Sunteți pe pagina 1din 374

DISTRICT

Complete Design Guide for District Heating Systems

HEATING GUIDE

District Heating Guide provides design guidance for all major aspects of district heating systems,
including central heating plants, distribution systems, and consumer interconnections, for both
steam and hot-water systems. It draws on the expertise of an extremely diverse international
team with current involvement in the industry and hundreds of years of combined experience.
In addition to design guidance, this book also includes a chapter dedicated to planning, with
information on cogeneration of heat and electric power as well as the integration of thermal
storage into a district heating system. Guidance on operations and maintenance is also provided
to help operators ensure that systems function as intended. In addition, examples are presented
where appropriate, and three detailed case studies are included in an appendix.
This guide is a useful resource for both the inexperienced designer as well as those immersed
in the industry, such as consulting engineers with campus specialization, utility engineers, district
heating system operating engineers, central plant design engineers, and steam and hot-water
system designers. District Heating Guide fills a worldwide need for modern and complete design
guidance for district heating systems.

ISBN 978-1-936504-43-5

ASHRAE
1791 Tullie Circle
Atlanta, GA 30329-2305
404-636-8400 (worldwide)
www.ashrae.org

ASHRAE_DH_Guide_On-Template.indd 1

DISTRICT HEATING GUIDE

Comprehensive Reference
Planning & System Selection Central Plants Distribution Systems
Heat Transfer Calculations System O&M Consumer Interconnection

9 781936 50443 5

Product code: 90562

8/13

7/30/2013 9:58:51 AM

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District
Heating
Guide

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

This publication was developed as a result of ASHRAE Research Project RP-1267 under the auspices of
ASHRAE Technical Committee 6.2, District Energy. It is not a consensus document.

CONTRIBUTORS
Gary Phetteplace, PhD, PE (Principal Investigator)
(Chapters 1, 2, 4, 5, 6, 8; Appendices B, C)
GWA Research LLC
Lyme, NH
Donald Bahnfleth, PE
(Chapter 2)
Bahnfleth Group Advisors LLC
Cincinnati, OH

Vernon Meyer, PE
(Chapters 4, 6)
Heat Distribution Solutions
Omaha, NE

Peter Mildenstein
(Chapter 7)
Rambll
DK-2300 Kbenhavn S

Ishai Oliker, PhD, PE


(Chapters 2, 3, 8; Appendix C)
Joseph Technologies Corp. Inc.
Montvale, NJ

Jens Overgaard
(Chapter 7)
Rambll
DK-2300 Kbenhavn S

Pernille M. Overbye
(Chapter 7)
Rambll
DK-2300 Kbenhavn S

Kevin Rafferty, PE
(Chapter 2)
Modoc Point Engineering
Klamath Falls, OR

Steve Tredinnick, PE, CEM


(volunteer contributor)
(Chapters 2, 5)
Syska Hennessy Group, Inc.
Verona, WI

David W. Wade, PE
(volunteer contributor)
(Appendix C)
RDA Engineering, Inc.
Marietta, GA

PROJECT MONITORING COMMITTEE


Steve Tredinnick, PE, CEM (Chair)
Syska Hennessy Group, Inc.
Verona, WI
Moustapha Assayed
Empower Energy Solutions
Dubai, United Arab Emirates

Lucas Hyman, PE
Goss Engineering, Inc.
Corona, CA

Samer Khoudeir
Empower Energy Solutions
Dubai, United Arab Emirates

Victor Penar, PE
Hanover Park, IL

David W. Wade, PE
RDA Engineering, Inc.
Marietta, GA

Updates/errata for this publication will be posted on the


ASHRAE website at www.ashrae.org/publicationupdates.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

RP-1267

District
Heating
Guide

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

ISBN 978-1-936504-43-5
2013 ASHRAE
1791 Tullie Circle, NE
Atlanta, GA 30329
www.ashrae.org
All rights reserved.
Cover Design by Laura Haass
ASHRAE is a registered trademark in the U.S. Patent and Trademark Office, owned by the
American Society of Heating, Refrigerating, and Air-Conditioning Engineers, Inc.
ASHRAE has compiled this publication with care, but ASHRAE has not investigated, and
ASHRAE expressly disclaims any duty to investigate, any product, service, process, procedure, design, or the like that may be described herein. The appearance of any technical data or
editorial material in this publication does not constitute endorsement, warranty, or guaranty by
ASHRAE of any product, service, process, procedure, design, or the like. ASHRAE does not
warrant that the information in the publication is free of errors, and ASHRAE does not necessarily agree with any statement or opinion in this publication. The entire risk of the use of any
information in this publication is assumed by the user.
No part of this publication may be reproduced without permission in writing from ASHRAE,
except by a reviewer who may quote brief passages or reproduce illustrations in a review with
appropriate credit, nor may any part of this publication be reproduced, stored in a retrieval system, or transmitted in any way or by any meanselectronic, photocopying, recording, or
otherwithout permission in writing from ASHRAE. Requests for permission should be submitted at www.ashrae.org/permissions.
Library of Congress Cataloging-in-Publication Data
District heating guide.
pages cm
Includes bibliographical references.
Summary: "Guidance for district heating system planning, design, operation, and maintenance for inexperienced designers
and complete reference for those immersed in district heating industry; includes terminology for district heating"-- Provided
by publisher.
ISBN 978-1-936504-43-5 (softcover)
1. Heating from central stations--Handbooks, manuals, etc. I. American Society of Heating, Refrigerating and AirConditioning Engineers.
TH7641.D4724 2013
697'.03--dc23
2013012052

ASHRAE STAFF SPECIAL PUBLICATIONS

Mark S. Owen, Editor/Group Manager of Handbook and Special Publications


Cindy Sheffield Michaels, Managing Editor
James Madison Walker, Associate Editor
Roberta Hirschbuehler, Assistant Editor
Sarah Boyle, Assistant Editor
Michshell Phillips, Editorial Coordinator

PUBLISHING SERVICES

David Soltis, Group Manager of Publishing Services and Electronic Communications


Jayne Jackson, Publication Traffic Administrator
Tracy Becker, Graphics Specialist

PUBLISHER

W. Stephen Comstock

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Contents

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv
Acronyms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

Chapter 1 Introduction
Purpose and Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1
District Heating Background. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1
Applicability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2
Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2
Benefits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3
Environmental Benefits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1.3
Economic Benefits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1.4
Initial Capital Investment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.5
Concept Planning. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1.5
Design. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1.5
Construction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1.5
Consumer Interconnection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1.6
Typical Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.6
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.6

Chapter 2 Planning and System Selection


Planning Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1
Establish and Clarify the Owners Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3
Development of the Database. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4
Sources of Data for Existing Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.4
Heating Load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.4
Demand Diversity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.5
Heat Load Density and Piping Costs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.5
Alternative Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.6
Codes and Standards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.6
Local and Institutional Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.7
Choice of Heating Medium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.7
Integrated Processes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.10

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Central Plant Siting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.10


Heat Distribution Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.11
Construction Considerations and Cost . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.11
Consumer Interconnection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.12
Energy Costs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.13
Operation and Maintenance Costs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.13
Economic Analysis and User Rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.14
Master Planning Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.14
Alternative Development for Heat Supply. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.16
Methods of Heat Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.16
Conventional Heat-Only Boiler Plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.16
District Heat Supply from Cogeneration Steam-Turbine-Based Stations. . . . . . . . . . . . . . . . . . .2.19
Retrofit of Single-Purpose Electric Generating Steam Turbine to District Heat Supply. . . 2.21
Overview. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.21
Configuration and Control of Steam Turbine Retrofits for District Heating . . . . . . . . . . . . . . . . .2.22
Cycle Efficiencies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.26
Examples of Power Plants Retrofitted to Cogeneration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.27
District Heating/Cogeneration from Stationary Gas Turbines . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.34
Reciprocating-Engine-Based Cogeneration/District Heating Systems . . . . . . . . . . . . . . . . . . . .2.37
Large Heat Pumps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.37
Integration of Heating, Cooling, and Electric Generation . . . . . . . . . . . . . . . . . . . . . . . . . . 2.39
Centralized Chilled-Water Generation by Thermal Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.40
Centralized Chilled-Water Generation by Electric Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.41
Decentralized Chilled-Water Generation by Thermal Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.43
Geothermal District HeatingDirect Use . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.43
U.S. Experience . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.44
European Experience. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2.46
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.46
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.48

Chapter 3 Central Plant Design for Steam and Hot Water


Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1
Higher Thermal Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.1
Use of Multiple Fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.2
Environmental Benefits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.2
Operating Personnel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.2
Insurance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.2
Usable Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.2
Equipment Maintenance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.2
Use of Cogeneration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.2
Central Plant Advantages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3
Central Plant Disadvantages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3
Heating Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4
Central Plant Heating Medium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4
Heat Capacity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.5
Pipe Sizes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.5
Condensate Return System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.5
Pressure and Temperature Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.5
Boilers Pressure and Temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6

vi

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Contents

Construction Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6


Selection Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6
Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.7
Combustion Efficiency. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.7
Overall (or Thermal) Efficiency. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.7
Seasonal Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.7
Performance Codes and Standards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.8
Commercial Heating Boilers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.8
Central Plant Design for Steam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.8
Typical System Arrangements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.8
Selection Criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.10
Construction Cost Estimate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.11
Feasibility Analysis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.12
Environmental Regulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.12
Water Supply . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.12
Site Development. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.13
Plant Access. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.14
Plant Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.15
Heating, Ventilating and Air-Conditioning (HVAC) Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.17
Drainage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.18
Plant Safety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.18
Central Plant Security . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.19
Steam Generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.19
Boiler Design and Type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.20
Boiler Construction Options . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.20
Available Fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.21
Combustion Technology Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.21
Burners . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.23
Boiler Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.24
Primary Air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.25
Economizer. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.25
Water Treatment and Makeup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.25
Sound and Vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.26
Seismic Issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.26
Breeching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.26
Plant Stack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.27
Fuel Train . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.28
Oil Supply System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.28
Piping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.29
Combustion Air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.29
Maintenance and Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.30
Commissioning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.30
Central Plant Design for Medium- and High-Temperature Water . . . . . . . . . . . . . . . . . . . . 3.31
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.31
HTW Plant Arrangement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.32
Basic System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.33
Design Considerations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.33
Direct-Fired HTW Generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.34
Direct-Contact Heaters (Cascades) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.41
System Circulating Pumps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.41

vii

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Controls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.44
Water Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.46
Heat Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.46
Safety Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.46
Other Design Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.46
Central Plant Design for Low-Temperature Water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.46
Typical System Arrangements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.47
Energy Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.49
System Temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.50
Heat Exchangers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.52
Thermal Storage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.52
Auxiliaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.52
Expansion Tanks and System Pressurization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.52
Pumping System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.55
Pump Curves and Water Temperature for Constant-Speed Systems . . . . . . . . . . . . . . . . . . . . .3.55
Parallel Pumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.57
Series Pumping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.59
Multiple-Pump Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.59
Standby Pump Provision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.59
Variable-Speed Pumping Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.60
Pump Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.61
Flow Design Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.62
Design Guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.63
Makeup and Fill-Water Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.64
Other System Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.64
Emission Control and Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.65
Pollutants and Control Techniques. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.65
Nitrogen Oxides (NOx) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.65
Sulfur Oxides (SOx) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.67
Carbon Monoxide (CO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.67
Particulate Matter (PM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.67
Volatile Organic Compounds (VOCs)/Hydrocarbons (HCs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.67
Calculation of Annual Emissions for District Heating Boilers . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.68
Current Emission Standards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.69
Compliance Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.69
Instrumentation and Controls for District Heating Plants . . . . . . . . . . . . . . . . . . . . . . . . . 3.72
General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.72
Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.74
District Heating Plant Controls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.79
Boiler Controls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.80
Non-Boiler Controls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.82
Control Panels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.83
Energy Management and Control Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.85
Control Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.87
Controls for Boilers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.88
Boilers Supervisory Control Strategies and Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.89
Supply Water and Supply Pressure Reset for Boilers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.90
Supervisory Control and Data Acquisition (SCADA) System. . . . . . . . . . . . . . . . . . . . . . . . . . . . .3.91
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.93

viii

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Contents

Chapter 4 Distribution Systems


Hydraulic Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1
Objectives of Hydraulic Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.1
Water Hammer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.1
Pressure Losses. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.2
Pipe Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.2
Diversity of Demand. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.2
Network Calculations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.3
Condensate Drainage and Return. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.3
Distribution System Construction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.4
Aboveground System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6
Underground Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.7
Site-Fabricated Underground Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.11
Prefabricated Conduit Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.16
Geotechnical Trenching and Backfilling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.27
Piping Materials and Standards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.29
Supply Pipes for Steam and Hot Water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.29
Condensate Return Pipes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.29
Pipe Expansion and Flexibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.29
Pipe Bends and Loops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.31
Cold Springing of Pipe. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.34
Computer-Aided Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.35
Analyzing Existing Piping Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.35
Expansion Joints and Expansion Compensating Devices. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.36
Cathodic Protection of Direct Buried Conduits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.38
Sacrificial Anode Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.38
Impressed Current Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.39
Design, Maintenance, and Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.39
Leak Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.40
WSL Conduit Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.40
DDT-Type Conduit Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.40
Valve Vaults and Entry Pits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.40
Valve Vault Penetrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.41
Ponding Water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.42
Crowding of Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.44
High Humidity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.44
High Temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.44
Deep Burial. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.45
Freezing Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.45
Safety and Access . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.45
Vault Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.46
Vault Covers and Venting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.47
Construction of Systems without Valve Vaults. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.48
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.49

Chapter 5 Consumer Interconnection


Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1
Connection Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1

ix

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Direct Connection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5.1


Indirect Connection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5.3
Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4
Heat Exchangers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5.4
Flow Control Devices. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5.7
Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5.7
Controller . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5.8
Pressure Control Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5.8
Heating Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.8
Steam Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5.8
Hot-Water Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5.10
Building Conversion to District Heating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.13
Temperature Differential Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.13
Metering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.13
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.16

Chapter 6 Heat Transfer Calculations for Piping Systems


Thermal Design Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1
Soil Thermal Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.2
Soil Thermal Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.2
Temperature Effects on Soil Thermal Conductivity and Frost Depth . . . . . . . . . . . . . . . . . . . . . . .6.3
Specific Heat of Soils. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.4
Undisturbed Soil Temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.5
Heat Transfer at Ground Surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.7
Insulation Types and Thermal Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.11
Steady-State Heat Loss/Heat Gain Calculations for Systems. . . . . . . . . . . . . . . . . . . . . . . 6.12
Single Buried Uninsulated Pipe. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.15
Single Buried Insulated Pipe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.17
Single Buried Pipe in Conduit with Air Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.18
Single Buried Pipe with Composite Insulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.20
Two Pipes Buried in Common Conduit with Air Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.26
Two Buried Pipes or Conduits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.28
Pipes in Buried Trenches or Tunnels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.30
Pipes in Shallow Trenches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.34
Pipes in Loose Fill Insulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.35
Buried Pipes with Other Geometries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.37
Pipes in Air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.38
Economical Thickness for Pipe Insulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.38
Calculating Temperatures of System Components and Surrounding Soil Temperatures. . 6.42
Line Source of Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.43
Spherical Source of Heat. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.44
Superposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.45
Spherical Heat Leak. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6.45
Spherical Heat Leak with Superimposed Parallel Line Source of Heat . . . . . . . . . . . . . . . . . . . .6.47
Thermal Impacts on Utilities Adjacent to Buried Heat Distribution Systems . . . . . . . . . . . . . . . .6.49
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.49

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Contents

Chapter 7 Thermal Storage


Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1
What is Thermal Storage? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.1
The Purpose of the Thermal Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.1
How Thermal Storage Works. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.2
Benefits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3
CHP Plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.3
Heat Production Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.3
Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4
Principles of Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.5
Storage Tank Monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.7
ChargingDirectly Connected Heat Storage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.8
DischargingDirectly Connected Heat Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.8
Charging/DischargingPressurized and Decentralized Tank . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.9
Water Quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.10
Specific Design Issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.11
Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.11
Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.12
Sizing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.12
Regulatory Requirements (Europe) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.13
Economics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.14
Net Present Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.14
Seasonal Thermal Storage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.14
Examples of Thermal Storage. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.15
Denmark . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.15
Other Countries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .7.16
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.17
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.17

Chapter 8 Operation and Maintenance


Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.1
Workplace Safety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2
Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.2
Hazards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.3
System Security . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3
System Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3
Water Treatment and Filtration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.4
Corrosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.4
Corrosion Protection and Preventive Measures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.5
White Rust on Galvanized Steel Cooling Towers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.6
Scale Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.7
Nonchemical Methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.8
External Treatments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.8
Biological Growth Control. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.8
Control Measures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.9
Legionnaires Disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.12
Suspended Solids and Depositation Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.13
Mechanical Filtration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.13

xi

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Selection of Water Treatment Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.15


Once Through Systems (Seawater or Surface-Water Cooling) . . . . . . . . . . . . . . . . . . . . . . . . . . .8.16
Open Recirculating Systems (Cooling Towers) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.16
Closed Recirculating Systems (District Heating Distribution Systems) . . . . . . . . . . . . . . . . . . . .8.17
Medium- and High-Temperature Hot-Water Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.17
European Practice in Closed Distribution Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.18
Steam Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.18
Steam Distribution and Condensate Collection Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.19
Maintenance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.20
CMMS Functionality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .8.22
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.26

Appendix A 96-Hour Boiling Water Test


Tests of Complete System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A.1
Apparatus and Specimens for Other than Wet Poured Systems . . . . . . . . . . . . . . . . . . . . . . . . . A.1
Test Procedure for Other than Wet Poured Systems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A.2

Appendix B Climatic Constants


Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . B.1
Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . B.2

Appendix C Case Studies


CASE STUDY A: DISTRICT ENERGY/CHP SYSTEM IN JAMESTOWN, NEW YORK. . . . . . . . . C.1
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.1
System Description. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.1
Central Energy Plant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.1
Transmission and Distribution Network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.3
Buildings. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.4
Retrofitting Electrically Heated Buildings to Hot Water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.5
System Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.6
Phased Implementation Philosophy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.7
Marketing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.7
Ownership . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.8
System Economics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.8
Rates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.8
Financing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.9
System Benefits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.9
Customer Savings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.9
Environmental Advantages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.9
Demand-Side Management Application. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.9
Future Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.10
District Cooling System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.10
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.10
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.11
CASE STUDY B: CROTCHED MOUNTAIN BIOMASS DISTRICT HEATING SYSTEM . . . . . . . C.12
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.12
Background on the Crotched Mountain Facility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.12

xii

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Contents

Details of the Retrofit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.12


Fuel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.12
Central Plant. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.13
Distribution System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.17
Building Interconnection and Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.17
Operational Experience . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.18
Economic Benefits to the User . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.18
Environmental Benefits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.19
Societal Benefits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.19
Contact Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.19
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . C.19
CASE STUDY C: DISTRICT HEATING CONVERSION FROM STEAM TO HOT WATER
AT THE SAVANNAH REGIONAL HOSPITAL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Original System Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
System Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Replacement Options . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Construction Program . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Revised Central Plant Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Energy Use Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

. C.20
. C.20
. C.20
. C.20
. C.22
. C.22
. C.23
. C.23
. C.25

Appendix D Terminology for District Heating


. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . D.1

xiii

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Acknowledgments

The principal investigator and authors would like to thank the members of the Project
Monitoring Subcommittee (PMS) for their patience through the long process of creating
this document. This includes many discussions of scope and content as well as the actual
review. This document has benefited tremendously from the careful review of the PMS
members and their many suggestions based on the vast and diverse knowledge of district
heating that their composite experience represents.
The chair of the PMS, Steve Tredinnick, deserves special recognition for the countless hours he has invested in this effort both in his role as the PMS chair and as a major
unpaid contributor and a sounding board for the principal investigator.

Gary Phetteplace
January 2013

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Acronyms

ACFB
ASME
CAD
CCS
CHP
CMMS
CO
CO2
DDT
EOR
EPA
FGD
FGR
FRP
GDH
HAP
HDPE
HRSG
HTW
IDEA
LCC
LCCA
LCD
LFI
LTW
MTW
NFPA
NOx
NPV
O&M
OSHA

atmospheric circulating fluidized bed


American Society of Mechanical Engineers
computer-aided design
combustion control system
combined heat and power
computerized maintenance management system
carbon monoxide
carbon dioxide
drainable, dryable, testable
Engineer of Record
U.S. Environmental Protection Agency
flue gas desulfurization
flue gas recirculation
fiberglass-reinforced plastic
geothermal district heating
hazardous air pollutant
high-density polyethylene
heat recovery steam generator
high-temperature water
International District Energy Association
life-cycle cost
life-cycle cost analysis
liquid-crystal display
loose fill insulation
low-temperature water
medium-temperature water
National Fire Protection Association
nitrogen oxides
net present value
operation and maintenance
Occupational Safety and Health Administration

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

PC
PM
SCADA
SOx
VFD
VSD
WSL

xviii

pulverized coal
particulate matter
supervisory control and data acquisition
sulfur oxides
variable-frequency drive
variable-speed drive
water spread limiting

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

1
Introduction

PURPOSE AND SCOPE


The purpose of this design guide is to provide guidance for all major aspects of district heating system design. The guidance is organized in such a manner as to be of use to
the inexperienced designer of district heating systems as well as to provide a comprehensive reference to those immersed in the district heating industry. In addition to design
guidance, information on operation and maintenance has also been included.

DISTRICT HEATING BACKGROUND


District heating distributes thermal energy normally in the form of hot water or steam
from a central source to residential, commercial, and/or industrial consumers for use in
space heating, domestic hot water heating, process heating, cooking, and humidification.
Thus, the heating effect comes from a distribution medium rather than being generated on
site at each facility.
Whether the system is a public utility or user owned, such as a multi-building campus, it has economic and environmental benefits depending largely on the particular
application. Political feasibility must be considered, particularly if a municipality or governmental body is considering a district heating installation. Historically, successful district heating systems have had the political backing and support of the community.
District heating was proposed as early as 1613 in London; however, it would seem
that the first known examples where hot-water systems installed in St. Petersburg, Russia,
in the 1840s (Pierce 1994). In the United States it appears that the first district heating
system was installed at the Naval Academy in Annapolis, Maryland, in 1853 (Pierce
1994). Birdsell Holly is often credited with the invention of district heating in Lockport,
New York, in 1877, but he was just the first to really make a commercial enterprise of the
practice, a business that he ultimately grew to cities beyond Lockport (Pierce 1994). Most
major cities in the U.S. soon were served by steam-based district heating systems, with
the majority of those systems still surviving today, the largest being in New York City.
While steam systems were also built in a number of major European cites, hot-waterbased district heating saw significant growth in Europe, in part due to the reconstruction
after World War II. Also following World War II, the construction/expansion of many
U.S. military bases was an ideal application for district heating, and both steam and hightemperature hot-water systems were built. In Europe, especially Scandinavia, hot-water-

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

based district heating systems adopted low temperatures for supply; that trend continues
today with systems using supply temperatures of 158F (70C) or lower.

APPLICABILITY
District heating systems are best used in markets where the thermal load density is
high and the annual load factor is high. A high load density is needed to cover the capital
investment for the transmission and distribution system, which usually constitutes a significant portion of the capital cost for the overall system, often amounting to 50% or more
of the total cost. This makes district heating systems most attractive in serving densely
populated urban areas and high-density building clusters with high thermal loads, especially tall buildings. Urban settings where real estate is very valuable are good settings for
district heating systems since they allow building owners to make maximum use of their
footprint by moving most of the heating equipment off-site. Low-density residential areas
do not represent the most attractive markets for district heating. However, through the use
of low-temperature hot water, Denmark has one million detached family homes on district heating (Zinko et al. 2008).
The annual load factor is important because the district heating system is capital
intensive and maximum utilization of the equipment is necessary for cost recovery. Thus,
in general district heating systems are more economic in colder climates. However, the
use of absorption air conditioning by customers can also provide summer load for systems operating at high enough supply temperatures. Other methods to increase the annual
load factor for district heating systems are discussed by Zinko et al. (2008).

COMPONENTS
District heating systems consist of three primary components (see Figure 1.1):
Central plant or production source
Distribution network piping (steam or hot water)
Consumer interconnection (direct or indirect)
In the central plant (see Chapter 3), hot water or steam is produced by one or more of
the following methods/fuels:
Heat only, fuel fired. Fuel sources: all fossil fuels, biomass, refuse-derived fuel
(RDF).
Cogenerating heat and electric power, steam cycle. Fuel sources: all fossil fuels,
biomass, RDF.
Cogenerating heat and electric power, combustion turbine. Fuel sources: oil,
gas, syngas from coal, biomass, RDF, etc.
Cogenerating heat and electric power, internal combustion engine/generator.
Fuel sources: oil, gas, syngas from coal, biomass, RDF, etc.
Heat pump. Heat sources: sewage, seawater, geothermal (ground water, groundcoupling, surface water), industrial waste heat, heat recovery from cooling/
refrigeration, etc.
Geothermal direct use. Heat sources: geothermal hot water or steam.
The second component is the insulated distribution or piping network that conveys
the hot water or steam (see Chapter 4). The piping may be the most expensive portion of a
district heating system. These networks require substantial permitting and coordinating
with nonusers of the system for right-of-way if not on the owners property. Because the
initial cost is high, it is important to maximize use of the distribution system.

1.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

1 Introduction

Figure 1.1 Basic components of a district heating system.

The third component is the consumer system, which includes in-building equipment.
Hot water or steam may be used directly by the building systems or isolated by a heat
exchanger (see Chapter 5).

BENEFITS
Environmental Benefits
Generating heat in a central plant is normally more efficient that using in-building
equipment and, thus, the environmental impacts are normally reduced. The greater efficiencies arise due to the larger equipment and the ability to stage that equipment to
closely match the load yet remain within the equipment range of highest efficiency. District heating systems may take advantage of diversity of demand across all users in the
system and may also implement technologies such as thermal storage more readily than
individual building heating systems.
When fuels are burned to generate district heat within a central plant, emissions are
easier to control than those from individual plants, and control technologies that are not
technically or economically viable on a small scale may be used. For example, a central
plant that burns high-sulfur coal can economically remove noxious sulfur emissions,
whereas individual combustors could not. With concerns about global warming it is also
clear that carbon capture and sequestration will be easier to accomplish for a central plant
than for many smaller in-building heating plants.
District-scale systems also offer the ability to use fuels that would not be viable to use
on an individual building scale. For example, the thermal energy from municipal wastes
can provide an environmentally sound system, an option not likely to be available on a
building-scale system. Biomass may be burned in a district system with well-controlled
emissions; on an individual building scale this is being done in some areas residentially,
but the results from an emissions perspective have not been favorable. Geothermal

1.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

resources, normally uneconomical to develop on an individual building scale, can sometimes be more effectively applied as a source for district heating.
Finally, from an emissions and efficiency standpoint generating heat in a central plant
has the additional advantage of higher quality of equipment, higher seasonal efficiencies
due to demand diversity, less system heat loss from cycling, and higher levels of maintenance to maintain performance.

Economic Benefits
A district heating system offers the following economic benefits. Even though the
basic costs are still borne by the central plant owner/operator, because the central plant is
large, the customer can realize benefits of economies of scale. The economic benefits are
outlined below and are discussed in more detail in Chapter 2.
Operating Personnel
One of the primary advantages of a district heating system for a building owner is that
operating personnel for the HVAC system can be reduced or eliminated. Most municipal
codes require operating engineers to be on site when high-pressure boilers are in operation. Some older systems require trained operating personnel to be in the boiler/mechanical room at all times. When hot water or steam is brought into the building as a utility,
depending on the sophistication of the building HVAC controls there will likely be opportunity to reduce or eliminate building-plant-related personnel.
Insurance
Both property and liability insurance costs may be significantly reduced with the
elimination of boilers from within the building and fuel storage from the site since risks
of fire, accident, and fuel spill/release are reduced or eliminated.
Usable Space
Usable space in the building increases when a boiler and related equipment are no
longer necessary. The noise associated with such in-building equipment is also eliminated. In retrofit applications, this space cannot usually be converted into prime office
space; however, it does provide the opportunity for increased storage or other use.
Equipment Maintenance
With less mechanical equipment, there is proportionately less equipment maintenance, resulting in less expense and a reduced maintenance staff.
Higher Efficiency
A larger central plant can achieve higher thermal and emission efficiencies than can
several smaller units. When strict regulations must be met for emissions, water consumption, etc., control equipment is also more economical for larger plants. Partial-load performance of central plants may be more efficient than that of many isolated small systems
because the larger plant can operate one or more capacity modules as the combined load
requires and can modulate output.
Available Primary Energy
As noted in the Environmental Benefits section, on a building scale it may not be
practical to generate heat from some fuel sources that may be used in a district system
central plant.

1.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

1 Introduction

INITIAL CAPITAL INVESTMENT


The initial capital investment for a district heating system is usually the major economic driving force in determining if there is an acceptable payback for implementation.
Normally, the initial capital investment includes four components:
Concept planning
Design
Construction
Consumer interconnection

Concept Planning
In concept planning, three areas are generally reviewed. First, the technical feasibility
of a district heating system must be considered. This includes factors such as looking at
the heating load density, potential locations for the central plant, and an acceptable corridor for the distribution system. Financial feasibility is the second consideration. For
example, a municipal or governmental body must consider availability of bond financing;
a private entity must be able to secure financing at an acceptable rate. Alternative choices
of potential customers for heat generation must be reviewed because consumers are often
asked to sign long-term contracts in order to justify district heating systems and their economics must be understood.

Design
There are many decisions that must be made in the detailed design of the heating
plant, distribution system, and consumer interconnect, as discussed in Chapters 3, 4,
and 5, respectively. For the central plant these decisions will be driven by the location
(i.e., availability of water, land, etc.) and size of the heating plant. The distribution system
accounts for a significant portion of the initial investment. Distribution design depends
largely on the type of distribution system construction that will be used. The routing of
the system will also be very important in the design of the distribution system. Failure to
consider these key variables results in higher-than-planned installation costs. An analysis
must be done to determine how much insulation is warranted; this process is outlined in
Chapter 6.

Construction
The construction costs of the central plant and distribution system depend on the
quality of the concept planning and design. Although the construction cost usually
accounts for most of the initial capital investment, neglect in any of the other three areas
could mean the difference between economic success and failure. Field changes usually
increase the final cost and delay start-up. Even a small delay in start-up can adversely
affect both economics and consumer confidence. It is extremely important that the successful contractors have experience commensurate with the project. Capital costs of district heating projects range greatly and are dependent upon site conditions such as
labor rates,
the construction environment (slow or busy period, high- or low-cost area, etc.),
the distance for shipping of equipment,
fuel choice/availability,
permits and fees (franchise fees),
local authorities (traffic control, allowed times for construction in city streets),
soil conditions (clay, bedrock),
the quality of equipment and controls (commercial or industrial),
the availability of materials,

1.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

the size of piping in the distribution system,


the type of distribution system installation (insulated, uninsulated, welded/fused
or mechanical joints, cathodic protection, etc.),
the depth of bury and restoration of existing conditions (city streets, green
areas),
below-grade conflict resolutions, and
economies of scale.
Lead time needed to obtain equipment may determine the time required to build a district heating system. In some cases, lead time on major components in the central plant
can be over a year. Installation time of the distribution system depends in part on the routing interference with existing utilities. A distribution system in a new development is simpler and requires less time to install than a system being installed in an established
business district.

Consumer Interconnection
Interconnection costs are usually borne by the consumer. High interconnection costs
may favor an in-building plant instead of a district heating system. The nature of the consumers in-building heating system may have a large impact on the ability to connect to a
district heating utility (see Chapter 5 for details on building conversion). For this reason,
in either new construction or retrofit situations, particular attention must be paid not just
to the consumer interface but in addition the remainder of the heating system within the
building.

TYPICAL APPLICATIONS
District heating has seen a wide variety of applications; some case studies of examples are provided in Appendix C. These applications span all major sectors of the building
market: residential, commercial, institutional, and industrial. For many of the applications, such as college campuses, medical complexes, and military bases, the loads are
captive (i.e., there is a common owner for the district heating plant and the buildings
being served). At the opposite end of the spectrum are district heating systems that operate as commercial enterprises in urban areas and compete with in-building equipment for
cooling loads. Between these two extremes are many other business models such as district heating providers, who operate under contract, plants owned by developers of real
estate projects, etc. Business models and business development for district heating enterprises are largely analogous to those for district cooling enterprises; information on the
latter may be found in IDEA (2008).

REFERENCES
IDEA. 2008. District Cooling Best Practices Guide. Westborough, MA: International
District Energy Association.
Pierce, M. 1994. Competition and cooperation: The growth of district heating and cooling, 11821917. Proceedings of the International District Energy Association
LXXXV: 1929.
Zinko, H., B. Bhm, H. Kristjansson, U. Ottosson, M. Rm, and K. Sipil. 2008. District
heating distribution in areas with low heat demand density. Report No. 2008: 8DHC08-03, Annex VIII, International Energy Agency, Paris, France.

1.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2
Planning and
System Selection
PLANNING OVERVIEW
For buildings, the cost of utility services and infrastructure needed to provide these
services is significant and may even exceed the cost of the buildings themselves over their
lifetimes. Planning has the potential to reduce both the initial and the future costs. The
objective of planning should be to guide decision making such that cost savings in providing utilities over the life expectancy of the building and the district heating system are
realized. Due to their capital-intensive nature, the need for planning is significant for district heating systems.
The term master plan will be used to refer to planning in this design guide, but there
are several levels of planning. A master plan covers all levels of planning and is typically
integrated with the development of both new and existing sites. Throughout this section
the development of utility master plans will typically be applicable to private, public, and
utility-owned systems. Differences in planning for the different systems will be identified
whenever those differences are significant.
For totally new systems in a greenfield project, master planning is an essential first
step to ensure that the owners requirements are fulfilled in the delivered system. Appropriate planning will have significant impacts not just on the first cost of the project but
also on the future operation and maintenance (O&M) costs. The master plan also will provide valuable information to those responsible for the O&M of the system; thus, it is
essential that the individuals who will be responsible for the O&M of the system be
involved in the planning process. For sites with existing systems/other utilities, involvement of the system operators in the development of a master plan is essential; there is no
substitute for the corporate knowledge when dealing with buried systems. A plan properly prepared will also be of benefit to the users served by the system, so it may be prudent to engage these representatives in the planning process. Potential stakeholders in the
development of a master plan include the following:
Building/campus/site owner
Owners project engineer
Site master planners
Utility system operators
Operators/engineers of other utilities on site or within utility right-of-way
Potential contractors

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

District heating consumers (users)


Adjacent residential neighborhood associations and business enterprises
This list is by no means exhaustive; special circumstances could result in many different and varied stakeholders. The planning process should be all-inclusive of everyone
involved in the design, construction, operation, and use of the facilities to be provided
with district heating.
Utility master plans for existing facilities also must consider whether to correct deficiencies built into them initially or during periods of rapid expansion. Utility plans for
existing systems also must account for aging of the existing infrastructure and consider
appropriate timing for replacement and upgrading efforts. Furthermore, building functions can change over the life of the building (i.e., college dormitories can be converted to
offices, etc.); therefore, the service lines to the building and distribution mains must be of
adequate size to accommodate future facility remodeling or expansion.
Planning is a difficult task that takes time and creative energy to explore the many
variables that must be considered to develop a system that will operate in accordance with
design intent and is sustainable, reliable, energy efficient, environmentally friendly, and
easily operated and maintained and supports expansion to serve new loads. One additional variable, which may be the most important of all considerations, is the development
of a concept-level budget construction cost estimate (opinion of probable cost). Due to
poor preliminary cost estimates, often before construction begins, systems have to be
modified during the bidding process to use lower-quality materials. These modifications
ultimately prove detrimental to the usability and life of the system. Early planning and
concept designs of adequate detail to develop more accurate cost estimates can mitigate
many of these issues. In this chapter, guidance is given for developing a system utility
master plan that is sustainable, meets the owners performance needs and desires, and can
be constructed within the owners constraints for overall capital cost and cash flow.
What should a utility master plan provide an owner? At a minimum, a utility master
plan should provide a prioritized program for long-term guidance for building, expanding, and upgrading the district systems that are typically built incrementally. A good utility master plan serves as a technically sound marketing tool for the owners engineers to
present needs and solutions to management or to prospective customers. Unfortunately,
many owners view utility master plans as an interesting technical exercise with a life of
one or two years. When this has been an owners experience, it usually results for one or
more of the following reasons:
Failure to involve the owners staff
Failure to provide intermediate owner reviews
Use of an unverified database
Lack of creativity in developing technically sound system alternatives for
screening and final selection by the owner
Inaccurate cost estimating often related to overly optimistic estimating using
unit costs that do not include all elements of the systems
The process for developing a utility master plan may be likened to a pyramid (Bahnfleth 2004); see Figure 2.1. The success of the plan depends upon the foundation: a strong
database that includes discovery and verification. The development of a strong and accurate database is the foundation on which all other aspects of a master plan stand and get
their credibility. With the database in place, identification of alternatives and preliminary
estimates of costs (screening grade) for each alternative are used to select, with the owner,
the most promising alternatives. These alternatives are then subjected to more intense

2.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Figure 2.1 The master planning pyramid.

analysis before the final decision is made of how the new plant (or expansion of an existing plant) will be developed as well as how a plan for future projects will be laid out.
Thus, the pinnacle of the pyramid is a prioritized, priced list of projects needed to keep
pace with the physical growth of a facility and to provide replacements and upgrades to
the existing system, if there is one.

ESTABLISH AND CLARIFY THE OWNERS SCOPE


Before the planning process can begin in earnest, as with every project the owners
scope and expectations for the project must be fully clarified. While a scope may have
been written in the preliminary stages of the owners development of the project, often
this scope will need clarification and refinement. Meeting with the owner, the staff who
will inhabit the facilities, those performing the O&M, and those responsible for budgeting
for the project and its O&M costs will help clarify and refine the scope and ultimately can
eliminate many potential issues and challenges later. In fact, such open meetings can
bring new insights to the owner, sometimes resulting in a totally new scope being developed. During the initial scope review, other factors affecting the system planning can be
determined through proper interaction with those present. Among the things that should
be discussed are the level of system quality expected, the budget and cash flow limitations, and anticipated potential long-term expansion of the system. With the present
emphasis on sustainability, this is an appropriate time to discuss the owners interests and
desires with respect to this important aspect of system design and development.
Throughout the process of establishing and refining the scope it is paramount that all
parties recognize that every system is subject to three basic constraints: budget (owners),

2.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

scope limitations (jointly developed by the engineer and the owner), and quality versus
cost (a champagne-quality system cannot be built on a beer budget.)

DEVELOPMENT OF THE DATABASE


Sources of Data for Existing Systems
In existing systems, inventories of available equipment that will become a part of the
new or expanded system must be obtained from the owners records and/or by gathering
data from available shop drawings and equipment nameplates. Additionally, condition
assessments of equipment that may be reused in a new system should be conducted.
Equipment conditions can be found in maintenance records, in code inspections (where
required), and from operating personnel interviews. Site utility maps showing locations of
existing utilities (water, sewer, electric, telephone, steam, hot water, etc.) are essential for
planning distribution system routing to avoid interferences to minimize the expense of
installing tunnels, shallow trenches, or direct burial piping. The site electrical drawings
will identify the locations of existing electrical system feeders and available substations.
A single-line electrical drawing showing system loads and feeder capacities, among other
things, should be obtained when available. As the planning process proceeds, components
of the database will be expanded by the planning team, but early identification of the need
for them can assist in meeting deadlines established by the owner.

Heating Load
Because gathering information for developing a suitable system planning database
takes some time from the owners staff that already have their regular jobs to do, it is
timely to provide at the outset a list of information and data that will be required to complete a good system utility master plan. The list includes whatever heating load data exists,
either in the form of estimated loads for new buildings to be served by the district heating
system or recorded load data for existing systems. When considering historic heating load
data it is particularly important to bear in mind the bias towards reduced heating requirements that has resulted in recent years from the combined impacts of increasing electronics
use in offices, dormitories, and other spaces as well as the tightening of building envelopes
by retrofit measures. Conversely, it is also important to consider the increases in heating
loads that may have resulted from increased ventilation rates, to the extent they have not
been moderated by the addition of ventilation heat recovery equipment.
Establishing the heat production capacity needed during the life established for the
master plan is one task that too often is complicated by attempts to develop computerbased load profiles for each and every building. But for large systems serving tens or even
hundreds of buildings, experience indicates that computer-based load analyses of each
building may not be necessary and can be very deceptive. It is an example of precision
exceeding the accuracy needed, especially when future additions of large numbers of
minimally defined buildings are added to the mix. Good examples of such additions were
new chemistry facilities in the 1990s and the large number of biological research buildings under construction today. Furthermore, too much credence can be given to loads calculated by a computer program without adequate scrutiny for reality. The use of
appropriate unit loads in Btu/hft2 (W/m2) often provides estimates of adequate accuracy,
providing consideration is given to the mix of buildings being served. Hyman (2010) provides a summary of the options for obtaining load data:
Energy metering data from an energy management system (EMS)
Meter readings at the building or equipment level
Analysis of utility bills
Computer energy modeling; requires calibration for existing buildings

2.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Installed equipment capacity where available


Norms for capacity per unit area, i.e., Btu/hft2 (W/m2)
While all of these methods may be applicable to existing buildings, only computer
modeling and the use of norms per unit area may be readily applied to planned future
buildings.
When determining plant capacity it is also important to include the load that is placed
on the heating plant by heat losses from the distribution system. Chapter 6 covers calculation of heat losses from all the principal types of insulated heat distribution system piping
and provides examples.

Demand Diversity
Regardless of the method(s) used to establish individual building loads, when establishing system capacity the diversity of consumer demands (i.e., that the various consumers maximum demands do not occur at the same time) should be considered. Thus, the
heat supply and main distribution piping may be sized for a maximum load that is less
than the sum of the individual consumers maximum demands. For steam systems, Geiringer (1963) suggests diversity factors of 0.80 for space heating and 0.65 for domestic
hot-water heating and process loads. Geiringer (1963) also suggests that these factors
may be reduced by approximately 10% for high-temperature water (HTW) systems,
owing to the significant heat capacity of the high-temperature hot water within the piping
network. Werner (1984) conducted a study of the heat load on six operating low-temperature hot-water systems in Sweden and found diversity factors ranging from 0.57 to 0.79,
with the average being 0.685.
The appropriate diversity factor is directly dependent upon the number and character
of the buildings being served, recognizing that different building functions will not peak
at the exact time of day. Establishing the correct diversity factor is more of an art than a
science, and its development should involve the system owner and operators so they are
aware of the assumptions created. The diversity factor also does not have to be accurate to
the third decimal point since accuracy at that level is not useful in accommodating future
growth and is often misleading. If the district heating system is existing and each building
or customer has an energy meter, then the diversity for a specific year or years can be calculated by determining the peak load day at the central plant and comparing it to when
each building actually peaked during the season.

Heat Load Density and Piping Costs


Often for district heating systems the distribution system represents the majority of
the capital cost investment; this is particularly true for systems of low heat demand density, such as those serving residential or light commercial neighborhoods or even some
campus settings such as military bases or universities, where building are separated by
significant green space. In fact, distribution costs are normally the factor preventing
expansion of district heating outwards from cities that already have established systems
to the less densely populated suburbs. Much effort has been expended, particularly in
Northern Europe, to develop methods that would allow district heating to be expanded
into areas of low heat load density. A report on the topic by Zinko et al. (2008) provides a
compendium of useful information on established approaches in the Scandinavian countries as well as many novel approaches that hold promise. When a new district heating
system or a significant expansion of an existing system is being contemplated, the
approaches discussed by Zinko et al. (2008) will likely be necessary for the economics to
be acceptable. Cost data presented by Zinko et al. (2008) are significantly lower than

2.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

those given in Table 2.1, which is based largely on U.S. experience with larger and
higher-temperature systems. For example, smaller-diameter piping using flexible nonmetallic piping as discussed in Chapter 4 might be expected to cost around $100 per foot
($328 per metre) for installed supply and return piping. This cost can be contrasted to the
minimum cost of $500 per linear foot of trench ($1640 per linear metre of trench) given
in Table 2.1. The cost data of Table 2.1 are more typical of larger-diameter steel piping
that would be expected in the U.S. and cover a range of installation conditions more typical to those found in the U.S., which include congested cities in the worst case and college campus-type settings in the best case. Most district heating applications in the U.S.
to date consist of large buildings connected rather than the relatively small buildings
found in the Scandinavian villages upon which the data of Zinko et al. (2008) is based.
In Scandinavia, a useful parameter has been developed to compare the relative load
density of district heating systems. This parameter quantifies the load density for a heat
distribution system in relation to the amount of piping that must be provided. This parameter is referred to as the line heat density (LHD) by Zinko et al. (2008). With more than
one million detached homes served by district heating, Denmark has the lowest LHD of
the Scandinavian counties and most likely the world. The average value of the LHD for
Denmark is around 0.3 MWth/ftyr (1 MWth/myr). Lower LHD would normally indicate a less viable system economically. Zinko et al. (2008) cite several studies for Danish
and Swedish conditions that suggest that LHDs as low as 0.06 MWth/ftyr (0.20 MWth/
myr) are feasible, although those approaching this lower limit are very rare.

ALTERNATIVE DEVELOPMENT
Codes and Standards
Once loads have been established, both existing and planned, and all data on existing
system(s), if any, as well as that of other utilities, have been gathered, identification of
alternatives can begin. Identifying potential sites for new central plant(s) will be the first
task. This task should begin with a review of voluntary and mandatory codes, standards,
and regulations; it is a necessary step in planning district heating systems that will preclude potential conflicts that could result in wasted effort and resources in planning and
design. Codes include local and state codes applicable to
construction of the central heating plant or any satellite plants,
construction of piping systems above and below ground,
determining limits of emissions,
preparation of construction and operating permits,
selection of equipment to meet limits on emissions, noise, waste water quality,
etc., and
system performance with emphasis on safety and energy efficiency as it relates
to sustainability.
Standards from organizations such as ASHRAE, American Society of Mechanical
Engineers (ASME), and Air-Conditioning, Heating, and Refrigeration Institute (AHRI),
as well as the National Fire Protection Association (NFPA) National Electrical Code
(2011), should be consulted
to ensure systems and equipment employed in the district heating plant meet
minimum construction and performance specified in design documents;
to ensure that performance testing is carried out using accepted test procedures
and instrumentation;
to ensure safety of any refrigeration systems, including various water chillers
and refrigerant storage systems;

2.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

to provide adequate ventilation of any building housing the refrigerants to protect operators from exposure to excessive amounts of the chemical; and
to select chillers, boilers, hot-water generators, and various pressure vessels that
meet safety requirements.
Regulations that must be complied with may be local or national. Federal regulations
such as those from the U.S. Department of Energy (DOE) and the U.S. Environmental
Protection Agency (EPA) may impact
emissions (this is especially applicable to systems burning fuels where new
sources of emissions may require permitting),
fuel handling,
energy efficiency, and
carbon footprint.
State and local regulations that may impact the proposed project include building
codes and standards covering
materials and safety requirements,
emission limits for heating plants,
construction and operating permits,
boiler and pressure piping codes, and
right-of-way for installing piping and electrical systems, where required.
The significance of reviewing and becoming very familiar with various codes, standards, and regulations should not be underestimated. Codes, standards, and regulations
impact the design and cost of district heating systems. They are especially important
when meeting special building and operating requirements, such as restrictive emission
limits and chemical storage and handling. In addition, codes may significantly impact the
aesthetic and mechanical design of a utility plant and its distribution system.

Local and Institutional Constraints


Institutional/community practices and priorities may also be considered as an additional level of code that must be complied with. For example, local practice may
require that the central plant be constructed with visual screening so as to disguise equipment or fuel storage. Another important institutional or local factor that must be considered is the disruption that construction/expansion of a buried heat distribution system may
cause. While routing of piping across a common area, park, or green space might be the
most expeditious route, doing so might be unacceptable. Trees are usually considered
valuableand well they should be when they are special and/or rare species, not to mention their role as good oxygen producers. Good practice is to search out the clients forester or arborist, if there is one, before routing large underground piping near trees. Trees
and other campus-type monuments may have greater impact upon pipe routing than congestion from existing active and abandoned utilities. If a forester or arborist is not available, avoiding the trees drip line is good practice. A good composite utility map is the
starting point for pipe routing, but ultimately a three-dimensional representation of the
underground utilities is key; it should include a representation of the surface factors such
as buildings, trees, paved surfaces, etc.

Choice of Heating Medium


In plants serving hospitals or industrial customers, or those also generating electricity,
steam is the usual choice for production in the plant and, often, for distribution to customers. For systems serving largely commercial buildings, hot water is an attractive medium.

2.7

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Hot water has also been used successfully to serve areas of low load density, such as residential neighborhoods; see the discussion under Heat Load Density and Piping Costs.
From the standpoint of distribution, hot water can accommodate a greater geographical
area than steam due to the ease with which booster pump stations can be installed. The
common attributes and relative merits of hot water and steam as heat-conveying media are
discussed in the following subsections.
Heat Capacity
Steam relies primarily on the latent heat capacity of water rather than on sensible
heat. The net heat content for saturated steam at 100 psig and a saturation temperature of
338F (690 kPa gage and 170C saturation temperature) condensed and cooled to 180F
(82C) is approximately 1040 Btu/lbm (2400 kJ/kg). Hot water cooled from 350F to
250F (175C to 120C) has a net heat effect of 103 Btu/lbm (240 kJ/kg), or only about
10% as much as that of steam. Thus, a hot-water system must circulate about 10 times
more mass than a steam system of similar heat capacity. While this might appear to be a
disadvantage, in reality it is not since pumping costs in properly designed hot-water systems are low (approximately 6% or less) and the additional mass in the system provides a
thermal flywheel within the network that has the ability to meet transient peak loads at
consumers locations. A rapid increase in load at one or more consumer locations within a
steam network may cause pressure to temporally be reduced, which may in turn cause
problems with other customers and/or pressure regulation equipment.
Pipe Sizes
Despite the fact that less steam is required for a given heat load and flow velocities
are greater, steam usually requires a larger pipe size for the supply line due to its lower
density (Aamot and Phetteplace 1978). This is compensated for by a much smaller condensate return pipe. Therefore, piping costs for steam and condensate are often comparable with those for hot-water supply and return given equal heat transport.
Load Modulation
Hot water offers the advantage of simple reset of temperature as a means of load
modulation. For steam systems, major adjustments in pressure are normally much more
difficult for pressure regulating and terminal equipment to accommodate. Flow is also
easily modulated in hot-water systems, normally by pumps with variable-speed drives
(VSDs) at the central heating plant that are controlled to maintain a minimum pressure
differential at the most hydraulically distant consumer location.
Return System
Condensate return systems require more maintenance than hot-water return systems
since hot-water systems function as a closed loop with very low makeup water requirements. For condensate return systems, corrosion of piping and other components, particularly in areas where feedwater is high in bicarbonates, is a problem. Nonmetallic piping
has been used successfully in some applications, such as systems with pumped returns,
where it has been possible to isolate the nonmetallic piping from live steam. However,
largely the initial promise of nonmetallic condensate return has not been realized.
Additional concerns are associated with condensate drainage systems (steam traps,
condensate pumps, and receiver tanks) for steam supply lines. Condensate collection and
return should be carefully considered when designing a steam system. Although similar
problems with water treatment occur in hot-water systems, those problems present less of
a concern because makeup rates are much lower. Makeup rates for hot-water district heating systems typically are less than 6%, with many systems operating at 3% or lower
makeup rates. For steam systems that return condensate, the best systems achieve make-

2.8

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

up rates of approximately 12% to 15%, and on average makeup rates are much higher,
perhaps 25%. Its not uncommon to find older steam systems with makeup rates of 50%,
especially where distribution systems are extensive. The difficulty in condensate return
led early district steam systems, as discussed in Chapter 1, to operate without return systems and discard the condensate at the consumer location. Of course, lower fuel and water
costs provided less incentive for condensate return in early steam district heating systems
as well. In addition to the increased costs associated with water treatment and energy lost
in the condensate not returned, there is the issue of labor. Condensate systems require
constant attention to maintenance, the primary source being steam traps and the appurtenant piping.
Pressure and Temperature Requirements
Flowing steam and hot water both incur pressure losses. Hot-water systems may use
intermediate booster pumps to increase the pressure at points between the plant and the
consumer. Due to the higher density of water, pressure variations caused by elevation differences in a hot-water system are much greater than in steam systems. This can
adversely affect the economics of a hot-water system by requiring the use of a higher
pressure class of piping and/or booster pumps.
Regardless of the medium used, the temperature and pressure used for heating should
be no higher than needed to satisfy the consumer requirements. This cannot be overemphasized. Higher temperatures and pressures require additional planning, engineering,
and construction expenses to avoid higher heat losses. Safety and comfort levels for operators and maintenance personnel also benefit from lower temperatures and pressures.
Higher temperatures may require higher pressure ratings for piping and fittings and may
preclude the use of materials such as polyurethane foam insulation and nonmetallic conduits. See Chapter 4 for additional information on piping systems for both steam and hot
water.
Hot-water systems are divided into three temperature classes:
High-temperature systems supply temperatures over 350F (175C)
Medium-temperature systems supply temperatures in the range of 250F to
350F (120C to 175C)
Low-temperature systems supply temperatures of 250F (120C) or lower
The temperature drop at the consumer end should be as high as possible, preferably
40F (22C) or greater. A large temperature drop allows the fluid flow rate through the
system, pumping power, return temperatures, return line heat loss, and condensing temperatures in cogeneration power plants to be reduced. A large customer temperature drop
can often be achieved through the cascading of loads operating at different temperatures.
Summary on Heating Medium Choice
In a few instances, existing equipment and processes will require the use of steam.
However, unless there is a need for steam that cannot be met by hot water, hot water is
clearly the superior choice of medium. Where a steam need exists, it may be prudent to
look at other means of meeting the needfor example, converting steam cooking appliances to gas or electricas opposed to letting a relatively small need for steam drive the
choice of medium.
Because of the advantages of hot water, many legacy steam systems in Europe have
converted to hot water, as have some in the United States. The case studies in Appendix C
include two systems that were converted from steam to hot water. Conversion of a building heating system from steam to hot water is discussed in Chapter 5.

2.9

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Integrated Processes
Where heating requirements are being met with a district system, it may be advantageous to investigate generation of electric power and even chilled water for district cooling. Cogeneration of heat and electric power is covered later in this chapter and in
Chapter 3, as well as in Chapter 7 of ASHRAE HandbookHVAC Systems and Equipment (2012). Absorption-based cooling may be combined with vapor compression chillers to accomplish the chilled-water production. If the generation of heat and electricity
will be combined with the generation of chilled water, the array of possibilities that must
be considered is greatly expanded. In addition, the selection of fuels may be much
broader, which will impact plant siting, as discussed in the next section. In the planning
stage, consideration should be given to thermal storage as both a means of reducing electric costs and as a chiller capacity reduction possibility, as well as for the backup capacity
it provides. Thermal storage should also be considered for district heating supply where
intermittent heat sources, capacity limitations, and/or the need for backup exist. Thermal
storage for district heating is covered in Chapter 7 of this design guide. For all issues of
district cooling including thermal storage, the reader is referred to this books companion
design guide on district cooling, District Cooling Guide (ASHRAE 2013a).

Central Plant Siting


Because site constraints are so highly varied it is difficult to provide general guidance
on plant siting. For larger systems it may be prudent to consider multiple heating plants
which may include phased construction of the plants. In addition to the code and standards constraints previously discussed, the following factors should be considered in
choosing a plant site.
Aesthetics
While ideally the plant would be located at the centroid of the loads to be served, siting will be heavily influenced by aesthetics and arrangement of the buildings. In evaluating alternative sites, it may be necessary to conduct preliminary hydraulic analysis
(discussed later) to evaluate the distribution system first-cost and operating-cost impacts
of competing plant locations.
Acoustics
The plant should be sited away from sound-sensitive adjacencies such as residential
areas, music halls, libraries, etc., and all measures should be taken to keep all local equipment noises from transmitting through the central plant building openings. Scrutiny and
calculations relating to acoustical and vibration abatement should be included at the planning stage.
Topography
Topographical factors may play a large role in plant siting: a plant located at the low
point in the system will be subjected to system hydrostatic heads, which may be significant where elevation differences are great. Where a thermal energy storage system is used
(see Chapter 7), it may be desirable to locate it at a higher elevation than the building
equipment served, whether that equipment consists of the heat exchangers of an indirectly
connected system or the coils of a directly connected system.
Fuel Availability, Storage, and Handling
Obviously, a natural-gas-fueled plant will require either location near an adequate
existing supply pipeline or construction of a supply pipeline from a nearby pipeline main.
For systems that use liquid or solid fuels, adequate space must be available on site for fuel
unloading and storage. Fuels such as biomass that have relatively low heat content per

2.10

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

unit volume will require frequent deliveries or increased storage capacity on site. Solid
fuels such as biomass or coal may be delivered on trucks or by rail where available, and
specialized facilities may be required at the plant site to unload the trucks or rail cars.
Adequate electric infrastructure will also be a consideration in plant siting.

Heat Distribution Systems


Once the loads at the buildings have been established, the distribution system can be
laid out (i.e., a plan view can be developed). Chapter 4 provides detailed information on
the design of heat distribution systems and Chapter 6 provides details on the thermal analysis of these systems. Heat distribution systems must supply sufficient energy at the
appropriate temperature and pressure to meet consumer needs. Within the constraints
imposed by the consumers end use and equipment, the required thermal energy can be
delivered with various combinations of temperature and pressure (see the previous section
entitled Choice of Heating Medium for discussion on temperatures, T, and pressure).
Computer-aided design (CAD) methods are available for thermal piping networks
(Bloomquist et al. 1999; COWIconsult 1985; Rasmussen and Lund 1987; Lund-Hansen
2010; Lund-Hansen and Hansen 2011). The use of CAD methods allows the rapid evaluation of many alternative designs. General steam system design can be found in
Chapter 11 of ASHRAE HandbookHVAC Systems and Equipment (2012) as well as in
the International District Heating Association (IDHA) District Heating Handbook
(1983). For water systems, consult Chapter 13 of ASHRAE HandbookHVAC Systems
and Equipment (2012) and District Heating Handbook (IDHA 1983).
With the assumed T and the loads, the design flows are calculated and the hydraulic
analysis of the distribution network may proceed. Pipe sizes must be chosen; that will be
the critical aspect of the planning process (see Chapter 4 for further details on pipe sizing). Where there is high uncertainty in the planning process regarding future loads, overdimensioning of pipes is generally preferable to the converse. Excavating to replace an
undersized existing segment of piping can be very costly after surfaces, vegetation, and
buildings are established and occupied. The replacement is further complicated by the
fact that there are buildings connected that must have continuous service and cannot be
without heating for an extended period of time.
Hydraulic analyses using the CAD tools discussed above should also look for the
potential to establish loops within the system. Loops within distribution systems are to be
favored when the routes are available for them as the loops provide alternate flow paths
that may be used during disruption to flow via the primary path and when the primary
path is in use; secondary paths via loops will reduce flow and thus pressure losses via the
primary path. Hydraulic simulation of the network is essential if loops are to be used and
their functioning under various load conditions understood. Hydraulic analyses are also
useful tools for selecting control valves at end users locations. In essence, valves that are
capable of dissipating the excess available head must be selected at the building interface
with the distribution system; see Chapter 5.
In existing systems, the hydraulic analysis should look closely at bottlenecks within
the system. The critical flow paths must be identified by the analyses; otherwise efforts to
re-dimension the network for future requirements or to correct existing deficiencies will
fail. The addition of additional loops will often be the solution to bottleneck problems and
reduce overall system pressure drop.

Construction Considerations and Cost


A key element in the decision making process in any master planning effort is the cost
estimation of the equipment, materials, and labor required to implement the proposed
alternatives. Thinking like a mechanical contractor during the cost estimating process is

2.11

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

essential if the results are to be appropriation-grade estimates. The master plan and its
ultimate implementation will fail if proper attention is not given to the estimating process.
Although systems have not been designed in the master planning process, enough data are
generated to take estimates to a fair amount of detail. Experience indicates that estimates
made today can be used with cost escalation data out as far as ten years or more when
they are based on knowledge of the systems proposed and not generalized unit cost data
that are readily available. Chapter 1 includes a list of factors that impact construction
costs and a discussion on the importance of accurate estimates.
Examples of construction costs for heating-only and cogenerating central plants are
provided later in this chapter. For chilled-water plants and distribution systems, sample
construction cost unit pricing is summarized in ASHRAE HandbookHVAC Systems and
Equipment (2012) and is provided in Table 2.1. However, the designer is cautioned that
costs can vary widely based on the various factors that impact construction costs as outlined in Chapter 1. For example, very large chiller plants that have been constructed in the
Middle East can cost as little as $1250 per ton while more typical values for chiller plants
of around 10,000 tons capacity are in the $2500 to $3000 range. In the case of distribution
systems, large-diameter piping can increase costs significantly, as can urban construction
conditions; however, as noted previously in this chapter in the section titled Heat Load
Density and Piping Costs, small-diameter low-temperature hot-water piping can be much
less costly than Table 2.1 would indicate.
Lead time needed to obtain equipment generally determines the time required to build
a district heating system. In some cases, lead time on major components in the central
plant can be over a year. Installation time of the distribution system depends, in part, on
the routing interferences with existing utilities. A distribution system in a new industrial
park is simpler and requires less time to install than a system being installed in an established business district.

Consumer Interconnection
The interconnection of the buildings with the distribution system will be a major cost
component of the system that must not be ignored in the planning phase. Consumer interconnection costs will vary widely depending on the type of existing system in the existing
buildings (if any) and the type of building interconnection, direct or indirect. Table 5.1
provides a summary of the relative merits of direct versus indirect connections. Consumer
interconnection costs are usually borne by the consumer for a system that is a commercial
venture. However, the magnitude of interconnection costs must be considered in the economic analysis discussed in the Economic Analysis and User Rates section of this chapter. High interconnection costs may favor an in-building plant for the customer instead of
a district heating system.
Table 2.1 Sample Cost Information (ASHRAE 2012)
Item

Cost Range per Unit*

Unit of Measure

Chiller plants
(including building, chillers, cooling towers, pumps, piping, and controls)

$1800$3500

ton of capacity

Direct buried distribution piping


(including excavation, piping, backfill, surface restoration)

$500$1500
($1640$4920)

linear foot (metre)


of trench

Distribution system for buried inaccessible tunnels

$700$1500
($2300$4920)

linear foot (metre)


of trench

Distribution system for buried walk-through tunnels

$3500$15,000
($11,480$49,200)

linear foot (metre)


of trench

* Costs include design fees, contingencies, and taxes.

2.12

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Retrofit costs for buildings will vary widely. For buildings that were steam heated,
retrofit can be costly but is possible; details are provided in Chapter 5.
Metering is a necessary component in nearly all consumer interconnections, and
appropriate allowance should be made for such costs in budget construction cost estimates. While it is tempting to not include metering when there is a common owner for all
buildings (i.e., a college, military, or institution campus), experience has shown that troubleshooting both network and building systems operational problems is greatly enhanced
by metering. Metering should include the ability to obtain flow and temperature data as
well as energy usage; many systems will allow for remote monitoring of this data at a
central point such as the district heating central plant. See Chapter 5 for a detailed discussion on metering and remote systems monitoring.

Energy Costs
Input energy will be the major operating cost of a district heating system; thus, where
multiple alternatives are available, each should be carefully considered. The largest component of the energy costs will be fuel for the equipment that generates the hot water or
steam. For hot-water systems, pumping of distribution system water will also require significant amounts of energy.
It is important to also include, as part of the operating energy cost, the heat loss
within the heat distribution system; calculating the heat losses and their associated costs
is covered in Chapter 6. While distribution system heat losses may represent as little as
2%3% of total system load when the system is at capacity, the impact on annual energy
consumption will be greater, as the heat losses will persist with little or no reduction at
times when heating demand by the consumers is significantly lower.

Operation and Maintenance Costs


Aside from the major cost of energy, there will be other significant operation and
maintenance (O&M) costs that should be considered. At the planning stage a detailed
analysis of these costs is normally not warranted. A method of making a first-order
accounting for these costs is to assume they are a percentage of the capital cost of the system on an annual basis. For heat distribution systems, Phetteplace (1994) has suggested
2% of the capital cost as the annual costs, excluding the cost of heat loss and pumping
energy. For central plant O&M costs, a higher percentage of capital costs may be justified
on an annual basis, perhaps 2%3% exclusive of energy costs.
Equipment lifetimes are an important factor in an economic analysis. Chapter 37 of
ASHRAE HandbookHVAC Applications (2011) provides data on service life estimates
for boilers, pumps, and other major equipment as would be used in a district heating central plant as well as the expected life of alternative systems to district heating, such as
smaller in-building boilers. As ASHRAE (2011) points out, much of this data is quite old
and recent efforts to update it have yet to provide a large enough database to provide high
confidence estimates for many types of equipment. Currently ASHRAE maintains an
online database, ASHRAE Owning and Operating Cost Database at www.ashrae.org/
database (2013b), where users can add data and query to find the most recent statistics. At
the time of writing this book, there were significant amounts of data for boiler and chiller
life, for example, and even a limited sample of data on the lifetime of district heating
building systems. In the case of the later, it is assumed that this is from a building owners
perspective rather than a district heating providers perspective, which of course may be
identical in the case of campus systems, for example.

2.13

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

ECONOMIC ANALYSIS AND USER RATES


This section covers economic topics in more detail and assists consultants and building owners in comparing the costs of self-generated (in-building) heating plants to those
of a district heating contract or offer, since a proper analysis will include more inputs than
just energy and construction costs. The evaluation is extremely site specific and, while not
complex, will have many input parameters, some that are economically quantifiable and
others that are qualitative and cannot be assigned a monetary value yet do add/subtract
value.
For the parameters that are quantifiable, the most common method of evaluation is a
net present value (NPV) or life-cycle cost analysis (LCCA) that encompasses all the
major costs and charges associated with each option and includes the time value of money
over the life of the project or life of the contract. The following discussion is intended to
assist the evaluator in understanding the big picture and identifying all the costs to prepare a fair comparison. Typically a district heating contract is for a minimum 20-year
period and the contract includes all conceivable charges, including charges that the evaluator may not be aware of. The duration of the contract enables the district heating provider to recoup their costs for the expenses incurred in connecting the building or
reserving the capacity at the district heating plant for the buildings load. Table 2.2 summaries the key parameters that are input into a detailed economic comparison calculation.
Since there are benefits to the building owner that cant be assigned a monetary value,
the comparison between the two alternatives should be value-based. In other words, the
analysis should include not only the quantitative variables itemized in Table 2.2 but also
the qualitative variables and benefits that have intrinsic value to the proposition. The following are examples of some qualitative variables that add value to a building connecting
to a district heating system:
Reuse of the space vacated by the heating equipment since some of the mechanical and electrical space can be rented out for uses other than storage, such as
office space, a clean roof area that could be used for more sustainable purposes
such as a roof garden or pool, etc.
No plumes from boiler stacks
Increased thermal energy source reliability
More stability in energy costs
Other than a possible demand charge, customer is only billed for the energy used
(metered)
Less greenhouse gas emissions and lower carbon footprint
Not having any equipment in hot standby that is idling and using energy
Freeing up maintenance staff to perform duties other than in-building plant operations
Of course, the building owner would have to determine from the above list if any or
all of the parameters would be pertinent or valuable to his/her building(s).
Refer to Chapter 37 of ASHRAE HandbookHVAC Applications (2011) for a more
detailed explanation of preparing a life-cycle cost (LCC) calculation.

MASTER PLANNING CONCLUSIONS


Master planning is as much art as science. It is only as good as the knowledge, creativity, and interest of the individuals that complete the tasks outlined above and how
capable they are of establishing strong client/engineer relationships that lead to honest
and open communication. The owner must be free to challenge the engineer and concurrently the engineer must be straightforward in discussing sensitive issues that challenge

2.14

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Table 2.2 Summary of Economic Analysis Factors


Capital Costs
Construction costs of the building
plant vs. energy transfer station (ETS)
equipment

Includes the materials and labor for boilers, stack, piping, pumps, heat exchangers,
valving, instrumentation, and controls; the cost of electric service; the cost of
additional structural support due to equipment weight on the roof of the building; fuel
storage, etc.

Value of increased mechanical and


electrical space that would house the
plant equipment

Includes value of penthouse, basement, roof and vertical chases for flues, etc.

Value of any equipment screening

Many times municipalities require screening for any equipment mounted on grade or
on the roof.

Cost of financing

Amount of project that is financed at the loan interest rate of the duration of the loan.

Construction permits and fees

Typically a percentage of construction.

This could include the replacement or overhaul of boilers, pumps, etc. over the life of
Life of major equipment overhauls and the analysis/contract duration. If the district heating contract is for 20 years and a
replacement costs
piece of equipment must be replaced or overhauled at, for example, 15 years, this
cost must be accounted for.

Contract vs. installed capacity

The district heating contract capacity will most likely be less than the planned
installed in-building equipment capacity as dictated by the consultant due to many
reasons, but mostly oversizing. Typically the estimated peak loads can be reduced
to 70%.

Cost of redundant equipment for


emergency or standby capacity

Similar to the above, N+1 redundancy requirements would be accommodated and


added to the first cost of the in-building alternative.

Energy and Utility Costs


Electric rate

From usage of each option from energy model or other estimate.

Natural gas rate

From usage of each option from energy model or other estimate.

Water and sewer charges for steam

From usage of each option from energy model or other estimate. Water is increasingly
becoming an important resource, hence makeup water and equipment blowdown/
sewer discharge amount are estimated. It is not uncommon for this utility to have a
different escalation rate.

Operation and Maintenance Costs


Labor and benefits of operations staff
assigned to in-building plant activities

This would include any staff assigned to the duties of maintaining and operating the
in-building plant, including supervisors and overtime due to unplanned outages.

Insurance costs

Added insurance costs associated with on-site fuel-burning equipment and fuel
storage where applicable.

Spare parts and supplies

Boiler and auxiliary equipment require replacement of parts for normal maintenance
procedures including tubes, seals, bearings, etc.

Cost of chemical treatment for


steam or hot-water systems

Includes scale and corrosion inhibitors, biocides, oxygen scavengers, etc. These
costs could be considerable for steam systems.

Cost of contracted maintenance

Some owners outsource specific tasks, such as boiler maintenance and overhauls, to
service companies.

Energy and Resource Usage


Peak heating and cooling
thermal loads

Used to apply the energy demand rate and the sizing of the plant equipment (boilers,
pumps, electrical service, water service, etc.).

Annual heating usage

Used to apply the energy consumption rate of the utilities to the equipment meeting
the thermal loads.

Annual water and sewer usage

Quantify makeup water usage and blowdown discharge pertinent to the boilers.

Other Costs
Architectural and engineering design
services

Specifically for new or retrofit applications.

Fees and licenses

Air and water permits, high-pressure steam operator licenses, city franchise fees for
running piping in streets, etc.

Insurance of equipment

Typically a percentage of construction costs.

2.15

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

the owner. The most effective plans are built on communication, credibility, integrity, and
trust of the parties involved.
Plans will need updating as the master planning database is modified by growth of the
owners requirements and facilities, development of new technology or system concepts,
and delays in implementing the prioritized plan. If the master plan has been developed
with an understanding of engineering fundamentals and it is well supported by the data
gathered, as well as the system flow and control diagrams and the hydraulic modeling,
then plan updates are straightforward and relatively low in cost.
It is recommended that, at the early planning phase in a project, an experienced
designer be consulted if those undertaking the master plan and/or design do not have wide
experience with district heating systems. Early in the design process, wide and varied
experience will normally provide large dividends.

ALTERNATIVE DEVELOPMENT FOR HEAT SUPPLY


Methods of Heat Generation
District heat can be generated by heat-only boilers (steam or hot water), combined
heat and power (CHP) generation plants (or cogeneration plants), and heat pumps (using
low- to moderate-temperature geothermal and other water resources). Cogeneration is the
simultaneous production of electrical and thermal energy from a single energy source. By
capturing and using recovered thermal energy from an effluent stream that would otherwise be rejected to the environment, cogeneration systems operate at utilization efficiencies greater than those achieved when heat and power are produced in separate processes.
The prime movers of the CHP systems could be steam turbines, gas (combustion) turbines, and/or reciprocating engines. This section provides the description and guidance
for selection of these systems.

Conventional Heat-Only Boiler Plants


The development of a district heating system is typically initiated by construction of a
heat-only boiler plant. This option allows gradual district heating development with minimal capital investment. As the district heating system is expanded and the thermal load
enlarged, it could be economical to build larger, more efficient cogeneration plants that
can be fueled by natural gas, oil, coal, or renewable sources such as refuse and biomass.
Several factors must be evaluated to select a suitable boiler plant. A basic decision is
whether the plant should use hot-water or steam boilers. If the district heating system supplied by the boiler plant is strictly a heating system, the most logical choice is hot water.
Other uses may require steam, however, such as kitchen, laundry, sterilizing, process, or
cooling. With steam loads of this type, it may be desirable to generate steam for the specific requirements and use steam-to-water heat exchangers for generating hot water.
Another alternative is using hot-water generators for the district heating load and utilizing
a separate steam boiler in the building to serve the steam load. A high-temperature hotwater system can be used to generate low-pressure steam in buildings for required local
uses.
There are two basic types of boilers used in district heating plantsfire tube boilers
and water tube boilers. Both of these are normally purchased as package boilers, which
means they are factory assembled (including controls), tested, and then shipped assembled. The package boilers require only rigging and connections and are typically natural
gas or oil fired and have steam production pressures in the 100 to 250 psig (690 to
1725 kPa) range.
In general, a fire tube boiler has twice the water-holding capacity of a water tube
boiler of comparable output. Therefore, a water tube boiler reaches steaming capacity

2.16

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

more rapidly from a start than does a fire tube boiler of comparable size. Also, water tube
boilers cool down more rapidly than comparably sized fire tube boilers. These two features are usually considered desirable in O&M.
On the other hand, a fire tube boiler is more forgiving in operation, especially with
boiler water chemistry control. Poor water chemistry control causes scaling on the water
side of tubes, which reduces heat transfer and can cause tube failure due to high fire- or
gas-side metal temperatures and can happen much faster in a water tube boiler than in a
fire tube boiler.
Additionally, a fire tube boiler has a greater steam release surface and is more resistant to priming and foaming, which cause carryover of solids and moisture into the steam
line. The large volume of water and large steam release surface of a fire tube boiler make
it an excellent choice for stable or slowly varying load conditions. Rapid load changes are
best handled by the faster response of a water tube boiler.
Below 20,000 lbm/h (9070 kg/h) steam output, a fire tube boiler is typically less
expensive than a water tube boiler. The cost advantage is reduced above 20,000 lbm/h
(9070 kg/h). Fire tube boilers are not available in sizes much above 28,000 lbm/h
(12,700 kg/h). In general, water tube boilers require less boiler room space than comparably sized fire tube boilers. This is primarily due to maintenance space requirements and
not actual package space requirements.
It is generally acknowledged that field-erected water tube boilers have longer lives
than package water tube boilers, and both have longer lives than package fire tube boilers.
Field-erected water tube boilers commonly have 40-year lives, package water tube boilers
have 20- to 25-year lives, and package fire tube boilers have 15- to 20-year lives. These
life estimates assume observance of O&M practices recommended by the vendors.
Detailed information on heat-only boilers is presented in Chapter 3, Central Plant
Design for Steam and Hot Water.
Cost of Heat Supply from a Heat-Only Boiler Plant
Experience has demonstrated that use of low-temperature hot water up to 250F
(121C) for district heat supply offers the opportunity to substantially reduce the cost of
heat. Therefore, when a heat-only production option is considered, it is recommended to
use hot-water boilers for heat supply. Low-pressure fire tube hot-water boilers have high
reliability, low cost, and a maintainability record. In addition, these do not require 24hour licensed operators because of the low-pressure operation. The boiler plant capital
cost estimating methodology for a 100 MMBtu/h (106 GJ/h) hot-water boiler plant is presented in Table 2.3. The presented cost does not include any district heating piping distribution cost.
The heat production cost at plant boundary includes the capital cost component
(annualized by applying a capital recovery factor of 7.82%, which corresponds to 6%
interest for 25-year plant life), fuel cost, and an O&M component. Production cost at the
plant boundary for 1700 equivalent full-load operating hours and a natural gas cost of
$7/MMBtu ($6.63/GJ) is presented in Table 2.4. The O&M cost of $3/MMBtu ($2.84/GJ)
is based on actual operating data.
Example of a Heat-Only Boiler Plant
The development of the district heating system in Buffalo, New York, started with the
construction of a heat-only boiler plant (Griffin et al. 1987). The district system included
the steam-based 30-story city hall and a number of hot-water buildings in close proximity.
The boiler plant included three package gas-fired boilers: one 500 hp (373 kW) steam
boiler and two 800 hp (597 kW) hot-water boilers. The steam boiler supplies the city hall
with steam while the hot-water boilers supply the rest of the buildings with hot water. The

2.17

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Table 2.3 Cost Estimate for 100 MMBtu/h (105.5 GJ/h) District Heating Hot-Water Boiler Plant
Cost Estimate,
$1000

Item
Hot-water boilers (dual fuel)

2,500

Stack, breaching

125

Expansion tank

25

Variable-speed district heating pumps

150

Mixing pumps

45

Control system

150

Power supply

200

Piping, valves, meter

350

Water treatment equipment

80

Chemical shot feeder

10

Air separator

30

Oil storage tanks

100

Fuel oil delivery system

40

Equipment foundations

60

Gas supply system

55

Building

180

Water and sewer services

45

Landscaping and fencing

55

Land cost

200

Subtotal

4,400

Engineering, construction management

660

Contingency

660

Total Cost

5,720

Table 2.4 Unit Cost of Heat at a Hot-Water Boiler Plant


Components of Unit Heating Cost

Unit Heat Cost,


$/MMBtu ($/GJ)

Capital component:
$5,720,000 0.0782/100 MMBtu load factor 1700 h
($5,720,000 0.0782/105.5 GJ load factor 1700 h)

2.63
(2.49)

Fuel cost:
($7/MMBtu)/efficiency 0.8
([$6.63/GJ]/efficiency 0.8)

8.75
(8.29)

O&M

3.00 (2.84)

Total

14.38 (11.53)

2.18

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

total district heating load is 45 MMBtu/h (47.5 GJ/h). The backup for the city hall is provided by the old existing steam boiler. The boilers are equipped with economizers. The
plant includes a water treatment system and hot-water circulating pumps. The district
heating piping includes four pipes: one steam, one condensate, and two hot-water pipes.

