Sie sind auf Seite 1von 194

SINGLE USER LICENSE AGREEMENT

THIS IS A LEGALLY BINDING AGREEMENT BETWEEN YOU AND THE ELECTRIC POWER RESEARCH INSTITUTE, INC.
(EPRI). PLEASE READ IT CAREFULLY BEFORE BREAKING OR TEARING THE WARNING LABEL AND OPENING THIS
SEALED PACKAGE.
BY OPENING THIS SEALED REPORT YOU ARE AGREEING TO THE TERMS OF THIS AGREEMENT. IF YOU DO NOT AGREE
TO THE TERMS OF THIS AGREEMENT, PROMPTLY RETURN THE UNOPENED REPORT TO EPRI AND THE PURCHASE
PRICE WILL BE REFUNDED.

1. GRANT OF LICENSE
EPRI grants you the nonexclusive and nontransferable right during the term of this agreement to use this
report only for your own benefit and the benefit of your organization. This means that the following may use
this report: (I) your company (at any site owned or operated by your company); (II) its subsidiaries or other
related entities; and (III) a consultant to your company or related entities, if the consultant has entered into
a contract agreeing not to disclose the report outside of its organization or to use the report for its own
benefit or the benefit of any party other than your company.
This shrink-wrap license agreement is subordinate to the terms of the Master Utility License Agreement
between most U.S. EPRI member utilities and EPRI. Any EPRI member utility that does not have a Master
Utility License Agreement may get one on request.
2. COPYRIGHT
This report, including the information contained in it, is either licensed to EPRI or owned by EPRI and is protected by United States and international copyright laws. You may not, without the prior written permission
of EPRI, reproduce, translate or modify this report, in any form, in whole or in part, or prepare any derivative work based on this report.
3. RESTRICTIONS
You may not rent, lease, license, disclose or give this report to any person or organization, or use the information contained in this report, for the benefit of any third party or for any purpose other than as specified
above unless such use is with the prior written permission of EPRI. You agree to take all reasonable steps
to prevent unauthorized disclosure or use of this report. Except as specified above, this agreement does not
grant you any right to patents, copyrights, trade secrets, trade names, trademarks or any other intellectual
property, rights or licenses in respect of this report.
4. TERM AND TERMINATION
This license and this agreement are effective until terminated. You may terminate them at any time by
destroying this report. EPRI has the right to terminate the license and this agreement immediately if you fail
to comply with any term or condition of this agreement. Upon any termination you may destroy this report,
but all obligations of nondisclosure will remain in effect.
5. DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES
NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, NOR ANY PERSON OR ORGANIZATION 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 REPORT, 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 PARTYS INTELLECTUAL PROPERTY, OR (III) THAT THIS REPORT IS SUITABLE TO ANY
PARTICULAR USERS 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 REPORT
OR ANY INFORMATION, APPARATUS, METHOD, PROCESS OR SIMILAR ITEM DISCLOSED IN THIS
REPORT.
6. EXPORT
The laws and regulations of the United States restrict the export and re-export of any portion of this report,
and you agree not to export or re-export this report or any related technical data in any form without the
appropriate United States and foreign government approvals.
7. CHOICE OF LAW
This agreement will be governed by the laws of the State of California as applied to transactions taking place
entirely in California between California residents.
8. INTEGRATION
You have read and understand this agreement, and acknowledge that it is the final, complete and exclusive
agreement between you and EPRI concerning its subject matter, superseding any prior related understanding or agreement. No waiver, variation or different terms of this agreement will be enforceable against EPRI
unless EPRI gives its prior written consent, signed by an officer of EPRI.

Guidelines for Upgrading


Electrostatic Precipitator
Performance
Volume 1: Optimizing an Existing Electrostatic
Precipitator
TR-113582-V1

Final Report, September 1999

EPRI Project Manager


R. Altman

EPRI 3412 Hillview Avenue, Palo Alto, California 94304 PO Box 10412, Palo Alto, California 94303 USA
800.313.3774 650.855.2121 askepri@epri.com www.epri.com

DISCLAIMER OF WARRANTIES AND LIMITATION OF LIABILITIES


THIS DOCUMENT WAS PREPARED BY THE ORGANIZATION(S) NAMED BELOW AS AN
ACCOUNT OF WORK SPONSORED OR COSPONSORED BY THE ELECTRIC POWER RESEARCH
INSTITUTE, INC. (EPRI). NEITHER EPRI, ANY MEMBER OF EPRI, ANY COSPONSOR, THE
ORGANIZATION(S) BELOW, NOR ANY PERSON ACTING ON BEHALF OF ANY OF THEM:
(A) MAKES ANY WARRANTY OR REPRESENTATION WHATSOEVER, EXPRESS OR IMPLIED, (I)
WITH RESPECT TO THE USE OF ANY INFORMATION, APPARATUS, METHOD, PROCESS, OR
SIMILAR ITEM DISCLOSED IN THIS DOCUMENT, INCLUDING MERCHANTABILITY AND FITNESS
FOR A PARTICULAR PURPOSE, OR (II) THAT SUCH USE DOES NOT INFRINGE ON OR
INTERFERE WITH PRIVATELY OWNED RIGHTS, INCLUDING ANY PARTY'S INTELLECTUAL
PROPERTY, OR (III) THAT THIS DOCUMENT IS SUITABLE TO ANY PARTICULAR USER'S
CIRCUMSTANCE; OR
(B) ASSUMES RESPONSIBILITY FOR ANY DAMAGES OR OTHER LIABILITY WHATSOEVER
(INCLUDING ANY CONSEQUENTIAL DAMAGES, EVEN IF EPRI OR ANY EPRI REPRESENTATIVE
HAS BEEN ADVISED OF THE POSSIBILITY OF SUCH DAMAGES) RESULTING FROM YOUR
SELECTION OR USE OF THIS DOCUMENT OR ANY INFORMATION, APPARATUS, METHOD,
PROCESS, OR SIMILAR ITEM DISCLOSED IN THIS DOCUMENT.
ORGANIZATION(S) THAT PREPARED THIS DOCUMENT
Grady Nichols Enterprises, Inc.
Black and Veatch, Inc.
Southern Research Institute
Bevilacqua Knight, Inc.

ORDERING INFORMATION
Requests for copies of this report should be directed to the EPRI Distribution Center, 207 Coggins
Drive, P.O. Box 23205, Pleasant Hill, CA 94523, (800) 313-3774.
Electric Power Research Institute and EPRI are registered service marks of the Electric Power
Research Institute, Inc. EPRI. POWERING PROGRESS is a service mark of the Electric Power
Research Institute, Inc.
Copyright 1999 Electric Power Research Institute, Inc. All rights reserved.

CITATIONS
This report was prepared by
Grady Nichols Enterprises, Inc.
400 Kiowa Street
Montevallo, Alabama 35115
Principal Investigator
G. Nichols
The following consultants provided major sections of the report
J. Gooch
A. Ferguson
G.H. Marchant
Bevilacqua Knight, Inc. were instrumental in organizing and writing the document.
This report describes research sponsored by EPRI.
The report is a corporate document that should be cited in the literature in the following manner:
Guidelines for Upgrading Electrostatic Precipitator Performance: Volume 1, Optimizing an
Existing Electrostatic Precipitator, EPRI, Palo Alto, CA,: 1999. TR-113582-V1.

iii

REPORT SUMMARY

The first of a two-volume set, this guide presents a systematic procedure to optimize a
chronically under-performing electrostatic precipitator (ESP) without conducting a major
upgrade. The guide focuses on ESPs that require only moderate improvements (less than $10
$20/kW) to achieve their emissions goals. The second volume of this report, which will appear at
the end of 1999, will cover more extensive upgrades, as well as flue gas conditioning.
Background
Various factors are spurring power producers to improve the performance of their electrostatic
precipitators. Equipment aging causes ESP performance to drop off from original levels, even as
emissions limits become more stringent. New PM25 standards under consideration could require
increased capture of fine particles, which are hardest for ESPs to collect. SO2 compliance
measures are another driver: low-sulfur coals generally produce fly ash particles with higher
electrical resistivity, which are more difficult for an ESP to collect and cause ESP performance
to drop as a result of the coal switch. Another compliance option, dry SO2 control systems,
dramatically increases the mass loading into the precipitator and can also alter the resistivity of
fly ash such that particle reentrainment becomes a serious problemagain reducing ESP
collection efficiency. Power plant engineers need tools for determining the most strategic
repair/replace/redesign option to meet their particulate emission targets.
Objective
To provide plant operators and engineers with a systematic method for (1) determining whether
emissions limits can be achieved with an existing ESP or whether an upgrade will be necessary;
(2) diagnosing the cause(s) of ESP under-performance; and (3) identifying the best corrective
actions.
Approach
The project team, comprised of experts on ESP design and operation, summarized published
EPRI research and drew on their own expertise to define systematic procedures to optimize
chronically under-performing ESPs without conducting a major upgrade. They produced a guide
that starts with a quick screening procedureusing computer-generated charts based on field
data from coal-fired unitsto determine whether emissions goals can be achieved by optimizing
the existing precipitator or whether an equipment upgrade will be needed. For units that can get
by with the current equipment, the remainder of the manual provides step-wise instructions for
determining the cause(s) of suboptimal performance and for taking corrective actions. The team
designed the manual to be used with or without an ESP modelsuch as ESPM or ESPert
although a model is recommended.

Results
The guide includes step-by-step procedures for diagnosing the cause(s) of suboptimal
performance, as well as a discussion of corrective measures. On-line diagnostics include
assessment of the rapping system and tests to identify problems such as back corona and
misalignment. Off-line diagnostics include air load tests under dirty and clean conditions,
physical inspections, and gas flow velocity measurements. The guide provides numerous
illustrations, including a data entry checklist for using an ESP model, a component checklist for
inspections, and sample opacity traces under normal and abnormal conditions. Special emphasis
is placed on interpreting electrical readings, with illustrations of power supply meter displays,
oscilloscope waveforms, and V-I curves under normal and abnormal conditions. Appendices
discuss the physics behind ESP operation, and can serve as a training guide or refresher.
EPRI Perspective
Power producers face the difficult task of meeting increasingly stringent pollution abatement
regulations while simultaneously reducing costs to compete in a deregulated market. Adding to
the challenge, various measures to meet acid rain regulations have had a negative impact on ESP
performance. This performance drop, combined with the new compliance assurance monitoring
(CAM) regulation for particulate emissions, requires many utilities to restore, or even improve,
precipitator performance to stay in compliance. EPRI has developed a vast body of information
to help its members meet this challenge cost-effectively. Much of that information has been
synthesized into this report, to help plant operators get the most out of their existing ESPs
(Volume 1). If further improvement is needed, power producers can consider the newest upgrade
options, analyzed and presented in Volume 2 to identify the least-expensive option for their
situation. Together, these two volumes help the user streamline the evaluation process and
identify the optimum solution to any ESP performance problem.
TR113582-V1
Keywords
Air Pollution Equipment Upgrades
Electrostatic Precipitator Optimization
ESP Upgrade
ESP Troubleshooting
ESP Performance Improvement
Particulate Control

vi

EPRI Licensed Material

ACKNOWLEDGMENTS
Gerry Klemm and Wallis Harrison of the Southern Company made many valuable contributions
to this guidebook. They contributed to the original discussions on how the manual should be
organized, and reviewed drafts of the document, providing valuable guidance and additional
material for inclusion. Both provided valuable insight into the sections dedicated to internal
inspections and troubleshooting. Gerry Klemm provided the original figures with appropriate
discussion about the behavior of power supplies and controls as well as the opacity traces used to
illustrate the ESP rapper behavior. Special thanks to them.
Dr. John P. Gooch and Mr. G. H. (Wim) Marchant of Southern Research Institute provided the
information about determining the current performance of an ESP in relation to that theoretically
possible, as well as material pertinent to the discussion of gas velocity distribution
characteristics. Mr. Alan W Ferguson of Black and Veatch provided the material for the section
describing the internal inspection and some discussions of upgrade options and their
applicability.
Mr. Richard L. Roberts of Entergy and Dr. Leonard N. Lapatnick and Mr. Edward G. Waugh of
Public Service Electric and Gas provided valuable reviews and comments as well. Their
contributions are enthusiastically acknowledged.
Mr. Charles (Chuck) Altin is remembered as well. Chuck was part of the original team, but
passed away before the project had progressed very far.

vii

EPRI Licensed Material

CONTENTS

1 INTRODUCTION.................................................................................................................. 1-1
1.1 Purpose of This Guide ................................................................................................. 1-1
1.2 Optimization Approach ................................................................................................ 1-1
1.3 Organization and Overview of Guide ........................................................................... 1-2
1.4 Tips for Using This Guide ............................................................................................ 1-4

2 CURRENT VS. EXPECTED PERFORMANCE COMPARISON........................................... 2-1


2.1 Determine Existing ESP Performance ......................................................................... 2-1
2.1.1 Inlet Loading ........................................................................................................ 2-1
2.1.2 Outlet Emissions .................................................................................................. 2-2
2.2 Determine Expected Optimum ESP Performance ..................................................... 2-2
2.2.1 Plotted Curves Show Best Expected Performance............................................... 2-2
2.2.2 Data Requirements .............................................................................................. 2-4
Inlet Particle Size Distribution.................................................................................... 2-4
Fields and Plate Spacing........................................................................................... 2-6
Specific Collection Area ............................................................................................ 2-6
Current Density ......................................................................................................... 2-6
2.2.3 Using the Performance-Estimating Graphs .......................................................... 2-9
Using the Performance-Estimating Curves: An Example........................................... 2-9
2.3 Determine Whether ESP Is Candidate for Performance Optimization........................ 2-17

3 PERFORMANCE ESTIMATING WITH AN ESP MODEL .................................................... 3-1


3.1 About the ESP Model .................................................................................................. 3-1
3.2 Site-Specific Inputs...................................................................................................... 3-2
3.2.1 Electrical Operating Data ..................................................................................... 3-5
3.2.2 Flue Gas Composition.......................................................................................... 3-6
3.2.3 Ash Loading......................................................................................................... 3-6
3.2.4 Fly Ash Resistivity................................................................................................ 3-6

ix

EPRI Licensed Material

3.2.5 Particle Size Distribution ...................................................................................... 3-7


3.2.6 Gas Flow Distribution ........................................................................................... 3-8
3.2.7 Gas Sneakage and Reentrainment ...................................................................... 3-8
3.3 Determining Best Possible Performance ..................................................................... 3-9
3.4 Model Calibration......................................................................................................... 3-9

4 ON-LINE DIAGNOSTICS..................................................................................................... 4-1


4.1 Retrieve Baseline Data: The ESP Log Book ................................................................ 4-1
4.2 Document Current Operating Data .............................................................................. 4-2
4.3 Assess the Rapping System........................................................................................ 4-3
4.3.1 Investigating Rapping Reentrainment .................................................................. 4-3
4.3.2 Analyzing Reentrainment ..................................................................................... 4-4
4.3.3 Inspecting Rapping Equipment ............................................................................ 4-9
4.4 Obtain Electrical DataV-j Curves and Waveformsvia Gas Load Test .................... 4-9
4.4.1 Diagnostic Value .................................................................................................. 4-9
4.4.2 How to Perform the Gas Load Test .................................................................... 4-10
4.4.3 Plotting the V-j Curves ....................................................................................... 4-13
4.5 Analyze the Electrical Readings ................................................................................ 4-16
4.5.1 Interpreting Meter Readings............................................................................... 4-16
Normal Meter Readings for ESP in Good Condition ................................................ 4-16
Meter Readings for Abnormal Conditions ................................................................ 4-18
4.5.2 Interpreting Secondary V-I Waveforms............................................................... 4-27
4.5.3 Interpreting V-j Curves ....................................................................................... 4-28
V-j Curves at Part Load ........................................................................................... 4-31

5 OFF-LINE DIAGNOSTICS................................................................................................... 5-1


5.1 General Notes on Shutdown and Inspections .............................................................. 5-1
5.1.1 Safety First........................................................................................................... 5-1
5.1.2 Checklist and Map ............................................................................................... 5-2
5.2 Conduct Dirty Air Load Test......................................................................................... 5-6
5.2.1 Procedure for Air Load Test ................................................................................. 5-6
5.2.2 Interpretation of Air Load V-j Curves .................................................................... 5-7
5.3 Conduct Dirty Inspection.............................................................................................. 5-9
5.3.1 Plates, Discharge Electrodes, and Other Main Structures.................................... 5-9
5.3.2 Ash Hoppers ...................................................................................................... 5-10

EPRI Licensed Material

5.3.3 High-Voltage Support Insulators ........................................................................ 5-10


5.3.4 Inlet and Outlet Ducts......................................................................................... 5-10
5.3.5 Discharge Rapping and Rapping Force.............................................................. 5-11
Discharge Rapping.................................................................................................. 5-11
Transmission of Rapping Force............................................................................... 5-11
5.4 Clean the ESP........................................................................................................... 5-12
5.4.1 Dry Blasting ....................................................................................................... 5-12
5.4.2 Water Wash ....................................................................................................... 5-13
5.5 Conduct Clean Air Load Test..................................................................................... 5-13
5.6 Conduct Clean Inspection.......................................................................................... 5-14
5.6.1 Collecting and Discharge Electrodes.................................................................. 5-16
Check Alignment ..................................................................................................... 5-16
Measure Plate Thickness ........................................................................................ 5-16
Inspect Discharge Electrodes.................................................................................. 5-16
5.6.2 Rapper Attachments .......................................................................................... 5-17
5.6.3 Casing and Structural Elements......................................................................... 5-17
5.6.4 Ash Handling System......................................................................................... 5-19
5.6.5 Ductwork............................................................................................................ 5-19
5.7 Gas Flow Distribution ................................................................................................ 5-19
5.7.1 Obtaining Measurements .................................................................................... 5-20
5.7.2 Analyzing Measurements................................................................................... 5-21

6 CORRECTIVE MEASURES FOR COMMON PROBLEMS.................................................. 6-1


6.1 Integrating Repairs With ESP Replacement Cycle....................................................... 6-1
6.2 Leakage ...................................................................................................................... 6-3
6.3 Non-optimum Rapping................................................................................................. 6-4
6.3.1 Rapping System Optimization .............................................................................. 6-4
Collection Plate Rapping ........................................................................................... 6-5
Discharge Electrode Rapping.................................................................................... 6-5
6.3.2 Design Modifications ............................................................................................ 6-6
6.4 Misalignment or Warping of the Collection Plates ........................................................ 6-6
6.4.1. Dead Misalignment ............................................................................................. 6-7
6.4.2. Live Misalignment ............................................................................................... 6-7
6.4.3 Repair and Replacement Options ........................................................................ 6-8
Replacement Sometimes Necessary......................................................................... 6-8

xi

EPRI Licensed Material

Repair More Common ............................................................................................... 6-8


6.5 Bent or Broken Discharge Electrodes .......................................................................... 6-9
6.5.1 Weighted-Wire Designs ....................................................................................... 6-9
6.5.2 Rigid Wire Frame Systems................................................................................... 6-9
6.5.3 Rigid Discharge Electrodes ................................................................................ 6-10
6.6 Suboptimal Power Supply and Controls..................................................................... 6-10
6.6.1 Automatic Voltage Control ................................................................................. 6-12
6.6.2 Current-Limiting Reactor .................................................................................... 6-13
6.7 Inadequate Electrical Sectionalization ....................................................................... 6-13
6.8 Non-optimum Flue Gas Parameters .......................................................................... 6-14
6.8.1 Volumetric Flow Rate ......................................................................................... 6-14
6.8.2 Flue Gas Temperature ....................................................................................... 6-14
6.8.3 Gas Flow Distribution ......................................................................................... 6-15
6.9 Undesirable Fly Ash Properties ................................................................................. 6-18
6.9.1 High Resistivity .................................................................................................. 6-18
Coal Switching ........................................................................................................ 6-18
Flue Gas or Fly Ash Conditioning ............................................................................ 6-19
6.9.2 Excessively Low Resistivity................................................................................ 6-19
6.9.3 Excessive Fine Particles .................................................................................... 6-22

A DATA SETS USED TO GENERATE PERFORMANCE-ESTIMATING CURVES................A-1


B PRECIPITATOR PERFORMANCE EVALUATION USING ESP MODEL ...........................B-1
Step 1. Determine the estimated collection efficiency for Vgas = 0.25 and S = 0.10. ............ B-2
Step 2. Compare the estimated performance to the measured collection efficiency........... B-2
Step 3. If measured collection efficiency is only slightly less than the estimated value,
investigate operating parameters........................................................................................ B-3
Step 4. If measured collection efficiency is significantly lower than estimated
performance, determine cause of poor performance. ......................................................... B-3
Step 4.1. Determine if the operating currents are completely useful in the
precipitation process. ..................................................................................................... B-4
Excessive Sparking................................................................................................... B-4
Back Corona ............................................................................................................. B-4
Electrode Misalignment ............................................................................................. B-5
Step 4.2. Determine if performance is limited by non-uniform inlet gas temperature
or fly ash loading............................................................................................................ B-5

xii

EPRI Licensed Material

Step 4.3. Determine if performance is limited by non-ideal effects. ............................... B-5


Summary ....................................................................................................................... B-7

C ELECTROSTATIC PRECIPITATOR PRINCIPLES .............................................................C-1


C.1 General Process Description.......................................................................................C-1
C.1.1 Single-Stage vs. Two-Stage Designs ..................................................................C-2
C.1.2 Hot-Side vs. Cold-Side Designs ..........................................................................C-4
C.2 Theory of ESP Operation ............................................................................................C-6
C.2.1 Gaseous Conduction and the Corona Discharge.................................................C-6
C.2.2 Current and Voltage Relationships ......................................................................C-7
C.2.3 The Electric Field.................................................................................................C-9
C.3 Particle Charging ......................................................................................................C-11
C.3.1 Field Charging ...................................................................................................C-11
C.3.2 Diffusion Charging.............................................................................................C-13
C.3.3 Combined Field and Diffusion Charging ............................................................C-13
C.3.4 Practical Aspects of Charging............................................................................C-14
C.4 Particle Collection .....................................................................................................C-15
C.4.1 Forces Acting on the Particles ...........................................................................C-16
C.4.2 Particle Collection With Laminar Flow ...............................................................C-16
C.4.3 Particle Collection With Turbulent Flow .............................................................C-17
C.4.4 Factors Modifying Particle Collection.................................................................C-19
C.5 Collected Particle Removal .......................................................................................C-25

D CONDUCTION MECHANISMS IN FLY ASH ......................................................................D-1


D.1 R&D History ................................................................................................................D-1
D.2 Current Theories of Electrical Conduction in Fly Ash ..................................................D-4
D.2.1 Volume Conduction .............................................................................................D-4
D.2.2 Surface Conduction .............................................................................................D-6

E SI AND U.S. UNIT CONVERSION FACTORS .................................................................... E-1

xiii

EPRI Licensed Material

LIST OF FIGURES
Figure 1-1 Flow Chart of Analytical Procedure ........................................................................ 1-3
Figure 2-1 Coarse (Eastern Bituminous) and Fine (Western Subbituminous) Particle
Size Distribution Used in Figures 2-3 Through 2-8 .......................................................... 2-5
Figure 2-2 Average Values of Current Density vs. In-Situ Resistivity....................................... 2-8
Figure 2-3 Coarse Ash: Full-Load Collection Efficiency vs. Average Values of Useful
Current Density for ESP With Three Fields (plate spacing = 10 inches or 25 cm).......... 2-11
Figure 2-4 Coarse Ash: Full-Load Collection Efficiency vs. Average Values of Current
Density for ESP With Five Fields (plate spacing = 10 inches or 25 cm) ......................... 2-12
Figure 2-5 Coarse Ash: Full-Load Collection Efficiency vs. Average Values of Useful
Current Density for ESP With Seven Fields (plate spacing = 10 inches or 25 cm) ......... 2-13
Figure 2-6 Fine Ash: Full-Load Collection Efficiency vs. Average Values of Useful
Current Density for ESP With Three Fields (plate spacing = 10 inches or 25 cm).......... 2-14
Figure 2-7 Fine Ash: Full-Load Collection Efficiency vs. Average Values of Useful
Current Density for ESP with Five Fields (plate spacing = 10 inches or 25 cm) ............. 2-15
Figure 2-8 Fine Ash: Full-Load Collection Efficiency vs. Average Values of Useful
Current Density for ESP With Seven Fields (plate spacing = 10 inches or 25 cm) ......... 2-16
Figure 4-1 Opacity Trace Indicating Baseline Opacity Before and After Rappers Are
Turned Off ....................................................................................................................... 4-4
Figure 4-2 Opacity Trace Indicating Appropriate Levels of Rapping ........................................ 4-5
Figure 4-3 Opacity Trace Indicating Excessive Rapping Puffs ................................................ 4-6
Figure 4-4 Opacity Trace Suggesting Excessive Rapping Forces Causing Rapping
Reentrainment to Raise the Baseline Opacity.................................................................. 4-7
Figure 4-5 Opacity Trace Suggesting Localized Reentrainment From a Particular Region
of the ESP ....................................................................................................................... 4-8
Figure 4-6 Secondary Voltage Waveforms for Normal Resistivity With No Back Corona....... 4-12
Figure 4-7 Normal Gas Load V-j Curves for Healthy Four-Field ESP .................................... 4-14
Figure 4-8 Normal V-j Curves From a Microprocessor Control .............................................. 4-15
Figure 4-9 Example Secondary Meters for a Four-Field ESP ................................................ 4-17
Figure 4-10 Normal Readings on a Microprocessor Power Supply Control ........................... 4-20
Figure 4-11 Microprocessor Ramping After Control Start or Control Regulation.................... 4-21
Figure 4-12 Minor Sparking Under Normal Operation ........................................................... 4-22
Figure 4-13 Microprocessor Responding to Spark................................................................. 4-23
Figure 4-14 Sluggish Response to Multiple Sparks ............................................................... 4-24
Figure 4-15 Suppression and Restart After an Arc ................................................................ 4-25

xv

EPRI Licensed Material

Figure 4-16 Sustained Arc Such as Caused by Broken Wire Shorting................................... 4-26
Figure 4-17 Secondary Voltage Waveform (Voltage vs. Time) at Corona Start ..................... 4-27
Figure 4-18 Secondary Voltage Waveform With No Back Corona......................................... 4-27
Figure 4-19 Secondary Voltage Waveform With Heavy Back Corona ................................... 4-28
Figure 4-20 Typical Gas Load V-j Curves for a Healthy Four-Field ESP................................ 4-29
Figure 4-21 Back Corona and Premature Sparking Due to High-Resistivity Ash (1012 :cm) ................................................................................................................................ 4-32
Figure 4-22 V-j Curve From Microprocessor Control With High Resistivity and Heavy
Back Corona ................................................................................................................. 4-33
Figure 4-23 Example Problem V-j Curves........................................................................... 4-34
Figure 5-1 Diagram of ESP for Inspection Use........................................................................ 5-3
Figure 5-2 Normal Air Load V-j Curves From Healthy ESP ..................................................... 5-8
Figure 5-3 Example of an Inspection Report With Photographs ............................................ 5-15
Figure 5-4 Relationship Between Collecting Efficiency and Gas Velocity Non-uniformity
for Different Resistivities (Gas Sneakage Factor = 10%) ............................................... 5-23
Figure 5-5 Relationship Between Collecting Efficiency and Gas Sneakage Factor (Gas
Velocity Vgas = 25%) ....................................................................................................... 5-24
Figure 6-1 Example Condition Assessment Curve: Plot of Metal Thickness vs. Time for
Collecting Plates (four casing arrangement) .................................................................... 6-2
Figure 6-2 Example Relationship Between Power, Rapping Intensity, and Opacity............... 6-11
Figure 6-3 Fly Ash Resistivity as a Function of Temperature................................................. 6-16
Figure 6-4 Electric Field as a Function of Position in the Inter-electrode Space .................... 6-21
Figure A-1 Current Density vs. Ash Resistivity for Inlet Section of an ESP............................. A-4
Figure A-2 Current Density vs. Ash Resistivity for Second Section of an ESP........................ A-5
Figure A-3 Current Density vs. Ash Resistivity for Third Section of an ESP ........................... A-6
Figure A-4 Voltage vs. Current Curve Composite From Seventeen ESP Tests Used for
Correlation (All Normalized to 10-inch Plate Spacing) ..................................................... A-7
Figure C-1 Depiction of Single-Stage and Two-Stage Electrostatic Precipitators ...................C-3
Figure C-2 Schematic Example of Wire-in-Pipe Electrostatic Precipitator ..............................C-4
Figure C-3 Example of Fly Ash Resistivity as a Function of Temperature ..............................C-5
Figure C-4 Idealized Secondary Voltage vs. Current Curve ...................................................C-8
Figure C-5 Idealized Electric Field vs. Radial Position, With and Without Current Flow
(Pipe Diameter = 8 inches, Wire Diameter = 0.109 inches)............................................C-10
Figure C-6 Particle Charging Sequence Depicted for Field Charging ...................................C-12
Figure C-7 Comparison of Field Charging Rates for Half-Wave and Pure DC Electrical
Energization ..................................................................................................................C-14
Figure C-8 Illustration of ESP Collection With Laminar Flow ................................................C-17
Figure C-9 Illustration of the Development of a Particle Concentration Gradient for ESP
Collecting a Wide Particle Size Range With Laminar Flow ............................................C-18
Figure C-10 Two Examples of Methods for Providing Wet Collecting Electrodes .................C-27

xvi

EPRI Licensed Material

Figure C-11 Illustrative Relationship Between Collecting Efficiency and Rapping


Intensity for Two Values of Ash Resistivity ....................................................................C-29
Figure D-1 Resistivity vs. Temperature for Ash Used in Electrical Conduction
Mechanisms Study ..........................................................................................................D-2
Figure D-2 Resistivity vs. Time for Experiment With 9 ppm of Sulfur Trioxide Injected
Into Resistivity Cell for a Period of Time and Then Turned Off ........................................D-7

xvii

EPRI Licensed Material

LIST OF TABLES
Table 1-1 Problem Identification Index .................................................................................... 1-5
Table 2-1 Process Variables and ESP Parameters Used to Generate PerformanceEstimating Curves ........................................................................................................... 2-3
Table 3-1 Information Required to Run an ESP Computer Simulation Model (Model
accepts only English, not metric, inputs. A conversion table is included in Appendix
E.) ................................................................................................................................... 3-3
Table 4-1 Example Power Supply Readings for Four-Field ESP (No Resistivity
Limitation)...................................................................................................................... 4-18
Table 5-1 ESP Inspection Area Checklist................................................................................ 5-4
Table 5-2 Suppliers of Gas Flow Measurement Equipment................................................... 5-20
Table A-1 Fly Ash Resistivity and Estimated Useful Current Densities (see Appendix E
for metric conversion factors) .......................................................................................... A-2
Table A-2 Operating Secondary Voltages for Each Electrical Field (kilovolts) (see
Appendix E for metric conversion factors) ....................................................................... A-3
Table C-1 Performance Comparison of an ESP Alone vs. an ESP With a Cyclone (see
Appendix E for SI conversion factors)............................................................................C-21
Table C-2 Example Showing Actual Migration Velocity With Changes in SCA (see
Appendix E for SI conversion factors)............................................................................C-24
Table C-3 Collecting Efficiencies for Selected Particle Sizes as a Function of SCA (see
Appendix E for SI conversion factors)............................................................................C-25
Table E-1 Unit Conversion Factors ....................................................................................... E-1

xix

EPRI Licensed Material

1
INTRODUCTION

1.1 Purpose of This Guide


This guideline, the first of a two-volume set of Guidelines for Upgrading Electrostatic
Precipitator Performance, provides the information necessary to optimize the collection
efficiency of a chronically under-performing electrostatic precipitator (ESP) without conducting
a major overhaul. It is aimed at ESPs requiring only minor to moderate performance
improvements; the solutions presented can generally be implemented for $10$20/kW or less.
The second volume, Electrostatic Precipitator Upgrade Options, addresses situations requiring
more extensive (and expensive) solutionswhether due to tighter emissions limits, coal
switching, or simple equipment aging. This companion volume covers such technologies as flue
gas conditioning, ESP enlargement, retrofit of polishing devices in series such as a compact
hybrid particulate collector (COHPAC) or wet ESP, and conversion to fabric filtration. While
Volume 1 can be used by itself, Volume 2 relies upon operating data and diagnostic information
collected via procedures presented in Volume 1. Please note that neither volume in this set is
intended as a troubleshooting guide for sudden malfunctions; EPRIs computer model ESPert
can assist in such matters.
Seeking to optimize an existing ESP, this volume leads the user through an analysis to determine
(1) whether performance goals can be met without a major upgrade and (2) if so, what steps to
take to improve performance as much as possible. Information is provided in sufficient detail
that a person with average experience in operating and maintaining a power plant ESP will be
able to conduct the study without consulting outside experts.

1.2 Optimization Approach


As outlined in Figure 1-1, the basic approach involves comparing the actual performance of the
ESP in question with the best performance that could reasonably be expected for an ESP of
similar design operating under similar conditions (i.e., same ash resistivity, same particle size
distribution, etc.).
Most readers will find that their measured collection efficiency is suboptimal. If restoring your
unit to its best performance would meet your emission goals, then it makes sense to proceed
through the full optimization study presented in this volume, which includes on-line diagnostics,
off-line diagnostics, and repair.

1-1

EPRI Licensed Material


Introduction

However, if the best performance your unit could hope to attain would still fall short of your
performance goalsfor example, if you expect more restrictive emissions limits or plan to burn
lower-sulfur coal, which produces higher-resistivity fly ashit will be necessary to conduct a
major overhaul or implement new control measures such as flue gas conditioning. In this case,
refer to Volume 2.
If your ESPs current collection efficiency matches or exceeds its best performance, the
optimization steps in this volume are unlikely to make any appreciable improvement. A
significant performance boost would require the technologies discussed in Volume 2.

1.3 Organization and Overview of Guide


Chapter 2 of this volume quickly screens whether optimization will be sufficient to meet
performance goals. The chapter provides approximate estimates of best possible precipitator
performance in the form of computer-projected curves of collection efficiency as a function of
ESP design and operating parameters. (These curves are based on field data from 33 wellperforming ESPs.) If the existing ESP is not performing as well as the curves suggest, then it is
appropriate to conduct a more detailed analysis using an ESP computer model such as EPRIs
ESPM or ESPert.
Chapter 3 describes how to collect input data for a computer model study. The model provides
a more definitive answer as to whether the existing ESP can reach performance goals through
relatively inexpensive optimization or whether more aggressive measures will be required. If
optimization is sufficient, the model can be used to prioritize candidate repairs by calculating the
performance improvement that would result from each repair. If optimization will not be
sufficient, then the model is needed to select and design the best upgrade approach, as described
in Volume 2.
Chapter 4 describes diagnostics to be conducted on-line with the power station operating in
normal dispatch. Such information is obtained from plant opacity meters, observations of the
rapper control system, voltage and current readings, and oscilloscope waveforms and voltagecurrent curves from a gas load test. Example opacity traces, meter displays, waveforms, and V-j
curves are provided to assist in data interpretation.
Chapter 5 covers diagnostic procedures to be conducted during an outage. These procedures
include an air load test and inspection under dirty conditions, to provide insight into abnormal
ash buildup and its effects on unit performanceand, after the internals have been cleaned
another air load test and inspection to check for mechanical damage (misalignment, warpage,
leakage, corrosion, etc.) Chapter 5 also describes the procedure for measuring gas velocity
distribution within the ESP to determine if performance can be improved by modifying the gas
flow characteristics.

