replace more sophisticated
evaluation methods. However it is a quick, cost-effective way to get a better
picture of what is going on within the liner during the initial stages of evaluating
the sources leading to SLD.
An additional test that can provide valuable insight into a stack experiencing
undesirable levels of SLD is to collect droplet samples downwind of the stack at
ground level. This test can be performed by placing metal collection pans, such as
a cookie sheet or a square piece of plywood covered in aluminum foil, on the
ground in the area where the droplets are hitting the ground. This approach can
be used to establish the number and size of droplets collected as a function of
distance to stack and unit load.
5-9

Section 6: References
1.

Entrainment in Wet Stacks, EPRI, Palo Alto, CA: 1982. CS-2520.

2.

Wet Stacks Design Guide. EPRI, Palo Alto, CA: 1996. TR-107099.

3.

FGD Mist Eliminator System Design and Specification Guide. EPRI, Palo
Alto, CA: 1993. GS-6984.

4.

Guidelines for the Fluid Dynamic Design of Power Plant Ducts. EPRI, Palo
Alto, CA: 1998. TR-109380.

5.

Task Committee on Steel Chimney Liners, Fossil Power Committee,


Power Division, American Society of Civil Engineers, Design and
Construction of Steel Chimney Liners. New York, NY, 1975.

6.

D. K. Anderson, P. E., The Planning and Design of Effective Wet Duct/Stack


Systems for Coal Fired Utility Power Plants. CICIND Report, Vol. 24, No.2,
July 2008, pp.3742.

7.

American Society of Testing and Materials, Standard Guide for Design,


Fabrication, and Erection of Fiberglass Reinforced (FRP) Plastic Chimney
Liners with Coal Fired Units. (ASTM D 5364-08), West Conshohocken,
PA, November 2008.

8.

American Concrete Institute, Code Requirements for Reinforced Concrete


Chimneys. (ACI 30708), Farmington Hills, MI, November 2008.

9.

VGB PowerTech Service GmbH, Chimneys for Operation without Flue Gas
Reheating After FGD. Instruction Sheet VGB-M 643-Ue. Essen, Germany,
February 2007.

10. Bernhardt Hertlein, ed., American Society of Civil Engineers, Chimney and
Stack Inspection Guidelines. Reston, VA, 2003.
11. Guidelines for FGD Materials Selection and Corrosion Protection. EPRI, Palo
Alto, CA: TR-100680, Vols. 1 and 2.
12. Acid Deposition on Ductwork. EPRI, Palo Alto, CA: 1983. CS-3240.

6-1

13. American Society for Testing and Materials, Standard Specification for
Design and Fabrication of Flue Gas Desulfurization System Components for
Protective Lining Application. (ASTM D4618-92), Philadelphia, PA, 2010.
14. T. S. Clark. Chimneys Subjected to Acid Gases, Industrial and
Engineering Chemistry. Vol.15, No.3, March 1923, pp. 227230.
15. R. Mongia, A. Reza, et al. Effect of Exhaust Stack Geometry on the
Amount of Liquid Condensate During Plant Start-Up, Proceedings of the
95th Annual Conference of Air and Waste Management Association. Baltimore,
MD (2002).
16. H. S. Rosenberg. Wet Stacks: Friend or Foe?, Power Engineering.
November 1998, pp.7681.
17. American Society for Testing and Materials, Standard Specification for
Inspection of Linings in Operating Flue Gas Desulfurization Systems. (ASTM
D4619-96), Philadelphia, PA, 2004.
18. American Society for Testing and Materials, Standard Guide for Design and
Construction of Brick Liners for Industrial Chimneys. (ASTM C1298-95),
Philadelphia, PA, 2007.
19. N. J. Gardner and I. Owen, The Behavior of Liquid Films and Drops in
Relation to Liquid/Gas Separators, Proc. Institute Mechanical Engineers.
Vol. 211, Part E, 1997, pp. 5359.
20. D. S. Miller, Internal Flow Systems. 2nd Edition, Gulf Publishing, Houston,
TX 1990.
21. Leaning Brick Stack Liners. EPRI, Palo Alto, CA: 1989. GS-6520.