District Heat Supply from Cogeneration Steam-Turbine-Based Stations


New District Heating/Cogeneration Steam Turbines
and Controls for District Heating Supply
Steam turbines in district heating power plants may supply a combination of space
heating, domestic hot water, and industrial loads. These loads usually vary hourly,
weekly, or seasonally and require turbines of special design to meet the particular load
duration curve while generating electric power. District heating turbines, therefore, have a
number of special features not incorporated in single-purpose turbines (Oliker 1979,
1980). The maximum combined electric and heat generation efficiency is achieved when
the turbine is designed to match the heat consumption, i.e., following the thermal load.
District heating turbines are designed for base load operation and do not supply the peak
load. Peaking boilers are usually used for this purpose.
The pressure of the steam extracted from the turbine must change to correspond with
the required district heating water temperatures. Therefore, the district heating turbine
must be designed to follow the district heating loads though it must also operate in a pure
condensing (power-only) mode. The turbine should provide economical operation
throughout the year in both district heating and power-only modes.
A steam turbine specially designed for district heating/cogeneration enables the turbine to operate across the entire performance spectrum from maximum to zero heat
extraction. The turbine has controls that stabilize operation in any mode while providing
the usual safeguards to prevent turbine overspeed and water induction. The turbine design
is limited only by boiler output and the extractions required for feedwater heaters and
other plant loads (e.g., air-ejector and deaerating feedwater tank loads).
Cogeneration turbines designed for hot-water production provide extractions from
two pressure levels and staged heating of the district hot water for higher plant efficiencies. The two pressure district heaters are arranged in series. The lower-pressure steam is
used to preheat the return water while the higher-pressure steam elevates the district water
for either direct distribution to end users or for further heating in a peaking boiler, as
shown in Figure 2.2 (Oliker 1979, 1980; Oliker and Armor 1992). Under extreme cold
weather conditions heat loads beyond the maximum capability of the district heating turbine are satisfied by peaking boilers by further raising the district heating water supply
temperature. When the heating season ends, the turbine must satisfy only the domestic
water heating requirements.
A comparison of two-stage district heating turbine arrangements is shown in
Figure 2.3 (Oliker and Mulhauser 1980). The top diagram shows a typical West European
district heating arrangement. In this design, all steam passes through the intermediatepressure (IP) section and is partially extracted for district heating when required. These
turbines use butterfly valves in the crossover pipes between the IP and low-pressure (LP)
sections in order to control heat and electrical generation.
The middle diagram in Figure 2.3 shows the district heating arrangement using a typical U.S. turbine. This machine has a separate two-flow asymmetrical LP element that
provides large quantities of extraction steam to the district heating heat exchangers from
its exhausts at two different pressures. This turbine section is located between the IP and
the LP casings. Crossover steam flow from the IP exhaust to the asymmetrical (heating)

2.19

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 2.2 New district heating/cogeneration turbine arrangement.

HP, IP, and LP = high-, intermediate-, and low-pressure sections of the steam turbine;
H1 and H2 = district heating heat exchangers; R1 = crossover butterfly control valves;
R2 = control diaphragms; R3 = bypass valves

Figure 2.3 Arrangements of specially designed district heating/cogeneration turbines.

2.20

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

LP and condensing LP sections is controlled by large butterfly valves located in the inlet
pipes of these elements. Depending on the outdoor temperature, the district heat exchangers require lower or higher pressure steam. Variation of steam extraction pressures is
achieved by adjusting the flow split between heating and condensing LP sections. During
maximum heat load conditions, the butterfly valve in the inlet to the heating LP element
is fully open while the valve in the inlet to the condensing element is fully closed. In the
pure condensing operation (at zero heat load conditions), all steam is diverted to the condensing LP section while the butterfly valve in the inlet to the heating LP section is fully
closed. It should be noted, however, that even when the crossover butterfly valves are
fully closed a small amount of steam must still pass through them for blade cooling purposes.
It can be seen that in the U.S. arrangement not all steam leaving the IP section enters
the heating LP section to be partially extracted. The flow split is adjusted by the butterfly
valves before entering the section from which district steam is provided.
The bottom diagram in Figure 2.3 shows a typical East European district heating turbine arrangement with a variable diaphragm type of control. This type of turbine also has
two extractions for heating district water in series. It has one pressure regulator, a gridtype variable control diaphragm located in the LP cylinder. The pressure in the district
heat exchanger is controlled by adjusting the position of the diaphragm. The maximum
heat load from both extractions is provided when the diaphragm in the LP element is
closed and the pressure in the controlled extraction is increased.
The thermal efficiency of a single-purpose electric generating plant with steam turbines may range between 30% and 38% depending on the initial steam parameters. The
overall efficiency of cogeneration plants with steam turbines ranges between 65% and
80%. The use of small district heating steam turbines with capacities below 10 MW is
usually not economical.
Cost of Heat Supply from a New District Heating/Cogeneration Steam Turbine
The cost of heat supply from a new district heating/cogeneration steam turbine is typically allocated between heat and electric generation. There are a number of cost allocation methodologies reviewed and approved by each states Public Service Commission.
One of a number of cost allocation methods is the physical method, wherein the fuel use
is allocated corresponding to each of the two forms of energy by properly accounting for
the physical processes involved.

RETROFIT OF SINGLE-PURPOSE ELECTRIC GENERATING STEAM TURBINE


TO DISTRICT HEAT SUPPLY
Overview
Fossil-steam-turbine-based electric generating plants located in relative proximity to
the district heating load centers (up to five miles) can be retrofitted to supply steam for
district heating systems that deliver energy (usually steam or hot water) to nearby customers. This type of cogeneration boosts the plants utilization of fuel energy considerably,
increasing its competitiveness in both electricity and thermal energy markets. Such district heating retrofits can be especially beneficial to fossil steam plants with high fuel
prices or relatively low thermal efficiency. New thermal energy sales provide added revenue. By retrofitting plants located near customers in an industrial park, university, or
downtown area, utilities can enhance energy services, helping to retain existing customers, attract new thermal energy customers, and provide new business opportunities for the
community. Such district heating systems also reduce overall emissions by replacing mul-

2.21

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

tiple heating units with a single central system. Regulators may even allow utilities to
apply emissions reductions as offsets to expand other generating facilities.
District heating retrofits of electricity-only fossil steam plants usually involve diverting steam from the turbine cycle for direct use or to feed specially designed heat exchangers that condense the steam to raise the temperature of water supplied to a district heating
system. Steam for district heating systems can be extracted at existing feedwater heating
extraction points, at new extraction points made in the turbine casing, or at crossovers
between turbine sections. Plants can also convert a low-pressure turbine to back-pressure
mode by eliminating or bypassing the last stages and exhausting the total steam flow to a
district heating condensing heat exchanger.
Retrofitting electric generating plants to provide thermal energy can increase total
plant energy utilization efficiency to 65% or more, as much of the heat previously
rejected to the condenser becomes marketable. Only in back-pressure mode operation is
electricity production appreciably reduced.
District heating cogeneration also provides electric system utilization benefits. For
utilities with winter-peaking loads, district heating can offset electrical heating loads,
reducing the need for new generating capacity or purchased on-peak power while retaining utility revenues. Plants used for peaking loads in the summer that are little-used in the
winter can be retrofitted to provide relatively high heat steam output and low-power output in the winter, improving year-round plant utilization. Further, district heating can be
used to diversify loads through such strategies as providing chilled water for cooling in
place of electricity, which flattens peak capacity needs.

Configuration and Control of Steam Turbine Retrofits for District Heating


There are a number of methods by which single-purpose steam turbine cycles can be
retrofitted to include district heating in addition to the generation of electricity (Oliker
1996):
Using existing feedwater heaters for district heating
Bypassing existing feedwater heaters and using the steam currently supplied by
the cycle to the feedwater heaters for new district heat exchangers
Using steam from the external crossover(s) of turbines
Increasing extraction flow from given stages and/or tapping into the existing
low-pressure turbine section(s)
Converting a condensing turbine into a back-pressure unit
Utilizing the condenser for district heating
To minimize the reduction in electric power output during district heating, steam
should be extracted from the turbine at the lowest pressure that will provide the required
water temperature. Additionally, when the district heating load is disconnected, the original electric output capability of the unit should be restored. The selection of an approach
from possible alternatives depends on the type of turbine, the pressures available at the
extraction stages, the maximum condenser operating pressure, the heat load, and the budgeted cost for the modification.
For steam turbines, the best retrofit uses steam from an external crossover
(Figure 2.4), thus avoiding modification to the turbine casing. This requires a butterfly
control valve in the crossover and a tee section with a new pipe and another butterfly control valve, or possibly a flow limiter, in the new line connected to the district heat
exchanger.
Naturally, the quantity of steam that can be extracted for district heating is limited by
the boiler size and minimum turbine flow requirements. The greatest overall fuel effi-

2.22

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

1 = boiler, 2 = HP turbine, 3 = IP turbine, 4 = LP turbine, 5 = condenser,


6 = district heating stage one heat exchanger, 7 = district heating stage two heat exchanger,
8 = return from district heating system, 9 = supply to district heating system,
10 = butterfly valves, 11 = bypass

Figure 2.4 District heating configuration with extractions from intermediate-pressure crossovers.

ciency can be achieved in applications where a large fraction of the turbine throttle flow is
eventually extracted. District heating system designs vary widely. Somemostly older
systems circulate steam to locations throughout the system and return condensate. Other
older systems do not return the condensate because it would be unacceptable for reuse
without extensive treatment. Modern district heating systems designed to provide commercial and residential space heating generally employ recirculated hot water. Maximum
temperatures are about 250F (121C), allowing the system to be designed with low-pressure equipment (i.e., a non-pressure vessel) throughout. The low district heating temperatures also allow the use of relatively low-pressure steam to provide heat to the system,
which, in turn, enables the greater part of the available enthalpy of the steam to be converted to electric power before extraction from the cycle.
Turbine retrofits can be accomplished in many different ways, depending on the
nature of the plant and its role in the utilitys operations. The retrofit may be performed
using a minimum-cost approach or done as a turbine cycle upgrade where every effort is
made to maximize the power and efficiency of the cycle. The modification to the turbine
cycle may be simple, utilizing a previously unused steam extraction point. Conversely, for
a large plant with a long projected useful life the retrofit may include reblading the entire
turbine, altering the diaphragms and nozzles, changing the number of stages to be used,
adding or altering feedwater heaters, and upgrading or replacing the control system. In
general, however, minimal disturbance of the turbine casing is the recommended design
approach. In cases involving an old and underperforming turbine in a plant with a useful
projected life, such a wholesale renovation may be justified.
Some significant turbine modifications may be required just to permit reliable operation with greater steam extraction. For example, drawing steam for the district heating
system can result in reduced pressures at the extraction points, which can cause unaccept-

2.23

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

able stresses on the original blades and diaphragms of the stages upstream of the extraction points. Reduced flows in the low-pressure stage downstream of the extraction points
can produce unacceptable temperature distributions. These problems generally can be
overcome with minor modifications. There are several proven methods to retrofit steam
turbines to feed district heating heat exchangers. Both steam extraction and use of a backpressure cycle have been successful. However, converting to a back-pressure cycle generally entails major modifications to the low-pressure section of the turbine, which prevents
the turbine from generating the original rated electrical output. The use of steam extraction allows more flexible operation and permits the electrical output to approach the original megawatt rating when extraction for district heating is not required.
In extraction-type designs, the number and location of points from which steam is
drawn can vary widely, depending on the original cycle and equipment design, the
required district heating temperature, anticipated power and heat loads, and economic
considerations. In smaller and older plants, equipment limitations may govern extraction
point selection. For example, configuration choices may be dictated by the availability of
existing unused extraction connections or by the observation that some stages are relatively inefficient (these are good candidates for extraction points as less power is lost by
reducing steam flow to these stages). Where there are no external crossovers, the use of
existing extraction points is strongly favored. If needed, these points may even be
enlarged. Figure 2.5 illustrates a typical configuration using existing extraction points for
heating district steam/hot water supply.
Figure 2.4 shows a typical configuration for large plants with external crossovers and
no design limitations that would restrict extraction point selection. The two-stage heaters
are both supplied with extraction steam from the exhaust from the intermediate-pressure
sections. The connections are made in the crossover lines between the intermediate- and

1 = boiler, 2 = HP turbine, 3 = IP turbine, 4 = LP turbine, 5 = condenser,


6 = district heating stage one heat exchanger, 7 = district heating stage two heat exchanger,
8 = return from district heating system, 9 = supply to district heating system, 10 = bypass

Figure 2.5 District heating configuration for a turbine with internal crossover using
modified existing extraction points.

2.24

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

low-pressure sections. Note that the intermediate-pressure sections are asymmetrical,


with the section having more stages supplying the lower-temperature heater. This permits
the maximum amount of energy to be extracted from the steam before it is diverted to the
district heating system. Such a configuration reflects the design of specially built district
heating turbines offered by major manufacturers. For retrofits to existing electricity-only
turbines, this approach could require substantial modification.
Another alternative, shown in Figure 2.6, provides steam from a low-pressure extraction point to the first-stage district heating heat exchangers and steam from the intermediate-pressure exhaust crossover to the second-stage heat exchanger. This configuration
maximizes the extraction of useful energy while providing greater steam flow to a portion
of the low-pressure section. Such a configuration requires an available extraction point of
sufficient size at a suitable pressure. Typically, in the configurations shown in Figures 2.4
and 2.5, the two district heating heat exchanger stages are approximately the same size,
which yields the best overall efficiency.
The proper apportioning of steam between the district heating heat exchangers and
turbine stages downstream of the extraction points requires an active control system to
meet the varying electric power and district heating and cooling loads. The control system
shown in Figure 2.5 is typical for a large modern cogeneration turbine. The butterfly
valves in the crossovers from the intermediate-pressure sections close with increasing
demand for district heating and cooling and are equipped with stop points to maintain a
minimum flow to the low-pressure turbines. The control network also ensures that the
valves maintain sufficient pressure in the intermediate-pressure section to prevent unacceptable temperatures in the low-pressure section. Bypass valves around the district heating and cooling heat exchangers allow the system to maintain proper district heating
system flow and temperature conditions during periods of reduced heating demand. In

1 = boiler, 2 = HP turbine, 3 = IP turbine, 4 = LP turbine, 5 = condenser,


6 = district heating stage one heat exchanger, 7 = district heating stage two heat exchanger,
8 = return from district heating system, 9 = supply to district heating system,
10 = butterfly valves, 11 = bypass

Figure 2.6 District heating configuration with extractions from


intermediate-pressure crossover and low-pressure stage.

2.25

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

cases where external crossovers are not available, valves in the extraction lines to the heat
exchangers represent the simplest control alternative, although internal valving arrangements may provide better performance.
In addition to providing for steam extraction points, a district heating retrofit design
must accommodate the mechanical effects of reduced flows, altered pressure differentials,
and unbalanced thrusts. Turbine internals may have to be modified to both address these
effects and maintain reliable service. For example, the extraction of additional steam for
district heating generally produces higher than normal pressure differentials across the
blades and diaphragms upstream of the extraction point, potentially requiring replacement of these components. Reduced flow to the low-pressure stages can raise operating
temperatures; therefore, added provisions for expansion may be needed even if the control system maintains the requisite minimum flow. Steam extraction for district heating
can also lead to increased asymmetrical thrust loads, necessitating adjustments at bearings. Fortunately, these modifications can normally be made at a moderate cost.
Changes in the feedwater system may be needed to achieve optimum performance
and accommodate reduced extraction steam for feedwater heating as well as to provide
for efficient condensate return from the district heating exchangers to the feedwater system. Designers should also review the consequences of contamination by heat exchanger
leaks and the need for any additional treatment equipment.
Reduced steam flow to the condenser can also necessitate mechanical changes. For
example, with lower flow rates it may be possible (and usually advantageous) to operate
at reduced back pressure. Such operation may create a need for additional vacuum pumping capacity to accommodate the higher specific volumes and increased in-leakage resulting from lower-pressure operation.
In developing a retrofit design for supplying steam to a district heating system, physical plant layout is also of prime importance. Not only must space be available for the new
extraction lines and district heating heat exchangers, but the spatial relationships must be
appropriate. Locations and arrangements must allow for drain flows to be gravity-assisted
where possible. Adequate net positive suction head (NPSH) must be provided for the various condensate return pumps. Any additions to the feedwater heating system must still
allow adequate plot space for tube-pulling and turbine lay-down areas.

Cycle Efficiencies
District heating retrofits reduce steam flow to the stages downstream of extraction
points and decrease the feedwater heating. The associated mechanical and thermodynamic effects resulting from these changes must be evaluated when selecting a retrofit
configuration.
The principal thermodynamic effect of district heating retrofits is the dramatic
decrease in the waste heat rejected from the cycle. Because efficiency is most sensitive to
the ratio of extracted heat to power generated, when only a small portion of the steam is
extracted the overall efficiency improvement will be minor. Of course, much greater efficiency improvements can be obtained when a significant portion of the steam flow is
extracted for district heating.
The loss of electric generating capacity from reduced flow in downstream stages is
inevitable. To minimize capacity loss and maximize generating efficiency, extractions
should be made at the point where steam conditions closely match those required by the
district heating heat exchangers. This maximizes power generation by the steam before it
leaves the turbine. However, the ideal extraction points are not always accessible, and
steam may have to be extracted at higher temperatures and pressures than necessary for
district heating. For example, the intermediate-pressure/low-pressure section crossover

2.26

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

lines are often the most convenient point for extraction, but the temperatures and pressures may be well above those required by the district heating heat exchangers.
This mismatch can be mitigated by passing the steam through depressurizing and
desuperheating equipment, which makes the steam better suited for the heat exchangers
but does not recover the energy that would otherwise be available for power generation.
Alternatively, it may be more economical to install an additional back-pressure turbine
between the extraction points and the heat exchangers.
Diverting steam for district heating can reduce the amount available for feedwater
heating. However, this reduction is partially offset by the higher enthalpy condensate
returned from the district heating heat exchangers relative to the condensate from the
main condenser. Nevertheless, modifications to the feedwater system may be required. If
they are, such modifications could range from adding a simple heater bypass to the installation of additional heaters. Factors affecting the necessary modifications include design
objectives (e.g., minimizing first cost versus maximizing fuel efficiency), plant size,
expected load patterns, remaining plant life, fuel costs, and plant design details. An
important design consideration is the incorporation of energy from the heat exchanger
condensate. Oliker (1996) recommends that drain coolers and heat exchanger condensate
return points downstream of the main condensers be used judiciously.
The reduced steam flows and altered pressure profiles resulting from mechanical
modifications required for reliable operation and alterations to the feedwater system generally have only minor effects on cycle efficiency. As a result, most plants retrofitted for
district heating cogeneration can still essentially match their pre-retrofit electrical outputs
and efficiencies at zero extraction. If the retrofit is accompanied by upgrades or repairs,
an overall improvement in turbine performance may be achieved.

Examples of Power Plants Retrofitted to Cogeneration


In order to demonstrate the technology of retrofitting electricity-only power plants
to provide steam for district heating, two examples of such retrofits are presented: one
for a single-casing steam turbine and one for a double-casing steam turbine.
Retrofit of a Single-Casing Turbine
The selected power plant included four coal-fired boilers and two 25 MW steamdriven turbine-generator units, both with non-reheat turbines (Oliker and Champ 2000).
One turbine was selected for district heating modification. This turbine was a 25,000 kW,
3600 rpm, 15-stage single-flow condensing unit designed to operate at 850 psig
(5865 kPa) steam pressure, 900F (482C) temperature, and 1.5 in. Hg (3.8 cm Hg) condenser pressure. It has a throttle flow of 238,072 lbm/h (107,966 kg/h). The heat balance
of the unit is presented in Figure 2.7. The turbine had one blanked-off extraction point at
the 11th stage. Steam was extracted from this turbine extraction point for use in a new district heat exchanger for loads up to 7 MWt; heating loads in excess of 7 MWt are served
with additional steam from the auxiliary steam header connected to the existing auxiliary
heat exchanger, which is arranged in series with the new district heating heat exchanger.
The new district heating heat exchanger operates throughout the year, providing hot
water for both space heating and domestic use. The auxiliary heater provides peaking and
backup supply. A schematic of the district heating retrofit is presented in Figure 2.8. During peak heat load operation, the return water temperature is 160F (85C) with a district
heating water supply of 250F (121C). The district heating water flow rate corresponding to the peak heating load is 498,000 lbm/h (225,843 kg/h). The maximum extraction
flow available from the turbines 11th stage provides heating for 379,000 lbm/h
(171,877 kg/h) of district circulating water to 223F (106C). At the maximum heat load

2.27

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 2.7 Heat balance of a single-casing turbine.

conditions, 119,000 lbm/h (53,967 kg/h) of district heating water bypass the district and
auxiliary heat exchangers. The balance of the flow, which amounts to 379,000 lbm/h
(171,877 kg/h), passes through the district heat exchanger, is raised in temperature to
223F (106C), and subsequently enters the auxiliary heat exchanger, where it is
increased in temperature above 250F (121C), such that when mixed with the bypassed
water the final temperature is 250F (121C), the design condition.
The maximum operating pressure of the district heat exchanger is about 20 psia
(138 kPa). The auxiliary heat exchanger operates approximately one-third of the year.
The maximum operating pressure of the auxiliary heat exchanger is 60 psia (414 kPa)
during district heating operation. The maximum steam flow to the auxiliary heat
exchanger during district heating operation is 18,900 lbm/h (8571 kg/h).
When the heat load is less than 7 MWt, the auxiliary heat exchanger is out of service
and the maximum extraction steam flow to the district heat exchanger is 23,813 lbm/h
2.28

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Figure 2.8 District heat supply from the single-casing steam turbine.

(107,992 kg/h) at an extraction stage pressure of 21.8 psia (150.4 kPa). The maximum
electrical load reduction of the turbine-generator is about 1.24 MWe during district heating operation. This reduction corresponds to a conversion ratio of 5.6:1 (reduction of one
unit in electric generation results in a 5.6 unit gain in heat generation), which is a very
good ratio for an existing single-purpose unit. Modifications to the turbine were not
required since the redistribution of extraction flows is minimal and all existing feedwater
heaters remain in service without modification.
Retrofit of a Double-Casing Turbine
The selected power plant is equipped with two steam turbines and two coal-fired
boilers (McIntire et al. 1994). The steam turbine selected for retrofit is of the tandemcompound, two-cylinder, impulse-reaction type, designed for steam throttle conditions
of 850 psig (5,865 kPa), 900F (482.2C), and 1.25 in. Hg (3.18 cm Hg) absolute back
pressure. The output of the generator under normal operating conditions is 33,360 kW.
The two-cylinder arrangement of the turbine consists of a high-pressure cylinder and a
low-pressure cylinder connected by an external crossover. The heat balance diagram of
the turbine is presented in Figure 2.9.
To supply a district heating system, the best available turbine extraction location is a
location where the pressure is 40 psia (276 kPa)the external crossover between the
high-pressure and low-pressure sections. The next highest pressure of 104.1 psia
(718.3 kPa) is too high; the next lowest pressure of 12.3 psia (84.9 kPa), in the low-pressure section, is too low.

2.29

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 2.9 Heat balance of the double-casing turbine.

A steam cycle analysis confirmed that sufficient steam could be extracted from the
turbine to meet the heating demand. These estimates were performed by computer modeling of the cycle heat balance.
Extraction of steam from the turbine crossover results in a reduced steam flow to the
low-pressure end of the turbine. Sufficient steam must pass through the low-pressure turbine section to cool the last stages of blades. While specially designed district heating turbines can operate with a minimum of 5% cooling steam for their last stages, retrofits of
existing single-purpose turbines typically use 15% to 20% of the throttle flow for cooling
these stages. The amount of steam that could be extracted from the turbine was determined based on providing the maximum district heating load of 150 MMBtu/h (158.3 GJ/
h), or 44 MWt. For a turbine throttle flow of 301,867 lbm/h (136,897 kg/h) and 20% steam
for last turbine stage cooling, the required minimum steam flow to the condenser is
60,300 lbm/h (27,346 kg/h).
Because of the lower condensate flow during the maximum heat load mode, the
extraction steam flow to feed water heater No. 3, the lowest pressure heating stage, is significantly reduced. As a result of the pressure difference between the new district heat

2.30

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Figure 2.10 Retrofit diagram of the double-casing steam turbine.

exchanger and the deaerator and subcooling of the district heat exchanger drains, the
drains can flow to the deaerator without drain pumps even though the deaerator is located
at a higher elevation than the district heat exchanger.
During maximum district heating load operation, the reduced steam flow to the condenser will:
reduce the condenser heat load, which must be removed by the circulating water;
reduce the stage pressure at the lowest extraction, which accounts for the lower
condensate temperature to the deaerator; and
reduce the electric output.
The new electric output of the unit is 24,637 kW during maximum district heating
load operation. The reduction in electric load is thus 8,723 kW, which corresponds to a
conversion ratio of 44,000/8723 = 5.04:1, which is a good ratio for an existing singlepurpose unit. This cogeneration ratio remains fairly constant over the load range.
Flow diagrams for the station district heating retrofit and the supply from the doublecasing steam turbine are presented in Figures 2.10 and 2.11. The proposal was to install
two heaters for the production of hot water for the district heating system. These are an
extraction heater and an auxiliary heater that provides backup to the extraction heater.
Both heat exchangers are sized to enable operation from the turbine crossover and boiler
steam headers for backup purposes.

2.31

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

AV = flow control, based on analyzer read; PRV = pressure-reducing valve; PCV = pressure control valve;
TCV = temperature control valve; LC = level control; FE = flow element; AE = analyzer (O1);
PE = pressure element; TE = temperature element; TI = temperature indication;
M = pump with constant-speed electric motor; VS = pump with variable-speed electric motor

Figure 2.11 District heating supply from the retrofitted double-casing steam turbine.

The drains from the district heat exchanger are routed to the deaerator at about 10F
(5.6C) below the temperature of the water in the deaerator. The large quantity of drain
flow from the district heat exchanger at this relatively high temperature compensates for
the much lower temperature and reduced condensate flow to the deaerator during maximum district heat load operation compared to pure condensing operation. This compensation in temperature enables the feedwater flow leaving the deaerator to remain nearly
constant in both the maximum heat load mode and the pure condensing mode.
Steam from a turbine crossover, shown in Figure 2.11, implies cogeneration and is the
source of steam to the heaters during normal operation. In the event that steam from the
crossover becomes unavailable, steam from the existing boiler headers is made available.
Steam from these boilers could be directed to either the extraction or the auxiliary heater
and would be controlled by a pressure-reducing valve in series with a desuperheating
valve. The extraction steam line from the crossover contains a flow element, motor operated shutoff valve, and non-return valve. The shutoff valve is designed to close when the
measured extraction flow exceeds the allowable steam flow to protect the turbine and
maintain downstream turbine steam flow requirements. The non-return valve prevents
back flow to the turbine in the event that the heater pressure exceeds the turbine extraction
pressure. A butterfly-type pressure control valve should be installed in the crossover to

2.32

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

maintain a set pressure. Each heater is equipped with a hot-well to collect condensate and
provide a level control to operate the condensate return pumps. Condensate from the heaters may be routed to several destinations. The primary sink for the condensate is the normal makeup to the deaerator. Use of deaerators, rather than a return to the condensers, for
the return of condensate to the cycle is thermodynamically advantageous. An emergency
drain from the heaters to the condenser is provided to establish a secondary path for the
condensate. The valves are intended to prevent water induction back into the turbine in
the event that the heater shell floods. An optional deaerator bypass arrangement is
included for emergency operations during scheduled deaerator tank inspection.
The district heating hot-water circulation system is a closed loop. Water discharges
from the district heating pumps and is heated in either the extraction or auxiliary heater.
Temperature control on the water side of both heaters is accomplished by varying the
flow through the heaters, either by a valve in the discharge or by a bypass around the heaters. Valving enables isolation of each heater from the system. The heated water then circulates through the underground distribution system to consumers. The water returns to
the suction side of the district heating pumps.
Suction pressure at the district heating pumps is maintained by two pressurizing
pumps. The pressurizing pumps take suction from the storage reservoir, which is near
atmospheric pressure. The discharge side of the pressurizing pumps is the point of zero
pressure change in the hot water system and provides the required pressure to prevent
flashing in any part of the system. This pressure is 6570 psig (448.5483 kPa). The pressurizing pumps provide positive suction pressure to the district heating pumps. Since it is
vital that these pumps operate continuously, their electric motors are generally interconnected to a diesel engine backup. Discharge flow from the pressurizing pumps circulates
back to the storage reservoir through a set of pressure control valves, which control the
pressure in the suction header to the district heating pumps.
The steam to the district heating deaerator, which operates near atmospheric pressure,
is supplied from an existing steam header in the plant. A pressure control valve is
installed to limit the deaerator operating pressure.
The water source is the existing softened water makeup. Based on the level in the
storage reservoir, the level control valve in the water makeup line will open or close to
maintain a setpoint. A pressure control valve is installed in the line to protect the deaerator from excessive pressure. This source of water is principally for makeup to the hotwater district heating system and will normally have minimal use.
Deaerator steam flow will result in additional hot-water makeup, tending to overflow
the deaerator storage tank. A control valve is placed at the drain connection from the
deaerator to maintain the water level. This overflow is directed to the water storage tank
or to the waste sump.
Each extraction heater is sized for 100% of the 150 MMBtu/h (158.3 GJ/h) thermal
load and will operate at the pressure in the crossover. Steam from the boilers is reduced in
pressure and temperature to simulate these conditions in the unit crossovers to enable
proper operation of the heaters. The auxiliary heater is located in a parallel steam extraction arrangement and is intended to provide backup to the extraction heater. Flexibility
exists to supply either heat exchanger. Under normal operating conditions, and based on
the flow and temperature requirements of the district heating system, steam flow to the
heater is controlled by the butterfly valve in the crossover.
Each heater is provided with a hot-well for controlling the suction pressure to the
condensate return pumps. Each heater shell is equipped with a vent to purge noncondensibles during start-up and normal operation. These vents are routed to the condenser where
they will be ejected. Each heater shell has two drains, one leading to the condenser and

2.33

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

the other to the deaerator. Condensate is normally collected and pumped to the deaerator
since this method improves the heat rate by an estimated 0.6% at 100% thermal load. The
drain to the condenser ensures fail-safe operation.
A source of softened water provides makeup to the system, thus ensuring removal of
those compounds that form scale on the heating surfaces. However, secondary scale
results when iron reacts with oxygen in the system. Thus, it is necessary to treat the water
through chemical and nonchemical techniques to remove any noncondensibles.
A storage tank provides an expansion volume for the hot-water district heating system.
It is sized according to the estimated pressure and temperature differentials in the system.
Attached to the top of this tank is an atmospheric deaerator sized to deaerate both makeup
to the system and a side stream of the systems normal flow. Since the deaerator operates at
near atmospheric pressure, the pressure in the storage reservoir is also near atmospheric.
Since the storage reservoir, as part of the deaeration system, is not pressurized with either
steam or nitrogen, the system static pressure is established by the pressurizing pumps.
Three 50% district heating pumps are installed to distribute the hot water to the district heating system. Two pumps are variable speed to compensate for the difference in
flow requirements between the summer and the winter loads. Two 100% pumps are
installed for returning condensate from the extraction heat exchangers to the deaerator.
Operation of these pumps is controlled by the hot-well level in the heat exchangers.
Cost of Heat Supply from the Retrofitted Power Plant
Table 2.5 presents the CHP plant capital cost estimate for the district heating system
supplied with heat from the retrofitted double-casing steam turbine plant example. The
primary components and modifications are the extraction and auxiliary heat exchangers;
the turbine crossover modifications; the condensate receiver and return pumps; the district
heating distribution pumps; the deaerator and storage reservoir; the pressurization pumps;
and the associated valves, piping, and instrumentation. The system components were
sized for a peak thermal load of 150 MMBtu/h (158.3 GJ/h) supply. The presented cost
does not include any district heating piping distribution cost.
The heat production cost at the plant boundary (Table 2.6) includes the capital cost
component (annualized by applying a capital recovery factor of 7.82%, which corresponds to 6% interest for 25-year plant life) for plant retrofit to cogeneration, an electric
penalty that is a fuel component charged to district heating for reduced electrical output
from the plant (8,723 kWe), peaking boiler fuel for use when the turbine is off-line, and an
O&M component to account for incremental O&M.

District Heating/Cogeneration from Stationary Gas Turbines


Gas-turbine district heating/cogeneration has flourished in the last 20 years. Major
performance improvements have enabled cogenerators to meet Public Utility Regulatory
Policies Act (PURPA) requirements, resulting in a proliferation of projects. The industry
has grown in response to improvements in engine design and performance and the growth
of independent power producers. Several technologies are used to meet district heating
efficiency and environmental standards. Figure 2.12 is a schematic of these technologies
and their integration with the gas turbine plant (Oliker 1981, 1982; Oliker and Silaghy
1987). To capture the substantial waste heat liberated by the gas turbine, a heat recovery
steam generator (HRSG) is installed at the turbine exhaust. The generator usually has two
or three steam pressure levels for efficient heat recovery from the turbines exhaust gases.
A duct burner is often installed at the entrance to the HRSG. The duct burner adds temperature to the exhaust gases at high combustion efficiency to provide superheat to the
high-pressure steam that can be generated, and it provides the HRSG with a backup ther-

2.34

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Table 2.5 Capital Cost Estimate


Item

Cost Estimate,
$1000

Extraction heat exchanger

225

Auxiliary heat exchanger

225

Heat exchanger support steel

45

Variable-speed district heating pumps

380

Pressurizing pumps

30

Condensate return pumps

45

Pump concrete pads

30

Motor control panel

68

Power wiring

53

District heating deaerator

52

Storage reservoir

75

Deaerator support steel

30

Reroute of existing piping and valves

550

New crossover piping section

83

Hot-water central control system

70

Plant Btu meter

18

Control modifications

75

Crossover diverter valve

38

Extraction line isolation valve

24

Extraction line non-return valve

34

Extraction line drain valves

18

Extraction heater steam control valve

56

Extraction heater drain control valve

28

Auxiliary heater steam control valve

70

Auxiliary heater drain valve

28

Feedwater bypass control

88

Deaerator steam control valve

14

Deaerator makeup control valve

10

Deaerator side stream control valve

12

Pressurizing pump recirculating control valve

13

Condensate pump control valve

17

Instrumentation/miscellaneous control/wiring

187

Plant Modification Subtotal

2,691

Engineering, construction management

404

Contingency

405

Total Modification Cost

3,500

2.35

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Table 2.6 Unit Heat Production Cost


Components of Unit District Heating Cost

Unit District Heating Cost,


$/MMBtu ($/GJ)

Capital component:
$3,500,000 0.0782 (capital recovery factor)/150 MMBtu/h load factor 1700 h
($3,500,000 0.0782 [capital recovery factor]/158 GJ/h load factor 1700 h)

1.07
(1.01)

Electric generation penalty:


8723 kWe $0.08/kWh/150 MMBtu/y
(8723 kWe $0.08/kWh/158 GJ/y)

4.65
(4.40)

Peaking fuel component:


150 MMBtu 400 h x $3.0/MMBtu (coal)/efficiency 0.8 150 MMBtu/h load factor 1700 h
(150 MMBtu 400 h x $2.84/GJ [coal]/efficiency 0.8 158 GJ/h load factor 1700 h)

0.88
(0.83)

O&M:
$3,500,000 0.05/150 MMBtu/h load factor 1700 h
($3,500,000 0.05/158 GJ/h load factor 1700 h)

0.69
(0.65)

Total

7.29 (6.89)

Figure 2.12 District heating supply from a combined cycle gas turbine plant.

mal source when the gas turbine is off-line. Typically a selective catalytic reduction
(SCR) system for emission control is also located inside the HRSG.
The high-pressure steam generated in the HRSG can be used for several purposes. It
can be supplied to a steam turbine to generate electricity to add to that produced by the
gas turbine; hence the term combined cycle. The high-pressure steam can also be injected
in the gas turbine for power enhancement (steam-injected gas turbines) or to control emissions of nitrogen oxides (NOx). Steam generated in the low-pressure sections of the
2.36

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

HRSG can be used to supplement the steam turbine or used for other process-related
tasks. The steam turbine can be designed with extraction ports for district heating and
other auxiliary loads. An economizer coil can be installed at the tail of the HRSG to preheat makeup water for a district heating system.
In applications where water recovery is important, there are techniques to condense
the water vapor in the exhaust gas, particularly when steam or water must be injected in
the gas turbine for performance reasons and substantial makeup is required.
To account for the behavior of the gas turbine at high and low ambient temperatures,
a heater and/or chiller is often installed to optimize the electric output of the gas turbine.
This device is installed in the inlet-air duct to the turbine. The heater element is supplied
with low-pressure steam from the HRSG or the steam turbine extraction port. Many
installations use chilled water produced in electric or absorption chillers for cooling. In
some cases, a refrigerant such as ammonia is expanded directly in the cooling coil insert
of the inlet-air duct.
The overall efficiency of cogeneration plants with combined cycles ranges between
70% and 85%.
Cost of Heat Supply from a New Gas-Turbine-Based Cogeneration Plant
The methodology of estimating the cost of heat supply is demonstrated for a combined cycle power plant with a nominal capacity of 53 MW consisting of a 40 MW gas
turbine and a 13 MW steam turbine. Systems of this size are usually installed at large university campuses or airports. Steam for district heating is extracted from the steam turbine
and supplied to heat exchanger to heat the district water. The capital cost estimate for this
CHP plant retrofit is presented in Table 2.7. The presented cost does not include any district heating piping distribution cost. The peak heat supply from the steam turbine is 100
MMBtu/h (105.5 GJ/h). The total capital and operating costs of the plant should be allocated between the cogenerated electricity and heat. The methodology for such allocation
is presented in Table 2.8.
Heat Supply from a Small Gas-Turbine-Based Cogeneration Plant
Figure 2.13 presents a heat balance diagram of a 3.5 MW small gas turbine equipped
with a waste heat recovery boiler with a duct burner. The heat recovery boiler generates
steam that is supplied to the existing district heating system. Table 2.9 presents the performance data and Table 2.10 provides the capital cost of the unit. Please note the capital
cost does not include a number of system components that are already available in the
existing district heating plant (deaerator, water treatment system, feedwater pumps, building and control room, and electric components).

Reciprocating-Engine-Based Cogeneration/District Heating Systems


Reciprocating engines fired with fuel may be used to drive a generator to produce
electrical power. The waste heat available in engine exhaust is used in a heat recovery
boiler to generate hot water or steam for district heating. The engine jacket cooling water
heat exchanger and lube-oil cooler are other sources of waste heat recovery to produce
hot water. The overall cogeneration system efficiency is typically 75% to 85%. The reciprocating engines are available with low-, medium-, and high-speed options. More information about reciprocating-engine-based cogeneration systems is provided in ASHRAE
HandbookHVAC Systems and Equipment (2012).

Large Heat Pumps


A heat pump works by extracting heat from a low-temperature sourcewater, air, or
the groundand using it to provide useful energy for district heating (Figure 2.14). In

2.37

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Table 2.7 Cost Estimate for District Hot-Water Supply from a 53 MW Combined Cycle
Cogeneration Plant
Item

Cost Estimate,
$1000

40 MW package gas turbine

19,700

Heat recovery steam generator (HRSG)

8,670

13 MW steam turbine

3,890

Selective catalytic reduction (SCR) system

1,250

Natural gas compressor

1,950

Demineralization package

618

Backup boiler

836

Hot-water convertors

465

Deaerator/storage tank

130

Pressurizing pumps

35

Condenser

380

Cooling tower

520

Fuel oil storage tank

300

Condensate return pumps

110

High-pressure boiler feedwater pumps

280

Intermediate-pressure/low-pressure boiler feedwater pumps

140

Deaerator heater

170

Mechanical and piping installation

1,870

Electrical/instrumentation and control installation

2,100

Civil/structural

2,900

Subtotal Plant Cost


Electrical interconnection
Total Direct Cost

46,300
1,700
47,960

Engineering/construction management

7,200

Contingency

7,240

Developer cost

3,500

Total Project Cost

65,900

Europe, a large number of electric-driven water-to-water heat pumps are used to supply
the base load for district heating systems.
Heat pumps use a refrigerant with a low boiling point that circulates in a closed system and is vaporized and condensed by the equipment. The main components of a heat
pump are the evaporator, the compressors, the condenser, the economizer, and the expansion valves. The pressure in the evaporator is low, causing the refrigerant to boil and
change to gas at low temperature. The compressors raise the pressure of this gas and
increase its temperature to about 215F (101.7C) depending upon the refrigerant used.
The hot refrigerant gas is cooled in the condenser, giving up its heat to the circulating district water, and leaves the condenser at a temperature of 165F to 185F (73.9C to 85C).
The gaseous refrigerant that has become condensed to the liquid state passes through the
economizer and expansion valves. The liquid then flows into the evaporator to complete

2.38

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Table 2.8 Breakdown Unit Cost for the 53 MW Cogeneration Plant


Cost Component

Data

Annual Cost,
$

Capital component
Capital cost of project, $
Capital recovery factor

65,848,000
0.0782

Maintenance

5,149,000
3,520,000

Peaking boiler fuel


Fuel quantity, MMBtu (GJ)
Unit fuel cost, $/MMBtu ($/GJ)

125,261 (132,150)
6.00 (5.69)

Peaking boiler fuel cost

752,000

Gas turbine fuel cost


Fuel quantity, MMBtu (GJ)
Unit fuel cost, $/MMBtu ($/GJ)

3,140,689 (3,313,426)
6.00 (6.33)

Gas turbine fuel cost

18,844,000
28,265,000

Total Annual Cost


Revenue
Heat generation
Quantity, MMBtu (GJ)
Unit heat cost, $/MMBtu ($/GJ)

846,742 (893,313)
11.70 (12.34)

Steam sales

9,907,000

Electric generation
Quantity, MWh

407,882

Unit electric cost, $/MWh


Electricity sales

45
18,358,000

Total Annual Revenue

28,265,000

the cycle. The largest heat pumps already in operation have a capacity of 120 MWt. The
use of large heat pumps is typically justified when low-cost electricity is available. The
economics of these installations is also very sensitive to the cost of displaced gas or oil.
A widely used source of thermal energy is the effluent of city sewage plants (StalLaval 1983). The heat pump uses heat from treated wastewater with temperatures of 50F
to 70F (10C to 21.1C) and delivers district heating with water temperatures of 160F to
180F (71.1F to 82.2C). The coefficient of performance (COP) for such plants ranges
between 3.5 and 5. Equally suitable is the reject heat from power plants and industrial
processes. Use of river, sea, and lake water as a thermal energy source may also be economical. In addition to economic advantages, the use of heat pumps results in reduction
of carbon dioxide (CO2) release to the environment (in comparison with conventional
combustion technologies).

INTEGRATION OF HEATING, COOLING, AND ELECTRIC GENERATION


The district heating systems discussed in this chapter include combined electric and
heat generation (cogeneration or CHP) systems (IEA 1996). Addition, or integration, of
cogeneration with district cooling can offer a number of additional advantages, including
the following:

2.39

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 2.13 Diagram of the 3.5 MW gas-turbine-based cogeneration system.

Improved opportunities for reducing peak electric demand through nonelectric


cooling systems
Cost avoidance due to reduced requirements for electricity distribution infrastructure serving downtown areas
Improved ability to produce chilled water from plant capacity that is underused during the summer
Improved energy efficiency resulting from optimal loading of chillers (and, as
applicable, CHP) equipment
Improved ability to use low-temperature heat sinks for condenser cooling
Chilled water can be generated by using the cogenerated steam, hot water, or electric
energy by the methods discussed in the following sections. For more detail on chilledwater generation and district cooling, refer to the companion book to this guide, District
Cooling Guide (ASHRAE 2013a).

Centralized Chilled-Water Generation by Thermal Energy


Chilled water is generated by extracted steam using steam-driven or absorption chillers at the cogeneration plant boundary and distribution of chilled water by a separate
chilled-water piping system. In this case, the district energy system will include a fourpipe distribution system: two pipes for heating and two pipes for cooling. In order to provide effective chilled-water generation, the steam extracted from the cogeneration unit
should be at much higher pressure than that for the district heating supply, which will

2.40

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Table 2.9 Performance of the 3.5 MW Gas-Turbine-Based Cogeneration System


Gas Turbine
Gross output at ISO 2314 conditions (2009)

3510 kW

Site ambient temperature for performance analysis

59F (15C)

Site elevation for performance analysis

45 ft (13.7 m)

Site ambient relative humidity for performance analysis

60%

Turbine inlet pressure loss

4.0 in. H2O (10.2 cm H2O)

Turbine outlet pressure loss

10.0 in. H2O (25.4 cm H2O)

Turbine fuel consumption at specified site conditions (low heating value)

42.5 MMBtu/h (44.8 GJ/h)

Gross output at specified site conditions

3359 kW

Gas compressor power consumption

39 kW

Turbine auxiliary power consumption

10 kW

Total auxiliary power consumption

49 kW

Net turbine power production

3310 kW

Black start kilowatt requirement (turbine generator set only)

200 kW

Heat Recovery Steam Generator (HRSG)


Process steam pressure

235 psig (1620 kPa gage)

Process steam temperature

401F (205C)

Condensate return

60%

Condensate temperature

212F (100C)

Makeup water temperature

70F (21C)

Steam contributed by gas turbine

18,762 lb/h (8509 kg/h)

Steam contributed by ductburners

61,238 lb/h (27,772 kg/h)

Ductburner fuel consumption (low heating value)

60.0 MMBtu/h (63.3 GJ/h)

Steam flow to process

80,000 lb/h (36,281 kg/h)


Net Gas Turbine Nominal Output with Inlet Cooling

Ambient Temperature,
F (C)

Gas Turbine Output,


kW

Gas Consumption,
MMBtu/h (GJ/h)

0 (17.8)

4136

49.8 (52.5)

20 (6.7)

3891

47.4 (50.0)

40 (4.4)

3631

44.9 (47.4)

60 (15.6)

3377

42.5 (44.8)

80 (26.7)

3092

40.1 (42.3)

100 (212)

2814

37.8 (39.9)

result in a substantial reduction of electric generation by the cogeneration plant. The economics of such system will be influenced by the seasonal demand of chilled water and
electric power and should be evaluated for each site. Centralized chilled-water systems
serve a large area through one integrated chilled-water distribution system fed by one or
several larger cogeneration plants using any type of chiller technology.

Centralized Chilled-Water Generation by Electric Energy


Generation of chilled water by electric-driven chillers may be combined with cooling
storage supplied with electricity from the cogeneration plant during off-peak hours. In
this case, the district energy system will include a four-pipe distribution system: two pipes
for heating and two pipes for cooling.

2.41

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Table 2.10 Capital Cost Estimate for the 3.5 MW Gas Turbine Cogeneration Plant
Item

Cost,
$1000

Dual fuel-fired gas turbine generator set

2,674

Heat recovery steam generator (HRSG) with duct burners

1,151

Gas compressor

303

Gas filter/separator

213

Liquid fuel centrifuge

76

Plant control system

142

Switchgear and auxiliary power transformer

191

Utility tie-in

213

Air compressor

104

Emission control equipment

440

Continuous emission monitoring system

138

Gas turbine inlet cooling

150

Commissioning parts, startup and site testing

156

Equipment shipping

101

Plant construction

1,691

Engineering

390

Construction management

180

Taxes

379

Contingency

1,300

Total

9,992

Figure 2.14 Heat pump/district heating diagram.

2.42

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Decentralized Chilled-Water Generation by Thermal Energy


Generation of chilled water at the end-user buildings is accomplished by using steamdriven chillers supplied from the steam distribution piping system or by absorption chillers activated with hot water or steam from the district heating distribution system. This
reduces the investment in chilled-water distribution but will require that higher temperatures be used in the district heating system than the lowest possible temperatures at which
heating-only services can be supplied.
Examples of Integrated District Energy Systems
City of New York
For many years, the Con Edison Company of New York has distributed high-pressure
steam (200 and 400 psig [1378 and 2756 kPa]) throughout a hundred-mile piping system
to about 1700 buildings located in Manhattan, New York. The peak heat demand of the
system is about 3500 MWt. A substantial amount of steam is cogenerated. The steam system supplies district heating and decentralized cooling (about 700,000 cooling tons
[2,461,000 kW]) to the buildings utilizing steam-driven and absorption chillers.
City of Gothenburg, Sweden
Gothenburg is served by a hot-water district heating system. The hot-water temperature varies from 170F (78C) in summer to a maximum of 250F (121C) during the
winter. The system derives its energy from cogeneration, waste heat, heat recovered from
wastewater via heat pumps, and heat from refuse incineration. Supply of district cooling
services was initiated in 1995. Decentralized cooling is provided with absorption chillers
driven by the hot-water district heating system with a utilization time of 1000 hours per
year. In order to increase the efficiency of cooling generation, the district heat temperature in the summer is maintained at 195F (90.6C).
City of Seoul, South Korea
Korea District Heating Corp. (KDHC) operates a district hot-water system supplying
the city customers with a peak heat demand of about 6000 MWt. District hot water has a
supply temperature of 240F (115.6C) in the winter and 205F (96.1C) in the summer.
Approximately 90% of the heat is produced in combined cycle and steam turbine cogeneration facilities. Since 1992 KDHC has been providing decentralized district cooling,
using the heating network to deliver hot water to absorption chillers located in customer
buildings.
City of Trenton, New Jersey
Electricity and high-temperature district hot water are cogenerated by two 6 MWe
diesel engine generator sets. Most of the district heat is supplied as 350F (176.7C) hot
water, but 220F and 400F (104.4C and 204.4C) heating are also provided through
separate supply and return pipes. The system provides district heating and decentralized
chilled water utilizing absorption chillers located in the buildings.
Cities of Chemnitz and Mannheim, Germany
Chemnitz and Mannheim are supplied with district heating systems that also provide
decentralized absorption chillers with district hot water (150F/212F [65.6C/100C).

GEOTHERMAL DISTRICT HEATINGDIRECT USE


The term geothermal has evolved in the U.S. to define two very different types of systems. Traditionally, the term denoted those systems that involved drilling for anomalously
high-temperature fluids (water or steam typically in the 90F to 400F [32C to 204C]
range) to be used for either electric power generation or other direct uses, which include

2.43

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

such applications as district heating, greenhouse heating, aquaculture, and various industrial uses. More recently, the term geothermal has also been associated with groundsource heat pump systems in which the natural temperature of the earth is used in conjunction with heat pumps to provide heating and cooling for buildings. This section
addresses the direct-use applications; heat pump applications were covered in the Large
Heat Pumps section.
District heating using geothermal resources offers the advantage of a renewable heat
source, the principal operating cost of which is the amortization of well construction
costs. There are zero air emissions associated with the heat source, and the design practices are well established based on the many systems currently in operation.

U.S. Experience
Some key challenges associated with district heating using geothermal resources in
the U.S. include geographic limitations for resource availability, drilling risks (depth,
temperature, fluid availability), regulatory impediments (particularly on federal lands),
poor overall economics, interference and/or litigation with existing resource users over
use of the resources, and disposal of fluids (injection).
In the U.S., geothermal resources available at drilling depths and temperatures suitable for district heating applications are limited to specific areas in the western states (see
Figure 2.15), and all geothermal district heating (GDH) development to date has occurred
here. U.S. systems generally provide 100% of the heating with the geothermal resource
(no conventional fuel peaking) and tend to be quite small in comparison to European
GDH and conventional district heating systems in general.

Figure 2.15 Geothermal resources and GDH locations in the U.S.

2.44

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Geothermal systems in the U.S. operate with production wells in the temperature
range of 130F to 210F (54C to 99C) and well depths in the range of approximately
300 to 1800 ft (91 to 548 m). Production well pumps must be carefully designed to
accommodate the higher temperature fluids. The geothermal fluids, because they often
contain chemical constituents that can be detrimental to standard materials of construction, are normally isolated from customer systems with plate heat exchangers. Distribution systems can be designed as open, in which the geothermal fluids are delivered to the
customer and heat exchange occurs there, or closed, in which a central heat exchanger
facility provides this function. Generally, the central heat exchange approach is more economical when more than six buildings will be served. Distribution piping is generally
direct buried and pre-insulated, fiberglass-reinforced plastic (FRP) or ductile iron being
the most common carrier pipe materials. Early systems often used asbestos cement piping. Details of the design of direct-use geothermal systems can be found in Chapter 34 of
ASHRAE HandbookHVAC Applications (2011). More in-depth coverage of the topic
appears in Geothermal Direct-Use Engineering and Design Guidebook, 3rd ed. (Lund
1998), A Materials and Equipment Review of Selected US Geothermal District Heating
Systems (Rafferty 1989), Advanced District Heating Technologies (Morck and Pedersen
1989), and the transactions of the Geothermal Resources Council.
In order to conserve the resource, most systems now inject the spent geothermal fluids back into the producing aquifer down gradient (downstream in the context of aquifer
flow) from the production wells. Drilling and testing of both production and injection
wells prior to final system design is strongly recommended so as to confirm available
flows and temperatures. This testing should be conducted by individuals experienced in
geothermal geology/hydrology.
Because of the location of geothermal resources, primarily in the western U.S., most
system development has occurred in small- to medium-sized cities; see Table 2.11 for
selected data on some U.S. systems. As a consequence, most systems serve relatively
small buildings (<15,000 ft2 [<1400 m2]). As a result, many customers do not have hotwater-based heating systems and retrofitting to accommodate the use of a hot-water
source is necessary. The cost of a retrofit for converting to hot water and the prospect of
achieving an acceptable payback is poor for small buildings. To date, the economics for
customers has been a significant impediment to full subscription of many existing GHD
systems. In addition, the risk associated with drilling wells and the high cost of installing
distribution infrastructure have resulted in most current systems being installed with
heavy subsidies in the form of government grants. Only 2 of the approximately 12 larger
systems are operated privately, with the balance operated by municipalities.
Table 2.11 Selected Cost and Capacity Data for U.S. GDH Systems
System

Capacity,
MBtu/h (MWt)

Well Cost,
2012 USD

District
System
Cost,
2012 USD

Well Depth,
ft (m)

Production
Well
Temperature,
F (C)

Production
Well Flow,
gpm (L/s)

Susanville, CA

12.2 (3.5)

$3,200,000

$5,000,000

925 (282)

174 (79)

700 (44)

Elko, NV

13.0 (3.8)

$2,000,000

$1,900,000

858 (261)

177 (81)

650 (41)

163 (73)
172 (78)
176 (80)

1500 (95)
2000 (127)
750 (47)

218 (103)
212 (100)

500 (32)
500 (32)

Boise, ID

85 (25)

$5,000,000

$7,500,000

880 (268)
1897 (578)
1103 (336)

Klamath Falls, OR

20 (5.8)

$850,000

$4,900,000

367 (119)
900 (274)

2.45

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

European Experience
The setting for GDH in Europe, Turkey, and Iceland is typically more positive than in
the U.S. The wider use of hot-water-based heating systems in buildings, denser heating
demand arising from more urban application sites, and the more highly developed use of
district heating, in general, combine to result in a much greater level of development
internationally than in the U.S. Iceland in particular has been the most successful in GDH
implementation, with most space heating (98%) in the country provided in this manner.
The Reykjavik system alone has a capacity of 2,730,000,000 Btu/h (800 MWt) and provides heat for a population of 180,000 using 167F (75C) water. The growth of the
Reykjavik GDH system has absorbed all the capacity of the local low-temperature geothermal resources, and recent GDH capacity expansion has been met by waste heat from
geothermal electric power generation at two high-temperature fields near Reykjavik.
Most Icelandic geothermal power plants are of the flashed steam type, but recent development has included Organic Rankine Cycle plants to take advantage of the potential in
lower-temperature fluids.
European systems tend to be somewhat smaller in capacity and operate from much
deeper wells than the U.S. systems. The most concentrated development is in the Paris
Basin, where some 34 systems with a combined capacity of 785,000,000 Btu/h
(230 MWt) serve a population of approximately 100,000 people. These systems produce
fluids of approximately 158F (70C) from depths of 4900 to 5900 ft (1500 to 1800 m).
GDH systems are also located in Sweden, Denmark, Finland, United Kingdom, Germany,
Macedonia, Poland, Lithuania, Hungary, Slovak Republic, Slovenia, Switzerland, and
Italy.
Turkey has embarked on an aggressive GDH development plan over the past 25 years,
with at least 20 systems in operation with a total capacity of 1,348,000,000 Btu/h
(395 MWt) serving approximately 64,000,000 ft2 (6,000,000 m2) of building area. These
systems operate with resource temperatures in the range of 153F to 257F (67C to
125C) (Serpen et al. 2010). Problems have developed in some systems with respect to
poor planning, insufficient geothermal capacity, and competition from low-cost natural
gas (Serpen 2004).

REFERENCES
Aamot, H., and G. Phetteplace. 1978. Heat transmission with steam and hot water. Publication H00128, American Society of Mechanical Engineers, New York.
ASHRAE. 2011. ASHRAE HandbookHVAC Applications. Atlanta: ASHRAE.
ASHRAE. 2012. ASHRAE HandbookHVAC Systems and Equipment. Atlanta:
ASHRAE.
ASHRAE. 2013a. District Cooling Guide. Atlanta: ASHRAE.
ASHRAE. 2013b. ASHRAE Owning and Operating Cost Database, www.ashrae.org/
database. Atlanta: ASHRAE.
Bahnfleth, D.R. 2004. A utility master planning pyramid for university, hospital, and corporate campuses. Heating/Piping/Air Conditioning 76(5):7681.
Bloomquist, R.G., R. OBrien, and M. Spurr. 1999. Geothermal district energy at colocated sites. WSU-EEP 99007, Washington State University Extension Energy Program, Olympia, WA.
COWIconsult. 1985. Computerized planning and design of district heating networks.
COWIconsult Consulting Engineers and Planners AS, Virum, Denmark.
Geiringer, P.L. 1963. High-Temperature Water Heating: Its Theory and Practice for District and Space Heating Applications. New York: John Wiley & Sons.

2.46

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2 Planning and System Selection

Griffin, J.D., C.F. Rosenow, J.C. Friedline, F. Strnisa, and I. Oliker. 1987. Development of
district heating in Buffalo. District Heating and Cooling Magazine 4.
Hyman, L.B. 2010. What building owners and designers need to consider for utility master planning for multiple building sites. Presented at the 2010 ASHRAE Annual Conference (Seminar 55), June, Albuquerque, NM.
IDHA. 1983. District Heating Handbook, 4th ed. Washington, DC: International District
Heating Association.
IEA. 1996. Integrating District Cooling with Combined Heat and Power. Paris, France:
International Energy Agency.
ISO. 2009. ISO 2314:2009, Gas turbines -- Acceptance tests. Geneva: International Organization for Standardization.
Lund, J., ed. 1998. Geothermal Direct-Use Engineering and Design Guidebook, 3rd ed.
Klamath Falls, OR: Geo-Heat Center, Oregon Institute of Technology.
Lund-Hansen, T. 2010. Practical use of real-time hydraulic modeling in day-to-day system operation and optimization: Examples from utilities in the US, and around the
world. Presented at the 101 Annual Conference of the IDEA, June 15, Indianapolis,
Indiana.
Lund-Hansen, T., and J.O. Hansen. 2011. Optimization of system operation and maintenance using advanced real-time monitoring. Presented at the Distribution Workshop
of the IDEA, February 222, Miami, Florida.
McIntire, M.E., D. Hall, D.J. Beal, I. Oliker, and W. Major. 1994. Assessment of district
heating and cooling supply from Goudey Generating Station. Proceedings of the Heat
Rate Improvement Conference, Baltimore, MD, May 35.
Morck, O., and T. Pedersen, eds. 1989. Advanced District Heating Technologies. Copenhagen: DK-teknik/International Energy Agency.
NFPA. 2011. National Electrical Code (NFPA 70). Quincy, MA: National Fire Protection Association.
Oliker, I. 1979. Steam turbines for cogeneration power plants. Transactions of the ASME
Journal of Engineering for Power, 102(2).
Oliker, I. 1980. Steam turbines for cogeneration power plants. Engineering for Power
102(2).
Oliker, I. 1981. Utilization of gas turbines for district heating applications. Presented at
the Gas Turbine Conference, Houston, Texas, March.
Oliker, I. 1982. Gas and steam turbines for district heating. Presented at the 27th International Gas Turbine Conference, London, England, April.
Oliker, I. 1996. Retrofitting power plants to provide district heating and cooling. EPRI
TR-106027, Electric Power Research Institute, Palo Alto, CA.
Oliker, I., and A.F. Armor. 1992. Supercritical power plants in the USSR. EPRI TR100364, Electric Power Research Institute, Palo Alto, CA.
Oliker, I., and D.V. Champ. 2000. Fifteen years of experience with district heat supply
from a retrofitted power plant in Jamestown, NY. Caddet Newsletter. International
Energy Agency Publication.
Oliker, I., and H.J. Mulhauser. 1980. Technical and economic aspects of coal-fired district
heating power plants in USA. Proceedings of the American Power Conference,
Vol. 42.
Oliker, I., and F. Silaghy. 1987. Combustion turbine repowering. EPRI AP-5493, Electric
Power Research Institute, Palo Alto, CA.
Phetteplace, G. 1994. Optimal design of piping systems for district heating. PhD dissertation, Mechanical Engineering Department, Stanford University, Stanford CA.

2.47

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Rafferty, K. 1989. A materials and equipment review of selected US geothermal district


heating systems. DE-FG07-87ID-12693, U.S. Department of Energy, prepared by
Geo-Heat Center, Oregon Institute of Technology.
Rasmussen, C.H., and J.E. Lund. 1987. Computer aided design of district heating systems. District Heating Research and Technological Development in Denmark, Danish
Ministry of Energy, Copenhagen, Denmark.
Serpen, U. 2004. Present status of geothermal energy and its utilization in Turkey 2004.
Proceedings of the 28th Annual Geothermal Reservoir Engineering Conference,
Stanford University, February.
Serpen, U., A. Niyazi, and O. Tahir. 2010. Present status of geothermal energy and its utilization in Turkey 2010. SGP-TR-188, Proceedings of the 34th Annual Geothermal
Reservoir Engineering Conference, Stanford University, February.
Stal-Laval. 1983. A new way to replace oil and coal with non-polluting energy sources.
Sweden: Stal-Laval Turbin AB.
Werner, S.E. 1984. The Heat Load in District Heating Systems. Gteborg, Sweden:
Chalmers University of Technology.

BIBLIOGRAPHY
Oliker, I. 1982. Combined electric and heat generation at coal fired plants. Presented at
Energy, Resources and Environment, U.S.-China Conference on Energy, Resources
and Environment, Beijing, China, November.
Oliker, I. 1983. District heating supply from retrofitted power plants. Proceedings of the
74th International District Heating Association, Vol. 74.
Oliker, I., W. Major, and D. Gray. 1993. District heating brings cogeneration to an existing municipal power plant. Power 137(1).
Petrill, E., and I. Oliker. 1994. EPRI assessment of power plant retrofits for district heating and cooling. Presented at the Heat Rate Improvement Conference, Baltimore,
MD, May 35.

2.48

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3
Central Plant Design
for Steam and
Hot Water
INTRODUCTION
District heating systems distribute thermal energy from a central source to residential,
commercial, and/or industrial consumers for use in space heating, domestic water, process heating and sometimes cooling. The thermal energy is distributed to users through
steam or hot-water underground piping. Whether the system is a public utility or user
owned, such as a multiple-building campus or a military base, it has economic and environmental benefits depending on the particular application.
District heating systems are best used in areas where the thermal load density and the
annual load factor are high. The annual load factor is important because the total system
is capital intensive. These factors make district heating systems most attractive in serving
densely populated urban areas, high-density building clusters, and industrial complexes
with high thermal loads. District heating is best suited to areas with a high building and
population density in relatively cold climates.
The central heating or production plant may include any type of boiler, a refuse incinerator, a geothermal source, solar thermal energy, or thermal energy developed as a byproduct of electrical generation. The last approach, called cogeneration or combined heat
and power (CHP), has high energy utilization efficiency. The central heating plant contains major mechanical and electrical equipment, such as boilers, pumps, switchgear, and
auxiliaries. In terms of cost, the central plant can represent up to 25% of the total capital
investment and is responsible for the largest segment of building operating costs. In comparison with individual heating, central plants offer several benefits (see Chapters 3
and 11 of ASHRAE HandbookHVAC Systems and Equipment [2008]), discussed in the
subsections that follow.

Higher Thermal Efficiency


A larger central plant can achieve higher thermal and emission efficiencies than can
individual smaller units. Cogeneration of heat and electric power results in much higher
overall efficiencies than are possible from separate heat and power plants.
Partial-load performance of central plants may be more efficient than that of many
isolated small systems because the larger plant can operate one or more capacity modules
as the combined load requires and can modulate output. Central plants generally have
efficient base-load units and less costly peaking equipment for use in extreme loads or
emergencies.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Use of Multiple Fuels


Individual heating plants are usually designed for one type of fuel, which is generally
gas or oil. Central plants can operate on less expensive coal or refuse. Larger facilities can
often be designed for more than one fuel (e.g., coal, oil, or biomass) and combined with
power generation.

Environmental Benefits
Emissions from central plants are easier to control than those from individual plants
and, in aggregate, are lower because of higher quality equipment, seasonal efficiencies,
and level of maintenance and lower system heat loss. When strict regulations must be
met, additional pollution control equipment is also more economical for larger plants. A
central plant that burns coal can economically remove noxious sulfur emissions, whereas
individual combustors could not. Similarly, the thermal energy from municipal wastes can
provide an environmentally sound system. Cogeneration of heat and electric power allows
for combined efficiencies of energy use that greatly reduce emissions and also allow for
fuel flexibility.

Operating Personnel
One of the primary advantages of a central plant to a building owner is that operating
personnel for the individual boiler plant can be eliminated. Most municipal codes require
licensed operating engineers to be on site when high-pressure boilers are in operation.

Insurance
Both property and liability insurance costs are significantly reduced with the elimination of a boiler in the mechanical room, because risk of a fire or accident is reduced.

Usable Space
Usable space in the building increases when a boiler and related equipment are no
longer necessary. Although this space usually cannot be converted into prime office
space, it does provide the opportunity for increased storage or other use. The noise associated with such in-building equipment is also eliminated.

Equipment Maintenance
With less mechanical equipment at the user site, there is proportionately less equipment maintenance, resulting in less expense and a reduced maintenance staff.

Use of Cogeneration
If a new central plant is being considered, a decision must be made of whether to
cogenerate electrical and thermal energy or to generate thermal energy only. Cogeneration will normally require higher temperatures and pressures than heat-only plants. In
either instance, however, the selection of temperature and pressure is crucial because it
can dramatically affect the economic feasibility of a central plant design. For a discussion
on choice of medium for district heating systems, see Chapter 2.
Central plants are characterized by large heating equipment located in one standalone location. Equipment configuration and ancillary equipment vary significantly,
depending on the facilitys use. Primary equipment (steam or hot-water generators) is
available in different sizes, capacities, and configurations to serve a variety of building
applications. Operating a few pieces of primary equipment (with backup equipment)
gives central plants different benefits from decentralized systems. Multiple types of
equipment and fuel sources may be combined in one plant. The primary fuel energy is

3.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

converted into steam or hot water that is distributed through an underground piping system to the users.
A central plant can be customized without sacrificing the standardization, flexibility,
and performance required to support the primary heating equipment through careful
selection of ancillary equipment, automatic control, and facility management. These
plants require more extensive engineering, equipment, and financial analysis than decentralized systems do. For larger plants, operation and maintenance (O&M) is performed
with multi-shift operating crews, providing 24-hour shift coverage all year round.

CENTRAL PLANT ADVANTAGES


The advantages of central plants are the following:
Heat energy can be provided at all times, independent of the operation mode of
equipment and systems outside the central plant.
Using larger but fewer pieces of equipment reduces the facilitys overall O&M
cost. It also allows wider operating ranges and more flexible operating
sequences.
A centralized location minimizes restrictions on servicing accessibility.
Energy-efficient design strategies, cogeneration, energy recovery, thermal storage, and energy management can be simpler and more cost-effective to implement.
Multiple energy (fuel) sources can be applied to the central plant, providing
flexibility and leverage when purchasing fuel.
Standardizing equipment can be beneficial for redundancy and stocking replacement parts. However, strategically selecting different-sized equipment for a central plant can provide better part-load capability and efficiency.
Standby capabilities (for firm capacity/redundancy) and back-up fuel sources
can easily be added to equipment and plant when planned in advance.
Equipment operation can be staged to match load profile and taken offline for
maintenance.
A central plant can be economically expanded to accommodate future growth
(e.g., adding new users).
Load diversity can substantially reduce the total equipment capacity requirement.
Major vibration and noise-producing equipment can be grouped away from
occupied spaces, making acoustic and vibration controls simpler. Acoustical
treatment can be applied in a single location instead of many separate locations.
Plant emissions are centralized, allowing a more economic and lower emission
release solution.
Opportunities for improving energy efficiency include staging multiple boilers
for part-load operation. Using correctly sized equipment is imperative to accurately provide the more flexible and economic sequencing of equipment.

CENTRAL PLANT DISADVANTAGES


The disadvantages of central plants are the following:
Equipment may not be readily available, resulting in long lead time for production and delivery.
Equipment is more complicated than decentralized equipment and thus requires
a more knowledgeable equipment operator.
A central location within the users area is needed.

3.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Depending on the fuel source, large underground or surface storage tanks may
be required on site. If coal is used, space for storage bunkers is needed.
Access may be needed for large deliveries of fuel (oil or coal).
Heating plants require a chimney and emission permits, monitoring, and treatments.
Multiple equipment manufacturers are required when combining primary and
ancillary equipment.
System control logic may be complex.
Special permitting is required.
Safety requirements are increased.
The central plant will have some piping distribution losses.

HEATING LOADS
Design heating load of the plant is determined by considering individual and simultaneous loads. The simultaneous peak or instantaneous load of the users is less than the sum
of the individual users (e.g., buildings do not receive peak solar load on the east and west
exposures at the same time). This difference between design load and peak load, called
the central equipment diversity factor, can be as little as 5% less than the sum of the individual loads (e.g., 95% diversity factor) or represent a more significant portion of the
load, as is common in large district energy applications (see Chapter 2 for further discussion of demand diversity). The peak central plant load can be based on this diversity factor, reducing the total installed equipment capacity needed to serve the users heating
loads. It is important for the design engineer to evaluate the full point-of-use load requirements of each facility served by the district system.
When selecting the peak heating load it should be taken into account that this load
may occur when the building must be warmed back up to a higher occupied space temperature after an unoccupied weekend setback period. Peak demand may also occur during
unoccupied periods when the ambient environment is harshest and there is little internal
heat gain to assist the heating system or during occupied times if significant outdoor air
must be preconditioned or some other process (e.g., process heating) requires significant
heat. In hot-water systems, variable-flow may be the best economical choice to accommodate part-load condition and energy efficiency. It is important for the designer to evaluate plant operation and system use.

CENTRAL PLANT HEATING MEDIUM


Central plants may generate steam or high-, medium-, and low-temperature water.
Central plants are divided into four major categories:
Steam plants generating steam
High-temperature plants generating water with temperatures over 350F
(176.7C)
Medium-temperature plants with temperatures in the range of 250F to 350F
(121.1C to 176.7C)
Low-temperature systems with temperatures of 250F (121C) or lower
In plants serving hospitals or industrial customers or plants with decentralized cooling or also generating electricity, steam is the usual choice for production in the plant and
for distribution to customers. For systems serving largely commercial buildings, hot water
is an attractive medium. From the standpoint of distribution, hot water can accommodate
a greater geographical area than steam because of the ease with which booster pump stations can be installed.

3.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Several factors must be evaluated to select a suitable medium. A basic decision is


whether the plant should use hot water or steam boilers. Another involves the operating
conditions that should be used. If the system supported by the boiler plant is strictly a
heating system, the most logical choice is hot water. Other uses, however, may require
steam, such as kitchen, laundry, sterilizing, process, or absorption cooling. With steam
loads of this type, it may be desirable to generate steam for the specific requirements and
use steam-to-water heat exchangers for generating heating hot water. Another alternative
is using hot-water generators for the building heating load and using a separate steam
boiler to serve the steam load.
When the basic system has been selected, the subject of operating temperature and
pressure must be addressed. The common attributes and relative merits of hot water and
steam as heat-conveying media are described in the following subsections.

Heat Capacity
Steam relies primarily on the latent heat capacity of water rather than on sensible
heat. The net heat content for saturated steam at 100 psig and 338F (690 kPa and 170C)
condensed and cooled to 180F (82C) is approximately 1040 Btu/lbm (2419 kJ/kg). Hot
water cooled from 350F to 250F (176.7C to 121.1C) has a net heat effect of 103 Btu/
lbm (239.6 kJ/kg), or only about 10% as much as that of steam. Thus, a hot-water system
must circulate about 10 times more mass than a steam system of similar heat capacity.

Pipe Sizes
Despite the fact that less steam is required for a given heat load and flow velocities
are greater, steam usually requires a larger pipe size for the supply line because of its
lower density. This is compensated for by a much smaller condensate return pipe. Therefore, piping costs for steam and condensate are often comparable with those for hot-water
supply and return.