1-2

EPRI Licensed Material


Introduction

Collect data on present operating


conditions and performance (Chapter 2)

Compare with optimum performance


expected for current design (Chapters 2, 3)

Current performance less


than optimum

Current performance equal to


or better than optimum, and
meets performance goal

Current performance equal to or


better than optimum, but does not
meet performance goal

Evaluate upgrade options


from Volume 2

Determine reason(s) for


performance shortfall

Evaluation complete

Gather data off-line


(Chapter 5)

Gather data on-line


(Chapter 4)

Rapper testing
(4.3)

Gas load test and


V-j curve plotting
(4.44.5)

Dirty air load test


and inspection
(5.25.3)

Clean air load test


and inspection
(5.55.6)

Gas flow
measurements
(5.7)

Analyze data - identify problems (Chapters 4, 5, and 6)

Poor rapping
(4.3, 5.3.5, 5.6.2, 6.3)

Plate and electrode


damage/misalignment
(4.4.1, 4.5.3, Chap. 5, 6.4, 6.5)

Inadequate controls
(6.6 and Vol. 2)

Leakage
(5.6.35.6.5, 6.2)

Poor sectionalization
(6.7 and Vol. 2)

Gas maldistribution
(5.7, 6.8.3; also Vol. 2)

Poor ash resistivity


(6.8.2, 6.9)

Repair/upgrade and retest

Performance still below optimum

Optimum performance achieved

Optimization complete

Figure 1-1
Flow Chart of Analytical Procedure

1-3

EPRI Licensed Material


Introduction

Chapter 6 presents guidelines for problem correction. Detailed cost estimates are not included,
as the costs for these types of upgrades are highly site specific. Typically, optimization should
cost no more than $10$20 per kilowatt of generating capacity, and in many cases, considerably
less. If needed repairs are so extensive they would cost more than $20/kW, a more complete
ESP upgrade should be considered.
Appendix A provides the data used to develop the performance-estimating graphs in Chapter 2.
Although provided for background purposes, these data can also assist in the selection of input
data for the ESP model if there are data gaps for the ESP under study.
Appendix B discusses the use of the ESP model in an optimization analysis program. The
appendix offers an example of a step-by-step procedure for diagnosing the cause of suboptimal
ESP performance.
Appendix C provides background theory of electrostatic precipitation to ensure that the
optimization program is based on a proper understanding of the physics of ESP operation.
Although a review for most readers, it is nonetheless useful as reference material for the
experienced technologist as well as a training document for those with primarily practical
experience.
Appendix D discusses the electrical conduction mechanisms in fly ash, providing the
background needed to understand the concept behind flue gas conditioning. The material
provided allows the user to evaluate the potential improvement in ash resistivity that the most
common commercial conditioning agent, SO3, could provide.
Appendix E provides a table to assist in converting from English to metric units.

1.4 Tips for Using This Guide


This guide was designed to be used step-wise; however, a certain amount of skipping around can
save time. If the nature of ESP under-performance is uncertain, proceed through this volume
chapter by chapter. However, if the likely cause of substandard performance is known or
strongly suspected, it is appropriate to skip directly to that item. For example, if inadequate
rapping is suspected, go directly to that section in the manual and continue. If the problem is
determined to be low power input, begin with the on-line electrical readings and continue. Table
1-1 shows the location of key topics by chapter section.
If you determine in Chapter 2 or 3 that mere optimization will not bring your unit to the desired
performance level, skip ahead to Volume 2 of this report. Note that Volume 2 will refer you
back to this report to conduct certain diagnostic tests, which will vary depending on the upgrade
technology you select. Note, too, that any technology investment discussed in Volume 2
warrants good, detailed modeling as discussed in Chapter 3 of this volume.

1-4

EPRI Licensed Material


Introduction
Table 1-1
Problem Identification Index
Problem

Discussion

Air In-leakage

5.3, 5.6.3, 6.2

Ash Buildup

Fig. 4-23, 5.3

Close Clearance / Misalignment

4.5.3, 5.2.2, 5.5, 5.6.1, 6.4

Insulators/Feedthrough

5.3.3, 5.6.3

Gas Flow Problems

3.2.6, 5.3.15.3.4, 5.7, 6.8.3

Uneven Gas Temperature

6.8.2

Gas Sneakage

3.2.7, 5.7.2

High Carbon in Ash

5.3.3, 6.9.2

High Resistivity

4.5.3, 5.2, 6.9.1, Appendix D

Very Low Resistivity

6.9.2

Power Supply Controls

4.5.1, 6.6

Rapping Problems

4.3, 5.3.5, 5.6.2,6.3

Reentrainment

3.2.7, 4.3, 6.3, 6.9.2

1-5

EPRI Licensed Material

2
CURRENT VS. EXPECTED PERFORMANCE
COMPARISON

This chapter describes how to determine if your ESP will achieve a sufficient performance gain
from the optimization program presented in this report, or whether it will have to undergo a more
thorough upgrade to reach performance goals. The method is simple: Compare your ESPs
actual performance to the best performance that could reasonably be expected for a wellfunctioning ESP of similar design operating under similar conditions. This comparison gives a
rough estimate of the potential efficiency gain that could be expected from an optimization
program.
This chapter features model-derived curves for estimating optimum performance, discusses the
data inputs required to use the curves, and provides an example of how to use them.

2.1 Determine Existing ESP Performance


To determine the collection efficiency of your ESP, you need values for the inlet mass loading,
which can either be measured or estimated, and the outlet emissions, which must be measured.

2.1.1 Inlet Loading


If no measurements are available, the inlet loading can be estimated from a coal analysis and a
combustion calculation. The calculation is fairly simple to perform now that the flue gas flow
rate is readily available from a continuous emissions monitor (CEM). The coals ash content and
heating value are the primary determining factors for the inlet loading; the boiler type affects
inlet loading as well.
Inlet Loading to ESP = Total Ash Production Rate x Fly Ash Fraction
Flue Gas Flow Rate
For example, a pulverized-coal unit usually produces 65% to 85% of its total coal ash as fly ash,
with the remainder as bottom ash. A cyclone boiler produces the reverse split, with about 35%
fly ash and 65% bottom ash.
A typical pulverized-coal boiler, burning a good grade of bituminous coal, will have an inlet
loading to the ESP of about 2 grains per actual cubic foot (gr/acf) or 4.6 g/m 3 of flue gas at a
temperature around 300OF (150OC); this is for a cold-side ESP. A cyclone boiler will have a
significantly lower inlet loading, on the order of 0.7 gr/acf (1.6 g/m3).
2-1

EPRI Licensed Material


Current vs. Expected Performance Comparison

2.1.2 Outlet Emissions


The outlet emissions value should be a measured quantity from a mass test conducted within the
last 12 months. The preferred test procedure is EPA Method 17 (in-duct filter operating at flue
gas temperature). EPA Method 5 data can be used if Method 17 data are not available, although
Method 5 measurements can potentially overestimate particulate loading by inadvertently
counting droplets from condensables. (Method 5 measurements take place at temperatures lower
than flue gas temperatures.)

2.2 Determine Expected Optimum ESP Performance


2.2.1 Plotted Curves Show Best Expected Performance
The optimum performance that can be expected from your ESP can be determined from the
performance estimation curves at the end of this chapter (see Figures 2-3 through 2-8), which
present predicted collection efficiency as a function of current density. These curves were
generated with an ESP computer modeldeveloped by Southern Research Institute and funded
by EPA and EPRIbased on data collected by EPRI and DOE from 33 operating ESPs. The
curves were originally presented in EPRI report CS-4145, A Manual on the Use of Flue Gas
Conditioning for ESP Performance Enhancement.
Naturally, a more accurate estimate of potential ESP performance can be obtained by running an
ESP model with site-specific data, as discussed in Chapter 3 of this guide. EPRIs ESPM
software works well for this purpose. But even if you plan to run a model, Figures 2-3 through
2-8 provide a useful initial check on how much performance improvement you can hope to attain
from optimizing your existing ESP.
Table 2-1 lists the key process variables and precipitator parameters that were used to generate
the curves in Figures 2-3 through 2-8. The mechanical parameters and gas flow parameters
represent an older cold-side utility ESP that is conservatively designed; newer installations will
have better gas velocity distribution, with a standard deviation lower than 0.25, and a sneakage
factor lower than 0.10.

2-2

EPRI Licensed Material


Current vs. Expected Performance Comparison
Table 2-1
Process Variables and ESP Parameters Used to Generate Performance-Estimating Curves
O

Gas Pressure
Gas Volume Flow

300 F (150 C)

Gas Temperature

1.0 atm
3

200,000 acfm (94 m /s)

Plate Height

33.33 ft (10.16 m)

Plate Length

10.0 ft (3.08 m)

Plate Spacing

10.0 inches (25.4 cm)

Parallel Lanes

30
2

Plate Area per Electrical Section

20,000 ft (1860 m )

Plate SCA per Electrical Section

100 ft /kacfm (19.7 m -s/m )

Wire Spacing

6 inches (15.4 cm)

Wire Diameter

0.1 inch (2.5 mm)


3

2.0 gr/acf (4.6 g/m )

Inlet Particle Mass Loading


Inlet Particle Size Distribution Parameters

Coarse

Log-Normal Mass Mean Diameter

15 m

Log-Normal particle

3.0

Fly Ash Resistivity (-cm)

Fine
20 m
5.0
10

11

12

13

10 , 10 , 10 , 10

Electrical Sections in Direction of Gas Flow

3, 5, 7

Baffled Sections in Direction of Gas Flow

3, 5, 7

Normal Std. Deviation of Gas Velocity Distribution

0.25

Gas Sneakage per Baffled Section

0.10

Reentrainment Correction Factor


Log-Normal Mass Mean Diameter of
Reentrained Mass Loading
particle

6 m
2.5

Although the data points used to create Figures 2-3 through 2-8 will probably not exactly match
those encountered in any particular ESP, they should be close enough in most cases to provide a
useful indication of what a units collection efficiency should be. Development of these
graphs is discussed in Appendix A, along with a summary of input data and plots that allow you
to check how well the data represent your particular precipitator.
2-3

EPRI Licensed Material


Current vs. Expected Performance Comparison

2.2.2 Data Requirements


To cover different ESP sizes and operating conditions, there are six different performanceestimating graphsthree for ashes with a high percentage of coarse particles and three for
ashes with a preponderance of fine particles. The graphs are further differentiated by the
number of collecting fields. Each graph presents multiple curves, each representing a different
SCA. Thus, to know which estimation curve to use for your unit, you will have to collect the
following information:
x

Type of coal being fired (to estimate the inlet particle size distribution)

Number of fields and spacing between plates

Specific collection area (SCA)

Current density

Direct measure from the power supply controls

Estimate based on ash resistivity

These data inputs are discussed in detail below.


Inlet Particle Size Distribution
To select the appropriate estimation graph, you must know whether your inlet particles are
predominantly coarse or fine. This distinction can be approximated from the type of coal being
burned. As shown in Figure 2-1, eastern bituminous coal is generally characterized by coarse fly
ash particles, which are easier to collect in an ESP, while western subbituminous coal has a
greater percentage of fine fly ash particles, which are harder to collect.
Figure 2-1 shows the log-normal distributions that were used for the coarse and fine inlet particle
size distributions used in Figures 2-3 through 2-8. The log-normal parameters reflect the average
size distributions in microns (Pm) for particles with diameters in the range of 0.3 to 30 Pm. The
parameters for the coarse distribution (mmd = 15 Pm, Vparticle = 3.0) represent the average of
fly ash particle size distributions from 17 pulverized-coal utility boilers (without cyclonic precollectors) burning bituminous coals. The parameters for the fine distribution (mmd = 20 Pm,
Vparticle = 5.0) represent the average of particle size distributions of fly ash from 16 pulverized
coal boilers (no cyclonic precollectors) burning subbituminous coals. (Mmd is the mass mean
diameter of the size distribution, i.e., D50, and Vparticle is the geometric standard deviation.)

2-4

EPRI Licensed Material


Current vs. Expected Performance Comparison

Figure 2-1
Coarse (Eastern Bituminous) and Fine (Western Subbituminous) Particle Size
Distribution Used in Figures 2-3 Through 2-8

Note that even though the mmd of the fine distribution is greater than that for the coarse one,
the larger geometric standard deviation is great enough to represent a greater amount of fine
particles. In general, a greater mass concentration of fine particles (that is, a smaller mmd and/or
a larger Vparticle) results in lower values of precipitator collection efficiency, because the smaller
particles are more difficult to collect.
If you have unit-specific data for fly ash particle size distribution, compare it to Figure 2-1 and
select the closest curve. Otherwise, simply choose a coarse or fine designation based on the rank
of coal being fired.

2-5

EPRI Licensed Material


Current vs. Expected Performance Comparison

Fields and Plate Spacing


The estimating graphs present curves for ESPs with 3, 5, and 7 fields. If your unit has 4 or 6
fields, you can interpolate.
The voltages used to derive these curves were normalized for 10-inch (25-cm) plate spacing. If
your plate spacing differs, you need not make an adjustment. The curves are valid for any plate
spacing in the range of 9 to 12 inches (23 to 30 cm), given the level of approximation involved in
the correlations of electrical data summarized in Appendix A. The applied voltage simply varies
in proportion to the plate spacing to give the same average electric field and the same useful
current density for a given fly ash resistivity. Actual plate spacing does, however, affect the
figure you use for SCA, as described below.
Specific Collection Area
Clearly, it is preferable to use your current operating SCA (gas flow can be obtained from the
CEM). But if for some reason the actual SCA is not available, the design SCA will suffice.
Before selecting an SCA curve on the chart, first adjust for plate spacing if your spacing is not at
10 inches (25 cm). For example, a precipitator with 9-inch (23 cm) plate spacing and an SCA of
250 ft2/kacfm (49 m2-s/m3) would have an equivalent SCA of 225 ft2/kacfm (44 m2-s/m3)i.e.,
ratio of 9:10if the plates were spaced at 10 inches within the same casing. Thus, if you were
using Figure 2-3, you would interpolate between the SCA curve for 200 ft2/kacfm and 250
ft2/kacfm.
Current Density
Current density can, of course, be determined based on direct meter readings. (For each section,
simply divide the measured operating secondary current by the sectional plate area to obtain the
current density for that section; subsequently average the current densities for all sections in the
ESP.)
However, for the purpose of predicting your units best possible performance, a direct measure is
not necessarily the most valid approach. If your ESPs condition is off-normal (e.g., there is
plate misalignment) or the unit is operating in back corona, your current readings are very likely
off-normal as well. In particular, actual current readings will be higher than the useful current
values if back corona is present. In this case, use the electrical values corresponding to the
formation of back corona; refer to Figures 4-19 and 4-22 and associated discussion to select
appropriate currents. For an ESP in off-normal condition or with significant back corona, using
the actual readings would likely mis-estimate (typically overestimate) your best possible
performance. (As a quick check on whether your meter readings appear normal, see if the
secondary voltage and current values for a given field fall along the curve for that field in Figure
A-4 of Appendix A. If your values lie far away from the curves, your ESP is operating in offnormal mode).

2-6

EPRI Licensed Material


Current vs. Expected Performance Comparison

Consequently, for the sake of reading Figures 2-3 through 2-8, it is recommended that you
compare your measured current density (determined from directly averaging the meter readings)
to a useful current density value based on your ash resistivity, using Figure 2-2. If your
measured current density differs from the estimate in Figure 2-2, use both the actual and Figure
2-2based current densities (when applying Figures 2-3 through 2-8) to obtain the range of
predicted collection efficiencies. The two resulting efficiency values should bracket the
expected optimum performance for your ESP. Note that if you find a substantial difference
between your measured current density and what Figure 2-2 says it ought to be for your ashs
resistivity, it is a tip-off that your ESP may be suffering from excessive back corona.
To use Figure 2-2, you will obviously need to know the resistivity of your ash. Three methods
are available:
1. In situ measurement. This method yields the most accurate value of ash resistivity, and
thus is preferred.
2. Laboratory measurement. Laboratory analyses can be conducted on either isokinetic or
blended hopper samples. This is the next most accurate means of determining ash resistivity.
3. Estimate from ESP model. If direct data are not available, EPRIs ESPM or ESPert models
can be used to predict resistivity based on data from an ultimate analysis and elemental ash
analysis of your coal (see Chapter 3).

2-7

EPRI Licensed Material


Current vs. Expected Performance Comparison

Figure 2-2
Average Values of Current Density vs. In-Situ Resistivity

Whatever the method for determining resistivity, note that if you have switched coals since
measuring the ESPs collection efficiency, you should use the resistivity that was current at the
time collection efficiency was determined. This will give the most accurate estimate of the
difference between your ESPs actual performance and its optimum performance. If you are
firing a blend, it is best to obtain an in situ measurement or a laboratory analysis of the blend, as
the blends resistivity is not simply the weighted average of the individual resistivities of the
fuels being fired.
2-8

EPRI Licensed Material


Current vs. Expected Performance Comparison

2.2.3 Using the Performance-Estimating Graphs


Having collected the foregoing data (particle size, fields, plate spacing, SCA, and current
density), you are now ready to select the appropriate performance-estimating graph from Figures
2-3 through 2-8. Each graph plots predicted ESP efficiency vs. current density for several
different values of specific collection area. Two sets of graphs were generated, one for each fly
ash particle size distribution (coarse and fine). Each set contains three figures, showing
estimates for 3, 5, and 7 ESP fields. All graphs assume conservative, old ESP values for gas
flow (gas = 0.25) and gas sneakage and reentrainment (S = 0.10).
In each graph, the SCA is fixed at a nominal value of 100 ft2/kacfm (about 20 m2-s/m3) for each
separately energized electrical section in the direction of gas flow (i.e., for each field). Thus, a
three-field ESP would have an SCA of 300 ft2/kacfm (59 m2-s/m3) while a five-field unit would
have an SCA of 500 ft2/kacfm (98 m2-s/m3). On each figure, the curves showing estimated
precipitator performance with SCA values lower than the nominal value were obtained by
mathematically increasing the gas velocity above the nominal value of 4.0 ft/s (1.2 m/s).
Note that with the increasing number of fieldsfrom 3 to 5 to 7there is an obvious step
change in collection efficiency for the same SCA and current density from one figure to the next
(e.g., compare Figures 2-3, 2-4, and 2-5). This discontinuity is principally due to the abrupt
change in the number of baffled sections in the direction of gas flow. (The baffled sections force
the gas that has bypassed an electrified collecting region to be remixed for collection in the next
collecting zone. Thus, the number of baffled sections specified in the ESP model used to
generate these curves has a strong effect on collection efficiency.)
Using the Performance-Estimating Curves: An Example
As an example of how to use the curves, consider a five-field ESP with an SCA of 320 ft2/kacfm
and a plate spacing of 9 inches. The collecting plate area is 22,000 ft2 per field. The expected
fly ash resistivity is 1 x 1011 -cm; the ash is from a bituminous coal with a heating value of
12,000 Btu/lb, ash content of 10%, and sulfur content of 0.8%. The flue gas temperature is
300OF. The ESP efficiency is measured to be 99%, with an outlet loading of 0.02 gr/acf (about
0.06 lb/MBtu). The emission limit is 0.03 lb/MBtu. Can this ESP achieve its limit through
optimization, or will it require a more substantial upgrade? (Note that this example provides
more information than is actually needed to apply the performance-estimating graphs.)

First, find the appropriate chart based on the number of fields and the particle size distribution.
A five-field ESP firing bituminous coal (i.e., coarse particles) will use Figure 2-4.
Next, adjust the SCA to correspond to the 10-inch spacing used in the curves by multiplying by
the ratio of plate spacing (9/10). The equivalent SCA becomes 288 ft2/kacfm at 10-inch
spacingthus, the estimated best performance will be just below the curve for SCA = 300.
Finally, determine current density. Figure 2-2 indicates a current density of about 18 A/ft2.
Compare this to the actual current density measurements. The secondary readings are:
2-9

EPRI Licensed Material


Current vs. Expected Performance Comparison

Field 1 (inlet): 42.7 kV and 320 mA

Field 2: 40.8 kV and 350 mA

Field 3: 37 kV and 400 mA

Field 4: 36.5 kV and 500 mA

Field 5 (outlet): 33.5 kV and 600 mA

These data provide an average current density of about 19.7 A/ft2.


Reading Figure 2-4 using both current densities (18 A/ft2 and 19.7 A/ft2), we find that the
ESP can be optimized to achieve a collecting efficiency of 99.7% to 99.75%. Actual present
measured efficiency is only 99.0%, so there is a lot of room for improvement. But will that
improvement be sufficient to meet the emission limit?
At 99.0% efficiency, the ESP had an outlet loading of 0.06 lb/106 Btu. The emission limit is half
that, at 0.03 lb/106 Btu. Thus, the emitted fraction needs to be cut in half, from 1.0% to 0.5%.
Accordingly, the ESP needs to achieve a collection efficiency of 99.5%, which is less than the
performance level it appears capable of achieving (99.7%).
Because the ESPs measured efficiency is less than expected and its desired performance falls
within expected capabilities, an optimization study is appropriate. Follow the sequence of steps
presented in the remainder of this volume to determine the cause(s) of the substandard collecting
efficiency.
Since the electrical current densities were found to be appropriate for a resistivity of 1x1011 cm, the problem should be something other than high resistivity. However, note that lowering
the ash resistivity (e.g., through flue gas conditioning) after all other problems have been
corrected would provide a further improvement in performance. To estimate this additional
improvement, use Figure 2-4 with the appropriate current density for a resistivity of 1 x1010 cm, the optimum resistivity. From Figure 2-2, the current density would be about 39 A/ft2.
Using this value in Figure 2-4 yields a predicted collection efficiency between 99.8% and 99.9%.

2-10

EPRI Licensed Material


Current vs. Expected Performance Comparison

Figure 2-3
Coarse Ash: Full-Load Collection Efficiency vs. Average Values of Useful Current
Density for ESP With Three Fields (plate spacing = 10 inches or 25 cm)

2-11

EPRI Licensed Material


Current vs. Expected Performance Comparison

Figure 2-4
Coarse Ash: Full-Load Collection Efficiency vs. Average Values of Current Density for
ESP With Five Fields (plate spacing = 10 inches or 25 cm)

2-12

EPRI Licensed Material


Current vs. Expected Performance Comparison

Figure 2-5
Coarse Ash: Full-Load Collection Efficiency vs. Average Values of Useful Current
Density for ESP With Seven Fields (plate spacing = 10 inches or 25 cm)

2-13

EPRI Licensed Material


Current vs. Expected Performance Comparison

Figure 2-6
Fine Ash: Full-Load Collection Efficiency vs. Average Values of Useful Current
Density for ESP With Three Fields (plate spacing = 10 inches or 25 cm)

2-14

EPRI Licensed Material


Current vs. Expected Performance Comparison

Figure 2-7
Fine Ash: Full-Load Collection Efficiency vs. Average Values of Useful Current
Density for ESP with Five Fields (plate spacing = 10 inches or 25 cm)

2-15

EPRI Licensed Material


Current vs. Expected Performance Comparison

Figure 2-8
Fine Ash: Full-Load Collection Efficiency vs. Average Values of Useful Current
Density for ESP With Seven Fields (plate spacing = 10 inches or 25 cm)

2-16

EPRI Licensed Material


Current vs. Expected Performance Comparison

2.3 Determine Whether ESP Is Candidate for Performance Optimization


Having determined your precipitators current collection efficiency from direct measurements,
and its estimated optimum collection efficiency from the appropriate curve in Figures 2-3
through 2-8, you can now make an informed decision as to whether to proceed with the complete
diagnostic, inspection, and repair steps in this volume, or whether you will require a more
extensive upgrade and should therefore simply gather the data necessary for ESP model
calibration and move onto Volume 2 of this guide.
Simply put, if the curves indicate that your unit has the potential to be operating well enough to
meet your desired emissions limit, then continue with this volume step by step. If the chart
shows that even if your ESP were operating optimally, it still would not meet your emissions
target, then you will need to consult the second volume of this guide, Electrostatic Precipitator
Upgrade Options, which discusses candidate technologies and their costs and benefits. First read
Chapter 3 of this volume, as computer modeling will certainly be required for the more extensive
upgrade. Other sections of this volume may also prove useful; Volume 2 will refer you back to
this first volume as necessary for data gathering, unit testing, and inspection.
Be sure to consider whether your precipitator will be operating under the same flue gas
conditions as when collection efficiency was last measured. Changes in inlet particle size
distribution, fly ash electrical resistivity, flue gas temperature, or gas volume flow rate all
influence the collection efficiency. Thus, even a precipitator in tip-top condition may not be able
to attain the collection efficiency it once achieved, or was designed to achieve, simply because of
changes in the process variables. In this case, the ESP is no longer properly sized for the actual
operating conditions encountered; again, refer to Volume 2.
If you anticipate switching to a coal with higher-resistivity ash, check the optimum collection
efficiency for both the current and the expected ash resistivities. That is, when applying the
performance-estimating graphs, select current densities corresponding to the two different ash
resistivities: (1) the resistivity of the ash that was being collected at the time of ESP performance
measurement, to determine whether your ESPs components are operating as well as they should
be; and (2) the anticipated resistivity of the new ash, to determine whether your unit will be able
to meet emission limits under the new flue gas conditions.

2-17

EPRI Licensed Material

3
PERFORMANCE ESTIMATING WITH AN ESP MODEL

If the charts in Chapter 2 indicate that unit optimization (rather than a more extensive upgrade)
will successfully meet your ESP performance goal, it is possible to skip straight to Chapter 4.
However, it is recommended to use a computer model to refine your estimates of best-attainable
performance. Modeling is especially valuable if the estimated best performance from the
charts in Chapter 2 is near your ESP performance goal, or if your performance goal is
particularly demanding. The model provides a more precise prediction of the potential
improvement, so you can be sure that sufficient improvement is possible before undertaking
optimization procedures described in the remainder of this volume.
Moreover, once you have completed the diagnostic procedures in Chapters 4 and 5, the model is
a valuable tool to determine which repairs are worth making. And if you determine that the
existing ESP, as designed, will not achieve your target emission limits, then the model is
necessary to evaluate candidate upgrade options.
This chapter discusses the site-specific data inputs required to obtain the most accurate, useful
estimates from an ESP computer model. Even if you do not intend to run the model, scan this
chapter (especially Section 3.2), as it contains valuable information on data collection needed for
the optimization study.

3.1 About the ESP Model


There are several versions of the leading computer model for simulating the performance of an
electrostatic precipitator. This model was initially developed by the Southern Research Institute
(SRI) in the late 1960s; subsequent evolution to its current form was funded, in part, by the U. S.
Environmental Protection Agency and the Electric Power Research Institute.
EPRI offers two ESP modeling packages, ESPM and ESPert, which are both built around the
same core SRI-EPA-EPRI model, but differ from the original model and from each other in data
entry routine. ESPM was designed to be more user friendly than the original model and places
less of a requirement on the user to have a thorough, detailed knowledge of ESP theory and
practice. ESPM uses approximations to replace some of the more detailed calculations included
in the original version. Tests comparing ESPM with the original model show that the data
developed by either model are suitable for analyzing the behavior of any ESP system and for
predicting the effects of many optimization and upgrade options.
ESPert is designed for use as a data monitoring package that works with the power plant data
gathering system. Like ESPM, ESPert can be used to evaluate what if changes in plant/fuel
3-1

EPRI Licensed Material


Performance Estimating With an ESP Model

parameters and ESP operation. In addition, ESPert serves as a useful tool for monitoring ESP
operation and alerting plant personnel to potential problems. However, for conducting the
optimization analysis in this guide, we recommend using ESPM due to its straightforward
simplicity, unless you have experience using one of the other versions of the model.
The core SRI-EPA-EPRI model is based on the Deutsch-Anderson equation (discussed in
Appendix C), which describes the behavior of a particle with a known value of electrical charge
in a given electric field being collected from a gas stream with fully developed turbulent flow,
assuming an idealized uniform flow field. Empirical correction factors have been added to the
model to account for the non-ideal conditions. These correction factors address non-uniform gas
flow, gas sneakage around electrified regions through hoppers and the space above the collecting
electrode assembly, and rapping reentrainment (rapping reentrainment correction factors were
developed independently for hot-side and cold-side units). The correction factors were
developed on the basis of particle size and loading measurements on several operating ESP units.
The model has been validated through many measurements from operating full-scale and pilotscale precipitators.

3.2 Site-Specific Inputs


The model calls for detailed data about the ESP and fly ash being investigated, as well as
information on the boiler design, fuel characteristics, and general plant configuration. Table 3-1
lists the required data inputs. Data entry requires a few hours, but once the model is set up, the
operation of the system to evaluate a variety of conditions is very straightforward.
For operating data, recent measurements are, of course, the best data source. If recent
measurements are not available, the next best source is earlier measurements, followed by
measurements from a nearly identical (sister) unit. As a last resort, if none of these measures are
available, you can use the models default values (ESPM and ESPert each come with default data
sets for all the variables necessary to operate the model). You may also wish to compare your
defaults to the data set suggested in Table 2-1. Obviously, model results are only as accurate and
up-to-date as the source data.

3-2

EPRI Licensed Material


Performance Estimating With an ESP Model

Table 3-1
Information Required to Run an ESP Computer Simulation Model
(Model accepts only English, not metric, inputs. A conversion table is included in Appendix E.)
Boiler and Plant Description
Plant Name
Unit Number
Gross Rating (MW)
Type of Furnace
Heat Rate (Btu/kWh)
Coal Usage (ton/hr)
Coal Heating Value (Btu/lb)
Coal Type and Mine Name
Overall ESP Parameters
Specific Collecting Area
2
(ft /kacfm)
2
Total Collecting Area (ft )
Design Gas Velocity (ft/s)
Number of Electrical Fields
Total Plate Length (ft)
Plate Height (ft)
ESP Width (ft)
Corona Electrode Diameter (in)
Stack Diameter at Top (ft)
Field Number
2
Plate Area (ft )
Field Length (ft)
Wire Diameter (in)
Wire-to-Plate Spacing (in)
Wire-to-Wire Spacing (in)
Secondary Voltage (kV)
Secondary Current (mA)
Corona Start Voltage (kV)
Rated Peak Voltage (kV)

ESP Parameters (By Field)


1
2

Gas Properties at ESP Inlet


O

Gas Temperature ( F)
Gas Volumetric Flow Rate (acfm/s)
% Nitrogen (N2)
% Oxygen (O2)
% Carbon Dioxide (CO2)
% Moisture Content (H2O)
ppm Sulfur Dioxide (SO2)
ppm Sulfur Trioxide (SO3)

3-3

EPRI Licensed Material


Performance Estimating With an ESP Model
Coal Properties
Coal Ultimate Analysis (Moisture Free)
%Carbon (C)
% Hydrogen (H2)
% Oxygen (O2)
% Nitrogen (N2)
% Sulfur (S)
Coal Proximate Analysis (As Received)
% Fixed Carbon (FC)
% Volatile Matter (VM)
% Moisture (H2O)
% Ash Content
Ash Properties
Inlet Ash Concentration (grains/acf)
3
Ash Density (g/cm )
Elect. Resistivity and Temperature
(-cm, F)
Mass Mean Diameter (m)
Geometric Standard Deviation ()
Ash Mineral Analysis
% Lithium Oxide (Li2O)
% Sodium Oxide (Na2O)
% Potassium Oxide (K2O)
% Magnesium Oxide (MgO)
% Calcium Oxide (Lime) (CaO)
% Iron Oxide (Fe2O3)
% Aluminum Oxide (Al2O3)
% Silicon Dioxide (SiO2)
% Titanium Dioxide (TiO2)
% Barium Oxide (BaO)
% Phosphorus Pentoxide (P2O5)
% Sulfur Trioxide (SO3)

3-4

EPRI Licensed Material


Performance Estimating With an ESP Model

3.2.1 Electrical Operating Data


Most ESP power supply controls are equipped with primary voltage and current meters and
secondary current meters. Many also display the secondary voltage, either from voltage dividers
across the secondary circuit or from an estimate of the secondary voltage based on the primary
voltage and the turns ratio of the step-up transformer. Modern power supply controls are
equipped with secondary kilovolt meters, which display a voltage proportional to the ESP
operating voltage from a set of calibrated voltage dividers. This is, of course, the most accurate
method for obtaining the actual operating voltage on the ESP collecting system.
Sometimes the secondary current meter is actually a voltmeter connected across a lowimpedance resistor (on the order of several ohms) in the ground return circuit from the ESP to
the power supply. The impedance of this resistor should be checked periodically to verify that it
meets the specification, assuring that the current is correctly displayed.
For installations without a secondary voltage divider to provide the ESP operating voltage, an
acceptable estimate can be made from the primary readings of voltage and current and the
secondary current. A reasonable estimate is that the secondary kVA (voltage times current) is
approximately 70% of the primary kVA. This estimate applies when the system is operating
with full wave rectification and little or no back corona.
If the ESP is in very poor mechanical condition or operates with substantial back corona
common condition for ESPs with old analog controls collecting a high-resistivity ashthe
observed electrical current reading will be higher than the useful current value (off-normal
operation). In this situation, voltage readings may also be off, typically lower than normal.
These off-normal readings should not be entered into the model if the purpose is to estimate the
best possible performance. (Clearly, if the purpose is to calibrate the model to existing ESP
performance, you should enter the actual secondary voltage and current.) Instead, for estimating
best performance, enter approximate good electrical data either from (1) a current vs. voltage
curve (the useful current density is the highest value before the current-voltage curve starts to
back-slant, indicating the presence of back corona; see Figures 4-22 and 4-23) or from (2)
Appendix A, which provides useful operating current densities and corresponding voltages for
different values of electrical resistivity, based upon statistical analysis of more than a dozen
operating ESPs.
If using Appendix A, obtain a current density value for the first field of your ESP based on your
ash resistivity and the solid curve fit line in Figure A-1. Figures A-2 and A-3 provide current
density values for the second and third fields, respectively. Note that current densities are
presented for 10-inch (25.4 cm) plate spacing. However, the values are reasonable
approximations for 9-inch spacing as well. After obtaining the current density values for each
field, use the appropriate curves in Figure A-4 to get an estimate of the corresponding secondary
voltage for each field.

3-5

EPRI Licensed Material


Performance Estimating With an ESP Model

3.2.2 Flue Gas Composition


Flue gas composition is a necessary part of the data set required to run the ESP model. All gas
constituents are usually determined as a part of a mass emissions test, except for SO3, which is
determined by a measurement method referred to as either the Cheney-Homalya or Controlled
Condensation Method. If a direct measurement cannot be obtained, the SO3 concentration is
estimated to be about 0.4% of the sulfur dioxide content for eastern bituminous coals and from
0.3% to 0.1 % for western subbituminous coals, depending on the CaO (lime) content in the ash.
The SO2 concentration, if not available from a measurement, is usually about 700 parts per
million (by volume) for each percent of sulfur (by weight) in the fuel. Naturally, actual
measurements are far preferable to these rules of thumb.

3.2.3 Ash Loading


The inlet mass loading to an ESP depends on several factors, including basic boiler type (e.g.,
wall fired, tangential, cyclone, etc.), coal characteristics (ash content, heating value, etc.), and
whether the power station is equipped with a multiclone or other type of pre-collector. A recent
mass test conducted with in situ filters provides the most useful input data. EPA Method 17,
with in-duct filters operated over a full duct traverse, is the best choice.
If such measurements are not available, the ash loading can be estimated by a combustion
calculation. This calculation requires knowledge of ratio of fly ash to bottom ash. As previously
noted, a wall-fired pulverized-coal boiler will have 6585% of the ash appearing as fly ash at the
inlet to the ESP, with 1535% as bottom ash. By contrast, a cyclone boiler produces about 65%
of the coal ash as bottom ash, with only 35% appearing as fly ash. The actual split between fly
ash and bottom ash depends upon the flue gas velocity (e.g., full-load vs. part-load operation),
and the ash collecting characteristics of the furnace (e.g., a cyclone boiler acts as a cyclonic
collector to remove a greater portion of the larger particles).
Overall, the size of the ESP required to reach a given outlet loading and opacity does not depend
much upon the type of boiler, as all boilers produce about the same amount of fine particles (10
microns and smaller). These small particles determine the SCA of the ESP needed to meet a
given emission limit.

3.2.4 Fly Ash Resistivity


Electrical resistivity is probably the most important fly ash property affecting ESP performance.
As discussed in Appendix D, the resistivity of the ash to be collected in a hot-side ESP is
determined by the operating temperature of the ESP and the chemical components in the fly ash.
In cold-side units, flue gas composition is also a determining factor in establishing the ash
resistivity. For cases where the flue gas composition is a significant factor, in situ measurement,
using a point-to-plane resistivity device, is the most accurate method for determining the value of
the ash resistivity. A laboratory measurement is as good as the in situ method for hot-side units
and is the second best method for cold-side units. The preferred laboratory measurement
technique is described in IEEE Standard P-548, which makes use of the ASME PTC 28
3-6

EPRI Licensed Material


Performance Estimating With an ESP Model

measurement cell. If a measured value of resistivity cannot be obtained, estimation methods are
available, including the resistivity prediction method in EPRIs ESPM and ESPert computer
models.