6-2

Appendix A: Glossary
A.1 Definitions
absolute humidity - the weight (or mass) of water vapor in a gas water-vapor
mixture per unit volume of space occupied.
absorber - general term for those gas/liquid contacting devices designed primarily
for the removal of SOx pollutants, i.e., scrubber.
absorption - the process by which gas molecules are transferred to a liquid phase
during scrubbing.
air cubic feet per minute (acfm) - a gas-flow rate expressed with respect to
operating conditions (temperature and pressure).
ambient - pertaining to the conditions (pressure, air quality, temperature, etc.) of
the surrounding environment of a plant or scrubbing system.
annual outage - a scheduled period of time (generally four to six weeks) set aside
by the utility once per year to shut down the boiler and/or FGD system for
inspection and maintenance.
annulus - the space between a chimney liner and a chimney shell.
base load - a generating station that is normally operated to take all or part of the
normal load of a system and that, consequently, operates at a constant output.
breeching - the section of ductwork in an FGD system between the absorberoutlet duct and the stack.
British thermal unit (Btu) - the amount of heat required to raise the temperature
of one pound of water 1F, averaged from 32212F.
bypass gas - flue gas that bypasses a scrubber for the purpose of raising wet flue
gas temperatures above the saturation temperature.
bypass reheat - a system that increases the temperature of the saturated flue gas
leaving an FGD system above dew point by ducting a slipstream of particlecleaned flue gas from the ESP exit duct past the FGD system to the absorberoutlet duct or directly to the stack.
A-1

capacity factor - the ratio of the average load on a boiler for the period of time
considered to the capacity rating of the boiler (actual kWh produced/theoretical
kWh produced x 100).
carryover - entrained solids, slurry droplets, and/or liquid droplets that leave with
the flue gas stream exiting a particular stage of a scrubber or absorber.
chimney - a vertical structure at a power plant that encloses one or more flues
which exhaust combustion gases. A chimney is typically constructed out of
reinforced concrete.
chloride - a compound of chlorine with another element or radical.
choke - the constricted upper section of a steel stack or chimney liner.
cladding - a thin sheet of corrosion-resistant alloy (usually nickel alloy or
titanium) that is either resistance-welded or roll-bonded to carbon-steel plate.
closed water loop - the water loop of an FGD system is closed when the fresh
makeup water added exactly equals the evaporative water loss leaving via the
stack and the water chemically or physically bonded to the sludge product.
column - a reinforced-concrete chimney shell. Usually encloses one or more
chimney liners. The purpose of the column is to protect the liner from weather
and to act as a wind shield.
computational fluid dynamics (CFD) - the use of finite elementanalysis
methods to simulate heat transfer, temperature profiles, and fluid and particle
movement in boilers and air-pollution- control equipment.
corrosion - the deterioration of a metallic material by electrochemical attack.
damper - a plate or set of plates or louvers in a duct used to stop or regulate gas
flow.
dew point - the temperature at which vapor contained in saturated flue gas begins
to condense.
efficiency - ratio of the amount of a pollutant removed to the total amount
introduced to the normal operation.
entrainment - the suspension of solids, liquid droplets, or mist in a gas stream.
erosion - the action or process of wearing away of a material by physical means
(friction).
electrostatic precipitator (ESP) - an air-pollution device used to remove particles
from an exhaust stream by initially charging them with electrodes and then
collecting them on oppositely charged plates.
A-2