Condensate Return System


Condensate return systems require more maintenance than hot-water return systems
because hot-water systems function as a closed loop with very low makeup water requirements. For condensate return systems, corrosion of piping and other components, particularly in areas where feedwater is high in bicarbonates, is a problem. Nonmetallic piping
has been used successfully in some applications, such as systems with pumped returns,
where it has been possible to isolate the nonmetallic piping from live steam.
Similar concerns are associated with condensate drainage systems (steam traps, condensate pumps, and receiver tanks) for steam supply lines. Condensate collection and
return should be carefully considered when designing a steam system. Although similar
problems with water treatment occur in hot-water systems, they present less of a concern
because makeup rates are much lower.

Pressure and Temperature Requirements


Flowing steam and hot water both incur pressure losses. Hot-water systems may use
intermediate booster pumps to increase the pressure at points between the plant and the
consumer location. Because of the higher density of water, pressure variations caused by
elevation differences in a hot-water system are much greater than for steam systems. This
can adversely affect the economics of a hot-water system by requiring the use of a higher
pressure class of piping and/or booster pumps.
Regardless of the medium used, the temperature and pressure used for heating should
be no higher than needed to satisfy consumer requirements; this cannot be overemphasized. Higher temperatures and pressures require additional engineering and planning to

3.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

avoid higher heat losses. Higher temperatures may require higher pressure ratings for piping and fittings and may preclude the use of materials such as polyurethane foam insulation and nonmetallic conduits. Safety and comfort levels for operators and maintenance
personnel benefit from lower temperatures and pressures.

BOILER PRESSURE AND TEMPERATURE


With few exceptions, boilers are constructed to meet the sections on constructing
heating and power boilers of the ASME Boiler and Pressure Vessel Code (2013). Lowpressure boilers are constructed for maximum working pressures of 15 psig (103.6 kPa)
steam and up to 160 psig (1104 kPa) hot water. Low-pressure hot-water boilers are limited to 250F (121.1C) operating temperature. High-pressure boilers are designed to
operate above 15 psig (103.6 kPa) steam or above 160 psig and/or 250F (1104 kPa and/
or 121.1C) for water boilers.
Traditionally, boilers are rated by boiler horsepower, a unit of measurement with
1 boiler horsepower being equal to 33,475 Btu/h (35,309 kJ/h) or the evaporation of
34.5 lbm(15.65 kg) of water per hour at standard atmospheric pressure 14.7 psia and
212F (101.4 kPa and 100C). Every steam or water boiler is rated for a maximum working pressure that is determined by the applicable boiler code under which it is constructed
and tested. When installed, it also must be equipped at a minimum with operation and
safety controls and pressure/temperature-relief devices mandated by such codes.

CONSTRUCTION MATERIALS
Most central plant boilers are made with steel. Steel boilers can be any size starting
from 50,000 Btu/h (52,740 kJ/h). Designs are constructed according to ASME Boiler and
Pressure Vessel Code (2013) or other applicable code requirements. They are fabricated
into one assembly of a given size and rating, usually by welding. The heat exchange surface past the combustion chamber is usually an assembly of vertical, horizontal, or
slanted tubes.
Both fire-tube and water-tube designs are used. Larger boilers usually incorporate
horizontal or slanted tubes. Boilers of the fire-tube design contain flue gases in tubes
completely submerged in fluid. Water-tube boilers contain fluid inside tubes. A popular
horizontal fire-tube design for medium and large steel boilers is the scotch marine, which
is characterized by a central fluid-backed cylindrical combustion chamber surrounded by
fire-tubes accommodating two or more flue gas passes, all within an outer shell. In
another horizontal fire-tube design, the combustion chamber has a similar central fluidbacked combustion chamber surrounded by fire tubes accommodating two or more flue
gas passes.
Noncondensing heat plant boiler efficiency may be improved with the use of external
flue gas-to-water economizers. The condensing medium may include domestic hot-water
preheat, steam condensate or hot-water return, fresh-water makeup, or other fluid sources
in the 70F to 130F (21C to 54.4C) range. The medium can also be used as a source of
heat recovery. Care must be taken to protect the noncondensing boiler from the low-temperature water return in the event of economizer service or control failure.

SELECTION PARAMETERS
Boiler selection should be based on a competent review of the following parameters:
Applicable code under which the boiler is constructed and tested
Gross boiler heat output
Combustion chamber (furnace volume) and burner type
Internal flow pattern of combustion products
Combustion air and venting requirements

3.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Fuel availability/capability
Part-load versus full-load efficiency (life-cycle cost)
Total heat transfer surface area
Water content weight or volume
Auxiliary power requirement
Cleaning and service access provisions for fire-side and water-side areas
Space requirement and piping arrangement
Water treatment requirement
Operating personnel capabilities and maintenance/operation requirements
Regulatory requirements for emissions and fuel usage/storage

The codes and standards include requirements for minimum efficiency, maximum
temperature, burner operating characteristics, and safety control. Test agency certification
and labeling, which are published in boiler manufacturers catalogs and shown on boiler
rating plates, are generally sufficient for determining boiler steady-state operating characteristics. However, for noncondensing commercial and industrial (district heating) boilers,
these ratings typically do not consider part-load or seasonal efficiency, which is less than
steady-state efficiency.

EFFICIENCY
The efficiency of fuel-burning boilers is defined in the following ways: combustion,
overall, and seasonal (ASHRAE 2008, Chapter 31). However, manufacturers are not
required to test or publish efficiencies that coincide with these industry definitions.

Combustion Efficiency
Combustion efficiency is input minus stack (flue gas outlet) loss divided by input and
generally ranges from 75% to 86% for most noncondensing boilers. Approximate combustion efficiency for noncondensing boilers can be determined under any operating condition by measuring flue gas temperature and percentage of carbon dioxide (CO2) in the
flue gas and by consulting a chart or table for the fuel being used. The approximate combustion efficiency of a condensing boiler must include the energy transferred by condensation in the flue gas.

Overall (or Thermal) Efficiency


Overall (or thermal) efficiency is gross energy output divided by energy input. Gross
output is measured in the steam or water leaving the boiler and depends on the characteristics of the individual installation. Overall efficiency is lower than combustion efficiency
due to the heat lost from the outside surface of the boiler (radiation loss or jacket loss) and
by off-cycle energy losses (for applications where the boiler cycles on and off). Overall
efficiency can be precisely determined only under controlled laboratory test conditions,
directly measuring the fuel input and the heat absorbed by the water or steam of the
boiler. Precise efficiency measurements are generally not performed under field conditions because of the inability to control the required parameters and the high cost involved
in performing such an analysis.

Seasonal Efficiency
Seasonal efficiency is the actual operating efficiency that the boiler will achieve during the heating season at various loads. Because most heating boilers operate at part load,
the part-load efficiency, including heat losses when the boiler is off, has a great effect on
the seasonal efficiency. The difference in seasonal efficiency between a boiler with an on/
off firing rate and one with a modulating firing rate can be appreciable if the airflow

3.7

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

through the boiler is modulated along with the fuel input. This increase in efficiency is
due to the increase in the ratio of heat exchanger surface area to heat input as the firing
rate is reduced. Since larger boilers normally regulate airflow and modulate firing rate,
this is an advantage of central plant scale applications.

PERFORMANCE CODES AND STANDARDS


Commercial Heating Boilers
Commercial heating boilers (i.e., boilers with inputs of 300,000 Btu/h [316,440 kJ/h]
and larger) in the United States at present are only tested for full-load steady-state efficiency according to standards developed by either
the Hydronics Institute Division of the Gas Appliance Manufacturers Association (GAMA) (formerly the Institute of Boiler and Radiator Manufacturers and
the Steel Boiler Institute),
the American Gas Association (AGA), or
Underwriters Laboratories (UL).

CENTRAL PLANT DESIGN FOR STEAM


Typical System Arrangements
Steam boilers are generally available in standard sizes in a wide range of capacities,
from 500 to over 100,000 lb/h (217 to over 43,350 kg/h) (50,000 to over 100,000,000 Btu/h
[63.3 MJ/h to over 105.5 GJ/h]), many of which are used for space-heating applications in
both new and existing systems. In larger installations, they may also provide steam for
auxiliary uses, such as hot-water heat exchangers, deaeration, absorption cooling, laundry, and sterilizers. In addition, many steam boilers provide steam at various temperatures
and pressures for a wide variety of industrial processes.
Selection of central steam plants equipped with boilers rated between 20,000 and
250,000 lb/h (9,070 and 113,375 kg/h) should take into account the considerations in this
section (DOD 2007). Where special conditions and problems are not specifically covered
in this section, acceptable industry standards should be followed. The selection of one
particular type of design for a given application, when two or more types of design are
known to be feasible, should be based on the results of an economic study. Central steam
plants can be fired by gas, oil, gas and oil, coal, biomass, or waste fuels. Coal-fired plants
can use any combination of three commercially proven coal-firing technologies: atmospheric circulating fluidized bed (ACFB), pulverized coal (PC), and stoker-fired boilers.
Stokers are designed to burn any one of the different types of anthracite, bituminous, subbituminous, or lignite type coals. ACFB boilers offer reduced sulfur dioxide emissions
without use of scrubbers while firing a range of lower-cost fuels.
The principal diagram of a central steam plant is presented in Figure 3.1. The steam
supply to the users is controlled by a control valve located on the major steam header.
Examples of typical boiler connections are demonstrated in Figure 3.2. Similar connections are used in larger district heating steam plants.
Steam plant reliability standards should be equivalent to a one-day forced outage in
ten years with equipment quality and redundancy selected during plant design to conform
to this standard. This standard also requires quality engineering, equipment, and O&M
personnel.
In order for boilers to have high availability it is mandatory that a good water treatment program be implemented. Availability guarantees offered by boiler manufacturers
for coal-fired units are in the range of 85% to 90% and for gas-fired boilers are in the

3.8

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

G = gas line, FOR = fuel oil return, FOS = fuel oil supply, MPC = medium-pressure condensate,
MPS = medium-pressure steam, HPS = high-pressure steam, CPD = condensate pump discharge,
MWL = makeup water line, BD = blowdown line, CF = chemical feed (water treatment)

Figure 3.1 Principal diagram of a central steam plant.


Courtesy of William Tao and Associates (Tao and Laird 1984)

range of 90% to 95%. A planned outage of a minimum of one day per year is normal for
water-side inspection.
Central steam plant arrangement should permit reasonable access for O&M of equipment. Careful attention should be given to the arrangement of equipment, valves, mechanical specialties, and electrical devices so that rotors, tube bundles, inner valves, top works,
strainers, contactors, relays, and like items can be maintained or replaced. Adequate platforms, stairs, handrails, and kick plates should be provided so that operators and maintenance personnel can function conveniently and safely.
The specific site selected for the central steam plant and the physical arrangement of
the plant equipment, building, and support facilities such as natural gas supply lines, coal
and ash handling systems, coal storage, circulating water system, trackage, and access
roads should be arranged insofar as practicable to allow for future expansion.

3.9

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 3.2 Central boiler plant typical connections.


Adapted from ASHRAE HandbookHVAC Systems and Equipment, Chapter 10, Figure 2 (2008)

Selection Criteria
The following information is required for the central steam plant equipment selection:
A forecast of annual and monthly diversified peak loads to be served by the
plant.
Typical, seasonal, weekly, and daily load curves and load duration curves of the
load to be served.
A forecast of peak loads encountered during the steam plant full mobilization.
If the plant is to operate in conjunction with any existing steam generation on
the district heating system or is an expansion of an existing facility:
An inventory of major existing steam generation equipment, including
principal characteristics such as type, capacities, steam characteristics,
pressures and like parameters.
Incremental thermal efficiency of existing boiler units.
Historical operating data for each existing steam-generating unit, including
energy generated, fuel consumption, and other related information.
Existing or recommended steam distribution systems to support base operations.
Fuel sources and cost. The type, availability, and cost of fuels will be determined
in the early stages of design, taking into account environmental regulatory
requirements that may affect the fuel selection for the plant. A complete fuel
analysis for each fuel being considered for use in the plant should be performed.
Coal analyses shall include proximate analysis, ultimate analysis, grindability,
higher heating value, ash analysis, ash fusion temperature, and agglomerating
classification.
Plant site conditions, including ambient temperature ranges, maximum expected
wind conditions, snow load, seismic conditions, and any other site conditions
that could affect the design of the boiler and its accessories.

3.10

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Freshwater analysis for the boiler cycle makeup. The quantity of makeup will
vary with the type of boiler plant, the amount of condensate return for any
export steam, and the maximum heat rejection from the cycle. Water analysis
should include such things as total dissolved solids, pH, etc.
All codes that pertain to designing, constructing, and placing into operation the
boiler and its accessories.
Stack emissions. A steam plant should be designed for stack gas cleanup equipment that meets federal, state, and local emission requirements. For a coal-fired
boiler, this involves an electrostatic precipitator or bag house for particulates or fly
ash removal and a scrubber for flue gas desulfurization (FGD) unless fluidizedbed combustion or compliance coal are employed. If the design is based on compliance coal, it should include space and other required provisions for the installation of equipment. Boiler design should be specified as required for nitrogen
oxides (NOx) control. ACFB boilers have low NOx emissions because of their relatively low combustion temperatures. Selective non-catalytic reduction (SNCR)
systems are required by many states, even on ACFB boiler installations, to meet
new NOx allowable emissions limits (AELs) that are more stringent.
Available waste disposal methods. Both solid and liquid wastes should be handled and disposed of in an environmentally acceptable manner. The wastes can
be categorized generally as follows: solid wastes including both bottom ash and
fly ash from boilers; liquid wastes including boiler blowdown, cooling tower
blowdown, acid and caustic water-treating wastes, coal pile runoff, and various
contaminated wastes from chemical storage areas, sanitary sewage, and yard
areas.
Other environmental considerations including noise control and aesthetic treatment of the project. The final location of the project within the site area should
be reviewed in relation to its proximity to hospital and office areas and the civilian neighborhood, if applicable. Also, the general architectural design should be
reviewed in terms of coordination and blending with the style of surrounding
buildings. Any anticipated noise or aesthetics problems should be resolved prior
to the time that final site selection is approved.
If any of the above data that is required for performing the detailed design is unavailable, the designer should develop this data.

Construction Cost Estimate


The following items should be considered in the construction cost estimate for the
central steam plant:
Steam generators, burners, particulate control equipment, sulfur removal equipment, stacks, continuous flue gas monitoring equipment
Coal, limestone, inert bed and ash storage and handling systems
Panels, instruments and controls
Water treatment equipment
Deaerators, feedwater heaters, boiler feed pumps, air compressors, power piping, electrical equipment, power wiring, plant substation
Fire protection
Substructure, structural steel, superstructure, foundations for all equipment,
building foundations
Roads, grading and site improvements
Ash pond, coal runoff pond and coal pile stabilization

3.11

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Railroad siding
Water and sewers
Painting, fencing
Operator training
Fuels and chemicals
Engineering fees, supervision and inspection for the construction
Contingencies
Profit

Feasibility Analysis
Each potential plant site should be evaluated to determine whether it is the most economically feasible for the size of plant being considered. The process of evaluating and
selecting a steam plant site must include the following activities:
Identify location of interest.
Apply exclusionary criteria to eliminate unsuitable areas within location of
interest.
Identify candidate sites.
Prepare site-specific layouts and conduct conceptual engineering studies of the
candidate sites.
Rank candidate sites.
Based on analyses, select most suitable site.
The evaluations of candidate sites should be performed by specialists from the fields
of environmental engineering, biology/ecology, ground and surface water hydrology,
geology and seismology, soils and foundation engineering, meteorology, demography,
land use and zoning, sociology and economics, system planning, and law.
Selection of the site is based on the availability of usable land for the plant, including
yard structures, fuel handling facilities, and any future expansion. Other considerations
include soil information, site drainage, wind data seismic zone, and ingress and egress.
For economic purposes and operational efficiency, the plant site should be located as
close to the load center as environmental conditions permit.
The amount of lead time necessary to perform the studies and to submit the required
environmental impact statements should also be considered in site selection, since this
lead time will vary from site to site. Environmentally sensitive areas will probably require
lengthier studies, delaying construction. If steam load demands must be met within a limited time period, this factor becomes more important. Generally, the lead time interval
should be 18 to 24 months.

Environmental Regulations
Environmental regulations address the air, water, and waste disposal control.
Air quality control. Central steam boiler plants must meet current federal, state
and local regulations.
Water quality control. Waste water must comply with current regulations including discharge of free available chlorine, pH, total suspended solids (TSS), oil/
grease, copper (Cu), iron (Fe), polychlorinated biphenyl (PCB), and heat.
Waste material control. State regulations control disposal of coal ash and FGD
wastes and the selection of landfill site.

Water Supply
Water supply should be adequate to meet present and future plant requirements. The
supply may be available from a local municipal or privately owned system, or it may be

3.12

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

necessary to use surface or subsurface sources. Plant requirements must be estimated for
all uses, including feedwater makeup, auxiliary cooling, ash transport, coal dust suppression, fire protection, and domestic uses. A backflow preventer upstream of water treatment, ash handling equipment, and FGD system should be provided.
Water quality and type of treatment required should be compatible with the type of
plant to be built. Surface and groundwater sources must be evaluated for sufficient quantity and quality. If water rights are required, it will be necessary to ensure that an agreement for water rights provides sufficient quantity for present and future use. If the
makeup to the closed system is from water wells, a study to determine water table information and well drawdown will be required. If this information is not available, test well
studies must be made.
For large district heating plants that may have a once-through cooling system, the following should be determined:
the limitations established by the appropriate regulatory bodies that must be met
to obtain a permit to discharge heated water to the source
maximum allowable temperature rise permissible as compared to system design
parameters; if system design temperature rise exceeds permissible value, a supplemental cooling system (cooling tower or spray pond) must be incorporated
into the design
maximum allowable temperature for river or lake after mixing of cooling system
effluent with source; if mixed temperature is higher than allowable temperature,
a supplemental cooling system must be added.
It should be determined if extensive or repetitive dredging of a waterway will be necessary for plant operations. From the historical maximum and minimum water level and
flow readings it should be verified that adequate water supply is available at minimum
flow and that the site will not flood at high level.

Site Development
Soils Investigation
An analysis of existing soils conditions should be made to determine the proper type
of foundation for the plant. Soils data should include elevation of each boring, water table
level, description of soil strata including the group symbol based on the Unified Soil Classification System (see Table 4.3 of this design guide), and penetration data (blow count).
Geological conditions must support the foundations of plant structures at a reasonable
cost, with particular attention being paid to bedrock formations, unstable soils, and faults.
The soils report should include recommendations as to type of foundations for various
purposes; excavation, dewatering, and fill procedures; and suitability of on-site material
for fill and earthen dikes, including data on soft and organic materials and rocks, and
other pertinent information as applicable.
Grading and Drainage
Grading and drainage considerations should cover the following topics.
Basic criteria. Determination of final grading and the drainage scheme for a new
steam plant will be based on a number of considerations, including the size of
the property in relationship to the size of plant facilities, desirable location on
site, and plant access based on topography.
Storm water drainage. Storm water drainage will be evaluated based on rainfall
intensities, runoff characteristics of soil, facilities for receiving storm water discharge, and local regulations.

3.13

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Erosion prevention. All graded areas should be stabilized to control erosion by


designing shallow slopes to the greatest extent possible and by means of soil stabilization such as seeding, sod, stone, riprap, and retaining walls.
Meteorology
Precipitation, wind conditions, evaporation, humidity, and temperature will affect
emissions dispersion, coal storage and handling, and other aspects of plant operation.
Area Requirements
Plant area requirements should be figured on the type of plant; capacity; urban, suburban, or rural location; design of fuel storage and handling facilities; disposal of solid
waste and treatment of wastewater; condenser cooling; and the plant structure and miscellaneous requirements. Acquisition of land should include plant access roads, rail access,
and space for future additions to the plant. In addition, the following requirements should
be satisfied:
Space must be provided for the long-term coal storage pile. Maximum allowable
height of the coal pile and methods of stockout and reclaim will further affect
space requirements. If oil is used as a primary fuel or as a backup fuel, storage
tank space must be calculated.
Space for the disposal of wet and dry ash should be provided.
Control of runoff from material storage areas is required by U.S. Environmental
Protection Agency (EPA) regulations. Space for retention ponds must be provided.
Flood Protection
Larger heat-only or cogenerating power plants are often located near major water
bodies for fuel delivery or access to cooling water. Ideally, when buildings are sited in
flood-prone locations, they should be elevated above expected flood levels to reduce the
chances of flooding and to limit the potential damage to the building and its contents
when it is flooded as well as to control potential spills of toxic substances. The Whole
Building Design Guide (NIBS 2012) suggests the following flood mitigation techniques:
Elevation of the building so that the lowest floor is above the flood level
Dry flood-proofing (making the building watertight to prevent water entry)
Wet flood-proofing (making uninhabited or non-critical parts of the building
resistant to water damage)
Relocation of the building
Incorporation of floodwalls into the site design to keep water away from the
building (note that levees require a significant amount of care and are discouraged as a mitigation measure; berms for frequent events are usually a better
choice)
Other Requirements
Space for parking, warehouses, cooling water systems, environmental systems, construction laydown areas, and other requirements should be provided.

Plant Access
Plant Roadway Requirements
Layout of plant roadway should be based on volume and type of traffic, speed, and
traffic patterns. Proximity to principal highways should permit reasonably easy access for
construction crews and deliveries. Roadway design should be in accordance with American Association of State Highway and Transportation Officials (AASHTO) standard

3.14

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

specifications. Roadway material and thickness will be based on economic evaluations of


feasible alternatives. Vehicular parking for plant personnel and visitors should be located
in areas that will not interfere with the safe operation of the plant. Turning radii should be
adequate to handle all vehicle categories.
Railroads
If a railroad spur is selected to handle fuel supplies and material and equipment deliveries during construction or plant expansion, the design should be in accordance with the
American Railway Engineering and Maintenance-of-Way Association (AREMA) Manual for Railway Engineering (2012):
Spur layout should accommodate coal handling facilities, including a storage
track or loop track for empty cars.
If liquid fuel is to be handled, unloading pumps and steam connections for tank
car heaters may be required in frigid climates. In extreme cold climates, heated
enclosures may be required for thawing railroad cars for unloading.

Plant Structures
Size and Arrangements
The steam plant main building size and arrangement depend on the selected plant
equipment and facilities, including the following:
whether steam generators are indoor or outdoor type;
coal bunker or silo arrangement in the cases of pulverized coal (PC) and stokerfired plants;
coal, limestone, and inert bed silo arrangement in the case of ACFB boiler
plants;
source of cooling water supply relative to the plant;
provisions for future expansion; and
aesthetic and environmental considerations.
Generally, the main building should consist of a steam generator bay (or firing aisle
for semi-outdoor units); an auxiliary bay for feedwater heaters, pumps, and switchgear;
and general spaces as may be required for machine shop, locker room, laboratory, and
office facilities. For very mild climates, the steam generators may be outdoor type (in a
weather-protected, walk-in enclosure), although this arrangement presents special maintenance problems. If incorporated, the elevator should have access to the highest operating level of the steam generator (drum levels).
Layout Considerations
The layout of the structural system should identify specific requirements relative to
vertical and horizontal access, personnel needs and convenience, equipment and respective maintenance areas, floors and platforms.
Subsurface Exploration
A subsurface exploration program should be conducted. Design information regarding the interaction of the structure and the surrounding ground is required. In addition, the
sources of construction materials and the types and extent of materials that will be
encountered during construction should be investigated. Information required for design
includes extent of each identifiable soil stratum, depth to top of rock and character of the
rock, elevation of normal groundwater at site, and engineering properties of the soil and
rock.

3.15

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Foundation Design
Selection of the type of foundation to be used for each component of the plant (i.e.,
main building, boiler, stacks, and coal handling structure) will be determined from the
subsurface exploration data, cost considerations, and availability of construction trades.
Structural Design
The plant should be designed utilizing conventional structural steel for the main
steam building. Separate structural steel should be provided to support building floors and
platforms; boiler steel should not be used in support of the building structure. The pedestal for supporting the turbine-driven boiler feed pump, if one is used, should be of reinforced concrete. Reinforced concrete or masonry construction may be used for the
building framing (not for boiler framing); special concrete inserts or other provisions
must be made in such an event for support of piping, trays, and conduits. An economic
evaluation should be made of these alternatives.
Structure Loading
Buildings, structures, and all portions thereof should be designed and constructed to
support all live and dead loads without exceeding the allowable stresses of the selected
materials in the structural members and connections. Typical live loads for steam plant
floors are as follows:
Basement and operating floors, 200 lb/ft2 (9.6 kPa)
Mezzanine, deaerator, and miscellaneous operating floors, 200 lb/ft2 (9.6 kPa)
Office, laboratory, instrument shop, and other lightly loaded areas, 100 lb/ft2
(4.8 kPa)
Live loads for actual design should be carefully reviewed for any special conditions
and actual loads applicable. Live loads for equipment floors should be based on the
assumption that the floor will be used as laydown for equipment parts during maintenance. Load is normally based on the heaviest piece removed during maintenance.
In addition to the live and dead loads, the following loadings should be provided for:
Piping load. Weight of major piping sized and routed should include weight of
pipe, insulation, and hydraulic weight (pipe full of water) in addition to any
shock loads. Pipe hanger loads should be doubled for design of supporting steel.
In congested piping areas, increase live load on the supporting floor by 100 lb/ft2
(4.8 kPa). Structural steel should be provided to adequately support all mechanical piping and electrical conduit. Provision should be made to accommodate
expansion and contraction and drainage requirements of the pipe. Piping connections must be made to preclude rupture under the most adverse conditions
expected. Pipe supports should be close coupled to supporting structures when
severe seismic conditions are expected.
Wind loading. The building should be designed to resist the horizontal wind
pressure available for the site on all surfaces exposed to the wind.
Seismic loading. Buildings and other structures should be designed to resist seismic loading in accordance with the zone in which the building is located.
Equipment loading. Equipment loads are furnished by the various manufacturers
of each equipment item. In addition to equipment dead loads, impact loads,
short-circuit forces for generators, and other pertinent special loads prescribed
by the equipment function or requirements should be included. Ductwork, flue
gas breeching, stacks, and other hot equipment must be such that expansion and
contraction will not impose detrimental loads and stresses on related structures.

3.16

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Snow loading. The building should be designed to resist the snow loading in
accordance with local building codes.
Architectural Treatment
The architectural treatment should be developed to harmonize with the site conditions, both natural and man-made. Depending on location, the environmental compatibility may be the determining factor. In other cases the climate or user preference, tempered
with aesthetic and economic factors, will dictate architectural treatment.
Special circumstances, such as areas where extended periods of very high humidity,
frequently combined with desert conditions giving rise to heavy dust and sand blasting
action, will require indoor construction with pressurized ventilation. In general, steps to
take for architectural treatment are the following.
Control rooms, offices, locker rooms, and some outbuildings should be enclosed
regardless of the enclosure selected for the main building.
Equipment room size should provide adequate space for equipment installation,
maintenance, and removal. A minimum of aisle space between items should be
4 ft (1.2 m) if feasible. Provide a minimum of 8 ft (2.4 m) clearance between
boilers 60,000 lb/h (27,210 kg/h) and larger. If future expansion is planned,
room size should be based on future requirements. Equipment room construction should allow for equipment removal and include double doors and steel
supports for chain hoists. Adequate ventilation and heating should be included.
Attenuation of noise should be considered in room design.
Provide openings or doorways for passage of the largest equipment units. Make
openings for ventilation louvers, breechings, and piping where necessary. Fire
doors, fire shutters, or a combination of both may be required.
For multiple-floor installations, provide a freight elevator.
Furnish necessary shower room, toilet room, and locker facilities for operating
personnel for both sexes in buildings. The plant should contain a sampling laboratory space, storage area, small repair area, control room (in larger plants), generator room, lunch room, compressor room, chemical storage area, and office
space for supervisors and clerks. Parking spaces for plant personnel and visitors
should be provided near the boiler plant.
Finish of plant interior walls and tunnels should have a coating that will permit
hose down or scrubbing of areas.
Concrete should be in accordance with American Concrete Institute (ACI) standards.
Special Considerations
Other considerations for plant structures are the following.
Crane bay. Provide removable openings in floor above major equipment for
removal by station crane (if provided).
Provide hoists and supports for maintenance on pumps, fans, and other heavy
equipment. Provide a beam into the plant to hoist equipment to an upper level.
All anchor bolts for equipment should be sleeved to allow adjustment for final
alignment of equipment.

Heating, Ventilating, and Air-Conditioning (HVAC) Systems


System analysis and design procedures for HVAC system design provided by
ASHRAE should be followed, unless otherwise stated or specifically directed by other
criteria.

3.17

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

When a computer program is required, provisions for showing an estimate of the


hourly space heating requirements and hourly performance of the heating system can be
structured. When manual computation is used, the heating load estimates should be in
accordance with the current edition of ASHRAE HandbookFundamentals.

Drainage
The following drainage issues should be considered.
Drains that may contain coal or oil should have suitable separators to separate
the coal or oil from drainage before being connected to sewer lines.
Provide drains and connect to storm water drainage system for diked areas for
aboveground oil storage.
Maximum temperature of effluent to the drain system is determined by governing codes. A temperature regulating valve is normally be used to inject potable
water into the high-temperature waste stream from the blowdown tank. This
technique may also be used on other equipment as required.

Plant Safety
The safety features described in the following paragraphs should be incorporated into
the steam plant design to assist in maintaining a high level of personnel safety.
Equipment should be arranged with adequate access space for operation and for
maintenance. Wherever possible, auxiliary equipment should be arranged for
maintenance handling by monorails, wheeled trucks, or portable A-frames if
disassembly of heavy pieces is required for maintenance.
Safety guards should be provided on moving or rotating parts of all equipment.
All valves, specialties, and devices needing manipulation by operators should be
accessible without ladders, and preferably without using chain-wheels.
Valve centers should be mounted approximately 7 ft (2.1 m) above floors and
platforms so that rising stems and bottom rims of hand wheels will not be a hazard. Provide access platforms for operations and maintenance of all valves and
equipment over 7 ft (2.1 m) above the floor.
Stairs with conventional riser-tread proportions should be used. Vertical ladders
should be installed only as a last resort.
All floors, gratings, and checkered plates should have nonslip surfaces.
No platform or walkway should be less than 30 in. (762 mm) wide.
Toe plates, fitted closely to the edge of all floor openings, platforms, and stairways, should be provided in all cases. Handrails should be provided on platforms and floor openings.
Not less than two exits should be provided from catwalks, platforms longer than
10 to 15 ft (3.1 to 4.6 m) in length, boiler aisles, floor levels, and the steam plant.
Emergency lighting should be provided for all modes of egress.
All floors subject to wash down or leaks should be sloped to floor drains.
All areas subject to lube oil or chemical spills should be provided with curbs and
drains.
If the plant is of semi-outdoor or outdoor construction in a climate subject to
freezing weather, weather protection should be provided for critical operating
and maintenance areas such as the firing aisle, boiler steam drum ends, and soot
blower locations.
Adequate illumination should be provided throughout the plant. Illumination
should comply with requirements of the Illuminating Engineering Society of
North Americas (IES) The Lighting Handbook (IES 2011).

3.18

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Comfort air conditioning should be provided throughout control rooms, laboratories, offices, and similar spaces where operating and maintenance personnel
spend considerable time.
Mechanical supply and exhaust ventilation should be provided for all of the
steam plant equipment areas to alleviate operator fatigue and prevent accumulation of fumes and dust.
The noise level should be reduced to at least the recommended maximum levels
of the Occupational Safety and Health Administration (OSHA) (2013). Use of
fan silencers, compressor silencers, mufflers on internal combustion engines,
and acoustical material will often be required.
Color schemes should be psychologically restful except where danger must be
highlighted with special bright primary colors.
Each equipment item should be clearly labeled in block letters identifying it by
both equipment item number and name.

Central Plant Security


Restricted access and proper location of exposed intakes and vents must be designed
into the central plant layout to protect the facility from attack and protect people from
injury. Care must be taken in locating exposed equipment, vents, and intakes, especially
at ground level. Aboveground and at least 10 to 15 ft (3.1 to 4.6 m) from access to intake
face is preferred. When this is not possible, fencing around exposed equipment, such as
central plant intakes, should be kept locked at all times to prevent unauthorized access.
Ensure that fencing is open to airflow so it does not adversely affect equipment performance. Air intakes should be located above street level if possible, and vents should be
directed so they cannot discharge directly on passing pedestrians or into an air intake of
the same or an adjacent facility.

Steam Generators
This section describes the specifics of gas, oil, and coal fuel-fired, steam-generating
water-tube boilers and components with steam capacities between 20,000 and
250,000 lb/h (9070 and 113,375 kg/h) and maximum pressures of 450 psig (3105 kPa)
saturated and 400 psig/700F (2760 kPa/371C) superheated.
Typical stoker boiler efficiency is between 80% and 82%. ACFB boiler efficiency is comparable to pulverized coal (PC) boiler efficiency (86% to 88%).
Fluidized-bed boilers have gained acceptance in the industrial and utility sectors
by providing an economical means of using a wide range of fuels while meeting
emissions requirements without installing FGD systems, such as wet and dry
type scrubbers.
Emission reductions. Sulfur capture is accomplished by injecting a sorbent, such
as limestone or dolomite, into the furnace along with coal and other solid fuels.
The boiler plant should include storage and equipment for handling limestone.
Optimum sulfur capture and reduced thermal NOx emissions are achieved by
maintaining a combustion temperature at approximately 1550F (843C), which
is lower than other coal-firing technologies. Sulfur is removed as calcium sulfate
in the baghouse and either landfilled or sold.
Unique components. ACFB boilers, in addition to having components common
to other combustion technologies (superheater, air heater, steam coil air preheater, economizer, sootblowers, etc.), have unique components. The following

3.19

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

unique ACFB boiler components are described in more detail in individual sections later in this chapter:
Lower combustor
Upper combustor and transition zone
Solids separator
Solids reinjection device
External heat exchanger (optional)

Boiler Design and Type


Boilers are designed and constructed in accordance with Section 1 of the ASME
Boiler and Pressure Vessel Code (2013). The boilers are normally natural circulation and
two-drum design. Coal-fired boilers are balanced draft, and gas/oil-fired boilers are
forced draft.

Boiler Construction Options


Construction Types
The specific type of boiler construction will depend on the boiler size, type of firing,
and life-cycle costs (LCCs). Three boiler construction types are available: shop-assembled package units, field-assembled modular units, and field-erected units.
Package Units
Package units are completely assembled before leaving the boiler manufacturers factory. For this reason, the quality of workmanship is generally better and the field installation costs are considerably lower than those for the modular and field-erected units.
Package units covered by this chapter are limited to stoker-fired boilers with steam capacities of approximately 50,000 lb/h (22,675 kg/h) and below and gas- and/or oil-fired units
of approximately 200,000 lb/h (90,700 kg/h) and below.
Modular Units
Modular units are too large for complete shop assembly. Some of the components,
such as the boiler furnace, superheater, boiler tube bank and economizer, and air heater
are assembled in the manufacturers shop prior to shipment for final erection at the field
site. Modular units should be subject to better quality control due to manufacturing plant
conditions. Since component assembly has taken place in the manufacturers shop, the
field man-hour erection time will be reduced. Modular units are limited to stoker-fired
boilers ranging in steam capacity from approximately 50,000 to 150,000 lb/h (22,675 to
68,025 kg/h).
Field-Erected Units
Field-erected boilers have numerous components, such as the steam drum, the lower
(mud) drum, furnace wall panels, superheater sections, generating tube banks, economizer, and air heater, plus the flue gas and air duct sections, that are assembled at the job
site. Therefore, they take longer to install than either a package or a modular unit. Fielderected units are available from about 40,000 to 250,000 lb/h (18,140 to 113,375 kg/h)
and, if required, much larger. Field-erected stoker-fired boilers are available in this size
range, and PC-fired units may be specified for boilers with capacities of 100,000 lb/h
(45,350 kg/h) and above. Field-erected ACFB boilers are 80,000 lb/h (36,280 kg/h) or
larger. Gas- and oil-fired boilers are field erected for capacities of 200,000 lb/h
(90,700 kg/h) or larger. Field-erected units are the only boilers available for any of these
technologies above 200,000 lb/h (90,700 kg/h).

3.20

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Available Fuels
Natural Gas
Natural gas is the cleanest burning of the widely used commercially available fuels. It
contains virtually no ash, which reduces design, building, and operating costs. This also
eliminates the need for particulate collection equipment such as baghouses or electrostatic precipitators. Thorough mixing with combustion air allows low excess air firing.
The high hydrogen content of natural gas compared to oil and coal causes more water
vapor to be formed in the flue gas. This water takes heat away from the combustion process, making less heat available for steam generation, which lowers the boiler efficiency.
Natural Gas Analysis
Two types of analyses of natural gas are commonly used. Proximate analysis provides the percentage content by volume of methane, ethane, carbon dioxide, and nitrogen.
Ultimate analysis provides the percentage content by weight of hydrogen, carbon, nitrogen, and oxygen.
Fuel Oil
Compared to coal, fuel oils are relatively easy to handle and burn. Ash disposal and
emissions are negligible. When properly atomized, oil has characteristics similar to those
of natural gas. Even though oil contains little ash, other constituents such as sulfur,
sodium, and vanadium present problems. These concerns include emission of pollutants,
external deposits, and corrosion.
Fuel Oil Analysis
Historically, petroleum refineries have produced six different grades of fuel oil. Fuel
oils are graded according to gravity and viscosity as defined by ASTM D396 specifications (2013) with No. 1 being the lightest and No. 6 being the heaviest.
Coal
For the purpose of boiler design, domestic U.S. coals are divided into four basic classifications: lignite, sub-bituminous, bituminous, and anthracite. Anthracite, however, is
not normally used in the U.S. for boiler fuel because it requires special furnace and burner
designs due to its low volatile content. In general, these coal classifications refer to the
ratio of fixed carbon to volatile matter and moisture contained in the coal, which increases
with the action of pressure, heat, and other agents over time as coal matures. Volatile matter consists of hydrocarbons and other compounds that are released in gaseous form when
coal is heated. The amount of volatile matter present in a particular coal is related to the
coals heating value and the rate at which it burns. The volatile matter to fixed carbon
ratio greatly affects boiler design, because the furnace dimensions must allow the correct
retention time to properly burn the fuel.
Biomass
The most common types of biomass fuel are wood chips, wood pellets, waste products from mills and malt houses, and biomass grown for fuel purposes (straw and grain)
(Penny 2000). Biomass combustion technologies include the traditional grate stokers and
gasifiers. Gasifiers provide better combustion control and lower emission levels (Inwood
2011).

Combustion Technology Selection


Exclusionary Factors
Gas- and oil-fired boilers are available over the entire size range of boilers. Their use
is limited to areas where these fuels are economically available. Stoker-fired boilers are

3.21

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

available for the entire load range covered by this chapter. PC boilers are available in
capacities of 100,000 lb/h (45,350 kg/h) and above. ACFB boilers are available in capacities of 80,000 lb/h (36,280 kg/h) and above. PC-fired units were used in capacity ranges
below 100,000 lb/h (45,350 kg/h) prior to the advent of ACFB package boilers, but with
the new ACFB designs seldom is the conclusion that PC firing is the preferred method of
firing coal.
When rapid load swings are expected, stoker-fired units may be eliminated because
of their inferior response to these conditions. When economics dictate the use of lowgrade fuels, including those of high or variable ash content or high sulfur content, then
stoker-fired and PC systems may be eliminated in favor of ACFB systems. If none of
these firing systems are excluded by these factors, then the choice between firing systems
must be made on the basis of a life-cycle cost analysis (LCCA).
Base Capital Cost
The base capital cost of a dual firing system is the total price of purchasing and
installing the entire system, including the boiler, furnace, stoker or pulverizer or fluidization system, fans, flues, ducts, and air quality control equipment. PC-fired and ACFB
boilers are more expensive than stoker-fired boilers of a given capacity, in part because
they have a larger furnace to provide space and time for the combustion process to go to
completion. Approximately 60% to 90% of the ash content of coal passes through the unit
along with the gases of combustion. Tube spacing within the unit has to be provided in
order to accommodate this condition and the ability of this ash to cause slagging and fouling of the heating surface. These factors can increase the size of the boiler and its cost.
PC-fired units have been replaced by package boilers in capacity ranges below
100,000 lb/h (45,350 kg/h). Gas, oil, and PC boilers require a flame failure system, which
increases their cost. The total cost of an ACFB boiler addition is offset by not requiring
flue gas desulfurization (FGD) or selective catalytic reduction (SCR). Selective noncatalytic reduction (SNCR) is required on ACFB boilers in place of SCR.
Average Boiler Duty
The remaining expenses calculated for a LCCA are all functions of the average boiler
duty. This value is based on the estimated annual boiler load during the expected life of
the plant. It is calculated as follows:
(average load [lb/h or kg/h] / rated capacity) (hours of operation/8760)
= average boiler duty
For example, if a 100,000 lb/h (45,350 kg/h) boiler operates at an average load of
75,000 lb/h (34,013 kg/h) for 8000 hours per year out of a possible 8760 hours, the average boiler duty is 68%.
Fuel Flexibility
When economically feasible, satisfying steam requirements with more than one type
of fuel offers significant advantages. Problems with only one fuels source, transportation,
handling, or firing system will not stop steam production. The flexibility of alternative
fuel supplies can be a powerful bargaining tool when negotiating fuel supply contracts.
Fuel Selection Considerations
The use of natural gas has the lowest first cost provided there is adequate supply in a
nearby suppliers pipeline. Natural gas does not require storage facilities; however, it is
subject to interruption and possible curtailment. The use of propane fuel systems can mitigate the problems associated with interruptible natural gas supply. Although diesel oil

3.22

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

burns more efficiently than natural gas, oil requires on-site storage and pumping facilities.
Because oil has the potential to contaminate ground water, storage facilities are required
to include spill containment and leak detection systems.
Power Cost
Gas-fired boilers have the lowest electrical energy requirements. Oil-fired units are
next due to oil pumping and heating needs. Auxiliary power requirements for gas and oil
boilers are considerably less than for coal-fired units because ash handling, coal handling,
sorbent handling, and pollution control systems are not needed. A comparison can be
approximated by listing the fans and drives with their respective duties and the sizes of
each.
Maintenance Costs
Gas-fired boilers have the lowest maintenance cost. Oil-fired boiler installations have
higher costs than gas-fired systems due to oil pumping needs, oil storage requirements,
and boiler corrosion and external deposits on heat transfer surfaces resulting from sulfur,
sodium, and vanadium in the oil. ACFB boilers have higher maintenance when compared
to PC boilers. The abrasive action of the solids circulating through the combustor and solids separator causes wear. ACFB systems are more complicated, with more components,
which add to the maintenance cost. Maintenance costs for a PC-fired unit are generally
higher than those for a stoker-fired boiler due to the higher-duty requirements of such
items as the pulverizers, primary air fans or exhausters, electric motors, and coal lines and
the greater number of sootblowers. Maintenance costs are also a reflection of the hours of
operation and average boiler duty.
Operating Costs
Operating expenses include manpower, sorbent, and other costs incurred on a continuing basis while the plant is in operation. Manpower requirements for oil-fired boilers
are somewhat higher than those for gas boilers because of the fuel storage and increased
handling concerns associated with oil. Coal-fired technologies require considerably more
manpower than either oil or gas. Fuel handling, ash handling, and pollution control systems account for the majority of the increase in operating costs. Even though stoker-fired
and PC boilers are less complicated than ACFB boilers, stoker and PC boilers, unlike
ACFB boilers, must include scrubbers. The evaluation of operating costs among coal firing technologies is site specific and must include all relevant factors.

Burners
Burners Classification
Burners may be classified into atmospheric (natural draft) and power types.
Atmospheric burners are used for small- and medium-size gas boilers. This type
of burner depends on natural draft provided by the stack for the introduction of
combustion air. Where adequate stack height cannot be provided and where the
breeching must be abnormally long, induced draft fans can be used to provide
the necessary draft. Atmospheric burners are generally of the ribbon type with
one or more pilots strategically located so as to provide a smooth light-off.
These burners are by far the simplest available and operate on low gas pressure
(3 to 5 in. H2O [76 to 127 mm H2O]).
A power burner incorporates a blower that supplies combustion air at a pressure
adequate to compensate for the airflow resistance offered by the boiler. The
blower also provides a means to include pre-purge and post-purge cycles. This

3.23

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

adds an element of safety in that the boiler is scavenged of any possible combustibles prior to lighting off and subsequent to the burning cycle.
Burners used for firing oil are equipped with means to atomize the oil for proper combustion. With small burners, pressure atomization is used. With larger modulating type
burners, atomization is accomplished with the use of compressed air or steam.
Combination gas/oil burners are popular due to the uncertainty of fuel availability and
cost fluctuations and restrictions imposed by utilities for limiting gas consumption during
certain high-demand periods. They incorporate the features of both gas and oil burners
and can be easily switched from one fuel to the other. A well-designed combination
burner should allow changeover of fuel without adjustment of dampers and nozzles.
Burner Design
State-of-the-art burner design calls for low excess air operation to improve the boiler
thermal efficiency (reduced exit gas temperature and dry gas loss) as well as to reduce
NOx emissions. Coal burners should be specifically designed for PC and compatible with
the gas or oil ignitors to be supplied and produced by a qualified manufacturer. Note that
in the case of gas, oil, and PC-fired units a flame safety system is also required.
Emission Control
Emission control options are presented in the Instrumentation and Control section of
this chapter.

Boiler Control
Boilers should be controlled to provide the required operating steam pressure and/or
water temperature while maintaining safe operation. The following controls are applicable for district-heating-scale boilers.
Operating temperature/pressure controls are two-position or modulating type
controls arranged to start and stop or modulate the burner when applicable.
Operating limit controls serve to shut down the burner if the temperature or
pressure of the system continues to increase with the boiler in the low-fire or
minimum-fire mode.
High-limit temperature/pressure controls have manual reset devices arranged to
shut down the burner upon occurrence of abnormal conditions.
Low water control often has two functions with steam boilers: to cycle a boiler
feedwater pump at a given water level and to shut down the burner upon further
lowering of the boiler water level. When used in this manner, the cutoff should
be of the automatic reset type. An auxiliary low water level control set at an
abnormally low water level to shut down the burner is generally adequate.
High/low gas/oil pressure controls are arranged to shut down the burner when
abnormal pressures occur. These are manual reset devices. A time delay may be
required on the low oil pressure switch to allow time for pressure buildup.
Flame failure controls are coupled with a programmer and arranged to shut
down the burner upon failure of either pilot or main flame. Insurers and underwriters require manual reset for this control function. When applied to a forceddraft burner, the aforementioned programmer also serves the function of programming the firing cycle.
With power burners, the combustion air damper can be controlled by the position of the fuel valves through a linkage and cam arrangement. As the fuel
valves modulate in response to firing demand, the air damper modulates proportionally. This type of control provides reasonably good results. With large boil-

3.24

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

ers it is desirable to maintain even closer control of the combustion air to


improve burner efficiency and to control the quality of flue gas emissions. This
is accomplished by independent control of the combustion air damper. With this
system, a flue gas analyzer is used that continuously measures the flue gas products of combustion and adjusts the combustion air damper or fuel valve for maximum combustion efficiency.
Sequencing control for multiple boilers as required by load demand is desirable
when two or more boilers are installed in parallel. This can be accomplished by
utilizing a steam flowmeter or water temperature sensor, when applicable, to
energize an additional boiler when the load cannot be met by the boilers already
on-line. This piloting phase of the control would only serve to activate the controls of an additional boiler. Additional pressure or temperature sensors located
in the common steam or hot-water header can be used via a controller to modulate all activated boilers on-line in parallel. With this type of control, all boilers
will fire at approximately the same rate and will feed the common header in a
uniform manner.

Primary Air
Pulverized coal firing requires heated primary air to dry the surface moisture on the
coal fed to the pulverizers and to provide the conveying medium for getting the finely
ground PC from the pulverizer to the burner and out to the furnace. This primary air is
supplied either by the forced-draft fan for a hot primary air system or by a separate cold
primary air fan for pressurized pulverizer systems or is drawn through the pulverizer by
an exhauster fan in the negative-pressure pulverizer system.

Economizer
The purpose of the economizer is to raise the temperature of the boiler feedwater
while lowering the flue gas temperature. Adding equivalent heat transfer surface in the
economizer is usually less expensive than adding the heating surface in either the furnace
area or the boiler convection tube bank.
Economizers are generally used on all gas-, oil-, and stoker-fired boilers and may be
the only type of heat recovery used on some stoker-fired boilers and the only type used on
gas/oil-fired boilers. As a rule of thumb, with the common fuels (coal, oil, and gas), steam
generator efficiency increases about 2.5% for each 100F (37.8C) drop in exit gas temperature. By putting flue gas to work, air heaters and economizers can improve boiler unit
efficiency by 6% to 10% and thereby improve fuel economy.

Water Treatment and Makeup


Boiler feedwater treatment has a direct bearing on equipment life. Condensate receivers, filters, polishers, and chemical feed equipment must be accessible for proper management, maintenance, and operation. Depending on the temperature, pressure, and quality
of the heating medium, water treatment may require softeners, alkalizers, deaerators, and/
or demineralizers (for systems operating at high temperatures and pressures). Deaerators
eliminate oxygen and carbon dioxide from the feedwater. Chemicals can be fed using several methods or a combination of methods, depending on the chemical(s) used (e.g., chelants, amines, oxygen scavengers). See Chapter 8 for more information on water treatment
methods.
Steam condensate is collected in receivers or surge tanks and pumped back to the
boiler. The receivers and associated pump(s) serve as a reservoir for condensate and
makeup water. The boiler feed pump provides system condensate and water makeup back

3.25

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

to the boiler. Calculations must be made to determine the necessary capacity of the
receiver tanks to accommodate the condensate required to fill the distribution system plus
cold startup reserve. On startup with a cold system, the boiler should operate at full
capacity and there should be a period of time between system startup and the return of
condensate to the receiver tank. During this time, there should be enough receiver reserve
capacity to satisfy the makeup water requirements of the boiler. This approach minimizes
the required amount of city makeup water and prevents any waste of water due to overflow of the tank.

Sound and Vibration


Proper space planning is key to sound and vibration control in central plant design.
For example, central plants are frequently located at or below grade, which provides a
very stable platform for vibration isolation and greatly reduces the likelihood of vibration
being transmitted into the occupied structure, where it can be regenerated as noise. Also,
locating the central ventilation louvers or other openings distant from noise- and vibration-sensitive areas greatly reduces the potential for problems in those areas. Louvers
should be installed to prevent unwanted ambient air or water from being entrained into a
facility. Louvers should be installed high enough to prevent security breaches (above
ground level, where possible). Vibration and noise transmitted both into the space served
by the plant and to neighboring buildings and areas should be considered in determining
how much acoustical treatment is appropriate for the design, especially if the plant is
located near sensitive spaces such as conference rooms or sleeping quarters. See the
Space Considerations section in this chapter for further discussion of space planning.
Acoustical considerations must also be considered for equipment outside the central
plant. For example, roof-mounted cooling tower fans sometimes transmit significant
vibration to the building structure and generate ambient noise. Many communities limit
machinery sound pressure levels at the property line, which affects the design and placement of equipment.

Seismic Issues
Depending on code requirements and the facilitys location with respect to seismic
fault lines, seismic bracing may be required for the central plant equipment. For instance,
a hospital located in a seismically active area must be able to remain open and operational
to treat casualties in the aftermath of an earthquake. Most building codes require that
measures such as anchors and bracing be applied to the plant equipment. Refer to the
local authority responsible for requirements.

Breeching
Breeching is defined as the horizontal duct from the flue gas outlet on the boiler to the
vertical flue stack or chimney (Tao and Laird 1984). Breechings are commonly fabricated
of 10 gage steel and covered with high-temperature insulation. Manufactured breechings
are also available. Breechings are subject to operating temperature variations, ranging
from the ambient temperature of the boiler room to temperatures approaching the that of
the flue gases.
Due to the wide temperature range, provisions to compensate for expansion must be
considered. Manufactured expansion joints are available for this purpose and should be
used whenever the breeching design cannot otherwise accommodate the calculated
changes in length.

3.26

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Plant Stack
The plant stack alone can influence fuel selection and location of plant. The stack
must extend vertically above the highest part of the building or of any adjacent buildings.
In the case of high-rise construction, the cost of the stack and the net space occupied by
the stack passing through each floor can become prohibitively expensive when the boiler
is located at the base of the building. The stack flue gas temperature should be below the
dew point a considerable amount of the time due to low loads, start-up and shutdown, and
normal weather conditions. Drainage of water should be provided due to operating conditions as well as rain and snow. Some of the factors to consider in stack design are:
Flue gas conditionsthe erosive and corrosive constituents, dew-point temperature, and maximum temperature if bypassing the economizer or air preheater
Temperature restrictions that relate to the methods of construction and the type
of stack lining material to be used
Stack and lining material selected to withstand corrosive gases (related to sulfur
in the fuel)
Wind, earthquake, and dead loads
After structural adequacy has been determined, both static and dynamic analyses
of the loads
Potential large deflections of the stack during steady wind conditions.
The plant location, adjacent structures, and terrain
The stacks should be equipped with cleanout doors, ladders, painter trolleys, EPA
flue gas testing ports and platforms, lightning protection, and aviation warning lights.
Stack Design
The stack height calculations are for the effective stack height rather than the actual
height; this is the distance from the top of the stack to the centerline of the opening of the
stack where the flue gas enters. Air and gas flow losses through the inlet air duct, air
heater (air side), windbox, furnace and passes, air heater (gas side) or economizer, gas
cleanup equipment, and other losses through duct and breeching should be plotted and
overcome with fans. The kinetic discharge head, friction losses at the entrance to the
stack, and friction losses in the stack should be provided by the natural draft of the stack.
Barometric pressures adjusted for altitude and temperature must be considered in determining air pressure. The following stack parameters must be determined:
The extreme and average temperatures of ambient air and gas entering the stack
All heat losses in the stack (to find mean stack temperature)
Altitude and barometric corrections for specific volume
Gas weight to be handled; also the infiltration of air and combustion air into the
stack casing and ductwork
Stack draft losses due to fluid friction in the stack and kinetic energy of gases
leaving the stack
The most economical stack diameter and the minimum stack height to satisfy
dispersion requirements of gas emissions
The stack height for required draft (Where scrubbers are used, the temperature
may be too low for sufficient buoyancy to overcome the stacks internal pressure
losses.)
Static and dynamic structural analyses of the wind, earthquake, dead, and thermal loads (Vortex shedding of wind loads must be considered to be assured that
destructive natural frequency harmonics are not built into the stack.)

3.27

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Stack Construction
The stack height and diameter, support, corrosion resistance, and economic factors
dictate the type of construction to be used.
Stacks are generally made of concrete or steel because of the high cost of radial brick
construction. If stack gases are positively pressurized, or if flue gases will be at or below
the dew point of the gases, corrosion-resistant linings must be provided; linings must be
able to withstand temperature excursions that may be experienced in the flue gas if flue
gas scrubbers are bypassed.
Stacks of steel or concrete construction should be insulated to avoid condensation by
not allowing the internal surfaces to drop below 250F (121C). This requirement does
not apply when scrubbers are used with low-temperature discharge (150F to 180F
[65.6C to 82.2C]) into the stack because the flue gas is already below the dew-point
temperature.
A truncated cone at the top of the stack will decrease cold air downdrafts at the
periphery of the stack and will thus help maintain stack temperature, but stack draft will
decrease considerably.

Fuel Train
Fuel trains consist of regulators, gages, valves, test cocks, pressure switches, and
related appurtenances to convey fuel to burners safely. Insurers and underwriters have
standardized the acceptable fuel trains for both oil and gas that are designed for optimum
safety.
These trains are available factory furnished on all commercial and industrial boilers.
Two separate motor-operated valves, in addition to any modulating valves, represent the
key elements. In the case of gas trains, a motor-operated or solenoid vent valve is provided between the two gas valves. This type of train provides a high degree of safety.
Note that the two motor-operated, two-position gas valves in the main gas piping are
slow-opening, fast-closing valves that close in less than 1 s. These valves respond in unison to boiler start-up and shutdown commands and always close on a signal of boiler malfunction, i.e., high pressure, high temperature, low water, flame failure, or high or low gas
or oil pressure.
A main gas pressure regulator should be provided upstream of the gas train. A gas
strainer should be considered and may or may not be required depending on the gas distribution system.

Oil Supply System


Each central boiler plant must be individually analyzed to determine the requirements
of the oil pumping system.
A simple suction system uses the pump normally furnished with the boiler to pump
oil from the storage tank to the boiler/burner. This system works reasonably well as long
as the suction lift of the pump is not exceeded and the tanks are reasonably close to the
boiler room (within approximately 50 ft [15.2 m]). The maximum lift of a gear-driven
pump is equivalent to 28 in. (711 mm) Hg vacuum; however, it is not practical to attempt
to operate at more than 12 in. (305 mm) Hg vacuum.
A pressurized loop system is used whenever the pumping head and suction lift of the
boiler burner pump are exceeded by the piping friction and lift requirements. This system
will ensure that oil is always available at adequate pressure at the point of usage. It also
eliminates the loss of prime at the burner pump as is often experienced with systems utilizing the burner pump only.

3.28

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

When prime is lost (usually due to a leaky check valve or foot valve or an air leak in
the piping), it becomes necessary for the pump to purge the line of air and to draw oil
from the tank into the line. This can require several minutes and create an undesirable
start-up problem.
A pressurized loop system circulates oil from the tank to the point of usage and
returns unused oil to the tank. A pressure relief valve in the loop will maintain positive
pressure in the segment of the loop used to supply the boiler burners. The burner pumps
become secondary pumps and derive oil from the pressurized portion of the loop and
return unused oil to the return side of the loop. The amount of oil storage is dependent on
the desired frequency of delivery, availability, and the rate of fuel consumption. Storage
capacity equal to 3 or 4 weeks of peak consumption is common.
The most acceptable method of storage is the use of buried fuel oil tanks placed in an
accessible location for servicing from a delivery truck. Inside tanks can be used, but these
are subject to stringent requirements dictated by the National Fire Protection Association
(NFPA) and local codes. In general, inside tanks must be housed in three-hour fire-rated
enclosures, be properly vented, and be constructed to retain oil within the enclosure in the
event of a tank rupture or leak. Typically fiberglass tanks are used.
An oil heating system is required for boilers utilizing low-cost heavy oil (generally
No. 5 or 6.) Heating must be provided at the following locations to ensure satisfactory circulation:
Electric or steam heaters should be provided in the tank to reduce the viscosity
of oil for easy pumping. The oil temperature should be maintained at approximately 110F (43.3C).
Underground/outside fuel oil piping should be electrically or steam traced and
insulated. If these lines are not traced and no circulation occurs for a period of
time during severe weather, it may be impossible to pump the system.

Piping
For a central steam heating plant, the major piping includes steam supply, condensate
return, pumped condensate, boiler feed, city water, natural gas, fuel oil, vent, and drain
pipe systems. Additionally, attention to the design of an efficient and maintainable condensate return system for a central steam distribution plant is important. A well-designed
condensate return system increases overall efficiencies and reduces both chemical and
makeup water use.
Piping design must include consideration for thermal expansion. Piping expands with
an increase in temperature. The change in linear and circumferential dimensions must be
compensated for. This can be controlled by the use of expansion joints, expansion loops,
and to some degree by natural piping configuration.
The main steam header should be uniformly sized over its entire length and should
have negligible pressure drops. Steam headers should be dripped to remove condensate.
All steam distribution piping should be taken from the top to ensure distribution of dry
steam. Provision should be made for future connections by providing capped tees with or
without valves.

Combustion Air
Careful consideration should be given to every boiler plant design regarding the provision of combustion air. Even though ventilation air can be used for combustion air purposes, the requirements for each do not always coincide. When properly understood,
combustion and ventilation air can be consolidated into a common system with the neces-

3.29

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

sary redundancy and fail-safe characteristics to ensure the satisfaction of both requirements.
Combustion air requirements. Depending on the boiler efficiency and type of fuel,
6 to 7 cfm (0.17 to 0.19 m3/min) of combustion air is required per generated boiler horsepower. In practice, however, 15% to 30% more air is required to ensure proper fuel-air
mixing and complete combustion. This quantity of excess air is approximately the same
for oil, gas, and solid fuels and can be provided by means of forced ventilation or a direct
connection to the outside.
Code requirements. Regardless of the approach used, the designer should check with
the applicable codes to ensure that proper procedures are being implemented to meet the
code requirements. It is very important that combustion air be provided in a manner that
will not result in the boiler room being under a negative pressure, which can affect
burner performance. Also, in cold-climate regions, care should be exercised to prevent
cold air from coming in contact with water piping and other components and causing
freezing.
Systems. In some boiler rooms, combustion air can be adequately provided by an
intake louver communicating directly with the outdoors. In larger plants, a more positive
method may be desirable. This can be accomplished by installing a makeup air unit
arranged to heat and deliver tempered outdoor air to the plant. Combustion air will eventually be heated in the fire box. Preheating the air, as with a tempered air unit, will not
consume additional energy and will eliminate any possible freeze problems. Return and
outdoor air dampers can be installed at the makeup air unit, allowing control to provide
both ventilation and combustion air. With this type of system, there should be a stationary
louver to serve as a relief means for any quantity of air delivered by the fan unit that is not
used for combustion. This louver will also serve as a secondary combustion air inlet if the
fan unit is out of service. This type of system maintains the boiler room at a neutral or
slightly positive pressure.

Maintenance and Operation


Proper consideration should be given during the preliminary space planning of a project for providing a convenient and efficient plant from the standpoint of O&M. Several
factors that should be considered to enhance the overall design of the boiler plant are:
Storage. Space for storage of tools, materials, and spare parts should be considered. In accordance with owner preference, shelves and cabinets can be included
to allow organized storage and planning.
Work space. Space for performing minor repairs to boiler plant components
should be included. The minimum provision includes space for a work bench,
tool cabinets, etc., with adequate lighting for repair and maintenance work.
Office space. Most major building or campus facilities have a staff of maintenance and operating people to maintain plant records and logs and to perform
the plant duties inherent for efficient operation of the boiler plant. Space should
be provided for this staff so that their activities may be conducted in an area
arranged to limit noise and dust penetration and to provide security.
Facilities. Toilets and wash and locker facilities for both sexes are required for
plants with full-time operating staff and should be considered in plants where
nearly continuous supervision is required.

Commissioning
Because a central plant consumes a major portion of the annual energy operating budget, system commissioning is imperative for new construction and expansion of existing

3.30

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

installations. During the warranty phase, district heating plant performance should be
measured, benchmarked, and course-corrected to ensure design intent is achieved. If an
energy analysis study is performed as part of the comparison between decentralized and
centralized concepts, and/or if life-cycle comparison of the study is part of a Leadership
in Energy and Environmental Design (LEED) project, the resulting month-to-month
energy data should be a good electronic document to benchmark actual energy consumption using the measurement and verification plan implementation.
Ongoing commissioning or periodic recommissioning further ensures that the design
intent is met and that heating is reliably delivered after the warranty phase. Many district
heating systems are designed and built with the intent to connect to facilities in a phased
program, which also requires commissioning. Retrocommissioning or recommissioning
should be considered whenever the plant is expanded or an additional connection is made
to the existing systems to ensure the original design intent is met.
Initial testing, adjusting, and balancing (TAB) also contribute to sustainable O&M.
The TAB process should be repeated periodically to ensure levels are maintained. When
completing TAB and commissioning, consider posting laminated system flow diagrams at
or adjacent to the district heating equipment that indicate operating instructions, TAB performance, commissioning functional performance tests, and emergency shutoff procedures. These documents also should be filed electronically in the district heating plant
computer server for quick reference.
Original basis of design and design criteria should be posted as a constant reminder of
design intent and to be readily available in case troubleshooting, expansion, or modernization is needed. For maintainable, long-term design success, building commissioning
should include the system training requirements necessary for building management staff
to efficiently take ownership and operate and maintain the HVAC systems over the useful
service life of the installation.

CENTRAL PLANT DESIGN FOR MEDIUM- AND HIGH-TEMPERATURE WATER


Introduction
Central plants generating medium-temperature water (MTW) operate with temperatures between 250F and 350F (121.1C and 176.7C), with pressures not exceeding
160 psig (1104 kPa) (ASHRAE 2008, Chapter 14). The usual design supply temperature
is approximately 250F to 325F (121.1C to 162.8C), with a usual pressure rating of
150 psi (1035 kPa) for boilers and equipment. Central plants generating high-temperature
water (HTW) operate at temperatures over 350F (176.7C) and usual pressures of about
300 psig (2070 kPa). The maximum design supply water temperature is usually about
400F (204.4C), with a pressure rating for boilers and equipment of about 300 psig
(2070 kPa). The pressure-temperature rating of each component must be checked against
the systems design characteristics. The rapid pressure rise that occurs as the temperature
rises above 400F (204.4C) increases cost because components rated for higher pressures are required. At the higher temperature level, the heating capacity of a given distribution system is increased drastically. For example, the heat transfer that can be provided
with 200F (93.3C) supply water and 120F (48.9C) return water is equal to the difference in enthalpy at the two conditions, i.e., approximately 80 Btu/lb (186.1 kJ/kg) of
water. But with 350F (176.7C) supply water and 120F (48.9C) return water, the heating capacity transmitted by the distribution system can be nearly tripled. However, it
should be noted that the higher-temperature system will be more expensive and there will
likely be higher heat losses in distribution. Additionally, without proper design and con-

3.31

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

trol, return temperatures are often much higher and thus the benefits of higher T may not
be fully realized.
MTW and HTW systems are widely used at campus-type facilities such as colleges or
universities and military bases. However, equipment for MTW and HTW systems is more
expensive than for low-temperature water (LTW) systems. Additional components and
materials such as nitrogen tanks for system pressurization, additional expansion joints,
heavier insulation, and high-pressure-rated boilers are also required. With respect to the
distribution systems it should also be noted that desirable materials (from a cost standpoint), such as polyurethane foam insulation and high-density polyethylene (HDPE) jackets, may not be suitable for the higher temperatures.
The design principles for both medium- and high-temperature plants are basically the
same. Throughout this section, HTW refers to both MTW and HTW plants. This section
presents the general principles and practices that apply to HTW plants and distinguishes
them from central LTW plants operating at 250F (121.1C) or below.

HTW Plant Arrangement


An HTW system has the following characteristics.
The system is a completely closed circuit with supply and return mains maintained under pressure. There are no losses from flashing, and heat that is not
used at the customer facilities is returned to the HTW generator. Tight systems
have minimal internal corrosion.
Mechanical equipment that does not control performance of individual customers is concentrated at the central plant.
Piping can slope up or down or run at a variety of elevations to suit the terrain
and the architectural and structural requirements without provision for trapping
at each low point. This may reduce the amount of excavation required and eliminate drip points and return pumps required with steam.
Greater temperature drops are possible and less water is circulated than in LTW
systems, given proper design and control of the consumer interface and the inbuilding equipment.
The pressure in any part of the system must always be well above the pressure
corresponding to the temperature at saturation in the system to prevent the water
flashing into steam.
Customers requiring different water temperatures can be served at their required
temperatures by regulating the flow of water, modulating water supply temperature, placing some units in series, and using heat exchangers or other methods.
The high heat content of the water in the HTW circuit acts as a thermal flywheel, evening out fluctuations in the load. The heat storage capacity can be further increased by adding heat storage tanks or by increasing the temperature in
the return mains during periods of light load.
The high heat content of the heat carrier makes HTW unsuitable for intermittent
operation if rapid start-up and shutdown are desired, unless the system is
designed for minimum water volume and is operated with rapid response controls.
Higher engineering skills are required to design an HTW system than are
required to design a comparable steam or LTW system.
HTW system design requires careful attention to basic laws of chemistry and
physics, as these systems are less forgiving than standard hydronic systems.

3.32

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Basic System
HTW systems require a heat source (which can be a direct-fired generator, a steam
boiler, or a heat exchanger) to heat the water. The expansion of the heated water is usually
taken up in an expansion vessel, which simultaneously pressurizes the system. Heat transport depends on circulating pumps. The distribution system is closed, comprising supply
and return pipes under the same sustaining pressure, although circulating terminal equipment losses will result in a pressure differential between supply and return. Heat transfer
at the customer facilities is through heat transfer surfaces. The basic system is shown in
Figure 3.3.
Most plants are either (1) a saturated steam cushion system, in which the HTW develops its own pressure or (2) a gas- or pump-pressurized system, in which the pressure is
imposed externally. HTW generators and all auxiliaries (such as water makeup and feed
equipment, pressure tanks, and circulating pumps) are located in a central plant. Cascade
HTW generators sometimes use an existing steam distribution system and are installed
remote from the central plant.

Design Considerations
Selection of the system pressure, supply temperature, temperature drop, type of HTW
generator, and pressurization method are the most important initial design considerations.
The following are some of the determining factors.
Type of load (space heating and/or process); load fluctuations during a 24 h
period and a 1 year period (Process loads might require water at a given minimum supply temperature continuously, whereas space heating can permit temperature modulation as a function of outdoor temperature, other climatic
influences, or occupancy effects.)
Customer temperature requirements
Distance between the central plant and space or process heating requirement
Quantity and pressure of steam used for power equipment in the central plant
Elevation variations within the system and the effect of basic pressure distribution
Usually, distribution piping is the major investment in an HTW system. A distribution
system with the widest temperature spread between supply and return will have the low-

Figure 3.3 Elements of HTW system.

3.33

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

est initial and operating costs. Economical designs have a temperature spread of 150F
(65.6C) or higher. The requirements of user systems determine the system selected.
Because the danger of water hammer is always present when the pressure drops to the
point at which pressurized hot water flashes to steam, the primary HTW system should be
designed with steel valves and fittings of 150 psi (1035 kPa) rating. The secondary water,
which operates below 212F (100C) and is not subject to flashing and water hammer,
can be designed for 125 psi (863 kPa) and standard HVAC equipment. Theoretically,
water temperatures up to about 350F (176.7C) can be provided using equipment suitable for 125 psi (863 kPa). But in practice, unless push-pull pumping is used, maximum
water temperatures are limited by the system design, pump pressures, and elevation characteristics to values between 300F and 325F (148.9C and 162.8C).
Many HTW central plants designed for self-generated steam pressurization have a
steam drum through which the entire flow is taken and that also serves as an expansion
vessel. A circulating pump in the supply line takes water from the tank. The temperature
of the water from the steam drum cannot exceed the steam temperature in the drum that
corresponds to its pressure at saturation. The point of maximum pressure is at the discharge of the circulating pump. If, for example, this pressure is to be maintained below
125 psig (863 kPa), the pressure in the drum that corresponds to the water temperature
cannot exceed 125 psig (863 kPa) minus the sum of the pump pressure and the pressure
that is caused by the difference in elevation between the drum and the circulating pump.
Most plants are designed for inert gas pressurization. In most of these plants, the pressurizing tank is connected to the system by a single balance line on the suction side of the
circulating pump. The circulating pump is located at the inlet side of the HTW generator.
There is no flow through the pressurizing tank, and a reduced temperature will normally
establish itself inside. A special characteristic of a gas-pressurized system is the apparatus
that creates and maintains gas pressure inside the tank.
In designing and operating an HTW system, it is important to maintain a pressure that
always exceeds the vapor pressure of the water, even if the system is not operating. This
may require limiting the water temperature and thereby the vapor pressure or increasing
the imposed pressure. Elevation and the pressures required to prevent water from flashing
into steam in the supply system can also limit the maximum water temperature that may
be used and must therefore be studied in evaluating the temperature-pressure relationships and method of pressurizing the system.
The properties of water that govern design are as follows:
Temperature versus pressure at saturation
Density or specific volume versus temperature
Enthalpy or sensible heat versus temperature
Viscosity versus temperature
Type and amount of pressurization
The relationships among temperature, pressure, specific volume, and enthalpy are all
available in steam tables.

Direct-Fired HTW Generators


The direct-fired HTW generators used in central plants are comparable to central
steam boiler plants operating within the same pressure range. The generators should be
selected for size and type in keeping with the load and design pressures as well as the circulation requirements peculiar to HTW. In some systems, both steam for power or processing and HTW are supplied from the same boiler; in others, steam is produced in the

3.34

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

boilers and used for generating HTW; and in many others, the burning fuel directly heats
the water.
HTW generators can be the water-tube or fire-tube type and can be equipped with
any conventional fuel-firing apparatus. Water-tube generators can have either forced circulation, gravity circulation, or a combination of both. The recirculating pumps of
forced-circulation generators must operate continuously while the generator is being
fired. Steam boilers relying on natural circulation may require internal baffling when
used for HTW generation. In scotch marine boilers, thermal shock may occur, caused by
a sudden drop in the temperature of the return water or when the temperature difference
exceeds 40F (8C). Forced-circulation HTW generators are usually the once-through
type and rely solely on pumps to achieve circulation.
Circulation must be maintained at all times while the generator is being fired, and the
flow rate must never drop below the minimum indicated by the manufacturer. Where
gravity circulation steam boilers are used for HTW generation, the steam drum usually
serves as an expansion vessel. In steam-pressurized forced-circulation HTW generators, a
separate vessel is commonly used for expansion and for maintaining the steam pressure
cushion. A separate vessel is always used when the system is cushioned by an inert gas or
auxiliary steam. Proper internal circulation is essential in all types of boilers to prevent
tube failures caused by overheating or unequal expansion.
In HTW plants, the generator is a steam boiler with an integral steam drum used for
pressurization and for expansion of the water level. A dip pipe removes water below the
water line. This dip pipe should be installed so that it picks up water at or near the saturation temperature, without too many steam bubbles. If a pipe breaks somewhere in the system, the boiler must not empty to a point where heating surfaces are bared and a boiler
explosion occurs. The same precautions must be taken with the return pipe. If the return
pipe is connected in the lower part of the boiler, a check valve should be placed in the
connecting line to the boiler to preclude the danger of emptying the boiler. When two or
more such boilers supply a common system, the same steam pressure and water level
must be maintained in each.
Water and steam balance pipes are usually installed between the drums (Figure 3.4).
These should be liberally sized. A difference of only 0.25 psi (1.73 kPa) in the system
pressure between two boilers operated at 100 psig (690 kPa) would cause a difference of
9 in. (229 mm) in the water level. The situation is further aggravated because an upset is
not self-balancing. Rather, when too high a heat release in one of the boilers has caused
the pressure to rise and the water level to fall in this boiler, the decrease in the flow of
colder return water into it causes a further pressure rise, while the opposite happens in the
other boiler. It is therefore important that the firing rates match the flow through each
boiler at all times. Modern practice is to use either flooded HTW heaters with a single
external pressurized expansion drum common to all the generators or the combination of
steam boilers or a direct-contact (cascade) heater.
Proper safety devices for high and low water levels and excessive pressures should be
incorporated in the expansion tank and interlocked with combustion safety and water flow
rate controls. The connection point of the expansion tank used for pressurization should
receive careful consideration, as it greatly affects the pressure distribution throughout the
system and the avoidance of HTW flashing.
The four fundamental methods in which pressure in a given hydraulic system can be
kept at a desired level are a) elevating the storage tank, b) steam pressurization, c) nitrogen, and d) pump pressurization. These methods are discussed in detail in the following
subsections.

3.35

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 3.4 Piping connections for two or more boilers in an HTW system pressurized by
steam expansion and pressurization.