3.2.5 Particle Size Distribution


The best source for particle size distribution data is an in situ measurement using either inertial
impactors or ganged cyclones. With impactors, use devices with a low flow rate for determining
inlet size distributions, because the run times for high-flow-rate devices are so short that accurate
measurements are extremely difficult to make. Ganged cyclones are a good choice for inlet size
measurements because the sample volume can be large enough to allow sufficient run time to
collect representative data, and the sample volume will also be large enough to chemically
analyze the collected ash. High-flow-rate impactors should only be used for outlet size
measurements.
A meaningful particle size distribution can also be determined from laboratory measurements, if
a representative ash sample is available. Acceptable laboratory techniques include analysis with
BAHCO or Shimadzu instruments. Care must be exercised in sampling the ash. The best
sampling method employs filters operated isokinetically in the flue gas stream, as isokinetic
sampling ensures a representative particle sample is collected. If the sampling system is
operated at a velocity lower than the velocity of the gas stream, the larger particles will be overrepresented, as they are inertially driven into the sampling nozzle when the gas stream diverts
around the nozzle. Conversely, if the sampling velocity is greater than that of the gas stream, the
larger particles will be under-represented, as the fine particles tend to follow the gas stream more
readily and the larger fraction will inertially pass the nozzle.
A composite blend of hopper ash is also an acceptable means for obtaining an ash sample. The
accuracy of the result will depend upon actual ESP operation, and the proportionate blend must
approximate the collection efficiency for each field in the ESP. A suggested 100-gram ash
sample for an ESP with four hoppers in the direction of gas flow would comprise 66 grams from
the inlet field, 26 grams from the second field, 6 grams from the third field, and 2 grams from the
outlet hopper.
Both of these ash sampling methods usually provide an acceptable characterization of the inlet
particle size distribution for input to the ESP model. (That is, they will yield the mass mean
diameter value and the geometric standard deviation of the data points.) The critical size range
that must be obtained correctly is from about 0.510 mthe fine particles that are most
difficult for the ESP to collect and therefore establish the outlet loading and opacity
characteristics.
Finally, if no actual data are available, use the models default values. Alternatively, for
pulverized-coal boilers, you may use the approximations shown in Figure 2-1: a mass mean
diameter (mmd) of 15 m and a standard deviation () of 3.0 for eastern bituminous coal, and an
mmd of 20 m and of 5.0 for western subbituminous coal.

3-7

EPRI Licensed Material


Performance Estimating With an ESP Model

3.2.6 Gas Flow Distribution


Another input needed for operating the ESP model is a gas flow distribution correction factor, to
account for areas of high and low gas velocity in the ESP. This parameter was not addressed in
Table 3-1, the data collection checklist, because the gas flow distribution is not usually
measured. As discussed in Chapter 5, gas flow distribution can only be measured during a unit
shutdown with an obstruction-free flue gas path, and the test requires specialized
instrumentation.
Thus, for a first run of the model, it is far more practical to estimate the gas flow distribution.
Ideally, gas velocity would be uniform throughout the precipitator. Because the ESP operates
with a collecting characteristic that is exponential in nature, regions with a higher gas velocity
carry a greater proportion of the particles (more gas flows through these regions) and their
collecting efficiency is lower (SCA is less). Regions with lower velocity (lower flow), while
having a higher collecting efficiency, do not collect enough particles to compensate for those
uncollected in the high-velocity regions. Therefore, from first principles, the more uniform the
gas velocity distribution, the higher the ESP collecting efficiency.
In actuality, gas flow through the ESP is not uniform. Warped plates, ash buildup, and other
imperfections can create areas of higher and lower gas flow. The ESP model accounts for this
uneven flow distribution by using the correction factor gas, which is the geometric standard
deviation of the actual gas velocity distribution measured in the ESP.
For a precipitator more than 15 years old that is still in good mechanical condition, select gas =
0.25 as a reasonable starting value. For a new precipitator where careful design attention was
given to gas flow quality, use gas = 0.10 to 0.15.

3.2.7 Gas Sneakage and Reentrainment


Like gas flow distribution, gas sneakage and reentrainment are not readily measured, but are
required inputs to the ESP model. Again, it is best to start with reasonable estimates and finetune them as needed in the modeling process. The models correction factor S simultaneously
compensates for (1) gas sneakage (the amount of flue gas and ash that bypasses the collecting
zones by flowing through the hopper regions and into the superstructure where the high-voltage
insulators are housed), and (2) both non-rapping reentrainment, and, to some degree, rapping
reentrainment. These factors are truly empirical corrections to bring the models performance
prediction into conformance with the actual or expected performance of an ESP. In particular,
the rapping reentrainment correction factors were developed from measurements conducted on
several full-scale and pilot-scale installations using mass and particle size distribution sampling
equipment, which measured the difference between periods when the rappers were operating and
not operating.
Older ESPs (both hot- and cold-side) built more than 15 years ago usually match S values of 0.10
to 0.12. New ESPs, where much attention has been given to gas flow quality and rapper design,
are represented by S = 0.05 to 0.08.
3-8

EPRI Licensed Material


Performance Estimating With an ESP Model

In addition, the model incorporates a correction for rapping reentrainment. The total reentrained
mass loading is automatically calculated by the model as a function of the mass collected by the
outlet electrical section. This reentrained mass is divided into the particle size bands according
to a log-normal particle size distribution with mmd = 6.0 m and particle = 2.5, as determined by
a curve fit to the field test data. These typical values are based on data from six full-scale and
four pilot studies, and are programmed into the SRI model. If detailed particle size data are
available for your unit, the SRI model can accept specific particle size distributions for the
reentrained material (i.e., you can adjust the mmd and particle); however, it does not allow you to
input a different amount of reentrainment. The SRI model uses a different reentrained fraction
for hot-side and cold-side units determined in a rapping reentrainment study using some pilot and
full-scale installations. ESPM, on the other hand, allows you to input specific values for all of
the reentrainment characteristics.

3.3 Determining Best Possible Performance


After the entering the foregoing data, run the model to determine the best possible performance
of your precipitator design under those process conditions. Note that if your electrical readings
are not within normal bounds, you should not use the actual readings. Rather, enter in normal
current and voltage values in order to determine the best your precipitator can do once it is fully
repaired and optimized.

3.4 Model Calibration


Before moving on to use the model for problem diagnosis or for deciding which repairs or
upgrades to implement, calibrate the model to ensure that the adjustable parameters in the model
appropriately reflect the specific unit under study. Here, it is crucial to use actual electrical
readings, especially if they are off-normal.
To calibrate, simply compare the models predicted values for outlet loading and opacity with
actual recent measurements. If the predicted values do not match the actual measurements,
adjust parameters as needed (see below), starting with gas sneakage and reentrainment (S) and
gas velocity standard deviation (gas). When predicted and measured values agree, the model is
calibrated to match the ESP under study. Candidate optimization options can then be evaluated
with confidence. Costly modifications to the precipitator that will not result in a significant
improvement in performance can be avoided.
Calibration is typically an iterative procedure. The first model run will be based on the ESP
design and operating data, collected as described in Section 3.2. Again, be sure to enter
measured secondary voltage and current readingsusing the average values of voltages for each
field and the sum of the currents for each field to establish the average for your ESP. Use the
default values for gas sneakage and reentrainment. Compare the result of this initial model run
with measured performance values.
If the model projections differ from the measured values by a small amount, adjust the
parameters for S and gas to obtain a near match. The default values for gas and S imply that
precipitators that have non-uniform gas flow distributions (large gas values) usually have
3-9

EPRI Licensed Material


Performance Estimating With an ESP Model

relatively high sneakage values (large S values). Therefore, it is a general practice to increase or
decrease both gas and S values together. Changing the values of these factors by 0.05 will
usually lead to a match after a few iterations. If the model under-predicts performance, both gas
and S should be decreased; if the model over-predicts performance, these input values should be
increased. It should not be necessary to reduce either factor to 0 or increase either factor to a
value of more than 0.5 to achieve calibration. In fact, if a match is not achieved within these
limits, it is likely that one or more model input parameters is incorrect or that there exists an
unusual operating condition within the ESP.
Adjusting S and gas is appropriate if model projections are fairly close to measured values.
However, if the model-predicted performance is significantly greater than the measured
performance, you will need to investigate the cause(s) for the difference. Review the electrical
operating voltages and currents for out of normal values. If some electrical sections are low,
the model may need to be exercised for the different values to evaluate the influence on
collecting efficiency. To make this comparison, model different portions of the ESP
individually, and use a weighted average of the emissions (i.e., loading x flow rate)not of the
collecting efficiencywhen combining the different portions of the ESP.
Ash resistivity, temperature, gas flow distribution and mass loading may also be factors in
suboptimal performance, as may be the particle size distribution. Investigate such potential
limiting factors to identify the changes necessary to provide a match between the actual and
model-projected performance.
If the performance calibration requires input parameters well outside of the norm, it may be
necessary to conduct detailed diagnostic tests as described in this manual. Pay particular
attention to the on-line evaluations to determine that the electrical parameters are appropriate for
the resistivity, and to the off-line evaluations to assure the mechanical integrity of the ESP. By
using the model in combination with the data-gathering exercises presented in this manual, the
existing ESP should be completely characterized.
Note that modeling is most often done to simulate full-load operation because outlet emissions
are usually greatest at full load, and consequently, this condition is of the greatest concern. The
models can be used to simulate operation at lower loads, but sometimes recalibration will be
needed. The model parameter most likely to change at low load is reentrainment. When the
velocity in the precipitator is reduced, losses due to both continuous and rapping reentrainment
are reduced. ESP power levels at low load may also change due to temperature-induced fly ash
resistivity changes, and so power levels appropriate for low-load operation must be entered into
the model.
After the model has been calibrated to the ESP, the parameters descriptive of the existing ESP
can be modified to upgraded values to evaluate the potential for improving the performance of
the ESP. It is very important to have correct values for the ash resistivity, loading, and particle
size distribution for the ash into the ESP. The electrical voltages and currents should be
representative of those for the resistivity and size distribution as well. If these values are used,
the model will provide reliable data for evaluating the potential improvement in the existing
ESP.
3-10

EPRI Licensed Material


Performance Estimating With an ESP Model

Appendix B provides a step-by-step example of using the model to evaluate an existing ESP.
The data-gathering exercises do not exactly follow the sequence described in the body of the
manual, but all of the necessary steps are included. After some experience using the model
together with the data gathering, a user will become proficient in model calibration.

3-11

EPRI Licensed Material

4
ON-LINE DIAGNOSTICS

To diagnose the cause(s) of poor ESP performance, begin by evaluating key metrics while the
unit is on-line. This chapter describes the diagnostic procedures that can be performed while the
ESP is in steady-state operation, typically at or near full load; Chapter 5 covers diagnostics that
require the ESP to be shut down. During steady-state operation, evaluate the following:

Basic operating data, e.g., opacity traces and voltage and current meter readings; significant
changes from previous measurements signal a likely problem area

Rapping performance

Detailed electrical datai.e., secondary voltage vs. current data graphs and oscilloscope
waveforms, obtained from a gas load test

4.1 Retrieve Baseline Data: The ESP Log Book


Finding and understanding the problem(s) underlying poor ESP performance will be much easier
if you have kept faithful records of ESP performance data. Such historical data provide a
valuable baseline of ESP metrics when the unit is operating normally. Any discrepancies
between current performance readings and historical norms signal a need for further
investigation.
Ideally, your company will have kept an ESP log book, beginning with data collected during
ESP commissioning and maintaining continuity throughout the life of the unit. The log book
should include electrical readings from the power supply control cabinets together with
concurrent stack opacity readings, secondary voltage vs. current curves, and secondary voltage
and current waveforms obtained with an oscilloscope.
This electrical information should be collected routinely. (If you have been suspecting a
problem, you may have been recording meter readings every day and collecting V-I curves and
V-I waveforms as often as weekly.) The log book should contain data collected under part-load
as well as full-load operation. (Boiler load affects flue gas temperature, which in turn alters the
electrical resistivity of the fly ash and therefore the electrical operating points and the appearance
of V-I curves.) Ideally, voltage, current, and opacity data will have been collected concurrently
with any inlet or outlet mass emissions tests, providing valuable reference data for the
performance optimization study.
The log book should also include any notes on opacity excursions or load ramp rate limitations
caused by opacity, as well as notes on ash accumulation, hopper plugging, misalignment, and
insulator failure, and any repair or replacement activities. Maintenance of an ESP log book is
4-1

EPRI Licensed Material


On-Line Diagnostics

described in detail in EPRI report CS-5198 Vol. 2, Electrostatic Precipitator Guidelines,


Operation, and Maintenance.
If, in addition to the basic ESP log, you have also kept records of certain auxiliary data, you will
be in a truly excellent position to begin diagnosing the problem(s) with your unit. These
auxiliary data can be collected during routine regulatory compliance tests and include:

Inlet and outlet mass loading

Particle size distributionboth inlet and outlet

Standard boiler, fuel, and fly ash dataload level (MW), heat rate, gas flow rate, gas
temperature, coal composition, fly ash resistivity, mills in service, excess O2, and other data
typically logged by the data acquisition system. Be sure to collect these data concurrent with
the mass loading and particle size tests. This auxiliary information provides guidance about
what may have caused test results to fall outside expected ranges. If there is a glitch in the
test program, these data are very useful to determine what went awry.

Coal proximate and ultimate analyses plus ash mineral analyses. These data can be used to
determine if changes in coal and ash properties are affecting ESP performance.

As the repository of the units electrical, opacity, and general operating data history, the ESP log
book serves as the starting point and baseline reference for the ESP optimization study.
Naturally, all tests and measurements undertaken as part of the optimization study should be
recorded in the log book as well.

4.2 Document Current Operating Data


Begin the diagnostic data collection process by recording standard operating data:

Opacity traces

Secondary voltage

Secondary current

Spark rate

Intermittent energization ratio (if used)

Auxiliary operational data such as boiler load, flue gas spatial temperature distribution at the
ESP inlet, and oxygen and sulfur dioxide levels in the flue gas

Check the key ESP data against historic norms; variations may indicate a problem area. As for
the auxiliary data, most of it will never be needed. However, detailed concurrent data are useful
for determining what happened in the plant when the ESP test data do not match expected
values. Seldom does a test program in an operating power station go off without a hitch. A
coal mill will fail, air heaters stop, clinkers fall into the furnace, unanticipated coal changes
occur, etc. When interpreting what happened in the ESP test, it is invaluable to know whether
the apparent problem with the ESP was actually due to another event elsewhere in the power

4-2

EPRI Licensed Material


On-Line Diagnostics

plant. Thus, be sure to collect these data during the rapping reentrainment test described in
Section 4.3.1 and during or just before beginning the gas load test described in Section 4.4.

4.3 Assess the Rapping System


Particle reentrainment during rapping is a leading factor in ESP outlet emissions, accounting for
about 3040% of the emissions from a high-efficiency cold-side ESP and as much as 50% from
hot-side units (normal percentages will vary from plant to plant). This level of emissions is
expected from well-designed and -operated ESP systems in good condition. If the rapping
system is poorly designed or improperly adjusted, even greater particle emissions from
reentrainment are expected.
For optimum collection efficiency, the frequency and intensity of rapping must be kept in
balance. Insufficient rapping of the collecting plates may result in low short-term opacity, but
can ultimately increase the overall opacity due to large puffs. Conversely, excessive rapping can
result in unacceptably high opacity due to excessive particle reentrainment. Excessive rapping
also wastes energy and can lead to premature mechanical failure of the rappers and the discharge
electrodes.
The good news is that, by and large, rapping performanceat least, performance of the
collecting plate rapperscan be assessed and adjusted while the unit is in service (see Chapter
6), although certain problems, such as improper transmission of rapping force, will require
diagnosis via visual inspection inside the de-energized ESP (see Chapter 5).
The first step, of course, is to determine whether or not there is a rapping problem by evaluating
the contribution of rapping reentrainment to total outlet emissions.

4.3.1 Investigating Rapping Reentrainment


To evaluate the contribution of rapping reentrainment to total outlet emissions, simply compare
outlet opacity with the rappers on and then off. First, observe the outlet opacity meter trace for a
few hours while operating at a steady load. As shown in Figure 4-1, this trace will typically
consist of a baseline trace with rapping spikes superimposed. After a recording a representative
trace, turn off the rappers for the last two fields. Just the outlet fields are adequate for this test,
as the most significant reentrainment typically occurs from these fields. (Particles reentrained in
the inlet fields are usually recollected by the downstream fields, whereas particles reentrained in
the outlet fields can pass directly into the duct and on up the stack.)
Naturally, when the rappers are turned off, the rapping spikes on the opacity trace will disappear,
as depicted in the right-side portion of Figure 4-1. Estimate (by eye) the difference in the
average opacity value with the rappers turned on and off. If the rappers-on value is more than
about 1.4 times the rappers-off value for cold-side units, or more than 1.5 times the rappers-off
value for hot-side units, reentrainment is likely excessive (see Sections 4.3.2 and 4.3.3 to
diagnose the cause).

4-3

EPRI Licensed Material


On-Line Diagnostics

A word of caution: If the rappers are off for a long time, a significant rapping puff could occur
when the rappers are placed back in service, which could create an opacity violation. In
general, try to avoid leaving the rappers off for more than a few hours.

4.3.2 Analyzing Reentrainment


Figure 4-2 illustrates a healthy rapping trace with moderate average opacity and occasional
rapping spikes low enough to avoid risk of an opacity violation. The ratio of base and peak
opacities, or opacity aspect ratio, correlates rapping, instantaneous opacity, and six-minute
opacity. The values shown in Figure 4-2 are typical for standard precipitator and duct
configurations for cold-side ESPs in good operating condition. Adjustments must be made if the
physical distance between the opacity monitor and the precipitator increases, because of the
potential for gas flow mixing to reduce the peak opacity values.
Figures 4-3, 4-4, and 4-5, in contrast, indicate rapping problems requiring attention.
Figure 4-3 illustrates insufficient rapping: the rapping puffs (spikes in the trace) are excessively
high compared with the base-level opacity. Rapping will need to be done more frequentlyand
possibly more intenselyto avoid the possibility of an opacity violation. Note that if prior
rapping settings are no longer adequate and there has been no change of coal, there is some
underlying problemsuch as inadequate rapper force transmission or a change in gas flow
distributionthat warrants investigation during shutdown.

Figure 4-1
Opacity Trace Indicating Baseline Opacity Before and After Rappers Are Turned Off

4-4

EPRI Licensed Material


On-Line Diagnostics

Figure 4-2
Opacity Trace Indicating Appropriate Levels of Rapping

4-5

EPRI Licensed Material


On-Line Diagnostics

Figure 4-3
Opacity Trace Indicating Excessive Rapping Puffs

4-6

EPRI Licensed Material


On-Line Diagnostics

Figure 4-4
Opacity Trace Suggesting Excessive Rapping Forces Causing Rapping Reentrainment to
Raise the Baseline Opacity

4-7

EPRI Licensed Material


On-Line Diagnostics

Figure 4-5
Opacity Trace Suggesting Localized Reentrainment From a Particular Region of the ESP

4-8

EPRI Licensed Material


On-Line Diagnostics

Figure 4-4 depicts the opposite problem, over-rapping. Here, rapping is so frequent and intense
that the rapping puffs are very small, but the overall base level of reentrainment has risen from
about 10% to about 15% (compare to Figure 4-2).
Figure 4-5 shows uneven rapping, with significant reentrainment occurring in a relatively
small portion of the ESP. Such a trace can indicate the need to vary rapping intensity across the
direction of gas flow.
If your trace appears similar to Figure 4-2, and rapping reentrainment is no more than 40% (coldside) or 50% (hot-side), then your rapping system is adequately adjusted. If reentrainment values
are higher than these bounds, you may need to fix the system, starting with visual inspection of
the rapping equipment. If you want an exact measure of the contribution of rapping
reentrainment to outlet emissions, conduct a mass emissions test using two independent sets of
measurement instruments; one set is operated with the rappers on and the other with the rappers
off. However, it is not necessary to measure the rapping reentrainment contribution this
precisely, as the rappers can be optimized without determining their specific contribution.

4.3.3 Inspecting Rapping Equipment


If your opacity traces indicate non-optimal rapping, inspect the general condition of the rapping
system components, including auxiliary equipment such as drive motors and controls. This
inspection should include a functional test, which entails manually sequencing through each
group or row while an observer notes any abnormalities. For solenoid rappers, the weight drop
height should be verified by measurement rather than relying on the sound of impact alone. This
on-line inspection allows you to check all aspects of rapper mechanical operation except for the
proper transmission of rapping force, which requires observing rapping while standing inside a
de-energized, dirty precipitator during shutdown.
Rapper condition should have been inspected periodically, providing a historical record in the
ESP log book against which you can compare results. Obviously broken or defective equipment
should be replaced. More often, however, the equipment is intact, and all that is needed is
adjustment of rapping frequency and intensity; refer to Section 6.3 of this guidebook.

4.4 Obtain Electrical DataV-j Curves and Waveformsvia Gas Load Test
4.4.1 Diagnostic Value
Electrical datanamely, (1) secondary voltage vs. current curves and (2) secondary voltage and
current waveformsare key tools for diagnosing problems with ESP operation. The remainder
of this chapter focuses on collecting and interpreting these electrical data.
Note that this chapter will refer to voltage vs. current density (V-j) rather than straight voltage vs.
current (V-I) curves. V-j curves cover the case where different transformer-rectifier sets feed
different collecting plate areas, by normalizing all electrical sections for comparison purposes. If
4-9

EPRI Licensed Material


On-Line Diagnostics

all your T-R sets feed equal plate areas, you can simply plot straight V-I data without the extra
step of calculating current density.
Voltage vs. current data provide valuable insight into ESP operation. Variations in the shape of
the secondary V-j curves or the secondary waveforms can indicate a variety of problems,
including:

Misalignment of plates or discharge electrodes

Electrical shorts (due, for example, to cracked insulators, carbon tracking on insulators,
broken wires, or overly full hoppers)

Abnormal sparking

Back corona (an undesirable condition that contributes to significantly degraded ESP
performance; caused by high-resistivity ash)

4.4.2 How to Perform the Gas Load Test


Both V-j data and oscilloscope waveforms are obtained by conducting a gas load test with the
ESP in service. This test entails incrementally increasing the secondary voltage from 0 V to its
limitor to the commencement of multiple sparking or back coronafor each electrical section,
one section at a time. At each increment, note the secondary current and corresponding voltage;
progressive data points furnish the V-j curve for each electrical section.
Important: Because a gas load test alters (and momentarily disables) ESP operation section
by section, it is possible to incur an opacity violation. If you are not assured of a sufficient
operating margin, you will need to obtain permission from your local air pollution control
agency before conducting the test. If your regulator refuses, try to obtain approval for a test at
part load. (Note that part-load results will not be as accurate.) Failing that, you will have to
settle for an air load test during a plant shutdownpreferably under both dirty and clean
conditionsas discussed in Chapter 5. If you must resort to a part-load test or none at all, at
least obtain oscilloscope waveforms at full load; these waveforms will indicate if you have a
problem with back corona.
As stated above, a gas load test entails incrementally increasing the secondary voltage until a
limitation occurseither the voltage limit, or the commencement of multiple sparking, or the
commencement of back corona, whichever comes first. Conduct the test one electrical section at
a time, beginning with the outlet fields and working back toward the inlet. This precaution
ensures that ash layer disturbances caused by changing energization do not influence data from
other fields. If the inlet field were tested first, dustcake released from the collecting plates as the
operating voltage is reduced would be recollected on the downstream fields, creating an
unrepresentative dustcake on those plates. This unrepresentative dustcake could skew test results
by causing unrealistic current readings for a given voltage.
Throughout the test, collect the plant operating data described earlier in Section 4.2. As
discussed, this auxiliary information may be able to explain any unusual test results. Stepwise
procedures for conducting the gas load test are as follows.
4-10

EPRI Licensed Material


On-Line Diagnostics

1. Record the following preliminary information for each electrical section:

Total collecting plate surface area

Pertinent mechanical design information, including plate design, discharge (corona)


electrode configuration, and plate-to-plate spacing. Also note the diameter of the
discharge electrode; its shape and size are important because electrodes that have sharp
points or are very slender exhibit a lower corona starting voltage than electrodes with a
larger radius of curvature.

Ambient temperature and barometric pressure

Meter readings for primary and secondary operating voltage and current. Note that if the
power supply controls are an older analog model, only average values of secondary
voltages and currents are available. Some of the newer control sets indicate the peak,
average, and trough (minimum) values of voltage for each value of current. If your
controls provide such information, record it for this preliminary measure and all
subsequent measurements throughout the gas load test.

2. Hook up an oscilloscope to obtain the secondary waveforms. Record the operating


waveforms before starting the gas load test, and then record the waveform at key points
throughout the test. These waveforms confirm the corona start voltage and the onset of
sparking or arcing during the gas load test. They also serve as diagnostic tools by indicating
the presence of back corona.
If you are not sure how to set up the oscilloscope for these readings, consult the power supply
manufacturer. Normally the oscilloscope will be connected to a voltage of around 5 V that is
proportional to the secondary voltage. If you install voltage dividers, remember that the
actual voltage in the ESP, where the connection is made, is on the order of 50,000 V. If you
do not have experience installing voltage dividers, it is important to obtain assistance.
3. Record the corona start voltage, i.e., the secondary voltage at which the secondary current
meter just moves off of zero, indicating the beginning of corona current flow. On an
oscilloscope, the secondary voltage will appear as essentially pure DC until corona start,
because the discharge and collecting electrode system provide sufficient distributed
capacitance to filter and smooth the secondary voltage to a constant value until corona
current begins to flow. The general shape of the secondary voltage as a function of time, at
corona start, is shown as Curve 1 in Figure 4-6.

4-11

EPRI Licensed Material


On-Line Diagnostics

Curve 1. Secondary Voltage vs. Time at Corona Start

Curve 2. Secondary Voltage vs. Time for Normal Operation


Figure 4-6
Secondary Voltage Waveforms for Normal Resistivity With No Back Corona

4. Incrementally increase the secondary voltage through its range, and record the
corresponding current at each interval. In addition, record oscilloscope traces at key points
of the test: near corona start, midway through the voltage range, and at the voltage limit (end
of the test), or at whatever point sparking or back corona is detected.
Increasing the secondary voltage can be accomplished by either (1) increasing the conduction
angle of the silicon-controlled rectifier (SCR) in the power supply control, (2) increasing the
secondary current limits, or (3) using the control microcomputer to run the power supply
through its range.
Beginning at the corona start voltage, measure the corresponding secondary voltages and
currents at regular intervals of secondary current (i.e., adjust the applied voltage until you get
a secondary current reading at a pre-designated current interval, then measure the
corresponding secondary voltage). As is evident from the sample V-j curves later in this
4-12

EPRI Licensed Material


On-Line Diagnostics

chapter (e.g., Figure 4-7), the plot of secondary voltage vs. current density changes curvature
quite rapidly near corona start. Consequently, it is best to take frequent measurements at
lower values, to ensure enough data points to define the critical beginning part of the curve.
Thereafter, it is sufficient to use intervals corresponding to current densities of 25
microamperes per square foot of collecting surface, or 25 nanoamperes per square
centimeter of collecting surface.
Note that each time you proceed to a new test point, it takes a minute or so for the ESP to
stabilize. Wait for this steady state before recording the V-I data point. As stated earlier, if
your controls provide peak and trough electrical readings, record them as well; the trough
values provide the earliest indication of the onset of back corona.
As the secondary voltage is raised above corona start, its oscilloscope waveform changes in
appearance to that indicated as Curve 2 in Figure 4-6. The rising portion of the waveform
shows that the secondary voltage is increasing with the applied voltage until the ESP
secondary voltage reaches the peak value of the applied voltage waveform. During this time,
current is flowing from the power supply to charge the distributed capacitance of the ESP
system and to supply the corona current that is flowing. The letter A on Curve 2 indicates
this region.
As the input voltage from the power supply begins to fall below the voltage across the ESP
electrode system, the back-biased diode stack in the transformer secondary effectively
disconnects the power supply from the ESP electrodes until the next half cycle of
energization. This point is indicated as point B on Curve 2. During the remainder of this
half cycle of input from the power supply, the secondary voltage on the ESP decays as the
corona current flow discharges the energy stored in the distributed capacitance of the
electrode system. This decay in secondary voltage continues until the next half cycle of
energization arrives from the power supply. The decay in voltage from the stored energy is
indicated as segment C on Curve 2.
5. Continue increasing voltage until you reach a limitation. This limitation may be the
power supply voltage or current limit, electrical sparkover, or back corona. If there is no
problem with high-resistivity ash, the voltage is usually limited by sparking.
Sparkover can be detected by the secondary meters and the oscilloscope waveforms (see
Figure 4-14). Back corona can also be detected by meter readings (the current will continue
to increase while the voltage will fail to increase commensurately, or even decrease), as well
as oscilloscope waveforms (see Figure 4-19).

4.4.3 Plotting the V-j Curves


Plot the data for all electrical sections on linear graph paper as depicted in Figure 4-7. If all your
ESPs sections have the same plate area, simple V-I plots will suffice. If your sections vary in
plate area, it is necessary to normalize all ESP sections by dividing the secondary current
readings for each electrical section by the collecting plate area for that section (thereby creating a
V-j instead of a V-I plot).
4-13

EPRI Licensed Material


On-Line Diagnostics

Some digital power supplies automatically plot V-I results. However, many of the new controls
jump the gun by logging the V-I values too soon (e.g., before the current has a chance to
stabilize in response to each incremental voltage increase). Some systems allow you to specify
the wait interval before taking a reading after a change in voltage or current. If you can set the
wait for 12 minutes, then your automatic V-I plot should be reliable. If not, log the stabilized
data points by hand and create your own graph.
The types of V-j curve that can be generated vary depending on the type and age of the power
supply controls. Figure 4-7 shows V-j curves for a precipitator with old analog controls, which
give only average values of secondary voltages and currents. Figure 4-8 shows a normal V-j
curve from a newer digital control set, which can record peak and trough values as well. (Both
Figures 4-7 and 4-8 correspond to a normally operating ESP without any limitation imposed by
high-resistivity ash.)

Figure 4-7
Normal Gas Load V-j Curves for Healthy Four-Field ESP

4-14

EPRI Licensed Material


On-Line Diagnostics

Figure 4-8
Normal V-j Curves From a Microprocessor Control

4-15

EPRI Licensed Material


On-Line Diagnostics

4.5 Analyze the Electrical Readings


This section provides examples to assist in analyzing ESP electrical readings. The discussion
first addresses meter readings, then oscilloscope waveforms, and finally, the V-j curves obtained
through gas load testing.

4.5.1 Interpreting Meter Readings


The meters on the power supply controls provide a wealth of information about the ESPs
operating conditions. Power supply controls operating conventionally act to maintain the
average voltage on the ESP electrode system as high as possible. The controller raises the
operating voltage from the high-voltage rectifier until a sparkover occurs between ground and
the high-voltage section. The secondary voltage is then reduced by a small amount (by reducing
the conduction angle on the control SCR), then increased again until the spark is repeated. This
process repeats to maintain the average voltage near to sparkover. The voltage ramp rate is
controllable in most modern controls.
It is important to determine whether the operating currents are completely useful in charging and
collecting particles. If excessive sparking or back corona is occurring, then the measured
currents will not be complete useful and may, in fact, be detrimental to ESP performance.
Normal Meter Readings for ESP in Good Condition
Figure 4-9 suggests typical meter readings for a healthy four-field ESP with 9- or 10-inch plate
spacing. Such an ESP, collecting pulverized-coal fly ash with a resistivity of 5 x 1010 -cm to 1
x 1011 -cm, will have inlet sections operating at voltages of about 4244 kilovolts and current
densities of about 15 microamperes per square foot (16 nanoamperes per square centimeter).
The second field would likely operate at 3842 kV with a current density of 25 A/ft2 (27
nA/cm2). The third and fourth fields would operate with secondary voltages of 3238 kV with
current densities approaching 4050 A/ft2 (4354 nA/cm2).
Note that the variation in electrical readings from field to field results from the action of the ESP.
The inlet field operates with a higher secondary voltage and lower current because of the greater
amount of electrical space charge from the uncollected particles that have been charged. The
fields further into the ESP operate with progressively higher currents and lower voltages as fewer
charged particles remain uncollected.

4-16

EPRI Licensed Material


On-Line Diagnostics

Figure 4-9
Example Secondary Meters for a Four-Field ESP

4-17

EPRI Licensed Material


On-Line Diagnostics

Table 4-1 provides an example set of electrical data from another ESP in good condition. This
unit is similar to the one whose readings are shown in Figure 4-9: four fields, 10-inch (25-cm)
plate spacing, pulverized bituminous coal, and no ash-resistivity limitation. Note that the values
are quite different from those in Figure 4-9, but the trend in electrical readings from field to field
is the same. This variation in voltage and current readings between these two normal ESPs
operating under similar conditions underscores the importance of keeping a historical log book
so you can compare your units meter readings to their own norm, rather than to some generic
ESP.
Table 4-1
Example Power Supply Readings for Four-Field ESP (No Resistivity Limitation)
Field No.

Primary
Voltage (V)

Primary
Current (A)

Primary
Power (W)

Secondary Secondary Secondary


Voltage (kV) Current (mA) Power (W)

335

55

18,425

47

280

13,160

325

80

26,000

44

420

18,480

315

110

34,650

41

600

24,600

310

130

40,300

39

700

27,300

Meter Readings for Abnormal Conditions


To appreciate meter readings, a correlation can be made between them, oscilloscope waveforms,
and the events occurring in the precipitator itself. Figures 4-10 through 4-16 each illustrate a
different condition in the ESP. Each figure depicts the meter readings, the secondary voltage and
current waveforms as they might appear on an oscilloscope, and a suggested picture of the
activity in the ESP concurrent with these readings.
Figure 4-10 depicts normal operation with a modern microprocessor power supply and control
system cycling through the control function. The control function is set to regulate power to the
section on the basis of spark detection.
Figure 4-11 shows the condition where power is increasing after the equipment is initially
energized or is recovering from a normal discharge spark. The meter direction arrows suggest
that the power is being ramped up to the section.
Figure 4-12 shows a minor increase in secondary current accompanied by a minor decrease in
secondary voltage associated with a normal light spark (referred to as a spit spark). The
decrease in secondary voltage results from the current surge discharging some of the electrical
energy stored in the ESP distributed capacitance. The control system does not normally respond
to this type of discharge.

4-18

EPRI Licensed Material


On-Line Diagnostics

Figure 4-13 represents the response to a somewhat heavier spark. This condition causes a larger
increase in current and reduction in voltage. In this case, the control system responds by setting
the voltage back slightly to prevent the occurrence of a spark re-strike. The rate of sparking
should be measured with spark rate meters. Acceptable levels of sparking are in the range of 10
60 sparks per minute.
Figure 4-14 shows the occurrence of several sparks and the response of the control system. In
this case, the spark has repeated. This multiple spark condition is undesirable, as it tends to drive
the average secondary voltage down, reducing the electrical collecting force and encouraging
particle reentrainment. The reentrainment, in particular, comes from a disruption of the surface
of the collected fly ash layer by the sparks and by an overall reduction in the electrical holding
force over large regions of the layer because of decreased current to those regions.
Figure 4-15 suggests the formation of an arc. A single occurrence such as this usually cannot be
avoided. This undesirable condition may be either the result of unstable electrodes or a process
upset. The control system must reduce the voltage to the ESP to near zero to quench the arc.
Figure 4-16 illustrates a recurring full-conduction arc existing over many half cycles with a
significant reduction in secondary voltage. This depiction suggests a series of unregulated sparks
that are increasing in amplitude. This is a very damaging event that can lead to wire failure. The
broken wire can then cause a short, which is itself manifest by sustained arcing.