expansion joint - a small section of ductwork or piping that is designed to


passively expand or contract as required by the flexing of more rigid duct runs,
piping, or pieces of equipment as such components are exposed to varying
external and internal temperatures.
flue gas desulfurization (FGD) system - an SO2 removal system that uses a wet
or dry process downstream of a boiler to reduce sulfur dioxide emissions.
fly ash - fine solid particles of noncombustible ash carried out of the boiler by the
exiting flue gas.
forced outage - the FGD system is taken out or forced out of service to make
necessary repairs or modifications regardless of boiler availability, such that the
system is unavailable for service.
heat exchanger - device used to transfer sensible and/or latent heat from one
stream of material to another to raise or lower the temperature of one of the
materials.
induced draft (ID) - a fan used to move an enclosed stream of gas by creating a
negative relative pressure in the stream to effectively draw the gas through the
system.
indirect hot air - a flue gas reheat system in which reheat is achieved by heating
ambient air with an external heat exchanger using steam at temperatures of 350
450.
in-line reheater - a heat exchanger installed in the wet flue gas duct downstream
of the mist eliminator, usually consisting of hot water or steam coils used to boost
the wet flue gas temperature above dew point.
liner - the flue located inside of a chimney or stack that exhausts combustion
gases.
lining - a metal, organic, or inorganic type of material applied to a shell of an
FGD system component that is intended to protect the shell from abrasion, heat,
and/or corrosion.
liquid-collection devices - devices such as gutters, troughs, and drains used in
absorber-outlet ductwork or stack liners within an FGD system to collect liquid
carried over from a wet scrubber.
load factor - the ratio of the average load in kilowatts supplied during a
designated period to the peak or maximum load in kilowatts occurring in that
period.
mist - dispersion of liquid particles in a gas stream, carryover from a gas-liquid
contact operation.

A-3

mist eliminator - a piece or section of pollution hardware used to remove a


dispersion of liquid particles from a gas stream.
megawatt (MW) - unit used to describe the gross or net power generation of a
particular facility. One watt equals one joule per second. One megawatt equals
106 watts.
micrometer - unit of measure equivalent to 0.0000394 in or 0.001 mm.
new (as opposed to retrofit, for FGD systems) - the FGD unit and boiler were
designed at the same time, or space for the addition of an FGD unit was reserved
when the boiler was constructed.
NOx - a symbol meaning oxides of nitrogen (e.g., NO and NO2).
New Source Performance Standards (NSPS) - environmental regulations that
apply to a new installation, referring primarily to the Federal NSPS that applies
to installations beginning construction on or after August 17, 1971.
opacity - the degree to which emissions reduce the transmission of light and
obscure the view of an object in the background.
open water loop - the water loop of an FGD system is open when the fresh
makeup water added exceeds the evaporative water loss leaving via the stack and
the water chemically or physically bonded to the sludge product.
outage - that period of time when the boiler and/or FGD system is shut down
for inspection and maintenance. Outages may be either forced or scheduled.
particulate matter - finely divided solid particles entrained in the gas stream (fly
ash, coal fines, dried reaction byproducts, etc.).
peak load - a boiler that is normally operated to provide power during maximum
load periods.
pH - the hydrogen ion concentration of a water or slurry to denote acidity or
alkalinity.
plume (stack plume) - the visible emission from a flue (stack).
plume downwash - the phenomenon that occurs when the flue gas exits a stack
and the vapor plume drops below the top of the stack before evaporating or
dispersing into the atmosphere. Usually occurs on stacks that operate at a
relatively low exit velocity.
parts per million (ppm) - unit of concentration that in wastewater applications is
equal to milligrams per liter and in air-pollution applications is equal to moles of
pollutant to million moles diluent.

A-4

pressure drop - the difference in force per unit area between two points in a fluid
stream as a result of resistive losses in the stream.
rain hood - the component at the top of a stack that covers the annular space.
reheat - the process of increasing the flue gas temperature downstream of a wet
scrubber. Reheat can be supplied by in-line indirect hot air, direct combustion, or
by partial bypass of unscrubbed flue gas.
reheater - device used to raise the temperature of the scrubbed gas stream to
prevent condensation and corrosion of downstream equipment, avoid visible
plume, and/or enhance plume rise and dispersion.
relative humidity (also relative saturation) - the ratio of the weight (or mass) of
water vapor present in a unit volume of gas to the maximum possible weight (or
mass) of water vapor in unit volume of the same gas at the same temperature and
pressure. The term "saturation" refers to any gas-vapor combination, whereas
"humidity" specifically refers to an air-water system.
removal efficiency:
-

particulate matter - the actual percentage of particulate matter removed


by the emission-control system (mechanical collectors, ESP, or fabric
filter and FGD) from the
untreated flue gas.