Elevating the Storage Tank


This is a simple pressurization method, but because of the great heights required for
the pressure encountered, it is generally impractical.
Steam Pressurization
This method requires the use of an expansion vessel separate from the HTW generator. Because firing and flow rates can never be perfectly matched, some steam is always
carried. Therefore, the vessel must be above the HTW generators and connected in the
supply water line from the generator. This steam, supplemented by flashing of the water
in the expansion vessel, provides the steam cushion that pressurizes the system. The
expansion vessel must be equipped with steam safety valves capable of relieving the
steam generated by all the generators. The generators themselves are usually designed for
a substantially higher working pressure than the expansion drum, and their safety relief
valves are set for the higher pressure to minimize their lift requirement.
The basic HTW pumping arrangements can be either one pump, in which one pump
handles both the generator and system loads, or two pump, in which one pump circulates
HTW through the generator and a second pump circulates HTW through the system (Figures 3.5 and 3.6). The circulating pump moves the water from the expansion vessel to the
system and back to the generator. The vessel must be elevated to increase the net positive
suction pressure to prevent cavitation or flashing in the pump suction. This arrangement is
critical. A bypass from the HTW system return line to the pump suction helps prevent
flashing. Cooler return water is then mixed with hotter water from the expansion vessel to
give a resulting temperature below the corresponding saturation point in the vessel.
In the two-pump system, the boiler recirculation should always exceed the system circulation because excessive cooling of the water in the drum by the cooler return water
entering the drum, in the case of overcirculation, can cause pressure loss and flashing in
the distribution system. Backflow into the drum can be prevented by installing a check

3.36

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Figure 3.5 HTW piping for one-pump (combined) steam-pressurized system.

valve in the balance line from the drum to the boiler recirculating pumps. Higher cushion
pressure may be maintained by auxiliary steam from a separate generator.
Steam-pressurized vessels should be sized for a total volume VT, which is the sum of
the volume V1 required for the steam space, the volume V2 required for water expansion,
and the volume V3 required for sludge and reserve. An allowance of 20% of the sum of V2
and V3 is a reasonable estimate of the volume V1 required for the steam space. The volume V2 required for water expansion is determined from the change in water volume from
the minimum to the maximum operating temperatures of the complete cycle. It is not necessary to allow for expansion of the total water volume in the system from a cold initial
start. It is necessary during a start-up period to bleed off the volume of water caused by
expansion from the initial starting temperature to the lowest average operating temperature. The volume V3 for sludge and reserve varies greatly depending on the size and
design of system and generator capacity. An allowance of 40% of the volume V2 required
for water expansion is a reasonable estimate of the volume required for V3.
Nitrogen
Nitrogen, the most commonly used inert gas, can be used for gas pressurization. Air
is not recommended because the oxygen in air contributes to corrosion in the system. The
expansion vessel is connected as close as possible to the suction side of the HTW pump

3.37

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 3.6 HTW piping for two-pump (separate) steam-pressurized system.

by a balance line. The inert gas used for pressurization is fed into the top of the cylinder,
preferably through a manual fill connection using a reducing station connected to an inert
gas cylinder. Locating the relief valve below the minimum water line is advantageous
because it is easier to keep it tightly sealed with water on the pressure side. If the valve is
located above the water line, it is exposed to the inert gas of the system. To reduce the
area of contact between gas and water and the resulting absorption of gas into the water,
the tank should be installed vertically. It should be located in the most suitable place in
the central station.
Similar to the steam-pressurized system, the pumping arrangements can be one-pump
or two-pump (Figures 3.7 and 3.8). The ratings of fittings, valves, piping, and equipment
are considered in determining the maximum system pressure. A minimum pressure of
about 25 to 50 psi (172.5 to 345 kPa) over the maximum saturation pressure can be used.
The imposed additional pressure above the vapor pressure must be large enough to prevent steaming in the HTW generators at all times, even under conditions when flow and
firing rates in generators operated in parallel, or flow and heat absorption in parallel circuits within a generator, are not evenly matched. This is critical, because gas-pressurized
systems do not have steam separating means and safety valves to evacuate the steam generated. The simplest type of gas-pressurization system uses a variable gas quantity with or
without gas recovery (Figure 3.9). When the water rises, the inert gas is relieved from the

3.38

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Figure 3.7 Inert gas pressurization for one-pump system.

Figure 3.8 Inert gas pressurization for two-pump system.

expansion vessel and is wasted or recovered in a low-pressure gas receiver from which the
gas compressor pumps it into a high-pressure receiver for storage. When the water level
drops in the expansion vessel, the control cycle adds inert gas from bottles or from the
high-pressure receiver to the expansion vessel to maintain the required pressure. Gas
wastage can significantly affect the operating cost. The gas recovery system should be
analyzed based on the economics of each application. Gas recovery is generally more
applicable to larger systems.
The vessel should be sized for a total volume VT, which is the sum of the volume V1
required for pressurization, the volume V2 required for water expansion, and the volume
V3 required for sludge and reserve. Calculations made on the basis of pressure-volume
variations following Boyles Law are reasonably accurate, assuming that the tank oper-

3.39

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 3.9 Inert gas pressurization using variable gas quantity with gas recovery.

ates at a relatively constant temperature. The minimum gas volume can be determined
from the expansion volume V2 and from the control range between the minimum tank
pressure P1 and the maximum tank pressure P2. The gas volume varies from the minimum V1 to a maximum, which includes the water expansion volume V2. The minimum
gas volume V1 can be obtained from
V1 = P1V2/ (P2 P1)
where P1 and P2 are units of absolute pressure.
An allowance of 10% of the sum of V1 and V2 is a reasonable estimate of the sludge
and reserve capacity V3. The volume V2 required for water expansion should be limited to
the actual expansion that occurs during operation through its minimum to maximum
operating temperatures. It is necessary to bleed off water during a start-up cycle from a
cold start. It is practicable on small systems (e.g., under 1,000,000 to 10,000,000 Btu/h
[1.055 to 10.55 GJ/h]) to size the expansion vessel for the total water expansion from the
initial fill temperature.
Pump Pressurization
Pump pressurization in its simplest form consists of a feed pump and a regulator
valve. The pump operates continuously, introducing water from the makeup tank into the
system. The pressure regulator valve bleeds continuously back into the makeup tank. This
method is usually restricted to small-process heating systems. However, it can be used to
temporarily pressurize a larger system to avoid shutdown during inspection of the expansion tank.

3.40

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

In larger central HTW systems, pump pressurization is combined with a fixed-quantity


gas compression tank that acts as a buffer. When the pressure rises above a preset value in
the buffer tank, a control valve opens to relieve water from the balance line into the
makeup storage tank. When the pressure falls below a preset second value, the feed pump
is started automatically to pump water from the makeup tank back into the system. The
buffer tank is designed to absorb only the limited expansion volume that is required for the
pressure control system to function properly; it is usually small.
To prevent corrosion-causing elements, principally oxygen, from entering the HTW
system, the makeup storage tank is usually closed and a low-pressure nitrogen cushion of
1 to 5 psig (6.9 to 34.5 kPa) is maintained. The gas cushion is usually the variable gas
quantity type with release to the atmosphere.

Direct-Contact Heaters (Cascades)


High-temperature water can be obtained from direct-contact heaters in which steam
from turbine exhaust, extraction, or steam boilers is mixed with return water from the system. The mixing takes place in the upper part of the heater where the water cascading
from horizontal baffles comes in direct contact with steam. The basic systems are shown
in Figures 3.10 and 3.11. The steam space in the upper part of the heater serves as the
steam cushion for pressurizing the system. The lower part of the heater serves as the systems expansion tank. Whereas the water heater and the boiler operate under the same
pressure, the surplus water is usually returned directly into the boiler through a pipe connecting the outlet of the high-temperature water circulating pump to the boiler. The cascade system is also applicable where both steam and HTW services are required. Where
heat and power production are combined, the direct-contact heater becomes the mixing
condenser.

System Circulating Pumps


Forced-circulation boiler systems can be either one-pump or two-pump. These terms
do not refer to the number of pumps but to the number of groups of pumps installed. In
the one-pump system (Figure 3.5), a single group of pumps ensures both generator and
distribution system circulation. In this system, both the distribution system and the generators are in series.
However, to ensure the minimum flow through the boiler at all times, a bypass around
the distribution system must be provided. The one-pump method usually applies only to
systems in which the total friction pressure loss is relatively low, because the energy loss
of available circulating pressure from throttling in the bypass at times of reduced flow
requirements in the district can substantially increase the operating cost.
In the two-pump system (Figure 3.6), an additional group of recirculating pumps is
installed solely to provide circulation for the generators. One pump is often used for each
generator to draw water from either the expansion drum or the system return and to pump
it through the generator into the expansion drum. The system circulating pumps draw
water from the expansion drum and circulate it through the distribution system only. The
supply temperature to the distribution system can be varied by mixing water from the
return into the supply on the pump suction side. Where zoning is required, several groups
of pumps can be used with a different pressure and different temperature in each zone.
The flow rate can also be varied without affecting the generator circulation and without
using a system bypass.
In steam-pressurized systems, the circulating pump is installed in the supply line to
maintain all parts of the distributing system at pressures exceeding boiler pressure. This
minimizes the danger of flashing into steam. It is common practice to install a mixing

3.41

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 3.10 Cascade HTW system.

Figure 3.11 Cascade HTW system combined with boiler feedwater preheating.

connection from the return to the pump suction that bypasses the HTW generator. This
connection is used for start-up and for modulating the supply temperature; it should not
be relied on for increasing the pressure at the pump inlet. Where it is impossible to provide the required submergence by proper design, a separate small-bore premixing line
should be provided.
The push-pull pumping divides the circulating pressure equally between two pumps
in series. One is placed in the supply and is sized to overcome frictional resistance in the
supply line of the heat distribution system. The second pump is placed in the return and is
sized to overcome frictional resistance in the return. The expansion tank pressure is
impressed on the system between the pumps. The HTW generator is either between the
pumps or in the supply line from the pumps to the distribution system (Figure 3.12).

3.42

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Figure 3.12 Typical HTW system with push-pull pumping.

In the push-pull system, the pressures in the supply and the return mains are symmetrical in relation to a line representing the pressure imposed on the system by the pressurizing source (expansion tank). This pressure becomes the system pressure when the
pumps are stopped. The heat supply to equipment or secondary circuits is controlled by
two equal regulating valves, one on the inlet and the other on the outlet side, instead of the
customary single valve on the leaving side. Both valves are operated in unison from a
common controller; there are equal frictional resistances on both sides. Therefore, the
pressure in the user circuits or equipment is maintained at all times halfway between
the pressures in the supply and return mains. Because the halfway point is located on the
symmetry line, the pressure in the user equipment or circuits is always equal to that of
the pressurizing source (expansion tank) plus or minus static pressure caused by elevation
differences. In other words, no system distribution pressure is reflected against the user
circuit or equipment.
While the pressure in the supply system is higher than that of the expansion tank, the
pressure in the return system, being symmetrical to the former, is lower. Therefore, the
push-pull method is applicable only where the temperature in the return is always significantly lower than that in the supply. Otherwise, flashing could occur. This is critical and
requires careful investigation of the temperature-pressure relationship at all points. The
push-pull method is not applicable in reverse-return systems.
Push-pull pumping permits use of standard 125 psi (863 kPa) fittings and equipment
in many MTW systems. Such systems, combined with secondary pumping, can be connected directly to low-temperature terminal equipment in the building heating system.
Temperature drops normally obtainable only in HTW systems can be achieved with

3.43

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

MTW systems. For example, 330F (165.6C) water can be generated at 90 psig (621
kPa) and distributed at less than 125 psig (863 kPa). Its temperature can be reduced to
200F (93.3C) by secondary pumping. The pressure in the terminal equipment then is 90
psig (621 kPa) and the MTW is returned to the primary system at 180F (82.2C). The
temperature difference between supply and return in the primary MTW system is 150F
(65.6C), which is comparable to that of an HTW system. In addition, conventional heat
exchangers, expansion tanks, and water makeup equipment are eliminated from the secondary systems.
Piping and Valves
All pipes, valves, and fittings used in the HTW plant should comply with the requirements of ASME B-31.1, Power Piping (ASME 2012), and National Fuel Gas Code
(NFPA 2012). These codes state that the system should be designed for the highest pressure and temperature actually existing in the piping under normal operation. This pressure
equals cushion pressure plus pump pressure plus static pressure. Schedule 40 steel pipe is
applicable to most HTW systems with welded steel fittings and steel valves. A minimum
number of joints should be used. In many installations, all valves in the piping system are
welded or brazed. Flange connections used at major equipment can be serrated, raised
flange facing, or ring joint. It is desirable to have back-seating valves with special packing
suitable for this service.
The ratings of valves, pipe, and fittings must be checked to determine the specific rating point for the given application. The pressure rating for a standard 300 psi steel valve
operating at 400F (204.4C) is 665 psi (4589 kPa). Therefore, it is generally not necessary to use steel valves and fittings over 300 psi (2070 kPa) ratings in HTW systems.
Because HTW is more penetrating than LTW, leakage caused partly by capillary action
should not be ignored because even a small amount of leakage vaporizes immediately.
This slight leakage becomes noticeable only on the outside of the gland and stem of the
valve, where thin deposits of salt are left after evaporation. Avoid screwed joints and fittings in HTW systems. Pipe unions should not be used in place of flange connections,
even for small-bore piping and equipment.
Loop-type expansion joints, in which the expansion is absorbed by deflection of the
pipe loop, are preferable to the mechanical type. Mechanical expansion joints must be
properly guided and anchored.

Controls
On steam-pressurized cycles, the temperature of the water leaving the generator
should control the firing rate to the generator. A master pressure control operating from
the steam pressure in the expansion vessel should be incorporated as a high-limit override. Inert-gas-pressurized systems should be controlled from the generator discharge
temperature. In the water-tube generators most commonly used for HTW applications,
the flow of water passes through the generator in seconds. The temperature controller
must have a rapid response to maintain a reasonably uniform leaving water temperature.
In steam-pressurized units, the temperature variation must not exceed the antiflash pressure margin. At 300F (148.9C), a 5F (2.8C) temperature variation corresponds to a
5 psi (34.5 kPa) variation in the vapor pressure. At 350F (176.7C), the same temperature variation results in a vapor pressure variation of 10 psi (69 kPa). At 400F (204.4C),
the variation increases to 15 psi (103.5 KPa) and at 450F (232.2 C), to 22 psi
(151.8 kPa). The permissible temperature swing must be reduced as the HTW temperature increases or the pressure margin must be increased to avoid flashing.

3.44

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

The rapid response through the generator makes it necessary to modulate the combustion rate on all systems with a capacity of over a few million Btu/h (kJ/h). In the smaller
size range, this can be done by high/low firing. In large systems, particularly those used
for district heating applications, full modulation of the combustion rate is desirable
through at least 20% of full capacity. On/off burner control is generally not used in steampressurized cycles because the system loses pressurization during the off cycle, which can
cause flashing and cavitation at the HTW pumps. All generators should have separate
safety controls to shut down the combustion apparatus when the system pressure or water
temperature is high. HTW generators require a minimum water flow at all times to prevent tube failure. Means should be provided to measure the flow and to stop combustion
if the flow falls below the minimum value recommended by the generator manufacturer.
For inert-gas-pressurized cycles, a low-pressure safety control should be included to shut
down the combustion system if pressurization is lost. Figure 3.13 shows the basic schematic control diagram for an HTW generator.
Valve selection and sizing are important because of the relatively high temperature
drops and smaller flows in HTW plants. The valve must be sized so that it is effective
over its full range of travel. The valve and equipment must be sized to absorb, in the control valve at full flow, not less than half the available pressure difference between supply
and return mains where the equipment is served. A valve with equal percentage flow
characteristics is needed. Sometimes two small valves provide better control than one
large valve. Stainless steel trim is recommended, and all valve body materials and packing should be suitable for high temperatures and pressures. The valve should have a closeoff rating at least equal to the maximum pressure produced by the circulating pump. Generally, two-way valves are more desirable than three-way valves because of equal percentage flow characteristics and the smaller capacities available in two-way valves. Singleseated valves are preferable to double-seated valves because the latter do not close tightly.

Figure 3.13 Control diagram for HTW generator.

3.45

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Water Treatment
Water treatment for HTW systems is of prime importance. Oxygen introduced in
makeup water immediately oxidizes steel at these temperatures, and over a period of time
the corrosion can be substantial. Other impurities can also harm boiler tubes. Solids in
impure water left by invisible vapor escaping at packings increases maintenance requirements. The condition of the water and the steel surfaces should be checked periodically in
systems operating at these temperatures. See Chapter 8 for a discussion of water treatment methods.

Heat Storage
The high heat storage capacity of water produces a flywheel effect in most HTW systems that evens out load fluctuations. Systems with normal peaks can obtain as much as
15% added capacity through such heat storage. Excessive peak and low loads of a cyclic
nature can be eliminated by an HTW accumulator or storage tank, based on the principle
of stratification (see Chapter 7 for details on thermal storage). Heat storage in an extensive
system can sometimes be increased by bypassing water from the supply into the return at
the end of the mains or by raising the temperature of the returns during periods of low load.

Safety Considerations
A properly engineered and operated HTW system is safe and dependable. Careful
selection and arrangement of components and materials are important. Piping must be
designed and installed to prevent undue stress. When HTW is released to atmospheric
pressure, flashing takes place, which absorbs a large portion of the energy. Turbulent mixing of the liquid and vapor with room air reduces the temperature well below 212F
(100C). With low mass flow rates, the temperature of the escaping mixture can fall to
125F to 140F (51.7C to 60C) within a short distance, compared with the temperature
of the discharge of an LTW system, which remains essentially the same as the temperature of the working fluid. If large mass flow rates of HTW are released to atmospheric
pressure in a confined space (e.g., from the rupture of a large pipe or vessel), a hazardous
condition could exist, similar to that occurring with the rupture of a large steam main.
Failures of this nature are rare if good engineering practice is followed in system design
and proper maintenance is performed.

Other Design Considerations


Other design considerations for HTW central plants are similar to those for steam
plants as described in the section Central Plant Design for Steam.

CENTRAL PLANT DESIGN FOR LOW-TEMPERATURE WATER


Low-temperature hot water is the most common and uniform medium for providing
heating and domestic hot water to users. Discrete loads with steam or high-temperature
requirements are best served individually, with a small, independent system designed for
the discrete load in the most cost-effective approach.
Central plants generating low-temperature water (LTW) operate within the pressure
and temperature limits for low-pressure boilers stated in ASME Boiler and Pressure Vessel Code (2013). The maximum allowable working pressure for low-pressure boilers is
160 psig (1104 kPa), with a maximum temperature of 250F (121C). The major components of an LTW plant are
hot-water generators,
an expansion chamber,
the pumps, and
the water treatment and makeup system.

3.46

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Actual systems generally have additional components such as controls, valves, vents,
etc., but these are not essential to the basic principles underlying the system.

Typical System Arrangements


Hot-Water Generation
LTW can be generated by the following energy sources: hot-water generators or
boilers, steam boilers to water heat exchangers, combined heat and power (CHP)
plants, exhaust gas heat exchangers, and/or incinerator heat exchangers. Figure 3.14
presents a principal diagram of an LTW central plant equipped with steam boilers providing steam to water heat exchangers (Solovyev 1976). The plant includes four major
sections: I steam boilers, II equipment for continuous blowdown heat utilization,

1 = steam boilers, 2 = heat exchangers for heating the primary boiler air,
3= separator of continuous blowdown, 4 = heat exchanger of blowdown utilization,
5 = heat exchanger for preheating the makeup water, 6 = blowdown flash tank, 7 = drains flash tank,
8 = boiler feedwater deaerators, 9 = vapor condenser, 10 = makeup water heater, 11 = boiler feed pumps,
12 = pressure-reducing station, 13 = condensate tank, 14 = condensate pumps,
15 = steam-pressure-reducing valve for the deaerator, 16 = makeup water deaerator,
17 = vapor condenser, 18 and 19 = makeup water preheaters, 20 = makeup feed pumps,
21 = makeup storage tanks, 22 = water filter, 23 = district heating circulating pumps,
24 = district heating heat exchanger, 25 = condensate cooler, 26 = deaerator control valve,
27 = bypass, 28 = water flow control valves

Figure 3.14 Central LTW plant supplied with steam.


(Solovyev 1976)

3.47

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

III water treatment equipment, and IV makeup water equipment and district hot
water supply. The rest of the equipment is listed in the legend of the figure.
Figure 3.15 presents a principal diagram of an LTW central plant supplied with hotwater generators (Solovyev 1976). The plant arrangement is much simplified when compared with the steam supply option, and the capital and operating cost are substantially
reduced. The makeup water deaeration is performed in vacuum deaerators (Oliker 1972,
1989).
Water boilers are generally available in standard sizes from 35,000 to more than
100,000,000 Btu/h (36.92 to more than 105.5 GJ/h), many of which are in the low-pressure
class and are used primarily for space-heating applications in both new and existing systems. A boiler may be purchased as a package that includes the burner, fire chamber, heat
exchanger section, flue gas passage, fuel train, and necessary safety and operating controls.
Water-tube boilers can be field assembled, but fire-tube, scotch marine, and waste-heat
boilers are usually package units. Fire-tube and water-tube boilers are available for gas/oil
firing. If coal is used, either package-type coal-fired boilers in small sizes (less than 20 to
25 MMBtu/h [21.1 to 26.4 GJ/h]) or field-erected boilers in larger sizes are available.
Coal-firing underfeed stokers are available up to a 30 to 35 MMBtu/h (31.7 to 36.9 GJ/h)
capacity; traveling grate and spreader stokers are available up to a 160 MMBtu/h
(168.8 GJ/h) capacity in single-boiler installations. Water-tube boilers contain less water
than equivalent-capacity fire-tube boilers. Figure 3.16 presents a typical water-tube hot-

1 = hot-water generators, 2 = recirculation pumps, 3 = primary air heaters,


4 = heat exchanger heating potable water, 5 = heat exchanger for preheating the makeup water,
6 = potable water pumps, 7 = district heating pumps, 8 = strainer, 9 = vacuum deaerator, 10 = ejector,
11 = gas separator, 12 = ejector water pump, 13 = makeup water pumps,
14 = makeup water back-up pumps, 15 = water storage tank,
16 = makeup water intermediate storage tank, 17 = flow control valve

Figure 3.15 Central LTW plant supplied with hot-water generators.


(Solovyev 1976)

3.48

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

water gas/oil-fired boiler widely used in Europe for central hot-water plants. These boilers
are manufactured in sizes ranging from 16 to 720 MMBtu/h (759.6 GJ/h).

Energy Sources
The energy used by a plant may be natural gas, oil, coal, electricity, or combustible
waste material, though natural gas and fossil-fuel oil (No. 2, 4, or 6 grade) are most common, either alone or in combination. Selecting a fuel source requires detailed analysis of
energy prices and availability. Use of multiple fuels is recommended for plant redundancy
and reliability improvement.

1 = side radiant water-tube bundles, 2 = front radiant water-tube bundles,


3 = back furnace wall-tube bundles, 4 = convective surfaces, 5 = burners, 6 = stack
All dimensions are in metres.

Figure 3.16 Cross section of a dual-fuel (natural gas/oil) hot-water generator


with a capacity of 400 MMBtu/h (422 GJ/h).
(Sasanov 1974)

3.49

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

The availability of fuel oil or coal delivery affects road access and storage for deliveries. Security around the central plant must be considered in the design. The plant should
be equipped with standby electrical power for central plant operation in emergency situations.
Energy for heating may also come from waste heat from a CHP system (see
Chapter 2). Typically, electricity production efficiency is improved when waste heat can
be reclaimed and transferred to a heat source. A heat recovery generator converting the
heat by-product of electric generation to steam or hot water to meet a facilitys heating
needs when the thermal load meets or exceeds the heat rejection of the electric generation
can be cost-effective plant operation.
The two primary considerations in selecting the hot-water generators are the design
and the part-load capability, sometimes called the turndown ratio. The turndown ratio,
expressed in percent of design capacity, is:
Turndown Ratio (%) = 100 (Minimum Capacity/Design Capacity)
The reciprocal of the turndown ratio is sometimes used (for example, a turndown
ratio of 25% may also be expressed as a turndown ratio of 4). The turndown ratio has a
significant effect on system performance; lack of consideration of the heat source systems part-load capability has been responsible for many systems that either do not function properly or do so at the expense of excess energy consumption. The turndown ratio
has a significant effect on the ultimate equipment and/or system design selection.
Figure 3.17 presents a footprint arrangement of a central LTW plant utilizing natural
gas/oil-fired hot-water generators and small auxiliary steam boilers (for deaeration).

System Temperatures
Design temperatures and temperature ranges are selected by consideration of the performance requirements and the economics of the components. Most economic considerations relating to distribution and pumping systems favor using the maximum possible
temperature range T. For an LTW central plant, the maximum supply hot-water temperature is normally established by ASME Boiler and Pressure Vessel Code (2013) as 250F
(121.1C); however, consideration of the availability of distribution system alternatives
for lower temperatures is also a major consideration. The return temperature is typically
170F (76.7C). Further reduction of the return temperature is typically limited by the
users heating systems and hot-water generator limitations. The temperature drop at the
consumer end should be as high as possible. A large temperature drop allows the fluid
flow rate through the system, pumping power, return temperatures, return line heat loss,
and condensing temperatures in cogeneration power plants to be reduced. For new areas
with low heat demand density, the supply and return temperatures can be further reduced
(Zinko et al. 2008). The temperature decrease allows reduction of the piping distribution
cost and increases the cogeneration rate of the CHP plant. For new building construction,
however, the temperature decrease typically results in an increase in the heat transfer area
of the heating elements. Therefore, the reduction of the hot-water supply and return temperatures should be justified by feasibility analysis.
Typically the hot water flows through piping that connects a hot-water generator and
the users water to water heat exchangers. In some small district heating systems the hot
water from the central plant can be supplied directly to the users heat conditioning equipment. In both cases the water recirculates in a closed loop back to the central plant for
reheating.

3.50

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

1= hot-water generators, 2 = forced-draft fans, 3 = primary air fan, 4 = stack,


5 = auxiliary boilers, 6 = feedwater deaerator

Figure 3.17 1100 MMBtu/h (1160.5 GJ/h) footprint of an LTW central plant.
(Solovyev 1976)

Determining the optimum heating water supply and return temperatures requires
design consideration of equipment performance, particularly the energy required to produce the supply water temperature. When designing for energy management, supply
water reset may be considered when peak capacity is not needed, potentially reducing
energy consumption. Energy conservation and management can best be achieved with
computerized design and facility management resources to simulate delivery then monitoring and measuring the actual plant performance.
Figure 3.18 presents a typical supply water temperature and annual load duration
curve from a central plant. The maximum temperature of 250F (121C) in Figure 3.18a
corresponds to the 100% heat load in Figure 3.18b. The minimum temperature of 178F
(81C) in Figure 3.18a corresponds to the minimum heat load of 8% in Figure 2.18b. The
supply water temperature is reset according to the outdoor temperature.

3.51

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

(a)

(b)

Figure 3.18 District heating plant (a) water supply temperature and (b) annual load duration curve.

Heat Exchangers
Heat exchangers offer operational and energy-recovery opportunities for district heating plants. Heat exchangers also provide opportunities for economizing and heat recovery
in a district heating plant (e.g., flue gas exhaust heat recovery). Tube and shell and plate
and frame heat exchangers are typically used.

Thermal Storage
Hot-water thermal storage can be easily used for district hot-water plants. Depending
on the plant design and loading, thermal storage can reduce the equipment requirements
and lower operating costs. Thermal storage may offer multiple strategies for both partload and peak-load efficiency control. Thermal storage systems are covered in detail in
Chapter 7.

Auxiliaries
Numerous pieces of auxiliary support equipment related to hot-water generator operations are not unique to the production plant of a district system and are found in similar
installations. Some components of a central plant deserve special consideration because
of their critical nature and potential effect on operations.
Equipment and layout of a district heating plant should reflect what is required for
proper plant O&M. The plant should have an adequate service area for equipment and a
sufficient number of electrical power outlets and floor drains. Equipment should be
placed on housekeeping pads.

Expansion Tanks and System Pressurization


Expansion tanks serve the purpose of providing a means to pressurize the system and
to allow for expansion of the water when heated. Typically, the same nitrogen or pressurizing pumps are used as in HTW plants as previously described. In smaller systems two

3.52

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Figure 3.19 Typical makeup water and expansion tank piping configuration
for plain steel expansion tank.

basic types of tanks are available, the plain steel hydro-pneumatic tank and the diaphragm-type tank (ASHRAE 2008, Chapter 12):
The plain tank is installed empty. When the system is filled, water enters the tank,
which compresses the air or nitrogen inside the tank. Tank pressure is equal to the
static pressure resulting from the height of the system plus a residual pressure of
5 to 10 psi (34.5 to 69 kPa) at the top of the system. This pressure value may be
referenced as the cold static pressure. The operating pressure is equal to the cold
static pressure plus any pressure developed by the water as it expands. Note that
the pump has no effect on the pressure in the tank. To change tank pressure, water
must be added to the tank; this is not accomplished by pump operation.
The diaphragm-type tank consists of a steel tank enclosing a flexible diaphragm.
When the tank is installed, the diaphragm is fully expanded against the walls of
the tank and is pressurized or precharged to the value of the cold static pressure.
As the system is filled, no water will enter the tank due to the precharged pressure. As the water in the system is heated, the expanded water enters the tank.
Both tank types have suitable applications. The plain tank serves not only as an
expansion tank but also as a collection point for air separated from the heating medium.
The diaphragm tank cannot be used for air collection, and separate automatic air vents are
required. The diaphragm tank is generally smaller size than the plain tank and has a correspondingly reduced weight.
To function properly, the expansion tank must be the single point in the system where
no pressure change occurs. A makeup water pump generally is used to make up water
loss. The pump is typically controlled from level switches on the expansion tank or from
a desired pump suction pressure. Hence, only one tank should be present in the system.
A conventional water meter on the makeup line can show water loss in a closed system. This meter also provides necessary data for water treatment. The fill valve should be
controlled to open or close and not modulate to a very low flow so that the water meter
can detect all makeup.
Figure 3.19 shows a typical piping configuration for a system with a plain steel or air/
water interface expansion tank (ASHRAE 2008, Chapter 12). Note that no valves are

3.53

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

installed between the hydronic system piping and the safety relief valve. This is a mandatory design requirement if the valve in this location is also to serve as a protection against
pressure increases due to thermal expansion.
Equations for sizing the common configurations of expansion tanks follow (from
Chapter 12 of ASHRAE HandbookHVAC Systems and Equipment [2008]).
For closed tanks with air/water interface:
v 2 v 1 1 3 6.5 10 6 3t
V t = V s --------------------------------------------------------------------------------------------P P P P
a

(3.1)

For diaphragm tanks:


v 2 v 1 1 3 6.5 10 6 3t
V t = V s --------------------------------------------------------------------------------------------1 P P
1

(3.2)

where
Vt = volume of expansion tank, gal (L)
Vs = volume of water in system, gal (L)
Pa = atmospheric pressure, psia (kPa)
P1 = pressure at lower temperature, psia (kPa)
P2 = pressure at higher temperature, psia (kPa)
v1 = specific volume of water at lower temperature, ft3/lb (m3/kg)
v2 = specific volume of water at higher temperature, ft3/lb (m3/kg)
= linear coefficient of thermal expansion, in./in.F (m/mK); = 6.5 106 in./in.F
(11.7 106 m/mK) for steel
t = (t2 t1)
where
t1 = lower temperature, F (C)
t2 = higher temperature, F (C)

Example 3.1
Size an expansion tank for a heating water system that will operate at a design temperature
range of 180F to 220F (82.2C to 104.4C). The minimum pressure at the tank is 10 psig
(24.7 psia [170.4 kPa]) and the maximum pressure is 25 psig (39.7 psia [273.9 kPa]). Atmospheric
pressure is 14.7 psia (101.4 kPa). The volume of water is 3000 gal (11,355 L). The piping is steel.
1. Calculate the required size for a closed tank with an air/water interface.
Solution:
For lower temperature t1, use 40F (4.4C).
From Table 3 in Chapter 6 of the 2005 ASHRAE HandbookFundamentals,
v1 at 40F (4.4C) = 0.01602 ft3/lb (1.0001 103 m3/kg)
v2 at 220F (104.4C) = 0.01677 ft3/lb (1.0478 103 m3/kg)

3.54

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Working the problem in I-P units using Equation 3.1:


0.01677 0.01602 1 3 6.5 10 6 220 40
V t = 3000 -------------------------------------------------------------------------------------------------------------------------------------- 14.7 24.7 14.7 39.7
Vt = 578 gal
Working the problem in SI units using Equation 3.1:
1.0468 1.0 1 3 11.7 10 6 104.4 4.4
V t = 11 355 -------------------------------------------------------------------------------------------------------------------------------- 101.4 170.4 101.4 273.9
Vt = 2188 L
2. Determine the tank size if a diaphragm tank were used in lieu of the plain steel tank.
Solution:
Working the problem in I-P units using Equation 3.2:
0.01677 0.01602 1 3 6.5 10 6 220 40
V t = 3000 --------------------------------------------------------------------------------------------------------------------------------------1 24.7 39.7
Vt = 344 gal
Working the problem in SI units using Equation 3.2:
1.0468 1.0 1 3 11.7 10 6 104.4 4.4
V t = 11 355 --------------------------------------------------------------------------------------------------------------------------------1 170.4 273.9
Vt = 1302 L

Pumping System
Centrifugal pumps are the most common pump type used in central hot-water plants.
Circulating base-mounted or vertical pumps can handle water flows of hundreds or thousands of gallons (litres) per minute, with pressures limited only by the system characteristics. Pump operating characteristics must be carefully matched to system operating
requirements (ASHRAE 2008, Chapter 12).

Pump Curves and Water Temperature for Constant-Speed Systems


Performance characteristics of centrifugal pumps are described by pump curves,
which plot flow versus head or pressure, as well as by efficiency and power information,
as shown in Figure 3.20. Large pumps tend to have a series of curves, designated with a
numerical size in inches (millimetres) (10 to 13.5 in. in diameter [254 to 343 mm in diameter]), to represent performance of the pump impeller and outline the envelope of pump
operation. Intersecting elliptical lines designate the pumps efficiency.
The net positive suction head (NPSH) required line represents the required entering
operating pressure for the pump to operate satisfactorily. Diagonal lines represent the
required power at that operating point. The point at which a pump operates is the point at

3.55

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 3.20 Example of manufacturers published pump curve.

Figure 3.21 Pump curve and system curve.

which the pump curve intersects the system curve (Figure 3.21). In Figure 3.20, note that
each performance curve has a defined end point. Small circulating pumps, which may
exhibit a pump curve as shown in Figure 3.21, may actually extend to the abscissa showing a run-out flow and may also not show multiple impellers, efficiency, or NPSH.
Large pumps do not exhibit the run-out flow characteristics. In a large pump, the area
to the right of the curve is an area of unsatisfactory performance and may represent pump
operation in a state of cavitation. It is important that the system curve always intersects
the pump curve in operation, and the design must ensure that system operation stays on
the pump curve.
A pump may be selected by using the calculated system pressure drop at the design
flow rate as the base point value. Figure 3.22 illustrates how a shift of the system curve to

3.56

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Figure 3.22 Shift of system curve caused by circuit unbalance.

the right affects system flow rate. This shift can be caused by incorrectly calculating the
system pressure drop by using arbitrary safety factors or overstated pressure drop charts.
Variable system flow caused by control valve operation, a larger than required control
valve, or improperly balanced systems (subcircuits having substantially lower pressure
drops than the longest circuit) can also cause a shift to the right.
Pumps for closed-loop piping systems should have a flat pressure characteristic and
should operate slightly to the left of the peak efficiency point on their curves. This allows
the system curve to shift to the right without causing undesirable pump operation, overloading, or reduction in available pressure across circuits with large pressure drops.
Operating points may be highly variable, depending on (1) load conditions, (2) the
types of control valves used, and (3) the piping circuitry and heat transfer elements. In
general, the best selection in smaller systems is:
for design flow rates calculated using pressure drop charts that illustrate actual
closed-loop hydronic system piping pressure drops;
to the left of the maximum efficiency point of the pump curve to allow shifts to
the right caused by system circuit unbalance, direct-return circuitry applications,
and modulating three-way valve applications; and
a pump with a flat curve to compensate for unbalanced circuitry and to provide a
minimum pressure differential increase across two-way control valves.
As system sizes and corresponding pump sizes increase, more care is needed in analysis of the pump selection. The Hydraulic Institute offers a detailed discussion of pump
operation and selection to optimize life-cycle costs (HI 2000). The Hydraulic Institute
guide covers all types of pumping systems, including those that are much more sophisticated than a basic HVAC closed-loop circulating system and use much more power. There
are direct parallels, though. The discussion of pump reliability sensitivity shown in
Figure 3.23 is presented in ASHRAE HandbookHVAC Systems and Equipment,
Chapter 12 (2008).

Parallel Pumping
When pumps are applied in parallel, each pump operates at the same head and provides its share of the system flow at that pressure (Figure 3.24). Generally, pumps of
equal size are used, and the parallel-pump curve is established by doubling the flow of the
single-pump curve (with identical pumps).

3.57

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 3.23 General pump operating condition effects.


Courtesy of ASHRAE (2008, Chapter 12, Figure 7)

Figure 3.24 Operating conditions for parallel-pump installation.

Plotting a system curve across the parallel-pump curve shows the operating points for
both single- and parallel-pump operation (Figure 3.24). Note that single-pump operation
does not yield 50% flow. The system curve crosses the single-pump curve considerably to
the right of its operating point when both pumps are running. This leads to two important
concerns: (1) the pumps must be sized to prevent overloading during single-pump operation and (2) a single pump can provide standby service of up to 80% of design flow; the
actual amount depends on the specific pump curve and system curve. As pumps become
larger, or more than two pumps are placed in parallel operation, it is still very important to
ensure in the design that the operating system curve intersects the operating pump curve
and that there are safeties in place to ensure that, should a pump be turned off, the remaining pump and system curves still intersect one another.

3.58

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Figure 3.25 Operating conditions for series-pump installation.

Series Pumping
When pumps are operated in series, each pump operates at the same flow rate and
provides its share of the total pressure at that flow. A system curve plotted across the
series-pump curve shows the operating points for both single- and series-pump operation
(Figure 3.25). Note that the single pump can provide up to 80% flow for standby and at a
lower power requirement.
Series-pump installations are often used in heating and cooling systems so that both
pumps operate during the cooling season to provide maximum flow and head, whereas
only a single pump operates during the heating season. Note that both parallel- and seriespump applications require that the actual pump operating points be used to accurately
determine the pumping point. Adding artificial safety factor head, using improper pressure
drop charts, or incorrectly calculating pressure drops may lead to an unwise selection.

Multiple-Pump Systems
Care must be taken in designing systems with multiple pumps to ensure that if pumps
ever operate in either parallel or series such operation is fully understood and considered
by the designer. Pumps performing unexpectedly in parallel or in series have caused performance problems in hydronic systems such as the following.
In parallel. With pumps of unequal pressure capability, one pump may create a
pressure across the other pump in excess of its cutoff or shutoff pressure, causing flow through the second pump to diminish significantly or to cease. This can
cause flow problems or pump damage.
In series. With pumps of different flow capacities, the pump of greater capacity
may overflow the pump of lesser capacity, which could cause damaging cavitation
in the smaller pump and could actually cause a pressure drop rather than a pressure rise across that pump. In other circumstances, unexpected series operation
can cause excessively high or low pressures that can damage system components.

Standby Pump Provision


If total flow standby capacity is required, a properly valved standby pump of equal
capacity is installed to operate when the normal pump is inoperable. A single standby

3.59

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

may be provided for several similarly sized pumps. Parallel- or series-pump installation
can provide up to 80% standby, which is often sufficient.

Variable-Speed Pumping Application


Centrifugal pumps may also be operated with variable-frequency drives (VFDs),
which adjust the speed of the electric motor, changing the pump curve. In this application,
a pump controller and typically one VFD per pump are applied to control the required
system variable. The most typical application is to control differential pressure across the
hydraulically most distant branch of the piping network, i.e. the critical circuit or consumer.
The controller is given a setpoint equal to the pressure loss of the critical consumer at
design flow. As seen in Figure 3.26, as the control valve closes in reaction to a control signal from points 1 to 2, the differential pressure across the branch rises as the system curve
shifts left up the pump curve. The pump controller, sensing an increase in differential
pressure, decreases the speed of the pump to about point 3, roughly 88% speed. Note that
each system curve is shown with two representations. The system curve is the simple relationship of the flow ratio squared to the head ratio. Under control of a pump controller
with a single sensor across the control valve, the pump decreases in speed until the theoretical zero-flow point, at which the pump goes just fast enough to maintain a differential
pressure under no flow. In this case, the speed is about 88% and the power is about 68%.
In operation, expect there to be a difference in performance.
Figure 3.26 shows the change as a step function, opposite of the way in which the
pump controller functions. Based on the control method, the controller adjustment settings (gain, integral time, and derivative time), system hydraulics, valve time constant,
sensor sensitivity, etc., it is far more likely that a small incremental rise in differential
pressure will have a corresponding small decrease in pump speed as the valve repositions
itself in reaction to its control signal. These changes appear as a series of small sawtoothed system curve and pump curve intersection points and are not of real importance.
What is important is that many factors influence operation, and theoretical variable-speed
pumping is different from reality.

Figure 3.26 Example of variable-speed pump and system curves.

3.60

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Decreasing pump speed is analogous to using a smaller impeller size in the volute of
the pump, and the result is that motor power is reduced by the cube of the speed reduction, closely following the affinity laws. Design flow conditions should be minimal hours
of operation per year; as such, the energy savings potential for a variable-speed pump is
great. In the general control valve application, a reasonably selected equal percentage
valve with a 50% valve authority should reduce flow by about 30% when the valve is
positioned from 100% open (design flow, minimal hours per year) to 90% open. The 10%
change in stroke should reduce pump operating power about 70%.
Depending on system design, direct digital control may also allow more advanced
control strategies. There are various reset control strategies (e.g., cascade control) to optimize pump speed and flow performance. Many of these monitor valve position and drive
the pump to a level that keeps one valve open while maintaining a setpoint for comfort
conditions. Physical operational requirements of the components must be taken into
account. There are some concerns over operating pumps and their motor drives at speeds
less than 30% of design, particularly about maintaining proper lubrication of the pump
mechanical seals and motor bearings. From a practical perspective, 30% speed is in theory a scant 3% of design power, so it may be unnecessary to reduce speed any further.
However, it should be noted that at low speed the variable-speed drives (VSDs) themselves often become very inefficient and thus actual power reductions will often be far
short of the theoretical pumping power reductions indicated. Consult manufacturers for
information on device limitations and efficiencies to ensure that sustainable operating
conditions are being used and projected savings will be realized.
Exceptional energy reduction potential and advances in VFD technology that have
reduced drive costs have made the application of VSDs common on closed hydronic distribution systems. Successful application of a VSD to a pump is not a given, however,
and depends on the designers skill and understanding of system operation.

Pump Connections
Pump suction piping should be at least as large as the nozzle serving the pump, and
there should be minimal fittings or devices in the suction to obstruct flow. Typically,
pump manufacturers prefer five to eight pipe diameters of unobstructed (i.e., without fittings) straight pipe entering the pump. Fittings such as tees and elbows, especially when
there is a change in planar direction, cause water to swirl in the pipe; this can be detrimental to pump performance and may also lead to pump damage. Manufacturers may recommend special fittings (suction diffusers) in applications where piping space is unavailable
to overcome geometry factors. These fittings need to be carefully reviewed in application.
Piping to the pump should be independently supported, adding no load to the pump
flanges, which, unless specifically designed for the purpose, are incapable of supporting
the system piping. Supporting the pipe weight on the flange can cause serious damage
(e.g., breaking the flange) or may induce stresses that misalign the pump. Similar results
can occur when the pipe and pump are improperly aligned to each other or when pipe
expansion and contraction are unaccounted for. Flexible couplings are one way to overcome some of these issues, when both the pump and the pipe are supported independently
of each other and the flexible coupling is not arbitrarily connected to the pump and pipe.
When pumps are piped in parallel, these requirements are extended. Pump entering and
discharge pressures should be equal in operation. In addition to maintaining the recommended straight inlet pipe to the pump, manifold pipe serving the inlets should also have
a minimum of two manifold pipe diameters between pump suction centerlines. Pump discharge manifolds should be constructed to keep discharge velocities less than 10 to 15 fps

3.61

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

(3.05 to 4.6 mps); lower when a check valve is applied, because it is necessary to prevent
hydraulic shock (water hammer). Soft-seating discharge check valves are required on the
pumps to prevent reverse flow from one pump to another. Various specialty valves and fittings are available for serving one or more of these functions.
Evaluation of factors such as those discussed above can help determine the effects of
control and component selection in system operation. The designer cannot assume that
the control system will compensate for poor selections and unnoticed installation mistakes. Anecdotal data suggest that numerous adjustments are required when selection and
operation techniques are poor. There is also a need for system balance when the pump is
correctly sized (when no excess head is added to the pump); the balance device adds head
only for compensating the difference in friction loss as water goes from one terminal
takeoff to another. System flow is reduced to design, and no extra energy input is required
for the pump.

Flow Design Considerations


Two types of energy-efficient system flow designs used today are primary variable
flow and primary/secondary variable flow.
Primary variable-flow systems have variable flow through hot-water generators
and directly pump water to the users (ASHRAE 2008, Chapter 3). Variable flow
can be achieved using two-way automatic control valves at users equipment and
either VFD pumping or distribution pressure control with a bypass valve (Figure
3.27). Both concepts function based on maintaining system pressure, usually at
the hydraulically most distant critical consumer. The load in Figure 3.27 indicates multiple district heating customer loads.
Primary/secondary variable flow systems hydraulically decouple the primary
production system (heating-water source), which is commonly constant flow
(ASHRAE 2008, Chapter 3). A variable-flow secondary piping system distributes the heating water to the point of use (Figure 3.28).
When using either primary or primary/secondary variable-flow designs, the engineer
should understand the design differences between two- and three-way modulating valves.

Figure 3.27 Primary (limited) variable-flow system using distribution pressure control.
Courtesy of RDK Engineers

3.62

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Variable-flow designs vary the flow of heating water through the distribution loop. As terminal units satisfy demand, the valve modulates toward the closed position. In the distribution system, the pump, if at design operating conditions, increases system pressure
above the design point. To compensate, a VFD is typically installed on the distribution
pump. This VFD is usually controlled by a pressure differential sensor, which is located
at the farthest point in the piping loop, to maintain the minimum pressure needed to provide design flow through the last terminal device and associated control valve. As the
pressure differential increases, the sensor sends a signal to the VFD to reduce speed. As
system demand requires increased flow, the control valves modulate open, reducing system differential pressure. The reduction in pressure difference measured causes a corresponding increase in pump speed (and therefore flow) to meet the plant demand.
With primary variable-flow designs, the primary heating-water pumps are of a variable-flow design, again using a VFD. Primary variable-flow systems use modulating twoway control valves to reduce water flow across each consumer interface (as in the primary/secondary system).
With a primary/secondary design, minimum system flow for the heating plant is
accomplished by using a decoupler bypass across the primary and secondary systems.
Because the primary system is constant volume (i.e., the system maintains design flow
during operation as designed when production flow exceeds distribution flow), water
recirculates within the plant to maintain the minimum design flow required by the production plant. Care should be taken during design to ensure minimum flow is achieved
under part-load operation throughout the system. When selecting the distribution system
pump, ensure distribution flow does not exceed production flow, which could cause a
temperature drop if flow exceeds demand (low T syndrome). The engineer should evaluate operational conditions at part load to minimize the potential for this to occur.

Design Guidelines
Guidelines for plant design and operation include the following:
Variable-speed pumping saves energy and should be considered for distribution
system pumping.

Figure 3.28 Primary/secondary variable-flow pumping hot-water system.


Courtesy of RDK Engineers

3.63

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Design hot-water systems for a maximum temperature differential.


Limit the use of constant-flow systems to relatively small central hot-water
plants.
Larger central hot-water plants can benefit from primary/secondary or primary/
secondary/tertiary pumping with constant flow in the central plant and variable
flow in the distribution system.
All two-way valves at the user systems must have proper close-off ratings and a
design pressure drop of at least 20% of the maximum design pressure drop for
controllability. Commercial-quality automatic temperature control valves generally have low shutoff ratings, but industrial valves can achieve higher ratings.

Makeup and Fill-Water Systems


Generally, the district heating system is filled with water through a valved connection
to a domestic water source with a service valve, a backflow preventer, and a pressure
gage. (The domestic water source pressure must exceed the system fill pressure.) Because
the expansion chamber is the reference pressure point in the district heating system, the
water makeup point is usually located at or near the expansion chamber at the district
heating plant.
Safety relief valves should be installed at any point at which pressures may exceed
the safe limits of the system components. Causes of excessive pressures include
overpressurization from the fill system,
pressure increases caused by thermal expansion, and
surges caused by momentum changes (shock or water hammer).
Overpressurization from the fill system could occur because of an accident in filling
the system or the failure of an automatic fill regulator. To prevent this, a safety relief valve
is usually installed at the fill location.

Other System Components


Drain, Shutoff, and Overpressure Protection
All low points should have drains. Separate shutoff and draining of individual equipment and circuits should be possible so that the entire system does not have to be drained
to service a particular item. Whenever a device or section of the system is isolated and
water in that section or device could increase in temperature following isolation, overpressure safety relief protection must be provided.
Strainers
Strainers should be used where necessary to protect system elements and sparingly to
enhance energy efficiency. Strainers in the pump suction must be checked carefully to
avoid pump cavitation. Designers should consider strainer designs that can trap particles
during the commissioning (flush and clean) phase, and allow for mesh size modification
after commissioning, to reduce system pressure loss. Large separating chambers can
serve as main air venting points and dirt strainers ahead of pumps. Automatic control
valves or other devices operating with small clearances require protection from pipe
scale, gravel, and welding slag, which may readily pass through the pump and its protective separator. Individual fine-mesh strainers may therefore be required ahead of each
control valve.
Flexible Connectors
Flexible connectors are sometimes installed at pumps and machinery to reduce pipe
stress. Piping systems should be supported independently of flexible connectors.

3.64

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Insulation
Insulation should be applied to minimize pipe thermal loss and to prevent condensation during hot weather operation.
Other Design Considerations
Other design considerations for HTW central plants are similar to those for steam
plants as described in the section Central Plant Design for Steam.

EMISSION CONTROL AND INSTRUMENTATION


Pollutants and Control Techniques
Combustion of standard fossil fuels (natural gas and ASTM grade oil) in district heating plant boilers result in the following emissions of regulatory concern: nitrogen oxides,
carbon monoxide, sulfur oxides, volatile organic compounds, and particulate matter. The
subsections that follow describe the formation and control of each of the pollutants in
commercial and industrial (district heating) boilers (Cleaver Brooks 2010).

Nitrogen Oxides (NOx)


The principal nitrogen pollutants generated by boilers are nitric oxide (NO) and nitrogen dioxide (NO2), collectively referred to as nitrogen oxides (NOx). In boilers, there are
two types of NOx: thermal NOx (formed at high flame temperatures) and fuel NOx (determined by the amount of nitrogen in the fuel). Another factor affecting NOx formation is
the burner excess air level. Most NOx control technologies for industrial boilers with
inputs less than 100 MBtu/h (105.5 GJ/h) reduce thermal NOx and have little effect on
fuel NOx (Cleaver Brooks 2010).
Boiler NOx controls are of two types: post-combustion (addressing emissions after
NOx formation) and combustion control (preventing the formation of NOx during the combustion process). Post-combustion methods are more expensive than combustion control
techniques and typically are not used on boilers with inputs of less than 100 MMBtu/h
(105.5 GJ/h).
Post-combustion control methods include selective non-catalytic reduction (SNCR)
and selective catalytic reduction (SCR), described as follows.
SNCR requires injection of a NOx-reducing agent (ammonia or urea) into the
boiler flue gases.
SCR requires injection of ammonia into the boiler flue gases in the presence of a
catalyst.
Combustion control techniques include low excess air firing, low nitrogen fuel oil,
burner modifications, water/steam injection, and flue gas recirculation (FGR). Combustion control techniques reduce the NOx emissions by limiting the NOx formation during
the combustion process by lowering flame temperatures. Brief descriptions of the combustion control techniques follow.
Low excess air firing involves limiting the amount of excess air in order to limit
the amount of extra nitrogen and oxygen that enters the flame. This is accomplished through burner design and can be optimized through the use of oxygen
trim controls.
The method of reducing NOx levels from boilers firing distillate oils involves the
use of low nitrogen fuel oil, which can contain up to 20 times less fuel-bound
nitrogen than standard No. 2 oil.
Burner modifications for NOx control involve changing the design of a standard
burner in order to create lower flame temperatures and lower thermal NOx for3.65

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

mation. This technology is most effective when firing natural gas and distillate
fuel oil and has little effect on boilers firing heavy oil. If burner modifications
are used exclusively to achieve low NOx levels, adverse effects on boiler operating parameters such as turndown, capacity, carbon monoxide (CO) levels, and
efficiency may result (Cleaver Brooks 2010).
By injecting water or steam into the flame, flame temperatures are reduced,
thereby lowering thermal NOx formation. Water or steam injection can reduce
NOx up to 80% (when firing natural gas) and can result in lower reductions
when firing oils. However, under normal operating conditions, water or steam
injection can result in a 3%10% boiler efficiency loss.
FGR entails recirculating a portion of relatively cool exhaust gases into the combustion chamber in order to lower the flame temperature and reduce NOx formation. It is currently the most effective technology for fire-tube and water-tube
boilers with inputs below 100 MMBtu/h (105.5 GJ/h).
The current trend with low-NOx technologies is to design the boiler and low-NOx
equipment as a package. This allows the NOx control technology to be specifically tailored to match the boilers furnace design features and minimize the adverse effects of the
low-NOx technology on boiler operating parameters (turndown, capacity, efficiency,
excess air, and CO emissions).
Impact of NOx Control Technologies on Boiler Performance
Selecting the best low-NOx control package should be made with total boiler performance in mind. This section describes the relationships of boiler operating parameters
with NOx control technologies.
Turndown
It is desirable that the boiler has a turndown ratio of at least 4:1. Even with this turndown ratio the boiler will cycle as frequently as 12 times per hour, or 288 times a day,
because the boiler must begin to cycle at inputs below 25% capacity. Frequent cycling
reduces the boiler efficiency and deteriorates the boiler components. Low-NOx control
can reduce the turndown ratio of the boiler.
Capacity
When selecting the best NOx control, boiler capacity and turndown ratio should be
considered together because some NOx control technologies require boiler derating in
order to achieve guaranteed NOx reductions. However, the boilers capacity requirement
is typically determined by the maximum heating load. Therefore, the boiler may be oversized for the typical load conditions. If the boiler is oversized, its ability to handle minimum loads without cycling is limited. Normally district heating plants have multiple
boilers in order to modulate capacity and thus have an advantage over single boiler plants
in this regard.
Efficiency
Low flame temperature decreases the radiant heat transfer and could lower boiler efficiency. The efficiency loss can be restored by increasing the convective heat transfer surfaces (with increases in boiler cost) or by utilizing FGR.
Excess Air
It is best to select a NOx control technology that has a minimal effect on excess air.
The boiler burner typically operates with 10%20% excess air (2%4% oxygen). NOx
controls that require higher excess air levels can result in increased stack losses and

3.66

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

reduced boiler efficiency. NOx controls that require reduced excess air levels can result in
an oxygen-deficient flame and increased levels of CO or unburned hydrocarbons.
Carbon Monoxide Emissions
Low CO emissions require high flame temperatures and intimate air/fuel mixing.
Some control technologies reduce NOx levels by lowering flame temperatures through
modification of air/fuel mixing patterns. The lower flame temperature and decreased mixing intensity can result in higher CO levels. Use of an FGR package can lower NOx levels
by reducing the flame temperature without increasing the CO levels. But, the level of CO
depends on the burner design. Not all FGR applications result in lower CO levels.

Sulfur Oxides (SOx)


Sulfur compounds (oxides) react with water vapor (in the flue gas and atmosphere) to
form sulfuric acid mist. The level of emitted SOx depends directly on the sulfur content of
the fuel. The SOx pollution has been controlled by either dispersion (by a tall stack) or
reduction. SOx reduction is achieved by switching to low sulfur fuel, desulfurizing the
fuel, or utilizing a flue gas desulfurization (FGD) system.

Carbon Monoxide (CO)


In the boiler the carbon in the fuel oxidizes through a series of reactions, forming carbon dioxide (CO2). However, 100% conversion of carbon to CO2 is rarely achieved and
some carbon only oxidizes to the intermediate step, carbon monoxide (CO). In modern
boilers, high CO emissions can result from incomplete combustion due to poor burner
design or firing conditions. Through proper burner maintenance and operation or utilizing
an oxygen control package, the formation of CO can be controlled at an acceptable level
(Cleaver Brooks 2010).

Particulate Matter (PM)


Particulate matter (PM) emissions from combustion sources consist of many different
types of nitrates, sulfates, carbons, oxides, and uncombusted elements in the fuel. Particulate pollutants can be corrosive, toxic to plants and animals, and harmful to humans. PM
emissions are classified into two categories: PM and PM10 (particulate matter with a
diameter less than 10 microns). The greatest concern is with PM10 because of its ability
to bypass the bodys natural filtering system.
PM emissions are dependent on the grade of fired fuel. PM emissions from natural
gas are significantly lower than those of oils. Distillate oils result in much lower particulate emissions than residual oils. When burning heavy oils, particulate levels mainly
depend on four fuel constituents: sulfur, ash, carbon residue, and asphalenes.
Methods of particulate control vary for different types and sizes of boilers. Except for
the smallest systems, most district heating applications use utility-scale boilers. For PM
control in district heating applications, electrostatic precipitators, scrubbers, and baghouses are commonly used. Particulate emissions can be reduced through proper burner
setup, adjustment, and maintenance, but not to the extent accomplished by switching to
lower-PM fuels.

Volatile Organic Compounds (VOCs)/Hydrocarbons (HCs)


Volatile organic compounds (VOCs) are compounds containing combinations of carbon,
hydrogen, and sometimes oxygen. VOCs vaporize easily once emitted into the air and are of
concern because of their role in ground-level ozone formation. VOCs are often referred to as
hydrocarbons (HCs) and are divided into two categoriesmethane and non-methane.

3.67

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Formation of VOCs in boilers results from poor or incomplete combustion due to


improper burner setup and adjustment. Properly maintaining the burner/boiler package
(keeping the air/fuel ratio at the manufacturers specified setting, maintaining proper air
and fuel pressures at the burner, and maintaining the atomizing air pressure on burners at
the correct levels) will help keep VOC emissions at a minimum. An improperly maintained
boiler/burner package can result in VOC levels more than 100 times the normal levels.

Calculation of Annual Emissions for District Heating Boilers


The EPA has established nationally uniform source-specific regulations through the
New Source Performance Standards (NSPS), to which all applicable sources must comply (EPA 2013a). The standards, which set minimum requirements for individual sources,
address approximately 65 categories of new or modified stationary sources, including
industrial-scale boilers. The NSPS for industrial-scale boilers regulate levels for NOx,
SOx, and PM. The regulated pollutants and requirements vary for different fuels and
boiler sizes. There are currently three categories for the NSPS:
Boilers with inputs greater than 250 MMBtu/h (263.7 GJ/h)
Boilers with inputs between 100 and 250 MMBtu/h (105.5 and 263.7 GJ/h)
Boilers with inputs between 10 and 100 MMBtu/h (10.55 and 105.5 GJ/h)
The small-boiler NSPS apply to all new, modified, and reconstructed boilers with
inputs between 10 and 100 MMBtu/h (10.55 and 105.5 GJ/h) where construction, modification, or reconstruction commenced after June 9, 1989. They set emission standards for
SOx and PM for boilers firing coal, distillate and residual oils, and wood, and they also dictate record-keeping requirements regarding fuel usage for all fuels, including natural gas.
Annual emissions are usually expressed as tons per year (tpy). The following equation can be used to determine potential annual emissions:
Emission Factor Boiler Input Annual Hours of Operation
= Total Annual Emissions
The rest of this section discusses how to calculate the potential annual NOx emissions
for industrial boilers as an example; the annual emissions for other pollutants are estimated in a similar way.
To determine the annual NOx emissions for a U.S. industrial district heating boiler,
three details must be known: the NOx emission factor for the boiler, the maximum rated
input for the boiler, and the maximum allowable hours of operation for the boiler. In
accordance with the NSPS (EPA 2013), the potential to emit is the theoretical maximum total emissions from the facility operating at full capacity 24 hours per day, 7 days
per week. The calculation of the theoretical maximum total annual NOx emissions for an
800 hp (597 kW) gas-fired boiler operating 24 h/day, 365 days/yr and having a NOx level
of 30 ppm is as follows (Cleaver Brooks 2010):
Emission Factor = 30 ppm/850 ppm
(850 ppm of NOx is equivalent to 1 lb/MMBtu of natural gas)
= 0.035 lb/MMBtu (30 ppm = 0.035 lb/MMBtu) (0.0151 kg/GJ)
Boiler Input = 33.5 MMBtu/h (35.33 GJ/h) (Based on 80% efficiency)
Annual Hours of Operation = 8760 h/yr (24 h/day 365 days/yr)
Annual NOx emissions for this specific boiler
= 0.035 33.5 8760
=10,271 lb/yr or 5.1 U.S. tpy (4.67 mtpy)

3.68

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Current Emission Standards


Air permitting of central plant boilers is based on the EPAs AP-42, Compilation of Air
Pollutant Emission Factors (1995). However, in spring of 2010 EPA proposed stringent
new standards, National Emission Standards for Hazardous Air Pollutants (NESHAPs)
governing the emission of toxic air pollutants from boilers and process heaters at industrial, commercial, and institutional facilities (EPA 2013b). The new standards are a set of
proposed changes to the EPAs Industrial, Commercial and Institutional Boiler Maximum
Achievable Control Technology (MACT) rule (Crilley 2010). Also known as Boiler
MACT, this rule is intended to set emissions limits for certain hazardous air pollutants
(HAPs) from major and area (i.e., minor) source boilers and process heaters, as
required under Section 112 of the Clean Air Act (EPA 2012). The HAPs addressed in the
Boiler MACT rule are mercury, particulate matter, hydrogen chloride, carbon monoxide,
and dioxins and furans. The basis for the emissions levels in the rule is the maximum
achievable control technology, defined by the EPA as the average emissions limitations
currently achieved by the best-performing 12% of sources in a source category (Crilley
2010). The full compliance of the new regulation is planned for 2013.
For the purposes of the Boiler MACT, a boiler is defined as an enclosed device
using controlled flame combustion and having the primary purpose of recovering thermal
energy in the form of steam or hot water. The EPA considers a major source of emissions to be a facility that annually emits at least 10 tons of any one toxic air pollutant or at
least 25 tons of all toxic air pollutants. Boiler MACT also sets limits for area, or minor,
sources, which are those boilers and process heaters that emit HAPs below the major
source thresholds (Crilley 2010). A summary of the new standards are presented in Tables
3.1 and 3.2.
Once a central plant has been defined as either a major or area emissions source, the
next step is to determine the compliance status of the affected units. This can be achieved
by doing fuel analyses and/or emissions testing. It is important to keep in mind that it is
necessary to measure the fuel variability, as trace element concentrations can vary significantly over time. A plant may want to consider establishing a long-term fuel sampling
program to both quantify fuel variability and ensure the accuracy of its emissions measurement.

Compliance Solutions
There are various options for achieving compliance with the Boiler MACT rule, each
complicated by the interconnectedness and inverse relationships that emissions have with
each other. These solutions will be specific to each particular facility, fuel, unit configuration and, potentially, each operation. By no means are any of the following strategies to be
considered a total solution, as the variables and HAP levels can change with the implementation of any one of them; fixing one HAP can positively or negatively affect another.
Mercury
If mercury is based on a pollutant (in pounds) and on fuel heat input (MMBtu [GJ])
then the solution will be highly fuel- and unit-configuration-specific. The mercury in particulate can be in elemental or oxidized form. For example, if the emissions are minimal
in elemental mercury and higher in particulate-bound or oxidized mercury, and the unit
has an existing baghouse, then there is the potential for the mercury concentration in the
flue gas to be below the new regulatory requirement. However, if the mercury is oxidized
and the unit has a higher unburned carbon level, then the mercury level may or may not be
in compliance. This means the solution for mercury will clearly be both fuel- and configuration-specific.

3.69

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Table 3.1 Emission Limits for Major Source Boilers and Process Heaters
(Major source facility emits 10 tons/yr of a single HAP or 25 tons/yr of multiple HAPs)
Dioxins/Furans
(Total Toxicity
Carbon
Mercury (Hg),
Equivalent),
Monoxide
lb/TBtu
nanograms per
(CO),
(kg/TJ)
dry standard
ppm @ 3% O2
cubic metre
(ng/dscm)

Existing/
New

Fuel

Firing
Method

Particulate
Matter (PM),
lb/MMBtu
(kg/GJ)

Hydrogen
Chloride
(HCL),
lb/MMBtu
(kg/GJ)

Existing

Coal

Stoker

0.02
(0.0465)

0.02
(0.0465)

3
(7)

50

0.003

Existing

Coal

Fluidized Bed

0.02
(0.0465)

0.02
(0.0465)

3
(7)

30

0.002

Existing

Coal

Pulverized

0.02
(0.0465)

0.02
(0.0465)

3
(7)

90

0.004

Existing

Biomass

Stoker

0.02
(0.0465)

0.006
(0.014)

0.9
(2.1)

560

0.004

Existing

Biomass

Fluidized Bed

0.02
(0.0465)

0.006
(0.014)

0.9
(2.1)

250

0.02

Existing

Biomass

Suspension
(Dutch Oven)

0.02
(0.0465)

0.006
(0.014)

0.9
(2.1)

1010

0.03

Existing

Biomass

Fuel Cells

0.02
(0.0465)

0.006
(0.014)

0.9
(2.1)

270

0.02

Existing

Liquid

0.004
(0.0093)

0.0009
(0.0021)

4
(9.3)

0.002

Existing

Gas

0.05
(0.116)

0.000003
(0.000007)

0.2
(0.465)

0.009

New

Coal

Stoker

0.001
(0.0023)

0.00006
(0.00014)

0.2
(0.465)

0.003

New

Coal

Fluidized Bed

0.001
(0.0023)

0.00006
(0.00014)

2
(4.65)

30

0.00003

New

Coal

Pulverized

0.001
(0.0023)

0.00006
(0.00014)

2
(4.65)

90

0.002

New

Biomass

Stoker

0.008
(0.0186)

0.004
(0.0093)

0.2
(0.465)

560

0.00005

New

Biomass

Fluidized Bed

0.008
(0.0186)

0.004
(0.0093)

0.2
(0.465)

40

0.007

New

Biomass

Suspension
(Dutch Oven)

0.008
(0.0186)

0.004
(0.0093)

0.2
(0.465)

1010

0.03

New

Biomass

Fuel Cells

0.008
(0.0186)

0.004
(0.0093)

0.2
(0.465)

270

0.0005

New

Liquid

0.002
(0.00465)

0.0004
(0.00093)

0.3
(0.7)

0.002

New

Gas

0.003
(0.007)

0.000003
(0.000007)

0.2
(0.465)

0.009

Source: NARA (2010a).

3.70

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Table 3.2 Emission Limits for Area Source Boilers


(Area source facility emits HAPs below major source threshold)
Particulate Matter (PM),
lb/MMBtu
(kg/GJ)

Mercury (Hg),
lb/TBtu
(kg/TJ)

Carbon Monoxide (CO),


ppm

3 (7)

310 (@ 7% O2)

Existing/New

Fuel

Existing

Coal

Existing

Biomass

160 (@ 7% O2)

Existing

Oil

1 (@ 1% O2)

New

Coal

0.03 (0.07)

New

Biomass

0.03 (0.07)

100 (@ 7% O2)

New

Oil

0.03 (0.07)

2 (@ 3% O2)

3 (7)

310 (@ 7% O2)

Source: NARA (2010b).

Ash disposal must also be considered, as carbon-based injection solutions can compromise fly ash quality and impact revenues. Options may include an oxidizer, activated
carbon, new noncarbon sorbents, or a combination of such solutions. Units with desulfurization systems or wet particulate scrubbers may require different reduction methods or
may be in compliance without the need for any additional HAP-reducing systems.
Hydrogen Chloride
A typical solution for reducing hydrogen chloride is to inject a sorbent, which may be
sodium sesquicarbonate, sodium bicarbonate, hydrated lime, or various other sorbents
that are either readily available or proprietary to specific companies. Unit configurations
with semidry wet desulfurization systems or wet particulate scrubbers may provide compliance without the need for sorbent injections or additional systems.
Particulate Matter (PM)
The solution for reducing PM is to upgrade or change out a plants particulate collection device. Particulate removal systems include cyclone separation, wet particulate
scrubbers, electrostatic precipitators, and baghouses. Typically, baghouses will ensure
PM compliance. All other devices will require evaluation and a determination on compliance ability.
Carbon Monoxide (CO)
Solutions for CO are a bit more challenging. Elevated levels indicate poor or incomplete combustion. To reduce this HAP, the solution will likely be to install a system overfire and/or to optimize combustion components (burners, grates, etc.). These components
will allow hydrocarbons and CO to be reduced through more efficient and thorough combustion. There are also advantages for adding these systems, such as increased unit efficiency, lower NOx emissions, and reduced total plant operational costs (fuel, chemistry,
maintenance/operation, etc.).
Dioxins/Furans
Dioxins and furans are formed post-combustion during the cooling of the flue gas.
The likely strategy for dioxin/furan capture and reduction is sorbent injection. One of the
sorbents known to reduce dioxins/furans is activated carbon. Since the original MACT
rule did not include this HAP, initial measuring of emissions values of dioxins/furans and
testing sorbent injection may be required to determine the appropriate solution.

3.71

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Other Measures
Various other HAP reduction options exist, with the most widely investigated being
fuel switching. To do a sound evaluation of whether a fuel switch makes sense, each facility will need to analyze its emissions inventory, operational expenses, maintenance costs,
cost to convert or replace existing equipment, and outlook for the future.
Compliance Solutions Summary
When determining compliance strategies, district heating plants and other facilities
have many variables to consider. There are several types of products and services that a
plant can use to ensure compliance, each with advantages and disadvantages.
It is essential for a central plant facility to develop a compliance plan for the Boiler
MACT that allows for total multi-pollutant reduction. Such a multi-pollutant strategy
would provide unit flexibility, which is key in the industry. Finding a solution with multipollutant control should also be a priority given that additional regulations are on the horizon and are likely to be more stringent.
Future regulations include Area Source Boiler MACT, Sulfur Particulate Matter, and
Ozone National Ambient Air Quality Standards. Furthermore, the replacement rules for
both the Clean Air Interstate Rule Plus (regulating NOx and SOx) and the Clean Air Mercury Rule originally planned for a final ruling in November 2011 were delayed until such
time as judicial review is no longer pending or until the EPA completes its reconsideration
of the rules, whichever is earlier. It is believed that the district heating segment may also
be required to meet SOx and NOx rules under upcoming regulations.

INSTRUMENTATION AND CONTROLS FOR DISTRICT HEATING PLANTS


General
This section addresses the criteria for the selection of instruments and controls to
meet the requirements of the central heating plant (DOD 2007).
Microprocessor-Based Controls
Microprocessor-based control systems can provide sequential logic control and modulating control in one control device. This capability makes available boiler control systems, which use both sequential logic and modulating control, more flexible and reliable
as well as more cost-effective. Processing units can be used as single-loop controllers, or
more powerful processing units can be applied to individual control subsystems, such as
combustion control of ash-handling control. The major considerations for microprocessor-based controls are the following.
Inputs from and outputs to field devices may be multiplexed. Data highways
connect all processing units to data storage and acquisition components, liquidcrystal displays (LCDs) and operator consoles, loggers, and printers, providing
communication among all components. Communications between modulating
control devices and sequential logic flow freely. Data acquisition and operator
interface for control may be accomplished using LCDs, keyboards, and printers
or through control station indicators and recorders mounted on an operator console. When LCDs, keyboards, and printers are used, redundant microprocessors
are sometimes used depending on unit size.
Program logic can be changed or expanded readily with limited hardware revisions. System selection can range from programmable controller systems to
fully distributed digital control systems. The criteria for selection of the proper
microprocessor-based system should include unit size, the amount and type of

3.72

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

modulating control and sequential logic required, operator interface requirements, and system security requirements.
Distributed digital control systems include process input/output sensors and
actuators that are connected to the termination units, which condition and multiplex the signals for communication to the microprocessor unit controllers where
the logic resides for control of the process variable. All microprocessor unit controllers communicate with each other and to a data highway that includes nodes
for the operator interface station, an engineering work station, and programmable logic controllers (PLCs) for balance of plant (BOP) controls.
The distributed control system logic for the controls should be located in the
vicinity of the process. This process controller should communicate with other
process controllers, the engineers workstation, the data acquisition system, and
the operator interface. The control system should be configured to allow the process controller to continue to function upon loss of communication with the
operator interface, data acquisition system, and other process controllers.
Control System Reliability
The methods used to ensure control system reliability are based on unit size, importance to plant operation, and the cost of control system failure versus the cost of backup
hardware.
Power sources. Power to the control system must have a backup supply. Microprocessor-based systems must have a backup power supply either from a separate AC source or an uninterruptible power supply. Loss of either the primary
power source or the backup power service must be detected and alarmed.
Control system safeguards. Microprocessor-based controls are highly reliable,
but safeguards must be provided to limit the effects of component failure. For
critical subsystems, consideration should be given to redundant microprocessors
with automatic switching of inputs and outputs from one microprocessor to
another. The data highway should be looped or redundant so that failure of a
segment of the data highway will not result in the loss of communication. Control elements should be designed to fail in a safe condition upon loss of the electric or pneumatic power to the actuators or loss of the input signal. The loss of
power at the component or subsystem levels must cause the associated auto/
manual stations to switch to the manual mode of operation. The control logic
should have continuous self-diagnostic capability and, upon detection of component failure, transfer to manual and indicate the cause of the failure. Microprocessors should contain nonvolatile memory that will not be erased on power
failure.
Control System Expansion
The control system architecture should allow expansion at all levels of the system.
Additional nodes can be added to the local area network (LAN) to allow additional processing units, engineering workstations (EWSs), and operator interface LCDs to be added
to the control system.
Data Link
The process signals should be connected to the termination units and through signal
conditioners to the microprocessor controllers. The control system data highway for
exchange of data between microprocessor-based controllers and between microprocessorbased controllers, data acquisition systems, the operator interface, and EWSs will be
redundant. The data highways should use coaxial, or twines, of fiber optic cabling. The

3.73

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

speed of data transmission continues to increase and thus future requirements should be
investigated prior to writing specifications.