4-19

EPRI Licensed Material


On-Line Diagnostics

Figure 4-10
Normal Readings on a Microprocessor Power Supply Control

4-20

EPRI Licensed Material


On-Line Diagnostics

Figure 4-11
Microprocessor Ramping After Control Start or Control Regulation

4-21

EPRI Licensed Material


On-Line Diagnostics

Figure 4-12
Minor Sparking Under Normal Operation

4-22

EPRI Licensed Material


On-Line Diagnostics

Figure 4-13
Microprocessor Responding to Spark

4-23

EPRI Licensed Material


On-Line Diagnostics

Figure 4-14
Sluggish Response to Multiple Sparks

4-24

EPRI Licensed Material


On-Line Diagnostics

Figure 4-15
Suppression and Restart After an Arc

4-25

EPRI Licensed Material


On-Line Diagnostics

Figure 4-16
Sustained Arc Such as Caused by Broken Wire Shorting

4-26

EPRI Licensed Material


On-Line Diagnostics

4.5.2 Interpreting Secondary V-I Waveforms


Figures 4-10 through 4-16 depict secondary voltage and current waveforms under various normal
and abnormal operating conditions. Figures 4-17 through 4-19 show how to recognize a
common problemback corona.

Figure 4-17
Secondary Voltage Waveform (Voltage vs. Time) at Corona Start

Figure 4-18
Secondary Voltage Waveform With No Back Corona

4-27

EPRI Licensed Material


On-Line Diagnostics

Figure 4-19
Secondary Voltage Waveform With Heavy Back Corona

Caused by high-resistivity fly ash, the presence of back corona significantly degrades ESP
collection efficiency. Back corona generates ions that serve as charge carriers in addition to the
desirable charge carriers from the normal corona process. These additional positive ions
neutralize much of the negative space charge from the discharge electrode, causing a significant
increase in the total current flowing in the inter-electrode space. (The resulting increase in the
current in each field can be detected via a characteristic change in the shape of the V-j curves, as
discussed in Section 4.5.3.) Some of the positive ions formed in back corona flow to the fly ash
particles and neutralize a portion of the negative charge on them, thereby decreasing their
probability of being collected.
Figures 4-17, 4-18, and 4-19 give oscilloscope waveforms of secondary voltage as a function of
time for corona start, normal, and back corona conditions. Corona start voltage is indicated on
the operating waveforms for reference. Note that under conditions of heavy back corona, the
rectified voltage waveform dips below the corona-start voltage value.

4.5.3 Interpreting V-j Curves


A baseline set of V-j curves, taken under air load conditions, will have been obtained during ESP
commissioning (the vendor uses these data to help evaluate plate alignment during erection of
the ESP). The ESP log book should also contain more recent air load and gas load V-j curves.
These historical records provide a useful baseline against which to compare the V-j data taken
during the present gas load test. When it comes to actual values of voltage and current, it is far
more instructive to compare your data to earlier data from your unit rather than to the generic
graphs. The sample graphs in this guide are included to illustrate characteristic normal and
abnormal curve shapes.
As discussed below, the general shape of a V-j curve is influenced by several factors, including:

Electrical space charge (responsible for curve variations between fields; see Figure 4-20)

Ash resistivity (high resistivity causes back corona, which has a characteristic back slant
curve shape; see Figures 4-21 and 4-22)

Thickness of ash layer (see Figure 4-23)

4-28

EPRI Licensed Material


On-Line Diagnostics

Figure 4-20 illustrates a V-j plot for a healthy ESP with 10-inch (25.4 cm) plate spacing,
operating at a temperature of 300OF (150OC), with no significant resistivity limitations. (Note
that Figure 4-20 is a repeat of Figure 4-7.)

Figure 4-20
Typical Gas Load V-j Curves for a Healthy Four-Field ESP

4-29

EPRI Licensed Material


On-Line Diagnostics

In Figure 4-20, observe that the curve for the inlet field is to the right of the second field (i.e.,
exhibits a lower current density for a given secondary voltage), and that this pattern continues
from field to field in the ESP. This variation in the V-j curves is caused by the change in
electrical space charge among the different fields. Approximately 60% of the total fly ash
particles are collected in the first field of the ESP. Thus, the particle loadingand
corresponding electrical space charge from the particlesin the second field is about 40% of that
in the inlet section, and so on down the line, until there are relatively few charged particles
remaining in the outlet field. It is this successive reduction in space charge that causes the shift
in the V-j curves from field to field. Removing the space charge allows more current to flow for
a given applied voltage, resulting in a systematic decrease in voltage coupled with a systematic
increase in current density from the inlet to the outlet field.
The amount of space charge influences the amount of shift to higher currents at a given voltage
as you progress from inlet to outlet. If the fly ash contains an unusually large amount of very
fine material, which is harder to collect, the space charge will be higher and the resultant V-j
curve will shift to a higher voltage (shift to the right) for a given current density. This sometimes
occurs with combined ammonia and sulfuric acid conditioning or ammonia injection into a highsulfur gas stream.
To analyze the V-j data, compare the V-j curves within a field and between fields. Ideally, the
curves for each section within a field should be almost identical (data points within a few
percent) unless different sections are equipped with different corona wire designs or are operated
with different energization methods. In actuality, however, most ESPs have variations in gas
flow or temperature distribution that will cause variations between the V-j curves for sections
within the same field.
Curves for different fields should share the same basic upward curve shape but have different
values of voltage and current due to the difference in electrical space charge in the different
fields.
When comparing curves, look for differences in:

General shape of the V-j curves

Corona start voltage (the beginning point of the curve)

Sparking voltage

Also compare the curves to historical gas load V-j curves, if available, to determine if there has
been a change in their appearance. If the curves for some sections deviate from the majority of
others, or differ significantly from previous V-j data, mechanical damage is likely. For example,
sparking at lower voltages than on previous occasions would suggest the possibility of electrode
misalignment or an insulator problem in that electrical section. Operating at higher current for a
given voltage would suggest an electrical fault or mechanical misalignment. An internal ESP
inspection is warranted to determine the cause(s).

4-30

EPRI Licensed Material


On-Line Diagnostics

Changes in the general shape of the V-j curves provide useful clues as to potential problems in
the unit. For example, if the ESP is exhibiting early sparkover and back corona, the appearance
of the V-j curves should resemble those in Figure 4-21 rather than those in Figure 4-20. The
back slant appearance to the curves as current rises indicates the presence of back corona. The
arrow drawn at the top of the inlet curve indicates the sparkover.
For a microprocessor power supply, the shape of a V-j curve indicating back corona will appear
as depicted in Figure 4-22. In this plot, note that as the SCR firing angle is increasedthereby
increasing the operating voltage above corona startthe peak, average, and trough (minimum)
values of secondary voltage increase until back corona is formed. After back corona is initiated,
further increases in the conduction angle for the SCR cause the peak value to continue to
increase, the average value to increase initially, then decrease, and the trough value to decrease.
When back corona becomes severe, the trough value falls below the corona start voltage,
indicating that the back corona is supplying positive ions as charge carriers even though the
corona from the discharge electrode has been interrupted between energization cycles.
Other deviations from a normal set of curves are illustrated by Figure 4-23. In this figure,
Curves 5 and 6 are normal, whereas curves for the other sections indicate problems, as noted
below:

Ash buildup on discharge electrodes. If the discharge electrode have a large buildup of ash
on their surfaces, they behave as an electrode with a much larger diameter as suggested by
Curve 1.

Overfilled hopper or ash buildup on high-voltage insulators. If there is a significant ash


deposit on the surface of the high-voltage insulators, a relationship such as Curve 2 results.
In this case, current flows through the ash deposit on the insulator before the operating
voltage is high enough to generate a corona at the discharge electrode. The slope of the
linear portion of such a curve is related to the actual ash deposit thickness: the greater the
deposit thickness, the greater the slope. Such a curve shape may also indicate an overfilled
hopper, with current flowing from the discharge electrode through the hopper ash to the
collecting plate.

Back corona. Curves 3 and 4 indicate the presence of back corona as discussed above.

Curves 5 and 6 indicate no problems. Curve 5 represents a curve with either a near-clean
collecting electrode or one with an ash deposit with resistivity less than about 5 x 1010 -cm.
Curve 6 is expected to have a reasonable thickness of deposit with a resistivity on the order of 5
x 1011 -cm. Curve 6s shift to the right relative to Curve 5 represents the voltage drop across
the ash layer.
V-j Curves at Part Load
V-j curves taken at part load show the same characteristic shapes to indicate the same ESP
problems as curves taken at full boiler load. However, a gas load test at part load is not as
instructive as a test run at full load. Mechanical problems such as misalignment can generally be
detected just as well, but ash-related problems cannot. The lower flue gas temperature at part
4-31

EPRI Licensed Material


On-Line Diagnostics

load changes the ash resistivity, which can result in better-than-actual electrical readings for
cold-side units (e.g., back corona could be undetectable) and worse-than-actual readings for hotside units (because temperature reduction in a hot-side unit increases ash resistivity, back corona
is likely to develop in hot-side units at part load).
To detect back corona, evaluate secondary voltage waveforms obtained with an oscilloscope
while the plant is operating at full load. If the secondary voltage increases and the current
density decreases as you reduce load (and temperature), you likely had back corona in a coldside unit at full load. Misalignment can be detected in particular fields either at full or partial
load, if not obscured by back corona.

Figure 4-21
Back Corona and Premature Sparking Due to High-Resistivity Ash (1012 :-cm)

4-32

EPRI Licensed Material


On-Line Diagnostics

Figure 4-22
V-j Curve From Microprocessor Control With High Resistivity and Heavy Back Corona

4-33

EPRI Licensed Material


On-Line Diagnostics

Figure 4-23
Example Problem V-j Curves

4-34

EPRI Licensed Material

5
OFF-LINE DIAGNOSTICS

If findings from the on-line tests covered in Chapter 4 indicate problems with the ESP internals,
the next step in performance optimization is to take the unit off-line for further testing and visual
inspection. Off-line diagnostics consist of the following:

Air load test under dirty conditions, yielding V-j curves to indicate mechanical and
electrical problems. When compared with V-j curves obtained through gas load and clean air
load tests, these dirty V-j curves can isolate problems due to ash deposits.

Dirty visual inspection, to diagnose problems with ash buildup, electrical tracking, poor
gas flow or leakage, and inadequate rapping.

Air load test under clean conditions, for V-j curves to evaluate mechanical alignment.

Clean visual inspection, to locate misalignment, corrosion, leaks, and other mechanical
damage.

Gas velocity distribution measurements, to determine whether better gas flow distribution
could significantly improve ESP collection efficiency.

5.1 General Notes on Shutdown and Inspections


IMPORTANT: When shutting down the ESP, be sure to turn off the rapper system before you
turn off power to the transformer-rectifier sets (T-R sets) so that the ash layer will remain on
the collecting plates and discharge electrodes. This is essential for the dirty air load test and
inspection.

5.1.1 Safety First


Extreme caution is required when conducting an internal inspection of an ESP and the associated
ductwork. The first precaution is to make sure that the electrical power supplies are off and
tagged out. Short each individual discharge electrode structure to ground with an appropriate
clamp-type shorting shunt. Install adequate lighting when entering the ESP. Exercise care when
moving through the unit, especially on ladders and walkways. Ash falling from plates and
structures presents a hazard. Wear hard hats, safety glasses, and gloves as appropriate. Make
sure there is always someone in talking distance, either directly or by radio; establish a buddy
system for maintaining contact while inside. If you are new at inspections, work with an
experienced partner.

5-1

EPRI Licensed Material


Off-Line Diagnostics

5.1.2 Checklist and Map


Dont rely on memorybring a checklist, which can also be used to record observations. One
such list is included as Table 5-1. Another aid to the inspection is a map or cell diagram of the
internals, such as that illustrated in Figure 5-1. (Be sure the map is accurate and includes all
areas to be inspected.) This map can be used to mark areas of poor electrical clearance,
corrosion damage, or any other items that will require repair or replacement. Such a preprepared map allows the inspector to quickly and accurately note problem areas, rather than
spend inspection time developing a sketch. Moreover, having problem areas marked on a
graphical layout of the ESP makes it easier to envision and plan the repair work. It also serves as
a permanent record and can make it easier to detect any patterns to problems by examining the
maps from successive internal inspections.

5-2

EPRI Licensed Material


Off-Line Diagnostics

Figure 5-1
Diagram of ESP for Inspection Use

5-3

EPRI Licensed Material


Off-Line Diagnostics
Table 5-1
ESP Inspection Area Checklist
Area for Inspection
Weather Enclosure (Rooftop)
Insulator Compartments
Bus Ducts
Switchgear/Jumpers
Support Insulators
Support Insulator Sweep Vents
Set-off Insulators
Heaters
Penthouse
Bus Conductors (high voltage)
Clearances
Overall Cleanliness (no fly ash)
Transformer/Rectifiers
Casing
Fluid Level/Leaks
Rappers and Vibrators
Mounting
Energy Transmission Shafts
Ground Connections
Interlock System
Keys Accessible
Locking Mechanisms
Inspection Doors
Corrosion
Latch Mechanisms
Key Interlock Switch
Gaskets
ESP Internals
Structural Corrosion
Structural Cracking
Rapper Connections
Wire Support Frames
Plate Support Beams
Missing Bolts / Broken Welds
Plate (top)
Connections
Tears
Wires (top)
Loose or Damaged Connections
Broken/Missing
Rigid Discharge Electrodes

5-4

Inspected

Notes

EPRI Licensed Material


Off-Line Diagnostics
Loose Connection
Bent
Alignment
Plate to Wire Clearance
Wire Frame to Casing (ground)
Clearance
Casing
Cracks
Tears
Warped (deformed but not torn yet)
Lower Walkway Access
Access Doors
Sealing
Corrosion
Door Seal Integrity
Plates
Warped
Alignment Frame
Anti-sway Connection
Wires
Weights for each wire
Alignment Frame
Anti-sway Insulator
Clearance Between Wires and Plates
Clearance Between Wire Frame and Casing
(ground)
Rappers
Excessive Wear
Missing Hammers
Ash Buildup
External
Holes
Closing Connections (dog system on doors)
Pass-Through Boots/Insulators
Hopper Vibrators
Ash Valves
Leakage
Proper Function
Ductwork
Expansion Joints
Holes
Ash Buildup
Corrosion
Flow Distribution Vanes

5-5

EPRI Licensed Material


Off-Line Diagnostics

5.2 Conduct Dirty Air Load Test


Before opening up the ESP for inspection, it is useful to run a dirty air load test. Similar to the
gas load test described in Chapter 4, this test generates V-j curves that provide insight into the
precipitators electrical operation and mechanical integrity, and can indicate problem areas
warranting special attention during inspection.

5.2.1 Procedure for Air Load Test


The procedures for this test are the same as for the gas load test described in Chapter 4, except
that the unit is off-line with ambient air rather than flue gas flowing through the ESP. The test is
conducted with either the natural draft of the chimney or the induced draft fans operating at a
low power level (keep it low to avoid disturbing ash deposits). Air flow is necessary to purge
any ozone generated by the corona process out of the ESP system to minimize its effect on the
electrical conduction characteristics of the air in the inter-electrode space. Air flow is also
necessary to avoid ion formation in the air stream. Such ionization causes sparking during the
power-up cycle at a lower level than normal; this early sparking could give a false indication of
poor electrical clearances or regions of excessive ash buildup that do not actually exist. As in the
gas load test, it is important to begin obtaining data at the outlet electrical fields, working back
toward the inlet. This precaution ensures that ozone generated in the upstream fields does not
influence measurements in the subsequent fields.
Although the procedure for air load and gas load test is the same, the expected end of test
limitation is different. A gas load test typically terminates in sparking or back corona. Air load
curves will seldom terminate in sparking unless misalignment is present; they usually reach the
power supply current limit, typically almost 100 A/ft2 (108 nA/cm2) of collecting surface.
A notable exception can occur with certain very-high-resistivity ashes, such as Powder River
Basin ashes, in very dry climates; in this case, the resistivity is high at ambient temperatures,
causing either sparking or back corona. For such ashes in such climates, a dirty air load test is
inappropriate. If you begin conducting the test and get immediate sparking or back corona, this
is probably the reason. (Severe misalignment could also be the cause, but that would have
shown up in the gas load test as well.) Simply record the onset of back corona, discontinue the
test, and rely on the air load V-j curves to be collected under clean ESP conditions.
In general, however, an ESP inlet field with plate spacing at 9 or 10 inches (23 or 25 cm) should
withstand an applied secondary voltage of 4248 kV before sparking begins. (Voltages for
wider plate spacings should increase about proportionally to the plate spacing increase; thus a
12-inch (30 cm) spacing would reach about 60 kV.)
If sparking commences before the current limit is reached (i.e., at a lower voltage than 4248 kV
for 910 inch (2325 cm) plate spacing), mechanical damage is indicated. Plates may be
misaligned, or there may be protrusions from the collecting electrode, such as a plate clip that
has become displaced, corrosion spots that create rough protrusions from the collecting
electrodes, necked-down regions on the corona wire, or perhaps some welding rod tips that were
left behind after a repair job. When sparking occurs in one or a few electrical sections at
5-6

EPRI Licensed Material


Off-Line Diagnostics

voltages significantly less than those on the other sections, this suggests a problem in those
sections warranting inspection and repair.
Although it would be unusual for all sections to spark before reaching the current limit, there
have been instances of this occurring in hot-side ESPs where the support structure has deformed
because of localized overheating or marginal design of the original structure. If the support
structure is permanently deformed, there is a chance that some or even all of the electrical
sections will have regions of misalignment where the air load V-j curves would exhibit sparking
before reaching the current limit. Early sparking on all sections could also indicate a
malfunction of the power supply controls.

5.2.2 Interpretation of Air Load V-j Curves


In an air load test, there is no ESP activity to alter the space charge between fields, and thus there
should be no systematic shift between curves from one field to the next, as in Figure 4-20.
Rather, V-j curves from all fields should match fairly closely, as in Figure 5-2. Otherwise,
interpretation of air load curves is the same as discussed for gas load curves in Section 4.5.3.
The air load test can help isolate the cause of a malfunction. For example, if the gas load V-j
curves indicate plate misalignment, but the air load V-j curves are normal, that would indicate a
live misalignmenti.e., a warpage or other problem that (so far) is manifest only at operating
temperatures, such as would be caused by a plate

5-7

EPRI Licensed Material


Off-Line Diagnostics

Figure 5-2
Normal Air Load V-j Curves From Healthy ESP

constrained from expansion. Similarly, discrepancies between dirty air load V-j curves and clean
air load V-j curves can indicate the contribution of ash buildup to the problem (e.g., a problem
with electrical tracking on insulators). If you were not able to run a gas load test due to regulator
restrictions, the dirty air load test will be your only source of V-j data regarding ash-induced
problems.

5-8

EPRI Licensed Material


Off-Line Diagnostics

5.3 Conduct Dirty Inspection


The dirty inspection provides, as closely as possible, a view of the ESP internals in their
operating condition. The inspection reveals patterns of ash buildup, which provide insight into
gas flow distribution and rapping adequacy. This initial visual inspection will also reveal much
of the mechanical damage that may exist. Obvious mechanical damage (e.g., broken wires,
severely warped plates) will also be evident at this stage and should be noted. However, clean
conditions will allow a detailed mechanical inspection; the focus of the dirty inspection is ash
buildup.

5.3.1 Plates, Discharge Electrodes, and Other Main Structures


The first step is to inspect the plates, electrodes, struts, vanes, baffles, and gas distribution
devices for ash buildup to ascertain if there are any problems with gas flow or rapping.
Consequently, the inspector must exercise great caution while moving throughout the ESP to
avoid inadvertently disturbing the ash deposits (as well as for safety reasons). The level of ash
buildup on the collecting plates and discharge electrodes will give an idea of the success of the
collecting system in capturing the ash evenly and from the proper place in the ESP internals. For
example, if there is excessive ash on the plate area directly across from the discharge electrode
and none in the plate areas between electrodes, there may be an electrical field problem due to
poor electrode clearances or the particular design of the discharge or collecting electrode.
The ash layer should generally be less than about 1/8 inch (0.3 cm); a layer more than about 1/4
inch (0.6 cm) deep is probably excessive. Areas of high gas flow are indicated by the absence of
ash buildup and possible scouring of the steel. The leading edges of turning vanes and internal
struts for duct supports are the areas most prone to scouring. Ash buildup may be excessive in
areas of low gas flow, most likely to occur in corners and behind turning vanes. Blockage of
flow control devices may be caused by acid or moisture condensation as a result of a process
upset, air in-leakage, or insulation failure.
When an area of high or low gas flow is encountered, judge whether the evidence signifies a
problem that should be reviewed and resolved, or if the distribution is as expected and does not
warrant change. Bear in mind that ash buildup can change the gas flow path in a downstream
area. This is especially true if upstream capture devices such as mechanical collectors or air
heater and economizer hoppers have been taken out of service. When vanes and other
distribution devices become partially or totally plugged by fly ash, the flow pattern changes from
the design distribution, which was modeled under clean conditions, and the new gas flow
distribution may be very undesirable. The dirty inspection should note these areas. (Bear in
mind that a dirty inspection is not a definitive check on gas flow. The gas velocity distribution
can be far outside a good range without being evident in an inspection. Gas velocity distribution
measurements, discussed in Section 5.7, are required to ensure proper gas flow.)

5-9

EPRI Licensed Material


Off-Line Diagnostics

5.3.2 Ash Hoppers


Hoppers should be emptied before the dirty inspection. Leaving the hoppers full of ash
endangers the inspection personnel, who could be burned or otherwise injured, and corrosion
may also occur as the ash cools. Moreover, little can be learned from a full hopper. (The one
exception is that flow patterns in the deposited ash can indicate malfunctioning hopper baffles;
however, this evidence of maldistribution is often obscured by ash falling from the other
internals as the ESP is shut down.)
If, after supposedly emptying, there is still ash in the hopper, clearly there is a problem with the
ash removal system. Note the problem and then manually empty the hopper and continue the
inspection. An earlier EPRI technical report, CS-4880, In-Plant Ash-Handling Reference
Manual (now out of print), provides guidance on repairing ash removal systems.
Unusual buildup in the hopper corners or other trouble spots will still be evident after hopper
evacuation. Ash caking and wind trails on the ash can indicate cracking and air in-leakage.
Pay special attention to hopper valleys and upper flanges, as these are areas of great structural
stress concentration, especially in hot-side precipitators.

5.3.3 High-Voltage Support Insulators


The high-voltage electrical support insulators should be inspected for evidence of electrical
tracking, ash buildup, or cracks. Tracking can cause support insulator failure by concentrating
excessively high temperatures on the surface. Ash buildup on the insulators can allow the
tracking to occur during operation if the dew point temperature is sufficiently high.
There will always be a small amount of dry ash on the surface of the insulators unless they have
been recently cleaned. However, the ash layer should be thin and regular in appearance. If the
ash contains a significant amount of carbon (i.e., it appears blacker than the rest of the ash) or if
the ash is wet from either moisture or acid condensation, the problem must be investigated and
corrected; check for in-leakage and problems with the insulator heaters. Deposits of high-carbon
ash or moist ash will lead to insulator failure and poor ESP performance.

5.3.4 Inlet and Outlet Ducts


Inspect these components to determine the gas flow and ash buildup patterns. Excessive buildup
in a particular area indicates a low flow, and distribution devices should be assessed for
effectiveness. The biggest problem with fly ash buildup in corners is that it provides a medium
for moisture collection and a place for corrosion, which cannot easily be seen in a visual
inspection.

5-10

EPRI Licensed Material


Off-Line Diagnostics

5.3.5 Discharge Rapping and Rapping Force


As discussed in Chapter 4, almost all aspects of the rapping system can be assessed while the
unit is on-line. The exceptions are discharge rapping performance and transmission of rapping
force, which should be checked during the dirty inspection.
Discharge Rapping
Rapping adequacy cannot be easily determined by visual inspection, as the electrodes will often
appear fairly dirty. For rigid electrodes, only the tips need to be clean. Of course, if the tips are
not free of ash, then rapping frequency and/or intensity should be increased. However, do not
rely solely on visual inspection: Determine acceptable cleanliness by evaluating V-j curves taken
under both dirty and clean air load conditions during the shutdown. Optimization of discharge
rapping is discussed in Section 6.3.1.
Transmission of Rapping Force
To determine whether the force is being properly transmitted through the rappers to the plates
and discharge electrodes, simply stand inside the ESP and observe each rapper in operation.
Save this test until the end of the dirty inspection so as not to disturb the ash deposits before they
have been examined.
If rapping causes the whole collecting plate (or discharge electrode) to shake and drop ash,
rapping force is probably being transmitted effectively. If, on the other hand, the rappers seem to
affect only a small area near the rapper attachment, there is likely a problem with force
transmission. Turn up the rapping force (if possible); if rapping is still not effective at the outer
edges of the plate (or discharge electrode), a problem with force transmission is indicated.
If you encounter a problem with force transmission, you may wish to obtain an accelerometer
profile to determine the energy distribution within the system. Mount a number of verylightweight three-axis accelerometers at several points on the electrode system. Record the
acceleration in each plane and each location. If there is a large variation in acceleration from
point to point, determine why and see if a fix is possible. (Usually minimum acceleration is on
structure farthest from point of impact.)
When assessing rapping force, be aware that the rappers are not supposed to shake loose all the
ash. It is normal for a small amount of ash to remain on the plates and wires in an adequately
rapped ESP. It is not unusual for the ash to break off in patches; one rapper operation removes
material primarily from one location while the next rap removes material from another. This
type of removal pattern is normal. The material builds up to an appropriate thickness to break
off, while in another region the layer is not yet thick enough. Inadequate cleaning is best
determined through V-j curves rather than visual inspection (see Figure 4-23, Curve 1).

5-11

EPRI Licensed Material


Off-Line Diagnostics

5.4 Clean the ESP


After the dirty inspection is completed, the ESP internals should be completely cleaned so that an
ash-free inspection can be performed. This cleaning can be accomplished by dry blasting with
sand or grain, or by water washing. After blasting or washing the internals, clean the inlet and
outlet ducts as well, either by vacuuming or shoveling.
Be careful during cleaning because there is a possibility of creating an opacity violation: The
natural draft of the stack will carry some of the ash liberated during the cleaning process up and
out of the chimney. This problem is more likely when removing the ash with a dry process.

5.4.1 Dry Blasting


The blast operation should begin at the top of the ESP and proceed to the lower parts in a fieldby-field manner. Blasting should also include the inside of the high-voltage support insulators.
After cleaning, the ash removal system can be used to empty the ash and blast material out of the
hoppers for disposal either in the landfill or into trucks for transport to a suitable disposal area.
Dry blasting has several advantages over a wet wash:
1. Blasting is not hampered by freezing weather, as is water washing.
2. It does not introduce water into the ESP that could cause the ash to cake and stick.
3. Special handling of the waste is not required.
However, blasting has its disadvantages:
1. The effectiveness of blasting is velocity dependenti.e., cleaning is limited to the maximum
throw distance of the medium. Complete cleaning is often not possible on large precipitators
or on units with limited access.
2. During the cleaning process, operator visibility is extremely limited. It is likely that the
operator would inadvertently dwell in an area, which could cause localized electrode damage.
This problem can be mitigated by using grain, which is softer than sand and unlikely to
damage the electrodes and structural supports. Note that blasting with grain may take longer
due the softness of the blast material.
3. Suspension of ash in the air exiting up the stackand the concomitant potential for an
opacity violationis much more likely with dry blasting than with a wet wash.
4. While effective for restoring performance hindered by ash deposition, blasting may not get
the ESP clean enough for a thorough mechanical inspection. Residual electrical and van der
Waals forces can attract dust from the air after cleaning, requiring the inspector to wipe off
surfaces for proper inspection.

5-12

EPRI Licensed Material


Off-Line Diagnostics

5.4.2 Water Wash


An alternative to a blast cleaning is to wash the ESP internals with water. This either entails the
use of multiple high-volume, high-pressure water hoses or the installation of a permanent wash
header. Use plenty of water to avoid ash caking. Once this effort has started, there should be no
interruptions of the cleaning, especially in the presence of high-calcium ashes. The pozzolanic
activity in high-calcium ashes can cause the formation of cement-like substances that are very
difficult to remove. Use of high water volume should minimize this problem because large
quantities of water will dilute the mixture sufficiently to avoid pozzolanic activity. The source of
water is usually high-pressure ash sluice water (using fire control water would leave the plant
inadequately protected).
The advantages of water washing are as follows:
1. A more thorough cleaning is possible, because reach is not limited to throw distance.
2. It can be done quickly by few personnel. Because reach is not limited to throw distance,
most units can be completely washed from the top elevation.
3. There is less airborne ash, and thus less likelihood of a cleaning-induced opacity violation.
But water washing also has its disadvantages:
1. The ambient temperature must be above freezing.
2. A sluicing system is necessary for drainage and disposal.
Contrary to popularly held beliefs, water washing does not promote corrosion of ESP internals.
If fact, this process can actually prevent wastage in high-sulfur coal situations compared to the
surface corrosion (due to acid condensation) that can occur in an unwashed precipitator left offline for several weeks.

5.5 Conduct Clean Air Load Test


The clean air load test can be performed either before or after the clean visual inspection,
although conducting the test beforehand offers the advantage of alerting the inspection team to
the location of specific problems. Follow the same procedure as for the dirty air load test.
V-j curves obtained from a clean air load test are the most reliable indicators of permanent
(dead) alignment problems. Compare your results to V-j curves from previous air load tests,
starting with the original air load curves from ESP commissioning. Also compare test data to
gas load and dirty air load V-j data, to help isolate the cause of the problem(s).

5-13

EPRI Licensed Material


Off-Line Diagnostics

5.6 Conduct Clean Inspection


Table 5-1 earlier in this chapter provides a checklist of areas to be inspected. Note that this list is
meant as an illustrative starting point for inspection and repair records. If you know any other
potential problem areasfor example, holes that were deliberately drilled for observationadd
them to the list. Figure 5-3 shows a sample inspection report including photographs of ESP
internals.

5-14

EPRI Licensed Material


Off-Line Diagnostics

Figure 5-3
Example of an Inspection Report With Photographs

5-15

EPRI Licensed Material


Off-Line Diagnostics

5.6.1 Collecting and Discharge Electrodes


Fundamental to ESP performance, the collecting plates and discharge electrodes are critical
components to examine during the clean inspection. The air load V-j curves should alert
inspectors to areas likely to present problems.
Check Alignment
Proper alignment and clearance of these components is crucial to ESP performance. Use a ruler
or template to evaluate all clearances between plates and electrodes, plates and frames,
electrodes and the casing, and high-voltage busses and the casing. Alignment within 1/4 inch
(0.6 cm) is desirable; for a unit with 9 or 10 inch (23 or 25 cm) plate spacing, misalignment by
more than 1/2 inch (1.3 cm) warrants correction. Pay special attention to areas exhibiting
unusually high or low ash deposits in the dirty inspection; poor clearances could be the cause.
Most personnel performing precipitator inspections are accustomed to looking for misalignment
across the direction of gas flow. However, it is also important to look for misalignment
occurring in the direction of gas flow, often called longitudinal misalignment. Areas with
improper clearances should be marked directly on the component and on a map of the internals.
Measure Plate Thickness
A precipitators life expectancy is limited by its collecting plates, which degrade over time.
Take this opportunity to measure plate thickness. By plotting plate thickness over periodic
inspections, usually annually, you can predict the units expected service life. Be sure to take
multiple measurements at the bottom elevation of the outlet, as plate wastage usually begins
here. Also pay special attention to any areas that appeared scoured during the dirty inspection;
make sure there is enough material left intact for the component to work properly.
Inspect Discharge Electrodes
In weighted-wire ESPs, it is not uncommon to find wires broken or missing. These electrodes
are prone to breakage, and rather than replace a broken wire, most plant personnel just remove it
during a short outage and return to operation. This stopgap measure is acceptable because the
loss of a few wires (probably up to 10% of the total wires in the ESP) does not impact
performance in a measurable way unless all the missing wires are in the same gas path or area of
the ESP. During this longer outage, however, all such broken or missing wires should be noted
and replaced.
For electrodes with a frame, check the attachment integrity of the electrodes to the frames to
determine if a problem is developing. Often this takes the form of spark erosion, whereby each
spark vaporizes a small portion of the electrode, eventually causing failure. With rigid discharge
electrodes, check the attachment shunt straps installed to avoid sparking at the point of support.

5-16

EPRI Licensed Material


Off-Line Diagnostics

The suspension systems for the discharge electrodes often use stand-off insulators to provide a
stable position in the ESP box between the collecting plates. Carefully check these insulators, as
they can accumulate significant ash buildup that can cause electrical tracking and subsequent
cracking. Damaged insulators can no longer maintain proper positioning of the discharge
electrodes relative to plates and other parts, resulting in poor ESP performance.
The restraining grid is often held in place by an insulator that attaches to the hopper. The
insulator maintains electrical isolation for the discharge electrodes and prevents the entire
discharge electrode assembly from swaying either due to gas flow or electrical forces. This
insulator should be inspected and replaced if damaged or broken.

5.6.2 Rapper Attachments


Examine the rapper system for cracks, tears, rubbing, and leakage, paying special attention to
penetrations of the casing by the rapper shafts, energy transmission shafts, and boots. Also pay
special attention to the point where the rapper shafts connect to the discharge electrode frames or
plate supports; tears in the weld connection can occur when the rapper energy is greater than the
connection was designed to accept. (Sometimes plant personnel misinterpret normal ash buildup
as excessive, or decide there is insufficient energy in the rapper when the real problem is poor
energy transmission; in either case, the subsequent decision to use a bigger hammer can lead to
torn welds or broken attachments.)
Be sure to examine the connections of the energy transmission shafts and the couplings that
transmit the rapper energy from the rapper to the collecting plate frames or discharge wire
suspension system. These connections can become loose, and therefore inefficient at energy
transmission. When there are cracks in the coupling, the rapper energy is not fully transmitted.
If the rapping energy is dissipated in the coupling, degradationof the coupling, and eventually,
of ESP performancemay result. The pieces may not fit properly, further exacerbating the poor
energy transfer. Improper fit may be due to wear, or the use of retrofit parts, or perhaps missing
parts in the energy transmission system. Because of this, there may be an inappropriate
determination to replace with bigger rappers that can cause breakage of the welds and other
connections.

5.6.3 Casing and Structural Elements


Check the casing and internals for leakage, which can take the form of in- or out-leakage
depending on the operating pressure of the ESP. In the case of out-leakage, the cause is usually a
structural failure resulting in the discharge of flue gas and ash into the surrounding area. Such
leakage can cause secondary deterioration of lagging, rappers, and controls. Moreover, outleakage of high concentrations of SO2 poses a safety hazard to plant personnel.
In-leakage can occur with just air or a combination of air and water. Both are detrimental to
precipitator operation and reliability. Air in-leakage, if in sufficient volume, can cause corrosion
that takes years of service from the life of downstream internals in a period of weeks. In-leakage
near the bottom of an ESP poses added potential for problems because air flow into the ESP
5-17

EPRI Licensed Material


Off-Line Diagnostics

casing near the collected ash can reentrain some of that ash, thereby increasing emissions as well
as causing ash caking from moisture condensation. Water (primarily rainwater) in-leakage can
plug hoppers, thus resulting in electrical shorts if the discharge electrodes contact overflowing
ash, plus accelerated corrosion even more severe than that caused by air in-leakage.
Conduct a detailed inspection of all areas prone to leakage. As needed, clean off any remaining
ash residue to get a clear view of component surfaces.
Be sure to include the following:

Access doors. When properly fitted and maintained, the gaskets sealing the access doors in
the ductwork and casing will not allow air leakage. If the doors are not properly sealed, air
in-leakage can cause cold spots, with consequent condensation and eventual corrosion. The
problem self-perpetuates as damage increases with time, causing the holes to get larger and
thereby allowing more leakage. Eventually leakage increases to the point that the gas
distribution in the ESP is affected. The capability of the ID fans to draw the flue gas through
the ESP may also become a problem, resulting in unit de-rating. (Obviously a derate would
occur only after a long time, but there have been door seal leaks which eventually formed
holes about a square foot, or 0.1 m2, in size!) Carefully examine the door seal as you close
up the ESP after inspections or maintenance. Extra care as the doors are closed can minimize
repair costs in the future.