SO2 - the actual percentage of SO2 removed from the flue gas by the
FGD system.

total unit design - the designed percentage of mass of SO2 or particulate


matter entering the stack to the mass of the material in the flue gas
exiting the boiler, regardless of the
removal efficiency of an
individual component or the percentage of the exiting flue gas actually
being scrubbed.

retrofit - the FGD unit will be/was added to an existing boiler not specifically
designed to accommodate an FGD system.
saturated - the situation in which a gas or liquid is filled to capacity with a certain
substance. No additional amount of the same substance can be added under the
given conditions.
saturation temperature - the temperature to which flue gas drops when it is
saturated by scrubbing in a wet FGD system.
scale - deposits of slurry solids (calcium sulfite or calcium sulfate) that adhere to
the surfaces of FGD equipment, particularly absorber/scrubber internals and
mist-eliminator surfaces.
scheduled outage - a planned period of time periodically set aside for inspection
and maintenance of the boiler and/or FGD system.
A-5

scrubber - a device that promotes the removal of pollutant particles and/or gases
from exhaust streams of combustion or industrial processes by the injection of an
aqueous solution or slurry into the gas stream, i.e., absorber.
sludge - the material containing high concentrations of precipitated reaction
byproducts and solid matter collected and/or formed by the FGD process
(composed primarily of calcium-based reaction byproducts, excess scrubbing
reagent, fly ash, and scrubber liquor).
slurry - a watery mixture of insoluble matter (usually lime or limestone).
SOx - a symbol meaning oxides of sulfur (e.g., SO2 and SO3).
stack - a vertical structure at a power plant that encloses one or more flues which
exhaust combustion gases.
stack flue - the inner duct or liner in a stack through which the flue gas is
conveyed.
stack-exit velocity - the exiting velocity of the flue gas out the top of the stack.
stack-liner velocity - area average gas velocity inside the liner.
stack liquid discharge (SLD) - liquid that is discharged from a stack and falls to
the ground prior to evaporating.
standard conditions - a set of physical constants for the comparison of different
gas volume flow rates (68F, 29.92 in Hg, barometric pressure).
standard cubic feet per minute (scfm) - units of gas-flow rate at standard
conditions.
steel stack - a vertical structure at a power plant that exhausts combustion gases.
The primary supporting shell is made of steel.
superficial gas velocity - the area average flue gas velocity through a mist
eliminator or other component of an FGD system.
temperature, dry bulb (DB) - the temperature of a gas or mixture of gases
indicated by a thermometer after correction for radiation.
temperature, wet bulb (WB) - a measure of the moisture content of air (gas)
indicated by a wet bulb psychrometer.
total controlled capacity (TCC) - the gross rating (MW) of a unit brought into
compliance with FGD, regardless of the percent of flue gas treated at the facility.
turning vanes (i.e., vanes) - devices used in ductwork or chimney liners to control
gas-flow direction. Usually fabricated from flat or curved plates.
A-6

unit rating:
-

gross - maximum continuous generating capacity in MW

net - gross unit rating less the energy required to operate ancillary station
equipment, inclusive of emission-control systems.

video droplet analyzer (VDA) - uses on-the-fly video image analysis to detect
and measure the diameters of all in-focus droplets that are entirely within the
view of the camera in each video frame
wallpaper - thin sheets of corrosion-resistant alloy material welded to new or
existing carbon-steel plate.
water loop - all aqueous mass flows from inlet (e.g., seal water, quench water,
scrubber liquor) to outlet of an FGD system (e.g., evaporation via stack, pond
evaporation, waste disposal).
wet stack - a chimney, stack, or flue that exhausts saturated, completely scrubbed
flue gas. Wet stacks are located downstream from a wet FGD system. Wet stack
operation does not utilize any flue gas reheat system or partial bypass. Wet stacks
are equipped with corrosion-resistant liners for handling the wet, acidic flue gas
exiting the FGD system.
zero discharge - a pollution regulation requiring that no effluent waste stream be
discharged back into the environment, with the exception of evaporation via
ponds and stacks (e.g., pond runoff or direct piping of spent slurry or waste into
nearby waterways or tributaries would be prohibited).