Instrumentation
All instrument operations should have instrumentation calibration that is traceable to
the National Institute of Standards and Technology (NIST).
Transmitters, control drives, control valves, and piping instrumentation should be
provided to sense the process variables and allow the control system to position the valves
and dampers to control the process. All field devices shall be designed to operate in a
dust-laden atmosphere with temperature conditions varying from 20F to 160F (6.7C
to 71.1C), unless more stringent requirements are necessary due to the location of a particular sensor.
Field Transmitters
Electronic transmitters should produce a 4-20 mA DC signal that is linear with the
measured variable. Electronic transmitters should be the two-wire type except when
unavailable for a particular application. Encapsulated electronics are unacceptable in any
transmitter. Transmitters should be selected such that the output signal represents a calibrated scale range that is a standard scale range between 110% and 125% of the maximum value of the measured process variable. Transmitters should be designed for the
service required and will be supplied with mounting brackets. Purge meters and differential regulators should be used on transmitters for boiler gas service or coal-air mixture
service. A change in the load on a transmitter within the transmitter load limits will not
disturb the transmitted signal. The load limits of the transmitter should be a minimum of
600 ohms. Transmitters can be supplied with local indicators, either integral or field
mounted. Transmitters used for distributed control systems should be the smart types
that have duplex digital communication ability transparent to the analog signal. Smart
transmitters may be remotely calibrated via a handheld terminal. Data available at the
handheld terminal should include programmed instrument number, instrument identification or serial number, instrument location, date of last calibration, calibrated range, and
diagnostics. The types of transmitters and sensors to consider include the following.
Flow transmitters. Numerous types of flow transmitters are available. These
include differential pressure with square root extractor, turbine flowmeters,
nutating disk-type transmitters, ultrasonic flow transmitters, and magnetic flowmeters. The most common method for measuring flow is to measure differential
pressure across an orifice, flow nozzle, venturi, vortex shedding, pitot tube, or
piezometer ring. Square root extraction is necessary to linearize the output signal. Differential pressure measurement should be used for most steam plant flow
applications. Nutating disk with pulse to 4-20 mA transmitters are normally
used for fuel oil flow measurement. Flow transmitters should be accurate within
0.5% of span from 20% to 100% span with ambient temperature effect not to
exceed 1.0% per 100F (37.8C) variation. For large water pipe applications,
mag meters could be more applicable.
Level transmitters. Measuring elements for level transmitters will be diaphragm,
bellows, bourdon tube, strain gage transducer, caged float, or sealed pressure
capsule. Level transmitters should be accurate within 0.5% of span with ambient
temperature effect not to exceed 1.0% of span per 100F (37.8C) variation. The
output signal should be linear with the sensed level.
Pressure and differential pressure transmitters. Measuring elements for pressure
transmitters are normally diaphragm, bellows, bourdon tube, or strain gage

3.74

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

transducers. Pressure transmitters will be accurate within 0.5% of span with


ambient temperature effect not to exceed 1.0% of span per 100F (37.8C) variation. Measuring elements for differential pressure transmitters are normally
diaphragm, bellows, or sealed pressure capsule. Pressure differential transmitters should be accurate within 0.25% of span with ambient temperature effect
not to exceed 1.0% of span per 100F (37.8C) variation. Output signals for
pressure and differential pressure transmitters are normally linear with the
sensed pressure or differential pressure.
Temperature sensors. Several types of sensors can be used for temperature measurement. Thermocouples sense temperature by a thermoelectric circuit that is
created when two dissimilar metals are joined at one end. A wide variety of thermocouples are available for temperature sensing. Type J (iron-constantan) and
type K (chromel-alumel) are the most common types for boiler plant applications. Type J thermocouples can be used for temperatures from 32F to 1382F
(0C to 750C). Type K thermocouples can be used for temperatures from
328F to 2282F (200C to 1250C). The type of thermocouple to be used,
type J or type K, should be selected based upon the temperatures to be sensed.
All thermocouples in the boiler plant should be of the same type. Resistance
temperature detectors (RTDs) sense temperature based upon the relationship
between the resistivity of a metal and its temperature. The most common RTD
used in boiler plant applications is platinum RTD with a resistance of 100 ohms
at 32F (0C). Sealed-bulb and capillary sensors detect temperature by sensing
the change in volume due to changes in temperature of a fluid in a sealed system.
Temperature transmitters. Measuring elements for temperature transmitters
should be thermocouple, RTD, or sealed-bulb and capillary. Temperature transmitters should be accurate within 0.5% of span with ambient temperature effect
not to exceed 1.0% of span per 100F (55.6C) variation. Output signals for temperature transmitters are normally linear with the sensed temperature.
Oxygen analyzers. Oxygen analyzers should be direct-probe type utilizing an
in situ zirconium sensing element. The element will be inserted directly into the
gas stream and directly contacts the process gases. The sensing element should
be provided with a protective shield to prevent direct impingement of fly ash on
the sensing element. The analyzer should be equipped to allow daily automatic
calibration checks without removing the analyzer from the process. The cell
temperature in the analyzer should be maintained at the proper temperature by a
temperature controller. The analyzer should be certified for in stack analysis
technique in accordance with the FM Global Approval Guide (2013). The analyzer should be furnished with all accessories necessary for a complete installation.
Opacity monitors. Opacity monitors use the principle of transmissometry to
indicate level of particulate emissions. A beam of light is projected across the
flue gas stream and a detector registers variations in the light transmittance
caused by the particulate in the flue gas.
Flue gas monitors. Flue gas monitors should be provided for all items required
for reports to environmental authorities. Flue gas monitors are either in situ or
extractive. In situ monitors are attached directly to the stack or breeching and
access for maintenance should be provided. Extractive systems are wet, dry, or
diluted. Wet extractive system sample lines should be heated to avoid corrosion.
Dry systems use a cooler to remove water. Dilution systems use clean dry air to

3.75

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

dilute the sample, eliminating the need to heat the sample lines or dry the sample. Either in-situ or extractive flue gas monitors should be used and not a mixture of the two for the various gases to be analyzed. All analyzers should be
provided with self-calibration features and have contact outputs for control room
annunciation.
Control Drives
Control drives are used for positioning of control dampers, isolating dampers, and
other devices requiring mechanical activation. Control drives may be pneumatic or electric and are either open-shut type or modulating type depending on the application. Modulating drives may include position transmitters. Control drives should have adjustable
position limit switches wired to terminal blocks, hand wheels or levers for manual operation, hand locks (or be self locking), position indicators, and adjustable limit stops at
maximum and minimum positions. Drive arms and connecting linkages should be supplied with the damper drives. Control drives should have stroking times as required by the
service and by NFPA recommendations.
Electric control drives should consist of an electric motor, gear box, rigid support
stand, and wiring termination enclosure. Electric control drives should be weatherproof.
The gear box should be dust tight, weather tight, and totally enclosed. Electric control
drives should be self-locking on loss of control or drive power. Drives for outdoor installation should be designed to operate with ambient conditions of 20F (28.9C) and
40 mph (64.4 km/h) winds. Drives should have adjustable torque limit switches and position limit switches. Electric drives should be supplied with motor starters, position controllers, speed controllers, characterizable positioners, transformers, and other accessories
as required. The control drive should be sized to provide 150% of the torque required to
drive the controlled device.
Control Valves
The following should be considered for control valve selection.
Valve bodies. Control valve bodies are normally constructed in accordance with
the applicable American National Standards Institute (ANSI) codes. Control
valves should be globe type unless otherwise required for the particular process.
Butterfly valves may be used in low-pressure water systems. Globe valves
should have a single port designed to meet the design conditions. Restricted
ports should be used when necessary for stable regulation at all loads. Special
consideration should be given to valves that pass flashing condensate to ensure
adequate port and body flow area. The valve body size may be smaller than the
line size if the plug guide is sufficiently rugged to withstand the increased inlet
velocity, but valve body size should not be smaller than one half the line size.
End preparations should be suitable for the applicable piping system. Valves
should have ploytetrafluorethylene packing for temperatures not exceeding
450F (232.2C). Bonnet joints should be flanged and bolted type and designed
for easy disassembly and ensuring correct valve stem alignment. Valve trim
should be cage guided and removable through the top after bonnet removal. Seat
rings should be easily replaceable. Flow direction should be flow opening unless
otherwise required.
Valve operators. Control valve operators should be pneumatic diaphragm actuated type except where piston actuators are required. Valve operators should be
adequate to handle unbalanced forces that occur from flow conditions or maximum differential. Allowances for stem force based on seating surface should be

3.76

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

made to ensure tight seating. Diaphragms should be molded rubber and diaphragm housing should be pressed steel. Piston operators should use cast pistons
and cylinders with O-ring seals. Each valve operator should have an air supply
pressure filter regulator. Valve operators for modulating service valves in fastresponse control loops, such as flow control or pressure control, should have
electropneumatic valve positioners. Limit switches should be provided if needed
for remote indication or control logic.
Control valve sizing. Proper control valve sizing requires careful analysis of the
process and piping system in which each valve is to be used. It is necessary to
calculate the required valve flow coefficients based upon flow, valve inlet pressure, valve outlet pressure, and process fluid conditions. Calculations should be
based on recommendations of the International Society for Automation (ISA)
Control Valve Primer: A User Guide (2009). Valve flow coefficients should be
calculated at the maximum, intermediate, and minimum process flow conditions. The control valves should be selected such that the maximum flow coefficient occurs at a valve travel between 70% and 80%. The minimum flow
coefficient should occur at a valve travel between 10% and 20%. Control valves
should be selected with a flow characteristic that provides uniform control loop
stability over the range of process operating conditions. A quick opening flow
characteristic provides large changes in flow at small valve travels and should
primarily be used for on-off service applications. With a linear flow characteristic, the flow rate is directly proportional to valve travel. Valves with linear flow
characteristics will be used for liquid level control where the ratio of the maximum valve pressure differential to the minimum valve pressure differential is
less than five to one. Linear flow characteristic valves should also be used for
pressure control of compressible fluids and for flow control when the flow rate
varies but the valve pressure differential is constant. With an equal percentage
flow characteristic, equal increments of valve travel produce equal percentage
changes in the existing flow rate. Equal percentage flow characteristic valves
should be used for liquid level control when the ratio of the maximum valve
pressure differential to the minimum valve pressure differential is greater than or
equal to five to one. Equal percentage flow characteristic valves should also be
used for pressure control of liquids and for flow control when the valve pressure
differential varies but the flow rate is constant. Special inner valve trim characteristics are required on applications where flashing or cavitation exist in liquid
service and for noise control in steam or gas service.
Control valve stations. Control valve stations are used to install control valves in
piping systems and to provide a means of isolating and bypassing the control
valve for maintenance purposes. Control valve stations should conform to the
recommendations of ISAs Control Valve Primer (2009). Control valve stations
consist of a control valve, isolating valves, bypass valve, and bypass line. Since
control valves are normally smaller than the line size, reducers are required and
can be integral to the control valve on valves with butt weld end connections.
Isolation valves are required to isolate the control valve for repair, removal, or
calibration and will be installed on the inlet and outlet sides of the control valve.
Isolation valves should be gate valves or other non-throttling-type valves. A
bypass valve is necessary to provide a means of controlling the process when the
control valve is not operable. The bypass valve should be identical to the control
valve except it will be manually operated. Using an identical valve on the bypass
provides better control during manual operation since the valve should have the

3.77

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

proper flow coefficient and special valve trim. The bypass line that contains the
bypass valve must be smaller than the main line size. The bypass line may be the
same size as the bypass valve but in no case should the bypass line be smaller
than one half the main line size.
Piping Instrumentation
Piping instrumentation should include the following items.
Pressure switches. Pressure switches are used to monitor pressures for remote
indications, interlocking functions, and alarm conditions. Pressure switches may
have snap-acting switch contacts. Shutoff valves of the same pressure and temperature rating as the process piping should be provided on each switch for isolation purposes. Snubbers should be provided on switches when the pressure
connection is located within 15 pipe diameters of a pump or compressor discharge.
Pressure gages. Pressure gages are used to provide local and remote indication
of process pressures. Scale ranges should be selected such that the normal operating pressure is at approximately mid-scale. Shutoff valves of suitable rating
should be provided on each gage for isolation purposes. Snubbers should be provided on gages when the pressure connection is located within 15 pipe diameters
of a pump or compressor discharge. Siphons should be provided on pressure
gages for steam service. Pressure gages should be provided on the discharge of
all pumps and compressors, all boiler drums, all main process headers, and other
locations as required to monitor equipment and process operation.
Thermometers. Thermometers are used to provide local indication of process
temperatures. Thermometers are normally the bimetallic type for most applications. Scale ranges should be selected such that the normal operating temperature is at approximately mid-scale. Thermometers should be provided with
thermowells so the thermometer sensing element is not inserted directly into the
process. Thermowells should be designed to withstand the pressure, temperature, and fluid velocities of the process in which they are inserted. Thermowells
installed in piping should be long enough to extend to approximately the pipe
centerline. Thermowells will have extensions to clear insulation and lagging.
Thermocouples. Thermocouples are used to provide remote indication and control of process temperatures. Type J or type K thermocouples are normally suitable for steam plant applications. Thermocouples should be provided with
thermowells or protection tubes of suitable rating. Thermowell or protection
tube length should be sufficient to provide the necessary insertion length plus
the desired nipple length. Thermocouple assemblies should also include insulators and terminal head with cover.
Temperature switches. Temperature switches are used to monitor temperatures
for remote indications, interlocking functions, and alarm conditions. Temperature switches may have snap-acting switch contacts or mercury switch contacts
and may be bulb and capillary type or direct insertion type. Thermowells of suitable rating should be supplied so the sensing element is not inserted directly into
the process.
Pressure controllers. Pressure controllers will be pneumatic with a bourdon tube
or bellows sensing element. The sensing element should be suitable for the pressure and temperature of the process fluid to be controlled and will be an integral
part of the controller assembly. The sensing element should have adequate sensitivity and be able to withstand the maximum pressure under all conditions. Pres-

3.78

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

sure controllers should have adjustable proportional and reset control action,
control point adjustment, a calibrated pressure setting dial, an air supply filter
regulator, and gages that indicate air supply and controller output pressures.
Pressure controllers should be mounted on the operator of the valve to be regulated.

District Heating Plant Controls


Combustion Controls
The purpose of a combustion control system (CCS) is to modulate the quantity of fuel
and combustion air inputs to the boiler in response to a load index or demand (steam pressure or steam flow) and to maintain the proper fuel/air ratio for safe and efficient combustion for the boilers entire load range. Three types of CCSs are available: series, parallel,
and series-parallel.
Series control. A series control system uses variation in the steam header pressure (or any other master demand signal) from the setpoint to cause a change in
the combustion airflow which, in turn, results in a sequential change in fuel flow.
The use of series control is limited to boilers of less than 100,000 lb/h
(45,350 kg/h) that have a relatively constant steam load and a fuel with a constant Btu (kJ) value.
Parallel control. A parallel control system uses a variation from the setpoint of
the master demand signal (normally steam pressure) to simultaneously adjust
both the fuel flows and combustion airflows in parallel.
Series-parallel control. A series-parallel control system should be used to maintain the proper fuel/air ratio if the Btu (kJ) value of the fuel varies by 20% or
more, if the Btu (kJ) input rate of the fuel is not easily monitored, or if both of
these conditions are present.
System Categories
Combustion controls can be further divided into two categories within the basic
types: positioning control and metering control.
Positioning control. Positioning systems require that the final control elements
move to a preset position in response to steam pressure variations from a setpoint. Parallel positioning systems that use a mechanical jackshaft to simultaneously position fuel feed and airflow from a single actuator apply to package-type
gas/oil-fired boilers in the 20,000 to 70,000 lb/h (9070 to 31,745 kg/h) size. This
system allows the operator to load the boiler over its complete operating range.
The fuel valves and air damper are operated by the same drive through a
mechanical linkage. The gas and oil valves include cams which are adjusted at
start-up to maintain proper fuel air ratio over the operating range of the boiler.
Parallel positioning systems with fuel/air ratio control are suitable for use on
gas/oil- and stoker-fired units.
Metering control. Metering control systems regulate combustion based on
metered fuel and airflows. The master demand developed from steam pressure
error establishes the setpoints for fuel flows and combustion airflows at the controllers. The controllers drive the final control elements to establish proper fuel
and airflows, which are fed back to the controllers. Maximum and minimum signal selectors are used to prevent the fuel input from exceeding available combustion air on a boiler load increase and to prevent combustion air from decreasing
below fuel flow requirements on load reduction. This system is a cross-limiting

3.79

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

flow tie-back system with air leading fuel on load increase and fuel leading air
on load reduction.
Gas/Oil System Controls
Gas/oil system controls should include the following components.
Fuel flow control. Fuel flow in gas/oil-fired boilers is controlled by operation of
gas or oil control valves in the supply lines to the burners. The gas or oil control
valves are modulated to control fuel flow based on the demand signal generated
by the CCS. Gas flow to the burner is measured by taking the differential pressure across an orifice. Metering-type control systems use the fuel flow and unit
load in the CCS to properly modulate fuel flow in response to the system
demand.
Combustion air control. The combustion air is normally controlled at the forceddraft fan. Airflow for package boilers is normally controlled by outlet dampers
at the forced-draft fan. Other methods that are used include control of the
forced-draft fan inlet vanes or control of the forced-draft fan inlet damper. The
relationship between inlet vane or damper position and airflow will be determined for characterizing the final control element. When a metering-type control system is used, airflow is measured downstream of the forced-draft fan by
averaging pitot tubes.
Combustion airflow measurement. Accurate combustion airflow measurement is
also important in metering-type CCSs for gas/oil-fired boilers. The flow element
should be designed to provide a design differential pressure across the flow element of not less than 2 in. H2O (5.1 cm H2O) at full-load conditions. The flow
transmitter selected for combustion airflow should be a differential pressure
transmitter that is accurate in the range of differential pressure developed by the
flow element.
Oil atomization. The oil to the burner will be atomized utilizing steam or compressed air. A control valve installed in the atomizing steam or air line should be
controlled to maintain the atomizing medium pressure above the oil supply pressure to the burner.

Boiler Controls
Furnace Safety System
The main function of a furnace safety system is to prevent unsafe conditions to exist
in the boiler, including prevention of the formation of explosive mixtures of fuel and air in
any part of the boiler during all phases of operation. The system must be made to comply
with the appropriate NFPA regulations and the recommendations of the boiler manufacturer and include the following components.
Purging and safety. Regardless of the type of firing system, certain functions
must be included in the furnace safety system. These functions include a prefiring purge of the furnace, establishment of permissives for fuel firing, emergency
shutdown of the firing system when required, and a post-firing purge. Gas/oil
firing require additional functions such as establishment of permissives for firing the ignition system and continuous monitoring of firing conditions. The prefiring purge is required to ensure that all unburned fuel accumulated in the
furnace is completely removed; it is accomplished by passing a minimum of
25% to 30% airflow through the furnace for five minutes. The furnace safety
system should be designed to be fail safe (de-energize-to-trip).

3.80

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Flame detection and management. Gas/oil-fired systems require flame detection, which is the key to proper flame management. The basic requirements of
flame detectors are detection of the high-energy zone of a burner flame; ability
to distinguish between ignitor and main flames; and discrimination between the
source flame, adjacent flames, and background radiation.
Burner controls. Burner controls are the permissives, interlocks, and sequential
logic that are required for safe start-up and operation. Burner controls range
from manual to fully automatic. Regardless of the level of automation incorporated into the burner controls, the system logic must ensure that the operator
commands are performed in the correct sequence with intervention only when
required to prevent a hazardous condition. Gas/oil burner controls must provide
the proper sequential logic to completely supervise burner start-up and operation, including gas/oil fuel valves, air registers, ignitors, and flame detectors.
Feedwater Flow and Drum Level Control
Feedwater flow and drum level control can be two- or three-element controls,
described as follows.
Two-element control. Two-element feedwater control systems are characterized
by the use of steam flow as a feed-forward signal to reduce the effect of shrink
and swell of the boiler drum level during load changes. Without the steam flow
feed-forward signal, load changes will momentarily cause the drum level to
change in a direction opposite to the load change. The feed-forward signal provides the correct initial response of the feedwater valve.
Three-element control. Three-element feedwater control uses feedwater flow in
addition to steam flow to improve drum level control. In this system, feedwater
flow to the boiler is metered and the feedwater valve is positioned by summing
steam flow and drum level error through a controller. This system should be
used when multiple boilers are connected to a common feedwater supply system
since feedwater flow is a metered feedback signal and the control system
demands a feedwater flow.
Furnace Pressure Controls
Furnace pressure controls have the following features.
Single-element control. Furnace pressure controls are primarily single-element
type. The final control element is the induced-draft fan inlet damper, induceddraft fan inlet vanes, or adjustable-speed drive for the induced-draft fan. The
control loop also uses a feed-forward demand signal that is representative of
boiler airflow demand. This feed-forward signal may be fuel flow, boiler master,
or other demand index but will not be a measured airflow signal.
Furnace implosion protection. Boilers that have a large capacity and large draft
losses due to air quality control equipment may require induced-draft fans with a
head capacity large enough to exceed design pressure limits of the furnace and
ductwork. If this possibility exists, the furnace pressure control system must
include furnace implosion protection. The furnace implosion protection system
will comply with the guidelines established by NFPA 85G (1987). These guidelines include redundant furnace pressure transmitters and a transmitter monitoring system, fan limits or run-backs on large furnace draft error, feed-forward
action initiated by a main fuel trip, operating speed requirements for final control elements, and interlock systems.

3.81

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Steam Controls
Steam controls include pressure, flow, and temperature controls, described as follows.
Steam pressure control. Steam pressure is controlled by boiler firing rate. As
discussed in the section on combustion control, steam pressure is used to establish the master demand signal that controls fuel and combustion airflow.
Steam flow control. Steam flow is a function of boiler load demand. Steam flow
is also a function of fuel Btu (kJ) input and can be used to trim combustion airflow as discussed in the Combustion Controls section. Steam flow is also used to
calculate boiler load for use in oxygen trim controls and as a feed-forward signal
in feedwater controls.
Steam temperature control. Boilers that produce saturated steam do not require
steam temperature controls. Boilers that produce superheated steam require a
control loop to maintain superheater outlet temperature.
Blowdown Controls
Automatic blowdown systems continuously monitor the boiler steam drum water
conductivity and adjust the rate of blowdown to maintain the conductance of the boiler
water at the proper level. Control action can be two-position or modulating. The use of
automatic blowdown will be dependent on whether blowdown heat is to be recovered.
Sootblower Control
Sootblower control should be an operator-initiated automated sequence control. After
the start command, the system should step through the sequence for all sootblowers
including opening the valve for the sootblowing medium (steam or compressed air), timing the length of the blow, and closing the valve. The system should automatically move
to the next sootblower and continue the sequence until all sootblowers have been completed.

Non-Boiler Controls
Non-boiler controls include the following equipment.
Low-pressure steam controls
Turbine drives. The boiler feed-pump turbine drive is controlled by feedwater header pressure. The steam control valve on the turbine drive inlet is
controlled by a pressure transmitter on the feedwater header acting through
a controller.
Steam coil air heater. The steam coil air heater controls are based on maintaining the flue gas leaving the air heater above the acid dew-point temperature. This is accomplished by using an average cold-end temperature
control system. Air heater average inlet air temperature and average gas
outlet temperature are calculated. These two signals are averaged to arrive
at the average cold-end temperature, which is used to control the steam coil
control valve. Also, the control system should include an interlock that
opens the steam coil control valve 100% when the ambient air is below a
set temperature, usually 35F (1.7C).
Deaerator. Deaerator steam controls are a pressure control system to maintain deaerator pressure. A single element loop with feedback is adequate
for controlling deaerator pressure.
Deaerator level control
Two-element control. A two-element deaerator level control system uses
feedwater flow as a feedwater signal to make the system responsive to load

3.82

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

changes. A two-element system for deaerator level control can be used for
most unit installations that operate under steady load conditions.
Three-element control. A three-element deaerator level control system uses
a metered condensate flow feedback signal in a cascaded control loop. This
system should maintain the deaerator level on units that operate under
swinging load conditions.
Pump recirculation control
Pump recirculation controls are necessary to maintain the minimum flow
through a pump when required by the manufacturer. A breakdown orifice
plate sized to pass the required minimum flow can be installed in a line
from the pump discharge to the pump suction source. Since this system is a
constant recirculation type, it is a source of lost pump power. The lost
power can be eliminated by using automatic pump recirculation controls.
Automatic pump recirculation controls require pump flow to be metered
and an automatic valve to open when pump flow is at or below the minimum flow requirement.

Control Panels
Control Room
A control room isolated from the plant environment complete with heating and air
conditioning should be provided for all boiler plants. The boiler panels and auxiliaries
may be located at the boiler front for package boilers up to 70,000 lb/h (31,745 kg/h). A
data historian computer should be located in the control room. The control room should
be located at a central location in the plant to allow operating personnel good access to
the boilers and the auxiliary equipment. The control room should be large enough for the
operator interface for the boiler and auxiliaries and also allow room for a desk to be used
by operating personnel.
Operator Interface
The operator interface to the boiler and auxiliaries may be via computer workstations
with LCD flat-screen monitors and printers housed in a control console or operator stations, recorders, indicators, annunciators, and start/stop controls mounted on a control
panel. The specific characteristics of flue operator interface systems are as follows.
Distributed control system. Operator interface via LCD flat-screen monitors and
printers are normally used on larger units and are part of the distributed control
system. This system should always use redundant microprocessors, object recognize tests (ORTs), and printers. The system will automatically switch to the
backup system and annunciate failure of a component. The system should be
used to perform combustion control, data acquisition and trending, boiler efficiency calculations, graphic displays, boiler control motor start/stop, and ash
system controls. An auxiliary panel should also be required to mount critical
controls and monitoring equipment. Following are elements of a distributed control system.
I/O racks. The system should include remote-mounted input/output (I/O)
racks with redundant microprocessors for control. The information at the
I/O rack should be multiplexed to allow communication with other I/O
racks and the central control console. Redundant communication links
should be provided to allow communication when one link is lost. All field
wiring entering or leaving the I/O racks is to be connected to terminal
blocks with spare terminals provided. The equipment in the I/O racks

3.83

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

should be designed for installation in a dusty atmosphere with a maximum


ambient temperature of 120F (48.9C).
Operator interface. The control console should include the appropriate
number of LCD flat-screen monitors and printers required by the size and
complexity of the system. A minimum of two LCD flat-screen monitors
and two printers should be installed. The LCD flat-screen monitors and
keyboard or other means of operation should be mounted in a console that
allows the operators to access and operate the controls while sitting at a
chair in front of the LCD flat-screen monitors.
Auxiliary panel. Auxiliary panel construction must conform to the requirements of the National Electrical Code (NFPA 2011) and National Electrical Manufacturers Association (NEMA) standards. It will be constructed of
steel plate with adequate internal reinforcement to maintain flat surfaces
and to provide rigid support for the instrumentation to be installed. The
panel interior should have adequate bracing and brackets for mounting of
equipment to be installed within the panel. Electrical outlets should be provided in the panel. No pressure piping of process fluids is to be run in control panels. All field wiring entering or leaving control panels is to be
connected to terminal blocks with spare terminals provided. The items to
be mounted in the auxiliary panel should include a hardwired main fuel trip
push button, fan trips, drum level indication, soot blower controls, and
annunciation of critical items. The annunciator should include the following items: main fuel trip, drum level high-low, furnace pressure high, boiler
feedwater pressure low, and control system power failure.
Panel-mounted control system. The control and auxiliaries panel, where used,
should include operator stations, recorders, indicators, equipment start-stop controls, and annunciation. Control panel construction must conform to the requirements of the National Electrical Code (NFPA 2011) and NEMA standards.
Panels should be constructed of steel plate with adequate internal reinforcement
to maintain flat surfaces and to provide rigid support for the instrumentation to
be installed. The panel interior should have adequate bracing and brackets for
mounting of equipment to be installed within the panel. A walk-in door for
access to the panel interior should be provided on at least one end of the panel
or, where possible, on both ends of the panel. Electrical outlets will be provided
in the panel. No pressure piping of process fluids is to be run in control panels.
All field wiring entering or leaving control panels is to be connected to terminal
blocks with spare terminals provided.
Instrumentation Requirements
The boiler and auxiliary panel or control console should provide operator interface
required to properly control and monitor the operation of the boilers and auxiliary equipment in the steam plant. This should include operator interface to stations, records or
trending, indication, equipment start/stop controls, and annunciation.
Operator stations. Operator stations on control consoles should be accessed
through a keyboard and mouse or through an individual touch-screen operator
station on control panels. Hand automatic operator stations should provide
bumpless transfer from hand to automatic and automatic to hand without manual
balancing for transfer and have anti-reset windup characteristics. Operator stations with setpoints should indicate setpoints in engineering units. Operator sta-

3.84

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

tions with ratio or bias are to indicate the magnitude of the ratio or bias at all
times. Operator stations are to indicate the measured variable continuously in
engineering units and should indicate station output continuously in percent.
The indications on a station should be consistent with all other stations such that
all final control elements move closed to open from zero to 100%. For hardwired operator stations, the position of final control elements should not change
when an operator station is disconnected from or reconnected to its plug-in
cable. Changes in ratio or bias settings should not cause a process upset.
Records. Records should be stored on the historian server hard drive. When a
control console is used, the operator will have access through the human
machine interface (HMI) and supervisory control and data acquisition (SCADA)
workstation to display trends for parameters for which records are kept.
Indicators. When a control console is used, the parameter should be displayed
on the LCD flat-screen monitor. The display may be digital or graphic and
should have scale markings consistent with the measured variable and associated field transmitter. When indicators are located on a panel, the indicators may
be digital indicators or analog indicators. Digital indicators should have lightemitting diode (LED) uniplaner numerals, zero instrument zero drift with time,
and 0.1% full-scale or less span drift per year. Analog indicators should have
vertical edgewise scales, plus or minus 2% full-scale accuracy, and scale markings consistent with the measured variable and the associated field transmitter.
Integrators should include the signal converters necessary to provide scaled integrated readings. Integrators should have at least a six-digit readout.
Equipment start/stop controls. Equipment start/stop controls should be provided
for all major equipment. Start/stop controls on a control console should be performed utilizing the ORT. Indication of motor operation should be indicated on
the LCD flat-screen monitor. Start/stop controls should be indicating control
switches or indicating push buttons when boiler and auxiliary panels are used
for the operator interface.
Annunciator. Annunciators should be provided on the boiler and auxiliary panel
for visual and audible indication of alarm conditions. Annunciator windows
should have alarm legends etched on the windows and be backlighted in alarm
or test state. Each window should have at least two parallel connected bulbs and
front access for ease of bulb replacement. All annunciator circuits should be
solid state and compatible with microprocessor-based controls. The annunciator
should have an adjustable tone and volume horn. Split windows should be
avoided unless conservation of panel space is critical. The annunciator system
should have one of the alarm sequences as specified in ISAs Control Valve
Primer (2009). Test and acknowledge push buttons should be provided on the
panel. All alarms, except critical alarms, should be displayed on the ORT and
printed on the printer when LCDs are used as the operator interface.

Energy Management and Control Systems


One advantage of central heating plants is easier implementation of building automation because the major and ancillary equipment are consolidated in one location. Computerized automatic controls can significantly affect system performance. A facility
management system to monitor system points and overall system performance should be
considered for any large district heating system. This allows a single operator to monitor
performance at many points in the plant.

3.85

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Software to consider when designing, managing, and improving central plant performance should include the following:
Automatic controls that can interface with other control software (e.g., equipment manufacturers unit-mounted controls)
Energy management system (EMS) control
Hydraulic modeling as well as metering and monitoring of distribution systems
Integration of VFDs on equipment to improve system control and to control
energy to consume only the energy required to meet the design parameter (e.g.,
temperature, flow, pressure)
Computer-aided facility management (CAFM) for integrating other software
(e.g., record drawings, O&M manuals, asset database)
Computerized maintenance management system (CMMS)
Automation from other trades (e.g., fire alarm, life safety, medical gases, etc.)
Regulatory functions (e.g., from federal, state, and local agencies, etc.)
Automatic controls for central plants may include standard equipment manufacturers
control logic along with optional, enhanced energy-efficiency control logic. These specialized control systems can be based on different architectures such as distributed controls, programmable logic controllers (PLCs), or microprocessor-based systems. Beyond
standard control technology, the following control points and strategies may be needed
for primary equipment, ancillary equipment, and the overall system:
Discharge temperature and/or pressure
Return distribution medium temperature
Head pressure for distribution medium
Stack temperature
CO and/or CO2 level
Differential pressures
Flow rate of distribution medium
Peak and hourly heat energy output
Flow rate of water and fuel(s)
Reset control of temperature and/or pressure
Night setback
Variable flow through equipment and/or system control
VFD control
Thermal storage control
Heat recovery cycle
Computerized energy management and control systems provide an excellent means
of reducing utility costs associated with operation of central energy plants. These systems
can incorporate advanced control strategies that respond to changing weather, user conditions, and utility rates to minimize operating costs. The central plants are typically controlled using a two-level control structure. Lower-level local-loop control of a single
setpoint is provided by an actuator. For example, the pressure differential control valve
located at the furthest customer is controlled by adjusting the opening of a valve that provides hot water to the building. The upper control level, supervisory control, specifies setpoints and other time-dependent modes of operation.
The performance of central heating plants can be improved through better local-loop
and supervisory control. Proper tuning of local-loop controllers can enhance comfort,
reduce energy use, and increase component life. Setpoints and operating modes for plant
equipment can be adjusted by the supervisor to maximize overall operating efficiency.

3.86

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

In small plants, fixed-speed pumps are used with their control dedicated to boiler control. Dedicated control means that each pump is cycled on and off with the boiler that it
serves. Systems signal to the VFD to reduce speed. As system demand requires increased
flow, the control valves modulate open, reducing system differential with fixed-speed
pumps, and two-way heating valves often incorporate a water bypass valve to maintain
relatively constant flow rates and reduce system pressure drop and pumping costs at low
loads. The valve is typically controlled to maintain a fixed pressure difference between
the main supply and return lines. This setpoint is termed the hot-water loop differential
pressure. Sometimes, primary systems use one or more variable-speed pumps to further
reduce pumping costs at low loads. In such cases, water bypass is not used and pumps are
controlled directly to maintain a water loop differential pressure setpoint.
Multiple boilers, typically arranged in parallel with dedicated pumps, provide the primary source of heating for the system. Individual feedback controllers adjust the capacity
of each boiler to maintain a specified supply water temperature or, for steam boilers,
steam header pressure. Additional control variables include the number of boilers operating and the relative loading for each. For a given total heating requirement, individual
boiler loads can be controlled by using different water supply setpoints for constant individual flow or by adjusting individual flows for identical setpoints.
Primary/secondary hot-water systems are designed specifically for variable-speed
pumping. In the primary loop, fixed-speed pumps provide a relatively constant flow of
water to the boilers. This design ensures good boiler performance and reduces the risk of
water vapor condensation on the boiler surfaces. The secondary loop incorporates one or
more variable-speed pumps that are controlled to maintain water loop differential pressure setpoint. The primary and secondary loops may be separated by a heat exchanger.
However, it is more common to use direct coupling with a common pipe.
The water supply temperature is controlled by setpoints depending on the outdoor
temperature. Night setback is often used in winter to lower the building temperature during unoccupied times and reduce the heating requirements. Setup and setbacks can also
be used during occupied periods to temporarily curtail energy consumption.

Control Variables
Control of a hot-water heating system is as follows: as the heating load increases, the
user controller responds to lower temperatures by opening a control valve and increasing
the flow of hot water through the building. Increasing water flow reduces the temperature
of the water returned to the boiler. With lower return water temperature, the supply water
temperature drops, which causes the feedback controller to increase the boiler firing rate
to maintain the desired supply water temperature (ASHRAE 2011, Chapter 42).
An increase in the user heating loads results in an increase in water flow rate, which is
ultimately propagated through the district system. For fixed-speed hot-water pumps, the
differential pressure controller closes the hot-water bypass valve and keeps the overall
flow relatively constant. For variable-speed pumping, the differential pressure controller
increases pump speed. For a hot-water system, the return water temperature and/or flow
rate to the boilers decreases, leading to a decrease in the hot-water supply temperature.
The boiler controller responds by increasing the boiler heating capacity to maintain the
hot-water supply setpoint (and match the user heating loads).
For any of these scenarios, several local-loop controllers respond to load changes to
maintain specified setpoints. A supervisory controller establishes modes of operation and
chooses (or resets) values that cause the feedback controller to increase the water flow.
This increases the hot-water flow and heat transfer load.

3.87

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Typical input and output stream variables for thermal systems are temperature and
mass flow rate. Uncontrolled variables are measurable quantities that may not be controlled but affect component outputs and/or costs, such as ambient temperature.
Both equality and inequality constraints arise in the optimization of hot-water systems. One example of an equality constraint that arises when two or more boilers are in
operation and the sum of their loads must equal the total load. The simplest type of
inequality constraint is a bound on a control variable. For example, lower and upper limits
are necessary for the hot-water set temperature.

Controls for Boilers


Boiler heating plants can be operated automatically by EMSs to reduce utility costs
associated with maintaining proper environmental conditions. Boiler efficiency depends
on many factors, such as combustion airflow rate, load factor, and water temperature in
hot-water boilers (or pressure for steam boilers). Opportunities for energy and cost reduction in boiler plants include excess air control, sequencing and loading of multiple boilers, and resetting the hot-water supply temperature setpoint (for hot-water boilers) or the
steam pressure setpoint (for steam boilers).
Excess Air in the Combustion Process
To determine the minimum excess air for a particular boiler, flue gas combustible
content as a function of excess oxygen should be monitored. For a gas-fueled boiler, carbon monoxide should be monitored; for liquid or solid fuel, monitor the smoke spot number. Different firing rates should be considered because the excess air minimum varies
with the firing rate (percent load). For burners and firing rates with a steep combustible
content curve, small changes in the amount of excess oxygen may cause unstable operation. The optimal control setpoint for excess air should generally be 0.5% to 1% above
minimum to allow for slight variations in fuel composition, intake air temperature and
humidity, barometric pressure, and control system characteristics.
Carbon Monoxide Limits
Carbon monoxide upper control limits vary with the boiler fuel used. The CO limit for
gas-fired boilers may typically be set at 400, 200, or 100 ppm. For No. 2 fuel oil, the maximum smoke spot number is typically 1; for No. 6 fuel oil, the smoke spot number is 4.
However, for any fuel used, local environmental regulations may require lower limits.
Oxygen Trim Control
An oxygen trim control system adjusts the airflow rate using an electromechanical
actuator mounted on the boilers forced-draft fan damper linkage and measures excess
oxygen using a zirconium oxide sensor mounted in the boiler stack. The oxygen sensor
signal is compared with a setpoint value obtained from the boilers excess air setpoint
curve for the given firing rate. The oxygen trim controller adjusts (trims) the damper setting to regulate the oxygen level in the boiler stack at this setpoint. In the event of an electronic failure, the boiler defaults to the air setting determined by the mechanical linkages.
Carbon Monoxide Trim Control
In CO trim systems, the amount of unburned fuel (in the form of CO) in the flue gas
is measured directly by a CO sensor and the air/fuel ratio control is set to actual combustion conditions rather than preset oxygen levels. Thus, the system continuously controls
for minimum excess air. Carbon monoxide trim systems are also independent of fuel type
and are virtually unaffected by combustion air temperature, humidity, and barometric
pressure conditions. However, they cost more than oxygen trim systems because of the
expense of the CO sensor. Also, the CO level in the boiler stack is not always a measure

3.88

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

of excess air. A dirty burner, poor atomization, flame chilling, flame impingement on the
boiler tubes, and poor fuel mixing can also raise the CO level in the boiler stack.
Sequencing and Loading of Multiple Boilers
Generally, boilers operate most efficiently at a 65% to 85% full-load rating. Boiler
efficiencies fall off at higher and lower load points, with the decrease most pronounced at
low load conditions. Boiler efficiency can be calculated by means of stack temperature
and percent oxygen (or percent excess air) in the boiler stack for a given fuel type. Partload curves of boiler efficiency versus hot-water or steam load should be developed for
each boiler. These curves should be dynamically updated at discrete load levels based on
the hot-water or steam plant characteristics to allow the control strategy to continuously
predict the input fuel requirement for any given heat load. When the hot-water temperature or steam pressure drops below setpoint for the predetermined time interval (e.g., 5
min), the most efficient combination of boilers must be selected and turned on to meet the
load. The least efficient boiler should be shut down and banked in hot standby if its capacity drops below the spare capacity of the current number of boilers operating (or, for primary/secondary hot-water systems, if the flow rate of the associated primary hot-water
pump is less than the difference between primary and secondary hot-water flow rates for a
predetermined time interval [e.g., 5 min]). The spare capacity of the current online boilers
is equal to their full-load capacity minus the current hot-water load.
Resetting Supply Water Temperature and Pressure
Standby losses are reduced and overall efficiencies are enhanced by operating hotwater boilers at the lowest acceptable temperature. Energy savings are therefore possible
if the supply water temperature is maintained at the minimum level required to satisfy the
largest heating load. However, to minimize condensation of flue gases and consequent
boiler damage from acid, water temperature should not be reset below that recommended
by the boiler manufacturer.
Similarly, energy can be saved in steam heating systems by maintaining supply pressure
at the minimum level required to satisfy the critical consumer. In practice, reset control is
only possible if boiler controls interface with the energy management and control system.
Operating Constraints
There are practical limitations on the extent of automatic operation if damage to the
boiler is to be prevented. Control strategies to reduce boiler energy consumption can also
conflict with recommended boiler operating practice. For example, in addition to the flue
gas condensation concerns mentioned previously, rapid changes in boiler jacket temperature (thermal shock) brought about by abrupt changes in boiler water temperature or flow,
firing rate, or air temperature entering the boiler should be avoided. The repeated occurrence of such transient conditions may weaken the metal and lead to cracking and/or
loose tubes. It is therefore important to follow all of the recommendations of the American Boiler Manufacturers Association (ABMA).

Boiler Supervisory Control Strategies and Optimization


Load Conditions
The specifics of the strategy for bringing boilers online depend on the type of boiler.
Hot-water boilers have dedicated or nondedicated hot-water pumps; steam boilers do not
have hot-water pumps but rely on differences in steam pressure between the boiler steam
header discharge and the point of use to distribute steam throughout the system.

3.89

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Hot-Water Boilers with Dedicated Pumps


Another boiler should be brought online when operating boilers reach capacity,
because the efficiency of the boiler should include the power to drive its associated hotwater pump. This can be determined when the hot-water temperature drops below its setpoint for a predetermined time interval (e.g., 5 min). Hot-water boilers with dedicated
pumps can be brought online and offline with the following logic:
Continuously calculate the load ratio of each boiler or boiler combination.
Every sampling interval (e.g., 60 s), calculate the predicted input fuel requirement for each boiler combination and the efficiency of each boiler combination.
Continuously evaluate time-averaged values of the hot-water supply temperature
over a fixed time interval (e.g., 5 min).
If the hot-water supply temperature drops below its setpoint for a predetermined
time interval (e.g., 5 min), then from boiler part-load performance curves select
the boiler combination with a load ratio between 0.5 and 1.0 and with the least
input fuel requirement to meet the load and turn this combination of boilers on.
Note that this strategy greatly reduces the possibility of short-cycling boilers
because the new combination of boilers to be started likely includes boilers
already operating (i.e., only one additional boiler is likely to be added).
If the capacity of the least efficient online boiler drops below the spare capacity
of the current number of boilers operating (or for a primary/secondary hot-water
system, if the flow rate of the associated primary hot-water pump is less than the
difference between primary and secondary hot-water flow rates) for a predetermined time interval (e.g., 5 min), then shut down and bank this boiler in hot
standby.
Hot-Water Boilers with Nondedicated Pumps or Steam Boilers
For hot-water systems without dedicated hot-water pumps or for steam systems, the
optimal load conditions for bringing boilers online or offline do not generally occur at the
full capacity of the online boilers. For these systems, a new boiler combination should be
brought online whenever the hot-water supply temperature or steam pressure falls below
setpoint for a predetermined time interval (e.g., 5 min) and the part-load efficiency curves
of the boiler combination predict that the new combination of boilers can meet the
required load using significantly less (e.g., 5%) input fuel.
It is generally more economical to run fewer boilers at a high rating. However, the
integrity of the steam or hot-water supply must be maintained in the event of a forced outage of one of the operating boilers or if the facility experiences highly diverse load swings
throughout the heating season. Both conditions can often be satisfied by maintaining a
boiler in standby or live bank mode. For example, in this mode, a steam boiler is isolated from the steam system at no load but is kept at system operating pressure by periodic firing of either the igniters or a main burner to counteract ambient heat losses.

Supply Water and Supply Pressure Reset for Boilers


Simple control strategies can be used to generate a suboptimal hot-water temperature
(for hot-water boilers) or steam pressure (for steam boilers). An energy management and
control system must be interfaced to the boiler controls and be capable of monitoring the
position of the valve controlling the flow of hot water to the heating coils or steam pressure at the most critical zone. For a hot-water system, do the following:
Continuously monitor the hot-water valve position. If none of the hot-water
valves are greater than 95% open, lower boiler hot-water supply temperature by

3.90

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

a small increment (e.g., 1F [0.6C]) each reset time interval (a predetermined


interval established by system thermal lag characteristics, e.g., 15 to 20 min).
Once one hot-water valve opens beyond 95%, stop downward resets of the
boiler hot-water temperature setpoint.
If two or more hot-water valves open beyond 95%, raise the boiler hot-water
temperature setpoint by a small increment (e.g., 1F) each reset interval.
For a steam system, the steam header pressure should be lowered to a value that just
satisfies the highest pressure demand (ASHRAE, 2011, Chapter 42).

Supervisory Control and Data Acquisition (SCADA) System


This section provides an example of a supervisory control and acquisition system
developed and installed at a central energy plant. The plant is equipped with five gas/#2
fuel oil-fired boilers each generating 70,000 lb/h (31,745 kg/h) of steam and a 3.5 MW
combined heat and power (CHP) unit with a heat recovery generator.
Figure 3.29 shows the architectural arrangement of the system. Descriptions of the
system components follow.
The supervisory control and data acquisition (SCADA) server communicates
with the process control hardware via the plant Ethernet network. Server-1 contains the tag database, processes graphics files, and manages alarms and events,
which enables the operators to monitor and control the entire plant. For reliability, a second SCADA system, Server-2, is installed that mirrors the primary
server. Server-1 and Server-2 are configured to operate in a redundant fashion so
that if the primary fails the backup takes over in a bumpless transition. The other
advantage of having redundant servers is that maintenance can be performed on
one server at a time without ever losing the ability to monitor and control the
plant. Server-1 is also used as an engineering workstation (EWS). The EWS has
SCADA and programmable logic controller (PLC) configuration software
loaded on it whereby plant technicians can make programming changes and perform system diagnoses.
SCADA Server-2 is the backup for SCADA Server-1 and it is also used as a client machine from which plant operators can monitor and control the equipment.
The data historian software is used to archive plant operating parameters, fuel
data, emissions data, production information, and energy data. The historian
software is specifically designed to compress the data to minimize the amount of
hard-drive space used. Current data historian software has the ability to store
decades of information. Reporting is normally performed using a Microsoft
Excel add-in that is supplied with the historian software. Reports and trends
can be generated at any one of the computers on the system via the plant network.
The Web server is used to display plant information on the Internet. This is a
useful tool for the operations staff to monitor the plant remotely. Also, customers are able to see the chilled-water supply and return temperatures, chilledwater differential pressure and steam delivery pressure in real time, thus reducing phone calls to the plant when the customer suspects problems with the district energy system.
The dual-screen SCADA client provides another point of access into the plant
system. Normally the large screen displays an overview of all the major plant
parameters such as power demand, fuel usage, statuses of major equipment, dis-

3.91

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 3.29 Architectural arrangement of a SCADA system for a central steam plant.
Courtesy of Array Systems, LLC

tribution pressures, flows, and temperatures. The small display screen is used for
day-to-day plant operation and monitoring.
The printer is on the plant network so graphics and reports can be printed from
any one of the computers. Alarm reports can also be printed from the system. In
years past, most plants used a separate dot matrix alarm printer with continuous
fan-fold paper. All alarms and events would be sent to this dedicated printer and
the printouts would be stored in boxes and file cabinets at the plant. Of course
the paper tended to break and jam periodically, resulting in lost alarm and event
logs. Nowadays all alarms and events are stored electronically in the SCADA
server and/or plant historian and are retrieved as needed.
Plant equipment that has communication capability, such as variable-speed
drives (VSDs), package chillers, water treatment systems, metering devices,
etc., are typically integrated with the plant control system. The connection to
this equipment is normally through a serial interface or more commonly now
through an Ethernet connection.

3.92

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

The balance of plant (BOP) PLC system is used to monitor and operate all control system components outside of the major vendor equipment. Normally, for
ease of integration, the PLC used for this application is of the same manufacturer as the type supplied with the combustion turbine generator (CTG) and the
boilers.
The BOP remote input/output (I/O) cabinet is normally located at a considerable
distance from the main cabinet. This is done to reduce the amount of field wiring
for the instrumentation and controls.
The CTG package is furnished with its own preprogrammed PLC from the manufacturer. In addition a local touch-screen human machine interface (HMI) is
provided on the front of the CTG package so operators can monitor and control
this equipment locally.
The heat recovery steam generator (HRSG) control system consists of two
PLCs. One is used for the burner management system (BMS) and the other is
used for the combustion control system (CCS). The BMS provides the necessary
protection for the boiler and duct burner such as low-water trips, high-pressure
trips, and flame safeguard system trips. In accordance with NFPA 85G (1987),
the BMS must be a separate and independent control system from the CCS. The
CCS is used to modulate the duct burner fuel valve, to provide drum level control and emissions control, and for monitoring real-time parameters. A local
touch-screen HMI allows operators to monitor and control the HRSG from the
plant floor.
The auxiliary boiler control system consists of two PLCs. One is used for the
BMS and the other is used for the CCS. The BMS provides the necessary protection for the boiler and burner such as low-water trips, high-pressure trips, and
flame safeguard system trips. In accordance with NFPA 85G (1987), the BMS
must be a separate and independent control system from the CCS. The CCS is
used to control the fuel system, combustion air, flue gas recirculating system,
and drum level and for monitoring real-time parameters. A local touch-screen
HMI allows operators to monitor and control the HRSG from the plant floor.
The steam turbine-driven chiller package is furnished with its own preprogrammed PLC from the manufacturer. Typically this needs to be purchased
as a special order from the chiller manufacturer. The advantages of using a PLCbased system in lieu of a standard chiller control panel are: PLCs are more rugged and robust, custom site-specific programming is much easier to do, and integration with the plant system is better because the chiller PLC will be from the
same manufacturer as the other control system hardware.
The plant Ethernet network backbone is fiber optic in a ring configuration. If a
failure were to occur at any one point in the ring, communication is automatically rerouted through the managed network switches so the system remains online.

REFERENCES
AREMA. 2012. Manual for Railway Engineering. Lanham, MD: American Railway
Engineering and Maintenance-of-Way Association.
ASHRAE. 2008. ASHRAE HandbookHVAC Systems and Equipment. Atlanta:
ASHRAE.
ASHRAE. 2011. ASHRAE HandbookHVAC Applications. Atlanta: ASHRAE.
ASME. 2013. ASME Boiler and Pressure Vessel Code. New York: ASME.
ASME. 2012. ASME B-31.1-2012, Power Piping. New York: ASME.

3.93

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

ASTM. 2013. ASTM D396-13, Standard Specification for Fuel Oils. West Conshohocken, PA: ASTM International.
Cleaver Brooks. 2010. The Boiler Book, Section E, Emissions. Thomasville, GA: Cleaver
Brooks. www.cleaver-brooks.com/Reference-Center/Insights/Emissions-Guide.aspx
Crilley, J.S. 2010. Boiler MACT overview. District Energy, Fourth Quarter 2010.
DOD. 2007. Unified Facilities Criteria: Central Steam Boiler Plants. UFC 3-430-02FA,
U.S. Department of Defense, Washington, DC.
EPA. 1995. AP-42, Compilation of Air Pollutant Emission Factors, 5th ed. Washington,
DC: U.S. Environmental Protection Agency.
EPA. 2012. Clean Air Act. Washington, DC: U.S. Environmental Protection Agency.
www.epa.gov/air/caa/
EPA. 2013a. New Source Performance Standards (NSPS). Code of Federal Regulations,
40 CFR Part 60. Washington, DC: U.S. Environmental Protection Agency.
www.ecfr.gov/cgi-bin/text-idx?c=ecfr&tpl=/ecfrbrowse/Title40/40cfr60_main_02.tpl
EPA. 2013b. National Emission Standards for Hazardous Air Pollutants (NESHAP).
Washington, DC: U.S. Environmental Protection Agency. www.epa.gov/ttn/atw/mact
fnlalph.html
FM Global. 2013. Approval Guide. Online database. www.fmglobal.com/page.aspx?id=
50040000
HI. 2000. Pump Standards. Parsippany, NJ: Hydraulic Institute.
IES. 2011. The Lighting Handbook, 10th ed. New York: Illuminating Engineering Society
of North America.
Inwood, D. 2011. Biomass gasification: A viable option for campus district energy systems. District Energy, Fourth Quarter 2011.
ISA. 2009. Control Valve Primer: A User Guide, 4th ed. Research Triangle Park, NC:
International Society for Automation.
NARA. 2010a. National Emission Standards for Hazardous Air Pollutants for Major
Sources; Industrial, Commercial, and Institutional Boilers and Process Heaters (40
CFR Part 63, Subpart DDDDD); Federal Register, Vol. 75, No. 107, Proposed Rules,
p. 32012 (June 4). Washington, DC: Office of the Federal Register, National Archives
and Records Administration.
NARA. 2010b. National Emission Standards for Hazardous Air Pollutants for Major
Sources; Industrial, Commercial, and Institutional Boilers and Process Heaters (40
CFR Part 63, Subpart JJJJJJ); Federal Register, Vol. 75, No. 107, Proposed Rules, p.
32901 (June 4). Washington, DC: Office of the Federal Register, National Archives
and Records Administration.
NFPA. 1987. NFPA 85G, Standard for the Prevention of Furnace Implosions in Multiple
Burner Boiler-Furnaces. Quincy, MA: National Fire Protection Association.
NFPA. 2011. National Electrical Code (NFPA 70). Quincy, MA: National Fire Protection Association.
NFPA. 2012. National Fuel Gas Code (NFPA 54/ANSI Z223.1). Quincy, MA: National
Fire Protection Association.
NIBS. 2012. Whole Building Design Guide. Washington, DC: National Institute of Building Sciences. www.wbdg.org
Oliker, I. 1972. Deaeration of Water in Central Boiler Plants and District Heating Systems. Leningrad, Russia: Stroiizdat.
Oliker, I. 1989. Chapter 15, Deaeration. The ASME Handbook on Water Technology for
Thermal Power Systems. New York: ASME.
OSHA. 2013. Occupational Safety and Health Standards, 29 CFR Part 1910. Washington, DC: Occupational Safety and Health Administration, U.S. Department of Labor.

3.94

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

3 Central Plant Design for Steam and Hot Water

Penny, J. 2000. From woodland to warmth: Biomass district energy in Canada. District
Energy, Second Quarter 2000.
Sasanov, B.V. 1974. Combined Heat and Power Plants. Moscow: Energiya.
Solovyev, U.P. 1976. Design of Large Central Boiler Plants for District Heating. Moscow: Energiya.
Tao, W., and H.C. Laird. 1984. Modern boiler plant design. Heating/Piping/Air Conditioning 56(11):6982.
Zinko, H., B. Bhm, H. Kristjansson, U. Ottosson, M. Rm, and K. Sipil. 2008. District
heating distribution in areas with low heat demand density. Report No. 2008: 8DHC08-03, Annex VIII, International Energy Agency, Paris, France.

3.95

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4
Distribution Systems

HYDRAULIC CONSIDERATIONS
Objectives of Hydraulic Design
Although the distribution of a thermal utility such as hot water encompasses many of
the aspects of domestic water distribution, many dissimilarities also exist; thus, the
designs should not be approached in the same manner. Thermal utilities must supply sufficient energy at the appropriate temperature and pressure to meet consumer needs.
Within the constraints imposed by the consumers end use and equipment, the required
thermal energy can be delivered with various combinations of temperature and pressure.
Computer-aided design (CAD) methods are available for thermal piping networks
(Bloomquist et al. 1999; COWIconsult 1985; Rasmussen and Lund 1987; Reisman 1985;
Lund-Hansen 2010; Lund-Hansen and Hansen 2011). The use of such methods allows the
rapid evaluation of many alternative designs. These methods may also be used to optimize operation and maintenance (O&M) of distribution systems (Lund-Hansen 2010;
Lund-Hansen and Hansen 2011).
General steam system design can be found in Chapter 11 of ASHRAE Handbook
HVAC Systems and Equipment (2012) as well as in the District Heating Handbook from
International District Heating Association (IDHA) (1983). For water systems, consult
Chapter 13 of ASHRAE HandbookHVAC Systems and Equipment and IDHAs District
Heating Handbook.

Water Hammer
The term water hammer is used to describe several phenomena that occur in fluid
flow. Although these phenomena differ in nature, they all result in stresses in the piping
that are higher than normally encountered. Water hammer can have a disastrous effect on
a district heating utility by bursting pipes and fittings and threatening life and property.
In steam systems (IDHA 1983), water hammer is caused primarily by condensate collecting in the bottom of the steam piping. Steam flowing at velocities 10 times greater
than normal water flow picks up a slug of condensate and accelerates it to a high velocity.
At a point where flow changes direction, the slug of condensate will collide with the pipe
wall. To prevent this type of water hammer, condensate must be prevented from collecting
in steam pipes by the use of proper steam pipe pitch and adequate condensate collection
and return facilities.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Water hammer also occurs in steam systems due to rapid condensation of steam during system warm-up. Rapid condensation decreases the specific volume and pressure of
steam, which precipitates pressure shock waves. This type of water hammer event is
sometimes called condensation-induced water hammer. In some instances this form of
water hammer may be prevented by controlled warm-up of the piping. Valves should be
opened slowly and in stages during warm-up. Large steam valves should be provided with
smaller bypass valves to slow the warm-up.
Looped systems or systems where dead legs of condensate may collect are susceptible to condensation-induced water hammer and often present special challenges. It
becomes necessary to ensure condensate is properly drained without any subcooled pockets. This often will require careful analysis of the system and the start-up procedure,
which may vary, requiring analysis at each start-up incidence. Some of the potential situations that can result in condensation-induced water hammer are discussed by Vogler
(2011).
Water hammer in hot-water distribution systems is caused by sudden changes in flow
velocity, which causes pressure shock waves. The two primary causes are pump failure
and sudden valve closures. A simplified method to determine maximum resultant pressure may be found in Chapter 22 of ASHRAE HandbookFundamentals (2009). More
elaborate methods of analysis can be found in works by Streeter and Wylie (1979), Fox
(1977), and Stephenson (1981). Preventive measures include operational procedures and
special piping fixtures such as surge columns.

Pressure Losses
Friction pressure losses occur at the interface between the inner wall of a pipe and a
flowing fluid due to shear stresses. In steam systems, these pressure losses are compensated for with increased steam pressure at the point of steam generation. In water systems,
pumps are used to increase pressure at either the plant or intermediate points in the distribution system. The calculation of pressure loss is discussed in Chapters 3 and 22 of
ASHRAE HandbookFundamentals (2009).

Pipe Sizing
Ideally, the appropriate pipe size should be determined from an economic study of the
life-cycle cost (LCC) for construction and operation. In practice, however, this study is
seldom done due to the effort involved. Instead, criteria that have evolved from practice
are frequently used for design. These criteria normally take the form of constraints on the
maximum flow velocity or pressure drop. Chapter 22 of ASHRAE HandbookFundamentals (2009) provides velocity and pressure drop constraints. Noise generated by
excessive flow velocities is usually not a concern for district heating distribution systems
outside of buildings. For steam systems, maximum flow velocities of 200 to 250 ft/s
(60 to 70 m/s) are recommended (IDHA 1983). For water systems, Europeans historically
used the criterion that pressure losses should be limited to 0.44 psi per 100 ft (100 Pa/m)
of pipe (Bhm 1988). However, other studies have indicated that higher levels of pressure
loss may be acceptable (Stewart and Dona 1987) and warranted from an economic standpoint (Bhm 1986; Koskelainen 1980; Phetteplace 1989).

Diversity of Demand
When establishing design flows for thermal distribution systems, the diversity of consumer demands should be considered (i.e., the various consumers maximum demands do
not occur at the same time). Thus, the heat supply and main distribution piping may be
sized for a maximum load that is somewhat less than the sum of the individual consum-

4.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

ers maximum demands. For steam systems, Geiringer (1963) suggests diversity factors
of 0.80 for space heating and 0.65 for domestic hot-water heating and process loads. Geiringer also suggests that these factors may be reduced by approximately 10% for hightemperature water (HTW) systems, owing to the significant heat capacity of the high-temperature hot-water within the piping network. Werner (1984) conducted a study of the
heat load on six operating low-temperature hot-water systems in Sweden and found diversity factors ranging from 0.57 to 0.79, with the average being 0.685.

Network Calculations
Calculating flow rates and pressures in a piping network with branches, loops,
pumps, and heat exchangers can be difficult without the aid of a computer. Computeraided methods were initially developed primarily for domestic water distribution systems.
These may be applied to heat distribution systems with appropriate assumptions and supplemental calculations; however, methods for heat distribution systems are now available
(Lund-Hansen 2010; Lund-Hansen and Hansen 2011). CAD methods usually incorporate
methods for hydraulic analysis as well as for calculating heat losses and delivered water
temperature at each consumer location. Calculations are usually carried out in an iterative
fashion. For example, the calculation might start with constant supply and return temperatures throughout the network. After initial estimates of the design flow rates and heat
losses are determined, refined estimates of the actual supply temperature at each consumer location are computed. Flow rates at each consumer location are then adjusted to
ensure that the load is met with the reduced supply temperature, and the calculations are
repeated.

Condensate Drainage and Return


Condensate forms in operating steam lines as a result of heat loss. When a steam systems operating temperature is increased, condensate also forms as steam warms the piping. At system start-up, these loads usually exceed any operating heat loss loads; thus,
special provisions should be made.
To drain the condensate, steam piping should slope toward a collection point called a
drip station. Drip stations are located in access areas or buildings where they are accessible for maintenance. Steam piping should slope toward the drip station a minimum of
1 in. in 40 ft (2 mm/m). For direct burial systems where construction tolerances make it
difficult to maintain a uniform slope, higher slopes of 1 in. in 20 ft (4 mm/m) are recommended. If possible, the steam pipe should slope in the same direction as steam flow. If it
is not possible to slope the steam pipe in the direction of steam flow, increase the pipe size
to at least one size greater than would normally be used. This will reduce the flow velocity of the steam and provide better condensate drainage against the steam flow. Drip stations should be spaced no further than 500 ft (150 m) apart in the absence of other
requirements.
Drip stations consist of a short piece of pipe (called a drip leg) positioned vertically at
the bottom of the steam pipe as well as a steam trap and appurtenant piping. The drip leg
should be the same diameter as the steam pipe. The length of the drip leg should provide
a volume equal to 50% of the condensate load from system start-up for steam pipes of 4
in. (100 mm) diameter and larger and 25% of the start-up condensate load for smaller
steam pipes (IDHA 1983). Steam traps should be sized to meet the normal load from
operational heat losses only. Start-up loads should be accommodated by manual operation
of the bypass valve.
Steam traps are used to separate the condensate and noncondensable gases from the
steam. For drip stations on steam distribution piping, use inverted bucket or bimetallic

4.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

thermostatic traps. Some steam traps have integral strainers while others require separate
strainers. Ensure that drip leg capacity is adequate when thermostatic traps are used
because they will always accumulate some condensate.
If it is to be returned, condensate leaving the steam trap flows into the condensate
return system. If steam pressure is sufficiently high, it may be used to force the condensate through the condensate return system. With low-pressure steam or on systems where
a large pressure exists between drip stations and the ultimate destination of the condensate, condensate receivers and pumps must be provided.
Schedule 80 steel piping is recommended for condensate lines because of the extra
allowance for corrosion that it provides. Steam traps have the potential of failing in an
open position; thus, nonmetallic piping must be protected from live steam where its temperature/pressure would exceed the limitations of the piping. Nonmetallic piping should
not be located so close to steam pipes that heat losses from the steam pipes could overheat
it. Additional information on condensate removal may be found in Chapter 11 of ASHRAE
HandbookHVAC Systems and Equipment (2012). Information on sizing of condensate
return piping may be found in Chapter 22 of ASHRAE HandbookFundamentals (2009).

DISTRIBUTION SYSTEM CONSTRUCTION


The distribution of heat in the form of either heated water or steam is a vastly different enterprise than that of other utilities such as water or sewage. When the designer fails
to recognize this and design accordingly, the system is destined to fail prematurely.
Among the special considerations that must be taken into account in the design of a heat
distribution system are the following:
The need to deal with expansion and contraction of the system and the forces it
generates.
For buried systems, the adverse nature of the environment below the Earths surface where the nearly universal presence of water results in a monumental challenge in several ways:
Maintaining a dry environment for the thermal insulation
Providing corrosion protection for metallic portions of the buried system
Providing dry environments for appurtenances such as valves, drains,
vents, steam traps, etc.
The impacts of the elevated operating temperatures of these systems on materials, particularly plastics.
The difficulty of executing a proper field joint of a prefabricated system under
normal construction tolerance, practices, and field conditions.
The inability to inspect the vast majority of the system easily once construction
has been completed.
All these factors conspire to make it approximately an order of magnitude more difficult to design and construct a buried heat distribution system when compared to most
other buried utilities or normal building systems. The successful application of an underground heat distribution system relies upon proper design, system fabrication, installation, and O&M.
The combination of aesthetics, first cost, safety, and LCC naturally divide distribution systems into two distinct categories: aboveground and underground distribution systems. The materials needed to ensure long life and low heat loss further classify district
heating systems into low-temperature, medium-temperature, and high-temperature systems (see Chapter 2 for the delineation of these classes). The temperature range for
medium-temperature systems is usually too high for the materials that are used in low-

4.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

temperature systems; however, the same materials that are used in high-temperature systems are typically used for medium-temperature systems. Because low-temperature systems have a lower temperature differential between the working fluid temperatures and
the environment, heat loss is inherently less. In addition, the selection of efficient insulation materials and inexpensive pipe materials that resist corrosion is much greater for
low-temperature systems.
Aboveground systems (Figure 4.1) have a lower first cost and a lower LCC when
compared to underground distribution systems because the aboveground systems can be
maintained easily and constructed with materials that are readily available. Generally,
aboveground systems are acceptable where they are hidden from view or can be hidden
by landscaping. Poor aesthetics, physical security, and the risk of vehicle damage to the
aboveground system remove them from contention for many projects.
Common underground systems that are completely field fabricated include the walkthrough tunnel (Figure 4.2), the concrete surface trench (Figures 4.3 and 4.4), the deepburial small tunnel (Figure 4.5), and underground systems that use poured insulation (Figure 4.6) to form an envelope around the carrier pipes.
Field-assembled systems must be designed in detail, and all materials must be specified by the Engineer of Record (EOR). Evaluation of the project site conditions indicates
which type of system should be considered for the site. For instance, the shallow trench
system is best where utilities buried deeper than the shallow trench bottom need to be
avoided and where the covers can serve as sidewalks. Direct buried prefabricated conduit,
with a thicker steel casing, may be the only system that can be used for medium- or hightemperature systems on flooded sites. Even where other system constructions are the primary method used, a prefabricated conduit is often used for short distances between
buildings and the main distribution system and where the owner is willing to accept
higher LCCs.
Conditions that will result in surrounding soil temperatures being higher than those
normally expected include high system operating temperature, deep burial, dry soil, an
old system with lower levels of insulation, and/or a failed/failing system. Rigid extruded
polystyrene insulation may be used to insulate adjacent utilities from the impact of a buried heat distribution pipe; however, the temperature limit of the extruded polystyrene
insulation must not be exceeded. A numerical analysis of the thermal problem may be
required to ensure that the desired effect is achieved. Chapter 6 contains information on
thermal analysis methods for district heating and cooling systems.
Tunnels that provide walk-through or crawl-through access can be buried in nearly
any location without causing future problems because utilities are typically placed in the
tunnel. Regardless of the type of construction, it is usually cost-effective to route distribution piping through the basements of buildings, but only after liability issues are
addressed. In laying out the main supply and return piping, redundancy of supply and
return should be considered. If a looped system is used to provide redundancy, flow rates
under all possible failure modes must be addressed when sizing and laying out the piping.
Manholes are required to provide access to underground systems at critical points such as
where there are high or low points on the system profile that vent trapped air, where the
system can be drained, at elevation changes in the distribution system that are needed to
maintain the required slope, at major branches with isolation valves, at steam traps and
condensate drainage points on steam lines, and at mechanical expansion devices. Sump
pumps or other methods of positive manhole drainage are required in manholes (see the
Valve Vaults and Entry Pits section of this chapter for more details).
To facilitate leak location and repair and to limit damage caused by leaks, access
points generally should be spaced no farther than 500 ft (150 m) apart. Special attention

4.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

must be given to the safety of personnel who come in contact with distribution systems or
who must enter spaces occupied by underground systems. The regulatory authoritys definition of a confined space and the possibility of exposure to high-temperature or highpressure piping can have a significant impact on the access design, which must be
addressed by the EOR. Hence, gravity venting of tunnels is required. Moreover, manholes
and tunnels should have electric- or steam-driven sump pumps as well as lighting and
convenience outlets for inspection and maintenance of anchors, expansion joints, and piping. Sump pumps may not be required if the manholes can be drained to a sanitary system. French drains are usually not acceptable because groundwater will back flow into
the manhole when high groundwater levels occur.

ABOVEGROUND SYSTEM
The aboveground distribution system consists of an insulated carrier pipe and a structural support system. The aboveground system cross section consists of a distribution
pipe, the insulation that surrounds the pipe, and a protective jacket that surrounds the
insulation. The jacket may have an integral vapor retarder; however, in heating applications the vapor retarder is not needed or recommended. A watertight jacket is required to
keep storm water out of the insulation. The jacket material can be aluminum, stainless
steel, galvanized steel, reinforced plastic, a multi-layered fabric and organic cement composite, or a combination of these. Plastics and organic cements exposed to sunshine must
be ultraviolet-light resistant. Jacketing materials for systems that are installed near grade
level must be able to withstand any expected pedestrian traffic or vandalism.
Aboveground systems may be installed at any height. For example, some systems are
installed essentially at grade level and earth berms and/or plantings can be used to disguise their presence. The aboveground system is supported by structural columns typically made of wood, steel, or concrete. A crossbar is often placed across the top of the
column when more than one distribution pipe is supported from one column. Sidewalk
and road crossings require an elaborate support structure to elevate the distribution piping
above traffic. Pipe expansion and contraction is normally taken up in loops, elbows,
bends, and elevated traffic crossings; Figure 4.1 illustrates a traffic crossing for an
aboveground system that functions as an expansion loop as well. Manufactured expansion
joints may be used, but they are usually not recommended because of a shorter life or a
higher frequency of required maintenance than the remainder of the system. Supports that
attach the distribution pipes to the support columns are commercially available as
described in Manufacturers Standardization Society (MSS) standards MSS SP-58 (2009)
and MSS SP-69 (2003). The distribution pipes should have welded joints.
The aboveground system normally has the lowest first cost and the lowest LCC
because it can be maintained easily and constructed with materials that are readily available. Generally, aboveground systems are acceptable where they are hidden from view or
can be hidden by landscaping. Its major drawbacks are its poor aesthetics, the safety hazard if it is struck by vehicles or equipment, and its susceptibility to freezing in cold climates if circulation is stopped or if heat is not added to the working medium. These
drawbacks often remove this system from contention as a viable option.
Although portions of the aboveground system are sometimes partially factory prefabricated, more typically it is entirely field fabricated of components such as pipes, insulation, pipe supports, and insulation jackets or protective enclosures that are commercially
available. Field-assembled systems must be designed in detail, and all materials must be
specified with great care by the EOR.

4.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

Figure 4.1 Aboveground system traffic crossing that also functions as an expansion loop.
Note the truss support structure for the pipe at the crossing as well as
the high-point vents and low-point drains.

UNDERGROUND SYSTEMS
Underground systems are used where aesthetics and safety are major issues. The
safety issue is related to vehicle traffic and pedestrian risk. Burying a system causes more
problems with materials, design, construction, and maintenance that have historically
been difficult to solve. Direct buried prefabricated conduit, concrete surface trenches, and
other underground systems must be routed to avoid existing buried utilities, which
requires a detailed site survey, considerable design effort, and additional risk during construction. Both the route and the burial depth must be considered. In the absence of a
detailed soil temperature distribution study or calculations using the methods outlined in
Chapter 6, direct buried heating systems should be spaced more than 15 ft (4.6 m) from
other utilities constructed of plastic components because the temperature of the soil during dry conditions can be high enough to reduce the yield strength of plastic material to
an unacceptable level.
The U.S. federal government conducted investigations of buried heat distribution systems from 1956 to 1984. The reports produced are typically referred to as the BRAB
reports (BRAB 1957, 1958, 1963a, 1963b, 1963c, 1964, 1966, 1975). Some of the types
of underground systems investigated were systems with cast iron casings, prefabricated
metal sheathings, concrete trenches, asbestos cement casings, metal jacketed asphalt casings, poured insulating hydrocarbon, bell and spigot tile with internal drains, bell and
spigot tile with external drains, insulating concrete, insulating cement with fiber jackets,
sand and gravel concrete casings, and various schemes where the carrier pipe and insulation were wrapped with plastic sheet. The committee concluded that groundwater in the
insulation was the prime cause of failures and that a prefabricated conduit type system
with a pressure-testable steel casing and with provisions for drying the pipe insulation
when the insulation gets wet was the best choice for the most severe sites (see Figures 4.7,
4.9, 4.10, 4.11, and 4.13) because the prefabricated steel cased system can be tested to

4.7

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

15 psig (104 kPa) during construction and anytime during its life to ensure that the pipe
insulation can be kept dry. This became known as the drainable, dryable, testable (DDT)
system. The carrier pipe sections were joined with welded joints, a method currently recommended. Later investigations (Phetteplace et al. 1998, 2000) also revealed that nonwelded joints performed poorly in heat distribution systems. The BRAB reports
recommended stringent quality control during design, procurement, and construction and
rigorous maintenance throughout its life. The committee also recommended concrete
manholes with electric sump pumps, manholes spaced no more than 500 ft (150 m),
cathodic protection, safe manholes large enough for maintenance personnel, and condensate pipes with extra heavy wall thickness that are not located in the same insulating
envelope as the steam line. Based on the committee findings and later studies (Segan and
Chen 1984), the U.S. Department of Defense (DOD) delineated manhole requirements
(UFC 3-430-01FA, par. 36) (2003).
Tables 4.1 and 4.2 present the site classification criteria that grew out of the BRAB
efforts as well as subsequent federally sponsored research. While additional materials and
types of system construction have entered the market since the site classification criteria
of Tables 4.1 and 4.2 were developed, this site classification scheme is still valid. Thus,
the site classification criteria of Tables 4.1 and 4.2 should continue to be used as a guide.
For example, systems that are being designed for installation on sites classified as
severe by either Table 4.1 or Table 4.2 must be designed to deal with groundwater presence as a certainty. For the system to be successful on a severe site, the designer must
also ensure that additional supervision and inspection are provided during the construction process.
The designer needs to be aware of the special requirements. An underground heat
distribution system is not a typical utility like gas, domestic water, and sanitary systems.
The underground environment is a very inhospitable one due primarily to the potential
presence of groundwater, moisture found in nearly all soils regardless of the water table
level, and the corrosiveness of soils. As stated previously, heat distribution systems
require an order of magnitude more design effort and construction inspection accuracy
when compared to gas, water, and sanitary distribution projects. The thermal effects and
difficulty of keeping the insulation dry make it much more difficult to successfully
design, construct, maintain, and operate such systems. When metal components are present, corrosion protection also becomes a major issue. For these reasons, the underground
system can cost many times as much as an aboveground system to build. In addition,
underground heat distribution systems require much more effort to operate and maintain,
all of which is compounded by the fact that access to the system is usually very difficult
and most of the system is hidden from casual inspection. Underground heat distribution
systems must be designed for zero leakage and must account for thermal expansion, degradation of material properties as a function of temperature, high pressure and transient
shock waves, heat loss restrictions, the presence of groundwater, wheel loads and differential ground settlement, and the potential for accelerated corrosion. These special problems add to the first cost of the system and also make it difficult to design a successful
system. Many approaches to underground heat distribution have been tried and nearly as
many have failed. In the past, solutions proposed to resolve one problem in underground
heat distribution systems often created new, more serious problems that were not recognized until premature failure occurred (see BRAB [1975]). Segan and Chen (1984)
describe the type of premature failures that may occur if this guidance is ignored. Failures of underground systems can be extremely expensive, as by the time the problem has
been identified and the system can be excavated for repair much of the system may be

4.8

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

Table 4.1 Site Classification Definition based on Known Underground Water Conditions
Site
Classification

General Conditions for Classification


1. The water table is expected to be frequently above the bottom of the system and surface water is
expected to accumulate and remain for long periods in the soil surrounding the system.