Corners near the inlet mouthpiece and outlet nozzles. These are prone to cracking and
tears.

Vertical columns, horizontal beams, turning vanesi.e., any structural element


completely immersed in flue gas. These are particularly prone to cause punctures in the
adjoining casing. This is because these members expand and contract with temperature
changes at a faster rate than the exterior casing. This relative movement pushes against the
casing wall like a piston until the wall finally yields.

Support stanchions, walkway anchors, conduit struts, and other casing attachments.
These are heat sinks that produce a relatively cold spot on the interior of the casing wall. If
conditions favor acid condensation, these areas will become perforated in time, allowing flue
gas to escape or air and rainwater to leak in.

Slide plates on the stub columns. If the plates do not move as they should (e.g., because of
grit in the sliding area or retainer (guide) bars not functioning properly), the seal welds at the
hopper attachment areas can tear.

Intentional holes or pass-throughs, such as those for opacity meters or sampling ports.
Make sure that seals are in good condition and there is no corrosion of the surrounding
casing.

One source of intentional air in-leakage is through the high-voltage support insulators. Ideally,
each insulator should be equipped with a heated purge/ventilation system to minimize ash and
highly conductive carbon buildup on these insulators, thereby preventing electrical tracking and
subsequent cracking. Such a system features holes in the insulator cap which allow ambient air
to enter, become warmed, and purge the ash deposits. (Note that the heater-only systems
5-18

EPRI Licensed Material


Off-Line Diagnostics

installed on some ESPs provide only minimum protection against condensation and are
ineffective against high concentrations of unburned carbon.)
Carefully check the insulator holes for corrosion or blockage. Cold air entering through the
purge system sometimes results in secondary corrosion of the insulator compartment floor
(where applicable) and the support shafts. Personnel entering the crawl space area (if equipped)
could be injured if the upper frame should collapse as a result of deteriorated support shafts. As a
rule, natural-draft purge systems should be replaced with forced-draft combination heat and
purge systems whenever possible.

5.6.4 Ash Handling System


In addition to checking for cracking, carefully inspect the valves and connections that attach each
hopper to the ash removal system. Poor sealing of these valves allows the transport air (on
pressure systems) to enter an adjacent hopper with a valve that does not seal well, causing an
increase in emissions by entrainment of hopper ash. Poor sealing of the hopper valves also
degrades the performance of the ash removal system. Pay particular attention to these areas
during inspection.

5.6.5 Ductwork
The entire ductwork system should be inspected as well. Examine the corners, seams, and access
doors for tears and corrosion, especially if there was a significant ash buildup in the area. Loose
or damaged turning vanes, expansion joints at inlet and outlet flanges, and connections should be
noted so appropriate repairs can be made.
Also check for holes in the ductwork. In many instances, leaks in ducts and expansion joints can
be determined from an external inspection of the duct while the unit is in service. However,
some holes may be too small to be visible from the outside or may be blocked by insulation and
lagging on the ducts; these holes can usually be detected from the interior of the duct by turning
off all the lights inside and observing the entry of sunlight. It is also helpful to check the ducts
for water leakage if it happens to rain during the outage.

5.7 Gas Flow Distribution


Proper flue gas distribution is essential to good ESP collection efficiency. To evaluate the
potential improvement for your unit, it is necessary to first measure the gas velocity distribution.
Model studies are very useful, especially for new installations, but upgrading an existing ESP
requires actual measurements of the gas velocity distribution in the ESP internals.
The data can be collected any time after the unit is cleanedi.e., before, after, or in between the
clean air load test and the clean visual inspection. Of all diagnostic activities, this will be the
most difficult to schedule. The test requires high air velocities and obstruction-free ducts,
restricting the type of work that can be conducted elsewhere in the gas path.
5-19

EPRI Licensed Material


Off-Line Diagnostics

5.7.1 Obtaining Measurements


Operate the induced draft fans with the boiler open to allow air flow through the ductwork. Set
fan currents at full-load settings to match mass flow (scfm or Nm3/s) as closely as possible.
Although the air velocity through the ESP will be lower than the velocity of flue gas during fullload operation, the measurements will nonetheless provide good data because the mass flow will
be about the same.
Make the measurements within the ESP proper using a thermal anemometer rather than a moving
vane instrument because of the greater sensitivity of the thermal anemometer. A list of suppliers
of gas velocitymeasuring instruments is presented in Table 5-2.
Table 5-2
Suppliers of Gas Flow Measurement Equipment

Company

Address

Telephone and Fax

Alnor Instrument Co.

7555-T N. Linder Ave.


Skokie, IL 60077

Tel. 847-677-3500
Fax 847-677-3539

Barnant Co.

28 W. 092 Commercial Ave.


Dept. TR
Barrington, IL 60010

Tel. 847-381-7050
800-637-3739
Fax 847-381-7053

Davis Instrument Mfg. Co.

4701 Mt. Hope Dr.


Baltimore, MD 21215

Tel. 800-548-9409
Fax 410-358-0252

Electric Speed Indicator Co.

12234 Triskett Rd.


Dept. T-16
Cleveland, OH 44111

Tel. 800-521-2541, Ext. 5


Fax 216-251-2641

KOBOLD Instruments, Inc.

1801 Parkway View Dr.


Pittsburgh, PA 15205

Tel. 800-899-9717
Fax 412-788-4890

Kurz Instruments, Inc.

2411 Garden Rd.


Monterey, CA 93940

Tel. 800-424-7356
Fax 408-646-8901

Omega Engineering, Inc.,


an Omega Tech. Co.

P.O. Box 4047


Stamford, CT 06907

Tel. 800-826-6342
Fax 203-659-7700

Scientific Sales, Inc.

P.O. Box 6725-T


Lawrenceville, NJ 08648

Tel. 800-788-5666
Fax 609-844-0466

Solomat, a Division of
Zellweger Analytics

4330 Thurmond Tanner Rd.


Flowery Branch, GA 30542

Tel 800-765-6628
Fax 770-967-1854

Testo, Inc.

35 Ironia Rd.
Flanders, NJ 07836

Tel. 800-227-0729
Fax 973-252-1729

5-20

EPRI Licensed Material


Off-Line Diagnostics

Gas flow measurements are customarily made either just before or after the inlet field
(downstream is the common choice for large ESPs) and possibly just before or after the outlet
field. Also measure gas flow above the electrified fields (typically midway through the field
length) and through the hoppers (typically mid-hopper).
At each measurement location, divide the ESPs internal cross section into an array of sampling
points and take gas velocity measurements at each of these points. For a typical installation,
measurements should be made in the center of the space between about every fifth set of
collecting plates across the ESP at vertical intervals of about every three to four feet. Test
personnel should stay well away from the measurement instrument location to avoid interfering
with the actual gas flow.

5.7.2 Analyzing Measurements


Compare measurements to gas velocity distribution guidelines published by the Institute of Clean
Air Companies (ICAC, formerly IGCI). The ESP model described in Chapter 3 can help you
evaluate the potential performance boost to be gained from improved gas velocity distribution.
If the measured value of gasi.e., the mathematical standard deviation of the matrix of velocity
measurements, divided by the average valueis greater than 0.25, then changes to make the gas
velocity distribution more uniform should improve ESP performance. If this is the case for your
unit, refer to Section 6.8.3 for discussion of flow correction. (There is some current interest in
establishing a skewed flow in an ESP as an effort to minimize reentrainment; this approach is
discussed in Volume 2.)
Estimates of the degradation of precipitator performance caused by poor inlet gas velocity
distribution are illustrated in Figure 5-4. These curves for collecting efficiency vs. gas were
computed by the ESP model using the input data listed in Table 2-1. The example curves were
computed for SCA values of 250 and 500 ft2/kacfm (49 m2-s/m3 and 98 m2-s/m3) corresponding
to three and six electrical sections in the direction of gas flow. The gas sneakage fraction was set
at 0.10. Each value of gas was set the same throughout each field of the ESP. For older units
with poor inlet ductwork or with partially plugged diffuser plates, it is not unusual to find gas
values of 0.40 or more. To estimate the degree of improvement by correcting gas flow
distribution, select the point on the curve for SCA 250 ft 2/kacfm and fly ash resistivity 1 x 1011
-cm with the collection efficiency of 98.6%. If the gas velocity distribution is improved to gas
= 0.25, the efficiency increases to 99.05%, with the emissions decreased to 68% of the original
value.
The fraction of gas bypassing the electrified fields should be determined from measurements of
the total gas flow and the gas flow above and below the electrified fields during the gas flow
distribution tests. If the sneakage fraction S is greater than 0.10, then baffling to control the gas
sneakage through the non-electrified regions should be installed or upgraded. The effect on ESP
performance by gas and fly ash particles passing above or below the electrified fields is similar
to the effect of non-rapping reentrainment. Both non-ideal effects are mathematically simulated
in the ESP model by an approximate correction to the precipitation rate parameter, using the
specified numerical value of the sneakage fraction S. Estimates of the degradation of
5-21

EPRI Licensed Material


Off-Line Diagnostics

precipitator performance by both kinds of non-ideal effects are illustrated below in Figure 5-5. It
should be noted that as precipitator performance is improved throughout the optimization
program, the relative impact of the various non-ideal effects escalates in importance. Indeed, for
applications where the ash resistivity is optimum (on the order of 1 x 1010 -cm) and all
electrical and mechanical equipment is in good condition, the non-ideal effects become the
limiting factor on ultimate ESP performance.

5-22

EPRI Licensed Material


Off-Line Diagnostics

Figure 5-4
Relationship Between Collecting Efficiency and Gas Velocity Non-uniformity for Different
Resistivities (Gas Sneakage Factor = 10%)

5-23

EPRI Licensed Material


Off-Line Diagnostics

Figure 5-5
Relationship Between Collecting Efficiency and Gas Sneakage Factor (Gas Velocity gas =
25%)

5-24

EPRI Licensed Material

6
CORRECTIVE MEASURES FOR COMMON PROBLEMS

This chapter reviews the following frequently encountered suboptimal conditions, as well as
correction options that can restore ESP performance:

Leakage (in or out, air or water)

Non-optimum rapping

Misaligned or warped collection plates

Bent or broken discharge electrodes

Suboptimal power supply and controls

Inadequate electrical sectionalization

Non-optimum flue gas parameters (flow rate, temperature, flow distribution)

Undesirable fly ash properties (high or very low resistivity, excessive fines)

Some of these problems will have been observed during on-line data gathering; others may have
become obvious during internal inspection. Naturally, repair and replacement decisions should
be made in the context of the remaining time to the next replacement of an ESPs internals. The
ESP models discussed in Chapter 3 can be used to simulate precipitator performance as is to
determine the value of the proposed correction (i.e., the expected increase in collection
efficiency).
Note that some problemssuch as high-resistivity ash or poor gas flow distributioncall for
solutions that may entail a significant capital investment. Such measures are covered more
thoroughly in Volume 2 of this guide. Indeed, optimization should typically cost no more than
$10$20 per kilowatt of generating capacity. If needed repairs are so extensive that costs are
estimated to be much higher, a more significant upgrade should be considered.

6.1 Integrating Repairs With ESP Replacement Cycle


Optimization decisions should be made in context of the remaining time to the next replacement
of an ESPs internals. The collecting plates are both the most expensive and least frequently
replaced component subject to replacement. Other items, such as discharge electrodes,
insulators, T-R sets, and controls, are replaced at shorter intervals and typically do not impact a
run/replace/retire decision. For example, it is not unusual to replace weighted-wire discharge
electrodes as often as four times in the life of a particular set of internals. However, when
collecting plates need to be replaced en masse, it is appropriate to consider the opportunity for
6-1

EPRI Licensed Material


Corrective Measures for Common Problems

coincident design changes that would improve environmental performance. Such improvements
include eliminating sigma-style plates, converting to rigid electrodes, rebuilding with wide-plate
spacing, re-sectionalizing, adding rappers or heated purge systems, etc. Rebuild options are
discussed in Volume 2 of this report.

Figure 6-1
Example Condition Assessment Curve: Plot of Metal Thickness vs. Time
for Collecting Plates (four casing arrangement)

The remaining time until collecting plates need to be replaced can be estimated by sampling the
plate thickness periodically, usually annually, and correlating the trend with the original
thickness and year installed. This will produce a curve that, in time, will approach zero
thickness, as depicted in Figure 6-1. Obviously, thinning plates will create problems before they
wear down to nothing; as a general rule, the minimum thickness for a sigma-type plate is about
17 mils (0.43 mm). Other plates, such as those with fabricated stiffeners, can function as thin as
15 mils (0.38 mm). These minimum levels allow the plate to be self-supporting. Original
performance cannot be expected at these thickness levels, as incidences of warpage, detachment,
and perforation are common. The thickness level at which these problems usually begin to occur
is about 25 mils (0.64 mm).

6-2

EPRI Licensed Material


Corrective Measures for Common Problems

Note that plate wastage does not occur evenly throughout a precipitator. Wastage frequently
begins at the bottom elevation of the outlet. While this is true for both hot- and cold-side
precipitators, the plate replacement cycle for a hot-side precipitator is generally longer than for a
cold-side unit. Of course, there are exceptions. Units that are energized before the flue gas
stabilizes near operating temperature usually experience deterioration beginning at the inlet and
traveling to the outlet. Also, units with unheated natural or forced purge air for the insulators
generally experience deterioration beginning at the top elevations. (The pattern in these
examples is greater wastage in cooler areasi.e., those more prone to moisture and acid
condensation.) Further, certain duct configurations that produce cross-sectional flow
maldistribution may create localized wastage within the precipitator.

6.2 Leakage
Leakage of ambient air or water into the ESP or ductwork will decrease particulate collection
efficiency. Air in-leakage disturbs flue gas flow adjacent to the hole and adds to the gas load
that must be handled by the ESP and ID fans. Further, because the in-leaked air is at a lower
temperature than the flue gas, it may cause the water and acid vapor in the flue gas to condense,
which in turn can cause corrosion. Condensation increases the likelihood of ash buildup, with
corrosion occurring under the cover of the ashhidden until significant damage has occurred.
Damage from water leakage is similar only worse.
In a positive-pressure ESP, out-leakage will occur, causing different but similarly serious
problems. As noted in Chapter 5, out-leakage can promote deterioration of lagging, rappers, and
controls; moreover, the hot, SO2-laden flue gas can pose a safety hazard. Although out-leakage
will probably increase the collecting efficiency of ESP proper, the fly ash leaked out with the
flue gas will increase total plant emissions of particulate matter.
Before you determine the best way to plug a hole, verify that it was not created by design. Small
holes are required in various places, such as for pass-through of an energy transmission device or
electrical power feed, or for air intake to the support insulator purge system. Naturally, if these
intentional holes have been enlarged by corrosion, they should be repaired down to their proper
size. Note that the purge holes in the top of the high-voltage insulator caps may be larger than
inspectors sometimes think they should be, but unless there is evidence of unintentional
enlargement, they should be left alone.
Welding can repair most holes in the casing, rooftop, or metallic ducts. Be sure that the weld
material does not reduce electrical clearances or leave a sharp point for sparking from the high
voltage distribution or discharge electrode systems. Nonmetallic joints can also be repaired as
long as the patch can be made to adhere to the joint material. Avoid patching that diminishes the
ability of the joint to allow dimensional growth. If the expansion joint can no longer function
properly, the ESP or ductwork could incur even worse damage than that caused by the hole. The
caution in any joint repair is to keep the joint flexible and allow for thermal expansion in the
appropriate directions.
Be sure to replace any damaged sealing boots on access doors and ports.

6-3

EPRI Licensed Material


Corrective Measures for Common Problems

6.3 Non-optimum Rapping


As discussed in Chapter 4, the most frequent rapping problems can be investigatedand fixed
without any need for shutdown. The chief exception is a problem with transmission of the
rapping force through the plate and/or discharge electrode system; this must be investigated
during an internal inspection, as discussed in Section 5.3.5.

6.3.1 Rapping System Optimization


In rapping optimization, the chief focus is on reducing reentrainment from the outlet collection
plates, as this is the major contributor to outlet emissions from the ESP. Fly ash removed from
the collecting plates (and to some degree from the discharge electrodes) will tend to break up
when the rappers are activated. Most of the ash will fall into the hoppers, while the remainder
will be reentrained into the flue gas with the uncollected fly ash. The reentrained ash will not
usually contain any sub-micron particles, as it comprises mostly agglomerates of the previously
collected material. The reentrained material from all but the last field will usually be recollected
in the remaining downstream fields. Because there is no downstream section to recollect this
reentrained material, proper adjustment of the rapping parameters for the outlet field is especially
important to obtain optimum ESP performance.
Optimal rapping frequency and intensity are achieved by trial and errormaking an adjustment,
then checking the opacity trace to see if it had the desired effect. The process usually requires
more than a week to establish the correct parameters. The goal is to minimize rapping puffs,
which can be observed visually by use of a light and observation port near the plates of the last
field or by obscuration or opacity meters (or similar instruments) placed in the outlet duct near
the precipitator. Because the primary location for rapping reentrainment is from the outlet field
and the ash buildup rate on the outlet field is very low, several days may be required for the
rapping reentrainment conditions to stabilize after the rapping system has been adjusted. An
ESP operating with a collecting efficiency of about 99.5% will have an average ash layer buildup
rate for the outlet field on the order of 0.03 inches (0.75 mm) per day. Thus, changes in the
equilibrium thickness for the ash layer on the outlet field may require several days to become
established.
Insofar as possible, for both discharge and collection rapping systems, optimize rapping intensity
before adjusting the frequency interval. For solenoid rappers, intensity refers to slug weight,
stroke length, and in some models, spring strength. For vibrators or for acoustic horns, if
installed, intensity refers to applied voltage, or air pressure. In the case of mechanical rappers
(tumbling hammers or dropped weights), intensity refers to hammer weight and swing arc or
drop weight and distance. Note that in many rapper models, the intensity variablessuch as
spring strength, hammer weights and swing arcare not easily changed. If you have such
rappers, first optimize the frequency, which is totally controllable. If that is insufficient to
achieve good performance, you may have to consider altering the weights or other intensity
variables.
If, after these measures, rapping is still substandard, the rapping system will require significant
modification or upgrade. Recommendations are discussed below in Section 6.3.2.
6-4

EPRI Licensed Material


Corrective Measures for Common Problems

Collection Plate Rapping


Determine the appropriate intensity for the collecting system by trial and errori.e., make an
adjustment, then check results with an opacity trace. The objective should be to find the
minimum shear force necessary to dislodge the dustcake at the extremities of the system, while
avoiding over-rapping some portion of the collecting electrode. For some top-rapped systems
with very tall collecting plates, the required rapping energy can be quite high. In these cases, a
compromise must often be reached between over-cleaning the upper portions of the system
(which would create unnecessary reentrainment) while providing adequate cleaning at the
bottom.
Intensity requirements usually decrease in the direction of gas flow. Along with shear forces
(linear to the plate), some normal forces (perpendicular to the plate) are developed. Normal
forces tend to disperse the dustcake, a condition that should be minimized in the outlet sections
to limit reentrainment. Sometimes it is necessary to vary intensity across gas flow, especially
when gas velocities are unbalanced or the resistivity of the ash changes across the face because
of temperature differences. When this condition is present, the opacity profile may appear like
the one shown in Figure 4-5, where significant reentrainment is occurring in a relatively small
portion of the ESP. The instantaneous trace indicates this localized reentrainment.
When the appropriate intensity has been determined, further improvement is gained through
optimizing the frequency. A good starting point for the rapping optimization program is rapping
intervals of about five minutes for the inlet field, 1015 minutes for the second, 3045 for the
third, and 12 hours for the outlet for a four-field ESP. The settings for the inlet sections are
usually not as critical as those for the outlet. Establish reasonable settings for the two inlet
fields, and then focus on fine-tuning the outlets. About a week of operation should allow the
new conditions to become established and provide a representative opacity trace.
Discharge Electrode Rapping
In contrast to collection rapping, discharge electrode rapping performance must be assessed offline and is evaluated by visual observation and air load V-j curves rather than opacity traces.
Discharge rapping is appropriately set for near-maximum cleaning levels, because, unlike
collection rapping, there is little or no significant reentrainment penalty associated with the
discharge system. However, excessive rapping of the discharge system wastes energy and may
lead to premature mechanical failure of the discharge electrodes, as well as the rappers
themselves. The appropriate rapping level produces sufficient cleaning while limiting wear and
tear on the discharge system.
Sufficient cleaning can be tricky to determine and requires testing during shutdown. Dont let
a visual inspection fool you. As discussed in Section 5.3.5, the electrodes neednt appear clean
in a dirty visual inspection; indeed, complete cleanliness is sometimes above the capability of
the rapping system. On rigid electrodes, for example, only the tips or barbs must be clean.
Instead of relying on visual inspection, determine acceptable cleanliness by evaluating currentvoltage curves taken under both dirty and clean air load conditions during the shutdown. As ash
6-5

EPRI Licensed Material


Corrective Measures for Common Problems

buildup increases, corona start voltage increases and the V-I curve shifts toward a secondary
voltage noticeably higher than historical norms.
As with the collecting plates, adjust the rapping frequency field by field. The same time
intervals suggested for collecting plate rappingor perhaps more frequent intervalsare good
starting points.

6.3.2 Design Modifications


If rapping equipment is in good condition and the intensity and frequency settings have been
fully optimized, and yet rapping performance is still substandard, design modifications or
upgrades will likely be necessary.
The most common deficiency in rapping system design is force (acceleration) distribution.
Ideally, rapping should be performed on an infinitely small segment of the collecting system to
minimize particle reentrainment. However, practical limitations of space and cost necessitate the
usual large groupings of collecting plates on a single rapper. Generally, additional resectionalization will reduce opacity, but the point of diminishing economic return is often
reached at about 1200 ft2 (110 m2) of collecting area per rapper.
Increasing the rapping intensity can also lower average opacity. As a general rule, single-blow
rappers provide better cleaning for collecting plates and rigid discharge electrodes. Vibratorytype cleaning is best suited for weighted-wire discharge electrodes.
Be careful, however, not to use too high a rapping intensity. Sometimes plant personnel
determine that there is not sufficient energy in the rapper as evidenced by ash buildup on the
collecting plates or wires. This buildup is often quite normal; on rigid discharge electrodes, for
example, only the tips must be clean. Even if the buildup is excessive, it could be because of
poor energy transmission rather than a lack of rapping energy. Unfortunately, the bigger
hammer approach is often used. The high rapping energy from larger retrofit rappers often
causes tearing of welds or breaking of the attachment of the rapper shaft to the plate or wire
suspension system.
Additional performance improvement may be gained by converting to digital or PC-based
controls, if not so equipped. Most modern controls provide the ability to prevent simultaneous
rapper actuations, referred to as anti-coincidence. Coincident operation of rappers, especially
within the same energized field or gas flow path, can result in objectionable opacity spiking.

6.4 Misalignment or Warping of the Collection Plates


Misalignment or warping of collecting plates can result in low secondary voltages with high
values of localized current. As mentioned in Chapter 5, misalignment by no more than 1/4 inch
(0.65 cm) is desirable. If the spacing between plates or between plates and discharge electrodes
is out of alignment more than 1/2 inch (1.3 cm), it should be corrected. ESPs with wider plate

6-6

EPRI Licensed Material


Corrective Measures for Common Problems

spacing (1216 inches or 3041 cm) can tolerate a greater degree of misalignment; if
misalignment is truly a problem, it should show up on the V-j curves.
It is important to determine the root cause of plate misalignment, because if the underlying
problem is not addressed, the misalignment will probably recur. Misalignment can be classified
into two categories: dead and live.

6.4.1. Dead Misalignment


A static, permanent condition, dead misalignment can be observed during an internal inspection
after all surfaces have cooled; it can also be diagnosed from air load V-j curves. Most often, this
type of misalignment results from gradual warping of the collecting platesin height, depth, or a
combination of both. Warping generally takes place on the stiffener, or edge, of the plate. Even
though bulging sometimes occurs in the pan, or center, of the plate, straightening is usually
performed on the stiffeners only, as the pan material will usually not retain the correction.
Ash hopper overfilling (due to failure of the ash removal system) is probably the leading cause of
dead misalignment. When the ESP section remains operational despite the buildup, the ash will
continue to collect in the hoppers below. Many times the rising ash will cause an electrical
problem, shorting the section, accompanied by a control trip. A full hopper that overflows into
the plate area will probably produce a clinkerwhich will cause localized high temperature
(>1500F, >800C) that can in turn cause plate warping or perforations.
The failure of a slide plate may also cause a dead misalignment, as can a fire or any other type of
very-high-temperature excursion event. If thermal expansion bends some of the structural steel,
misalignment may occur and remain.

6.4.2. Live Misalignment


Live misalignment is a dynamic condition that arises from a conflict in expansion forces. It is
unobservable during inspections, as it only occurs while components are at operating
temperatures. Rather, live misalignment can be inferred by comparing gas load to air load V-j
curves; if the curve indicates abnormally low voltage and high current under gas load, but normal
values under a dirty air load test, and there is no back corona, the problem is probably live
misalignment.
A precept of electrode system design is that components must be held in tension (or free)
throughout the thermal cycle of the unit. Otherwise, a component will usually experience a
bowing condition that varies with flue gas temperature (live misalignment) or a permanent
deformation (dead misalignment). In addition, misalignment in the direction of gas flow, or
longitudinal misalignment, can occur due to constraints on the ESP components preventing
with-gas-flow expansion.
The root cause(s) of vertical or longitudinal misalignment should be investigated to determine
what is binding or otherwise constraining the plates. For example, if the internals were not set
6-7

EPRI Licensed Material


Corrective Measures for Common Problems

properly during construction or are not level, a dead misalignment may develop over time
because of the improper installation. The installation errors should be corrected if possible. A
permanent correction may not be possible because other factors such as poor gas flow
distribution may also be responsible. Finally, there are instances where the root cause is difficult
to identify, such as when residual stress relief is responsible.

6.4.3 Repair and Replacement Options


Once you have determined that the plates are bent, estimate the value of straightening them. Run
an ESP model with and without the repairi.e., simulate the correct plate spacing and the
faulty plate spacingto determine the performance gain that could be had from correcting the
problem. (The significant parameters to change in the model are the operating voltage and
current. A bent electrode reduces the spacing between the discharge and collecting electrode,
causing a reduction in operating voltage, which in turn influences the current.)
Note that the model assumes warpage will not continue to get worse if left as-is, which seldom
happens. As discussed above, the mechanism causing the plates to bend should be identified to
determine if further bending or warping is likely to occur.
Replacement Sometimes Necessary
If the damage is severe, replacement will be necessary. Cutting off the damaged part and
welding on a new, straight portion in its stead can replace a badly corroded plate bottom. Entire
strips of collecting plates can be replaced if the new parts can be moved into the collecting area,
either by inserting them through the hopper access doors or a slot cut in the roof of the ESP near
the area where the damage has occurred. Sometimes other internals have to be temporarily
removed to allow the plate replacement to proceed.
Repair More Common
Fortunately, replacement can generally be avoided, as most mildly bent or warped collecting
plates can be repaired by either hot or cold methods. Each has its place in the arsenal of
possibilities. The most common hot method is shrinking, a procedure that involves annealing a
very small spot on the too close side of the plate and allowing it to air cool or cooling it with a
water spray. Experienced personnel best perform this technique, but plant staff can become
fairly skilled at straightening plates in a short time. Shrinking is considered to be a permanent
fix as long as the original cause for the bowing is rectified.
Cold straightening, while more limited in application, can produce faster results. The most
common form is crimping, a procedure used on plates incorporating flat stiffeners such as
Pittsburgh Lock SeamsTM. Crimping involves twisting the seam, or rib, at various elevations
with a long-handled slotted tool. The disadvantage of this approach is that every successive
correction progressively weakens the rib.

6-8

EPRI Licensed Material


Corrective Measures for Common Problems

Recurrence of plate warping is often inaccurately attributed to previous attempts at straightening.


Warping is actually a natural process to be expected in the life of most plate designs. All plates
are cold-formed, which leaves residual stresses in the boundary region of the plate cross section.
As the outer region is removed by wastage, the plate section deforms about its core. This is
especially true with one-piece or sigma-style fabrications. Although more costly to produce,
other designs are available that should be more resistant to deformation.
An alternative to hot or cold straightening procedures is to install stiff spacer bars to reestablish
the proper clearances. Bars of various designs are available from many suppliers. The spacer
bars must often be installed in areas other than the damaged area because the corrections depend
on the stiffness of the plate metal to maintain the corrections after straightening. The primary
disadvantage of these devices is that they may distribute rapping energy across a greater number
of plates, since they tie plates together. Such plate clusters generally include more area than each
rapper was designed to clean, and thus the resultant assembly is frequently not cleaned properly,
resulting in ESP performance degradation. For plate straighteners to be effective, near-perfect
longitudinal alignment must be present or electrode-to-straightener sparking may result.

6.5 Bent or Broken Discharge Electrodes


6.5.1 Weighted-Wire Designs
The advantage of a weighted-wire design is that the wire electrodes can be assured of hanging
vertically. Hanging wires usually do not become bent beyond the capability of the weight to
keep them straight. A bottom guide grid usually restrains the weights, so the wires hang plumb
within that limitation. However, severe misalignment of the restraining gridsuch as that
caused by overfilling of the hopperscauses the wires to be misaligned with respect to the
collecting plates.
One significant drawback to weighted wires is their tendency to break. The high voltage of the
discharge electrode system tends to cause greater sparking at the top and bottom edges of the
collecting plate. This sparking eventually erodes the wire to failure, and the wire then usually
falls against the collecting plate, causing a short circuit. Most new and replacement wires have
shrouds in the areas where they pass the collecting plates to avoid this type of failure.
Replace all broken or missing wires in a weighted-wire electrode system when possible. Make
sure the new wires are equipped with shrouds at the top and bottom where the wires pass the
edges of the plates. These shrouds will minimize sparking and extend the life of the wire.

6.5.2 Rigid Wire Frame Systems


Rigid-frame discharge electrode systems are generally free from breakage, although there have
been notable exceptions to this rule, often due to an installation problem or other deficiency.
Spark erosion can also destroy the attachment integrity of the discharge wires to the frames;
consequently, some systems are equipped with attachment shunt straps to prevent sparking at the
point of support.
6-9

EPRI Licensed Material


Corrective Measures for Common Problems

Rigid frames can warp for reasons similar to those that would bend the collecting plates.
Straightening them can also be accomplished by heating and bending or crimping as described
for the collecting plates (see Section 6.4.3). Replacement of badly damaged sections should be
considered. Often where collecting plates are severely damaged, the discharge electrode system
will also be damaged.
Entrance access for replacements can be obtained by cutting a slot in the roof or moving
replacement pieces in through the hopper. As with plates, be sure to consider longitudinal as
well as vertical misalignment, and investigate the root cause to ensure the long-term efficacy of
the repair or replacement procedure.
The frame restraining grid is often held in place by an insulator attached to the hopper wall. The
insulator maintains electrical isolation for the discharge electrodes and prevents the entire
discharge electrode assembly from shifting or swaying due to either gas flow or electrical forces.
This insulator should be replaced if damaged or broken.

6.5.3 Rigid Discharge Electrodes


The advantage of rigid discharge electrodes (RDE) is that they seldom break. In the unlikely
event of a failure, they should be replaced (or possibly repaired) and the cause of the failure
investigated. The decision to repair or replace rigid discharge electrodes depends on the mode of
failure, the severity of failure, and the number of failed units. The electrodes can be removed
and replaced through the hopper.

6.6 Suboptimal Power Supply and Controls


The purpose of an ESP power supply and control system is to provide the optimum amount of
power to achieve the particulate removal required. Less than optimum power results in
inadequate performance; more power than necessary increases plant auxiliary power usage.
Figure 6-2 suggests a relationship between applied power, rapping intensity, and opacity. Fly
ash resistivity, electrical alignment of the ESP, electrical clearances of the discharge electrodes,
and the type of control system employed all have significant effects on the amount of power that
must be supplied. Even the cleanliness of the internals affects the power input. With so large a
number of variables, some of which change over time, many ESP power supply systems are
intentionally oversized (i.e., capable of providing more than the optimum amount of power) to
provide operating margins over the life of the ESP.

6-10

EPRI Licensed Material


Corrective Measures for Common Problems

Figure 6-2
Example Relationship Between Power, Rapping Intensity, and Opacity

T-R sets, however, can sometimes be undersized for present conditions. This situation could
arise if an ESP designed for a relatively high-resistivity ash is now collecting an intermediateresistivity ashdue to coal switching, the addition of flue gas conditioning, or operational
changes that altered flue gas temperature. In this case, the power density in the ESP fields (watts
per unit of collecting plate area) could easily double or triple, e.g., 1.8 W/ft2 (19 W/m2) vs. 0.5
W/ft2 (5.4 W/m2). Also, older controls that permit operation in back corona can have very high
power densities, e.g., 1.52.0 W/ft2 (1622 W/m2), although not all of this power is useful. TR sets with inadequate power typically require some combination of the replacement of the
transformer with one of a higher rating and the addition of new power supplies. Such a retrofit
will enable higher power densities in the collecting region, which in turn will yield a small boost
in collection efficiency.
Replacement of an antiquated control system will often improve ESP performance, increase
reliability, and save energy. Volume 2 of this guide discusses available options. In general, the
big improvement comes from upgrading analog controls to digital. Changing from one type of
digital control to another may not appreciably change performance.
Note that upgraded controls offer limited performance improvement potential for units with
moderate ash resistivity; benefits are usually greatest for high-resistivity ashes. Accordingly, it
is worthwhile to consider specific precipitator challenges (e.g., back corona), control system
6-11

EPRI Licensed Material


Corrective Measures for Common Problems

response, and potential modifications or partial replacement before replacing power supply and
control systems en masse.

6.6.1 Automatic Voltage Control


Among other functions, the automatic voltage control (AVC) optimizes the power to the ESP
and protects the power supply and ESP internals. The control algorithm employed by the AVC
increases the voltage until the system sparks (or nearly sparks) and then backs down the voltage.
The voltage is again increased until the ESP sparks and then the voltage is stepped back. The
AVC is always searching for the optimum position and sits still only when the current or
voltage limit of the transformer-rectifier is reached and the ESP does not spark at that point.
AVCs are classified as either quench-only or differential. Quench-only controls reduce power to
zero immediately after a flashover of any type between the electrodes. This control design exists
in analog, digital, and microprocessor versions. Differential controlsavailable only in
microprocessor versionsregulate applied power according to the amplitude of the disturbance.
These controls attempt to differentiate between discrete levels of flashover, such as a spit spark
(low level), spark (normal level), and arc (highest level of electrical disturbance), and regulate
accordingly.
Many old ESPs were installed with earlier versions of analog controls, which attempt to
maximize the electrical field strength by increasing the applied voltage up to a predetermined
number of sparks per minute (the AVC counts the sparks). Although these older analog systems
are effective at maintaining acceptable ESP performance under many operating conditions, they
tend to consume a lot of power and in some instances may not adequately protect the ESP
internals from arcing damage. The risk of damage is greatest with AVC systems that are slow to
detect sparks. If excessive sparking or arcing is a problem, the typical remedy is to replace the
AVC.
A leading operational challenge to the AVC is back corona, a common problem for plants that
burn low-sulfur, low-sodium, high-resistivity coals. Back corona occurs when the current flow
through the collected ash layer establishes an electric field in the layer sufficient to cause the
formation of a corona. The actual breakdown occurs in the gas contained in the interstices of the
ash layer. The breakdown is analogous to the breakdown adjacent to the corona wire in normal
ESP operation.
There are two manifestations of high-resistivity ash related to back corona. In the instance of
incipient back corona (typical of moderately high ash resistivity), a localized breakdown in the
ash layer causes sparks to develop in the layer and travel from the plate to the discharge
electrode, at operating voltages less than would normally cause sparking. In the instance of
actual back corona (typical of high ash resistivity), there is a general electrical breakdown over a
substantial area in the ash layer.
Analog systems that only count sparks will allow operation in back corona. The spark rate
controller does not recognize conditions of impending back corona and increases the conduction
angle on the SCR trying to raise the voltage to establish sparking. However, instead of sparking,
6-12

EPRI Licensed Material


Corrective Measures for Common Problems

back corona forms, and the increase in conduction angle carries the power supply to current
limit. In contrast, a digital, PC-based system can detect operation tending toward back corona
and reduce the power input to maintain operation before back corona formation.
In other instances with higher ash resistivity, there is a general electrical breakdown in the entire
ash layer causing a great increase in current flow at lower voltages. Older power supply controls
will operate near current limit under these circumstances, while computer-based control systems
can be programmed to operate at a point just at the beginning of back corona formation. If back
corona is a persistent problem, upgrading to a digital control system is recommended.