A-7

A.2

Units and Conversion Factors

To Obtain

Multiply

By

Atmospheres

Feet of water @ 4C

0.0295

Atmospheres

Inches of mercury @ 0C

0.03314

Atmospheres

Pounds per square inch

0.068

Cubic meters

Cubic feet

0.02831685

Cubic meters

Gallons

0.00378541

Cubic meters
per second

Cubic feet per minute

0.0004719

Inches of mercury
@ 0C

Pounds per square inch

2.036

Kilograms per
square meter

Pounds per square foot

4.882

Kilowatts

Btu per minute

0.01757

Millimeter

Inches

25.4

Kilograms

Pounds

0.4565924

Kilogram per
cubic meter

Pound per cubic foot

16.01846

Liters

Gallons

3.785

Meters

Feet

0.3048

Meters per second

Feet per second

0.3048

Pascal

Pounds per square inch

6,894.757

Pascal

Pounds per square foot

47.88

Square meters

Square feet

0.09290304

Temperature Conversions:
o

F = 1.8 x C + 32

C = ( F - 32)/1.8

A-8

Liquid Load Conversion:


Liquid Load (grains / acf) = QL(gal / minx57,700)
QG(acfm)
6

(gal/min) = (grains/acf) x 17.24 x (acfm/10 )


2

(grains/acf) = (gpm/ft ) x (970.0/(ft/s))


3

(mg/m ) = 2,288 x (grains/acf)

A-9

Export Control Restrictions

The Electric Power Research Institute Inc., (EPRI, www.epri.com)

Access to and use of EPRI Intellectual Property is granted with the spe-

conducts research and development relating to the generation, delivery

cific understanding and requirement that responsibility for ensuring full

and use of electricity for the benefit of the public. An independent,

compliance with all applicable U.S. and foreign export laws and regu-

nonprofit organization, EPRI brings together its scientists and engineers

lations is being undertaken by you and your company. This includes

as well as experts from academia and industry to help address

an obligation to ensure that any individual receiving access hereunder

challenges in electricity, including reliability, efficiency, health, safety and

who is not a U.S. citizen or permanent U.S. resident is permitted access

the environment. EPRI also provides technology, policy and economic

under applicable U.S. and foreign export laws and regulations. In the

analyses to drive long-range research and development planning, and

event you are uncertain whether you or your company may lawfully

supports research in emerging technologies. EPRIs members represent

obtain access to this EPRI Intellectual Property, you acknowledge that it

approximately 90 percent of the electricity generated and delivered

is your obligation to consult with your companys legal counsel to deter-

in the United States, and international participation extends to more

mine whether this access is lawful. Although EPRI may make available

than 30 countries. EPRIs principal offices and laboratories are located

on a case-by-case basis an informal assessment of the applicable U.S.

in Palo Alto, Calif.; Charlotte, N.C.; Knoxville, Tenn.; and Lenox, Mass.

export classification for specific EPRI Intellectual Property, you and your
company acknowledge that this assessment is solely for informational

Together...Shaping the Future of Electricity

purposes and not for reliance purposes. You and your company acknowledge that it is still the obligation of you and your company to make
your own assessment of the applicable U.S. export classification and
ensure compliance accordingly. You and your company understand and
acknowledge your obligations to make a prompt report to EPRI and the
appropriate authorities regarding any access to or use of EPRI Intellectual Property hereunder that may be in violation of applicable U.S. or
foreign export laws or regulations.

Programs:
Integrated Environmental Controls

2012 Electric Power Research Institute (EPRI), Inc. All rights reserved. Electric Power
Research Institute, EPRI, and TOGETHER...SHAPING THE FUTURE OF ELECTRICITY are
registered service marks of the Electric Power Research Institute, Inc.

1026742

Electric Power Research Institute


3420 Hillview Avenue, Palo Alto, California 94304-1338 PO Box 10412, Palo Alto, California 94303-0813 USA
800.313.3774 650.855.2121 askepri@epri.com www.epri.com

Вам также может понравиться