Severe

or
2. The water table is expected to be occasionally above the bottom of the system and surface water is
expected to accumulate and remain for long periods in the soil surrounding the system.
1. The water table is expected to be occasionally above the bottom of the system and surface water is
expected to accumulate and remain for short periods (or not at all) in the soil surrounding the system.
or

Bad

2. The water table is expected never to be above the bottom of the system but surface water is expected
to accumulate and remain for short periods in the soil surrounding the system.
1. The water table is expected never to be above the bottom of the system but surface water is expected
to accumulate and remain for short periods (or not at all) in the soil surrounding the system.
or
2. The water table is expected never to be above the bottom of the system but surface water is expected
to accumulate and remain for brief or occasional periods in the soil surrounding the system.

Moderate

or
3. The water table is expected never to be above the bottom of the system and surface water is not
expected to accumulate or remain in the soil surrounding the system.

Table 4.2 Site Classification Criteria based on Subsurface Soil Investigation


Site
Classification

Severe

Bad

Water Table Level

Soil Types*

Terrain

Precipitation Rates or
Irrigation Practices in Area

Water table within 1 ft (300 mm)


of bottom of system

Any

Any

Any

Water table within 5 ft


(1500 mm) of bottom of system

GC, SC, CL, CH, OH

Any

Any

Water table within 5 ft


(1500 mm) of bottom of system

GW, GP, SW, SP

Any

Any

No groundwater encountered

GC, SC SW, CH, OH

Any

Equivalent to 3 in. (75 mm) or more


in any one month or 20 in. (500 mm)
or more in one year

No groundwater encountered

GM, SM, ML, OL, MH

Any

Equivalent to 3 in. (75 mm) or more


in any one month or 20 in. (500 mm)
or more in one year

GC, SC, CL, CH, OH

Any except
low areas

Equivalent to less than 3 in. (75 mm)


in any one month or less than
20 in. (500 mm) in one year

GW, GP, SW, SP,

Any

Any

GM, SM, ML, SM,

Any

Equivalent to less than 3 in. (75 mm)


in any one month or less than
20 in. (500 mm) in one year

or

or

or
No groundwater encountered
Moderate
or
No groundwater encountered
or
No groundwater encountered

* See Table 4.3 of this guide, TM 3-34.64 (DA 2012), or ASTM D2487 (2011a) for soil classifications.

4.9

Table 4.3 Unified Soil Classification System (DA 2012)

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

4.10

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

degraded to the point where it cannot be repaired. The reader is cautioned to avoid new or
unproven designs/concepts without a through introduction to the history of these systems.

Site-Fabricated Underground Systems


Common types of site-fabricated underground systems, in the order of O&M preference, are the walk-through tunnel, the concrete surface trench, the deep-buried small tunnel, and the poured insulation envelope system. Unfortunately these options will be in
about the same order when arranged by the typical first costs from highest to lowest,
which makes life-cycle cost analysis (LCCA) necessary if more than one type of system
is being considered. These systems are discussed in detail in the following subsections.
Walk-Through Tunnel
The walk-through tunnel system (Figure 4.2) consists of a field-erected tunnel that is
large enough for someone to walk through when the distribution pipes are in place. It is
essentially an aboveground system enclosed in a tunnel. The tunnel is buried deep enough
to cover the top with earth and is large enough so that routine maintenance and inspection
can be done easily without excavation. The preferred construction material for the tunnel
walls and top cover is cast-in-place reinforced concrete. Precast concrete floor and wall
sections have not been successful because of groundwater leakage at the large number of
oblique joints and nonstandard sections required to follow the earth topography and to
slope the floor to drain. Masonry units and metal preformed sections have been used to
construct tunnels and tops with less success due to groundwater leakage and metal corrosion. The distribution pipes are supported from the tunnel wall or floor with pipe supports
that are commonly used on aboveground systems or in buildings. Due to construction
costs, for walk-through tunnels, expansion joints and changes in direction are often used
to absorb pipe movements due to changes in temperature, as opposed to expansion loops.

Figure 4.2 Walk-through tunnel.


Courtesy of ASHRAE (2012, Chapter 12, Figure 12)

4.11

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Because ready access for maintenance and replacement is possible with the walk-through
tunnel, expansion joints are a reasonable solution on a LCC basis.
Some groundwater will penetrate the top and walls of the tunnel; therefore, a water
drainage system must be provided. Usually electric lights and electric service outlets are
provided for ease of inspection and maintenance. This system has the highest first cost of
all underground systems; however, it can have the lowest LCC due to ease of maintenance, the ability to correct construction errors easily, and an extremely long life. With
respect to safety, this type of tunnel is usually judged a confined space. If it is, the precautions delineated by the cognizant safety authority must be followed for entry.
Concrete Surface Trench
The concrete surface trench system (Figures 4.3 and 4.4) is a partially buried system.
The floor is usually about 3 ft (1 m) below surface-grade elevation. It is only wide enough
for the carrier pipes and the pipe insulation plus some additional width to allow for pipe
movement and possibly enough room for a person to stand on the floor. The trench usually is about as wide as it is deep. The top is constructed of reinforced concrete covers that
protrude slightly above the surface and may also serve as a sidewalk. The floor and walls
are usually cast-in-place reinforced concrete and the top is either precast or cast-in-place
concrete. Precast concrete floor and wall sections have not been successful because of
groundwater leakage at the large number of oblique joints and nonstandard sections
required to follow the surface topography and to slope the floor to drain. This system is
designed to handle the storm water and groundwater that enters the system, so the floor is
always sloped toward a drainage point. A drainage system is required at all floor low
points.
Crossbeams that attach to the side walls are the preferred method of supporting the
carrier pipes. This keeps the floor free of obstacles that would interfere with drainage and
allows the distribution pipes to be assembled before lowering them onto the pipe supports. Also, floor-mounted pipe supports tend to corrode. The carrier pipes, the pipe sup-

Figure 4.3 Concrete surface trench.


Courtesy of ASHRAE (2012, Chapter 12, Figure 13)

4.12

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

ports, the expansion loops and bends, and the insulation jacket are similar to those used in
aboveground systems, with the exception of the pipe insulation. Experience with these
systems indicates that flooding will occur several times during their design life; therefore,
the insulation must be able to survive flooding and boiling and then return to near its original thermal efficiency. The pipe insulation is covered with a metal or plastic jacket to
protect the insulation from abuse and from storm water that enters at the top cover butt
joints. Small inspection ports of about 12 in. (300 mm) diameter may be cast into the top
covers at key locations so that the system can be inspected without removing the top covers. All replaceable elements such as valves, condensate pumps, steam traps, strainers,
sump pumps, and meters are located in valve vaults. The first cost of this system is among
the lowest for underground systems because it uses typical construction techniques and
materials. The LCC is often the lowest because it is easy to maintain, correct construction
deficiencies, and repair leaks. Extensive details of the design of this type of system are
available from the DOD (2003).
Deep-Bury Tunnel
The tunnel in the deep-bury system (Figure 4.5) is only large enough to contain the
distribution piping, the pipe insulation, and the pipe supports. One type of deep-bury tunnel is the shallow concrete surface trench covered with earth and sloped independent of
the topography. Because the system is covered with earth, it is essentially not maintainable between valve vaults without major excavation. All details of this system must be
designed and all materials must be specified by the EOR. Precast concrete floor and wall
sections have not been successful because of groundwater leakage at the large number of
oblique joints and nonstandard sections required by changes in direction along the system
route and the requirement to slope the floor to drain.
Because this system is not readily maintainable between valve vaults, great care must
be taken to select materials that will last for the intended life and to ensure that the
groundwater drainage system will function reliably. This system is intended to be used on
sites where the groundwater elevation is typically lower than the bottom of the tunnel.

Figure 4.4 Concrete surface trench during construction. This system is being instrumented for
temperature and heat loss measurement (see Phetteplace et al. [1991] for details).

4.13

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 4.5 Deep-bury small tunnel.


Courtesy of ASHRAE (2012, Chapter 12, Figure 14)

The system can tolerate some groundwater presence depending on the watertightness of
the construction and the capacity and reliability of the internal drainage. But even in desert areas, storms occur that expose underground systems to flooding; therefore, as with
other types of underground systems, the tunnel must be designed to handle groundwater
or storm water that enters the system. The distribution pipe insulation must be of the type
that can withstand flooding and boiling and still retain its thermal efficiency.
Construction of this system is typically started in an excavated trench by pouring a
cast-in-place concrete base that is sloped so intruding groundwater can drain to the valve
vaults. The slope selected must also be compatible with the pipe slope requirements of the
distribution system. The concrete base may have provisions for the supports for the distribution pipes, the groundwater drainage system, and the mating surface for the side walls.
The side walls may have provisions for the pipe supports if the pipes are not bottom supported. If the upper portion is to have cast-in-place concrete walls, the bottom may have
reinforcing steel for the walls protruding upward. The pipe supports, the distribution
pipes, and the pipe insulation are all installed before the top cover is installed and covered
with earth.
The groundwater drainage system may be a trough formed into the concrete bottom, a
sanitary drainage pipe cast into the concrete bottom, or a sanitary pipe that is located
slightly below the concrete base. The top cover for the system is typically either of castin-place concrete or preformed sections such as precast concrete sections or half-round
clay tile sections. The top covers must mate to the bottom and each other as tightly as possible to limit the entry of groundwater. After the covers are installed, the system is covered with earth to match the existing topography.
Poured Insulation
The poured insulation system (Figure 4.6) is buried with the distribution system pipes
encased in an envelope of insulating material and the insulation envelope covered with a
thick layer of earth as required to match existing topography. This system is used on sites
where the groundwater is typically below the system. Like other underground systems,
experience indicates that the system will be flooded because the soil will become satu-

4.14

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

Figure 4.6 Poured insulation system.


Courtesy of ASHRAE (2012, Chapter 12, Figure 15)

rated with water several times during the design life; therefore, the design must accommodate flooded conditions.
The insulation material in this system serves several functions. It may support the distribution pipes and it must support earth loads. The insulation must prevent groundwater
from entering the interior of the envelope and it must have long-term resistance to physical breakdown due to heat and water. The insulation envelope must allow the distribution
pipes to expand and contract axially as the pipes change temperature. In elbows, expansion loops, and bends, the insulation must allow formed cavities for lateral movement of
the pipes or be able to migrate around the pipe without significant distortion of the insulation envelope while still retaining the required structural load carrying capacity. Special
attention must be given to corrosion of metal parts and water infiltration at anchors and
structural supports that penetrate the insulation envelope.
Hot distribution pipes tend to drive moisture out of the insulation as steam; however,
pipes used to distribute a cooling medium tend to condense water in the insulation, which
reduces the insulation thermal resistance. Therefore, this type of system construction
would not be a good choice for a cooling distribution system. A groundwater drainage
system may be required depending on the insulation material selected and the severity of
the groundwater; however, if such a drainage is needed, it is a strong indicator that this is
not the proper system for the site conditions.
This system is constructed by excavating a trench with a bottom slope that matches
the desired slope of the distribution piping. The width of the bottom of the trench is usually the same as the width of the insulation envelope because it serves as a form. The distribution piping is then assembled in the trench and supported at the anchors and by
blocks that are removed as the insulation is poured in place. The form for the insulation
can be the trench bottom and sides, wooden forms, or sheets of plastic depending on the
type of insulation used and the site conditions. The insulation envelope is covered with
earth to complete the installation.
The EOR is responsible for finding an insulation material that fulfills all of the previously mentioned requirements. At present, no standards have been developed for insulation used in this type of application. Hydrophobic powders, which are a special type of

4.15

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

pulverized rock that is treated to be water repellent, have been used successfully. The
hydrophobic characteristic of this powder prevents water from dampening the powder and
has some capability as a barrier for preventing water from entering the insulation envelope. This insulating powder typically has a much higher thermal conductivity than mineral wool or foam pipe insulation; therefore, the thickness of the poured envelope must be
significantly greater, possibly four times as thick as mineral fiber pipe insulation. The
EOR should expect great difficulty in obtaining ASTM certifications for the thermal conductivity of these materials, compressed or uncompressed, for the entire temperature
range needed. Calculation of the heat losses requires a number of assumptions as discussed in Chapter 6.

Prefabricated Conduit Systems


The term conduit has come to denote an entire assembly, which consists of a carrier
pipe, the pipe insulation, the casing, and the exterior casing coating (Figure 4.7). The conduit is assembled in a factory and shipped as unit called a conduit section. The pipe that
carries the working medium is called a carrier pipe and the outermost perimeter enclosure is called a casing.
Each conduit section is shipped in lengths up to 40 ft (12 m). Elbows, tees, loops, and
bends are factory prefabricated to match the straight sections. The prefabricated components are assembled at the construction site; therefore, a construction contract is typically
required for trenching, backfilling, connecting to buildings, connecting to distribution
systems, constructing valve vaults, and performing some electrical work associated with
sump pumps, power receptacles, and lights.
Much of the design work is done by the factory that manufactures the prefabricated
sections; however, the field work must be designed and specified by the EOR. Prefabricated components create a serious problem with accountability. For comparison, when
systems are entirely field assembled, the design responsibility clearly belongs to the EOR
and the assembly of the system is clearly the responsibility of the construction contractor.
When a condition arises where a conduit system cannot be built without modifying prefabricated components, or if the construction contractor does not follow the instructions from
the prefabricator, a serious conflict of responsibility arises. For these reasons, it is imperative that the EOR takes full responsibility for the design, including field changes.

Figure 4.7 Conduit system components at manhole termination.

4.16

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

One consideration in the design of the steel casing is crushing. Crushing loads have
been used (erroneously) to size the steel casing thickness assuming that corrosion was not
a factor. But for steel casings, the corrosion rate is usually the controlling factor because
the casing temperature can range from less than 100F (38C) to more than 300F
(149C), a range that encompasses the maximum corrosion rate of steel (Figure 4.8). As
shown in Figure 4.8, the steel casing of a district heating pipe experiences corrosion rates
several times that of domestic water pipes. The temperature of the casing varies with
burial depth, earth temperature, soil conditions, the carrier pipe temperature, and the pipe
insulation thickness. The casing must be strong enough and thick enough to withstand
expansion and contraction forces, earth load, and corrosion degradation. Cathodic protection systems (discussed later in this section) have often been used where the conditions
dictate that casing coating alone will not provide adequate protection against corrosion.
Corrosion is less of a problem with plastic casings; however, linear expansion and
decreasing yield strengths as the temperature rises are of major concern in high- and
medium-temperature systems. After detailed study (Marsh and Carnahan 1998; Marsh et
al. 1996), the federal government removed fiberglass-reinforced plastic (FRP) casing
from their preapproved drainable, dryable, testable (DDT) systems. This study found that
the likelihood of a FRP conduit leaking was four to five times higher than that of a steel
conduit. Surveys of users also found that problems were encountered with condensate
return piping of FRP (Phetteplace and Carbee 1992) as a result of not only issues with live
steam being discharged into the FRP piping but also the difficulty in completing the field
joint in the piping and differential settlement at various points within the buried piping
system. A later study (Marsh and Marshall 1998) found that the problems with FRP pipe
in condensate return service and associated maintenance to correct them far outweigh the
benefits derived from using FRP pipe and subsequently it was removed from acceptable
condensate piping materials for Army systems.
It is also important to note that crushing loads become very important in plastic casings of all types because the yield strength of these materials decreases rapidly as temperature rises. For instance, the yield strength of a polyvinyl chloride (PVC) casing is
approximately 20% of its room temperature value at 140F (60C). Plastic casings, how-

Figure 4.8 Corrosion rate in aggressive environment similar to mild steel casings in soil.
Courtesy of ASHRAE (2012, Chapter 12, Figure 18)

4.17

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

ever, may be used successfully in chilled-water systems and low-temperature hot-water


district heating systems when the limits of the materials are not exceeded.
All insulation must be kept dry for it to maintain its thermal insulating properties; the
exception is cellular glass in cold applications. Because underground systems may be
flooded several times during their design life, even on sites that are thought to be dry, a
reliable water intrusion removal system is necessary in the valve vaults. Two designs are
used to ensure that the insulation performs satisfactorily for the life of the system: the air
space system and the water spread limiting (WSL) system. In the air space system, an
annular air space, typically 1 in. (25 mm), between the pipe insulation and the casing
allows the insulation to be dried out if water enters. This system is often referred to as a
drainable, dryable, testable (DDT) system. One of the drawbacks of this type of system is
that when groundwater enters one section of conduit, it can easily pass through the annular airspace to other sections of conduit. In the second type, the WSL system, which has
no air space, the conduit is designed to keep water from entering the insulation. In the
event that water enters one section, a WSL system attempts to prevent the spread of water
to adjacent sections of conduit.
Air Space (Drainable, Dryable, Testable) Systems
Conduit Terminations
At the entrance to valve vaults, conduits must be sealed in order for the DDT concept
to be successful. In many cases where a DDT system fails from flooding, the breach of
the system occurs at the valve vault. There are two types of conduit terminations commonly used: a welded end plate (see Figure 4.7) or an end plate fitted with what is
referred to as a gland seal (see Figure 4.9). The gland seal allows pipe movement through
the end plate, thus allowing for expansion movements into the valve vault. The piping
design within the valve vault must account for this expansion. Gland seals unfortunately
have not proven to be airtight. In a recent study (Beitelman et al. 2011) that conducted air
pressure testing of 327 essentially new DDT conduit sections 100% of the sections of pip-

Figure 4.9 Gland seal components (right) and assembly (left).


Courtesy of Rovanco Piping Systems

4.18

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

ing with gland seals failed to pass the pressure test, even after efforts were made to tighten
the packing bolts. For this reason gland seals are not recommended. Their use can be
avoided by providing an anchor outside of each valve vault or building entry and accounting for all pipe expansions with loops and bends in the buried portion of the system.
At building entrances it is also necessary to provide an airtight termination of the conduit even where flooding conditions are not expected. If this is not done then it becomes
impossible to pressure test that entire segment of piping, defeating the DDT premise of
design.
Other important aspects of the conduit termination are the vents and drains. The
drains should be plugged with a galvanized iron or brass plug; countless DDT systems
have been destroyed by manholes flooding where the drain plugs were never installed or
inadvertently left out. The vents must be treated with equal respect; many DDT systems
have also been destroyed by improper termination of the vents. Leaving the plugs open or
providing a riser that terminates below the top of the manhole will both lead to failure in
the event of manhole flooding. Proper treatment of vents is to extend them to the top of
the manhole and terminate them with a gooseneck or u-bend, as shown in Figure 4.10.
Leaving the vents plugged is another common error that defeats their purpose of venting
and also acting as a telltale sign for water having entered the airspace. Furthermore, having both vents and drains plugged can create the dangerous and extremely high heat loss
circumstance of having a conduit flooded and pressurized either by a carrier pipe leak
and/or boiling water within the air space.
Field Joints
The field joints between successive lengths of conduit are the most critical aspect of
the system that must be executed on the construction site. This element can be the Achilles heel of an otherwise superior design and construction effort. This cannot be overemphasized. The method of joining the carrier piping and the quality control exercised in the

Figure 4.10 Proper termination for DDT conduit vents. The elevated frame has screened sides to
provide a degree of vault ventilation; the frame should have been galvanized.

4.19

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

process will largely determine the leakage that can be expected from the system. For insulated systems, the field joints in the carrier piping and the integrity of the casing splices at
those locations are very important and are details that must not be overlooked or underappreciated. If the conduit has external insulation on the exterior of the casing, the splice of
the exterior insulation jacket at the field joint must not be allowed to leak groundwater
into the insulation. When the field joint is completed, the casing of DDT systems should
be pressurized with air to 15 psig (104 kPa) for a period of at least 2 h. No air leakage
should be allowed. Some methods of insulating the casing joint provide a jacket for that
insulation allowing the joint to be pressure tested to ensure that the method of extending
the waterproof jacket over the joint is leak tight. These types of field joints are generally
preferable to designs that do not allow the jacket joint to be leak tested. When selecting an
insulated piping system, pay particular attention to the details that are provided regarding
the field joint. Normally the pre-insulated piping manufacturer will provide a kit with
the necessary materials including both the insulation and the jacketing materials. The
manufacturer of the system will normally also provide training to the construction crew.
The casing-to-casing seal and, if the conduit has external insulation, the jacket-to-jacket
seal are critical aspects of the field joint. Many methods of making these seals are used,
including heat-shrink materials and fusion of the jacket by electric heating welding in the
case of high-density polyethylene (HDPE) casing. For FRP casing, field lay-ups of
fiberglass are common; extra care must be exercised to ensure that the casings of the adjacent sections of pipe are properly prepared as well as clean and dry to ensure success of
the joint.
Because of the critical nature of the field joint on direct buried systems, a prudent
practice is to require training of the construction crew by representatives of the system
manufacturer/prefabricator. In addition, a contractual requirement will often be that a representative of the system manufacturer be on site to supervise some or all of the field joint
completion. For field joints that are pressure testable, the manufacturers field representative would normally supervise and attest to the success of that process. In addition, for
system constructions such as the DDT system, a contract requirement is normally that
successful pressure testing of the entire system is witnessed and attested to by a representative of the system manufacturer. However, as stated previously, the EOR ultimately
should assume responsibility for all aspects of the prefabricated system design and installation and thus it is in his/her best interest to independently verify the proper completion
of all critical phases of construction. It is important to note that while the system manufacturer/prefabricator may have the most knowledge of the system and the requirements
for its success, there is an inherent conflict of interest in manufacturer certification that
results from the contractor having purchased the product from the manufacturer. For
example, one potential conflict that results is the contractor withholding final payment
until the manufacturer provides certification.
The air space or DDT conduit system (Figures 4.7, 4.9, 4.10 4.11, 4.12, and 4.13)
should have insulation that can survive short-term flooding without damage. The conduit
manufacturer usually runs a boiling test with the insulation installed in a typical factory
casing. No U.S. standard has been approved for this boiling test; however, the federal
government and conduit manufacturers have been using a Federal Agency Committee
96-hour boiling water test for conduit insulation (see Appendix A). The insulation must
demonstrate that it can be dried with air flowing through an annular air space, and once
dried it must retain nearly new thermal insulating properties. Insulation manufacturers
also run this test to assure conduit manufacturers that their insulation can be used in DDT
conduit systems. Thermal properties of insulations tested under an ASHRAE Research
Project (Chyu et al. 1997a, 1997b, 1998a, 1998b) are discussed in Chapter 6. The testing

4.20

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

discussed did not follow the 96-hour boiling test protocol due to the range of insulation
and conditions tested, but it does provide much insight into the performance of insulation
during and after wetting.
DDT conduits with external insulation are being manufactured (Figure 4.12). This
type of conduit is essentially a conduit similar to Figure 4.11 or Figure 4.13 with an external layer of foam insulation and a jacket covering the foam insulation. The foam insulation is polyurethane, polyisocyanuate, or a mixture that may be intended to withstand a
temperature higher than 250F (120C). The jacket is usually HDPE or FRP. The jacket is
usually spliced between conduit sections with a heat-shrink product that is a processed
polyethylene. The intent of this arrangement is to eliminate the need for cathodic protection and require a smaller diameter steel casing because less pipe insulation is used on the
carrier pipe inside the steel casing. The exterior layer of foam insulation is not pressure
testable; therefore, groundwater that leaks through the exterior jacket cannot be detected
unless a leak detection system has been included in that insulation layer.

Figure 4.11 Conduit system with annular air space and single carrier pipe.
Courtesy of ASHRAE (2012, Chapter 12, Figure 19)

Figure 4.12 Conduit with external insulation.

4.21

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

There are some areas that require special attention when designing with this type of
conduit. Chyu et al. (1997a, 1997b, 1998a, 1998b) found that even though polyurethane
foam insulation is a closed-cell insulation, it absorbs water slowly and dries out very slowly.
This research suggests that the exterior jacket must be watertight. In an investigation of 313
conduits of this type (Beitelman et al. 2011), the air space was tested on all 313 conduits
between valve vaults and 10 conduits were excavated and their jackets were inspected.
These systems had both HDPE and FRP jackets, which were applied over the external foam
insulation that covered the steel casing. The HDPE was smooth and had a nicer appearance
than the FRP, but both appeared to be performing equally well. The shrink-wrap jacket
splice did not always bond well, especially at the wrap splice. It appeared that groundwater
could enter the foam insulation at the conduit splice. There were several locations where
cathodic protection testing indicated that cathodic protection was needed. The shrink-wrap
problem and the cathodic protection problem could have been caused by design error or
construction error.
The design of systems with foam exterior insulation is a compromise. The exterior
foam and the jacket are plastic materials that can be permanently damaged if either gets too
hot. If the pipe insulation is very thick, the steel casing must be larger in diameter than the
manufacturer would like. If the exterior foam insulation is very thick, the temperature at the
foam-to-steel-casing interface may be too hot for the foam material. A thermal analysis for
the dry earth condition is needed to determine what the temperature distribution is
through the conduit cross section (see Chapter 6). The EOR is required to investigate and
take responsibility for these potential problems. The exterior foam and jacket arrangement
can also be used on the conduit system with two carrier pipes and annular air space shown
in Figure 4.13. Note that this type of system is not recommended for steam and condensate as the condensate line will nearly always fail before the steam line and in this type of
system complete replacement will be required under those circumstances. The economics
of two pipes within a single conduit are most favorable for small pipe sizes.
Fiberglass and mineral wool insulation fails when flooded because the bonding
agents, called binders, fail to hold the principal insulation material fiber in the desired
shape. Some binders dissolve completely. When these fiber-type insulations fill with
water they conduct up to 50 times more heat then when dry (Chyu et al. 1997a, 1997b,

Figure 4.13 Conduit system with two carrier pipes and annular air space
(not recommended for steam and condensate; see text).
Courtesy of ASHRAE (2012, Chapter 12, Figure 20)

4.22

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

1998a, 1998b); see Chapter 6 for further information on the thermal performance of wet
pipe insulation. With foam insulation it is the parent material that fails.
Water Spread Limiting System
The water spread limiting (WSL) system, shown in Figure 4.14, encloses the insulation in an envelope that is intended to not allow water to contact the insulation. Some systems rely on a watertight field joint between joining casing sections to extend the
envelope to a distant envelope termination point where the casing is sealed to the carrier
pipe. Other systems form the waterproof insulation envelope in each individual prefabricated conduit section using the casing and the carrier pipe to form part of the envelope
and a waterproof bulkhead to seal the casing to the carrier pipe. The typical insulation is
polyurethane foam, which will be destroyed if excess water infiltration occurs, especially
in the presence of temperatures above atmospheric boiling. Polyurethane foam is limited
to a temperature of about 250F (120C) for a service life of 30 years or more. However,
it is important to note that the material properties for polyurethane foam can vary widely
depending on the formulation, blowing agent, blowing process, and process quality control; not all polyurethane foams are equal nor can that assumption be safely made.
The most successful WSL systems have been developed in Europe and are typically
used in low-temperature water (LTW) applications. Extensive research and development
and design effort have gone into the creation of the European LTW systems as well as
many years of experience and continuous product improvement. The European type of
WSL system is largely possible because of the smaller temperature differential and
because the thinner carrier pipe wall creates lower expansion forces due to its reduced
cross-sectional area. Some of the methods used in the installation of these systems require
preheating of the pipe or other methods of dealing with expansion and the forces it generates. The design basis for these systems is well established and is documented in
Randlv (1997) and EN 13941 (CEN 2010). These systems, which are available worldwide including in the United States, meet European Standard EN 253 (CEN 2009a) with
regard to all major construction details. Standards also have been established for fittings
(EN 448 [CEN 2009b]), pre-insulated valves (EN 488 [CEN 2011]), field joint assemblies (EN 489 [CEN 2009c]), and the design of the systems (EN 13941 [CEN 2010]).
With these systems, the carrier pipe, insulation, and casing are bonded together to form a

Figure 4.14 WSL conduit system with single carrier pipe and no air space.
Courtesy of ASHRAE (2012, Chapter 12, Figure 21)

4.23

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

single unit; special precautions are taken in the choice and preparation of the materials to
ensure the strength of the bonds. A significant portion of the forces caused by thermal
expansion are passed as shear forces to the mating components and ultimately to the soil;
thus, the system is restrained to a degree by the surrounding soil and hence the amount of
expansion is significantly reduced. The amount of restraining force that may be generated by the soil is limited, so systems that generate more force result in larger unrestrained movements. Thus, a system with thicker piping and/or higher temperature
differences will have higher unrestrained movements at the changes in direction for
equivalent lengths of piping. While it is practical to absorb some degree of movement at
changes in direction by use of cushioning material, the amount is limited, and where
larger movements are attempted the deformation of the cushion and/or surrounding soil
will generate higher restraining forces to the lateral movement. This in turn requires that
the piping system, notably the insulation and jacket as well as the field closures, must be
able to withstand higher compressive forces and resultant deformations. For systems
meeting the EN 253 (CEN 2009a) requirements, prescriptive design methods or, in more
complicated situations, well-detailed calculations are used to ensure that the system will
be able to withstand all of the prevailing forces as a system. While it is possible to design
such systems using thicker pipe such as Schedule 40/standard weight piping, absent the
strict adherence of the procedures for design of such systems as outlined in Randlv
(1997) and EN 13941 (CEN 2010), such systems are not recommended. In addition to
using the methods of Randlv and EN 13941 for design to ensure success, the components must also meet the strict requirements of EN 253 (CEN 2009a); see Table 4.4 for a
partial list of those requirements and note that many of the tests are of the assembly, not
simply of the materials used. This is an essential requirement for success; otherwise there
is no assurance that the system will perform as intended.
As noted, the European LTW systems meeting the EN 253 (CEN 2009a) requirements
and companion standards have proven to be very successful. However, attempts at similar
approaches for higher-temperature use have not been as successful. While a detailed disTable 4.4 A Partial Compilation of the Testing and Quality Control Requirements
of EN 253 (CEN 2009a)
System Component or Quality Control Measure
Minimum pipe wall thickness requirement (thicker walls are allowed, but they must be designed to EN 13941 [CEN 2010])
Pipe wall cleaning required
Jacket raw material and quality control
Jacket material thickness dependent on pipe size
Jacket material minimum elongation
Polyurethane foam material documentation
Maximum polyurethane foam voids
Polyurethane foam compressive strength
Polyurethane foam thermal conductivity
Polyurethane foam thermal conductivity, thermally aged
Polyurethane foam cell gas composition
Polyurethane foam cell size
Centerline deviation pipe/jacket, function of pipe diameter
Assembly axial shear or tangential shear strength
Assembly axial shear or tangential shear strength, thermally aged
Assembly axial shear or tangential shear strength, at operating temperature

4.24

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

cussion of the design basis for the EN 253 systems is beyond the scope of this book (refer
to Randlv [1997] or EN 13941 [CEN 2010]), just simply looking at the magnitude of the
expansion forces involved provides some insight into why extrapolation of the EN 253
(CEN 2009a) systems to higher temperatures and/or thicker piping has not been entirely
successful. Consider, for example, a high-temperature water (HTW) system using a 4 in.
(100 mm) nominal Schedule 40 steel pipe installed at 70F (21C) and operating at 400F
(204C). If completely restrained, a section of straight piping in such a system would generate approximately three times greater expansion force than the equivalent 4 in. (100 mm)
nominal EN 253 pipe operating at 250F (121C). Slightly less than half of this increase in
force comes about due to the greater wall thickness of Schedule 40 pipe versus the pipe
used for EN 253, which has a wall thickness that is approximately 60% of the Schedule 40
pipe for the diameters of this example. The remaining increase in force is due to the higher
temperature difference, 330F (183C) versus 180F (100C). In order for the high-temperature system to be successful, the bonding of the casing pipe to the insulation as well as the
bonding of the carrier pipe to the insulation would need to be as strong as what is required
for the EN 253 system, and in addition piping runs between changes in direction would
need to be much shorter, otherwise movement at those changes in direction would be much
greater, resulting in the issues discussed previously. Both the material and bonding
strengths would need to be able to meet the EN 253 criteria at higher temperatures, where
the strength properties likely will not be as favorable.
The European EN 253 (CEN 2009a) systems require extensive quality control procedures to be followed, and in the case of the polyurethane foam insulation, testing that
establishes the long-term (30-year) service life is also required (see Table 4.4). To the
knowledge of the authors, no such testing has been conducted on foams that are being
marketed as capable of service at temperatures up to 400F (204C) or more, let alone
testing of complete systems constructed with such foams.
Another piping system alternative that is available for low-temperature applications is
a flexible piping system. These are only available in smaller carrier pipe diameters, up to
approximately 4 in. (100 mm) nominal and are used primarily for connections between
the main distribution system and the consumer, although small district heating systems
could consist entirely of flexible piping. Carrier pipe materials that have been used
include corrugated steel and stainless steel, thin-walled steel, copper, aluminum, and
cross-linked high-density polyethylene (sometimes referred to as PEX). The PEX pipes
must be equipped with a diffusion barrier to prevent the diffusion of oxygen into the hot
water. The flexible piping is normally delivered on rolls in lengths up to approximately
330 ft (100 m). Also available are flexible piping systems where both the supply and
return piping are encased in the same jacket. The advantage of these flexible piping systems is the ease of installation resulting primarily from fewer field joints. In most
instances no expansion provisions are needed.
Conduit Design Conditions
Three design conditions must be addressed to have reasonable assurance that the system selected will have a satisfactory service life: the wet soil or high heat loss condition,
the dry soil or high-temperature condition, and the nominal or average condition. Because
the heat distribution system lasts many years, each of these three conditions will occur
sometime during the life of the system. Of course the nominal condition occurs often.
Methods for calculating the heat loss and temperatures of the materials for each of these
conditions are contained in Chapter 6.
Maximum heat loss occurs when the soil is wettest and the conduit is buried shallow
(minimum burial depth), usually with about 2 ft (0.61 m) of earth cover. This condition

4.25

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

represents the highest gross heat transfer and is used to size the distribution piping and
equipment in the district heating plant. For heating piping, since the casing is coldest during start-up, the relative movement (axial expansion) of the carrier pipe with respect to
the casing may be at its maximum during this condition.
A dry soil condition may occur in some circumstances and is of particular concern
when the conduit is buried deep. The soil plays a more significant role in the heat transfer
since it is quite thick when compared to the pipe insulation thickness. When wet, the soil
thermal conductivity is about 50 times higher than the pipe insulation; however, when dry,
the earth can be as little as three times higher. There is much less temperature drop across
the conduit insulation cross section, and much of the temperature drop is from the outside
of the conduit in the earth. The highest temperature of the insulation, casing, and casing
coating occur during this condition. Paradoxically, the minimum heat loss occurs during
this condition because the thick soil is acting as a good insulator. This condition is used to
select temperature-sensitive materials and to design for casing expansion. The relative
movement of the carrier pipe with respect to the casing may be at its minimum during this
condition because the casing expands linearly almost as much as the carrier pipe; however, if the casing is not restrained, its movement with respect to the soil will be maximum. If restrained, the casing axial stresses and axial forces will be highest and the
casing allowable stresses will be lowest because of the high casing material temperature.
Figure 4.15 shows the impact of burial depth on casing temperature as a function of
soil thermal conductivity for a typical system. Figure 4.16 shows the impact of insulation
thickness on casing temperature, again as a function of soil thermal conductivity. Analysis of Figures 4.15 and 4.16 suggest some design solutions that could lower the effects of
the dry soil condition. Possible solutions would be to lower the carrier pipe temperature,
use thicker carrier pipe insulation, provide a device to keep the soil wet, or minimize the

Figure 4.15 Effect of burial depth on casing temperature.


Courtesy of ASHRAE (2012, Chapter 12, Figure 22A)

4.26

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

burial depth. However, if these solutions are not feasible or cost-effective, a different type
of material or an alternate system should be considered.
While it is possible that the soil will never dry out, given the variability of climate in
most areas, it is likely that a drought will occur during the life of the system. The higher
temperatures of the heat distribution system will also have a tendency to cause moisture
in the soil to migrate away from the soil adjacent to the conduit. Only one very dry condition can cause permanent damage to the insulation and plastic components if the soil thermal conductivity drops below the nominal design value. On-site measurement of the
driest soil condition likely to cause damage is not normally feasible. As a result, the
designer is left with the conservative choice of using the lowest soil thermal conductivity
from (Table 6.1 in Chapter 6) to calculate the highest temperature to be used to select
materials.
The nominal or average condition occurs when the soil is at its average water content
and the conduit is buried at the average depth. This condition is used to estimate the
yearly energy consumption due to heat loss to the soil (see Table 6.1 in Chapter 6 for soil
thermal conductivity). This condition is used for economic analysis.

GEOTECHNICAL TRENCHING AND BACKFILLING


There are stringent burial requirements for underground district heating distribution
systems as compared to most other utilities associated with buildings. The elevation of the
trench bottom must not have slope reversals between valve vaults and building entry locations. These piping systems normally have coatings or jackets needed for corrosion protection or insulation protection that must not be damaged by rocks, debris, or construction
equipment. Thus, proper burial conditions must be established for the district heating distribution system to achieve its design life. Requirements will vary and manufacturers of
the piping system should be able to provide guidance specific to their system. It is recom-

Figure 4.16 Effect of insulation thickness on casing temperature.


Courtesy of ASHRAE (2012, Chapter 12, Figure 22B)

4.27

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

mended that the services of a licensed geotechnical engineer familiar with the local conditions be engaged to conduct a site survey in advance of construction and recommend
any soil testing required as well as develop the specifications for excavation and backfill.
The geotechnical engineer should also be responsible for design of any thrust blocks or
anchors that are needed based on forces provided to her/him by the EOR.
To allow for proper completion of the field joints/closures, the trench should be sized
such that at least 8 in. (200 mm) is allowed between the jackets of the insulated pipes. In
general trenches must be over-excavated by a minimum of 4 in. (100 mm) to remove any
unyielding material; over-excavation may need to be greater at the locations of the field
joints depending on the type of system being installed and the method of construction.
The over-excavation is generally filled with a select backfill material; normally this is a
sandy, noncohesive material free of any stones greater than 3/4 in. (19 mm). If unstable
materials are encountered in the excavation they should be removed and properly backfilled and compacted. The select backfill in the trench bottom should be prepared to
achieve the minimum slope for the carrier pipe of 1 in. in 20 ft (25.4 mm in 6.2 m) or
0.4% and compacted to 95% of laboratory maximum density per ASTM D698 (2012a).
Some of the methods of carrier pipe joining, such as welding, will require a working area
around the entire circumference of the field joint. One method of achieving this is to overexcavate under the pipe and potentially even at the sides of the trench at the locations of
the field joints. If this is done, extra care must be taken to fully compact the additional
backfill material under the field joint area. Another method used to provide working
clearance for making the field joint is to block the piping up off the bottom of the trench
during that process. When this method is used care must be taken to block the pipe sufficiently to achieve proper alignment for joining and to emulate the pipe as it will ultimately lie on the sloped trench bottom. Once the field joints have been completed, the
blocking should be removed and the piping carefully and uniformly placed in the trench
bottom that has been prepared as described. The blocking should not be left in place, as
this creates point loading on the piping and may contribute to differential settlement as
well. In some situations, when welded steel piping is being used, for example, it may be
possible to join two sections of piping together adjacent to the trench and then lift the
assembly in as a unit and thus reduce the work required in the trench.
After the piping is placed in the trench and all field joints and pressure tests have been
completed, immediately preceding the backfilling, the elevation of the top invert of the
pipe/jacket should be taken at each pipe section midpoint and field joint. These elevations
should be recorded and subsequently transferred to the as-built or record drawings. Backfill of the piping should then be accomplished in layers of no more than 6 in. (150 mm)
with the same select backfill material used for the pipe bedding. The select backfill
should be extended to approximately 12 in. (300 mm) above the top of the pipe or jacketing. Compaction of this backfill material should also be to 95% of laboratory maximum
density per ASTM D698 (2012a). Care should be exercised to ensure that the backfill
adequately fills the void created under the pipe and between the supply and return pipes.
Care should also be exercised not to damage any pipe coating or insulation jacketing
material; if any such damage does occur it should be repaired per the pipe system manufacturers field repair instructions. Final backfill of the remainder of the trench should be
accomplished using the native soil but removing any stones greater than 3 in. (75 mm),
compacted in layers of no more than 6 in. (150 mm). This final backfill should be compacted to 95% of laboratory maximum density per ASTM D698 for noncohesive soils or
90% of laboratory maximum density per ASTM D698 for cohesive soils. Note that it is
not advisable to complete the final backfilling process with other than native soil as its
permeability may be much different than the native material. For example, using a perme-

4.28

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

able backfill material in a native soil that is impermeable is essentially placing your district heating system in a drainage ditch for surface water.

PIPING MATERIALS AND STANDARDS


Supply Pipes for Steam and Hot Water
Adequate temperature and pressure ratings for the intended service should be specified for all piping. All piping, fittings, and accessories should be in accordance with
ASME B31.1 (2012) or with local requirements if more stringent. For steam and hot
water, all joints should be welded and pipe should conform to either ASTM A53/A53M
(2012b) seamless or electric resistance welded (ERW), Grade B, or ASTM A106/A106M
(2011b) seamless, Grade B. Care should be taken to exclude ASTM A53, Type F, because
of its lower allowable stress and because of the method by which its axial seams are manufactured. Mechanical joints of any type are not recommended for steam or hot water.
Pipe wall thickness is determined by the maximum operating temperature and pressure.
In the United States, most piping for steam and hot water is Schedule 40 for 10 in.
(254 mm) nominal pipe size (NPS) and below and standard weight for 12 in. (305 mm)
NPS and above.
Many European LTW systems have piping with a wall thickness similar to Schedule
10. Due to reduced piping material in these systems, they are not only less expensive but
as discussed previously they also develop reduced expansion forces and thus require simpler methods of expansion compensation, a point to be considered when making a selection between a high-temperature and a low-temperature system. Welding pipes with
thinner walls requires extra care and may require additional inspection. The extensive
U.S. experience with low-temperature systems has indicated that welding of thin-wall carbon steel piping (in order to pass the X-ray test) has to be performed in accordance with
special procedures specified by qualified metallurgical engineers and performed by
trained certified welders. Also, extra care must be taken to avoid internal and external corrosion because the thinner wall provides a much lower corrosion allowance.

Condensate Return Pipes


Condensate pipes require special consideration because condensate is much more
corrosive than steam. This corrosive nature is caused by the oxygen and carbon dioxide
that the condensate accumulates. The usual method used to compensate for the corrosiveness of condensate is to select steel pipe that is thicker than the steam pipe, extra heavy
or Schedule 80. For highly corrosive condensate, stainless steel and/or other corrosionresistant materials should be considered. Materials that are corrosion resistant in air may
not be corrosion resistant when exposed to condensate; therefore, a material with good
experience handling condensate should be selected. FRP pipe used for condensate return
has not performed well. Failed steam traps, pipe resin solubility, deterioration at elevated
temperatures, differential settlement, and thermal expansion are thought to be the causes
of premature failure.

PIPE EXPANSION AND FLEXIBILITY


The following material on expansion and pipe flexibility has been extracted in large
part from Chapter 45, Pipes, Tubes, and Fittings, of the 2008 ASHRAE Handbook
HVAC Systems and Equipment.
Temperature changes cause dimensional changes in all materials. Tables 4.5a and 4.5b
show the expansion for piping materials commonly used in heat distribution systems. For
systems operating at high temperatures, such as steam and hot water, the rate of expansion
is high and significant movements can occur in short runs of piping. Even though rates of

4.29

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

expansion may be low for systems operating in the range of low-temperature water, they
can cause large movements in long runs of piping, which are common in distribution systems. Therefore, in addition to design requirements for pressure, weight, and other loads,
district heating piping systems must accommodate thermal and other movements to prevent the following:
Failure of pipe and supports from overstress and fatigue
Leakage of jackets at field joints
Detrimental forces and stresses in connected equipment
Table 4.5a Thermal Expansion of Metal Pipe from 0F
Linear Thermal Expansion, in./100 ft
Saturated Steam
Pressure,
psig

Temperature,
F

Carbon
Steel

Type 304
Stainless Steel

Copper

212

1.62

2.38

2.43

2.5

220

1.69

2.48

2.52

10.3

240

1.85

2.71

2.76

20.7

260

2.02

2.94

2.99

34.6

280

2.18

3.17

3.22

52.3

300

2.35

3.40

3.46

75.0

320

2.53

3.64

3.70

103.3

340

2.70

3.88

3.94

138.3

360

2.88

4.11

4.18

181.1

380

3.05

4.35

4.42

232.6

400

3.23

4.59

4.87

666.1

500

4.15

5.80

5.91

1528

600

5.13

7.03

7.18

Adapted from ASHRAE HandbookHVAC Systems and Equipment, Chapter 45, Table 10 (2008)

Table 4.5b Thermal Expansion of Metal Pipe from 18C


Linear Thermal Expansion, mm/m

Saturated Steam
Pressure,
kPa (gage)

Temperature,
C

Carbon
Steel

Type 304
Stainless Steel

Copper

100

1.35

1.98

2.03

17.2

104

1.41

2.07

2.10

71.0

116

1.54

2.26

2.30

142.7

127

1.68

2.45

2.49

238.6

138

1.82

2.64

2.68

360.6

149

1.96

2.83

2.88

517.1

160

2.11

3.03

3.08

712.3

171

2.25

3.23

3.28

953.6

182

2.40

3.43

3.48

1249

193

2.54

3.63

3.68

1604

204

2.69

3.83

4.06

9039

304

4.11

5.65

5.77

Adapted from ASHRAE HandbookHVAC Systems and Equipment, Chapter 45, Table 10 (2008)

4.30

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

An unrestrained pipe operates at the lowest overall stress level. Anchors and restraints
are needed to support pipe weight and to protect branch connections. The anchor forces
and the bowing of pipe anchored at both ends are generally too large to be acceptable, so
general practice is to never anchor a straight run of steel pipe at both ends. Piping must
be allowed to expand and contract through thermal changes. Ample flexibility can be
attained by designing pipe bends and loops or by including supplemental devices, such as
expansion joints. Special care must be taken to account for linear (axial) expansion at
branch connections.
For buried heat distribution systems, anchors are normally located in the soil but may
be located within a building or manhole provided the structure of the building/manhole
has been designed to accommodate the forces developed.

Pipe Bends and Loops


Detailed stress analysis requires involved mathematical analysis and is generally performed by computer programs. However, such involved analysis is not required for simple systems because the piping arrangements and temperature ranges at which they
operate are simple to analyze.
L Bends
The guided cantilever beam method of evaluating L bends can be used to design
L bends, Z bends, pipe loops, branch take-off connections, and some more complicated
piping configurations.
Equation 4.1 may be used to calculate the length of leg BC shown in Figure 4.17
needed to accommodate thermal expansion or contraction of leg AB for a guided cantilever beam.
L =

3DE
----------------------------------------- 144 in. 2 ft 2 S A

(I-P)

(4.1a)

3DE
--------------SA

(SI)

(4.1b)

L =

where
L = length of leg BC required to accommodate thermal expansion of long leg AB
(Figure 4.17), ft (mm)
= thermal expansion or contraction of leg AB (Figure 4.17), in. (mm)
D = actual pipe outside diameter, in. (mm)
E = modulus of elasticity, psi (kPa)
SA = allowable stress range, psi (kPa)
For the commonly used ASTM A53 Grade B seamless or ERW pipe (ASTM 2012b),
an allowable stress SA of 22,500 psi (155 MPa) can be used without overstressing the
pipe. However, this can result in very high-end reactions and anchor forces, especially
with large-diameter pipe. Designing to a stress range SA of 15,000 psi (103 MPa) and
assuming E = 27.9 106 psi (193 GPa), Equation 4.1 reduces to Equation 4.2, which provides reasonably low-end reactions without requiring too much extra pipe. In addition,
Equation 4.2 may be used with ASTM A53 butt-welded pipe and ASTM B88 drawn copper tubing (2012c).

4.31

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 4.17 Guided cantilever beam.


Courtesy of ASHRAE (2012, Chapter 46, Figure 1)

Thus, for ASTM A53 (2012b) continuous (butt-) welded, seamless, and ERW pipe,
and ASTM B88 drawn copper tubing (2012c),
L = 6.225 D

(I-P)

(4.2a)

L = 75 D

(SI)

(4.2b)

The guided cantilever method of designing L bends assumes no restraints; therefore,


care must be taken in supporting the pipe. For horizontal L bends, it is usually necessary
to place a support near point B shown in Figure 4.17, and any supports between points A
and C must provide minimal resistance to piping movement; this is done by using slide
plates or hanger rods of ample length, with hanger components selected to allow for
swing no greater than 4.
For L bends containing both vertical and horizontal legs, any supports on the horizontal leg must be spring hangers designed to support the full weight of pipe at normal operating temperature with a maximum load variation of 25%.
The force developed in an L bend that must be sustained by anchors or connected
equipment is determined by the following equation:

where
F =
Ec =
I =
L =
=

12E c I
F = ------------------------------------------ 1728 in. 3 ft 3 L 3

(I-P)

(4.3a)

12E c I
F = -----------------10 6 L 3

(SI)

(4.3b)

force, lb (kN)
modulus of elasticity, psi (kPa)
moment of inertia, in.4 (mm4)
length of offset leg, ft (mm)
deflection of offset leg, in. (mm)

In lieu of using Equation 4.3, for L bends designed in accordance with Equation 4.2
for 1 in. (25 mm) or more of offset, a conservative estimation of force is 500 lb (90 N) per

4.32

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

Figure 4.18 Z bend in pipe.


Courtesy of ASHRAE (2012, Chapter 46, Figure 2)

diameter inch (millimetre) (e.g., a 3 in. [75 mm] pipe would develop 1500 lb [6750 N] of
force).
Z Bends
A Z bend, as shown in Figure 4.18, is very effective for accommodating pipe movements. A simple and conservative method of sizing Z bends is to design the offset leg to
be 65% of the values used for an L bend in Equation 4.1, which results in
L = 4 D

(I-P)

(4.4a)

L = 48.7 D

(SI)

(4.4b)

where
L = length of offset leg, ft (mm)
= anchor-to-anchor expansion, in. (mm)
D = pipe outside diameter, in. (mm)
The force developed in a Z bend can be calculated with acceptable accuracy as follows:
F = C 1 D L 2
where
C1 =
F =
D =
L =
=

(4.5)

4000 lb/in. (101 kN/mm)


force, lb (kN)
pipe outside diameter, in. (mm)
length of offset leg, ft (mm)
anchor-to-anchor expansion, in. (mm)

U Bends and Pipe Loops


Pipe loops or U bends are commonly used in long runs of piping. A simple method of
designing pipe loops is to calculate the anchor-to-anchor expansion and, using
Equation 4.1, determine the length L necessary to accommodate this movement. The pipe
loop dimensions can then be determined using W = L/5 and H = 2W. Note that guides must

4.33

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

be spaced no closer than twice the height of the loop, and piping between guides must be
supported, as described in the section on L Bends, when the length of pipe between guides
exceeds the maximum allowable hanger spacing for the size of pipe.
Table 4.6a lists pipe loop dimensions for pipe sizes 1 through 24 in. and anchor-toanchor expansion (contraction) of 2 through 12 in., and Table 4.6b lists pipe loop dimensions for pipe sizes 25 through 600 mm and anchor-to-anchor expansion (contraction) of
50 through 300 mm.
No simple method has been developed to calculate pipe loop force; however, it is generally low. A conservative estimate is 200 lb per diameter inch (35 N per diameter millimetre) (e.g., a 2 in. [50 mm] pipe will develop 400 lb [1.75 kN] of force and a 12 in. [300
mm] pipe will develop 2400 lb [10.5 kN] of force).

Cold Springing of Pipe


Cold springing or cold positioning of pipe consists of offsetting or springing the pipe in
a direction opposite the expected movement. Cold springing is not recommended for most
heat distribution piping because ASME B31.1 (2012) does not allow the design stress to be
reduced due to the reduced stresses attributed to cold springing. Furthermore, cold springing
does not allow designing a pipe bend or loop for twice the calculated movement. For examTable 4.6a Pipe Loop Design for ASTM A53 (2012b) Grade B Carbon Steel Pipe Through 400F

Anchor-to-Anchor Expansion, in.

Pipe
Size,
in.

4
H

6
H

8
H

10
H

12
H

3.5

4.5

10

10

5.5

11

12

14

3.5

10

12

6.5

13

7.5

15

16

5.5

11

6.5

13

7.5

15

8.5

17

18

10

6.5

13

16

18

10

20

11

22

5.5

11

7.5

15

18

10.5

21

12

24

13

26

10

12

8.5

17

10

20

11.5

23

13

26

14

28

12

6.5

13

18

11

22

12.5

25

14

28

15.5

31

14

14

9.5

19

11.5

23

13

26

15

30

16

32

16

7.5

15

10

20

12.5

25

14

28

16

32

17.5

35

18

16

11

22

13

26

15

30

17

34

18.5

37

20

8.5

17

11.5

23

14

28

16

32

18

36

19.5

39

18

12.5

25

14.5

29

17.5

35

19.5

39

21

42

24
Notes:

W and H dimensions are feet.


L is determined from Equation 4.1a. W = L/5, H = 2W, 2H + W = L
Approximate force to deflect loop = 200 lb/diameter in. (e.g., 8 in. pipe creates 1600 lb of force).

Adapted from ASHRAE HandbookHVAC Systems and Equipment, Chapter 45, Table 11 (2008)

4.34

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

ple, if a particular L bend can accommodate 3 in. (75 mm) of movement from a neutral position, cold springing does not allow the L bend to accommodate 6 in. (150 mm) of
movement.

Computer-Aided Design
When significant portions of a distribution system are being designed, computeraided design (CAD) software is normally used. When analyzing the system using CAD,
normally stresses of the entire run (valve vault to valve vault) are calculated. The run may
include changes in direction along with expansion loops. By analyzing the entire system
segment, the forces at intermediate anchors may balance and the forces that must be dealt
with may be reduced. Therefore, it may be possible to eliminate some intermediate
anchors and thus reduce the cost of the installation. This only applies to systems without
expansion joint devices, which have their own anchor and guide requirements.

Analyzing Existing Piping Configurations


Piping is best analyzed using a computer stress analysis program because these provide all pertinent data including stress, movements, and loads. Program vendors can perTable 4.6b Pipe Loop Design for ASTM A53 (2012b) Grade B Carbon Steel Pipe Through 200C

Pipe
Nominal
Outside
Diameter,
mm

Anchor-to-Anchor Expansion, mm
50

100

150

200

250
H

300
H

25

0.6

1.2

0.9

1.8

1.1

2.1

1.2

2.4

1.4

2.7

1.5

3.0

50

0.9

1.8

1.2

2.4

1.5

3.0

1.7

3.4

1.8

3.7

2.1

4.3

80

1.1

2.1

1.5

3.0

1.8

3.7

2.0

4.0

2.3

4.6

2.4

4.9

100

1.2

2.4

1.7

3.4

2.0

4.0

2.3

4.6

2.6

5.2

2.7

5.5

150

1.5

3.0

2.0

4.0

2.4

4.9

2.7

5.5

3.0

6.1

3.4

6.7

200

1.7

3.4

2.3

4.6

2.7

5.5

3.2

6.4

3.7

7.3

4.0

7.9

250

1.8

3.7

2.6

5.2

3.0

6.1

3.5

7.0

4.0

7.9

4.3

8.5

300

2.0

4.0

2.7

5.5

3.4

6.7

3.8

7.6

4.3

8.5

4.7

9.4

350

2.1

4.3

2.9

5.8

3.5

7.0

4.0

7.9

4.6

9.1

4.9

9.8

400

2.3

4.6

3.0

6.1

3.8

7.6

4.3

8.5

4.9

9.8

5.3

10.7

450

2.4

4.9

3.4

6.7

4.0

7.9

4.6

9.1

5.2

10.4

5.6

11.3

500

2.6

5.2

3.5

7.0

4.3

8.5

4.9

9.8

5.5

11.0

5.9

11.9

2.7

5.5

3.8

7.6

4.4

8.8

5.3

10.7

5.9

11.9

6.4

12.8

600
Notes:

W and H dimensions are metres.


L is determined from Equation 4.1b. W = L/5, H = 2W, 2H + W = L
Approximate force to deflect loop = 35 N/mm pipe diameter (e.g., 200 mm pipe creates 7600 N of force).

Adapted from ASHRAE HandbookHVAC Systems and Equipment, Chapter 45, Table 11 (2008)

4.35

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 4.19 Multiplane pipe system.


Courtesy of ASHRAE (2012, Chapter 46, Figure 3)

form such analysis if programs are not available in-house. However, many situations do
not require such detailed analysis. A simple yet satisfactory method for single and multiplane systems is to divide the system with real or hypothetical anchors into a number of
single-plane units, as shown in Figure 4.19, which can be evaluated as L and Z bends.

Expansion Joints and Expansion Compensating Devices


Mechanical expansion joints are discouraged because of maintenance, reliability, and
leakage problems; however, there are conditions where they must be used. Although the
inherent flexibility of the piping should be used to the maximum extent possible, expansion joints must be used where movements are too large to accommodate with pipe bends
or loops or where insufficient room exists to construct a loop of adequate size. Typical situations are tunnel piping and risers in high-rise buildings, especially for steam and hotwater pipes where large thermal movements are involved.
Packed and packless expansion joints and expansion compensating devices are used
to accommodate movement, either axially or laterally.
In the axial method of accommodating movement, the expansion joint is installed
between anchors in a straight line segment and accommodates axial motion only. This
method has high anchor loads, primarily because of pressure thrust. It requires careful
guiding, but expansion joints can be spaced conveniently to limit movement of branch
connections. The axial method finds widest application for long runs without natural offsets, such as tunnel and underground piping and risers in tall buildings.
The lateral or offset method requires the device to be installed in a leg perpendicular
to the expected movement and accommodates lateral movement only. This method generally has low anchor forces and minimal guide requirements. It finds widest application in
lines with natural offsets, especially where there are few or no branch connections.
In all cases, expansion joints should be installed, anchored, and guided in accordance
with expansion joint manufacturers recommendations.
Packed Expansion Joints
Packed expansion joints depend on slipping or sliding surfaces to accommodate the
movement and require some type of seals or packing to seal the surfaces. Most such
devices require some maintenance but are not subject to catastrophic failure. Further, with
most packed expansion joint devices, any leaks that develop can be repacked under full
line pressure without shutting down the system.
Packed expansion joints are preferred where long-term system reliability is of prime
importance (using types that can be repacked under full line pressure) and where major

4.36

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

leaks can be life threatening or extremely costly. Typical applications are risers, tunnels,
underground pipe, and distribution piping systems.
There are two types of packed expansion joints: packed slip expansion joints and
flexible ball joints.
Packed Slip Expansion Joints
Packed slip expansion joints are telescoping devices designed to accommodate axial
movement only. Some sort of packing seals the sliding surfaces. The original packed slip
expansion joint design used multiple layers of braided compression packing, similar to
the stuffing box commonly used with valves and pumps; this arrangement requires shutting off and draining the system for maintenance and repair. Advances in design and
packing technology have eliminated these problems, and most current packed slip joints
use self-lubricating semi-plastic packing, which can be injected under full line pressure
without shutting off the system. (Many manufacturers use asbestos-based packing, unless
requested otherwise. Asbestos-free packing, such as flake graphite, is available and,
although more expensive, should be specified in lieu of products containing asbestos.)
Standard packed slip expansion joints are constructed of carbon steel with weld or
flange ends in sizes from 1.5 to 36 in. (40 to 910 mm) for pressures up to 300 psig
(2.1 MPa) and temperatures up to 800F (425C). Larger sizes and higher-temperature,
higher-pressure designs are available. Standard single joints are generally designed for
4, 8, or 12 in. (100, 200, or 300 mm) axial traverse; double joints with an intermediate
anchor base can accommodate twice these movements. Special designs for greater
movements are available.
Flexible Ball Joints
Flexible ball joints are used in pairs to accommodate lateral or offset movement and
must be installed in a leg perpendicular to the expected movement. The original flexible
ball joint design incorporated only inner and outer containment seals that could not be
serviced or replaced without removing the ball joint from the system. The packing technology of the packed slip expansion joint has been incorporated into the flexible ball joint
design; now, packed flexible ball joints have self-lubricating semi-plastic packing that can
be injected under full line pressure without shutting off the system.
Standard flexible ball joints are available in sizes 1 1/4 through 30 in. (32 through
760 mm) with threaded (1 1/4 to 2 in. [32 to 50 mm]), weld, and flange ends for pressures
to 300 psig (2.1 MPa) and temperatures to 750F (400C). Flexible ball joints are also
available in larger sizes and for higher temperature and pressure ranges.
Packless Expansion Joints
Packless expansion joints depend upon the flexing or distortion of the sealing element
to accommodate movement. They generally do not require any maintenance, but maintenance or repair is not usually possible. If a leak occurs, the system must be shut off and
drained and the entire device must be replaced. Further, catastrophic failure of the sealing
element can occur and, although likelihood of such failure is remote, it must be considered in certain design situations.
Packless expansion joints are generally used where even small leaks cannot be tolerated (e.g., for gas and toxic chemicals), where temperature limitations preclude the use of
packed expansion joints, and for very large-diameter piping where packed expansion
joints cannot be constructed or the cost would be excessive.
Packless expansion joints suitable for use on district heating applications are metal
bellows expansion type.

4.37

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Metal Bellows Expansion Joints


Metal bellows expansion joints have a thin-wall convoluted section that accommodates movement by bending or flexing. The bellows material is generally Type 304, 316,
or 321 stainless steel, but other materials are commonly used to satisfy service conditions. Small-diameter expansion joints in sizes 3/4 through 3 in. (20 through 80 mm) are
generally called expansion compensators and are available in all-bronze or steel construction. Metal bellows expansion joints can generally be designed for the pressures and
temperatures commonly encountered in district heating distribution systems and can also
be furnished in rectangular configurations for ducts and chimney connectors as might be
found in the central plant.
Overpressurization, improper guiding, and other forces can distort the bellows element. For low-pressure applications, such distortion can be controlled by the geometry of
the convolution or the thickness of the bellows material. For higher pressures, internally
pressurized joints require reinforcing and externally pressurized designs are not subject to
such distortion and are not generally furnished without supplemental bellows reinforcing.
Single- and double-bellows expansion joints primarily accommodate axial movement
only, similar to packed slip expansion joints. Although bellows expansion joints can
accommodate some lateral movement, the universal tied bellows expansion joint accommodates large lateral movement. This device operates much like a pair of flexible ball
joints, except that bellows elements are used instead of flexible ball elements. The tie rods
on this joint contain the pressure thrust, so anchor loads are much lower than with axialtype expansion joints.
Expansion Joint Guides and Anchors
Pipe guides must be positioned per the expansion joint manufacturers recommendations, ensuring that the expansion joint will be aligned. Anchors must also be provided as
specified by the expansion joint manufacturer.

CATHODIC PROTECTION OF DIRECT BURIED CONDUITS


Corrosion is an electrochemical process that occurs when a corrosion cell is formed.
A corrosion cell consists of an anode, a cathode, a connecting path between them, and an
electrolyte (soil or water). The structure of this cell is the same as a dry cell battery, and,
like a battery, it produces a direct electrical current. The anode and cathode in the cell
may be dissimilar metals, and due to differences in their natural electrical potentials, a
current flows from anode to cathode. When current leaves an anode, it destroys the anode
material at that point. The anode and cathode may also be the same material; differences
in composition, environment, temperature, stress, or shape make one section anodic and
an adjacent section cathodic. With a connection path and the presence of an electrolyte,
this combination also generates a direct electrical current and causes corrosion at the
anodic area. The cathodic protection system generates a reverse voltage strong enough to
stop the corrosion cell.
Cathodic protection is a standard method used by the underground pipeline industry
to further protect coated steel against corrosion. Cathodic protection systems are routinely
designed for a minimum life of 20 years. Cathodic protection may be achieved by the sacrificial anode method or the impressed current method.

Sacrificial Anode Systems


Sacrificial anode systems are normally used with well-coated structures. A direct current is applied to the outer surface of the steel structure with a potential driving force that
prevents the current from leaving the steel structure. This potential is created by connecting the steel structure to another metal, such as magnesium, aluminum, or zinc, which

4.38

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

becomes the anode and forces the steel structure to be the cathode. The moist soil acts as
the electrolyte. These deliberately connected materials become the sacrificial anode and
corrode. If they generate sufficient current they adequately protect the coated structure
while their low current output is not apt to corrode other metallic structures in the vicinity.

Impressed Current Systems


Impressed current systems use a rectifier to convert an alternating current power
source to usable direct current. The current is distributed to the metallic structure to be
protected through relatively inert anodes such as graphite or high-silicon cast iron. The
rectifier allows the current to be adjusted over the life of the system. Impressed current
systems, also called rectified systems, are used on long pipelines in existing systems with
insufficient coatings, on marine facilities, and on any structure where current requirements are high. They are installed selectively in congested pipe areas to ensure that other
buried metallic structures are not damaged.

Design, Maintenance, and Testing


A National Association of Corrosion Engineers (NACE) registered engineer should
design the cathodic protection system. The design of effective cathodic protection
requires information on the diameter of both carrier pipe and conduit casing, length of
run, number of conduits in a common trench, and number of system terminations in
access areas, buildings, etc. Soil from the construction area should be analyzed to determine the soil resistivity, or the ease at which current flows through the soil. Areas of low
soil resistivity require fewer anodes to generate the required cathodic protection current,
but the life of the system depends on the weight of anode material used. The design life
expectancy of the cathodic protection must also be defined. All anode material is theoretically used up at the end of the cathodic protection system life. At this point, the corrosion
cell reverts to the unprotected system and corrosion occurs at points along the conduit
system or buried metallic structure. Anodes may be replaced or added periodically to continue the cathodic protection and increase the conduit life.
A cathodically protected system must be electrically isolated at all points where the
pipe is connected to building or access (manhole) piping and where a new system is connected to an existing system. Conduits are generally tied to another building or access
piping with flanged connections. Flange isolation kits, including dielectric gaskets, washers, and bolt sleeves, electrically isolate the cathodically protected structure. If an isolation flange is not used, any connecting piping or metallic structure will be in the
protection system, but protection may not be adequate.
Cathodic protections systems require maintenance. The EOR is responsible for a
Maintenance Manual that delineates the required maintenance.
The effectiveness of cathodic protection can only be determined by an installation
survey after the system has been energized. Cathodically protected structures should be
tested at regular intervals to determine the continued effectiveness and life expectancy of
the system. Sacrificial anode cathodic protection is monitored by measuring the potential
(voltage) between the underground metallic structure and the soil versus a stable reference. This potential is measured with a high-resistance voltmeter and a reference cell.
The most commonly used reference cell material is copper/copper sulfate. One criterion
for protection of buried steel structures is a negative voltage of at least 0.85 V as measured between the structure surface and a saturated copper/copper sulfate reference electrode in contact with the electrolyte (the soil). Impressed current systems require more
frequent and detailed monitoring than sacrificial anode systems. The rectified current and
potential output and operation must be verified and recorded at monthly intervals.

4.39

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

NACE SPO169 (NACE 2007) has further information on control of external corrosion on buried metallic structures.

LEAK DETECTION
Various techniques are available for detecting leaks in district heating and cooling
piping. They range from performing periodic pressure tests on the piping system to
installing a sensor cable along the entire length of the piping to continuously detect and
locate leaks. Pressure testing should be performed on all piping to verify integrity during
installation and during the life of the piping. Hot-water and steam systems should be
tested during the summer when an outage period is acceptable.
A leak is difficult to locate without the aid of a cable-type leak detector. Finding a
leak typically involves excavating major sections between valve vaults. Infrared detectors
and acoustic detectors can help narrow down the location of a leak, but they do not work
equally well for all underground systems. Also, they are not as accurate with underground
systems as they are with aboveground systems.

WSL Conduit Systems


WSL conduits are usually insulated with foam insulation with a water proof jacket
(HDPE, FRP, etc.). Special wires or cables can be installed during fabrication to aid in
detecting and locating liquid leaks. The wires may be insulated or un-insulated depending
on the manufacturer. Some systems monitor the entire wire length while others only monitor at the joints of the piping system. The detectors either look for a short in the circuit
using Ohms Law or monitor for impedance change using time domain reflectometry
(TDR). Some types of leak detection cable cannot be dried out after coming in contact
with water.

DDT-Type Conduit Systems


Air-gap designs, which have a gap between the inner wall of the casing and the pipe
insulation, can have probes installed at the low points of drains or at various points to
detect leaks. Leaks can also be detected with a continuous cable that monitors liquid leakage. The cable is installed at the bottom of the conduit with a minimum air gap required,
typically 1 in. (25.4 mm). Pull points or access ports are installed every 400 to 500 ft
(122 to 152 m) on straight runs with changes in direction reducing the length between
pull points. The detectors either look for a short on the cable using Ohms Law or monitor
the impedance on a coaxial cable using resistance temperature detectors (RTDs). During
installation care must be taken to keep the system clean and dry to keep any contamination from the leak detection system that might cause it to fail. The system must be sealed
airtight to prevent condensation from accumulating in the piping at the low points.

VALVE VAULTS AND ENTRY PITS


Manholes or valve vaults are required on underground distribution systems to provide
access to underground systems at critical points such as where there are
high or low points on the system profile that vent trapped air or where the system can be drained;
elevation changes in the distribution system that are needed to maintain the
required constant or nonreversing slope (a slope of 1:240 is a reasonable construction value that can be achieved);
major branches with isolation valves, steam traps, and condensate drainage
points on steam lines;
mechanical expansion devices; and/or
sump pumps.

4.40

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

To facilitate leak location and repair and to limit damage caused by leaks, access
points generally should be spaced no farther than 500 ft (150 m) apart. Special attention
must be given to the safety of personnel who come in contact with distribution systems or
who must enter spaces occupied by underground systems. The regulatory authoritys definition of a confined space and the possibility of exposure to high-temperature or highpressure piping have a significant impact on the access design, which must be addressed
by the project designer or EOR. Hence, gravity venting of tunnels is required. Moreover,
manholes and tunnels should have electric sump pumps with lighting and convenience
outlets to inspect and maintain anchors, expansion joints, and piping conditions. Electric
sump pumps may not be required if the manholes can be drained to a sanitary system.
French drains are usually not acceptable because groundwater will back-flow into the
manhole when high groundwater levels occur.
For most buried heat distribution systems, connections to the system are made in
manholes or valve vaults. For low-temperature hot-water systems, pre-insulated fittings
and valves suitable for direct burial may be used. An example of a pre-insulated valve
suitable for direct burial would be one conforming to EN 488 (CEN 2011) as discussed
previously. Any appurtenances to a pre-insulated system must ensure the integrity of the
waterproofing jacket. In addition, for any buried pre-insulated piping system, corrosion of
the components exposed to the soil must be considered and metal piping must be protected as discussed in the section on cathodic protection. This complicates direct burial of
fixtures such as vents and drains.
Valve vaults allow a user to isolate one segment of a system rather than analyze the
entire system or a large section thereof. Isolation may be for both the purpose of routine
maintenance as well as addressing operational problems or failures. The term vault is
used to eliminate confusion with sanitary manholes. Valve vaults are important when the
underground distribution system cannot be maintained without excavation, such as for
conduits and deep bury small tunnels. Valve vaults allow for step elevation change in the
distribution system piping while maintaining an acceptable slope on the system; they also
allow the designer to better match the topography and avoid unreasonable and expensive
burial depths. For all their positive attributes, valve vaults also have a significant potential
to generate maintenance requirements of their own as well as provide the potential for
problems, and, if left uncorrected, allow the problems to cascade into the adjacent buried
portions of the heat distribution system.

Valve Vault Penetrations


One of the basic functions of a valve vault is to provide a dry, corrosion-free environment for the piping and appurtenances that are located in the valve vault. This means that
the vault walls must not let groundwater enter the vault. Typical types of penetrations of
the vault walls are heat distribution conduits, electrical service conduits, sump pump discharge pipes, sanitary drains, conduit air space vents, and vault ambient air vents. Each
penetration must be designed to be watertight. Reference Figure 4.7note the leak plate
that is welded to the steel casing and cast into the concrete vault wall. All penetrations
must have a method to provide a positive water seal between the vault wall and the pipe or
electrical conduit. Precast or existing vault walls do not have leak plates cast into the
walls; thus, typically a link seal or a type of segmented adjustable compressed rubber seal
is used; see Figure 4.20. These rubber seals will work poorly where construction quality
control is poor. The holes in the vault wall are often the wrong diameter and the heat distribution conduit tends to deflect when the seal puts on a radial compressive load (needed
to develop the water seal). Wall penetrations should be provided with a galvanized steel
sleeve with a leak plate welded to it. Conduits that penetrate the vault wall at an angle

4.41

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 4.20 A segmented adjustable compressed rubber seal between the outer conduit of a
DDT system and a hole cored in a manhole wall. This installation lacks the recommended sleeve
with leak plate cast into the manhole wall as shown in Figure 4.7.

cause especially difficult sealing problems for precast construction because the hole in the
vault wall is not perpendicular to the conduit. Conduits with plastic casings and conduits
with external insulation and plastic jackets typically deflect radially inward when compressive seal loads are applied and thus let water past. When plastic jackets or casings are
used, a special design is needed to prevent radial deflection, or some other type of sealing
method is required. The following subsections discuss some of the issues associated with
valve vaults.

Ponding Water
The most significant problem with valve vaults is that water ponds in them. Ponding
water may be from carrier pipe leaks or intrusion of surface or groundwater into the valve
vault. When the hot- and chilled-water distribution systems share the same valve vault,
plastic chilled-water lines often fail because ponded water heats the plastic to failure. For
this reason it is not advisable to include both heating system and nonmetallic chilledwater piping in the same valve vaultor in close proximity outside of manholes, for that
matter. Historical experience indicates that water gathers in the valve vaults irrespective
of climate; therefore, a sound design strives to eliminate the water for the entire life of the
underground distribution system. Where possible, the most successful water removal systems are those that drain to sanitary or storm drainage systems; this technique is successful because the system is affected very little by corrosion and has no moving parts to fail.
Backwater valves are recommended in case the drainage system backs up.
Duplex sump pumps with lead-lag controllers and a failure annunciation system (see
Figure 4.21) are used when storm drains and sanitary drains are not accessible. Because
pumps have a history of frequent failures, duplex pumps help eliminate short cycling and
provide standby pumping capacity. Steam ejector pumps can be used only if the distribution system is never shut down because the carrier pipe insulation can be severely damaged during even short outages. A labeled, lockable, dedicated electrical service should

4.42

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

Figure 4.21 Valve vault duplex sump pump control and trouble annunciation system.
The aboveground location eliminates the components from the undesirable vault atmosphere
but will score low on aesthetics and vandalism potential where those are issues.

be used for electric pumps. The circuit label should indicate what the circuit is used for; it
should also warn of the damage that will occur if the circuit is de-energized.
Electrical components have experienced accelerated corrosion in the high heat and
high humidity of closed unventilated vaults. A pump that works well at 50F (10C) often
performs poorly, or not at all, at 200F (93C) and 100% relative humidity. To resolve this
problem, one approach specifies components that have demonstrated high reliability at
200F (93C) and 100% relative humidity with a damp-proof electrical service. The
pump should have a corrosion-resistant (when immersed in water) shaft and impeller and
have demonstrated 200,000 cycles of successful operation, including the electrical
switching components, at the referenced temperature and humidity. The pump must also
pass foreign matter; therefore, the requirement to pass a 3/8 in. (10 mm) ball should be
specified. The sump pump intake should be screened to prevent entry of foreign matter
that could prevent the pump from working as intended.
Another method drains the valve vaults into a separate sump similar to a sanitary
manhole located adjacent to the vault. The pumps are placed in this remote sump, which
is cool and more nearly a typical sump pump environment. Redundant methods may be
necessary if maximum reliability is needed or future maintenance is questionable. The

4.43

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

pump can discharge to the sanitary or storm drain or to a splash block near the valve vault.
Water pumped to a splash block has a tendency to enter the vault, but this is not a significant problem if the vault construction joints have been sealed properly. Extreme caution
must be exercised if the bottom of the valve vault has French drains. These drains work
backward when the groundwater level is high and allow groundwater to enter the vault
and flood the insulation on the distribution system. To control moisture, adequate ventilation of the valve vault is also important, as discussed later.