6.6.2 Current-Limiting Reactor


The current-limiting reactor (CLR) is a single-winding, iron core, air-cooled device installed in
series with the silicon controlled rectifier (SCR) and the transformer-rectifier. The CLR protects
the SCR and T-R against high current transients or a short circuit. Additionally, the CLR
provides wave shaping to help produce an optimum secondary waveform.
The CLR is a large inductor that is tapped at several increments of inductance. Generally,
50% impedance is selected to produce desirable results. Less impedance results in unstable
controller operation, which leads to a high incidence of broken weighted-wire electrodes.
Excessive impedance will result in a choking effect, reducing the capacity of the T-R. Select the
appropriate tap to match the conditions in a particular electrical section.

6.7 Inadequate Electrical Sectionalization


Your unit may suffer from inadequate sectionalization if (1) power supplies operate at or near
current limit when the V-j shape is good, or (2) there are opacity excursions when the spark rate
controller backs down voltage, or (3) your ESP exhibits sparking at a somewhat lower-thanexpected voltage, and yet high ash resistivity or mechanical misalignment are not responsible.
Ideally, ESPs would be designed with a transformer-rectifier set for each individual discharge
electrode. Since the cost would be ridiculously high, ESP suppliers strike a compromise,
providing the sectionalization needed for performance but maintaining cost at a reasonable level.
A rule of thumb is that a single electrical section should include about 20,000 ft2 (1860 m2) of
collecting plates, which should all be in the same electrical field. Of course, there are successful
ESPs that have more collecting area per T-R set than this rule of thumb suggests, and there are
those with less. Sectionalization design varies among suppliers based on their actual experience.
The performance improvement expected after re-sectionalizing the ESP can be estimated with an
ESP computer model. Re-sectionalizing can be fairly expensive unless the original design
included space and opportunities for re-sectionalizing. (In some of the older units, sections
across the ESP are fed by a single transformer, even though individual regions may have
independent insulators. These can be easily re-sectionalized by adding T-R sets.) For most
installations, it is necessary to add new support and feed-through insulators as well as new T-R

6-13

EPRI Licensed Material


Corrective Measures for Common Problems

setswork that is beyond the normal capability of plant maintenance personnel. In such cases,
this is an option for a serious ESP rebuild, and is described in Volume 2 of these guidelines.

6.8 Non-optimum Flue Gas Parameters


Flue gas physics and chemistry influence ESP behavior. An ESP is designed for a specified
volumetric flow rate, temperature and temperature distribution, and gas velocity distribution. In
addition, flue gas conditionssuch as sulfuric acid or SO3 concentrationsinfluence fly ash
resistivity for cold-side installations. If actual flue gas conditions fall outside the design ranges
(as could occur when switching to a low-sulfur coal), fly ash resistivity may lie outside the
expected ranges and ESP performance may be less than anticipated.

6.8.1 Volumetric Flow Rate


The flue gas volumetric flow rate is established by coal chemistry, boiler size/design and
combustion conditions, plus any in-leakage in the ductwork leading to the ESP and, for cold-side
units, across the air heater. The flow rate is critical, given that the collection capability of an
ESP can be correlated to its specific collecting area (SCA)an exponential relationship
involving the ratio of the collecting plate area to the gas volume flow rate. If the flue gas flow
rate is greater than the design value, the collection efficiency will drop. Consequently, if volume
flow rate is excessive, the cause(s) should be determined and corrected. Poor furnace
combustion control and/or holes in the ductwork, air heater seals, door seals, and expansion
joints are likely candidates for correction.

6.8.2 Flue Gas Temperature


Temperature affects ESP collection efficiency primarily by affecting the electrical resistivity of
the fly ash particles. If the temperature is higher than expected for cold-side units or lower than
expected for hot-side units, the resistivity will be higher than expectedsometimes making the
fly ash more difficult to collect.
Figure 6-3 shows a plot of the resistivity of a particular fly ash as a function of temperature and
flue gas SO3 concentration. If the ESP is expected to operate at a temperature of 275F (138C),
the anticipated resistivity is about 1 x 1011 -cm, if the fuel produces about 1 ppm of SO3 in the
flue gas. Although this value of resistivity is somewhat above optimum, the design SCA for the
ESP will be moderately small compared to that required for higher resistivities.
If, after the plant is built, changes in furnace design or combustion conditions cause the flue gas
temperature to rise above the design temperature to 300F (150C), ash resistivity will increase
to about 5 x 1011 -cm, making the particles much more difficult to collect. Note that the air
heater can contribute to increased ash resistivity for a portion of the flue gas flow. A Ljungstrom
air heater, for example, typically produces a temperature gradient of about 50F (28C) across
the ductwork. As shown in Figure 6-3, the resistivity on the high-temperature side of the
ductwork will be on the order of 1 x 1012 -cmmaking that portion of the fly ash extremely
6-14

EPRI Licensed Material


Corrective Measures for Common Problems

difficult to collect with an ESP. Such a high resistivity would require flue gas conditioning with
SO3 to maintain near-expected ESP performance.
You may be able to correct the problem by adding or modifying baffles to promote more even
flue gas mixing. If this fails to achieve the desired improvement, you will need to lower the ash
resistivitytypically through flue gas conditioning, which is discussed in detail in Volume 2 of
this guide.

6.8.3 Gas Flow Distribution


In optimizing gas flow, the aim is generally to create a more uniform flow, given that, when
considered from first principles, the collecting efficiency of an ESP is the best that can be
expected when the gas velocity distribution (and particle distribution) within the collecting zone
is uniform. The Institute of Clean Air Companies (ICAC, formerly IGCI) has set guidelines for
the appropriate gas velocity distribution in ESPs: 85% of the measurement points should have
velocities less than 1.15 times the average velocity, while 99% of the points should have
velocities not more than 1.4 times the average velocity. The ICAC standards are considered to
be very good. If the gas velocity distribution in an ESP meets ICAC guidelines, further
modifications to the gas flow distribution will provide very little, if any, incremental
improvement in collection efficiency.

6-15

EPRI Licensed Material


Corrective Measures for Common Problems

Figure 6-3
Fly Ash Resistivity as a Function of Temperature

However, in practice, many ESPs do not meet these standards and exhibit undesirable gas flow
characteristics that, when corrected, could significantly boost collection efficiency. Very poor
gas velocity distributions can arise because a careful gas flow model study was not performed
and implemented initially, or because events within the ESP or upstream of it are affecting the
gas flow.
6-16

EPRI Licensed Material


Corrective Measures for Common Problems

Two examples illustrate this point. In one case, the power station was equipped with a
multiclone particle collecting system preceding an ESP. The unit burned a higher-sulfur coal
that produced a flue gas with sufficient sulfuric acid to corrode the multiclone system. As the
multiclones deteriorated, some of the flue gas would flow through the holes in the multiclones,
causing a very poor gas velocity distribution in the ESP and severely degrading overall
precipitator performance. The solution to this problem was to remove the old multiclone.
Thereafter, the ESP alone provided about the same collecting efficiency as the previous
combination when both collectors were operating in tandem. This result is not unexpected, as an
ESP provides very high collecting efficiencies for the larger particles that were previously
removed in the cyclonic pre-collectors.
The second example concerns an ESP that was not equipped with anti-sneak baffles in the ash
hoppers. This particular plant burned a coal with an ash content approaching 50%, and thus
required very large ash hoppers. An ESP model study indicated that the addition of anti-sneak
baffles in the hoppers would reduce the outlet emissions by more than half. Thus, careful gas
velocity distribution control was critical.
Use one of the ESP models discussed in Chapter 3 to evaluate your units gas velocity
distribution measurements (discussed in Section 5.7). If the measured data are such that gas =
0.25 or more and S = 0.12 to 0.15 or higher, there is ample opportunity to improve ESP
performance by modifying the gas velocity distribution and gas flow baffles.
Before embarking on such an improvement, it will be necessary to model the gas flow in detail.
This complex exercisewhich requires modeling a significant portion of the boiler, ductwork,
particulate collecting, and stack systemwill usually require hiring an expert with previous
experience in gas flow studies. Several contractors specialize in either physical or computer
model studies; both types of study are probably equally successful in establishing a good gas
flow distribution and the costs are generally comparable. Alternatively, some ESP
manufacturers can also provide this modeling service, and bring not only modeling experience
but also the expertise to provide workable corrections for the unit.
The model study will provide a description of the modifications needed to correct the gas flow in
the ESP. These corrections can usually be installed during a short plant outage. Another round
of flow measurements is appropriate after modification, to verify that the new distribution
achieves the modeled performance projections.
Note that consideration is currently being given to the use of a non-uniform gas velocity
distribution to compensate for some rapping reentrainment situations. Data available are
inconclusive, but the concept seems to have validity. Volume 2 of this guide discusses this
skewed flow redesign concept in greater detail.

6-17

EPRI Licensed Material


Corrective Measures for Common Problems

6.9 Undesirable Fly Ash Properties


The most common characteristic limiting ESP performance is high electrical resistivity of the fly
ash. In some cases, very low resistivity is also a problem. It is important to know the resistivity
of your ash in order to determine whether certain improvement optionssuch as flue gas
conditioning or upgrading to digital controlsare likely to provide a significant boost in
collection efficiency. Other ash properties that impair ESP operation include an unexpectedly
high fraction of fine particles and unusually high amounts of unburned carbon (high LOI), as
often occurs following the retrofit of low-NOx burners.

6.9.1 High Resistivity


High-resistivity ashes create very poor electrical conditions, causing greater sparking or back
corona at relatively low voltages; such ashes may even impede effective rapping. If resistivity is
high enough to cause back corona, this will be evident from the V-j curves and secondary
waveforms from the gas load test.
In some cases, high resistivity is just a localized problem due to poor temperature distribution
(i.e., a hot spot), as discussed in Section 6.8.2. To confirm the problem, take readings from a
temperature probe traverse across the duct just upstream of the ESP inlet using a grid sampling
pattern. (If conducting this test, consider also making mass flow measurements to verify particle
distribution uniformity. High mass loading of very fine particles at portions of the inlet can
cause corona quenching, which hampers ESP performance.) The highest recorded temperature
should not exceed the mean value by more than about 25F (14C). If a more substantial hot
spot is observed, assess whether repositioning the guide vanes or ESP inlet flow distributors can
produce greater flue gas mixing.
Appendix D discusses fly ash conduction mechanisms, providing a good background into the
characteristics of fly ash that influence the electrical behavior of the ESP. For both hot-side and
cold-side ESPs, where the volume conduction mechanism is active, fly ash resistivity is
determined by the chemical composition of the fly ash material and the operating temperature of
the ESP. In cold-side ESPs, where the surface conduction mechanism is also active, resistivity
is also influenced by flue gas composition.
There are two basic approaches to reducing fly ash resistivity: changing the coal, or conditioning
the flue gas.
Coal Switching
In many cases, the most direct way to correct high resistivity is to change or blend the fuel
supply.
For hot-side installations, simply switching to a coal with a higher percentage of sodium in the
fly ash will usually alleviate the problem. Blending with a higher-sodium coal is also worth
6-18

EPRI Licensed Material


Corrective Measures for Common Problems

considering; however, blending adds some degree of complexity to power station operation.
Alternatively, sodium-containing compounds such as sodium carbonate or sulfate can be added
to the coal in the bunkers prior to combustion. A word of caution about increasing the sodium
concentration in the fuel supply: Higher concentrations of alkali metals in the fuel can sometimes
aggravate boiler slagging and fouling. Be sure to carefully monitor the furnace during any
experiments involving higher sodium content.
For cold-side units, options for coal switching are quite limited. Fly ash resistivity for cold-side
units is usually governed by the sulfur content of the coalthe higher the sulfur content, the
lower the resistivity. Yet switching to a higher-sulfur coal is generally not an option given acid
rain regulations restricting SO2 emissions. Thus, flue gas conditioning is the corrective measure
of choice for cold-side units suffering from high ash resistivity.
Flue Gas or Fly Ash Conditioning
In cold-side units, modifying the flue gaswith SO3, moisture, ammonia, or other agentscan
reduce fly ash resistivity and thereby improve ESP collection efficiency. To get a rough idea of
the potential performance improvement for your unit, assume a resistivity value of 1 x 1010 -cm
on Figure 2-2, select the corresponding average current density, and use the appropriate curve
among Figures 2-3 through 2-8. The ESP models discussed in Chapter 3 provide a much more
accurate prediction of unit performance improvement with flue gas conditioning.
Various flue gas conditioning technologies are discussed in detail in Volume 2 of this guide.

6.9.2 Excessively Low Resistivity


Very-low-resistivity ash particles (below 108 -cm) are collected very effectively. The problem
is that they are also readily reentrained.
Once collected in an ESP, ash particles are held on the collecting plates by a combination of
cohesive and adhesive mechanical forces, Van der Waals forces, and electrical forces. The
magnitude and direction for the electrical force is governed by the electrical resistivity of the ash
and the electrical current density in the ESP. The electric field in the collected ash layer is the
product of the ash resistivity and the current density. Thus, for ash with high resistivity, the
electric field in the ash layer will also be high. Conversely, if the ash resistivity is very low, the
electric field in the ash layer will be low. The electric field in the ESP inter-electrode space,
adjacent to the collecting electrode, is typically in the 513 kV/inch (25 kV/cm) range. This
value is established by a combination of discharge electrode geometry, electrode spacing,
particle and ionic space charge, and applied secondary voltage on the ESP.
If the electric field in the space adjacent to the collected ash layer is lower than the field in the
ash layer itself (this will be the case for mid- to high-resistivity ashes), there will be an abrupt
change (a step function) in the electric field at the ash surface. A step in the electric field
requires a surface layer of electrical charge to provide this abrupt change. When the electric
field is higher in the ash layer than in the space adjacent, the surface charge layer will be of
6-19

EPRI Licensed Material


Corrective Measures for Common Problems

negative polarity (for ESPs with negatively charged discharge electrodes), causing the collected
layer to be pressed toward the collecting electrode by the applied electric field. The electric field
distribution in the inter-electrode space and in the ash layer is illustrated in Figure 6-4.
Contrast this situation with the case where the ash resistivity is low and the electric field in the
ash layer is less than that in the adjacent space. This condition leads to the formation of a
positive surface charge layer on the ash. The action of the electric field on the positive surface
charge layer is to change the electrical force from a clamping force to one that is directed back
into the gas stream. When this electrical force is greater than the combined mechanical and Van
der Waals forces, the particles will be electrically reentrained. Erosive forces from the gas flow
can also contribute to this reentrainment.
This mechanism for particle reentrainment was demonstrated and identified in a pilot-scale test
of a spray dryer operating ahead of an ESP, when the system was operated near the moisture dew
point. These conditions produced a composite ash and spray dryer residue resistivity of about
107 -cm; significant electrical reentrainment occurred. A small increase in temperature to
increase the resistivity eliminated the problem.
This type of reentrainment mechanism is responsible for the high emission or opacity problems
sometimes encountered in plants with significant quantities of unburned

6-20

EPRI Licensed Material


Corrective Measures for Common Problems

Figure 6-4
Electric Field as a Function of Position in the Inter-electrode Space

carbon in the fly ash. Typically, if the loss on ignition (LOI) is less than 56%, the unburned
carbon does not cause a problem. (In some cases where the ash resistivity is marginally high, the
higher LOI conditions can actually improve ESP performance). But, if the LOI due to unburned
carbon exceeds 910%, the carbon particles may be preferentially reentrained and contribute
significantly to overall particulate emissions. In this case, the percent LOI for an outlet mass
sample from an ESP will be significantly greater than that for an inlet mass sample. The
problem of carbon carryover into the ESP outlet is frequently found in plants that have converted
to low-NOX burners.
6-21

EPRI Licensed Material


Corrective Measures for Common Problems

Combustion engineers may be pursuing solutions to reduce LOI for other performance
objectives, and in this case, high reentrainment due to low resistivity will likely be corrected as
well. Otherwise, flue gas conditioning with ammonia or a combination of ammonia and SO3
may be required; see Volume 2 for discussion of flue gas conditioning options.

6.9.3 Excessive Fine Particles


Fly ash with a very fine particle size distribution poses yet another limitation to ESP
performance, as smaller particles are more difficult to collect (see Appendix C). Western
subbituminous coals, for example, produce a high fraction of fine particles in the fly ash. An
increase in the proportion of fine particles can also occur after retrofitting low-NOX burners
(with LNB, the coal is often ground finer in an attempt to reduce the amount of unburned
carbon).
The fine particle fraction is formed by two mechanisms. First, as the coal burns, moisture in the
coal particles explosively evaporates, causing localized spalling of fine particles from the
surface of the coal. The finer fraction is considered to be formed as vaporized constituents in the
coal recondense as the flue gas cools. There is very little you can do to avoid these mechanisms.
In fact, there is very little you can do to increase the particle size distribution from a furnace.
Although there is a relationship between fineness of coal grind and particle size distribution,
changing the fineness of grind is usually not an option, since other factors are also changed.
Restoring ESP collection efficiency thus requires increasing the design capacity of the ESP
through a rebuild, addition of a polishing device, or wholesale replacement with a larger ESP or
fabric filter. Such upgrades are addressed in detail in Volume 2 of this report.

6-22

EPRI Licensed Material

A
DATA SETS USED TO GENERATE PERFORMANCEESTIMATING CURVES

The graphs presented in Chapter 2 for estimating the expected best performance of a normally
operating ESP were derived from a statistical analysis of performance data from numerous fullscale precipitators. These data were collected over several years in projects conducted for EPRI
and EPA.
The data sets consisted of measurements of inlet and outlet mass loadings, particle size
distributions, electrical resistivity as measured in situ and in the laboratory, and the electrical
readings from the power supply controllers. In most cases, the secondary voltages were
measured with calibrated voltage dividers and the current meters were verified by measuring the
value of the resistor in the ground return leg of the power supplies.
The electrical data were obtained from 17 field tests on 13 different cold-side ESP units at 12
plant sites. The ESP units had 3 to 6 electrical fields in the direction of gas flow. The electrical
resistivity of the fly ash ranged from the mid-1010 -cm to the mid-1012 -cm level. Four of the
units were tested with and without flue gas conditioning for resistivity modification. Table A-1
provides the actual value of current densities for each data set together with the in situ resistivity.

Where the ESPs were operating with back corona, the voltage vs. current curves were examined
to estimate the useful current density. The current density selected was that value where the V-I
relationship changed curvature noticeably.
The secondary voltages were evaluated in a similar fashion. The collecting fields plate-to-plate
spacing varied from 9 to 12 inches (23 to 30 cm). The operating voltages for each electrical field
are given in Table A-2.
The operating current densities for the first three fields of the ESPs in Table A-1 are plotted in
Figures A-1, A-2, and A-3 against the measured in-situ resistivity. The solid curve fit lines
(one per figure) indicate the relationships between current density and resistivity for wellperforming ESPs. They can be used to estimate the useful current density in a field when in situ
resistivity is known, or conversely, to estimate resistivity when a useful current density value
(i.e., no back corona or arcing) for one of the first three fields is known. Included in these
figures are lines representing an electric field in the ash layer of 1 kilovolt per centimeter and 15
kilovolts per centimeter, based on average values of current density (breakdown would occur
based on the higher local value of current density, but current density distributions were not
known).

A-1

EPRI Licensed Material


Data Sets Used to Benerate Performance-Estimating Curves

The data summarized in this appendix were analyzed to develop a set of generalized voltage vs.
current curves for the test ESPs. The voltages were adjusted for the different plate spacings to be
appropriate for a generic ESP with 10-inch (25-cm) plate spacing. These voltage vs. current
density curves are presented in Figure A-4.
Table A-1
Fly Ash Resistivity and Estimated Useful Current Densities
(see Appendix E for metric conversion factors)
Test

Resistivity cm

Inlet
2
A/ft

Section 2
2
A/ft

1.5E12

3.0

3.0

1.6E10

11.5

17.5

5.0E11

5.0

03.0

1.0E11

18.6

25.2

3.6E11

11.4

15.3

2.6E10

14.5

6.9E12

A-2

Section 3
2
/ft

Section 4
2
A/ft

Section 5
2
A/ft

Section 6
2
A/ft

10.8

16.3

26.8

31.1

16.2

16.2

16.4

14.5

32

0.5

1.5

3.0

5.5E12

0.7

0.8

2.0

2.2

5.4E11

4.0

5.0

8.0

13.0

10

2.2E10

18.8

28.4

38.6

46.0

11

1.0E12

1.7

4.0

7.0

8.0

12

6.0E10

6.2

11.0

24.3

25.3

13

2.0E11

8.5

6.0

8.0

12.0

13.0

14

7.0E11

4.4

6.0

16.0

22.0

23.0

15

2.1E10

12.8

45.5

53.0

54.0

58.0

49.4

16

5.8E11

6.0

10.0

10.0

10.0

17

4.1E10

16.1

24.9

13.8

23.8

29.8

36.8

23.6

17.0

17.0

EPRI Licensed Material


Data Sets Used to Benerate Performance-Estimating Curves
Table A-2
Operating Secondary Voltages for Each Electrical Field (kilovolts)
(see Appendix E for metric conversion factors)
Test No.

Plate Spacing (inches)

Inlet

Section 2

Section 3

Section 4

Section 5

Section 6

35

35

11

56

44

43

38

51

40

38

37

40

39

43

39

40

41

41

41

35

44

25

29

26

12

39

31

40

34

9.75

37

40

36

37

10

10

48

50

48

48

11

33

31

26

24

12

38

35

36

32

13

12

41

39

52

38

36

36

14

12

56

59

59

15

10

45

49

46

40

38

35

16

34

40

38

33

17

39

41

36

41

37

40

43

34

A-3

EPRI Licensed Material


Data Sets Used to Benerate Performance-Estimating Curves

Figure A-1
Current Density vs. Ash Resistivity for Inlet Section of an ESP

A-4

EPRI Licensed Material


Data Sets Used to Benerate Performance-Estimating Curves

Figure A-2
Current Density vs. Ash Resistivity for Second Section of an ESP

A-5

EPRI Licensed Material


Data Sets Used to Benerate Performance-Estimating Curves

Figure A-3
Current Density vs. Ash Resistivity for Third Section of an ESP

A-6

EPRI Licensed Material


Data Sets Used to Benerate Performance-Estimating Curves

Figure A-4
Voltage vs. Current Curve Composite From Seventeen ESP Tests Used for Correlation (All
Normalized to 10-inch Plate Spacing)

A-7

EPRI Licensed Material

B
PRECIPITATOR PERFORMANCE EVALUATION USING
ESP MODEL

This appendix offers a step-by-step example of using an ESP model to evaluate an existing ESP.
The example applies to all versions of the core model developed by Southern Research
Institutethe SRI/EPA version, and EPRIs ESPert and ESPM models. ESPM is the simplest to
use and therefore recommended unless you already have experience with one of the other
models.
The ESP model is a useful tool for comparing the actual performance of an ESP to that which
could be expected for the conditions existing in any coal-burning power station. If the actual
performance is less than predicted, the model is also useful for investigating the reasons for the
substandard performance. Before any upgrade is undertaken, a systematic evaluation program
such as that described in the body of this volumeshould be used to determine if the ESP is in
good mechanical and electrical condition.
In order to evaluate a precipitator, certain data are required for the analysis. Mandatory inputs
are listed in Table 3-1; these include gas temperature and velocity distributions in the inlet and
outlet ducts, inlet particle size distribution, fly ash resistivity, T-R secondary voltage-current
readings, and electrical operating points. These data, in addition to careful inspection of
precipitator components, can be used in a variety of methods to evaluate a precipitator.
Comparing measured precipitator data with the ESP performance predicted by the model
provides the opportunity to diagnose the possible reasons for a degraded level of precipitator
performance. There are various ways to conduct such an analysis; this appendix demonstrates
one such approach. Note that the sequence of data gathering referred to in this appendix does not
always match the sequence of data gathering described in the body of this guide; however, the
necessary steps are identified and can be re-sequenced as needed to match the logistics of a
particular situation.
The procedure adopted in this example begins with the assumption that the precipitator is in
good mechanical and electrical condition and is operating with maximum useful current density.
If the actual performance is less than the model-estimated performance under these conditions,
the following steps are taken to examine the potential causes for the degraded performance.

B-1

EPRI Licensed Material


Precipitator Performance Evaluation Using ESP Model

Step 1. Determine the estimated collection efficiency for gas = 0.25


and S = 0.10.
These values are considered to be reasonable for an older ESP that has been well maintained. (If
your ESP is newer with a better gas flow design, use gas = 0.10 to 0.15, and S = 0.05 to 0.08.)
The estimated collection efficiency for these starting assumption values of gas and S can be
obtained by exercising the ESP model with the specific collection area from the measured gas
flow and the known plate area and from the measured electrical operating points for each power
supply T-R set.
This model-projected collection efficiency should be attainable if the precipitator is in good
mechanical condition. It should be emphasized that the measured collection efficiency may not
necessarily represent the best possible performance for an ESP, even though it has good gas
velocity distribution, low gas sneakage, and low reentrainment losses without rapping.
Excessive accumulation of fly ash on the discharge and collection electrodes, broken discharge
electrodes, electrode misalignment, and warped collection plates will result in degraded electrical
operating points and degraded collection efficiency. It is important to take steps to optimize the
electrical conditions in the precipitator before obtaining data for use in this evaluation procedure.
These steps involve correcting any existing electrical problems mentioned above and perhaps
cleaning the electrodes.

Step 2. Compare the estimated performance to the measured collection


efficiency.
If the measured collection efficiency is slightly lower than the model-projected performance, the
discrepancy may be caused by operating conditions that are different from the typical estimates
used with the model; refer to Step 3. If the measured collection efficiency is significantly less
than the modeled value, the precipitator is performing poorly. Refer to Step 4.
If the measured collection efficiency is equal to or greater than the estimated value, the
precipitator can be assumed to be performing well for the given set of operating conditions. It
should be emphasized that the performance of a precipitator can change while all precipitator
components are functioning properly. Changes in the inlet particle size distribution, the fly ash
electrical resistivity, the gas temperature, or the gas volume flow can change the collection
efficiency. Thus, a precipitator may not be able to achieve the collection efficiency it once
achieved, or was designed to achieve, simply because of changes in the process variables. In this
case, the precipitator may no longer be sized properly for the actual operating conditions
encountered.
If it is verified that there are no serious mechanical problems, then the options that are available
for improving the performance of the precipitator are limited to an improvement in the electrical
operating conditions, an improvement in the gas velocity distribution, or a reduction in the gas
flow. A small improvement in performance may be gained by correcting minor electrical
problems (such as T-R sets with inadequate power ratings or malfunctioning automatic
controllers) that may limit the electrical operating points. However, if the measured fly ash
B-2

EPRI Licensed Material


Precipitator Performance Evaluation Using ESP Model

resistivity is greater than 1 x 1011 -cm, flue gas conditioning to reduce the resistivity may be
considered as an option. This should result in significantly higher operating voltages and
currents without excessive sparking or back corona.

Step 3. If measured collection efficiency is only slightly less than the


estimated value, investigate operating parameters.
A measured collection efficiency that is only slightly less than the estimated value may be
caused by operating conditions that are different from the typical data that were chosen for the
estimates used with the model. Check to see if small differences in the estimated and measured
efficiencies can be caused by differences in such factors as the inlet particle size distribution or
in the electrical input data selected. A greater mass loading of sub-micron particles or a lower
average electric field will result in a lower value of measured collection efficiency. If this is the
case, then the conclusions reached in Step 2 will still be applicable. If not, then the non-ideal
operating parameters should be investigated.
First, evaluate rapping reentrainment for the unit. Vary the rapping frequencies and intensities
(if possible) to minimize losses in collecting efficiency from rapping reentrainment. Increasing
the time between raps can sometimes reduce emissions. If this is not successful in bringing the
measured performance to match the estimated performance, then the gas velocity distribution
over the face of the ESP and the fraction of gas passing below and above the electrodes should
be measured under air load conditions. If the measured data are such that gas 0.25 and S
0.10, then the gas velocity distribution and gas flow baffles should be improved.
If these steps are not helpful, there is a possibility that non-rapping reentrainment is degrading
the precipitator performance. A high average gas velocity, excessive sparking, low fly ash
resistivity, low electrical operating points, or reentrainment this from the hoppers could be the
cause. These potential contributing factors should be examined. If non-rapping reentrainment is
sporadic, a continuous opacity monitor at the precipitator outlet can detect it.

Step 4. If measured collection efficiency is significantly lower than


estimated performance, determine cause of poor performance.
If the measured collection efficiency, with good mechanical conditions, is significantly less than
the modeled value, then the precipitator is performing poorly. Any one or a combination of
several factors, each of which can be analyzed using the ESP model, may cause poor
performance. These factors include electrical operating conditions, non-uniform gas velocity
distribution, gas sneakage past the electrified fields, particle reentrainment without rapping, and
rapping reentrainment. The following steps describe procedures to pinpoint other problem areas.

B-3

EPRI Licensed Material


Precipitator Performance Evaluation Using ESP Model

Step 4.1. Determine if the operating currents are completely useful in the
precipitation process.
At this point, the electrical operating conditions should be examined to determine whether or not
the operating currents are completely useful in charging and collecting particles. If excessive
sparking or back corona is occurring in the ESP, then the measured currents will not be
completely useful in the precipitation process. On the contrary, the currents may actually be
detrimental to the performance. ESP model input data, which use the values of currents
measured under these conditions, may result in the prediction of much higher collection
efficiencies than can be achieved in practice.
Excessive Sparking
Sparking consists of highly localized currents that are not effective in charging particles.
Furthermore, excessive sparking can lead to increased particle reentrainment. Excessive
sparking disrupts the surface of the collected fly ash layer and reduces the electrical holding
force over large regions of the layer because of decreased currents to those regions. If sparking
is occurring, then the extent of the sparking should be measured with spark rate meters.
Acceptable levels of sparking are in the range of 1060 sparks per minute. If excessive sparking
is occurring, then the operating voltage should be lowered; the power supply controllers should
be adjusted or upgraded. Although the operating voltages and currents will be lower, the ESP
performance may improve (unless the problem is caused by severe ESP mechanical problems)
and the use of the lower operating electrical conditions in the ESP model will give better
agreement between estimated and measured collection efficiencies.
Back Corona
In a condition of severe back corona, both positive and negative ions move in the inter-electrode
space. The measured current can be very much greater than the current that is actually useful in
the precipitation process. The detection of back corona was discussed in Chapter 4. The
measured value of fly ash resistivity can be used with Figure 2-2 to estimate the maximum
allowable average current density. The curves shown in that figure are based on the average of
17 sets of measurements of in situ fly ash resistivity and useful current density in each electrical
section of fly ash ESPs. If the average current density in the ESP greatly exceeds those values
indicated for a given fly ash resistivity, the ESP may be operating in back corona.
As a second check, the secondary voltage-current curve for each electrical section can be
examined to see if at some point on the curve increased current is obtained with decreased
voltage (see Figure 4-21, 4-22, or 4-23). If the ESP is operating on this portion of the voltagecurrent curve, then it is operating in back corona. As another check, the secondary voltage
waveform can be displayed on an oscilloscope. A rectified voltage waveform that periodically
dips near or below the voltage value at corona start (where ripples in the voltage waveform first
appear) indicates back corona (see Figure 4-19).

B-4

EPRI Licensed Material


Precipitator Performance Evaluation Using ESP Model

If severe back corona is occurring, the operating voltage should be lowered to operate with a
lower current density that will not lead to the formation of back corona. The reduced voltages
and currents will usually result in improved performance, and the use of the lower electrical
operating points in the ESP model will give better agreement between estimated and measured
collection efficiencies. These operating points suggested for high-resistivity fly ash can be more
readily obtained with one of the more modern PC-based power supply controls. Since back
corona and sparking caused by high fly ash resistivity will degrade the ESP collecting efficiency,
flue gas conditioning can be considered as a means for improving the performance.
Electrode Misalignment
If the spacing between plates or between discharge electrodes and plates is out of alignment by
over 1/2 inch (1/4 preferred), then corrective measures should be taken. Significant
misalignment of the electrodes is detectable from secondary voltage-current curves obtained
under air load conditions. The air load V-j curves should be the same for different electrical
sections with similar geometry, up to the current limits of the T-R sets. If these curves differ
significantly, then misalignment of some of the electrodes is expected. Warped plates can result
in very low secondary voltages with high values of localized currents. The electrode alignment
should always be checked to avoid reaching wrong conclusions. For example, low secondary
voltages may be attributed wrongly to high fly ash resistivity, when mechanical deformation of
the electrodes may be responsible.

Step 4.2. Determine if performance is limited by non-uniform inlet gas


temperature or fly ash loading.
Excessive temperature variation across the face of the ESP or mass loading at the inlet of the
precipitator can be detrimental to the ESP performance. The uniformity should be checked by
sampling the gas temperature and mass loading over the inlet duct of the precipitator, under
actual operating conditions. A temperature variation of 90OF
(50 OC) would cause an order-of-magnitude change in the electrical resistivity of most fly ashes
(see Figure 6-3). Even large diurnal variations in the ambient temperature can be detectable in
the ESP performance. Large increases in the particulate mass loading or perhaps the
condensation of sulfuric acid could cause corona quenching. Fly ash accumulation on some
parts of the inlet electrical sections may also cause disruptions. Either type of non-uniformity
can cause excessive sparking and a severe reduction in the useful current density in the ESP.

Step 4.3. Determine if performance is limited by non-ideal effects.


If poor performance cannot be traced to the electrical operating conditions, then the effects of
non-uniform gas velocity distribution, gas sneakage past electrified fields, and particle
reentrainment should be investigated. Certain measurements are required for this investigation.
The inlet gas velocity distribution should be measured by making velocity measurement
traverses over the face of the precipitator (see Section 5.7). The parameter gas is the
mathematical standard deviation of the matrix of velocity measurements, divided by the average
B-5

EPRI Licensed Material


Precipitator Performance Evaluation Using ESP Model

value. If the measured value of gas is greater than 0.25, then improving the gas velocity
distribution will likely improve the ESP performance. The gas flow above and below the
electrified fields should also be measured during the gas flow measurement. These data apply to
the following section.
Estimates of the degradation of precipitator performance caused by poor inlet gas velocity
distribution are illustrated in Figure 5-4. These curves for collecting efficiency vs. gas were
computed by the ESP model using the input data described in Table 2-1. The example curves
were computed for SCA values of 250 and 500 ft2/kacfm, corresponding to 3 and 6 sections in
the direction of gas flow. The gas sneakage fraction was set at 0.10. Each value of gas was set
the same throughout each field of the ESP. For older units with poor inlet ductwork or with
partially plugged diffuser plates, it is not unusual to find a gas values of 0.40 or more. At the
point on the curve for SCA 250 ft2/kacfm and fly ash resistivity 1 x 1011 -cm, the collection
efficiency is 98.6%. If the gas velocity distribution is improved to gas = 0.25 (efficiency
99.05%), the emissions would be reduced to 0.68 times the original value.
The fraction of gas bypassing the electrified fields should be determined from measurements of
the total gas flow and the gas flow above and below the electrified fields during the gas flow
distribution tests. If the sneakage fraction S is greater than 0.10, then the baffling to control the
gas sneakage through the non-electrified regions should be upgraded. The effect on ESP
performance by gas and fly ash particles passing above or below the electrified fields is similar
to the effect of non-rapping reentrainment. Both non-ideal effects are mathematically simulated
in the ESP model by an approximate correction to the precipitation rate parameter, using the
specified numerical value of the sneakage fraction S. Estimates of the degradation of
precipitator performance by both kinds of non-ideal effects are illustrated in Figure 5-5.
If the measured values of gas and S are not large enough to cause poor performance, or if
improvements in gas and S do not cause a match between measured and estimated efficiencies,
then particle reentrainment should be investigated. Collection efficiency measurements should
be made with and without the rappers in operation to determine the fraction of the total emissions
from rapping reentrainment. If rapping reentrainment accounts for more than 40% of the total
emissions, then efforts should be made to reduce the rapping reentrainment by changing the
rapping frequency or rapping intensity. (Note that ESP performance may take several days or
weeks to stabilize after changing the rapping in outlet fields.)
If rapping reentrainment is not a significant problem, then particle reentrainment without rapping
may be limiting the performance. Non-rapping reentrainment can result from high average gas
velocity, a very non-uniform gas velocity distribution, a low value of fly ash resistivity,
excessive sparking, low operating current densities, hopper ash buildup, air in-leakage through
the hoppers, or malfunction of the hopper emptying mechanism. The measured gas velocity
distribution and the measured gas flow below the electrified fields (above the hoppers) should be
analyzed to see if reentrainment is resulting from poor gas flow. The ash collection hoppers
should be checked to see if they are functioning properly. Sporadic reentrainment without
rapping can be detected by a continuous opacity monitor at the precipitator outlet.

B-6

EPRI Licensed Material


Precipitator Performance Evaluation Using ESP Model

As precipitator performance is improved with new design and new retrofit technology, the role
of the various non-ideal effects escalates in importance. Thus, Step 4.3 in the evaluation
procedure is of great significance in the upgrade studies for the higher efficiency designed ESPs.
The various non-ideal effects become the limiting factor on ESP performance as the efficiency is
increased.

Summary
In summary, the precipitator evaluation method described in this section can be a valuable aid in
diagnosing the causes of poor precipitator performance. Costly modifications to the precipitator
that will not result in a significant improvement in performance can be avoided. Again, a
precipitator should be placed in good mechanical condition and its performance should be
properly diagnosed before initiating flue gas conditioning or any other type of upgrade.

B-7

EPRI Licensed Material

C
ELECTROSTATIC PRECIPITATOR PRINCIPLES

C.1 General Process Description


Electrostatic precipitation differs from other gas cleaning technologies (e.g., fabric filters,
cyclones, and scrubbers) because the force to remove the particles is applied directly to the
particles rather than the entire gas stream. Thus, the energy required to accomplish the particle
removal is less in ESPs than in competing technologies.
The force to remove the particles results from the action of an electric field on an electrically
charged particle. The force acting on a charged particle is the product of the charge on the
particle and the value of the electric field in the vicinity of the particle.
Particles from industrial processes can acquire electrical charges in a number of ways. Particles
passing through a flame can be charged by flame ionization as electrons are removed from gas
molecules by the energy of the combustion process. Triboelectric charging results from particles
contacting the walls of the ductwork while being transported by the gas stream. (Even cosmic
rays and radioactive decay contribute to particle charge, although not to a significant degree.)
However, these mechanisms provide a very low level of charge on the particles. The electric
charge on particles to be collected in industrial ESPs primarily results from the intentional
formation of ions in a corona process. This provides a much higher level of charge on the
particles than would occur naturally, leading to more effective particle collection.
An electrical corona is an ionization process that involves an electric field strength strong enough
to accelerate free electrons to a velocity that will strip electrons from neutral gas molecules. The
corona forms a region of free electrons and a cloud of ions. Generation of a quasi-stable corona
requires a strong electric field in a localized region, surrounded by another region with lower
electric fields. In other words, there must exist a large gradient in electric field strength. This
type of electric field can be readily generated by applying a high voltage to an electrode system
consisting of a series of wires, or other structures with sharp points, suspended between parallel
plates or in large cylinders.
While the electric field is important for particle charging, it is also essential for particle
collection. An electric field is established by the presence of charged particles, ions, electrons,
or the high voltage applied to an electrode system. All of these factors contribute to the electric
field in an electrostatic precipitator.

C-1

EPRI Licensed Material


Electrostatic Precipitator Principles

C.1.1 Single-Stage vs. Two-Stage Designs


Electrostatic precipitators are classified as either single stage or two stage. The classification
depends on whether the charging and collection of the particles is accomplished in the same
region of the precipitator or in different stages. The two-stage unitused primarily for cleaning
air in occupied spacesconsists of a charging section with reasonably high values of corona
current density followed by a collection region with high electric fields and with low or
essentially no current flow in the collecting electrode system, depending upon the specific
design. Different electrode systems are used in the two stages to provide these distinct traits.
These types of devices are illustrated in Figure C-1.
The single-stage precipitator is almost exclusively used for industrial applications (including
power plants), although there are a few industrial units that are essentially two-stage devices
(e.g., United McGill provides a two-stage type of ESP for industrial applications). Single-stage
ESPs perform particle charging and collecting in the same physical region. The corona (or
discharge) electrodes are distributed throughout the particle control device with the collection
electric field established by the same electrode system that provides the charging ions.
The single-stage precipitation process is illustrated in Figure C-2. The sketch is based on a wirein-pipe configuration, in which the corona wire is suspended in the center of the cylindrical
collecting electrode by an insulator. A high voltage (usually negative polarity for industrial
units) is applied to the wire electrode. This electrode configuration results in a very high electric
field near the wire with the field decreasing with increasing radial distance toward the cylindrical
collecting electrode. This electric field distribution provides the functions necessary for the
electrical removal of solid or liquid particles from gas streams.
Whether one or two stages, the ESP must provide the following functions: (1) a corona system to
generate the ions for the electrical charge, (2) an electric field and/or thermal diffusion to drive
the ions from the corona process to the particles, (3) an electric field to cause the particles to be
removed from the gas stream and deposited on the collecting electrode, and (4) a means of
removing particles from the system. All of the factors in the precipitation process are discussed
in greater detail later in this appendix.

C-2

EPRI Licensed Material


Electrostatic Precipitator Principles

Two-Stage ESP

Single-Stage ESP
Figure C-1
Depiction of Single-Stage and Two-Stage Electrostatic Precipitators

C-3

EPRI Licensed Material


Electrostatic Precipitator Principles

Figure C-2
Schematic Example of Wire-in-Pipe Electrostatic Precipitator

C.1.2 Hot-Side vs. Cold-Side Designs


Most ESPs in the power generation industry are cold side units, located between the air heater
and the induced draft fan or stack. However, during the 1960s, several ESPs were installed in
utility boilers upstream from the air heaters, on the hot side of the air heater (and are thus
called hot-side units). The purpose for installing the ESP before the air heater was to take
advantage of the expected favorable ash electrical resistivity in the 700F (370C) temperature
range. As shown in Figure C-3, ash resistivity peaks between 300F (150C) and 400F (200C)
and decreases at both higher and lower temperatures.

C-4

EPRI Licensed Material


Electrostatic Precipitator Principles

Figure C-3
Example of Fly Ash Resistivity as a Function of Temperature

Selection of the hot-side units started with the increased utilization of low-sulfur coals, where the
electrical resistivity of the ash at the conventional temperature of 300F (150C) was higher than
that for the higher-sulfur coals customarily burned. In many instances, these hot-side units have
worked as expected, but there is a class of low-sodium coals that developed an increased ash
resistivity with operating timea phenomenon that has been associated with sodium depletion
in the collected ash layer. This characteristic, identified by Dr. Roy E. Bickelhaupt of Southern
Research Institute, is discussed in Appendix D on electrical conduction mechanisms in fly ash.

C-5

EPRI Licensed Material


Electrostatic Precipitator Principles

C.2 Theory of ESP Operation


A detailed understanding of the physics of ESP operation requires a considerable number of
mathematical equations to describe the spatial and time distribution of electric fields, particle
charging, and gas flow. However, it is not necessary to master the mathematics in order to have
a good working knowledge of an ESP. This appendix presents the general theory of ESP
behavior without going into the mathematics required to describe the behavior in detail. A
detailed discussion of the theoretical background can be found in other texts, notably Industrial
Electrostatic Precipitation by H. J. White, published by Addison-Wesley, Reading, MA, 1963
(now only available from The International Society for Electrostatic Precipitation); and
Electrostatic Precipitation by Sabert Oglesby and Grady B. Nichols, published by Marcel
Dekker, New York & Basil, 1978 (out of print).

C.2.1 Gaseous Conduction and the Corona Discharge


Electrical conduction in gases at conditions near standard temperature (32F; 0C) and pressure
(1 atm) is very small; accordingly, gases typically behave as very good insulators. Although
there are a few ions and free electrons formed in atmospheric gases by radioactive decay, highenergy photons, and cosmic rays, the number density of these carriers is very small and the
electrical resistivity of gases is normally very high.
In order for an ESP to operate efficiently, it is necessary to modify the electrical conditions in the
flue gas stream to provide for the generation of a significant quantity of ions to charge the fly ash
particles. These ions are formed from free electrons generated in the electrical corona that
occurs near the high-tension electrode in the ESP. The corona can exist in gases near standard
conditions when the applied electric field is on the order 20 kilovolts per centimeter. When the
electric field approaches this value, the free electrons formed naturally by the above-mentioned
mechanisms can be accelerated to a velocity adequate to cause electrical ionization of a
significant number of the neutral gas molecules in the flue gas stream through collision
processes. This ionization provides the electrons that later form negative ions in the gas stream
that charge the particles to be collected.
Electrical coronas can be established in ESPs with the corona electrode (i.e., the discharge
electrode, as opposed to the collecting plate) energized with either negative or positive polarity.
Negative corona (corona wire at negative potential with respect to the collecting electrode) is
used almost exclusively in industrial electrostatic precipitators. This is because negative corona
systems operate at higher voltages before sparkover than positive corona under flue gas
conditions in most industrial applications, including pulverized-coal-burning power stations.
However, under conditions of very high temperature and pressure, such as might be found in a
pressurized fluidized-bed combustion (PFBC) plant, positive corona provides better particle
collection. Positive corona is also used to clean the air in occupied living spaces, such as homes,
offices, and restaurants, because it generates less ozone than negative corona.
The physical appearance of positive and negative coronas is different. A negative corona has
individual tufts of corona distributed over the surface of the corona electrode, while a positive
corona appears as a steady glow covering the entire surface of the corona wire. The different
C-6

EPRI Licensed Material


Electrostatic Precipitator Principles

appearance is caused by the different directions of flow for the free electrons and positive ions
generated in the electrical breakdown process. In a negative corona, the electrons, which move
with much greater velocity than the more sluggish positive ions, are flowing toward the
collecting electrode. They must flow some distance before they form negative ions by attaching
to an electronegative gas molecule such as SO2, H2O, or oxygen. In a positive corona system,
conduction toward the collecting electrode is by the motion of positive ions formed in the corona
process. The positive ions provide a more ordered flow with a smooth transition from the corona
region to the quiescent zone in the electric field. The free electrons in positive corona flow
rapidly to the corona electrode without forming tufts.
There is one condition where negative corona appears as a smooth glow. When the ESP is
energized with very fast rise-time voltage pulses, the entire surface of the corona electrode is
brought well above the electric field strength for electrical breakdown, causing all the region of
space adjacent to the corona electrode to break down simultaneously. This fast rise-time
energization forces the negative corona to develop over the entire corona electrode surface,
yielding an appearance very similar to a positive corona.

C.2.2 Current and Voltage Relationships


Electrical current begins to flow in the ESP as soon as the applied voltage from the power supply
increases to the point that the electric field adjacent to the corona electrode is high enough to
ionize the surrounding flue gas. The actual field strength necessary to achieve this ionization is a
function of the flue gas composition, pressure, and temperature. The value of the electric field
near the corona wire is related to the applied voltage, the radius of curvature of the corona
electrode, and the distance between the corona and collecting electrodes. The electric field is
also influenced by the surface condition of the corona electrode. Very smooth wires break down
at a higher applied voltage than do rough wires. Irregularities on the corona electrode surface act
as stress raisers to increase the electric field in the vicinity of the irregularity. Therefore,
electrode systems with sharp points added to their surface will develop a corona at a lower
voltage, and will operate with a higher current than smooth, round wires. Typically, for coldside ESPs, electrodes with barbs will develop a corona at a voltage of 16 to 18 kV (12-inch plate
spacing). Standard-finish 0.109-inch diameter wires (a common size) will form a corona at 18 to
20 kV; polished, smooth wires of the same diameter will break down at around 28 kV. The
difference is totally attributable to the surface condition of the electrode system. Electrical
breakdown (corona formation) in hot-side units occurs at somewhat lower voltages because of
the reduced gas density, which increases the mean free path for the electrons. (Mean free path is
the average distance that an electron travels between collisions with gas molecules.)
In a negative corona system, the corona develops as a series of individual tufts along the surface
of the corona wire. As the applied voltage is increased, the electric field adjacent to the sharpest
point on the corona wire will reach breakdown strength. A flare develops and grows until a
cloud of negative ions develops in front of the tuft, electrically shielding the tuft from the applied
electric field. This space charge from the ions acts to stabilize the corona, and a steady-state
current is established.

C-7

EPRI Licensed Material


Electrostatic Precipitator Principles

As the applied voltage is further increased, the electric field near the next rough spot increases
until breakdown occurs there and proceeds as described above. The increased voltage
establishes a new equilibrium value of current from the first breakdown point as well. Increasing
the voltage from corona startup to the appropriate operating voltage, usually about 40 kV or
more, causes the corona electrode to develop tufts over its entire surface. An electrode system in
flue gas will reach an operating current density in the range of 40 to 50 microamperes of current
per square foot of collecting area (unless some other factor limits current). The numeric values
are about the same for metric units, as 1 A/ft2 equals 1.08 nA/cm2. Figure C-4 illustrates the
general shape of the voltage vs. current curve for the above sequence of events.

Figure C-4
Idealized Secondary Voltage vs. Current Curve

C-8

EPRI Licensed Material


Electrostatic Precipitator Principles

C.2.3 The Electric Field


The electric field is the primary factor that determines ESP collection efficiency. In particular,
the force that removes the particles from the flue gas stream results from the electric field
adjacent to the collecting electrode interacting with the electrical charge on the fly ash particles,
driving them toward the collecting electrode. The electrical charge on the ash particles is
proportional to the peak value of the electric field where the particle is charged (for particles
larger than 1 or 2 microns in diameter). The force acting on the particle is proportional to the
product of its electrical charge and the value of the electric field at the collecting electrode.
Thus, the electric field is very important to particle collection.
Several factors contribute to the establishment of an electric field in an ESP. Foremost is the
value of the voltage applied to the electrode system. The higher the voltage applied to an
electrode structure, the greater the resultant electric field. The presence of electrical charge also
establishes an electric field. A uniform charge between a parallel plate electrode system
establishes an electric field between those electrodes proportional to the distance between the
plates. Both the ions and charged fly ash particles contribute to the electric field in a utility ESP.
Figure C-5 illustrates the electric field associated with the applied voltage and the resultant field
modified by the presence of the charged particles and ions in the inter-electrode space. The
space charge from the charged particles is the primary reason that the operating voltages and
currents differ from the inlet through the outlet fields of an operating ESP. As the flue gas flows
through the ESP, charged fly ash particles are collected, reducing the space charge from those
particles. This is why inlet fields operate with higher voltages and lower currents than the
downstream fields.
The inter-electrode space (space between the corona and collecting electrodes) contains some
free electrons near the corona electrode, negative ions throughout the space, and charged
particles that have not yet been collected. The current flow in the inter-electrode space is the net
result of the migration of all these charges from the corona electrode toward the collecting
electrode. The electrical current at any point in the inter-electrode space is the sum of all
currents from the different charge carriers. The current from each charge carrier is the product of
the number (density) of the charge carriers, the electrical charge on those carriers, the electrical
mobility of the carrier, and the electric field (charged particles have a large number of
electronic charges on each particle). The mobility is a physical characteristic of individual
charge carriers and is defined as the ratio between the terminal velocity of the particle and the
applied electric field. The Deutsch migration velocity, a parameter used to characterize the rate
of collection of particles in an ESP, is similar to the mobility of the charged particle but the
numerical value is different for particles of different sizes.

C-9

EPRI Licensed Material


Electrostatic Precipitator Principles

Figure C-5
Idealized Electric Field vs. Radial Position, With and Without Current Flow (Pipe Diameter
= 8 inches, Wire Diameter = 0.109 inches)

The value of the electric field and the characteristics of the flue gas determine the velocity of
charged particles moving under the influence of that electric field. The electrical force on the
particle, ion, or electron causes the charge carrier to accelerate and increase in velocity. The
motion of the carriers through the gas stream causes a frictional retarding force caused by the
collisions between the charge carrier and the gas molecules in the gas stream. The carrier
velocity increases until this frictional drag force just balances the force from the electric field on
the charge carrier.
The electrical mobility of the different charge carriers varies. Free electrons are extremely small,
so the drag force from molecular collisions is small. The mobility of ions is on the order of 1000
times less than that of the free electrons, but much greater than that of the charged ash particles.
In most industrial gases and especially power station flue gases, there is an abundance of
electronegative gases (molecules that readily attach free electrons); thus the free electrons
flowing from the corona zone quickly become attached, forming ions. Therefore, free electrons
only contribute significantly to the flow of current in the very narrow region near the corona
electrode.

C-10

EPRI Licensed Material


Electrostatic Precipitator Principles

The charged particles are much more sluggish than the ions. Even though the total electrical
charge attached to fly ash particles may be approximately the same as, or perhaps greater than,
that which resides on the ions, the velocities of the ash particles are negligible relative to that of
ions. Therefore, the ions carry essentially all of the current in an operating utility ESP. The
charged particles do contribute significantly to the space charge between the electrodes, but
because of their very low electrical mobility, they carry a negligible amount of current.

C.3 Particle Charging


The corona process generates large numbers of ions in the inter-electrode region. These ions
move toward the collecting electrode approximately along electric field lines. Fly ash particles
introduced into the inter-electrode region receive their charge by attachment of these ions. Two
processes bring about the transport of these ions to the fly ash particles: field charging and
diffusion charging. In field charging, the ions flow to the particles along field lines, which
terminate on the surface of the particles. In diffusion charging, the random motion of the ions,
due to thermal energy, causes them to impact with particles and impart their charges to them.
In practical precipitators, both field and diffusion charging are significant. As mentioned
previously, large particles are charged principally by field charging, whereas diffusion charging
dominates for the very fine particles or fume (particles smaller than 0.1 micron in diameter).
Both mechanisms contribute significant charge to particles in the 0.1- to 2-micron diameter
range.

C.3.1 Field Charging


The sequence of events associated with field charging is suggested by the illustrations in Figure
C-6. In Figure C-6(a), an uncharged spherical particle with a dielectric constant significantly
greater than 1 is depicted in a uniform electric field. The presence of this particle causes a
distortion in the electric field in the vicinity of the particle as the electric field induces a dipole of
charge on the surface of the particle. As ions flow in the region, they tend to follow the electric
field lines or maximum field gradient, causing them to impact and be retained by the
particulate matter.
After a period of time, the ionic flow to the particulate causes a net charge to accumulate on the
particle. This accumulated charge produces an electric field that decreases as the reciprocal of
the square of the radius as suggested in Figure C-6(b). This self-field, added to the uniform
applied field, yields a resultant field as suggested in C-6(c).
Finally, the particle will acquire a saturation value of charge when the electric field from the
charged particle just balances the applied field. When this condition occurs, field charging
ceases. This situation is depicted in Figure C-6(d). Derivations for the spatial distribution of
charge and the charging rate have been covered by several authors (White and Oglesby &
Nichols, referenced earlier in this appendix). In these publications, an expression is developed
that relates the current flow to the particle with the free ion density, ion mobility, and applied
electric field. The expression indicates that as the particle acquires charge, the ionic flow rate to
C-11

EPRI Licensed Material


Electrostatic Precipitator Principles

the particle decreases with time at a rate diminishing to zero as the saturation value of charge is
approached. The charging rate (inversely related to the charging time constant) for particles is
proportional to the free ion density in the vicinity of the particles. The total charge on the
particles is proportional to the surface area of each particle. Thus, large particles will acquire
more charge than small ones, proportional to the square of the radius. Therefore, the electrical
force to remove large particles will be greater than for small ones. The charging rate (see above
on time constant) is usually such that large particles acquire a saturation charge after passing
only a few feet into an ESP.

Figure C-6
Particle Charging Sequence Depicted for Field Charging

C-12

EPRI Licensed Material


Electrostatic Precipitator Principles

C.3.2 Diffusion Charging


Diffusion charging differs from field charging in that the driving force for moving the ion to the
particle is the thermal energy that causes the ions and neutral gas molecules to move with
random velocities. The thermal energy in a gas stream causes the molecules to move and suffer
collisions between themselves and the walls of the vessel. These collisions between molecules
and ions impart velocity to the ions in the gas. The average velocity for all the molecules and
ions in the ESP is proportional to the square root of the absolute temperature. For these ions,
individual velocities range in value up to several orders of magnitude greater, as well as less,
than the root-mean-square (RMS) velocity. The velocity distribution is either approximately
Gaussian or perhaps log-normal. Thus, there is always a nonzero probability of finding an
individual ion with a velocity greater than any particular value of velocity. This means that there
is a nonzero probability of finding an ion with sufficient kinetic energy to overcome the energy
barrier caused by the field from a charged particle. This causes a continued flow of ions to
charged particles, the rate of which decreases with increasing charge but never reaches zero.
This mechanism is known as diffusion charging. The thermal velocity is in addition to any
velocity on the ion caused by the applied field. An applied electric field causes a bias on the
component of velocity in the direction of the field, which is often neglected in theoretical
descriptions of diffusion charging. In general, the charging rate is proportional to the absolute
temperature and the free ion density raised to some fractional power.
In contrast with the field charging mechanism, diffusion charging theory does not predict a
saturation value of charge. There is always a nonzero probability of finding an ion with
sufficient kinetic energy to overcome the repulsive force of the charged particle. However, a
saturation value of charge does exist even though it may be only of academic interest. This
saturation value is that amount of charge on the particle that establishes an electric field at the
surface of the particle high enough to cause the expulsion of an electron by field emissiona
condition that would not occur in industrial ESPs.

C.3.3 Combined Field and Diffusion Charging


Both field and diffusion charging are responsible for charging all the particles in an ESP.
However, for large particles, the total amount of charge from field charging is so great in
comparison to that from diffusion, that charge by diffusion can be neglected. Conversely, for the
very small particles, the saturation value of field charging is so small that only the diffusion
charging process need be considered. For the intermediate range of particles (i.e., those with
diameters of 0.1 to 2.0 m), both diffusion and field charging mechanisms must be considered.
Before the saturation value of field charging

C-13

EPRI Licensed Material


Electrostatic Precipitator Principles

Figure C-7
Comparison of Field Charging Rates for Half-Wave and Pure DC Electrical Energization

is attained, both mechanisms are active. In this case, the charging rates for both systems must be
added. After the saturation value for field charging is attained, only the diffusion charging
component need be considered. Thus, the charging behavior for this intermediate size range can
be described as augmented field charging initially, followed by diffusion charging. The charging
equations in the ESP models take this into account for the intermediate-size particles.

C.3.4 Practical Aspects of Charging


Particle charging theories have been developed for highly idealized conditions. The assumptions
in these theories include such factors as a uniform and constant electric field, spherical isolated
particles, and constant free ion densities. The actual situations encountered in practice include
heavy loading of uncharged particles into the inlet of a precipitator that is powered by either a
half-wave or full-wave rectified voltage waveform. The only filtering applied to this power
supply is from the distributed capacitance of the precipitator electrode system plus the resistive
C-14

EPRI Licensed Material


Electrostatic Precipitator Principles

and reactive components from the power supply and distribution system. Thus, the
instantaneous applied voltage and the free ion density change with time. Hence, the charging
rate changes with time, and in fact charging may be interrupted during portions of the voltage
cycle. The time-varying secondary voltage and an estimate of the charge vs. time for a range of
conditions are shown in Figure C-7.
A further complication is brought about by the introduction of large quantities of ash into the
precipitator inlet. The available charge is quickly bound to the particulate matter, which causes
an immediate reduction in the current. The highly mobile gas ions are quickly attached to the
relatively sluggish particles. Because the charging rate is proportional to free ion density, the
current quenching in the inlet section causes a decrease in the charging rate. For very low
current densities, such as those encountered in precipitators collecting high-resistivity ash or very
fine fume, charging times can be significantly long, resulting in a reduced collection efficiency.
If ash resistivity is low, charging times are usually short enough to be insignificant, often on the
order of a few milliseconds.

C.4 Particle Collection


Fundamental electrostatic theory states that the force acting on a charged particle under the
influence of an electric field depends on the amount of charge on the particle and the strength of
the electric field. The direction of the force depends upon the polarity of the particle charge and
the direction of the electric field.
Simplified concepts of particle collection often consider the electrostatic forces as dominant, and
that the motion of a particle toward the collecting electrode is governed primarily by the
electrostatic forces. However, in a full-size precipitator, the aerodynamic forces associated with
the highly turbulent gas flow are dominant until the charged particle approaches the collecting
electrode. This causes the motion of particles smaller than about 10 m in diameter to be almost
completely determined by the gas stream motion throughout most of the inter-electrode space.
Near the collecting electrode, where the influence of gas turbulence is reduced by the frictional
force of the collecting electrode, the electrostatic forces dominate. The collection of an
individual particle therefore depends on the probability that it will enter the region (boundary
layer) where the electrostatic forces result in its deposition on the collecting electrode surface.
Factors other than gas turbulence also influence particle collection. Once an ash layer is formed
on the collecting electrode, impingement of a particle being precipitated can cause those particles
previously collected to reentrain. Scouring of the ash layer, reentrainment of ash during rapping,
and unusual electrical conditions, including sparking and back corona, also alter the basic
collection process. These factors are neglected in describing the fundamental theory of particle
collection, in order to simplify the derivation of collection models and facilitate understanding of
the collection process.

C-15

EPRI Licensed Material


Electrostatic Precipitator Principles

C.4.1 Forces Acting on the Particles


The forces acting upon charged particles, other than the electrical force from the applied electric
field and the viscous drag force caused by the motion of the particle through the gas, can be
neglected. Throughout the inter-electrode space and particularly in the vicinity of the collecting
electrode, the electrical force accelerates the particle toward the collecting electrode, while the
viscous drag force of the gas opposes this motion. The final value of velocity attained will be
that which causes the viscous drag force to just balance the electrical driving force.
In the region adjacent to the collecting electrode, the frictional drag force from the collecting
electrode reduces the turbulence of the gas flow. The motion of the charged particle in this
reduced turbulent flow is a value on the same order of magnitude as that of the electrical velocity
for the intermediate and smaller particles. The trajectory of charged particles in this boundary
region is governed by the electrical velocity toward the collecting plate and by the frictionally
reduced gas velocity through the precipitator.
The velocity with which a charged particle is driven to the collecting electrode is related to the
charge on the particle, the dimensions of the particle, the value of the electric field driving the
particle, and the viscosity of the flue gas. The particle reaches a terminal velocity relative to the
gas where the viscous drag force from the gas just balances that from the applied field.

C.4.2 Particle Collection With Laminar Flow


For simplicity, first consider the highly idealized case of laminar flow. Then add the more
complex situation of turbulent flow, which exists in essentially all commercial ESP installations.
In the laminar flow case, consider a gas stream and entrained particles moving through the
precipitator at a uniform gas velocity, v. As described above, the particles are driven toward the
collecting plate with an electrical velocity, . The particle trajectory will then be the resultant
from the vector sum of these two velocities, as shown in Figure C-8. Also shown in the figure is
the development of an ash-free zone at the center of the precipitator duct that grows
progressively in size as the gas flows through the collector until point A, where all the particles
have been removed (i.e., 100% collection). This development assumes that the particles were
previously charged to saturation, no reentrainment of the collected ash occurs, and all particles
are of the same size, yielding identical migration velocities.

C-16

EPRI Licensed Material


Electrostatic Precipitator Principles

Figure C-8
Illustration of ESP Collection With Laminar Flow

Because the migration velocity for field charging is proportional to the particle radius, a
composite ash with a range of particle sizes would behave differently from one with a single
particle size. The distance required for 100% collection of small particles, say 1 m diameter,
would be 10 times as great as the distance for particles with a diameter of 10 microns. Thus, for
a particle size range that includes particles from the larger to the smaller, the ash concentration
would decrease as the precipitator was traversed, with the larger particles completely removed
first and progressively smaller ones collected subsequently. Thus, a concentration gradient, as
suggested in Figure C-9 would develop in the longitudinal as well as the transverse direction.
Research sponsored by DOE is currently being conducted (19961999) on the development of
laminar flow ESPs. The current interest is in developing a Laminar Flow Agglomerator that is
discussed in Volume 2 of this report.

C.4.3 Particle Collection With Turbulent Flow


The gas flow is turbulent rather than laminar in all practical industrial precipitators, with this
turbulent flow fundamentally determining the trajectory of the ash particles in most of the interelectrode space. Near the wall, the turbulent eddies are damped by the wall friction, forming a
boundary layer. Within this boundary region, the magnitude of the electrical velocity is the same
order of magnitude as the gas velocity. This is the region where precipitation actually takes
place, with the trajectory of the charged particle defined by the vector sum of the electrical and
gas flow velocity vectors.

C-17

EPRI Licensed Material


Electrostatic Precipitator Principles

Figure C-9
Illustration of the Development of a Particle Concentration Gradient for ESP Collecting a
Wide Particle Size Range With Laminar Flow

Within the inter-electrode space outside the boundary region, the magnitude of the electrical
migration velocity is small compared with the magnitude of the turbulent gas velocity. In this
region, the effect of the electrical velocity is to apply a small bias to the overall turbulent flow
toward the collecting electrode. The electrical velocity can be neglected in all but that region of
space immediately adjacent to the collecting electrode.
The consequence of this turbulent flow situation is that the collection efficiency, which could be
expressed as a linear function of the precipitator length for laminar flow, is modified to an
exponential relationship in the case of turbulent flow. In the distance where 100% collection
would occur with laminar flow, the concentration is reduced to [1 - 1/e] (~63%) for turbulent
flow, where e is the base of the natural logarithm system. The collection efficiency for the ideal
case where a given particle is moving with a given migration velocity is:
= 1 - exp{-(A /V)}
where is the efficiency, A is the collection surface area, V is the gas volume flow rate, and is
the migration velocity.
This exponential efficiency equation is known as the Deutsch-Anderson (D-A) equation.
Deutsch derived the exponential relationship, which was discovered experimentally by Anderson
in 1919, from theoretical considerations in 1922. Deutsch made several simplifying assumptions
in his original work. These include:
1. The particles are fully charged immediately on introduction into the collection system.
2. Turbulent and diffusive forces cause the particles to be distributed uniformly in any cross
section. (This assumption is more restrictive than necessary in actual practice.)
C-18

EPRI Licensed Material


Electrostatic Precipitator Principles

3. The velocity of the gas stream does not affect the particle migration velocity.
4. Particle motion is governed by viscous drag where Stokes law applies.
5. The particle always moves at its electrical terminal velocity relative to the gas stream.
6. Ash particles are separated enough for mutual repulsion to be neglected.
7. The effect of collisions between ions and gas molecules can be neglected.
8. There are no disturbing effects present such as erosion, reentrainment, uneven gas flow
distribution, or back corona.
Derivations of this equation are included in several texts (see Oglesby and Nichols, op. cit., and
H. E. Rose and A. J. Wood, Electrostatic Precipitation, London, Constable, 1956).
The D-A equation pertains specifically to the efficiency of the collection of particles with a given
migration velocity and hence of a given size. Because precipitators typically collect particles
with a wide range of sizes, no single equation is sufficient to describe the composite collection
efficiency for the system. A range of particle sizes will have a range of associated migration
velocities. The collection efficiency for the composite could be expressed as a summation of the
products of the ash load within an increment of size distribution with the collection efficiency
associated with that size interval, or by a definite integral for continuous particle size
distributions. Dr. Harry White performed this calculation for a log-normal size distribution in
Industrial Electrostatic Precipitation. Dividing the specific particle size distribution into a
number of small size increments and then determining the overall collection efficiency from a
summation of the individual incremental computations can make a very close approximation to
the actual performance. This technique forms the basis for computer models commonly used to
estimate ESP performance.