Crowding of Components
The valve vault must be laid out in three dimensions, considering standing room for
the worker, wrench swings, the size of valve operators, variation between manufacturers
in the size of appurtenances, and all other variations that the specifications allow with
respect to any item placed in the vault. To achieve desired results, the vault layout must be
shown to scale on the contract drawings.

High Humidity
High humidity develops in a valve vault when it has no positive ventilation. Gravity
ventilation is often provided in which cool air enters the valve vault and sinks to the bottom. At the bottom of the vault the air warms, becomes lighter, and rises to the top of the
vault, where it exits. In the past, some designers used a closed-top valve vault with an
exterior ventilation pipe with an elbow that directed the exiting air down. However, the
elbowed-down vent hood tends to trap the exiting air and prevent gravity ventilation from
working. Open structural grate tops are the most successful covers for ventilation purposes and safety (Figure 4.24). Open grates allow debris and rain to enter the vault; however, the techniques mentioned in the section on ponding water are sufficient to handle the
rainwater. Open grates with remote sump basins have worked well in extremely cold climates and in warm climates. Some vaults have a closed top and use screened, elevated
sides to allow free ventilation. In this design, the solid vault sides extend slightly above
grade and a screened window is placed in the wall on at least two sides (Figure 4.25). The
overall above-grade height may be only 18 in. (460 mm).

High Temperatures
The temperature in the valve vault rises when no systematic way is provided to
remove heat from the distribution system pipes in the valve vault. The gravity ventilation
rate is usually not sufficient to transport heat from the closed vaults. Part of this heat
transfers to the earth; however, an equilibrium temperature is reached that may be higher
than desired. Ventilation techniques discussed in the section on high humidity can resolve
the problem of high temperature if the heat loss from the distribution system is near normal. Typical problems that greatly increase the amount of heat released include the following:
Leaks from a carrier pipe, gaskets, packings, or appurtenances
Insulation that has deteriorated due to flooding or abuse
Standing water in a vault that touches the distribution pipe
Steam vented to the vault from partial conduit flooding between valve vaults
Vents from flash tanks
Insulation that was removed during routine maintenance and not replaced
To prevent heat buildup in a new system, a workable ventilation system must be
designed and specified. On existing valve vaults, the valve vault must be ventilated properly, all leaks must be corrected, and all insulation that was damaged or left off must be
replaced. Commercially available insulation jackets that can easily be removed from and

4.44

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

reinstalled on fittings and valves should be used to ease maintenance and increase the
likelihood that the valves and fitting will remain insulated. Vents from vault appurtenances that exhaust steam or vapor into the vault may have to be routed aboveground if
the ventilation technique is insufficient to handle the quantity of steam or vapor
exhausted.

Deep Burial
When a valve vault is buried too deeply, (1) the structure is exposed to potentially
greater groundwater pressures, (2) entry and exit often become a safety problem, (3) construction becomes more difficult, and (4) the cost of the vault is greatly increased. The
most common way to limit burial depth is to place the valve vaults closer together. Steps
in the distribution system slope are made in the valve vault (i.e., the carrier pipes come
into the valve vault at one elevation and leave at a different elevation). If the slope of the
distribution system is changed to more nearly match the earth topography, the valve
vaults will be shallower; however, the allowable range of slope of the carrier pipes
restricts this method. In most systems, the slope of the distribution system can be reversed
in a valve vault, but not within the system between valve vaults. The minimum slope for
the carrier pipes is 1 in. in 20 ft (25.4 mm in 6.1 m) or 0.4%. Lower slopes are outside the
range of normal construction tolerance. If the entire distribution system is buried too
deeply, the designer must determine the maximum allowable burial depth of the system
and survey the topography of the distribution system to determine where the maximum
and minimum depth of burial will occur. All elevations must be adjusted to limit the minimum and maximum allowable burial depths.

Freezing Conditions
Failure of distribution systems due to water freezing in components is common. The
designer must consider the coldest temperature that may occur at a site and not the 99%
or 99.6% condition used in building design (see ASHRAE [2009]). Drain legs or vent
legs that allow water to stagnate are usually the cause of failure. Insulation should be on
all items that can freeze, and it must be kept in good condition. Electrical heat tape and
pipe-type heat tracing can be used under insulation. If part of another type of water system is also in a ventilated valve vault, the water may have to be circulated or drained if not
used in winter.

Safety and Access


Some of the working fluids used in underground distribution systems can cause
severe injury or death if accidentally released in a confined space such as a valve vault.
Shallow valve vaults with large openings are desirable because the openings allow personnel to escape quickly in an emergency. The layout of the pipes and appurtenances
must allow easy access for maintenance without requiring maintenance personnel to
crawl underneath or between other pipes. The task of the designer is to keep clear work
spaces for maintenance personnel so that they can work efficiently and, if necessary, exit
quickly. Engineering drawings must show pipe insulation thickness to scale; otherwise
they will give a false impression of the available space.
The location and type of ladder is important for safety and ease of egress. It is best to
lay out the ladder and access openings when laying out the valve vault pipes and appurtenances as a method of exercising control over safety and ease of access. Ladder steps,
when cast in the concrete vault walls, may corrode if not constructed of the correct material. Corrosion is most common in steel rungs. Either cast iron or prefabricated, OSHAapproved, galvanized steel ladders that sit on the valve vault floor and are anchored near

4.45

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

the top to hold it into position are best (see Figure 4.22). If the design uses lockable
access doors, the locks must be operable from the inside or have some keyed-open device
that allows workers to keep the key while working in the valve vault.
Valve vaults in city streets or sidewalks present special challenges due to the inability
to provide venting; requirements that they be level (or nearly so) with the street surface,
which almost always results in surface water infiltration; and the limited amount of access
area that can be provided. Normally such valve vaults are provided with two access ports,
one that may be used for venting while the other is used for access. Venting may require
use of portable exhaust blowers to provide adequate ventilation to result in suitable working conditions.

Vault Construction
The most successful valve vaults are those constructed of cast-in-place reinforced
concrete. These vaults conform to the earth excavation profile and show little movement
when backfilled properly. Leakproof connections can be made with mating tunnels and
conduit casings even though they may enter or leave at oblique angles. In contrast, prefabricated valve vaults often settle and move after construction is complete. Penetrations for
prefabricated vaults, as well as the angles of entry and exit, are difficult to locate exactly.
As a result, much of the work associated with penetrations is not detailed and must be
done by construction workers in the field, which can greatly lower the quality and greatly
increases the chances of a groundwater leaking into the valve vault.
Normal construction deficiencies that go unnoticed in buildings can destroy a heating
and cooling distribution system; therefore, the designer must clearly convey to the contractor that a valve vault does not behave like a sanitary manhole. A design that is sufficient for a sanitary manhole will cause the appurtenances in a heat distribution system

Figure 4.22 Steel valve vault ladder, fabricated, galvanized, then mounted to the vault wall.

4.46

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

manhole to fail prematurely because many of the requirements mentioned herein are not
provided by a sanitary manhole.

Vault Covers and Venting


Many of the problems discussed here (e.g., access, high temperature, high humidity)
can be successfully addressed, at least in part, by proper vault cover design. Obviously
vault covers will be problematic in city streets, where little more than being certain that
they are at or slightly above grade is about all that can be done from the standpoint of the
cover itself. Alternate approaches to the cast-iron manhole cover are possible, such as the
option shown in Figure 4.23, with design and material selection sufficient to withstand
wheel loads and corrosive deicing chemicals where applicable. For applications in streets,
supplemental vaults are often located adjacent to the main valve vault to contain appurtenances such as steam traps, sump pumps, etc. This may help provide a better environment
for equipment requiring more frequent service as well as provide a location with better
access both physically and from a service environment and safety perspective.
Where the limitations of city streets do not apply, other approaches to vault covers
have proven quite successful. One design that has proven very successful is an open
grated cover, either at grade level (Figure 4.24) if dictated by the application or elevated if
possible. Note that in Figure 4.24 the sidewalk in the background is a shallow trench type
heat distribution system (also including chilled water) with removal lids.

Figure 4.23 Manhole cover suitable for streets. Note one of the two gooseneck vents
in the background.

4.47

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 4.24 Open grate type valve vault cover. Note sump pump discharge in foreground
and shallow trench heat distribution system in background. The grating is in
removable sections providing for excellent service access.

Figure 4.25 Perimeter-ventilated valve vault cover. Portions of the lids are hinged for easy access;
for major work the entire cover may be removed using the corner lifting eyes.

Another successful type of manhole cover uses perimeter ventilation, thus avoiding
most rainwater infiltration as will occur with the design shown in Figure 4.24. An example of this design is shown in Figure 4.25; note that this design must be elevated. It also
provides less ventilation than the design shown in Figure 4.24, but that is likely offset by
better isolation of rainwater in high rainfall climates.

Construction of Systems without Valve Vaults


Systems have been designed where features such as valves are directly buried and
then remotely operated, much the same as is frequently done for potable water distribution systems. This approach is prevalent with the European low-temperature hot-water
systems discussed previously. For these European systems, standards have been estab-

4.48

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

lished for all directly buried components as noted. In order for such an approach to be
successful, a continuous high-integrity jacket system, as well as the necessary seals for
operable shafts, etc., must be provided. This must be done with the additional confounding factor of thermal expansion and contraction of the system components. If the waterproofing is not entirely successful, groundwater will enter the system, causing corrosion
and deterioration of the insulation thermal properties. Achieving adequate waterproofing
of features such as direct burial valves and vents/drains will be very difficult, if not
impossible, to achieve in field-fabricated systems. High-temperature distribution systems
constructed without valve vaults have performed very poorly and thus are not recommended.

REFERENCES
ASHRAE. 2008. ASHRAE HandbookHVAC Systems and Equipment. Atlanta:
ASHRAE.
ASHRAE. 2009. ASHRAE HandbookFundamentals. Atlanta, GA: ASHRAE.
ASHRAE. 2012. ASHRAE HandbookHVAC Systems and Equipment. Atlanta:
ASHRAE.
ASME. 2012. ASME B31.1-2012, Power Piping. New York: ASME.
ASTM. 2011a. ASTM D2487-11, Standard Practice for Classification of Soils for Engineering Purposes (Unified Soil Classification System). West Conshohocken, PA:
ASTM International.
ASTM. 2011b. ASTM A106/A106M-11, Standard Specification for Seamless Carbon
Steel Pipe for High-Temperature Service. West Conshohocken, PA: ASTM International.
ASTM. 2012a. ASTM D698-12, Standard Test Methods for Laboratory Compaction
Characteristics of Soil Using Standard Effort (12 400 ftlbf /ft3 (600 kNm/m3)). West
Conshohocken, PA: ASTM International.
ASTM. 2012b. ASTM A53/A53M-12, Standard Specification for Pipe, Steel, Black and
Hot-Dipped, Zinc-Coated, Welded and Seamless. West Conshohocken, PA: ASTM
International.
ASTM. 2012c. ASTM A53/A53M, Standard Specification for Pipe, Steel, Black and HotDipped, Zinc-Coated, Welded and Seamless. West Conshohocken, PA: ASTM International.
Beitelman, A., C. Marsh, D. Neale, V. Meyer, J. Taylor, D. Butler, F. Mandan, L. Clark,
and T. Carlson. 2011. In situ corrosion and heat loss assessment of two nonstandard
underground heat distribution system piping designs. Final Report on Project F07AR01, U.S. Army Corps of Engineers Engineer Research and Development Center,
Construction Engineering Research Laboratory.
Bloomquist, R.G., R. OBrien, and M. Spurr. 1999. Geothermal district energy at colocated sites. WSU-EEP 99007, Washington State University Energy Office.
Bhm, B. 1986. On the optimal temperature level in new district heating networks. Fernwrme International 15(5):301306.
Bhm, B. 1988. Energy-economy of Danish district heating systems: A technical and
economic analysis. Laboratory of Heating and Air Conditioning, Technical University of Denmark, Lyngby, Denmark.
BRAB. 1957. Underground insulated piping systems (excluding walk-in tunnels). Technical Report No. 27, Federal Construction Council, Building Research Advisory Board,
National Research Council, National Academy of Sciences, Washington DC.

4.49

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

BRAB. 1958. Underground heat distribution systems. Technical Report No. 30, Federal
Construction Council, Building Research Advisory Board, National Research Council, National Academy of Sciences, Washington DC.
BRAB. 1963a. Field investigation of underground heat distribution systems. Technical
Report No. 47, Federal Construction Council, Building Research Advisory Board,
National Research Council, National Academy of Sciences, Washington DC.
BRAB. 1963b. Evaluation of components for underground heat distribution systems.
Technical Report No. 39-64, Federal Construction Council, Building Research Advisory Board, National Research Council, National Academy of Sciences, Washington
DC.
BRAB. 1963c. Field investigation of underground heat distribution systems. Technical
Report No. 47S, Federal Construction Council, Building Research Advisory Board,
National Research Council, National Academy of Sciences, Washington DC.
BRAB. 1964. Underground heat distribution systems. Technical Report No. 30R-64, Federal Construction Council, Building Research Advisory Board, National Research
Council, National Academy of Sciences, Washington DC.
BRAB. 1966. Proceedings of a symposium/workshop on underground heat distribution
systems. Symposium/Workshop Report Number 3, Federal Construction Council,
Building Research Advisory Board, National Research Council, National Academy
of Sciences, Washington DC.
BRAB. 1975. Criteria for underground heat distribution systems. Technical Report No.
66, Federal Construction Council, Building Research Advisory Board, National
Research Council, National Academy of Sciences, Washington DC.
CEN. 2009a. EN 253:2009, District heating pipes Preinsulated bonded pipe systems for
directly buried hot water networks Pipe assembly of steel service pipe, polyurethane thermal insulation and outer casing of polyethylene. Brussels: European Committee for Standardization.
CEN. 2009b. EN 448:2009, District heating pipes Preinsulated bonded pipe systems for
directly buried hot water networks Fitting assemblies of steel service pipes, polyurethane thermal insulation and outer casing of polyethylene. Brussels: European
Committee for Standardization.
CEN. 2009c. EN 489:2009, District heating pipes Preinsulated bonded pipe systems for
directly buried hot water networks Joint assembly for steel service pipes, polyurethane thermal insulation and outer casing of polyethylene. Brussels: European Committee for Standardization.
CEN. 2010. EN 13941:2009+A1:2010, Design and installation of preinsulated bonded
pipe systems for district heating. Brussels: European Committee for Standardization.
CEN. 2011. EN 488:2011, District heating pipes Preinsulated bonded pipe systems for
directly buried hot water networks Steel valve assembly for steel service pipes,
polyurethane thermal insulation and outer casing of polyethylene. Brussels: European Committee for Standardization.
Chyu, M.-C., X. Zeng, and L. Ye. 1997a. Performance of fibrous glass insulation subjected to underground water attack. ASHRAE Transactions 103(1):303308.
Chyu, M.-C., X. Zeng, and L. Ye. 1997b. The effect of moisture content on the performance of polyurethane insulation on a district heating and cooling pipe. ASHRAE
Transactions 103(1):30917.
Chyu, M.-C., X. Zeng, and L. Ye. 1998a. Behavior of cellular glass insulation on a DHC
pipe subjected to underground water attack. ASHRAE Transactions 104(2):16167.
Chyu, M.-C., X. Zeng, and L. Ye. 1998b. Effect of underground water attack on the performance of mineral wool pipe insulation. ASHRAE Transactions 104(2):16875.

4.50

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

4 Distribution Systems

COWIconsult. 1985. Computerized planning and design of district heating networks.


Virum, Denmark: COWIconsult Consulting Engineers and Planners AS.
DOD. 2003. Heating and cooling distribution systems, UFC-3-430-01FA, U.S. Department of Defense, July 25, available at www.wbdg.org/ccb/DOD/UFC/ufc_3_430
_01fa.pdf
DA. 2012. Military Soils Engineers, TM 3-34.64, Department of the Army, Washington,
DC.
Fox, J.A. 1977. Hydraulic Analysis of Unsteady Flow in Pipe Networks. New York: John
Wiley & Sons.
Geiringer, P.L. 1963. High Temperature Water Heating: Its Theory and Practice for District and Space Heating Applications. New York: John Wiley & Sons.
IDHA. 1983. District Heating Handbook, 4th ed. Washington, DC: International District
Heating Association.
Koskelainen, L. 1980. Optimal dimensioning of district heating networks. Fernwrme
International 9(4):8490.
Lund-Hansen, T. 2010. Practical use of real-time hydraulic modeling in day-to-day system operation and optimization: Examples from utilities in the US, and around the
world. Presented at the 101 Annual Conference of the IDEA, June 15, Indianapolis,
Indiana.
Lund-Hansen, T., and J.O. Hansen. 2011. Optimization of system operation & maintenance using advanced real-time monitoring. Presented at the Distribution Workshop
of the IDEA, February 22, Miami, Florida.
Marsh, C., and J. Carnahan. 1998. Comparison of failure rates in conduits for underground heat distribution. Journal of Transportation Engineering 124(6).
Marsh, C., and O. Marshall. 1998. Investigation of fiberglass-reinforced plastic (FRP)
condensate return carrier piping. Technical Report No 98/26, U.S. Army Corps of
Engineers, Construction Engineering Research Laboratory, Champaign, IL.
Marsh, C., N. Demetroulis, and J. Carnahan. 1996. Investigation of preapproved underground heat distribution systems. Report No 96/77, U.S. Army Corps of Engineers,
Construction Engineering Research Laboratory, Champaign, IL.
MSS. 2003. MSS SP-69-2003, Pipe Hangers and Supports Selection and Application.
Vienna, VA: Manufacturers Standardization Society.
MSS. 2009. MSS SP-58-2009, Pipe Hangers and Supports Materials, Design, Manufacture, Selection, Application, and Installation. Vienna, VA: Manufacturers Standardization Society.
NACE. 2007. NACE SP0169 (formerly RP0169), Control of External Corrosion on
Underground or Submerged Metallic Piping Systems. Houston: NACE International.
Phetteplace, G.E. 1989. Simulation of district heating systems for piping design. Paper
13. Proceedings of the International Symposium on District Heat Simulation, Reykjavik, Iceland.
Phetteplace, G., and D. Carbee. 1992. Survey of field experience with thermal pipe systems SUPER TEMP-TITE, IR 1113, U.S. Army Corps of Engineers, Cold Regions
Research and Engineering Laboratory, Hanover, NH.
Phetteplace, G., S.K. Monaghan, and G. Pedrick. 1998. Performance of water spread limiting and loose fill insulation of federal agency approved heat distribution systems.
Proceedings of the Eighty-Ninth Annual Conference of the International District
Energy Association LXXXVIX:18195
Phetteplace, G., S.K. Monaghan, G. Pedrick, and J. Lacombe. 2000. Performance of
water spread limiting and loose fill insulation of federal agency approved heat distri-

4.51

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

bution systems, US Army Corps of Engineers Cold Regions Research and Engineering Laboratory, Hanover, NH.
Randlv, P. 1997. District Heating Handbook. Fredericia, Denmark: European District
Heating Pipe Manufacturers Association.
Rasmussen, C.H., and J.E. Lund. 1987. Computer aided design of district heating systems. District Heating Research and Technological Development in Denmark, Danish
Ministry of Energy, Copenhagen, Denmark.
Reisman, A.W. 1985. District Heating Handbook: A Handbook of District Heating and
Cooling Models, Vol. 2. Washington, DC: International District Energy Association.
Segan, E.G., and C.-P. Chen. 1984. Investigation of tri-service heat distribution systems.
CERL Technical Report M-347. U.S. Army Construction Engineering Research Laboratory, Champaign, IL.
Stephenson, D. 1981. Pipeline Design for Water Engineers. New York: Elsevier Scientific.
Stewart, W.E., and C.L. Dona. 1987. Water flow rate limitations. ASHRAE Transactions
93(2):81125.
Streeter, V.L., and E.B. Wylie. 1979. Fluid Mechanics. New York: McGraw-Hill.
Vogler, M. 2011. Water hammer basics, a short refresher. Presented at the Distribution
Workshop of the IDEA, February 21, Miami, Florida.
Werner, S.E. 1984. The heat load in district heating systems. Chalmers University of
Technology, Gteborg, Sweden.

4.52

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

5
Consumer
Interconnection
INTRODUCTION
Thermal energy produced at the district heating plant is transported via the distribution network and is finally transferred to the consumer. This connection is commonly
called an energy transfer station or ETS. When district heating in the form of hot water or
steam is supplied, it may be used directly by the building HVAC system or process loads
or indirectly via a heat exchanger that transfers energy from one media to another. When
energy is used directly, it may be reduced in pressure that is commensurate to the buildings systems.
For commercially operated systems, a contract boundary or point of delivery divides
responsibilities between the energy provider and the customer. This point can be at a
piece of equipment, as in a heat exchanger with an indirect connection, or at flanges, as in
a direct connection. It is highly recommended that a chemical treatment analysis be performed (regardless of the type of connection) to determine the compatibility of each side
of the system (district and consumer) before energizing. Whether the connection is direct
or indirect, a cathodic isolation flange is recommended at the ETS to preserve system
pipe integrity, especially when connecting to older buildings (refer to Sperko [2009] and
Tredinnick [2008a]).

CONNECTION TYPES
Buildings (or other loads) may be connected either directly to the network, i.e., the
district heating system water is circulated within the building, or indirectly, where the
building water is separated from the district heating system water by a heat exchanger.
The design engineer must perform an analysis to determine which connection type is best.
Table 5.1 outlines the relative merits of each type of connection; additional details are
contained in the following subsections devoted to each type of connection.

Direct Connection
Because a direct connection offers no barrier between the district water and the buildings own system (e.g., air-handling heating coils, fan-coils, radiators, unit heaters, and
process loads), water circulated at the district plant has the same quality as the customers
water. Direct connections, therefore, are at a greater risk of incurring damage to equipment or contamination of the water based on the poor water quality of either party. Typically, district heating systems have contracts with water treatment vendors and monitor

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Table 5.1 Relative Merits of Direct and Indirect Consumer Interconnections


Issue

Direct Connection

Indirect Connection

Water
Quality

District heating system water is exposed to


building systems water, which may have lower
levels of treatment and filtering. Components within
existing building systems may have scale and
corrosion.

Water quality of district heating system is isolated


from building system and can be controlled.

Water
Consumption

Leakage and consumption of district heating


system water within the building may be difficult
to control and correct.

Water leakage is within control of


district heating utility.

Contractual

Demarcation of consumers building system


may not be clear.

Clear delineation between consumer and


district heating utility equipment.

Cost

Generally lower in overall cost due to absence of


heat exchanger and possible deletion of
building pumps and controls.

Higher cost due to heat exchanger and


additional controls.

Reliability

Failures within the building may cause problems or


potentially even outages for the district system.

District heating system is largely isolated from


any problems in the building beyond
the interconnection.

Pressure
Isolation

Building systems may need to be protected from


higher pressure in district heating system or,
for tall buildings, district heating system may be
subjected to higher pressures by building system.

Heat exchanger provides isolation of building


system pressure from district heating system
pressure and each may operate at their preferred
pressures without influence from the other.

Potential for greater T due to absence of


heat exchanger.

Approach temperature in heat exchanger


is a detriment to T.

In-Building
Space Requirements

Low space requirements.

Additional space required for heat exchanger


and controls.

water quality continuously. This may not be the case with all consumers. A direct connection is often more economical than an indirect connection because the consumer is not
burdened by the installation of heat exchangers, additional circulation pumps, or water
treatment systems; therefore, investment costs are reduced and return temperatures identical to design values are possible.
Figure 5.1 shows the simplest form of hot-water direct connection, where the district
heating plant is pumping the water through the consumer building. The installation in this
figure includes a pressure differential regulator (if required to reduce the system differential pressure to meet any lower building system parameters), a thermostatic control valve
on each terminal unit, a pressure relief valve, and a check valve. Most commercial systems have a flowmeter installed as well as temperature sensors and transmitters to calculate the energy used. Pressure transmitters may be installed as input for plant circulating
pump speed control. The location of each device may vary from system to system, but all
of the major components are indicated. The control valve is the capacity regulating device
that restricts flow to maintain either a supply or a return temperature on the consumers
side.
Particular attention must be paid to connecting high-rise buildings because they
induce a static head that will affect the design pressure of the entire system. Pressure control devices should be investigated carefully. It is not unusual to have a water-based district heating or cooling system with a mixture of direct and indirect connections in which
heat exchangers isolate the systems hydraulically based on the location of the ETS within
the building and the number of floors served.
In a direct system, the pressure in the main distribution system must meet local building codes to protect the customers installation and the reliability of the district system. To
minimize noise, cavitation, and control problems, constant-pressure differential control

5.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

5 Consumer Interconnection

Figure 5.1 Direct connection of building system to district hot water.


Adapted from ASHRAE HandbookHVAC Systems and Equipment, Chapter 12, Figure 23 (2012)

valves should be installed in the buildings. Special attention should be given to potential
noise problems at the control valves. These valves must correspond to the design pressure
differential in a system that has constantly varying distribution pressures because of load
shifts. Multiple valves may be required in order to serve the load under all flow and pressure ranges. Industrial-quality valves and actuators should be used for this application.
Depending on the size and design of the main system, elevation differences, and types
of customers and building systems, additional safety equipment, such as automatic shutoff valves on both supply and return lines, may be required.
When buildings have separate circulation pumps, primary/secondary piping, and
pumping, isolating techniques are used (cross-connection bridges between return and
supply piping, decouplers, and bypass lines). This ensures that terminal unit two-way
control valves are subjected only to the differential pressure established by the customers
building (tertiary) pump.

Indirect Connection
Many of the components of an indirect connection are similar to those used in directconnection applications, with the exception that a heat exchanger performs one or more
of the following functions: heat transfer, pressure interception, and/or buffering between
potentially different-quality water treatments.
Identical to that of a direct connection, the rate of energy extraction in the heat
exchanger of an indirect connection is governed by a control valve that reacts to the building load demand. Once again, the control valve usually modulates to maintain a temperature setpoint on either side of the heat exchanger, depending on the contractual agreement
between the consumer and the producer.
The three major advantages of using heat exchangers are (1) the static head influences of a high-rise building are eliminated, (2) the two water streams are separated, and
(3) consumers must make up all of their own lost water. The disadvantages of using an
indirect connection are (1) the additional cost of the heat exchanger and (2) the tempera-

5.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

ture loss and increased pumping pressure because of the addition of another heat transfer
surface.

COMPONENTS
Heat Exchangers
Aside from their isolation function, heat exchangers act as the line of demarcation
between ownership responsibilities of the different components of an indirect system.
They transfer thermal energy and act as pressure interceptors for the water pressure in
high-rise buildings. They also keep fluids from each system (which may have different
chemical treatments) from mixing. Figure 5.2 shows a basic building schematic including
heat exchangers and secondary systems.
Reliability of the installation is increased if multiple heat exchangers are installed.
The number selected depends on the types of loads present and how they are distributed
throughout the year. When selecting all equipment for the building interconnection, but
specifically heat exchangers, the designer should do the following:
Size the units capacity to match the given load and estimated load turndown as
closely as possible (oversized units may not perform as desired at maximum
turndown; therefore, several smaller units will optimize the installation).
Assess the critical nature of the load/operation/process to address reliability and
redundancy. For example, if a building has a 24-hour process load, consider adding a separate heat exchanger for this load. Also, consider operation and maintenance of the units. If the customer is a hotel, hospital, casino, or data center,

Figure 5.2 Basic heating-system schematic.


Courtesy of ASHRAE (2012, Chapter 12, Figure 24)

5.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

5 Consumer Interconnection

select a minimum of two units at 50% load each to allow one unit to be cleaned
without interrupting building service. Separate heat exchangers should be capable of automatic isolation during low-load conditions to increase part-load performance.
Determine the customers temperature and pressure design conditions. Some
gasket materials for plate heat exchangers have limits for low pressure and temperature.
Evaluate the customers water quality (i.e., use appropriate fouling factor).
Determine the available space and structural factors of the mechanical room.
Quantify the design temperatures. The heat exchanger may require rerating at a
lower inlet temperature during off-peak hours.
Calculate the allowable pressure drop on both sides of heat exchangers. The customers side is usually the most critical for pressure drop. The higher the pressure drop, the smaller and less expensive the heat exchanger. However, the
pressure drop must be kept within reasonable limits (15 psi [100 kPa] or below)
if the existing pumps are to be reused.

All heat exchangers should be sized with future expansion in mind. When selecting
heat exchangers, be cognizant that closer approach temperatures and low pressure drop
require more heat transfer area and hence heat exchangers that can supply these cost more
and take up more space. Strainers should be installed in front of any heat exchangers and
control valve to keep debris from fouling surfaces.
Plate, shell-and-coil, and shell-and-tube heat exchangers are all used for indirect connection. Whatever heat transfer device is selected must meet the appropriate temperature
and pressure duty and be stamped/certified accordingly as pressure vessels. It is advantageous and recommended that if one side of the unit has variable flow, the other side
should as well. This ensures that minimum pump and thermal energy is used to satisfy a
load (Tredinnick 2007). Refer to Chapter 48 of ASHRAE HandbookHVAC Systems and
Equipment (2012) for more information on heat exchangers.
Plate Heat Exchangers
Plate heat exchangers (PHEs), which are used for either steam or hot-water applications, are available as gasketed units and in two gasket-free designs (brazed and all- or
semi-welded construction). All PHEs consist of metal plates compressed between two
end frames and sealed along the edges. Alternate plates are inverted and the gaps between
the plates form the liquid flow channels. Fluids never mix, as hot fluid flows on one side
of the plate and cool fluid flows countercurrent on the other side. Ports at each corner of
the end plates act as headers for the fluid. One fluid travels in the odd-numbered plates
and the other in the even-numbered plates.
Because PHEs require turbulent flow for proper heat transfer, pressure drops may be
higher than those for comparable shell-and-tube models. High efficiency leads to a
smaller package. The designer should consider specifying that the frame be sized to hold
20% additional plates. PHEs require very little maintenance because the high velocity of
the fluid in the channels tends to keep the surfaces clean from fouling. However, larger
particles may become lodged in the fine cavities between the plates and choke flow; automatic back-flushing valves may be used to address this issue. PHEs generally have a cost
advantage and require one-third to one-half the surface required by shell-and-tube units
for the same operating conditions. PHEs are also capable of closer approach temperatures.

5.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Gasketed PHEs (also called plate-and-frame heat exchangers) consist of a number of


gasketed embossed metal plates bolted together between two end frames. Gaskets are
placed between the plates to contain the two media in the plates and to act as a boundary.
Gasket failure will not cause the two media to mix; instead, the media leaks to the atmosphere. Gaskets can be either glued or clip-on. Gasketed PHEs are suitable for steam-toliquid and liquid-to-liquid applications. Designers should select the appropriate gasket
material for the design temperatures and pressures expected. Plates are typically stainless
steel; however, plate material can vary based on the chemical makeup of the heat transfer
fluids.
PHEs are typically used for district heating with water. Double-wall plates are also
available for potable-water heating, chemical processes, and oil quenching. PHEs have
three to five times greater heat transfer coefficients than shell-and-tube units and are capable of achieving 1F (0.5C) approach temperatures, but for economic reasons the
approach is traditionally 2F (1C). Gasketed PHEs can be disassembled in the field to
clean the plates and replace the gaskets. Typical applications go up to 365F and 400 psig
(185C and 2.8 MPa). Plates are typically made from stainless steel but are available in
titanium for more corrosive uses such as geothermal waters.
Brazed PHEs are suitable for steam, vapor, or water solutions. They feature a close
approach temperature (within 2F [1C]), large temperature drop, compact size, and high
heat transfer coefficient. Construction materials are stainless steel plates and frames
brazed together with copper or nickel. Tightening bolts are not required as with the gasketed design. These units cannot be disassembled and cleaned; therefore, adequate strainers must be installed ahead of an exchanger and they must be periodically flushed clean in
a normal maintenance program. Brazed PHEs usually peak at a capacity of under 200,000
Btu/h (60 kW) (about 200 plates and 600 gpm [0.04 m3/s]) and are suitable for 435F and
435 psig (225C and 3 MPa). Double-wall plates are also available for domestic hot-water
heating. Applications where the PHE may be exposed to large, sudden, or frequent
changes in temperature and load must be avoided because of risk of thermal fatigue.
Welded PHEs can be used in any application for which shell-and-tube units are used
that are outside the accepted range of gasketed PHE units in hot-water or steam district
heating applications. Construction is very similar to that of gasketed units except the gaskets are replaced with laser welds. Materials are typically stainless steel, but titanium,
monel, nickel, and a variety of alloys are available. Models offered have design ratings
that range from 500F at 150 psig to 1000F at 975 psig (260C at 1 MPa to 540C at
6.7 MPa); however, they are available only in small sizes. Normally, these units are used
for ammonia refrigeration and aggressive process fluids. They are more suitable to pressure pulsation or thermal cycling because they are thermal-fatigue resistant. A semiwelded PHE is a hybrid of the gasketed and the all-welded units in which the plates are
alternatively sealed with gaskets and welds.
Shell-and-Coil Heat Exchangers
Shell-and-coil heat exchangers are European-designed heat exchangers that are suitable for steam-to-water and water-to-water applications and feature an all-welded-andbrazed construction. This counter/crossflow heat exchanger consists of a hermetically
sealed (free of gaskets), carbon-steel pressure vessel with hemispherical heads. Copper or
stainless steel helical tubes within are installed in a vertical configuration. This type of
heat exchanger offers a high temperature drop and close approach temperature. Its vertical arrangement requires less floor space than other designs and has better heat transfer
characteristics than shell-and-tube units.

5.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

5 Consumer Interconnection

Shell-and-Tube Heat Exchangers


Shell-and-tube heat exchangers are usually a multiple-pass design. The shell is usually constructed from steel and the tubes are often of U-bend construction, usually 3/4 in.
(20 mm) (nominal) outside diameter copper, but other materials are available. These units
are ASME U-1 stamped for pressure vessels (ASME 2013).

Flow Control Devices


In commercial systems, after the flowmeter the control valves are the most important
element in the interface with the district heating system because proper valve adjustment
and calibration save energy. High-quality, industrial-grade control valves provide more
precise control, longer service life, and minimum maintenance.
All control valve actuators should take longer than 60 seconds to close from full open
to mitigate pressure transients and water hammer, which occurs when valves slam closed.
Actuators should also be sized to close against the anticipated system pressure so the
valve seats are not forced open, causing water to bypass the closed valve and degrading
the temperature differential.
The wide range of flows and pressures expected makes selection of control valves
difficult. Typically, only one control valve is required; however, for optimal response to
load fluctuations and to prevent cavitation, two valves in parallel are often needed. The
two valves operate in sequence and for a portion of the load (i.e., one valve is sized for
two-thirds of peak flow and the other sized for one-third of peak flow). The designer
should review the occurrence of these loads to size the proportions correctly. The possibility of overstating customer loads complicates the selection process, so accurate load
information is important. It is also important that the valve selected operates under the
extreme pressure and flow ranges foreseen. Because most commercial-grade valves will
not perform well under these conditions, industrial-quality valves are typically specified
for district heating applications.
Electronic control valves should remain in a fixed position when a power failure
occurs and should be manually operable. Pneumatic control valves should close upon loss
of air pressure. A manual override on the control valves allows the operator to control
flow. All steam pressure-reducing valves should fail closed as well.
Oversizing results in a reduction of valve and actuator lifespan and causes valve hunting. Select control valves having a wide range of control; low leakage; and proportionalplus-integral control for close adjustment, balancing, temperature accuracy, and response
time. Control valves should have actuators with enough force to open and close under the
maximum pressure differential in the system. The control valve should have a pressure
drop through the valve equal to at least 10% to 30% of the static pressure drop of the distribution system. This pressure drop gives the control valve the authority it requires to
properly control flow. The relationship between valve travel and capacity output should
be linear, with an equal percentage characteristic.
In hot-water systems, control valves are normally installed in the return line because
the lower temperature in the line reduces the risk of cavitation and increases valve life.

Instrumentation
In many systems, where energy to the consumer is measured for billing purposes,
temperature sensors assist in calculating the energy consumed as well as in diagnosing
performance. Sensors and their transmitters should have an accuracy range commensurate
to the accuracy of the flowmeter. In addition, pressure sensors are required for variablespeed pump control (water systems) or valve control for pressure-reducing stations
(steam and water systems).

5.7

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Temperature sensors need to be located by the exchangers being controlled rather


than in the common pipe. Improperly located sensors will cause one control valve to open
and others to close, resulting in unequal loads in the exchangers. Temperature sensors
used in energy metering applications should be matched pair, four-wire, and 1000 ohm
for increased accuracy.

Controller
The controller performs several functions, including recording demand and the
amount of energy used for billing purposes, monitoring the differential pressure for plant
pump control, performing energy calculations, alarming for parameters outside normal,
and monitoring and controlling all components. Many times a small battery power supply
is present in the controller to allow preservation of settings and billing information in the
event of a power outage.
Typical control strategies include regulating district flow to maintain the customers
supply temperature (which results in a fluctuating customer return temperature) or maintaining the customers return temperature (which results in a fluctuating customer supply
temperature).

Pressure Control Devices


If the steam or water pressure delivered to the customer is too high for direct use, it
must be reduced. Similarly, pressure-reducing or pressure-sustaining valves may be
required if building height creates a high static pressure and influences the district systems return water pressure. Water pressure can also be reduced by control valves or
regenerative turbine pumps. The risk of using pressure-regulating devices to lower pressure on the return line is that if they fail, the entire distribution system is exposed to their
pressure and overpressurization will occur.
In high-rise buildings, all piping, valves, coils, and other equipment may be required
to withstand higher design pressures. Where system static pressure exceeds safe or economical operating pressure, either the heat exchanger method or pressure-sustaining
valves in the return line may be used to minimize the impact of the pressure. Vacuum
vents should be provided at the top of the buildings water risers to introduce air into the
piping in case the vertical water column collapses.

HEATING CONNECTIONS
Steam Connections
Although higher pressures and temperatures are sometimes used, most district heating systems supply saturated steam at pressures between 5 and 250 psig (35 and 1000 kPa
[gage]) to customers facilities. The steam is pretreated to maintain a neutral pH (see
Chapter 8), and the condensate is both cooled and discharged to the building sewage system or returned back to the central plant for recycling. Many consumers run the condensate through a heat exchanger to heat the domestic hot-water supply of the building before
returning it to the central plant or sometimes to the building drains if the district system
does not have a condensate return pipe. This energy-saving process extracts the maximum
amount of energy out of the delivered steam.
Interconnection between the district and the building is simple when the building uses
the steam directly in heating coils or radiators or for process loads (humidification,
kitchen, laundry, laboratory, steam absorption chillers, or turbine-driven devices). Other
buildings extract the energy from the district steam via a steam-to-water heat exchanger to
generate hot water and circulate it to the air-side terminal units. Typical installations are
shown in Figures 5.3 and 5.4. The type of steam chemical treatment should be carefully
considered in applications for the food industry and for humidification because some

5.8

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

5 Consumer Interconnection

Figure 5.3 District/building interconnection with heat recovery steam system.


Courtesy of ASHRAE (2012, Chapter 12, Figure 25)

chemicals are more harmful than others and therefore shouldnt be used in these types of
applications (Tredinnick 2008b).
Other components of the steam connection may include condensate pumps, flowmeters (steam and/or condensate), and condensate conductivity probes, which may dump
condensate if it is contaminated by unacceptable debris. Many times, energy meters are
installed on both the steam and condensate pipes to allow the district energy supplier to
determine how much energy is used directly and how much energy (condensate) is not
returned back to the plant. The use of customer energy meters for both steam and condensate is desirable for the following reasons:
Offers redundant metering (if the condensate meter fails, the steam meter can
detect flow or vice versa)
Bills customer accordingly for makeup water and chemical treatment on all condensate that is not returned
Meter is in place if customer requires direct use of steam in the future
Assists in identifying steam and condensate leaks
Having two meters improves customer relations (may ease customers fears of
overbilling because of a faulty meter)
Provides a more accurate reading for peak demand measurements and charges

5.9

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 5.4 District/building interconnection with heat exchange steam system.


Adapted from ASHRAE HandbookHVAC Systems and Equipment, Chapter 12, Figure 26 (2012)

Each level of steam pressure reduction should also be monitored, as well as the temperature of the condensate. Where conductivity probes are used to monitor the quality of
the water returned to the steam plant, adequate drainage and cold-water quenching equipment may be required to satisfy local plumbing code requirements (the temperature of
fluid discharging into a sewer). The probe status should also be monitored at the control
panel to communicate high-conductivity alarms to the plant and, when condensate is
being dumped, to notify the plant a conductivity problem exists at a customer location.

Hot-Water Connections
Figure 5.5 illustrates a typical indirect connection using a heat exchanger between the
district hot-water system and the customers system. It shows the radiator configurations
typically used in both constant- and variable-flow systems. Figure 5.6 shows a variation
of a typical direct connection between the district hot-water system and the building hotwater systems. It includes the typical configurations for both demand flow and constantflow systems and the additional check valve and piping required for the constant-flow
system. The differential pressure flow sensor controller indicated in the bypass bridge is
not found on many contemporary systems, and the bridge may contain a check valve to
prohibit any district supply water blending with the building return water to affect the system T. Figure 5.7 shows an indirect connection for both space and hot-water heating.

5.10

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

5 Consumer Interconnection

Figure 5.5 District/building indirect interconnection with hot-water system.


Courtesy of ASHRAE (2012, Chapter 12, Figure 27)

Lines on these systems must be sized using the same design used for the main feed
lines of an in-building power plant. In general, demand flow systems permit better energy
transfer efficiency and smaller line size for a given energy transfer requirement. Line sizing should account for any future loads on the building, etc. To keep return temperature
low, water flow through heating equipment should be controlled according to the heating
demand in the space.
The secondary supply temperature must be controlled in a manner similar to the primary supply temperature. To ensure a low primary return temperature and large temperature difference, the secondary system should have a low design return temperature. This
design helps reduce costs through smaller pipes, pumps, amounts of insulation, and pump
motors. A combined return temperature on the secondary side of 130F (55C) or lower is
reasonable, although this temperature may be difficult to achieve in smaller buildings
with only baseboard or radiator space heating.
A hot-water district heating distribution system has many advantages over a steam
system. Because of the efficiency of the relatively low temperature, energy can be saved
over an equivalent steam system. Low- and medium-temperature hot-water district heat-

5.11

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 5.6 District/building direct interconnection for a hot-water system.


Courtesy of ASHRAE (2012, Chapter 12, Figure 28)

ing allows greater flexibility in the heat source, lower-cost piping materials, and more
cost-effective ways to compensate for expansion and new customer connections.
Major advantages of a hot-water distribution system include the following:
Changing both temperature and flow can vary the delivered thermal power and
save energy.
Hot water can be pumped over a greater distance with minimal energy loss. This
is a major disadvantage of a steam system.
With outdoor air reset control, the system operates at lower supply temperatures
over the year. This results in lower line losses. Steam systems normally operate
at constant pressure and temperature year round, which increases energy production and system losses while decreasing annual system efficiency.
Hot-water distribution systems have much lower operating and maintenance
costs than steam systems (leaks at steam traps, chemical treatment, spares, etc.).
Hot-water distribution systems use simpler, less expensive prefabricated insulated piping, which has a lower initial capital cost as compared to the typical airspace conduit-type drainable, dryable, testable (DDT) steam systems.

5.12

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

5 Consumer Interconnection

Figure 5.7 Building indirect connection for both heating and domestic hot water.
Courtesy of ASHRAE (2012, Chapter 12, Figure 29)

BUILDING CONVERSION TO DISTRICT HEATING


Table 5.2 summarizes the suitability or success rate of converting various heating systems to be served by a district hot-water system. As can be seen, the probability is high
for water-based systems, medium for steam, and low for fuel-oil or electric-based systems. Systems that are low on suitability usually require the high expense of replacing the
entire heating terminal and generating units with suitable water-based equipment, including piping, pumps, controls, and heat transfer media.

TEMPERATURE DIFFERENTIAL CONTROL


The success of a water-based district heating (or cooling) system achieving high efficiency is usually measured in terms of the temperature differential. Proper control of the
district heating temperature differential is not dictated at the plant but at the consumer. If
the consumers system is not compatible with the temperature parameters of the district
heating system, operating efficiency will suffer unless components in the consumers system are modified.
Generally, maintaining a high temperature differential (T) between supply and
return lines is most cost-effective because it allows smaller pipes to be used in the primary
distribution system. These savings must be weighed against any higher building conversion costs that may result from the need for a low primary return temperature. Furthermore, optimization of the T is critical to the successful operation of the district heating
system. That is the reason the customers T must be monitored and controlled.

METERING
All thermal energy or power delivered to customers or end users for billing or revenue
by a commercially operated district heating system must be metered. The type of meter
selected depends on the fluid to be measured, accuracy required, and expected turndown
of flow to meet low-flow and maximum-flow conditions. It is important that the meter be
sized accurately for the anticipated loads and not oversized, which would lead to inaccuracies. Historical, metered, or benchmarked data should be used when available if the
actual load is not accurately known.

5.13

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Table 5.2 Conversion Suitability of Heating System by Type (Sleiman et al. 1990)
Type of System

Low

Medium

High

Steam Equipment
One-pipe cast iron radiation

Two-pipe cast iron radiation

Finned-tube radiation

Air-handling unit coils


Terminal unit coils

X
X

Hot-Water Equipment
Radiators and convectors

Radiant panels

Unitary heat pumps

Air-handling unit coils

Terminal unit coils

Gas/Oil-Fired Equipment
Warm-air furnaces

Rooftop units

Other systems

Electric Equipment
Warm-air furnaces

Rooftop units

Air-to-air heat pumps

Other systems

Steam may be used for direct comfort heating or to power absorption chillers for
cooling. It is typically measured by using the differential pressure across calibrated orifices, nozzles, or venturi tubes or by pitot tubes, vortex-shedding meters, or condensate
meters. In the United States, the customary commercial unit is pounds of steam per hour
with a heat equivalent typically assumed to be 1000 Btu/lb (the mass flow can be converted to heat equivalent by assuming 2300 kJ/kg), but for more precise thermal metering
the meters are coupled with devices that measure steam quality via temperature, pressure,
and differential heat content. Care must be taken to deliver dry steam (superheated or saturated without free water) to the customer. Dry steam is delivered by installing an adequate trap just ahead of the customers meter (or ahead of the customers process when
condensate meters are used).
When condensate meters are used, care must be taken to ensure that all condensate
from the customers process, but only such condensate, goes to the condensate meter.
This may not be possible if the steam is used directly for humidification purposes and
hence no condensate is returned. In this case, steam meters are preferable.
The International District Energy Association (IDEA 1969) and Stultz and Kitto
(1992) offer further information on steam metering. For steam, as with hot- and chilledwater system metering, electronic and computer technology provide direct, integrating,
and remote input to central control/measurement energy management systems (EMSs).
Hot-water systems are metered by measuring the temperature differential between the
supply and return lines and the flow rate of the water. Thermal (Btu or kWh) meters compensate for the actual volume and heat content characteristics of the water. Thermal trans-

5.14

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

5 Consumer Interconnection

Table 5.3 Flowmeter Characteristics


Meter Type

Accuracy

Range of Control

Pressure Loss

Straight Piping
Requirements
(Length in Pipe
Diameters)

Orifice plate

1% to 5% full scale

3:1 to 5:1

High (>5 psi [>35 kPa])

10 D to 40 D upstream;
2 D to 6 D downstream

Electromagnetic

0.15% to 1% rate

30:1 to 100:1

Low (<3 psi [<20 kPa])

5 D to 10 D upstream;
3 D downstream

Vortex

0.5% to 1.25% rate

10:1 to 25:1

Medium (3 to 5 psi [20 to 35 kPa])

10 D to 40 D upstream;
2 D to 6 D downstream

Turbine

0.15% to 0.5% rate

10:1 to 50:1

Medium (3 to 5 psi [20 to 35 kPa])

10 D to 40 D upstream;
2 D to D downstream

Ultrasonic

1% to 5% rate

>10:1 to 100:1

Low (<3 psi [<20 kPa])

10 D to 40 D upstream;
2 D to 6 D downstream

Adapted from ASHRAE HandbookHVAC Systems and Equipment, Chapter 12, Table 10 (2012)

ducers, resistance thermometer elements, or liquid expansion capillaries are usually used
to measure the differential temperature of the water in supply and return lines.
Water flow can be measured with a variety of meters, usually pressure differential,
turbine or propeller, or displacement meters. Chapter 36 of the 2011 ASHRAE HandbookFundamentals, District Heating Handbook (IDHA 1983), and Pomroys (1994)
Selecting Flowmeters article have more information on measurement. Ultrasonic
meters are sometimes used to check performance of installed meters. Various flowmeters
are available for district heating billing purposes. Critical characteristics for proper installation include clearances and spatial limitations as well as the attributes presented in
Table 5.3. Note that the data in the table only provide general guidance, and the manufacturers of meters should be contacted for data specific to their products.
Heat meters should be located upstream of the heat exchanger and the control
valve(s) should be downstream from the heat exchanger. This orientation minimizes the
possible formation of bubbles in the flow stream and provides a more accurate flow indication. The transmitter should be calibrated for zero and span as recommended by the
manufacturer.
Wherever possible, the type and size of meters selected should be standardized to
reduce the number of stored spare parts, technician training, etc.
Displacement meters are more accurate than propeller meters, but they are also larger.
They can handle flow ranges from less than 2% up to 100% of the maximum rated flow
with claimed 1% accuracy. Turbine-type meters require the smallest physical space for a
given maximum flow. However, like many meters, they require at least 10 diameters of
straight pipe upstream and downstream of the meter to achieve their claimed accuracy.
The United States has no performance standards for thermal meters. ASHRAE Standard 125 (2000) describes a test method for rating liquid thermal meters. Several European countries have developed performance standards and/or test methods for thermal
meters; CEN Standard EN 1434, developed by the European Community, is a performance and testing standard for heat meters (2007a, 2007b, 2007c, 2007d, 2007e, 2008).
District energy plant meters intended for billing or revenue require means for verifying performance periodically. Major meter manufacturers, some laboratories, and some
district energy companies maintain facilities for this purpose. In the absence of a single
performance standard, meters are typically tested in accordance with their respective
manufacturers recommendations. Primary measurement elements used in these laborato-

5.15

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

ries frequently obtain calibration traceability to the National Institute of Standards and
Technology (NIST).
For district heating cogeneration systems that send out and/or accept electric power to
or from a utility grid, the demand and usage meters must meet the existing utility requirements. For district heating systems that send out electric power directly to customers, the
electric demand and usage meters must comply with local and state regulations. American National Standards Institute (ANSI) standards are established for all customary electric meters.

REFERENCES
ASHRAE. 2000. ANSI/ASHRAE Standard 125-1992 (RA 2000), Method of Testing
Thermal Energy Meters for Liquid Streams in HVAC Systems. Atlanta: ASHRAE.
ASHRAE. 2011. ASHRAE HandbookFundamentals. Atlanta: ASHRAE.
ASHRAE. 2012. ASHRAE HandbookHVAC Systems and Equipment. Atlanta:
ASHRAE.
ASME. 2013. ASME Boiler and Pressure Vessel Code. New York: ASME.
CEN. 2007a. CEN EN 1434-1:2007, Heat meters - Part 1: General requirements. Brussels: European Committee for Standardization.
CEN. 2007b. CEN EN 1434-2:2007, Heat meters - Part 2: Constructional requirements.
Brussels: European Committee for Standardization.
CEN. 2007c. CEN EN 1434-4:2007, Heat meters - Part 4: Pattern approval tests. Brussels: European Committee for Standardization.
CEN. 2007d. CEN EN 1434-5:2007, Heat meters - Part 5: Initial verification tests. Brussels: European Committee for Standardization.
CEN. 2007e. CEN EN 1434-6:2007, Heat meters - Part 6: Installation, commissioning,
operational monitoring and maintenance. Brussels: European Committee for Standardization.
CEN. 2008. CEN EN 1434-3:2008, Heat meters - Part 3: Data exchange and interfaces.
Brussels: European Committee for Standardization.
IDEA. 1969. Code for Steam Metering. Washington, DC: International District Energy
Association.
IDHA. 1983. District Heating Handbook, 4th ed. Washington, DC: International District
Heating Association.
Pomroy, J. 1994. Selecting flowmeters. Instrumentation & Control Systems 67(3):6168.
Sleiman, A.H., M. Spurr, A. Rydaker, and M. Karnitz. 1990. Guidelines for converting
building heating systems for hot water district heating. Report by International
Energy Agency Programme for Research, Development, and Demonstration on District Heating. NOVEM, Sittard, The Netherlands.
Sperko, W. 2009. Dissimilar metals in heating and AC piping systems. ASHRAE Journal
51(4).
Stultz, S.C., and J.B. Kitto, eds. 1992. Steam: Its Generation and Use, 40th ed. Chapter
40-13, Measurement of steam quality and purity; Chapter 40-19, Measurement of
steam flow. Barberton, OH: Babcock & Wilcox.
Tredinnick, S. 2007. Tales of the paranormal: PHEs variable countercurrent flows, and
the X-Files. Inside Insights, District Energy, 93(4).
Tredinnick, S. 2008a. Opposites attract: The color purple and galvanic corrosion. Inside
Insights, District Energy 94(2).
Tredinnick, S. 2008b. Humidification with steam: Go ahead and breathe easy! Inside
Insights, District Energy 94(3).

5.16

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6
Heat Transfer
Calculations
for Piping Systems
THERMAL DESIGN CONSIDERATIONS
Three operating conditions must be evaluated to ensure satisfactory system performance.
The normal condition used for the life-cycle cost analysis (LCCA) determines
appropriate insulation thickness. Average values for the temperatures, burial
depth, and thermal properties of the materials are used for design. If the thermal
properties of the insulating material are expected to degrade over the useful life
of the system, appropriate allowances should be made in the cost analysis.
The operating condition where the maximum heat transfer rate occurs determines the load on the central plant due to the distribution system. It also determines the temperature drop (or increase, in the case of chilled-water
distribution), which determines the delivered temperature to the consumer.
Because heat transfer in piping is not related to the load factor, it can be a large
part of the total load. For this calculation, the thermal conductivity of each component must be taken at its maximum value and the temperatures must be
assumed to take on their extreme values, which results in the greatest temperature difference between the carrier medium and the soil or air. The burial depth
will normally be at its lowest value for this calculation.
A final operating condition is one where none of the thermal capabilities of the
materials (or any other materials in the area influenced thermally by the system)
exceed their design values. To satisfy this objective, each component and the
surrounding environment must be examined to determine whether thermal damage is possible. A heat transfer analysis may be necessary in some cases.
Each operating condition may require a thermal analysis. The parameters for these
analyses must be chosen to represent the worst-case scenario from the perspective of the
operating concern being examined. For example, in assessing the suitability of a coating
material for a metallic conduit, the thermal insulation is assumed to be saturated, the soil
moisture is at its lowest probable level, and the burial depth is maximum. These conditions, combined with the highest anticipated pipe and soil temperatures, give the highest
conduit surface temperature to which the coating could be exposed.
Uncertainty in heat transfer calculations for thermal distribution systems results from
the uncertainty in the thermal properties of the materials involved. Generally, the designer
must rely on manufacturers specifications and handbook data to obtain approximate val-

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

ues. The data in this chapter should only be used as guidance in preliminary calculations
until specific products have been identified; then specific data should be obtained from
the manufacturer of the product in question.

SOIL THERMAL PROPERTIES


Heat transfer in buried systems is influenced by the thermal conductivity of the soil
and by the depth of burial, particularly when the thermal insulation of the system has low
thermal resistance. Soil thermal properties are principally a function of three factors:
the type of soil (grain size and composition),
the moisture content of the soil, and
the density (state of compaction) of the soil.

Soil Thermal Conductivity


In the absence of specific information on the soil type, moisture content, and density,
the thermal conductivity factors provided in Table 6.1 may be used as estimates. Because
dry soil is rare in most areas, low moisture content should be assumed only for system
material design, or where it can be validated for calculation of heat losses in the normal
operational condition. Values of 0.8 to 1 Btu/hftF (1.4 to 1.7 W/mK) are commonly
used where soil moisture content is unknown. Because moisture migrates toward chilled
pipe, use a thermal conductivity value of 1.25 Btu/hftF (2.16 W/mK) for chilled-water
systems in the absence of any site-specific soil data.
Ideally, when confronted with an engineering problem requiring soil thermal properties one would obtain samples from the site and have them tested. For most applications
this is much too expensive, although in-situ thermal property testing has become very
popular for design of ground-source heat pump systems; see Kavanaugh (2000). If an
analysis of the soil is available or can be done, the thermal conductivity of the soil can be
estimated from published data (e.g., Kersten [1949], Farouki [1981], and Lunardini
[1981]) for soils with similar composition and gradation. For steady-state analyses, only
the thermal conductivity of the soil is required. If a transient analysis is required, the specific heat and density are also required. If neither laboratory tests for thermal properties or
soil composition and gradation are available, the approximate equations developed by
Kersten (1949) may be used. Kersten presents one set of equations for fine-grained soils
(silts and clays) and another set for coarse-grained soils (sands and gravels). These equations are presented here in Inch-Pound (I-P) and Systme International (SI) units.
In I-P units:
k s = 0.083 0.9 log w 0.2 10 0.01 d
k s = 0.00083 10 0.022 d + 0.0071 10 0.008 d w

unfrozen silts-clay

(6.1a)

frozen silts-clay

(6.2a)

Table 6.1 Soil Thermal Conductivities


Thermal Conductivity,
Btu/hftF (W/mK)

Soil Moisture Content (by Mass)


Sand

Silt

Clay

Low, <4%

0.17 (0.29)

0.08 (0.14)

0.08 (0.14)

Medium, 4% to 20%

1.08 (1.87)

0.75 (1.30)

0.58 (1.00)

High, >20%

1.25 (2.16)

1.25 (2.16)

1.25 (2.16)

Note: For consistency and simplicity, thermal conductivities in I-P units are given in Btu/hftF with all dimensions in feet (not the more traditional
Btuin./hft2F). By doing this, conversion factors within the equations themselves are not needed and in most instances the equations apply for
either I-P or SI units.

6.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

k s = 0.083 0.7 log w + 0.4 10 0.01 d


k s = 0.0063 10 0.013 d + 0.0027 10 0.0146 d w

unfrozen sand-gravel (6.3a)


frozen sand-gravel

(6.4a)

where
ks = thermal conductivity, Btu/hftF
d = dry density of soil, lbm/ft3
w = moisture content of soil, % (dry basis)
In SI units (Farouki 1981):
k s = 0.1442 0.9 log w 0.2 10 0.6243 d

unfrozen silts-clay

k s = 0.001442 10 1.373 d + 0.01226 10 0.4994 d w frozen silts-clay


k s = 0.1442 0.7 log w + 0.4 10 0.6243 d
k s = 0.01096 10 0.8116 d + 0.00461 10 0.9115 d w

(6.1b)
(6.2b)

unfrozen sand-gravel (6.3b)


frozen sand-gravel

(6.4b)

where
ks = thermal conductivity, W/mK
d = dry density of soil, g/cm3
w = moisture content of soil, % (dry basis)
These equations apply to moisture contents of 7% or more for silts and clays and 1%
or more for sands and gravels. Soils with more than 50% silt and clay fall into the finegrained group. For sandy soil with relatively high silt and clay content (i.e., 40%), Kersten (1949) suggests that the average of the fine-gained and coarse-grained equations
would provide a reasonable prediction. If applied judiciously, Kersten suggests that the
resulting thermal conductivity predictions from Equations 6.1 through 6.4 should be
within 25%.

Temperature Effects on Soil Thermal Conductivity and Frost Depth


It should also be noted that the thermal conductivity of soils, like most materials, is a
function of temperature. Kersten (1949) provides data for the soils he studied; however,
normally for moderate temperature ranges near normal ambient temperature it is not
necessary to compensate for this effect. What must not be ignored, however, is the dramatic increase in thermal conductivity between unfrozen and frozen soils when there is
significant moisture present. While frozen soils will not normally be found in the immediate vicinity of buried district heating pipelines in all but the coldest climates, such is
not the case for district cooling systems in areas of seasonal frost and permafrost. In
such climates, consideration must be given to freeze protection for chilled-water piping
that is not in operation during the periods when frost penetration is expected to reach the
burial depth of the pipelines. The calculation of frost depth is beyond the scope of this
book; the reader is referred to Lunardini (1981), ASCE (1996), or Andersland and
Ladanyi (2004). Furthermore, the reader is cautioned against using overly generalized
frost depth maps as such maps fail to capture microclimatic variations or factors such

6.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

as surface type, soil type, and soil moisture content, which can all have very substantial
impacts on frost depth.

Specific Heat of Soils


The specific heat of soils with moisture in them is a function of the soil solids and
water content and the specific heat of those two components in the soil/water system. The
following equation is a commonly used simple proration of these impacts:
w
c s = c ss + c w ---------
100
where
cs =
css =
cw =
w =

(6.5)

specific heat of soil and water mixture, Btu/lbmF (kJ/kgK)


specific heat of dry soil solids, Btu/lbmF (kJ/kgK)
specific heat of water, Btu/lbmF (kJ/kgK)
moisture content of soil, % (dry basis)

The specific heat of liquid water, cw , can be taken as 1 Btu/lbmF (4.18 kJ/kgK) for
our purposes here. Because the specific heat of dry soil is nearly constant for all types of
soil, css may be taken as 0.175 Btu/lbmF (0.73 kJ/kgK).

Example 6.1: Soil Thermal Property Calculations


As an example on the use of the soil thermal conductivity equations (Equations 6.16.4), consider the case of an unfrozen sandy soil with a moisture content of 10% at a dry density of 100 lbm/
ft3 (1.60 g/cm3).
Solution:
Working the problem in I-P units using Equation 6.3a:
k s = 0.083 0.7 log w + 0.4 10 0.01 d = 0.083 0.7 log 10 + 0.4 10 0.01 100 = 0.91 Btu/hftF
Repeating this calculation using SI units with Equation 6.3b:
k s = 0.1442 0.7 log w + 0.4 10 0.6243 d = 0.1442 0.7 log 10 + 0.4 10 0.6243 1.60 = 1.58 W/mK
Calculating the specific heat for this same soil using Equation 6.5 yields:
w
10
c s = c ss + c w --------- = 0.175 + 1.0 --------- = 0.275 Btu/lb m F
100
100

(I-P)

w
10
c s = c ss + c w --------- = 0.73 + 4.18 --------- = 1.15 kJ/kgK
100
100

(SI)

The calculation of the thermal diffusivity using the thermal conductivity, the specific heat, and
the density is illustrated in Example 6.2.

6.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

UNDISTURBED SOIL TEMPERATURES


Before any heat loss/gain calculations may be conducted, the undisturbed soil temperature at the site must be determined. The choice of soil temperature is guided primarily
by the type of calculation being conducted; see the previous Thermal Design Considerations section. For example, if the purpose of the calculation is to determine if a material
within the piping system will exceed its temperature limit, the maximum expected undisturbed ground temperature is used. The appropriate choice of undisturbed soil temperature also depends on the location of the site, time of year, depth of burial, and thermal
properties of the soil. Some methods for determining undisturbed soil temperatures and
suggestions on appropriate circumstances to use them are as follows.
Use the average annual air temperature to approximate the average annual soil
temperature. This estimate is appropriate when the objective of the calculation is
to yield the average heat loss over the yearly weather cycle. Average annual air
temperatures may be obtained from various sources of climatic data such as
ASHRAE HandbookFundamentals (2009), which contains data from weather
stations worldwide (the data are on the CD distributed with the Handbook;
Chapter 14 of the Handbook provides the explanatory notes).
Use the maximum/minimum air temperature as an estimate of the maximum/
minimum undisturbed soil temperature for pipes buried at a shallow depth. This
approximation is an appropriate conservative assumption when checking the
temperatures to determine if the temperature limits of any of materials proposed
for use will be exceeded. Maximum and minimum expected air temperatures
may be found in ASHRAE HandbookFundamentals (2009).
For systems that are buried at other than shallow depths, maximum/minimum
undisturbed soil temperatures may be estimated as a function of depth, soil thermal properties, and prevailing climate. This estimate is appropriate, for example,
when checking the temperatures in a system to determine if the temperature limits of any of the materials proposed for use will be exceeded. The following
equations may be used to estimate the minimum and maximum expected undisturbed soil temperatures.
For maximum temperature:

T s z = T ms + A s

z ----
e

(6.6)

z -----

(6.7)

For minimum temperature:

T s z = T ms A s e
where
Ts,z =
z =
=
=
Tms =
As =

temperature, F (C)
depth, ft (m)
annual period length, 365 days
thermal diffusivity of the soil, ft2/day (m2/day)
mean annual surface temperature, F (C)
surface temperature amplitude, F (C)

6.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Values for the climatic constants Tms and As have been computed for all
5564 weather stations, domestic and international, listed in Chapter 14 of
ASHRAE HandbookFundamentals (2009) and are available at http://tc62
.ashraetcs.org/pdf/ASHRAE_Climatic_Data.pdf. Appendix B of this book also
provides a method to find the climatic constants given observed climatic data,
which may be found from many sources or may be fabricated to represent some
special set of circumstances.
Thermal diffusivity for soil may be calculated as follows:
24k
= -----------s
s cs

(I-P)

(6.8a)

86.4k
= ----------------s
s cs

(SI)

(6.8b)

where s is the soil density (lbm/ft3 [kg/m3]).


For instances where a specific temperature other than the maximum or minimum is needed, the undisturbed soil temperatures may be estimated for any time
of the year as a function of depth, soil thermal properties, and prevailing climate.
This specific temperature, as well as those calculated with Equations 6.6 and
6.7, may be used in lieu of the soil surface temperature normally called for by
the steady-state heat transfer equations when estimates of heat loss/gain as a
function of time of year are desired. The substitution of the undisturbed soil
temperatures at the pipe depth allows the steady-state equations to be used as a
first approximation to the solution to the actual transient heat transfer problem
with its annual temperature variations at the surface. Equation 6.9 may be used
to estimate the undisturbed soil temperature at any depth at any point during the
yearly weather cycle (ASCE 1996).

T s z = T ms + A s

z ----2 t t lag

e
sin ---------------------------

-
z ----

(6.9)

Note: argument for sin function is in radians.


where
t
= Julian date, days
tlag = phase lag of soil surface temperature, days
Values for the climatic constants Tms, As, and tlag have been computed for
all 5564 locations, domestic and international, listed in Chapter 14 of ASHRAE
HandbookFundamentals (2009) and are available at http://tc62.ashraetcs.org/
pdf/ASHRAE_ Climatic_Data.pdf. These constants were derived by performing
a best fit to the average monthly temperatures using the least squares method.
Appendix B of this book provides a method to find the climatic constants given
an annual set of observed or fabricated climatic data.
Equation 6.9 does not account for latent heat effects due to freezing, thawing, or evaporation. However, for the soil adjacent to a buried heat distribution
system the equation provides a good estimate since the heat losses from the system tend to prevent the adjacent ground from freezing. For buried chilled-water

6.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

systems, freezing may be a consideration; thus, systems that are not used or
drained during the winter months should be buried below the seasonal frost
depth. For systems that are in use but are buried above the frost depth, freezing
may still occur if flow is low or nonexistent in portions of the system and/or the
burial depth is shallow when compared to the frost depth.
In the calculation of the climatic constants of Appendix B, the ground surface temperature is assumed to equal the air temperature, which is an acceptable assumption for
most design calculations. If a more accurate calculation is desired, the methods presented
in the Heat Transfer at Ground Surface section may be used to compensate for the convective thermal resistance to heat transfer at the ground surface and the impacts of the
type of surface cover.

Heat Transfer at Ground Surface


Heat transfer between the ground surface and the ambient air occurs by convection.
Heat transfer also takes place as a result of incident precipitation and radiation. The heat
balance at the ground surface is too complex to warrant detailed treatment in the design of
buried district heating and cooling systems; the interested reader is referred to Lunardini
(1981).
As a first approximation for the convective heat transfer, an effectiveness thickness of
a fictitious soil layer may be added to the burial depth to account for the effect of the convective heat transfer resistance at the ground surface. The effective thickness is calculated
as follows:
k
= -----s
h

(6.10)

where
= effectiveness thickness of fictitious soil layer, ft (m)
h = convective heat transfer coefficient at ground surface, Btu/hft2F (W/m2K)
The effective thickness calculated with Equation 6.10 is simply added to the actual
burial depth of the pipes in calculating the soil thermal resistance using the equations presented in the section of this chapter entitled Steady State Heat Loss/Heat Gain Calculations for Systems.
The surface type (i.e., asphalt, concrete, grass) can have a large impact on the heat
balance at the grounds surface and the resulting soil temperatures below. The type of surface impacts the heat transfer from radiation, convection, and precipitation. The impacts
are well known. McCabe et al. (1995) observed significant temperature variations due to
the type of surfaces and predicted significant impacts for district cooling systems. While
there has not been any detailed study beyond the work of McCabe et al. of the impacts of
surface type on soil temperatures surrounding district heating and cooling systems specifically, there has been significant study of the impacts of surface type on soil freezing and
thawing. For this application, a method of adjusting the air temperature to find an effective surface temperature has been developed. This method is referred to as the n-factor
method, with n-factors having been determined empirically by a number of investigators.
Because the impacts of solar radiation in particular are so important, n-factors have been
developed for the summer (thawing) and winter (freezing) seasons, and these factors vary
appreciably with surface and climate types. For more discussion of n-factors the reader is

6.7

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

referred to Lunardini (1978, 1981), who explains the theory and tabulates the values for
the n-factor. Freitag and McFadden (1997) also supply tabulated values of n-factors.
The reader is cautioned that when using n-factors one must recognize that they are
not only specific to the surface but they are also site-specific and thus one should only
extrapolate with caution and understanding. With that caveat as a first approximation,
lacking other data, the n-factor method can be used to estimate soil temperatures beneath
various surfaces. An example of the impact is provided in Example 6.2.

Example 6.2: Soil Temperature Calculations


To illustrate the use of the equations presented in the Undisturbed Soil Temperatures section of
this design guide, consider the most general case of determining soil temperatures as a function of
both time and depth using Equation 6.9. Well assume a climate typical of coastal Massachusetts:
T ms = 49.0F (9.5C)
A s = 20.6F (11.4 K)
t lag = 115.9 days
Solution:
For soil well use the same unfrozen sandy soil of Example 6.1 with a moisture content of 10%
at a dry density of 100 lbm/ft3 (1600 kg/m3), which yielded the following thermal properties:
k s = 0.91 Btu/hftF (1.58 W/mK)
c s = 0.275 Btu/lb m F (1.15 kJ/kgK)
Working the problem in I-P units using Equation 6.8a, the thermal diffusivity is:
24k
24 0.91
= -----------s = --------------------------- = 0.79 ft 2 day
100 0.275
s cs
Repeating this calculation using SI units with Equation 6.8b:
86.4k
86.4 1.58
= ----------------s = --------------------------- = 0.074 m 2 day
1600 1.15
s cs
Substituting these thermal property and climatic constant values into Equation 6.9 we have:*

T s z = 49.0 + 20.6e

z ----------------------2 t 115.9 -
0.79 365 sin --------------------------------

365

-
z ----------------------0.79 365

= 49.0 + 20.6e 0.10z sin 0.0172 t 115.9 0.10z


With this result we can now evaluate the soil temperature for any depth z (ft) and time t
(Julian day). A series of calculations have been made using this result; they are presented in
* In examples for this chapter where the equation is the same for I-P and SI units, the example will be worked in I-P
units and the result provided in both I-P and SI units. In the few cases where the equation is different for I-P and SI
units, the example will be worked in both sets of units.

6.8

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

Figure 6.1 Soil temperatures calculated with Equation 6.9 for a coastal Massachusetts
climate and the assumption that surface temperature is equal to air temperature.

Figure 6.1. These calculations are based on the assumption that the ground surface temperature is
equal to air temperature. To adjust for a convective coefficient, the value calculated by Equation
6.10 would simply be added to the burial depth of interest to calculate the appropriate depth value
in Figure 6.1.
To illustrate the use of the n-factor concept, Figure 6.2 has been prepared in a manner similar to
Figure 6.1. The same climate and soil have been assumed as for the calculation of Figure 6.1. A
surface of concrete pavement has been assumed and the n-factors have been estimated based on the
data provided by Lunardini (1981) as 0.66 during the freezing season and 1.7 during the thawing
season. The calculation is somewhat complex to detail here, but to summarize it proceeds by first
calculating surface temperatures using the air temperatures calculated by Equation 6.9 at zero
burial depth and the assumed n-factors. Subsequently, a sinusoidal curve is fitted to these surface
temperatures using the method in Appendix B. The constants from that sinusoidal curve fit are then
used in Equation 6.9 as before, noting that no adjustment is made to depths for the convective coefficient at the surface, as this impact has been included in the n-factor. The use of n-factors for this
purpose is a significant extrapolation of the technique discussed previously in this section and thus
these results should be taken as very approximate. That having been said, by comparing Figures 6.1
and 6.2 we can see that at 3 ft (0.91 m) of depth the highest temperature reached in the summer
under the concrete pavement is predicted to be about 23F (13C) greater than in our calculation
that ignored any surface type impacts and assumed the air temperature and surface temperature
were equal.
It is interesting to compare the results of this approximation method with the measurements of
McCabe et al. (1995), who found peak summer temperatures under pavement of approximately
82F (28C) at a 3 ft (0.91 m) depth. Using the n-factor method described previously with climatic
constants calculated with the method of Appendix B for the Ithaca, New York, area where the

6.9

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 6.2 Soil temperatures calculated with Equation 6.9 for a coastal Massachusetts
climate and the use of n-factors to adjust for a concrete pavement surface.

measurements of McCabe et al. were made, the peak ground temperature under a concrete pavement at a depth of 3 ft (0.91 m) is predicted to be 85F (29C). This is considered reasonable agreement given the approximate nature of the method outlined here as well as the difficulty in making
measurements of soil temperatures. Clearly, as McCabe et al. point out, consideration should be
given to surface impacts on subsurface soil temperatures when making calculations to determine
appropriate insulation thickness. In addition, other impacts such as those on the materials within
chilled-water, hot-water, and steam distribution systems should be considered.
Accurate undisturbed soil temperatures are much more of a concern in district cooling system
design than for district heating design, as for district cooling systems the temperatures of the carrier
fluid are much closer to the temperatures of the undisturbed soil. Thus, errors of similar magnitude
in undisturbed soil temperature estimation will result in much larger errors in estimated heat gain
for a district cooling system than for heat losses from a district heating system. Consider, for example, the peak heat gains from a 40F (4.4C) chilled-water supply pipe buried at 3 ft (0.91 m) in the
coastal Massachusetts climate used in Example 6.2. Peak temperature at that depth is estimated at
64F (18C) when the surface heat transfer impacts are excluded and 87F (31C) when the estimated impacts of a concrete surface pavement are included. The heat gains for the 87F (31C)
undisturbed soil temperature are 1.96 times greater [(87 40)/(64 40)] than those for the 64F
(18C) soil temperature. As an example of how important this difference is, imagine the lower
ground temperature had been used and the economic insulation thickness (discussed later) had been
determined with the assumption of the lower ground temperature; in such a case it is likely that the
resulting insulation thickness would not be valid for the higher ground temperature and more insulation would be indicated. A similar calculation for a low-temperature hot-water supply pipe operating at 250F (121C) shows that the heat loss would only be reduced by 12% under the same
assumptions.