C.4.4 Factors Modifying Particle Collection


ESP collection efficiency, as described by the D-A equation, is based on the highly idealized
case described by the original assumptions. These idealized factors seldom exist in practice, and
are typically modified by the following factors:
Migration velocity. The particle charging time is seldom negligible, which results in the
particle traversing some distance through the precipitator before acquiring a saturation value of
charge. The net result is that the migration velocity is less than would be predicted based on the
assumption of instantaneous charging.
Non-uniform concentration of particles. The assumption of a uniform particle concentration
across the individual collecting lanes as well as the total ESP may be significantly in error. Tests
suggest that the electrical forces tend to cause an increase in the particle concentration in the
vicinity of the collecting plate. This factor should cause an increase in the collection efficiency
over that predicted from the original assumptions. The turbulent flow transports particles into
the boundary region faster than they are electrically removed.
C-19

EPRI Licensed Material


Electrostatic Precipitator Principles

Reentrainment. Reentrainment can occur from erosion by either electrical or mechanical forces
during the collection process, from reentrainment during plate rapping, or from a pickup by the
gas stream of material in the ash hoppers. All forms of reentrainment constitute a reduction in
the collection efficiency by reintroducing previously collected particles into the exit gas stream.
Uneven gas flow. Uneven gas flow reduces the collection efficiency in two ways. First, the DA equation shows that the collection efficiency is inversely related to the gas velocity or volume
flow rate. Second, the region experiencing the high gas velocity also carries a greater percentage
of the total suspended fly ash material.
Sneakage. The hopper regions that store the collected fly ash in a utility ESP provide a nonelectrified region for the fly ashladen gas stream to bypass a portion of the ESP collecting
zones. This portion of the total gas flow is re-mixed with the main gas flow at the end of the
hoppers. The portion of the gas that bypassed through the hopper was not subjected to collection
in that particular collecting field. Gas sneakage was not considered in the original idealized
equations.
Particle size distribution. The particular particle size distribution fed into the precipitator also
has an effect on collection efficiency. If a mechanical collector precedes the precipitator, the
larger particles will be removed ahead of the precipitator. In the absence of a mechanical
collector, the large particles, which are easier to collect, will still be present in the precipitator.
The effect of particle size distribution can be illustrated by an example from an actual ESP in
service, operating with and without a mechanical collector. The power plant is a pulverized-coal
boiler burning an average bituminous coal with an ash loading of 2 grains per actual cubic foot
of flue gas. The ESP is a four-field unit with a specific collecting electrode area of 320 square
feet per thousand cubic feet per minute of flue gas. The particle size distribution from the
furnace has an mmd (mass mean diameter) of 14.4 microns with a geometric standard deviation
(p) of 3.6. There is no performance limitation imposed by high resistivity.
Particle collection was modeled for two cases: (1) the ESP alone, with no mechanical collector
and (2) a mechanical collector with a D50 of 8.5 microns and a collecting efficiency of 66.35%
installed before the ESP. The performance model built into EPRIs ESPert software was used to
compare the two examples. The particle collection characteristics for the mechanical collector
were estimated from the EPA student training manual Control of Particulate Emissions (Draft
September 1978), Air Pollution Training Institute, Research Triangle Park, NC 27711. The
ESPs were modeled with a gas sneakage of 10% and a gas velocity standard deviation of 12%,
indicative of an ESP that has been in service for several years, but is still in good condition.
In the first case, the ESP alone is the particle collection device. The ESP alone will provide a
collection efficiency of 99.917%, yielding an outlet loading of 0.0044 pounds per million Btu,
which corresponds to 0.0017 grains per actual cubic foot of gas and an opacity of 2%. The
model data are summarized in Table C-1.
The mechanical collector installed in front of the ESP will remove about 63.5% of
the fly ash before the entrance of the ESP, providing for an inlet loading of 0.73
C-20

EPRI Licensed Material


Electrostatic Precipitator Principles
Table C-1
Performance Comparison of an ESP Alone vs. an ESP With a Cyclone
(see Appendix E for SI conversion factors)
Comparison With Rapping Reentrainment
Item

Units

ESP

ESP + Cyclone

ESP + Cyclone

Gas Velocity

ft/sec

Spec. Col Area (SCA)

ft /kcfm

320

320

320

99.917

99.759

99.919

cm/sec

11.297

9.606

11.316

Inlet Mass Loading

gr/acf

2.00

0.673

2.00

Outlet Loading to Stack

gr/acf

0.00166

0.00162

0.00162

Efficiency w/o Rapping

99.950

99.837

99.945

ESP Efficiency
Precip. Rate Parameter

Performance Comparison for Individual Size Bands


Particle

Inlet gr/acf

Cyclone

ESP Alone

ESP +

ESP Alone

ESP+Cyc.%

0.0875

0.00018

0.00018

4E-7

3E-7

99.766%

99.768%

0.145

0.00034

0.00034

2.1E-6

2.2E-6

99.368%

99.372%

0.205

0.00086

0.00086

7.2E-6

8.3E-6

99.163%

99.169%

0.285

0.00182

0.00182

2.06E-5

2.55E-5

98.868%

98.875%

0.465

0.00990

0.00985

0.0001240

1.64E-4

98.747%

98.755%

0.725

0.01408

0.01397

0.0001408

1.85E-4

99.000%

99.006%

1.025

0.02524

0.02494

0.0001725

2.14E-4

99.317%

99.321%

1.45

0.04294

0.04208

0.0001773

2.1E-4

99.587%

99.590%

2.05

0.06660

0.06227

0.0001462

1.45E-4

99.781%

99.782%

2.85

0.08824

0.07942

0.0001019

8.6E-5

99.885%

99.885%

3.85

0.10464

0.08790

0.0000676

4.7E-5

99.935%

99.936%

5.20

0.13972

0.10200

0.0000333

1.88E-5

99.976%

99.977%

7.25

0.18644

0.10441

0.0000030

1.3E-5

99.998%

99.998%

10.25

0.20624

0.08662

0.0000002

1.0E-7

99.999%

99.9999%

14.5

0.21638

0.05842

0.0000002

1.0E-7

99.9999%

99.9999%

20.5

0.20694

0.03104

0.0000002

1.0E-7

99.9999%

99.9999%

28.5

0.17276

0.01555

0.0000002

1.0E-7

99.9999%

99.9999%

51.0

0.29626

0.00889

0.0000002

1.0E-7

99.9999%

99.9999%

84.5

0.22042

0.00220

0.0000002

1.0e-71

99.9999%

99.9999%

Total

2.00

0.733

0.0010

0.0011

Note: Appropriate significant figures are exceeded in large particle ESP outlet load and in
efficiency/migration velocity, effective migration velocity, and precipitation rate parameter

grains per actual cubic foot of gas. The mechanical collector exhibits a 50% collecting
efficiency for particles with 8.5-micron diameter, with higher collecting efficiencies for large
particles and lower for small ones. The mechanical collector removes the large particles
C-21

EPRI Licensed Material


Electrostatic Precipitator Principles

preferentially, changing the inlet particle size distribution to the ESP to an mmd of 4.5 microns
and a p of 2.8. This change in particle size distribution causes the collecting efficiency of the
ESP to drop to 99.76%, with an outlet loading of 0.0043 pounds per million Btu or 0.0016 grains
per actual cubic foot, essentially the same as that without the mechanical collector.
This example indicates that the collecting efficiency of a good ESP alone will essentially match
that of a mechanical collector and ESP in series. This is because the mechanical collector
primarily removes the larger particles that will be easily collected in the ESP. The smaller
particles establish what material will actually exit the stack. (It should be noted, however, that
the ash loading in the inlet field hoppers of the ESP would be significantly reduced with the
addition of a mechanical collector). Table C-1 summarizes the results of this ESP model
comparison.
The term migration velocity refers specifically to the terminal velocity with which an individual
particle moves through the gas stream under the forces resulting from the action of the local
electric field on the charged ash particle and from the viscous drag of the flue gas stream. This is
the migration velocity as defined in the Deutsch-Anderson equation. This Deutsch migration
velocity represents the actual velocity for particles used in electrostatic precipitators. Other
expressions are used that can have the units of distance per unit time, but these are in fact
measures of performance, rather than a real velocity.
A generic or overall migration velocity is sometimes used to describe the performance of a
given ESP in a particular situation. However, in this usage, even though the term has the units of
velocity, the parameter should be identified differently from migration velocity because in
these instances the parameter is only a relative measure of performance. A more precise
identification for this parameter is either effective migration velocity or precipitation rate
parameter (PRP). This parameter, usually presented with the units of centimeter per second or
feet per minute, results from a simple calculation of the value of the parameter by the use of the
Deutsch-Anderson equation. As such it is at best a relative measure of performance. If the
collecting electrode area, gas volume flow rate, and collection efficiency are known, a
migration velocity can be calculated. However, in this usage, the migration velocity is not that
as defined in the original derivation of the D-A equation. Its value is not constant with respect to
changes in efficiency in any given ESP installation.
The previous example for a given ESP collecting fly ash from a pulverized-coal boiler both with
and without a mechanical collector illustrates this point. For the case where no mechanical
collector is installed, the collection efficiency is 99.917% with an SCA of 320 square feet per
thousand cubic feet per minute of flue gas, which provides a precipitation rate parameter (PRP)
of 22.24 feet per minute (11.3 cm/sec). When the mechanical collector is installed, the ESP
efficiency drops to 99.76% for a PRP of 18.9 ft/min (9.6 cm/sec). Note that the collection
efficiency for the ESP alone and the combination of the ESP and mechanical collector is
calculated to be the same (i.e., about 99.95%).
A second example employs an ESP model to compute the collection efficiency for the ESP
above with the gas velocity varying from 3 to 8 feet per second (0.9 to 2.4 meter/sec.). The
results of this exercise are summarized in Table C-2, including those from the previous example
C-22

EPRI Licensed Material


Electrostatic Precipitator Principles

with and without the mechanical collector. As the gas velocity is decreased, the collecting
efficiency for the ESP increases as indicated in the table. However, the increase in collecting
efficiency is primarily because of an increase in the collection of the smaller size fraction of the
particles, which have a lower Deutsch migration velocity. This increase in collection of the finer
size fraction results in a decrease in the PRP or effective migration velocity with increasing
SCA and efficiency. The migration velocities reported in Table C-2 neglect rapping
reentrainment. The migration velocities corrected for rapping reentrainment will decrease for the
higher gas velocities because of the greater amount of material collected in the outlet field, where
the major portion of the emitted reentrained material originates. There is also a small increase in
the charge on the particles because of the increased space charge from the uncollected particles
in the higher-velocity model runs.
Changing the gas velocity through the ESP while maintaining other factors constant establishes a
range of specific collecting areas (SCA) from 200 at 8 ft/s to 400 at 4 ft/s. The collecting
efficiency ranges from 99.985% at 4 ft/s to 99.65% at 8 ft/s. Note that the precipitation rate
parameter (PRP) increases from 11.21 cm/sec at the higher collecting efficiency up to 14.36
cm/sec at the lower collecting efficiency. This apparent contradiction results from the fact that at
high gas velocities, the percentage of large particles collected is high, and the larger particles
have a higher Deutsch migration velocity. However, note that the Deutsch migration velocity for
the ultrafine particles (smaller than about 1.5 micron in diameter) actually increases as the gas
velocity decreases. This increase is because the fine particles are primarily charged by diffusion
charging and the longer residence time allows these small particles to actually attain a greater
value of charge.
This contrasts sharply with the larger particles. The particles larger than 5 microns have
essentially the same migration velocity for the complete range of SCAs, as these particles reach a
saturation value of charge rather quickly for the electrical conditions used in the model. There is
a small increase for the larger particles as the space charge contributes to the saturation value of
charge for these particles.
The reason why the PRP decreases with increasing specific collecting area and efficiency is
somewhat more subtle. Compare the computed collection efficiency for particles 20.5 microns
in diameter with the efficiency for particles at 5.2 microns and 2.05 microns, over the range of
gas velocities.

C-23

EPRI Licensed Material


Electrostatic Precipitator Principles
Table C-2
Example Showing Actual Migration Velocity With Changes in SCA
(see Appendix E for SI conversion factors)
Gas Velocity (ft/s)

5 (w/cyc)

SCA

399

319

319

266

228

200

Res. Time

9 sec

7.2 sec

7.2 sec

6 sec

5.14 sec

4.5 sec

Efficiency

99.985%

99.952%

99.86%

99.890%

99.79%

99.65%

PRP cm/s

11.21

12.17

10.46

13.01

13.74

14.36

Particle Dia.

Mig. Vel.

Mig. Vel.

Mig. Vel.

Mig. Vel.

Mig. Vel.

Mig. Vel.

0.0875

10.114

9.925

9.933

9.755

9.596

9.456

0.145

8.636

8.451

8.459

8.283

8.130

7.994

0.205

8.235

8.076

8.084

7.934

7.800

7.679

0.285

7.764

7.640

7.647

7.517

7.419

7.322

0.465

7.670

7.581

7.587

7.508

7.433

7.365

0.725

8.155

8.115

8.121

8.061

7.997

7.934

1.025

8.822

8.857

8.863

8.884

8.815

8.772

1.45

9.727

9.849

9.855

9.883

9.905

9.887

2.05

10.476

11.012

11.017

11.162

11.247

11.271

2.85

11.689

12.181

12.186

12.475

12.624

12.748

3.85

12.352

13.132

13.117

13.584

13.903

14.127

5.20

15.048

14.942

14.946

15.527

16.071

16.430

7.25

20.394

20.107

20.070

19.974

19.791

20.673

10.25

41.850

41.686

41.694

41.382

27.213

27.048

14.50

57.735

57.513

57.353

57.150

56.665

56.075

20.50

80.103

79.487

79.502

78.526

77.796

77.837

C-24

EPRI Licensed Material


Electrostatic Precipitator Principles

Table C-3
Collecting Efficiencies for Selected Particle Sizes as a Function of SCA
(see Appendix E for SI conversion factors)
SCA

399

319

266

228

200

Gas Velocity

4 ft/sec

5 ft/sec

6 ft/sec

7 ft/sec

8 ft/sec

Part. Size

Collecting Efficiency for Each Particle Size Band (%)

0.0875

99.964

99.804

99.395

99.653

97.562

0.145

99.886

99.504

99.691

97.395

95.665

0.285

99.772

99.169

98.033

96.396

94.326

0.465

99.751

99.129

98.001

96.390

94.380

1.025

99.887

99.580

98.962

97.979

96.656

2.05

99.959

99.853

99.622

99.218

98.606

5.20

99.983

99.950

99.893

99.804

99.670

10.25

99.989

99.973

99.950

99.920

99.884

28.5

99.997

99.991

99.984

99.975

99.964

51.0

99.998

99.996

99.992

99.988

99.983

Overall Sizes

99.985

99.952

99.89

99.79

99.65

Even though the specific collecting electrode area varies from 200 to 400 square feet per
thousand cubic feet per minute in the example, the collecting efficiency for the large particles
remains about 99.99%which means the collecting efficiency for these particles remains
essentially constant. However, the collecting efficiency for the 5.2 micron particles varies from
99.98% to 99.67% whereas the efficiency for the 2.05 micron particles varies from 99.96% to
just over 98.61%. Clearly, the difference in collection over this very wide range in specific
collecting areas and efficiencies is the difference in the collection of the small particles. Because
these small particles have much lower migration velocities, the ensemble average for all particles
causes the PRP or effective migration velocity to decrease with increasing SCA and efficiency.
The efficiency data for selected particle size bands are summarized in Table C-3.

C.5 Collected Particle Removal


Ash deposited on the collecting plates must be removed to complete the collection process. In
the case of liquid aerosol collectors, such as tar separators, acid mist collectors, etc., the material
coalesces and drains from the collecting electrodes under the force of gravity. Solid materials in
dry ESPs, on the other hand, require an external method for removing the ash deposits. Wet
collectors use water or some other liquid to wash the plates, carrying the collected material with
C-25

EPRI Licensed Material


Electrostatic Precipitator Principles

it. Several types of wet removal systems are in current use. One type uses a flooded header with
either cylindrical or parallel plate collecting electrodes. The tops of the electrodes are designed
to form a weir with water flowing down the electrode to carry the collected material with it.
Other types of wet collectors use water sprays to atomize water into the precipitator section,
where the water droplets precipitate onto the collecting surface and subsequently drain, carrying
the collected ash with them. Figure C-10 illustrates two configurations of precipitators with wet
removal systems.
The principal advantages claimed for wet removal systems are (1) reentrainment losses are
eliminated or kept to a minimum, and (2) resistivity problems are eliminated. These factors are
in part related to the wetability of the ash. Other claims are made for wet systems, including
condensation on small particles for better collection, and space charge enhancement. The
principal difficulties with wet collection systems are avoiding dry spots on the collecting
electrode, internal corrosion and scaling problems, loss of plume buoyancy (possibly requiring
reheat), and problems with handling the slurry that results from wet collection. Wet ESP aftercollectors are under consideration for applications downstream of wet scrubbers in units burning
high-sulfur coals, where acid mist condensation may present an opacity problem.
Dry ash removal is accomplished by either periodic or continuous rapping of the collecting
electrodes. If the ash deposit on the electrodes is allowed to accumulate until an appropriate
layer thickness is collected, it will tend to fall into the hopper in sheets or large clumps when
rapped. The majority of precipitators collecting solid particulate matter are of the dry ash
removal type. The principal problem with dry ash removal is reentrainment of the collected ash,
which must be maintained at a minimum for good overall collection efficiency.
Ash reentrainment can occur as a result of several factors. If gas velocity is high, ash can be
scoured from the surface. However, studies indicate that for most fly ash particles, scouring is
not a serious problem at gas velocities below about 12 ft/sec. Large particles tend to be scoured
more easily than small ones, so some scouring of large particles may occur if the gas velocity
distribution is very non-uniform (creating regions of high velocity).
Reentrainment losses, however, are primarily associated with ash removal during plate rapping.
Minimizing rapping reentrainment requires careful attention to establishing the appropriate
rapping frequency and intensity. Rapping frequency is of importance in preventing too thick a
deposit from accumulating on the plate. Under these conditions, the ash layer might fall of its
own weight. When ash voluntarily falls from the plate, it can achieve a relatively high velocity
in free fall, with reentrainment losses becoming quite severe as the ash breaks up when falling
into the hopper. If, on the other hand, rapping is too frequent, the ash layer will be too thin and
rapping will tend

C-26

EPRI Licensed Material


Electrostatic Precipitator Principles

Figure C-10
Two Examples of Methods for Providing Wet Collecting Electrodes

to powder the ash, where it will be picked up and carried away by the gas stream. Figures 4-2
through 4-5 illustrate opacity meter traces indicating reentrainment.
For rapping systems that permit rapping intensity adjustment, tests should be conducted to ensure
optimum performance. Too soft a rap can fail to dislodge the ash at the proper time, so that
before the succeeding rap, the ash can fall freely because of the weight of the collected ash layer.
Too severe a rap can also cause poor performance, as the ash layer may be fragmented and
thrown into the gas stream where a significant amount of the fly ash may be reentrained.

C-27

EPRI Licensed Material


Electrostatic Precipitator Principles

Figure C-11 shows the relationship between rapping intensity and efficiency. The curve,
developed by Wayne Sproull in the late 1960s, shows that the optimum rapping intensity is
affected by ash resistivity.
Because the collection efficiency of a precipitator is exponential in nature, the rate of ash layer
buildup on the first sections will be much higher than that on subsequent ones. Consequently,
rapping cycles are usually varied between sections, the first sections being rapped several times
for each rap of the last section. For typical fly ash installations, the inlet fields should be rapped
in intervals from 5 to 15 minutes while outlet fields in 99+% collectors may only require rapping
every several hours.
Ash properties determine optimum rapping conditions. A low-resistivity ash is held to the plates
principally by mechanical and molecular adhesive and cohesive forces and is relatively easy to
dislodge. Strong electrical forces in addition to those mentioned for low-resistivity ash, on the
other hand, hold a high-resistivity ash. The rapping intensity required to dislodge the highresistivity ash can be quite high. Very-high-resistivity ash often cannot be removed by
conventional rapping, and power-off or reduced-power rapping is used. In ESPs with 5 or 6
fields, power-off rapping for a single field at a time could be used with very little loss in
collecting efficiency, if the gas velocity is sufficiently low. In an ESP with 3 or 4 fields, the
reentrained ash might not be recollected and a rather heavy discharge of ash would follow such a
power-off rap. As noted below, reentrainment losses for very-low-resistivity material cannot be
effectively controlled and may be significantly higher than for an ash in a more favorable
resistivity range.
Reentrainment losses can adversely affect precipitator performance if they are unacceptably
high. In Industrial Electrostatic Precipitation, White points out several methods for detecting
abnormally high reentrainment losses: (1) particle size analysis, (2) analysis of the charge on the
particles, and (3) an abrupt change in the precipitation rate parameter with gas velocity. Because
large particles are more easily reentrained, their presence at the precipitator exit indicates
reentrainment losses. Reentrained particles of intermediate- to low-resistivity tend to take on the
opposite electrical charge polarity due to the pith-ball effect, so that the appearance of large
numbers of oppositely charged particles suggest significant reentrainment. Plots of the

C-28

EPRI Licensed Material


Electrostatic Precipitator Principles

Figure C-11
Illustrative Relationship Between Collecting Efficiency and Rapping Intensity for Two
Values of Ash Resistivity

precipitation rate parameter as a function of gas velocity can also indicate reentrainment. A
sudden decrease in precipitation rate parameter as the gas velocity is increased signifies
reentrainment.
Reentrainment losses usually can be kept within acceptable limits if the rappers are properly
designed and adjusted. However, even with a well designed and operated rapping system, poor
gas flow quality or unfavorable ash properties can cause a severe degradation in ESP
performance.
Very-low-resistivity ash can be electrically reentrained. An electric field (E) is established in the
ash layer, by the electrical current flow through the layer and the resistivity of the layer. This
field can be described by the following equation:

C-29

EPRI Licensed Material


Electrostatic Precipitator Principles

E = j
where is the resistivity of the deposited ash layer (ohm-cm) and j is the current density
(amps/cm2). If the resistivity of the layer is very low, the electric field in the layer is less than
that in the inter-electrode space adjacent to the layer. Thus there is a discontinuity in the electric
field at the ash surface. A surface charge layer brings about this discontinuity or step function in
the electric field at the ash surface. In negative-polarity systems with the electric field in the
layer less than in the adjacent space, the surface charge is positive, which develops an electrical
force on the surface of the ash to reentrain the material. This phenomenon was identified in a
pilot-scale project where a spray dryer preceded an electrostatic precipitator. Attempts to operate
the system near the moisture dew point resulted in significant reentrainment. Observing the
surface of the ash layer, particles were reentrained when reducing the temperature to the
moisture dew point reduced the resistivity. Increasing the temperature a small amount could stop
the reentrainment process. (See Figure 6-3 and corresponding discussion in Section 6.8.2.)

C-30

EPRI Licensed Material

D
CONDUCTION MECHANISMS IN FLY ASH

Because the resistivity of fly ash has such a large effect on ESP performance, it is helpful to
understand the electrical conduction mechanisms in the fly ash material. The flow of corona
current begins with the generation of free electrons in the negative corona process used in most
industrial applications. These free electrons quickly attach to electronegative gases, thereby
forming negative ions. These ions either flow directly across the inter-electrode space or attach
to ash particles that are subsequently collected when they impinge on the previously collected
ash layer. The electrical current must then flow through the ash layer to reach the grounded
collecting electrode, and back to the power supply, completing the circuit. The current flow in
the ash layer could be either as negative charges flowing toward the collecting electrode or as
positive charges flowing in the opposite direction.

D.1 R&D History


The need to understand fly ash conduction mechanisms became evident only when lower-sulfur
coals came into general use. Low-sulfur coals frequently produce a fly ash with an electrical
resistivity high enough to severely limit the performance of electrostatic precipitators.
Prior to specific studies aimed at identifying the electrical conduction mechanisms in fly ash, it
was generally accepted that the conduction mechanism was electronic in naturei.e., a direct
flow of electrons through the collected ash layer. It was already well known that, for cold-side
units, conduction through the fly ash layer was the result of two conduction modes: volume and
surface conduction. As the names imply, volume conduction involves the migration of the
current-carrying species through the bulk of the material, while surface conduction occurs along
the surface of the material. Surface conduction is the result of the interaction of the constituents
of the flue gas with the material on the surface of the fly ash, forming a conduction path in
parallel with that through the bulk of the material.
The volume conduction mode usually dominates at temperatures greater than about 400F
(200C), while at temperatures well below 400F (200C), surface conduction becomes the
dominant mechanism. These modes are evident in Figure D-1, which plots the electrical
resistivity of an ash layer as a function of temperature. The linear portion of the plot above about
400F (200C) is where volume conduction dominates. At the lower temperature range, where
the plot deviates from a straight line, surface conduction becomes evident and eventually
dominates when the temperature becomes sufficiently low.

D-1

EPRI Licensed Material


Conduction Mechanisms in Fly Ash

Figure D-1
Resistivity vs. Temperature for Ash Used in Electrical Conduction Mechanisms Study

Prior to about 1972, volume conduction was thought to be electronic. Harry White in Industrial
Electrostatic Precipitation described it as such, with the fly ash particles considered to be
intrinsic semiconductors or insulators. The conduction was thought to be related to the freeing of
electrons in the conduction bands by thermal energy, yielding the linear relationship between
resistivity and reciprocal absolute temperature. This belief was not challenged until 1974, when
Dr. Roy Bickelhaupt conducted studies at the Southern Research Institute (SRI) to develop
definitive relationships to better understand the conduction of electricity in fly ash. Bickelhaupt
conjectured that the prior attribution to thermal energy evolved from the misconception that fly
ash was principally crystalline oxides, sulfates, and alumino-silicates, rather than the glassy
material that it was later found to be. White placed special emphasis on the linearity between
resistivity and reciprocal temperature (Arrhenius relationship) as an indication of semiconductor
behavior, although this same linearity applies to ionic conduction in solids.
D-2

EPRI Licensed Material


Conduction Mechanisms in Fly Ash

A misconception also existed about the surface conduction mode. Surface conduction was
considered to be ionic or electrolytic in nature, related to the physical or chemical adsorption of
water vapor or sulfuric acid on the surface of inert fly ash particles. The conduction was
considered to be through the adsorbed film on the fly ash particles. Surface conduction was
thought to be essentially independent of the chemical composition of the fly ash.
In 1968, Shale, at the Bureau of Mines (now DOE) in Morgantown, West Virginia, noted that
even though volume conduction was thought to be electronic in nature, higher concentrations of
alkali metals led to reduced resistivities at higher temperatures. He suggested that composition
may also influence the resistivity of fly ash at lower temperatures. Bucher, in his dissertation
from the University of North Dakota, also noted a reduction in the resistivity of fly ash from
lignitic coal with higher sodium content. He concluded that the sodium was principally related
to the volume conduction regime, but did not consider sodium to actually be a charge carrier.
Bucher referred to Shales work and to Duries observation that the ESPs from the Liddell Power
Station in England experienced improved performance when collecting fly ash with higher
sodium content.
The first formal study of the electrical conduction mechanisms in fly ash was performed at SRI
for Transalta Utilities (at that time, Calgary Power, Ltd.), with Bickelhaupt and Nichols working
jointly on this project. Bickelhaupt, a ceramics engineer, recognized the similarity between fly
ash and glass, and designed and conducted experiments that proved conclusively that the
principal electrical charge carrier for volume conduction in fly ash is the actual migration of
sodium ions in the fly ash layer. These experiments suggested that the conduction was ionic
rather than electronic as previously assumed.
After laboratory work identified sodium migration as the conduction mechanism, sodium salts
were added to the feed coal at Transalta Utilities Wabamun Station. This material immediately
reduced the resistivity of the fly ash from the stations low-sulfur coal when operating at 340F
(170C), a temperature where surface conduction would be expected to dominate.
In 1973, Bickelhaupt and Nichols suggested the use of sodium conditioning as a means for
combating high-resistivity problems in ESPs. Holyoak also discussed the improvement in ESP
performance obtained when burning coal with high sodium contents in Burning Western Coals
in Northern Illinois at the November 1973 Winter Annual Meeting of the American Society of
Mechanical Engineers, referring to Bickelhaupt and Nichols work for an explanation. White
presented an extensive discussion of resistivity problems in the April 1974 Journal of the Air
Pollution Control Association. He mentioned the results of the SRI work at Transalta Utilities
and formally presented the results of this first sodium conditioning experiment conducted by
Bickelhaupt and Nichols.
Based on these observations and experiments, and subsequent research at SRI, Bickelhaupt
formulated theories for both volume and surface conduction mechanisms in amorphous fly ash
material. Fly ash was shown to be glass-like with the conduction primarily ionic in nature. The
explicit relationship between resistivity and alkali metals, principally lithium and sodium, was
verified. A phenomenological description of the conduction mechanisms was given by
paraphrasing the fundamental research on silicate glasses. The numerical data from chemical
D-3

EPRI Licensed Material


Conduction Mechanisms in Fly Ash

transference experiments and photographs of the sodium were offered as proof of the nature of
this conduction mechanism.

D.2 Current Theories of Electrical Conduction in Fly Ash


The current theories of conduction in fly ash are summarized below for both volume and surface
conduction. In volume conduction, the presence or absence of sulfur trioxide (which joins with
water vapor to form sulfuric acid) in the gas stream is immaterial. There are two regimes of
conduction in surface conduction, both of which depend upon the amount of sulfuric acid vapor
present and the temperature.

D.2.1 Volume Conduction


For temperatures above 575F (300C), the logarithm of the resistivity plotted against the
reciprocal of absolute temperature is linear for the volume conduction mode. The fly ash layer
residing on the collecting electrodes (for pulverized-coal boilers) consists principally of spherical
particles, with the surface area and porosity dependent on the particle size distribution and
packing density of the fly ash. (The combustion temperature in pulverized-coal boilers is usually
high enough to melt the ash particles so they coalesce into spherical particles.) The fly ash
material is amorphous in nature with perhaps small amounts of crystalline and combustible
material. Alkali metal ions, primarily lithium and sodium, are the active charge carriers, which,
when released by thermal energy, are freed from the amorphous structure so that the applied
electric field drives them from the anode to the cathode. The oxygen ions associated with those
metal ions are expected to drift toward the anode. Because the majority of industrial electrostatic
precipitators operate with negative polarity on the corona (or discharge) electrode, the corona
wire is the cathode while the collecting electrode is the anode. Thus, the alkali metal ions drift
from the collecting electrode toward the surface of the ash layer where they acquire negative ions
generated by the corona. The oxygen ions drift to the collecting electrode and convert to O2
molecules by giving up an electron to the collecting electrode. The charge-carrying alkali metal
ions physically move through (or over the surface of) the particles that form the ash layer.
The magnitude of the electrical resistivity of the ash layer in the volume conduction regime is
controlled by several factors, including:

Temperature of the layer. Higher temperatures contribute thermal energy to the glassy
material freeing a greater number of ions to participate in the conduction process as well as
facilitating their ability to migrate through the fly ash. This factor causes the decrease in
electrical resistivity of the fly ash with increasing temperature.

Alkali metal ion concentrations. Since the sodium ion is the principal charge carrier
(lithium concentrations tend to be negligibly small), the greater their concentration, the lower
the resistivity of the ash.

Factors that influence the glassy structure of the amorphous material in the ash. The
iron concentration in the ash, combustion zone characteristics (temperature, residence time,
coal particle size, etc.), and the cooling rate of the particles leaving the combustion zone also

D-4

EPRI Licensed Material


Conduction Mechanisms in Fly Ash

influence the amorphous structure of the ash and perhaps modify the number and migration
rates of the alkali metal charge carrier ions and the resulting resistivity.

Electric field strength. The resistivity of fly ash usually decreases with increasing electric
field strength. The electric field distorts the glass structure, reducing the energy barriers at
the charge carrier sites, which allows ions to be released and move more freely through the
layer. This reduction in the energy barrier causes a decrease in resistivity.

Particle size distribution. The finer the particle distribution, the lower the resistivity, as
there are a greater number of contact points between the particles. The greater the number of
conduction paths, the lower the resistivity.

Porosity of the ash layer. The greater the porosity, the less material is present in any cross
section of the conduction path. The resistivity of the fly ash material is less than the gas in
the interstitial regions of the deposit, causing the resistivity to increase with increasing
porosity. Typical fly ashes exhibit porosities in the range of 55% to 70% by volume. The
change in specific resistivity change caused by this range of porosity is small.

Fly ash resistivity was at one time considered low enough above 575F (300C) to pose no
limitation on ESP performance. However, this has proven not to be the case for certain ashes
that are low in sodium content. The resistivity of newly formed fly ash is usually in the mid-1010
ohm-cm range, a value that produces satisfactory ESP operation. However, a residual layer of
fly ash remains on the collecting electrodes, which is not removed by rapping. A continued flow
of electrical current through this layer, which is in fact a physical migration of sodium ions,
effectively removes a portion of the sodium ions in that residual layer. This depletion of charge
carriers in the residual layer causes an increase in the layers resistivity, so that a high-resistivity
behavior pattern develops over time. In cases where the sodium concentration in the ash layer is
sufficiently high, this sodium depletion phenomenon does not lead to degradation in ESP
performance. However, some fly ashes low in sodium concentration develop symptoms of highresistivity problems in periods from a few weeks to a few months.
The depletion of charge carriers adjacent to the collecting electrodes causes a carrier
concentration gradient to develop in the ash layer. The newly deposited layer has a higher
concentration of sodium than the depleted layer adjacent to the collecting electrodes. Thermal
energy tends to drive the sodium ions from the higher-concentration region into the lowerconcentration region by diffusion, while electrical migration tends to drive them from the lower
concentration region toward the surface of the ash. These competing mechanisms work toward
establishing an equilibrium condition, where one flow just matches the other. If the layer
adjacent to the plate develops a sufficiently high resistivity before this equilibrium is established,
the ESP performance becomes limited either by sparking at reduced voltages or by back corona.
The high-resistivity problem described above can be alleviated either by thoroughly cleaning the
collecting electrodes (i.e., water washing or blasting with sand or grain) or by injecting a
conditioning agent to modify the ash resistivity. Conditioning agents such as sodium carbonate
or sodium sulfate, as well as proprietary agents, are potential candidates.
The volume mode of conduction is active over the entire temperature range and is represented by
a straight line in a plot of log resistivity vs. reciprocal absolute temperature. Thus, at lower

D-5

EPRI Licensed Material


Conduction Mechanisms in Fly Ash

temperatures, the volume conduction becomes very small, leading to a very high resistivity from
volume conduction alone.

D.2.2 Surface Conduction


At temperatures below about 400F (200C), conduction becomes increasingly affected by
surface conductionthe transport of electrical charge carriers over the surface of the ash
particles. This mode of conduction depends on the concentrations of water and sulfuric acid (or
its precursor, SO3) in the flue gas stream. If the concentration of these constituents is high
enough, it is probable that hydrogen ions participate directly in the conduction process. For
high-SO3 conditions, electrical transference experiments do not indicate a physical migration of
fly ash constituents as occurs in volume conduction.
When the concentration of sulfuric acid is low, the moisture and dilute sulfuric acid are
considered to be adsorbed to the surface of the ash particles. The adsorbed material attacks the
surface, freeing alkali metal ions to move for electrical conduction. The reaction of the acid with
the ash surface produces sodium hydroxide, which in turn attracts more water vapor until a
surface layer is developed where the sodium ions become quite mobile. This concept has been
well documented with respect to silicate glasses and water vapor, but the actual role in fly ash
with high levels of sulfuric acid has not been conclusively determined. Surface conduction
through the electrolysis of sulfuric acid also cannot be completely ruled out. However,
transference experiments with fly ash and small quantities of sulfuric acid show that sodium ions
physically migrate.
Factors that control surface resistivity include the following:

Temperature

Specific surface area

Ash chemical composition

Electric field strength

Water vapor and sulfuric acid concentrations in the flue gas stream

With respect to ash composition, high alkali metal concentrations reduce resistivity. Also, high
iron concentrations may reduce the chemical durability of the ash, making it more susceptible
to attack by acid and water vapor, freeing larger numbers of ions to move under the influence of
the electric field. Alternatively, higher concentrations of silica and alkaline earth materials may
increase the chemical durability, causing an increase in the ash surface resistivity.
Chemical transference experiments for fly ash in a standard resistivity cell in the presence of 9
ppm of SO3 indicated no significant migration of sodium ions, as was apparent with water vapor
and lower values of SO3. With 38 coulombs of electricity transmitted through the cell, only
about 6% of the sodium ions migrated. This suggests that only volume conduction contributed to
sodium migration, while surface conduction did not. During this transference experiment, the
current remained constant for a fixed operating voltage across the cell, indicating that the
D-6

EPRI Licensed Material


Conduction Mechanisms in Fly Ash

resistivity remained constant during the experiment. However, another experiment, which also
started with 9 ppm of SO3, showed an immediate increase in resistivity with time when the acid
was removed. Figure D-2 shows the results of this experiment.
These experiments support the view that surface conduction results from two possible
mechanisms. For very low concentrations of sulfuric acid, the mechanism is the migration of
alkali metal ions, similar to that in volume conduction but enhanced by the water vapor and trace
acid attack of the ash surface. (It is possible that the sodium depletion phenomenon may occur
for this mechanism.) For higher concentrations of sulfuric acid, hydrogen ions may directly
participate in the conduction process with little or no physical migration of carrier ions from the
fly ash structure. There is also the possibility that there are acid concentrations where both
surface conduction mechanisms are active.

Figure D-2
Resistivity vs. Time for Experiment With 9 ppm of Sulfur Trioxide Injected Into Resistivity
Cell for a Period of Time and Then Turned Off

D-7

EPRI Licensed Material

E
SI AND U.S. UNIT CONVERSION FACTORS
Table E-1
Unit Conversion Factors
To Obtain
actual cubic feet/minute (acfm)
Btu/kWh
Btu/lb
centimeters (cm)
degrees Celsius (C)
3
cubic meters/second (m /s)
degrees Farenheit (F)
feet/second (ft/s)
grains/actual cubic foot (gr/acf)
3
grams/cubic meter (g/m )
inches (in)
inches (in)
kJ/kg
kJ/kWh
meters (m)
meters/second (m/s)
millimeters (mm)
millimeters (mm)
3
normal cubic meters/second (nm /s)
2
square centimeters (cm )
2
square feet (ft )
2
square feet (ft )
square feet per thousand actual
2
cubic feet/minute (ft /kacfm)
2
square meters (m )
square meters per cubic
2
3
meter/second (m -s/m )
standard cubic feet/minute (scfm)
tons (metric)
tons (short)

Multiply
3
cubic meters/second (m /s)
kJ/kWh
kJ/kg
inches (in)
degrees Farenheit (F)
actual cubic feet/minute (acfm)
degrees Celsius (C)
meters/second (m/s)
3
grams/cubic meter (g/m )
grains/actual cubic foot (gr/acf)
centimeters (cm)
millimeters (mm)
Btu/lb
Btu/kWh
feet (ft)
feet/second (ft/s)
inches (in)
mils
standard cubic feet/minute (scfm)
2
square feet (ft )
2
square centimeters (cm )
2
square meters (m )
square meters per cubic
2
3
meter/second (m -s/m )
2
square feet (ft )
square feet per thousand actual
2
cubic feet/minute (ft /kacfm)
3
normal cubic meters/second (nm /s)
tons (short)
tons (metric)

By
2120
0.948
0.43
2.540
9/5*
0.000472
5/9*
3.2808
0.437
2.288
0.3937
0.03937
2.325
1.055
0.3048
0.3048
25.4
0.02540
0.000438
929.0
0.00108
10.764
5.077
0.0929
0.197
2282
0.9072
1.1023

*For converting differential temperatures only. To convert actual temperatures,


F = C x 9/5 + 32; C = (F - 32) x 5/9.

E-1

Das könnte Ihnen auch gefallen