6.10

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

INSULATION TYPES AND THERMAL PROPERTIES


Insulation provides the primary thermal resistance against heat loss or gain in thermal
distribution systems. Thermal properties and other characteristics of insulations normally
used in thermal distribution systems are listed in Table 6.2. The material properties provided in Table 6.2 should only be used in the absence of data specific to the actual insulation material to be used/in use. Material properties such as thermal conductivity, density,
compressive strength, moisture absorption, dimensional stability, and combustibility are
typically reported in ASTM standards for the respective material. Some properties have
more than one associated standard. For example, thermal conductivity for insulation
material in block form may be reported using ASTM C177 (2010a), C518 (2010b), or
C1114 (2006). Thermal conductivity for insulation material fabricated or molded for use
on piping is reported using ASTM C335/C335M (2010c).
Table 6.2 contains the most common insulations used in the construction of district
heating and cooling piping systems. However, from time to time practitioners may
encounter insulation materials that are in essence now obsolete, yet their thermal properties values are still needed. An example of such an instance is calculating the heat loss
Table 6.2 Comparison of Commonly Used Insulations in Underground Piping Systems
Calcium Silicate
Type I/II,
ASTM C533 (2013)

Urethane
Foam

Cellular Glass,
ASTM C552 (2012a)

Mineral Fiber/
Preformed Glass
Fiber Type 1,
ASTM C547 (2012b)

Thermal Conductivity,a I-P Units (Btu/hftF)


The values in parentheses are the maximum permissible by the ASTM standard listed.
Mean temperature = 100F

0.028

0.013

0.033 (0.030)

0.022 (0.021)

200F

0.031 (0.038/0.045)

0.014

0.039 (0.037)

0.025 (0.026)

300F

0.034 (0.042/0.048)

0.046 (0.045)

0.028 (0.033)

400F

0.038 (0.046/0.051)

0.053 (0.054)

(0.043)

Thermal Conductivity,a SI Units (W/mK)


The values in parentheses are the maximum permissible by the ASTM standard listed.
Mean temperature = 40C

0.048

0.022

90C

0.054 (0.066/0.078)

0.024

150C

0.059 (0.073/0.083)

200C

0.057 (0.052)

0.038 (0.036)

0.067 (0.064)

0.043 (0.045)

0.080 (0.078)

0.048 (0.057)

0.066 (0.080/0.088)

0.092 (0.093)

(0.074)

Maximum density, lb/ft3 (kg/m3)

15/22 (240/360)

6.79.2 (107147)

811 (128178)

Maximum temperature, F (C)

1200 (650)

Compressive strength (minimum),b kPa


Dimensional stability, linear shrinkage
at maximum use temperature
Flame spread
Smoke index
Water absorption

800 (430)

850 (450)

700 at 5%
deformation

450

N/A

2%

N/A

2%

25

50

0.5

Water vapor sorption,


5% maximum
(by weight)

As shipped moisture
content, 20%
maximum (by weight)

250 (120)

a. Data in this table are from Chapter 12 of ASHRAE HandbookHVAC Systems and Equipment (2012). Thermal conductivity of insulation may
vary with temperature, temperature gradient, moisture content, density, thickness, and shape. ASTM maximum values given are comparative
for establishing quality control compliance and are suggested for preliminary calculations where other more specific values are not available.
They may not represent installed performance of insulation under actual conditions that differ substantially from test conditions. The insulation
manufacturer should be able to furnish design values.
b. Compressive strength for cellular glass shown is for flat material, capped as per ASTM C240 (2012c).

6.11

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

from an existing system for purposes such as providing a baseline for the calculation of
payback on an insulation retrofit/upgrade. Ideally in such cases a sample would be
removed and thermal conductivity tests would be conducted on the sample. In the absence
of such testing, Table 6.3 provides a compilation of values for various obsolete piping
insulations compiled from the references cited in the footnotes of the table.
One of the greatest challenges in any system with thermal insulation is preventing it
from becoming wet. This is especially difficult in buried heating and cooling piping systems. All buried systems make some efforts, and many take extreme precautions, to try to
maintain the dryness of the thermal insulation. However, it can be expected that water
may find its way into the insulation in some cases. Thus, it is important that the consequences of wet thermal insulation be understood. Chyu et al. (1997a, 1997b, 1998a,
1998b) studied the effect of moisture on the thermal conductivity of insulating materials
commonly used in underground district energy systems (ASHRAE Research Project RP721). The results are summarized in Table 6.4. The insulated pipe was immersed in water
maintained at 46F to 100F (8C to 38C) to simulate possible conduit water temperatures during a failure. The fluid temperature in the insulated pipe ranged from 35F to
450F (2C to 232C). All insulation materials were tested unfaced and/or unjacketed to
simulate installation in a conduit.
Painting or coating of chilled-water piping prior to insulating is recommended for
aboveground installations in areas of high humidity and all belowground insulations.
Insulations used today for chilled water include polyurethane and polyisocyanurate cellular plastics, phenolics, and fiberglass. With the exception of fiberglass, the rest can form
acidic solutions (pH 23) once they hydrolyze in the presence of water. The acids emanate from the chlorides, sulfates, and halogens added in the manufacturing process to
increase fire retardancy or expand the foam. Phenolics can be more than six times more
corrosive than polyurethane due to the acids used in their manufacture and can develop
environments to pH 1.8. The easiest way of mitigating corrosion is to paint the pipe exterior with a strong rust-preventative coating (two-coat epoxy) prior to insulating. This is
good engineering practice and most insulation manufacturers will suggest this, but it may
not be in their literature. For aboveground installations, a high-integrity vapor barrier is
also paramount in order to minimize the amount of moisture migrating into the insulation
and to the pipe surface. Belowground systems also require a high-integrity vapor barrier
formed by a jacket that will also provide protection from the rigors of burial. More information on piping systems and the methods used to protect the insulation from water
ingress are provided in this book in Chapter 4.

STEADY-STATE HEAT LOSS/HEAT GAIN CALCULATIONS FOR SYSTEMS


This section presents the formulas necessary to calculate steady-state heat losses from
piping system geometries typically used in district heating and cooling systems. The most
important factors affecting heat transfer are the difference between earth and fluid temperatures and the thermal insulation. Other factors that affect heat transfer are (1) depth of
burial, related to the earth temperature and soil thermal resistance; (2) soil thermal conductivity, related to soil moisture content and density; and (3) distance between adjacent
pipes.
For complex geometries and to compute transient heat gains or losses in underground
piping systems, numerical methods that may approximate any physical problem and
include such factors as the effect of temperature on thermal properties provide the most
accurate results. For most designs, numerical analyses are probably not warranted. While
a number of commercial computer-aided design (CAD) programs are available for
numerical heat transfer, few are able to model the true complexities of heat transfer in soil

6.12

1000
(538)

500
(260)
1600
(871)

16
(256)
1721
(272336)
3040
(480640)
22
(352)
2426
(384416)

Molded asbestosb

Laminated asbestosa
approximately
20 laminations per inch

Laminated asbestosa
approximately
3040 laminations per inch

Diatomaceous silicab

Diatomaceous earth
with asbestosa

a. King and S. Crocker (1967)


b. Strock and Koral (1965)

400
(204)

21
(336)

Corrugated asbestosa
8 plies per inch

16001900
(8711038)

500
(260)

400
(204)

1115
(176240)

Corrugated asbestosa
4 plies per inch

600
(316)

1517
(240272)

85% magnesiaa

Material

Approximate Approximate
Density,
Temperature
lb/ft3
Limit,
F (C)
(kg/m3)

0.052
(0.090)

0.030
(0.052)

0.031
(0.053)

0.028
(0.048)

0.042
(0.073)

0.041
(0.071)

0.033
(0.057)

100
(38)

0.055
(0.095)

0.035
(0.061)

0.036
(0.062)

0.032
(0.055)

0.050
(0.087)

0.052
(0.090)

0.036
(0.062)

200
(93)

0.058
(0.100)

0.039
(0.067)

0.046
(0.080)

0.036
(0.062)

0.058
(0.100)

0.062
(0.107)

0.040
(0.069)

300
(149)

0.061
(0.106)

0.053
(0.092)

0.044
(0.076)

0.052
(0.090)

0.040
(0.069)

0.066
(0.114)

0.073
(0.126)

0.043
(0.074)

400
(204)

0.064
(0.111)

0.055
(0.095)

0.049
(0.085)

0.059
(0.102)

0.044
(0.076)

0.046
(0.080)

500
(260)

0.067
(0.116)

0.057
(0.099)

0.048
(0.083)

0.050
(0.087)

600
(316)

Thermal Conductivity, Btu/hftF (W/mK),


at the Following Mean Temperatures, F (C)

Table 6.3 Thermal Properties of Obsolete Pipe Insulation Materials

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

0.070
(0.121)

0.058
(0.100)

700
(371)

0.072
(0.125)

0.060
(0.104)

800
(427)

6 Heat Transfer Calculations for Piping Systems

6.13

6.14
Cellular Glass

Mineral Woolb

Pipe at 420F (215C),


8 hours
Pipe temperature 36F,
water bath 46F to 58F
(Pipe temperature 2.2C,
water bath 8C to 14C)

Pipe temperature 37F,


water bath 52F
(Pipe temperature 2.8C,
water bath 11C)

Conduction

2 to 4 times; water absorption


minimal, ceased after 7 days
Conduction
Pipe at 37F (2.7C), data curve
extrapolated to 10+ days

Effective k-value increase from dry


conditions after steady state
achieved

Primary heat transfer mechanism

Length of time for specimen to


return to dry steady-state k-value
after submersion

Pipe at 35F (1.7C), data curve


extrapolated to 25 days

Conduction and convection

14 times; insulation completely


saturated at steady state

6 days

Pipe temperature 35F to 45F,


water bath 55F
(Pipe temperature 1.8C to 7C,
water bath 13C)

Pipe at 450F (230C),


9 days

Pipe at 35F (1.7C),


15 days

Conduction and convection

20 times; insulation completely


saturated at steady state

1/2 hour

Pipe temperature 39F to 43F,


water bath 51F to 59F
(Pipe temperature 3.9C to 6.1C,
water bath 10.6C to 15C)

Pipe at 380F (193C),


6 days

Conduction and convection

52 to 185 times; insulation


completely saturated at
steady state

2 hours

a. Polyurethane material tested had a density of 2.9 lb/ft3 (46 kg/m3).


b. Mineral wool tested was a preformed molded basalt designed for pipe systems operating up to 1200F (650C). It was specially formulated to withstand the Federal Agency Committee 96-hour boiling water test (see Appendix A).
c. Cracks formed during heating for all samples of cellular glass insulation tested. Flooded heat loss mechanism involved dynamic two-phase flow of water through cracks; the period of dynamic process was about 20 min. Cracks had a negligible effect on the thermal conductivity of dry cellular glass insulation before and after submersion. No cracks formed during the cooling test.

Pipe at 33F (0.6C),


no change

None; no water penetration

16 days

Length of submersion time to


reach steady-state conditions for
k-value

Cooling Test

Data recorded at 4 days


constant at 12 days

See Note C

Conduction
Pipe at 260F (127C), after 16
days moisture content 10% (by
volume) remaining

Primary heat transfer mechanism

Length of time for specimen to


return to dry steady-state k-value
after submersion

Conduction and convection

Up to 50 times at steady state;


insulation completely saturated

Average 10 times, process


unsteady (Note C); insulation
showed evidence of moisture
zone on inner diameter

14 to 19 times at steady state;


estimated water content of
insulation 70% (by volume)

70 days

10 days

Effective k-value increase from dry


conditions after steady state
achieved in submersion

Fibrous Glass

Pipe temperature 35F to 260F, Pipe temperature 35F to 420F, Pipe temperature 35F to 450F, Pipe temperature 35F to 450F,
water bath 46F to 100F
water bath 46F to 100F
water bath 46F to 100F
water bath 46F to 100F
(Pipe temperature 1.8C to 127C, (Pipe temperature 1.8C to 215C, (Pipe temperature 1.8C to 230C, (Pipe temperature 1.8C to 230C,
water bath 8C to 38C)
water bath 8C to 38C)
water bath 8C to 38C)
water bath 8C to 38C)

Polyurethanea

See Note Cc

Length of submersion time to


reach steady-state k-value

Heating Test

Characteristic

Table 6.4 Effect of Moisture on Underground Piping System Insulations


(Chyu et al. 1997a, 1997b, 1998a, 1998b)

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

systems, including moisture migration, freezing/thawing, and the surface heat balance.
Thus, the use of numerical methods is seldom justified given the difficulty or inability to
accurately model heat transfer in soils. This is especially true when taking into consideration the uncertainty that normally exists in parameters such as the soil thermal properties
that, as noted previously, may only be known to within 25% even when the soil type,
moisture content, and density are well known and homogeneous, which is seldom the
case. The steady-state methods presented here are based on the electrical resistance analogy to heat transfer and are normally adequate for design calculations. Approximate analytical methods for transient calculations may be of use under special circumstances
encountered in system operation; some of these methods are contained in the section Calculating Temperatures of System Components and Surrounding Soil Temperatures later
in this chapter.
Steady-state calculations are appropriate for determining the annual heat loss/gain
from a buried system if average annual earth temperatures are used. Steady-state calculations may also be appropriate for worst-case analyses of thermal effects on materials.
Steady-state calculations for a one-pipe system may be done without a computer, but calculations become increasingly difficult for two-, three-, and four-pipe systems. Software
called HeatConduitExpert is available that helps simplify many of the more complicated
calculations outlined in the following subsections (Meyer 2013).
The steady-state methods of analysis discussed in the following subsections use resistance formulations developed by Phetteplace and Meyer (1990) that simplify the calculations needed to determine temperatures within the system. Each type of resistance is
given a unique subscript and is defined only when introduced. In each case, the resistances are on a unit length basis so that heat flows per unit length result directly when the
temperature difference is divided by the resistance.**

Single Buried Uninsulated Pipe


For the case of a single uninsulated buried pipe (Figure 6.3), an approximation for the
soil thermal resistance has been used extensively. This approximation is sufficiently accurate (within 1%) for the ratios of burial depth to pipe radius indicated next to
Equations 6.11 and 6.12. Both the actual resistance and the approximate resistance are
presented, along with the depth/radius criteria for each.
ln d r o + d r o 2 1 1 / 2
R s = ---------------------------------------------------------------------------2k s
ln 2d r o
R s = ------------------------2k s
where
Rs =
ks =
d =
ro =
*

for d r o 2

for d r o 4

(6.11)

(6.12)

thermal resistance of soil, hftF/Btu (mK/W)


thermal conductivity of soil, Btu/hftF (W/mK)
burial depth to centerline of pipe, ft (m)
outer radius of pipe or conduit, ft (m)

Note: For consistency and simplicity, thermal conductivities in I-P units are given in Btu/hftF with all
dimensions in feet (not the more traditional Btuin./hft2F). By doing this, conversion factors within the
equations themselves are not needed and in most instances the equations apply for either I-P or SI units.

6.15

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 6.3 Single uninsulated buried pipe.


Courtesy of ASHRAE (2012, Chapter 12, Figure 6)

Include the thermal resistance of the pipe if it is significant when compared to the soil
resistance. The thermal resistance of a pipe or any concentric circular region is given by
ln r o r i
R p = ---------------------2k p

(6.13)

where
Rp = thermal resistance of pipe wall, hftF/Btu (mK/W)
kp = thermal conductivity of pipe, Btu/hftF (W/mK)
ri = inner radius of pipe, ft (m)

Example 6.3
Consider an uninsulated 3 in. (75 mm) Schedule 40 polyvinyl chloride (PVC) chilled-water
supply line carrying 45F (7C) water. Assume the pipe is buried 3 ft (0.91 m) deep in soil with a
thermal conductivity of 1 Btu/hftF (1.7 W/mK) and no other pipes or thermal anomalies are
within close proximity. Assume the average annual soil temperature is 60F (16C).
ri = 1.54 in. = 0.128 ft (0.0390 m)
ro = 1.75 in. = 0.146 ft (0.0445 m)
d = 3 ft (0.91 m)
ks = 1 Btu/hftF (1.7 W/mK)
kp = 0.10 Btu/hftF (0.17 W/mK)
Solution:
Calculate the thermal resistance of the pipe using Equation 6.13:
Rp = 0.21 hftF/Btu (0.12 mK/W)

6.16

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

Calculate the thermal resistance of the soil using Equation 6.12. (Note: d/ro = 21 is greater
than 4; thus, Equation 6.12 may be used in lieu of Equation 6.11.)
Rs = 0.59 hftF/Btu (0.34 mK/W)
Calculate the rate of heat transfer by dividing the overall temperature difference by the total
thermal resistance:
t f ts
45 60
q = -------------- = -------------------------------------- = 19 Btu/hft (18 W/m)
0.80 hftF/Btu
Rt
where
Rt = total thermal resistance (i.e., Rs + Rp in this case of pure series heat flow), hftF/Btu
(mK/W)
tf = fluid temperature, F (C)
ts = average annual soil temperature, F (C)
q = heat loss or gain per unit length of system, Btu/hft (W/m)
The negative result indicates a heat gain rather than a loss. Note that the thermal resistance of the
fluid/pipe interface has been neglected. This is a reasonable assumption as such resistances tend to be
very small for flowing fluids. Also note that in this case the thermal resistance of the pipe is a significant portion of the total thermal resistance. This results from the relatively low thermal conductivity of
PVC compared to other piping materials and the fact that no other major thermal resistances exist in
the system to overshadow it. If any significant amount of insulation were included in the system, its
thermal resistance would dominate and it might be possible to neglect that of the piping material.

Single Buried Insulated Pipe


Equation 6.13 can be used to calculate the thermal resistance of any concentric circular region of material, including an insulation layer. When making calculations using
insulation thickness, actual thickness rather than nominal thickness should be used to
obtain the most accurate results.

Example 6.4
Consider the effect of adding 1 in. (25 mm) of urethane foam insulation and a 1/8 in. (3 mm)
thick polyvinyl chloride (PVC) jacket to the chilled-water line in Example 6.3.
Solution:
Calculate the thermal resistance of the insulation layer from Equation 6.13 as follows:
ln 0.229 0.146
R i = ----------------------------------------- = 5.75 hftF/Btu (3.32 mK/W)
2 0.0125
For the PVC jacket material, use Equation 6.13 again:
ln 0.240 0.229
R j = ----------------------------------------- = 0.07 hftF/Btu (0.04 mK/W)
2 0.10

6.17

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

The thermal resistance of the soil as calculated by Equation 6.12 decreases slightly to Rs =
0.51 hftF/Btu (0.29 mK/W) because of the increase in the outer radius of the piping system. The
total thermal resistance is now:
R t = R p + R i + R j + R s = 0.21 + 5.75 + 0.07 + 0.51 = 6.54 hftF/Btu (3.78 mK/W)
The heat gain by the chilled-water pipe is reduced to about 2 Btu/hft (2 W/m). In this case, the
thermal resistance of the piping material and the jacket material could be neglected with a resultant
error of <5%. Considering that the uncertainties in the material properties are likely greater than
5%, it is usually appropriate to neglect minor resistances such as those of piping and jacket materials when insulation is present.

Single Buried Pipe in Conduit with Air Space


Systems with air spaces (Figure 6.4) may be treated by adding an appropriate resistance for the air space. For simplicity, assume a heat transfer coefficient of 3 Btu/hft2F
(17 W/m2C) applies in most cases, based on the outer surface area of the insulation. The
resistance due to this heat transfer coefficient is then:
Ra = 1/(3 2 roi) = 0.053/roi

(I-P)

(6.14a)

Ra = 1/(17 2 roi) = 0.0094/roi

(SI)

(6.14b)

where
Ra =

resistance of air space, hftF/Btu (mK/W)

roi =

outer radius of insulation, ft (m)

Figure 6.4 Single insulated buried pipe with air space.


Courtesy of ASHRAE (2012, Chapter 12, Figure 7)

6.18

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

A more precise value for the resistance of an air space can be developed with empirical relations available for convection in enclosures such as those given by Grober et al.
(1961). The effect of radiation within the annulus should also be considered when high
temperatures are expected within the air space. For the treatment of radiation, refer to Siegel and Howell (1981).

Example 6.5
Consider a 6 in. (150 mm) nominal diameter (6.625 in. [168 mm] outer diameter) high-temperature water (HTW) line operating at 375F (190C). Assume the pipe is insulated with 2.5 in.
(63.5 mm) of mineral wool with a thermal conductivity ki = 0.026 Btu/hftF (0.045 W/mK) at
200F (93C) and ki = 0.030 Btu/hftF (0.052 W/mK) at 300F (149C).
The pipe will be encased in a steel conduit with a concentric air gap of 1 in. (25 mm). The steel
conduit will be 0.125 in. (3.2 mm) thick and will have a corrosion-resistant coating approximately
0.125 in. (3.2 mm) thick. The pipe will be buried 4 ft (1.2 m) deep to pipe centerline in soil with an
average annual temperature of 60F (16C). The soil thermal conductivity is assumed to be 1 Btu/
hftF (1.7 W/mK). The thermal resistances of the pipe, conduit, and conduit coating will be
neglected.
Solution:
Calculate the thermal resistance of the pipe insulation. Assume a mean temperature of the insulation of 250F (121C) to establish its thermal conductivity, which is equivalent to assuming the
insulation outer surface temperature is 125F (52C). Interpolating the data listed previously, the
insulation thermal conductivity ki = 0.028 Btu/hftF (0.048 W/mK). Then calculate the insulation
thermal resistance using Equation 6.13:
ln 0.484 0.276
R i = ----------------------------------------- = 3.19 hftF/Btu (1.84 mK/W)
2 0.028
Calculate the thermal resistance of the air space using Equation 6.14:
R a = 0.053 0.484 = 0.11 hftF/Btu (0.064 mK/W)
Calculate the thermal resistance of the soil from Equation 6.12:
ln 8.0 0.589
R s = ----------------------------------- = 0.42 hftF/Btu (0.24 mK/W)
2 1
The total thermal resistance is:
R t = R i + R a + R s = 3.19 + 0.11 + 0.42 = 3.72 hftF/Btu (2.15 mK/W)
The first estimate of the heat loss is then:
q = 375 60 3.72 = 84.7 Btu/hft (81.4 W/m)
Now repeat the calculation with an improved estimate of the mean insulation temperature.
Begin by calculating the insulation outer surface temperature:
t io = t po qR i = 375 84.7 3.19 = 105F (41C)
where tpo is the outer surface temperature of pipe, F (C).

6.19

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

The new estimate of the mean insulation temperature is 240F (116C), which is close to the
original estimate. Thus, the insulation thermal conductivity changes only slightly, and the resulting
thermal resistance is Ri = 3.24 hftF/Btu (1.87 mK/W). The other thermal resistances in the system remain unchanged, and the heat loss becomes q = 83.8 Btu/hft (80.5 W/m). The insulation surface temperature is now approximately tio = 104 F (40C), and no further calculations are needed.

Single Buried Pipe with Composite Insulation


A number of systems are available that use more than one insulating material. The
motivation for doing so is usually the desire to use an insulation that has desirable thermal
properties or lower cost but is unable to withstand the service temperature of the carrier
pipe (e.g., polyurethane foam). Another insulation with acceptable service temperature
limits (e.g., calcium silicate or mineral wool) is normally placed adjacent to the carrier
pipe in sufficient thickness to reduce the temperature at its outer surface to within the limits of the insulation that is desirable to use. The calculation of the heat loss and temperatures within a composite system is straightforward using the equations presented thus far
as illustrated in Examples 6.6 and 6.7.

Example 6.6
The maximum operating temperature of the polyurethane foam in a system with composite
insulation is desired. Because it is the maximum operating temperature of the materials within the
system that is sought, we need to use the maximum expected soil temperature rather than the average annual soil temperature. We also need to assume the lowest anticipated soil thermal conductivity and the deepest burial depth to be encountered by the system. These conditions will result in the
maximum internal temperatures of the system components.
The system is to be installed in southern Texas and consists of a 6 in. (150 mm) nominal diameter (6.625 in. [168 mm] outer diameter) high-temperature water (HTW) line operating at 400F
(204C). The pipe is insulated with 1.5 in. (38 mm) of calcium silicate; in addition, on the exterior
of the calcium silicate insulation there is 1 in. (25 mm) of polyurethane foam insulation. The thermal conductivities of both insulations are provided in the table below. The polyurethane insulation
will be encased in a 0.5 in. (13 mm) thick fiberglass-reinforced plastic (FRP) jacket. The piping
system will be buried 10 ft (3 m) deep to pipe centerline. The soil is sandy and very dry with an
assumed moisture content of 1% by dry weight and a density of 95 lbm/ft3 (1520 kg/m3). The soil
thermal conductivity is found from Equation 6.3 as 0.30 Btu/hftF (0.52 W/mK). The thermal
resistances of the pipe and FRP jacket will be neglected.
Insulation Properties for Example 6.6

Calcium
Silicate

Polyurethane
Foam

6.20

Insulation Mean Temperature,


F (C)

Insulation Thermal Conductivity,


Btu/hftF (W/mK)

200 (93)

0.038 (0.066)

300 (149)

0.042 (0.073)

400 (204)

0.046 (0.080)

100 (38)

0.013 (0.022)

200 (93)

0.014 (0.024)

275 (135)

0.015 (0.026)

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

Solution:
First calculate the maximum soil temperature expected at the installation depth using
Equation 6.6. The values for climatic constants for all 5564 weather stations, domestic and international, listed in Chapter 14 of ASHRAE HandbookFundamentals (2009) are available at http://
tc62.ashraetcs.org/pdf/ASHRAE_Climatic_Data.pdf. The soil thermal diffusivity must first be estimated using Equation 6.8:
24k s
24 0.30
= --------------------------------------------------- = 0.41 ft 2 /day (0.038 m 2 /day)
= -------------------------------------95 0.175 + 1 100
s c s + w 100
Then Equation 6.6 is used to calculate the maximum soil temperature at the installation depth:
T s z = T ms = A s

z -----
e

= 71.8 + 15.0 exp 10 0.41 365 0.5 = 75.3F (24.1C)

Now calculate the first estimates of the thermal resistances of the pipe insulations. For the calcium silicate assume a mean temperature of the insulation of 300F (149C) to establish its thermal
conductivity. From the data listed previously, the calcium silicate thermal conductivity ki =
0.042 Btu/hftF (0.073 W/mK) at this temperature. For the polyurethane foam assume a mean
temperature of the insulation of 200F (93C). From the data listed previously, the polyurethane
foam thermal conductivity ki = 0.014 Btu/hftF (0.024 W/mK) at this temperature. Now the insulation thermal resistances are calculated using Equation 6.13:
ln 0.401 0.276
R i 1 = ----------------------------------------- = 1.42 hftF/Btu (0.821 mK/W)
2 0.042

(calcium silicate)

ln 0.484 0.401 - = 2.14 hftF/Btu (1.24 mK/W)


R i 2 = ---------------------------------------2 0.014

(polyurethane foam)

Calculate the thermal resistance of the soil from Equation 6.12:


ln 20 0.526
R s = ---------------------------------- = 1.93 hftF/Btu (1.11 mK/W)
2 0.30
The total thermal resistance is:
R t = R i 1 + R i 2 + R s = 1.42 + 2.14 + 1.93 = 5.49 hftF/Btu (3.17 mK/W)
The first estimate of the heat loss is then:
q = 400 75.3 5.49 = 59.1 Btu/hft (56.8 W/m)
Now that we have the first estimate of the heat loss we can calculate estimated insulation surface temperatures. First we find the temperature at the interface between the calcium silicate and
polyurethane foam insulations:
t io 1 = t po qR i 1 = 400 59.1 1.42 = 316F (158C)
where
tio,1 = outer surface temperature of first insulation (calcium silicate), F (C)
tpo = outer surface temperature of pipe, F (C)

6.21

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

and
t io 2 = t po q R i 1 + R i 2 = 400 59.1 1.42 + 2.14 = 190F (87.7C)
where
tio,2 = outer surface temperature of second insulation (polyurethane foam), F (C)
The new estimate of the mean insulation temperature is (400 + 316)/2 = 358F (181C) for the
calcium silicate and (316 + 190)/2 = 253F (123C) for the polyurethane foam. Thus, the insulation
thermal conductivity for the calcium silicate would be interpolated to be 0.044 Btu/hftF
(0.077 W/mK) and the resulting thermal resistance is Ri,1 = 1.35 hftF/Btu (0.781 mK/W). For
the polyurethane foam insulation the thermal conductivity would be interpolated to be 0.015 Btu/
hftF (0.026 W/mK) and the resulting thermal resistance is Ri,2 = 2.00 hftF/Btu (1.16 mK/W).
The soil thermal resistance remains unchanged, and the heat loss is recalculated as q = 61.5 Btu/hft
(59.1 W/m). The calcium silicate insulation outer surface temperature is now approximately tio,1 =
317F (158C) while the outer surface temperature of the polyurethane foam is calculated to be
tio,2 = 194F (89.0C). Since these temperatures are within a few degrees of those calculated previously, no further calculations are needed.
The maximum temperature of the polyurethane insulation of 317F (158C) occurs at its inner
surface, i.e., the interface with the calcium silicate insulation. This temperature clearly exceeds the
maximum accepted 30-year service temperature of polyurethane foam of 250F (120C) as stipulated by EN 253 (CEN 2009). Thus it would be necessary to increase the amount of calcium silicate
insulation significantly in order to achieve a maximum temperature for the polyurethane foam insulation within its long-term service temperature limit. Under the conditions of this example it would
take about 5 in. (127 mm) of calcium silicate insulation to reduce the insulation interface temperature to less than 250F (120C).
In Example 6.7 for a system with two insulation materials and an air space, the impact of
decreasing burial depth is also examined, as well as the impact of soil thermal conductivity.

Example 6.7
As in Example 6.6, we will assume a high-temperature water (HTW) line operating at 400F
(200C) is to be installed in southern Texas. Here it also consists of a 6 in. (150 mm) nominal
diameter (6.625 in. outer diameter [84 mm outer radius]) carrier pipe insulated, but the first layer
of insulation is initially proposed to be 1.5 in. (40 mm) of mineral wool. The carrier pipe and the
mineral wool insulation are contained inside of a 0.125 in. (3.2 mm) thick steel conduit with a 1 in.
(25 mm) air space between the insulation exterior and the conduit interior. On the exterior of the
steel conduit, 1 in. (25 mm) polyurethane foam insulation has been proposed. The polyurethane
insulation thermal properties will be the same as those assumed for Example 6.6 and provided in
the table within that example; the assumed thermal properties of the mineral wool insulation are
provided in the table below. The polyurethane insulation will be encased in a 0.25 in. (6.4 mm)
thick high-density polyethylene (HDPE) jacket. The piping system will be buried 10 ft (3 m) deep
to the pipe centerline. The initial assumption for the burial depth and soil thermal properties are the
same as for those used in Example 6.6 but the impact of those design parameters will be examined.
Neglect the thermal resistances of the pipe, conduit, and HDPE jacket.

6.22

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

Insulation Properties for Example 6.7


Insulation Mean Temperature,
F (C)

Insulation Thermal Conductivity,


Btu/hftF (W/mK)

200 (93)

0.026 (0.045)

300 (149)

0.033 (0.057)

400 (204)

0.043 (0.074)

Mineral
Wool

Solution:
Following Example 6.6, the maximum operating temperature of the materials is sought, thus the
maximum expected soil temperature rather than the average annual soil temperature must be used
along with the lowest anticipated soil thermal conductivity and the deepest burial depth. The effect
of these design parameters will be explored briefly once the proposed design has been evaluated.
Calculation of the thermal resistances of the insulation, soil, and air space follow the methods outlined in Examples 6.5 and 6.6 and thus will not be repeated here in detail. The results from the calculation as well as the final results for the thermal resistances after three iterations are summarized in the
tables that follow (I-P and SI) to allow the interested reader to repeat the calculations as an exercise.
In I-P units:
Insulation
Thickness,
in.

Thermal Resistances,
Air
Burial
hftF/Btu
Space, Depth,
ft
Mineral Urethane in.
Mineral
Air
Urethane
Soil
Wool
Foam
Wool
Space
Foam

Foam
HDPE
Heat
Maximum
Casing
Loss,
Temperature, Temperature,
Btu/hft
F
F

1.5

10

1.55

0.132

1.69

1.86

62.1

296

191

2.5

10

2.44

0.109

1.50

1.79

55.6

258

175

10

2.85

0.101

1.41

1.76

53.0

244

169

3.5

10

3.23

0.0934

1.34

1.73

50.8

231

163

10

3.60

0.0870

1.27

1.70

48.7

220

158

In SI units:
Insulation
Thickness,
mm

Thermal Resistances,
Air
Burial
mK/W
Space, Depth,
m
Mineral Urethane mm
Mineral
Air
Urethane
Soil
Wool
Foam
Wool
Space
Foam

Heat
Loss,
W/m

Foam
HDPE
Maximum
Casing
Temperature, Temperature,
C
C

38.1

25.4

25.4

3.05

0.895

0.0764

0.977

1.08

59.6

146

88

63.5

25.4

25.4

3.05

1.41

0.0632

0.864

1.04

53.5

126

79.5

76.2

25.4

25.4

3.05

1.65

0.0582

0.817

1.02

51.0

118

75.9

88.9

25.4

25.4

3.05

1.87

0.0540

0.774

1.00

48.8

111

72.9

102

25.4

25.4

3.05

2.08

0.0503

0.736

0.98

46.8

105

70.2

As can be seen from the tables above, for the design initially proposed with 1.5 in. (38 mm) of
mineral wool insulation the maximum temperature of the urethane foam (at its interface with the

6.23

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

steel casing) is 296F (146C). This clearly exceeds the long-term temperature capabilities of urethane foam. Additionally note that the temperature of the HDPE casing would be 191F (88C),
which would likely be well in excess of its capabilities as well. The tables above provide results for
other thicknesses of the mineral wool insulation, and as can be seen from these results it would
require 3 in. (76 mm) of mineral wool insulation to reduce the maximum temperature seen by the
urethane foam to within its 250F (121C) limit for long-term exposure. Even with this insulation
thickness the HDPE casing temperature would be at 169F (75.9C), which could be problematic
depending on the HDPE formulation. As will be shown by some additional calculations in this
example, the casing temperature could be reduced by adding additional thickness to the urethane
insulation. However, this has the unfortunate consequence of increasing the maximum temperature
of the urethane insulation, in the absence of any other changes to the design.
The burial depth and soil thermal conductivity chosen in this example and in Example 6.6
could be considered extreme; however, they are encountered in practice. In fact, deeper burials have
not been entirely infrequent in practice, and the results of these heat transfer calculations should
only reinforce the cautions in Chapter 4 regarding deep burial. Low soil thermal conductivity may
result not just from low rainfall and deep water tables but also from soil type and the impact of the
constant heat loss from the buried heat distribution system. The tables below (I-P and SI) illustrate
the impact of soil thermal conductivity and burial depth on the design of this example.
In I-P units:
Insulation
Thermal Resistances,
Foam
HDPE
Thickness,
Soil Thermal
Air
Burial
Heat
hftF/Btu
Maximum
Casing
in.
Conductivity,
Space, Depth,
Loss,
Temperature, Temperature,
Btu/hftF Mineral Urethane in.
ft
Btu/hft
Mineral Air Urethane
F
F
Soil
Wool
Foam
Wool Space Foam
0.30

3.0

10

2.85

0.101

1.41

1.76

53.3

243

167

0.80

2.0

10

2.08

0.120

1.65

0.685

72.0

241

123

0.30

2.5

2.47

0.109

1.51

1.42

59.2

247

158

0.80

2.0

2.09

0.120

1.65

0.547

73.9

236

114

0.30

3.0

10

2.76

0.101

2.66

1.70

45.2

271

151

In SI units:
Insulation
Thermal Resistances,
Thickness,
Soil Thermal
Air
Burial
mK/W
mm
Conductivity,
Space, Depth,
W/mK
m
Mineral Urethane mm
Mineral Air Urethane
Soil
Wool
Foam
Wool Space Foam

6.24

Foam
HDPE
Heat
Maximum
Casing
Loss,
Temperature, Temperature,
W/m
C
C

0.52

76.2

25.4

25.4

3.05

1.65

0.0582

0.818

1.02

51.2

117

75.3

1.4

50.8

25.4

25.4

3.05

1.20

0.0692

0.951

0.396

69.2

116

50.6

0.52

63.5

25.4

25.4

1.52

1.43

0.0632

0.874

0.82

56.9

120

70.0

1.4

50.8

25.4

25.4

1.52

1.21

0.0692

0.956

0.316

71.0

114

45.6

0.52

76.2

50.8

25.4

3.05

1.60

0.0582

1.54

0.98

43.4

133

65.9

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

The tables above contain the results of several calculations with increased soil thermal conductivity and reduced burial depth. For each of these cases the thickness of the mineral wool insulation
has been adjusted in 0.5 in. (12.7 mm) increments to find the minimum mineral wool insulation
thickness that would reduce the maximum temperature of the polyurethane insulation to 250F
(121C) or lower. The results show that where the soil conditions result in the soil thermal conductivity being increased to 0.80 Btu/hftF (1.4 W/mK) the thickness of the mineral wool insulation
may be reduced from 3 to 2 in. (75 to 50 mm). Where the soil conditions result in a thermal conductivity of 0.3 Btu/hftF (0.52 W/mK), as was the case in the original calculation of the example, if
the burial depth is decreased from 10 to 5 ft (3.05 to 1.52 m) the thickness of the mineral wool insulation may be reduced only by 0.5 in. (12.7 mm), from 3 to 2.5 in. (75 to 63.5 mm). For a case
where the lower burial depth and the high soil thermal conductivity would hold, the thickness of the
mineral wool insulation would still need to be 2 in. (50 mm); thus, the decrease in burial depth from
10 to 5 ft (3.05 to 1.52 m) results in slightly lower maximum temperature for the urethane foam but
no net reduction in the mineral wool insulation thickness required.
The last entry in the above tables illustrates the impact of increasing the thickness of the polyurethane insulation on the temperature of the HDPE jacket. While the addition of 1 in. (25 mm) of
urethane insulation to bring its thickness to 2 in. (50 mm) reduces the maximum temperature of the
HDPE jacket from 167F to 161F (75.3C to 65.9 C), it increases the maximum temperature of
the urethane foam insulation from 243F to 271F (117C to 133C), above the accepted limit of
250F (121C).

It should be noted that all of the calculations in Examples 6.6 and 6.7 are based on the
assumption of a single insulated buried pipe. In the case where supply and return lines are
buried side by side and both are operating at temperatures above the surrounding soil temperature, the heat losses will be reduced; however, the temperatures of the components in
the system will be elevated. The effect can be considered as somewhat analogous to burying the pipelines in a region with higher soil temperatures, which of course also reduces
their heat loss. The formulation for calculating heat losses for adjacent supply and return
lines that is presented later in this chapter does not allow for direct calculation of temperatures within the system. As an approximation it is suggested that the individual heat
losses for the supply pipe, where the limiting condition will occur, be first calculated
using Equations 6.23 through 6.30 presented later (and illustrated in Example 6.9). To
approximate the temperatures within the system, use the methods outlined in Example 6.7
as if the pipeline were buried by itself but adjust the soil temperature upward to achieve
the same rate of heat loss as determined using the method of Equations 6.23 through 6.30
for the supply pipe in the supply and return configuration. This will underestimate the
impacts of the adjacent return pipeline somewhat, so be conservative when using this
method.
The purpose of Examples 6.6 and 6.7 is to illustrate the need to fully consider the
burial conditions when assessing systems that have limitations on any of their component
temperatures that are below the operating temperature of the carrier fluid. While the examples are from systems with multiple insulations with temperature limitations on one of the
insulations, this consideration also applies to systems other than those with composite
insulation; it applies to any system. Nonmetallic jacket materials as well as corrosionresistant coatings are other examples of materials that often have temperature limitations
well below the carrier pipe temperatures encountered in practice. Deep burial of systems
and/or burial of systems in low-thermal-conductivity soil requires that calculations such as

6.25

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 6.5 Two pipes buried in a common conduit with an air space.
Courtesy of ASHRAE (2012, Chapter 12, Figure 8)

those from Examples 6.6 and 6.7 be conducted where materials with temperature limitations below the carrier pipe temperature exist.

Two Pipes Buried in Common Conduit with Air Space


For the case of two pipes buried in a common conduit with an air space (Figure 6.5),
make the same assumption as in the previous section regarding heat transfer with the air
space. For convenience, add some of the thermal resistances as follows:
R 1 = R p1 + R i1 + R a1

(6.15)

R 2 = R p2 + R i2 + R a2

(6.16)

The subscripts 1 and 2 differentiate between the two pipes within the conduit. The
combined heat loss is then given by:
t f 1 t s R1 + t f 2 t s R2
q = --------------------------------------------------------------------------------1 + R cs R 1 + R cs R 2

(6.17)

where Rcs is the total thermal resistance of conduit shell and soil, hftF/Btu (mK/W).
Once the combined heat flow is determined, calculate the bulk temperature within the
air space:
t a = t s + qR cs

(6.18)

Then calculate the insulation outer surface temperatures:

6.26

t i1 = t a + t f 1 t a R a1 R 1

(6.19)

t i2 = t a + t f 2 t a R a2 R 2

(6.20)

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

The heat flows from each pipe are given by:


q1 = t f 1 t a R1

(6.21)

q2 = t f 2 t a R2

(6.22)

When the insulation thermal conductivity is a function of its temperature, as is usually the case, an iterative calculation is required, as illustrated in Example 6.8.

Example 6.8
A pair of 4 in. (100 mm) nominal pipe size (NPS) medium-temperature hot-water supply and
return lines run in a common 21 in. (533 mm) outside diameter conduit. Assume that the supply
temperature is 325F (163C) and the return temperature is 225F (107C). The supply pipe is insulated with 2.5 in. (63.5 mm) of mineral wool insulation, and the return pipe has 2 in. (51 mm) of
mineral wool insulation. This insulation has the same thermal properties as those given in
Example 6.5. The pipe is buried 4 ft (1.2 m) to centerline in soil with a thermal conductivity of
1 Btu/hftF (1.7 W/mK). Assume the thermal resistance of the pipe, the conduit, and the conduit
coating are negligible. As a first estimate, assume the bulk air temperature within the conduit is
100F (38C). In addition, use this temperature as a first estimate of the insulation surface temperatures to obtain the mean insulation temperatures and subsequent insulation thermal conductivities.
Solution:
By interpolation, estimate the insulation thermal conductivities from the data given in
Example 6.5 at 0.0265 Btu/hftF (0.046 W/mK) for the supply pipe and 0.0245 Btu/hftF
(0.042 W/mK) for the return pipe. Calculate the thermal resistances using Equations 6.15 and 6.16:
R 1 = R i1 + R a1 = ln 0.396 0.188 2 0.0265 + 0.053 0.396
= 4.48 + 0.13 = 4.61 hftF/Btu (2.67 mK/W)
R 2 = R i2 + R a2 = ln 0.354 0.188 2 0.0245 + 0.053 0.354
= 4.11 + 0.15 = 4.26 hftF/Btu (2.46 mK/W)
R cs = R s = ln 8 0.875 2 1 = 0.352 hftF/Btu (0.203 mK/W)
Calculate first estimate of combined heat flow from Equation 6.17:
325 60 4.61 + 225 60 4.26- = 83.0 Btu/hft (79.8 W/m)
q = -------------------------------------------------------------------------------------1 + 0.352 4.61 + 0.352 4.26
Estimate bulk air temperature in the conduit with Equation 6.18:
t a = 60 + 83.0 0.352 = 89.2F (31.8C)
Then revise estimates of the insulation surface temperatures with Equations 6.19 and 6.20:
t i1 = 89.2 + 325 89.2 0.13 4.61 = 95.8F (35.4C)
t i2 = 89.2 + 225 89.2 0.15 4.26 = 94.0F (34.4C)

6.27

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

These insulation surface temperatures are close enough to the original estimate of 100F
(37.8C) that further iterations are not warranted. If the individual supply and return heat losses are
desired, calculate them using Equations 6.21 and 6.22.

Two Buried Pipes or Conduits


The case of two buried pipes or conduits (Figure 6.6) may be formulated in terms of
the thermal resistances used for a single buried pipe or conduit and some geometric and
temperature factors. The factors needed are:
1 = t p2 t s t p1 t s

(6.23)

2 = 1 1 = t p1 t s t p2 t s

(6.24)

d 1 + d 2 2 + a 2 0.5
1
P 1 = ------------ ln -------------------------------------
2k s d d 2 + a 2

(6.25)

d 2 + d 1 2 + a 2 0.5
1
P 2 = ------------ ln --------------------------------------
2k s d d 2 + a 2

(6.26)

where a is the horizontal separation distance between the centerline of two pipes, ft (m).
The thermal resistance for each pipe or conduit is given by:
R t1 P 12 R t2
R e1 = -------------------------------------1 P 1 1 R t2

Figure 6.6 Two buried pipes or conduits.


Courtesy of ASHRAE (2012, Chapter 12, Figure 9)

6.28

(6.27)

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

R t2 P 22 R t1
R e2 = -------------------------------------1 P 2 2 R t1

(6.28)

where
= temperature factor, dimensionless
P = geometric/material factor, hftF/Btu (mK/W)
Re = effective thermal resistance of one pipe/conduit in two-pipe system, hftF/Btu
(mK/W)
Rt = total thermal resistance of one pipe/conduit if buried separately, hftF/Btu
(mK/W)
Heat flow from each pipe is then calculated from:
q 1 = t p1 t s R e1

(6.29)

q 2 = t p2 t s R e2

(6.30)

Example 6.9
Consider buried supply and return lines for a low-temperature hot-water system. The carrier
pipes are 4 in. (100 mm) nominal pipe size (NPS) (4.5 in. [114 mm] outer diameter) with 1.5 in.
(38 mm) of urethane foam insulation. The insulation is protected by a 0.25 in. (6.4 mm) thick polyvinyl chloride (PVC) jacket. The thermal conductivity of the insulation is 0.013 Btu/hftF
(0.022 W/mK) and is assumed constant with respect to temperature. The pipes are buried 4 ft
(1.2 m) deep to the centerline in soil with a thermal conductivity of 1 Btu/hftF (1.7 W/mK) and a
mean annual temperature of 60F (16C). The horizontal distance between the pipe centerlines is 2
ft (0.61 m). The supply water is at 250F (121C), and the return water is at 150F (65.6C).
Solution:
Neglect the thermal resistances of the carrier pipes and the PVC jacket. First, calculate the
resistances from Equations 6.12 and 6.13 as if the pipes were independent of each other:
ln 8.0 0.333 - = 0.51 hftF/Btu (0.29 mK/W)
R s1 = R s2 = ---------------------------------2 1.0
ln 0.313 0.188 - = 6.25 hftF/Btu (3.61 mK/W)
R i1 = R i2 = ---------------------------------------2 0.013
R t1 = R t2 = 0.51 + 6.25 = 6.75 hftF/Btu (3.90 mK/W)
From Equations 6.25 and 6.26 the geometric and temperature factors are:
4 + 4 2 + 2 2 0.5
1
P 1 = P 2 = --------------- ln --------------------------------
= 0.225 hftF/Btu (0.130 mK/W)
2 1 4 4 2 + 2 2
1 = 150 60 250 60 = 0.474
2 = 1 1 = 2.11

6.29

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Calculate the effective total thermal resistances as:


6.76 0.225 2 6.76
R e1 = ------------------------------------------------------------- = 6.87 hftF/Btu (3.97 mK/W)
1 0.225 0.474 6.76
6.76 0.225 2 6.76 - = 7.32 hftF/Btu (4.23 mK/W)
R e2 = --------------------------------------------------------1 0.225 2.11 6.76
The heat flows are then:
q 1 = 250 60 6.87 = 27.7 Btu/hft (26.6 W/m)
q 2 = 150 60 7.32 = 12.3 Btu/hft (11.8 W/m)
q t = 27.7 + 12.3 = 40.0 Btu/hft (38.4 W/m)
Note that when the resistances and geometries for the two pipes are identical, the total heat flow
from the two pipes is the same if the temperature corrections are used or if they are set to unity. The
individual losses will vary somewhat, however.
These equations may also be used with air space systems. When the thermal conductivity of the
pipe insulation is a function of temperature, iterative calculations must be done.

Pipes in Buried Trenches or Tunnels


Buried rectangular trenches or tunnels (Figure 6.7) require several assumptions to
obtain approximate solutions for the heat transfer. First, assume that the air within the
tunnel or trench is uniform in temperature and that the same is true for the inside surface
of the trench/tunnel walls. Field measurements on operating shallow trenches
(Phetteplace et al. 1991) indicate maximum spatial air temperature variations of about
10F (5.6C). Air temperature variations of this magnitude within a tunnel or trench will
not cause significant errors for systems with normal operating temperatures when using
the following calculation methods.
Unless numerical methods are used, an approximation must be made for the resistance of a rectangular region such as the walls of a trench or tunnel. One procedure is to
assume linear heat flow through the trench or tunnel walls, which yields the following
resistance for the walls (Phetteplace et al. 1981):
xw
R w = -------------------------2k w a + b
where
Rw =
xw =
a =
b =
kw =
6.30

thermal resistance of trench/tunnel walls, hftF/Btu (mK/W)


thickness of trench/tunnel walls, ft (m)
width of trench/tunnel inside, ft (m)
height of trench/tunnel inside, ft (m)
thermal conductivity of trench/tunnel wall material, Btu/hftF (W/mK)

(6.31)

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

Figure 6.7 Pipes in buried trenches or tunnels.


Courtesy of ASHRAE (2012, Chapter 12, Figure 10)

As an alternative to Equation 6.31, the thickness of the trench/tunnel walls may be


included in the soil burial depth. This approximation is only acceptable when the thermal
conductivity of the trench/tunnel wall material is similar to that of the soil.
The thermal resistance of the soil surrounding the buried trench/tunnel is calculated
using the following equation (Rohsenow 1998):
ln 3.5d b o0.75 a o0.25
R ts = ------------------------------------------------------k s a o 2b o + 5.7
where
Rts =
ao =
bo =
d =

ao bo

(6.32)

thermal resistance of soil surrounding trench/tunnel, hftF/Btu (mK/W)


width of trench/tunnel outside, ft (m)
height of trench/tunnel outside, ft (m)
burial depth of trench to centerline, ft (m)

Equations 6.31 and 6.32 can be combined with the equations already presented for
insulation (Equation 6.13) and air spaces (Equation 6.14) to calculate approximations for
the heat flows and temperatures for trenches/tunnels. As with the conduits described in
earlier examples, the heat transfer processes within the air space inside the trench/tunnel
are too complex to warrant a complete treatment for design purposes. The thermal resistance of this air space may be approximated by several methods. For example, Equation
6.14 may be used to calculate an approximate resistance for the air space. Thermal resistances at the pipe insulation/air interface may also be calculated from published heat
transfer correlations. If the thermal resistance of the air/trench wall interface is also
included, use Equation 6.33:
R aw = 1 2h t a + b

(6.33)

where

6.31

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Raw = thermal resistance of air/trench wall interface, hftF/Btu (mK/W)


= total heat transfer coefficient at air/trench wall interface, Btu/hft2F (W/m2C)
ht
The total heat loss from the trench/tunnel is calculated from the following relationship:
t p1 t s R 1 + t p2 t s R 2
q = ----------------------------------------------------------------------1 + R ss R 1 + R ss R 2

(6.34)

where
R1, R2 = thermal resistances of two-pipe/insulation systems within trench/tunnel,
hftF/Btu (mK/W)
= total thermal resistance on soil side of air within trench/tunnel, hftF/Btu
Rss
(mK/W)
Once the total heat loss has been found, the air temperature within the trench/tunnel
may be found as:
t ta = t s + qR ss

(6.35)

where tta is the air temperature within the trench/tunnel, F (C).


The individual heat flows for the two pipes within the trench/ tunnel are then:
q 1 = t p1 t ta R 1

(6.36)

q 2 = t p2 t ta R 2

(6.37)

If the thermal conductivity of the pipe insulation is a function of temperature, assume


an air temperature for the air space before starting calculations. If the air temperature calculated with Equation 6.35 differs significantly from the initial assumption, it will be necessary to iterate the calculations.

Example 6.10
The walls of a buried trench are 6 in. (152 mm) thick, and the trench is 3 ft (0.91 m) wide and 2 ft
(0.61 m) tall. The trench is constructed of concrete, with a thermal conductivity of kw = 1 Btu/hftF
(1.7 W/mK). The soil surrounding the trench also has a thermal conductivity of ks = 1 Btu/hftF
(1.7 W/mK). The centerline of the trench is 4 ft (1.2 m) below grade, and the soil temperature is
assumed be 60F (16C). The trench contains supply and return lines for a medium-temperature water
(MTW) system with the physical and operating parameters identical to those in Example 6.8.
Solution:
Assuming the air temperature within the trench is 100F (37.8C), the thermal resistances for
the pipe/insulation systems are identical to those in Example 6.8, or:
R 1 = 4.61 hftF/Btu (2.67 mK/W)
R 2 = 4.26 hftF/Btu (2.46 mK/W)

6.32

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

The thermal resistance of the soil surrounding the trench is given by Equation 6.32:
ln 3.5 4 3 0.75 4 0.25
R ts = ----------------------------------------------------------------------- = 0.231 hftF/Btu (0.134 mK/W)
1 4 6 + 5.7
The thermal resistance of the trench walls is calculated using Equation 6.31:
R w = 0.5 2 3 + 2 = 0.050 hftF/Btu (0.029 mK/W)
If the thermal resistance of the air/trench wall is neglected, the total thermal resistance on the
soil side of the air space is:
R ss = R w + R ts = 0.050 + 0.231 = 0.281 hftF/Btu (0.162 mK/W)
Find a first estimate of the total heat loss using Equation 6.34:
325 60 4.61 + 225 60 4.26
q = --------------------------------------------------------------------------------------- = 85.4 Btu/hft (82.1 W/m)
1 + 0.281 4.61 + 0.281 4.26
The first estimate of the air temperature within the trench is given by Equation 6.35:
t a = 60 + 85.4 0.281 = 84.0F (28.9C)
Refined estimates of the pipe insulation surface temperatures are then calculated using Equations 6.19 and 6.20:
t i1 = 84.0 + 325 84.0 0.13 4.61 = 90.8F (32.7C)
t i2 = 84.0 + 225 84.0 0.15 4.26 = 89.0F (31.7C)
From these estimates, calculate the revised mean insulation temperatures to find the resultant
resistance values. Repeat the calculation procedure until satisfactory agreement between successive
estimates of the trench air temperature is obtained. Calculate the individual heat flows from the
pipes with Equations 6.36 and 6.37.
If the thermal resistance of the trench walls is added to the soil thermal resistance, the thermal
resistance on the soil side of the air space is:
ln 14 2 0.75 3 0.25
R ss = -------------------------------------------------- = 0.285 hftF/Btu (0.165 mK/W)
1 3 4 + 5.7
The result is less than 2% higher than the resistance previously calculated by treating the trench
walls and soil separately. In the event that the soil and trench wall material have significantly different thermal conductivities, this simpler calculation will not yield as favorable results and should not
be used.

6.33

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Pipes in Shallow Trenches


The cover of a shallow trench is exposed to the environment. Thermal calculations for
such trenches require the following assumptions: (1) the interior air temperature is uniform as discussed in the section Pipes in Buried Trenches or Tunnels, and (2) the soil and
the trench wall material have the same thermal conductivity. This assumption will yield
reasonable results if the thermal conductivity of the trench material is used because most
of the heat flows directly through the cover. The thermal resistance of the trench walls and
surrounding soil is usually a small portion of the total thermal resistance; thus, the heat
losses are not usually highly dependent on this thermal resistance. Using these assumptions, Equations 6.32 and 6.34 through 6.37 may be used for shallow trench systems.
Another method for calculating the heat losses in a shallow trench assumes an interior
air temperature and treats the pipes as pipes in air (see the later section Pipes in Air). Interior air temperatures in the range of 70F to 120F (21C to 49C) have been observed in
a temperate climate (Phetteplace et al. 1991).

Example 6.11
Consider a shallow trench having the same physical parameters and operating conditions as the
buried trench in Example 6.10, except that the top of the trench is at grade level.
Solution:
Calculate the thermal resistance of the shallow trench using Equation 6.32:
ln 3.5 1.5 2 0.75 3 0.25
R ts = R ss = --------------------------------------------------------------------------- = 0.134 hftF/Btu (0.0775 mK/W)
1 3 4 + 5.7
Use thermal resistances for the pipe/insulation systems from Example 6.10, and use
Equation 6.29 to calculate q:
325 60 4.61 + 255 60 4.26
- = 90.7 Btu/hft (87.2 W/m)
q = -------------------------------------------------------------------------------------1 + 0.134 4.61 + 0.134 4.26
From this, calculate the first estimate of the air temperature using Equation 6.35:
t ta = 60 + 90.7 0.134 = 72.2F (22.3C)
Then refine estimates of the pipe insulation surface temperatures using Equations 6.19
and 6.20:
t i1 = 72.2 + 325 72.2 0.13 4.61 = 79.3F (26.3C)
t i2 = 72.2 + 225 72.2 0.15 4.26 = 77.6F (25.3C)
From these, calculate the revised mean insulation temperatures to find resultant resistance values. Repeat the calculation procedure until satisfactory agreement between successive estimates of
the trench air temperatures is obtained. If the individual heat flows from the pipes are desired, they
may be calculated using Equations 6.36 and 6.37.

6.34

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

Pipes in Loose Fill Insulation


For the calculation of the heat loss from pipes encased in loose fill insulation (LFI),
the soil thermal resistance may be calculated using Equation 6.32. Calculation of the thermal resistance of the LFI material itself is unfortunately less straightforward due to the
geometry involved. While thermal resistance formulas can be found for many simple
geometries of interest, the geometry typically used for installations with LFI does not
have a commonly accepted resistance formulation available. Thus we are forced to use
approximate methods.
Ideally, a heat transfer problem with the relatively complicated geometry like what
we have here would be treated by numerical methods, such as finite differences or finite
elements. Because of the computation effort involved and the even greater effort required
to prepare the discretisation of the medium, numerical methods are probably not appropriate for such calculations when used for design purposes, although numerical methods
provide a means to access the validity of approximating assumptions. Several methods of
approximation are possible, as discussed in the following paragraphs.
For a single buried pipe in a LFI system, the closest geometry to the actual problem
for which a resistance formulation is available is that of a square insulation surrounding a
concentric cylinder (Lunardini 1981):
0.54x
ln ---------------s
ro
R lfi = --------------------------2 k lfi

(6.38)

where
xs = length of each side of the square insulation, ft (m)
Rlfi = thermal resistance of the LFI, hftF/Btu (mK/W)
klfi = thermal conductivity of the LFI, Btu/hftF (W/mK)
There are two differences between the typical district heating and cooling application
and this resistance relation: in most cases the LFI material is placed in a rectangular cross
section, and normally we have at least two pipes within the LFI material. Approximating
the rectangle with an equivalent square is straightforward and is expected to give reasonable results for rectangles that are reasonably close to square, i.e., with insulation widthto-height ratios from 1.0 to perhaps 1.25. The approximation for the square is simply:
xse = (ao + bo)/2

(6.39)

where xse is the side length of an effective square insulation cross section, ft (m).
The greatest approximation that must be made is using resistance relations for single
pipes in cases where multiple pipes are present. Treating each pipe separately and using
the insulation volume extending to the midpoint between it and the adjacent pipe is one
method. However, this will overestimate the heat loss/gain in most instances; thus, it can
be considered if a conservative approach is desired. Another possible approach is to combine the multiple pipes (two in most cases) into a single fictitious effective pipe. This is
the approach that we will use here. We begin by defining the effective radius of this fictitious pipe as the sum of the two pipe radii:
roe = ro1 + ro2

(6.40)

where roe is the outer radius of the effective pipe, ft (m).


6.35

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

It is also necessary to define an appropriate temperature for the effective pipe. An


approximation for this is to use the average of the two pipe temperatures weighted by
their respective radii:
r o1 t p1 + r o2 t p2
t pe = ------------------------------------r o1 + r o2

(6.41)

where
tpe = effective pipe temperature, F (C)
tp1 = temperature of pipe 1, F (C)
tp2 = temperature of pipe 2, F (C)
Now the thermal resistance of the equivalent square insulation cross section is:
0.54x se
-
ln ---------------- r oe
R lfi = ----------------------------2 k lfi

(6.42)

With this result the heat loss is calculated from:


t pe t s
q = --------------------R lfi + R ts

(6.43)

Since the thermal conductivity of the LFI is a function of temperature it will be necessary to iterate the calculation as has been done for the other types of systems discussed.

Example 6.12
Consider a system that has a 6 in. (150 mm) nominal diameter steam pipe (6.625 in. [168 mm]
outer diameter) and a 3 in. (75 mm) nominal diameter condensate pipe (3.5 in. [88.9 mm] outer
diameter) in a loose fill insulation (LFI) material. The steam pipe is carrying saturated 100 psig (6.9
bar) steam, which puts its temperature at 338F (170C). The condensate is assumed to be at 180F
(82.2C). The envelope of the LFI is 18 in. (457 mm) deep and 22 in. (559 mm) wide, and the depth
to the center of that envelope is 4 ft (1.2 m). The soil surrounding the LFI material is assumed to
have a thermal conductivity of 1 Btu/hftF (1.7 W/mK) and a temperature of 50F (10C). The
following table shows the thermal conductivity data assumed for the LFI material.
Mean Temperature,
F (C)

LFI Thermal Conductivity, klfi ,


Btu/hftF (W/mK)

100 (37.8)

0.050 (0.087)

200 (93.3)

0.054 (0.093)

300 (149)

0.058 (0.100)

Solution:
First we calculate the effective radius and temperature for the pipe using Equations 6.40 and
6.41, respectively:
r oe = r o1 + r o2 = 0.276 + 0.146 = 0.422 ft (0.129 m)

6.36

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

r o1 t p1 + r o2 t p2
0.276 338 + 0.146 180
- = ----------------------------------------------------------------------- = 283F (140C)
t pe = ------------------------------------0.276 + 0.146
r o1 + r o2
In order to determine a first estimate of the thermal conductivity of the LFI material well need
a first estimate of its bulk temperature. If we assume that the exterior of the LFI envelope is at
120F (48.9C), we can average that temperature with the effective pipe temperature to obtain an
average LFI temperature of 202F (94.4C); thus, its thermal conductivity can be taken as
0.054 Btu/hftF (0.093 W/mK).
The thermal resistance of the soil surrounding the LFI envelope is found using Equation 6.32:
ln 3.5d b o0.75 a o0.25
ln 3.5 4 1.5 0.75 1.83 0.25 -
R ts = ----------------------------------------------------------= --------------------------------------------------------------------------1 1.83 2 1.5 + 5.7
k s a o 2b o + 5.7
= 0.346 hftF/Btu (0.200 mK/W)
The thermal resistance of the LFI material is calculated using Equation 6.42:
0.54x se
0.54 1.67
-
ln ----------------ln --------------------------
0.422
r oe
R lfi = ----------------------------- = ------------------------------------- = 2.24 hftF/Btu (1.29 mK/W)
2 0.054
2 k lfi
The heat flow is then calculated using Equation 6.43:
t pe t s
283 50
q = --------------------- = ------------------------------ = 90.1 Btu/hft (86.6 W/m)
2.24 + 0.346
R lfi + R ts
Now that we have our first estimate of heat loss, an improved estimate of the outside temperature of the LFI material is obtained from:
t lf i o = t s + qR ts = 50 + 89.8 0.346 = 81.1F (27.3C)
where Tlfi,o is the temperature at the interface between the LFI material and the soil, F (C).
This improved estimate of the outside temperature of the LFI material, Tlfi,o , is then averaged
with the effective pipe temperature, Tep , for a new estimate of the bulk temperature of the LFI
material of 182F (83.4C). The thermal conductivity of the LFI material at this temperature is
interpolated to be 0.053 Btu/hftF (0.092 W/mK). When the heat loss calculation is repeated with
this value the resulting heat loss is slightly lower, at 88.7 Btu/hft (85.3 W/m). Since this change in
heat loss is insignificant it is not necessary to iterate the calculation.

Buried Pipes with Other Geometries


Other geometries that are not specifically addressed by the previous cases have been
used for buried thermal utilities. In some instances, the equations presented previously
may be used to approximate the system. For instance, the soil thermal resistance for a buried system with a half-round clay tile on a concrete base could be approximated as a circular system using Equation 6.11 or 6.12. In this case, the outer radius ro is taken as that
6.37

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

of a cylinder with the same circumference as the outer perimeter of the clay tile system.
The remainder of the resistances and subsequent calculations would be similar to those
for a buried trench/tunnel. The accuracy of such calculations varies inversely with the
proportion of the total thermal resistance that the thermal resistance in question comprises. In most instances, the thermal resistance of the pipe insulation overshadows other
resistances, and the errors induced by approximations in the other resistances are acceptable for design calculations.

Pipes in Air
Pipes surrounded by gases may transfer heat via conduction, convection, and/or thermal
radiation. Heat transfer modes depend mainly on the surface temperatures and geometry of
the system being considered. For air, conduction is usually dominated by the other modes
and thus may be neglected. The North American Insulation Manufacturers Association
(NAIMA) provides a complementary software tool that allows for evaluation of insulated
and uninsulated pipes, both horizontal and vertical, in air. It also has the capability to perform analyses for other objects in air and conduct economic analyses. This tool, 3E Plus,
uses the equations of ASTM C680 (2010d) and is available at www.pipeinsulation.org
(NAIMA 2012).

ECONOMICAL THICKNESS FOR PIPE INSULATION


Determining the most cost-effective insulation thickness is a classic engineering optimization problem where additional first costs must be balanced against reduced operating
costs (i.e., heat losses) to determine the option that provides the lowest cost over the
expected life of the system. Thus, a life-cycle cost analysis (LCCA) should be conducted
to determine the most economical thickness of pipe insulation. Because the insulation
thickness impacts other parameters in this type of system, each insulation thickness must
be considered as a separate system and evaluated independently. From the first-cost
standpoint, the most significant factors besides the insulation cost itself that will be
impacted by insulation thickness are the cost of the conduit or jacket around the insulation and the cost of pre-insulated fittings. Other significant first costs associated with
increased insulation thickness include increases in excavation, backfill, and compaction
costs. In addition, increases in shipping costs may also be experienced with increased
insulation thickness, although such costs are difficult to quantify as prefabricated insulated piping is normally shipped in truckload lots directly from the manufacturer to the
job site. From the operating cost (heat loss) perspective, other than the increase in insulation thermal resistance that results from increased insulation thickness, soil thermal resistances will also be affected slightly by changes in conduit/jacket diameter and burial
depth that come about from a fixed amount of burial cover desired.
The life-cycle cost (LCC) of a system is the sum of the initial capital cost and the
present worth of the subsequent cost of heat lost over the life of the system. Note that the
appropriate cost of heat should include not just the fuel costs but any other costs such as
labor, water treatment, makeup water, etc., that are increased by increased heat output
i.e., in most cases it should represent the marginal cost of heat production for the district
heating utility. The initial capital cost needs only to include those costs that are affected
by insulation thickness.
For pipes in air the tool 3E Plus (NAIMA 2012), mentioned previously in discussing
heat loss calculations, may be used for economic analyses. Regretfully 3E Plus does not
provide an option for district heating as the energy source, thus it will require some
manipulation of the inputs to achieve the desired result. Alternatively, for systems in air

6.38

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

the heat losses can be obtained from 3E Plus and then the methods that follow can be
used for the remainder of the analysis.
The basic equation of the LCC is:
LCC = CC + qt u C h PWF
where
LCC
CC
q
tu
Ch
PWF

=
=
=
=
=
=

(6.44)

present worth of LCCs associated with pipe insulation thickness, $/ft ($/m)
capital costs associated with pipe insulation thickness, $/ft ($/m)
annual average rate of heat loss, Btu/hft (W/m)
utilization time for system each year, h
cost of heat lost from system, $/Btu ($/Wh)
present worth factor for future annual heat loss costs, dimensionless

The present worth factor is the reciprocal of the capital recovery factor and may be
found from Equation 6.45:
CRF = i 1 + i n 1 + i n 1

(6.45)

where
CRF = cost recovery factor, dimensionless
i
= interest rate
n
= useful lifetime of system, years
If the cost of heat is expected to inflate over the lifetime of the project, the present
worth factor, PWF, may be computed using an effective interest rate computed from the
following equation (ASHRAE 2011):
i' = (i j)/(1 + j)

(6.46)

where
i' = effective interest rate
j = inflation rate

Example 6.13
Consider an application where supply and return high-temperature hot-water piping for a district
heating system are being installed. The piping construction is a steel conduit drainable, dryable, testable (DDT) type as described in Chapter 4. The supply temperature is 350F (177C) and the return
temperature is 250F (121C). The carrier pipe size is 4 in. (100 mm) nominal pipe size (NPS) (4.5
in. [114 mm] actual outside diameter). The burial depth is 3.5 ft (1.1 m) to the outer conduit, and the
horizontal separation distance between the supply and return pipe conduit exterior surfaces is 6 in.
(150 mm). Select backfill surrounds the pipe with 4 in. (100 mm) below, 12 in. (300 mm) above, and
6 in. (150 mm) on each side. The select backfill is compacted 50% by hand and 50% by a vibrating
plate compactor, which yields a backfill and compaction cost of $45.40 per cubic yard ($59.33 per
cubic meter) based on third-quarter 2012 costs from RSMeans Online cost data (RSMeans 2013).
The excavation extends a total of 1.5 ft (0.46 m) on the side of each conduit and the excavation cost
is taken as $6.50 per cubic yard ($8.50 per cubic meter) based on third-quarter 2012 costs from
RSMeans Online cost data. Compaction outside the zone of select backfill is by vibrating plate

6.39

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

compactor, which results in backfill and compaction costs of $38.79 per cubic yard ($50.70 per
cubic meter) again based on third-quarter 2012 costs from RSMeans Online cost data.
The average annual soil temperature for the location is 60F (16C) and the soil is classified as
a sand/gravel soil with a 6% moisture content and 115 lbm/ft3 (1.84 g/cm3) dry density yielding a
soil thermal conductivity of 1.11 Btu/hftF (1.92 W/mK) using Equation 6.3a. The insulation is
mineral fiber and the thermal conductivities are as stipulated in Table 6.2. The ground surface heat
transfer coefficient is 1.63 Btu/hft2F (9.24 W/m2C), which results in a convective layer thickness equivalence of 0.68 ft (0.21 m) from Equation 6.10.
For the baseline analysis the economic parameters are:
Useful system lifetime = 20 years
Cost of heat = $15/MMBtu ($14.22/GJ)
Interest rate = 6%
Inflation rate for heat cost = 3%
Annual operating hours = 8760
The assumed piping cost data are given in I-P units in the table below.

Case #

Nominal
Pipe Size,
in.

Pipe Outside
Diameter,
in.

Insulation
Thickness,
in.

Conduit Outside
Diameter,
in.

System
Cost, $/ft

4.5

1.5

10.75

$43.00

4.5

10.75

$45.00

4.5

2.5

12.75

$50.00

4.5

12.75

$52.00

4.5

3.5

14

$56.00

4.5

16

$63.00

4.5

4.5

16

$66.00

The heat losses are calculated using the methods outlined previously in this chapter; the table below
summarizes the heat loss calculations and the calculation of their present worth (in I-P units only).

Case #

Insulation
Thickness,
in.

Heat Loss
Supply,
Btu/hft

Heat Loss
Return,
Btu/hft

Heat Loss
Total,
Btu/hft

Annual Heat
Loss,
Btu/ft

Annual Heat
Loss Cost,
$/ft

Present Worth
of Lifetime
Heat Loss,
$/ft

1.5

72.6

42.1

114.7

1004474

$15.07

$225.98

2.0

60.8

35.6

96.4

844805

$12.67

$190.06

2.5

53.4

31.6

85.0

744581

$11.17

$167.51

3.0

47.9

28.5

76.4

669226

$10.04

$150.56

3.5

43.9

26.2

70.1

614479

$9.22

$138.24

4.0

40.9

24.5

65.3

572315

$8.58

$128.76

4.5

38.3

23.0

61.2

536446

$8.05

$120.69

The additional costs for excavation, backfill, and compaction that arise from the increases in
insulation thickness and conduit diameters are accounted for by calculating the volumes and associated costs above the base case of minimum insulation thickness. These costs are summarized in the
table that follows (in I-P units only).

6.40

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

Casing
Insulation
Outside
Case # Thickness,
Diameter,
in.
in.

Casing
r,
in.

Casing
r,
ft

Select
Select
Overall
Overall
Backfill
Backfill
Total
Excavation
Excavation
Envelope
Envelope
Cost,
V,
Cost,
V,
Cost,
$/linear ft
ft3/linear ft
$/linear ft
ft3/linear ft
$/linear ft

1.5

10.75

5.375

10.75

5.375

0.0000

0.00

0.00

$0.00

$0.00

$0.00

2.5

12.75

6.375

0.0833

1.46

2.62

$2.45

$2.31

$4.76

12.75

6.375

0.0833

1.46

2.62

$2.45

$2.31

$4.76

3.5

14

0.1354

2.54

4.43

$4.27

$3.79

$8.06

16

0.2188

4.54

7.60

$7.63

$6.23

$13.86

4.5

16

0.2188

4.54

7.60

$7.63

$6.23

$13.86

Finally, the cost of present worth of the heat loss, increases in excavation and backfill costs for
the increased insulation thicknesses and resulting conduits required, and the cost of the system
itself are added together to find the total life-cycle costs (LCCs) that we wish to minimize; the table
below summarizes those results (in I-P units only). Note that this is not the total cost of the systemit neglects all components of the installation cost beyond the increases only of those that
would be primarily impacted by insulation thickness.

Case #

Insulation
Thickness,
in.

Present Worth
Excavation and
System Cost,
Heat Loss Cost,
Backfill Cost,
$/ft
$/ft
$/ft

1.5

$225.98

2.0

2.5

4
5

Total Cost,
$/ft

$86.00

$311.98

$190.06

$90.00

$0.00

$280.06

$167.51

$100.00

$4.76

$272.27

3.0

$150.56

$104.00

$4.76

$259.32

3.5

$138.24

$112.00

$8.06

$258.30

4.0

$128.76

$126.00

$13.86

$268.62

4.5

$120.69

$132.00

$13.86

$266.55

For the assumptions given, the table above indicates that 3.5 in. (89 mm) of insulation offers the
lowest LCC. Note how little difference there is between the total LCC for the options with 3 to 4.5
in. (76 to 114 mm) of insulation. Also note how the limitation of discrete conduit sizes impacts the
costs; i.e., while the lowest LCC is for the 3.5 in. (89 mm) option, the LCC for the option with 4.5
in. (114 mm) of insulation is actually less than that for the option with 4 in. (102 mm) of insulation.
This results from the fact that they both require a larger conduit size than the 3.5 in. (89 mm)
option, yet the lower heat losses of the 4.5 in. (114 mm) option more than pay for its incremental
cost over the 4 in. (102 mm) option.
This analysis ignores other costs such as additional construction costs beyond excavation,
backfill, compaction, and shipping that would result for the larger sizes. Inclusion of these costs
would increase the differential cost between cases that have larger conduit sizes, which would in
turn make the options with higher insulation levels look less cost-effective. However, that must be
taken in view of the fact that future energy costs are not known and adding insulation at a later date
is not economically practical. The use of an inflation factor for energy costs, as was done in this
example, provides a more accurate analysis. Where incremental costs are low, i.e., where the total
cost is relatively flat near the lowest LCC, history would suggest that it would be prudent to err on

6.41

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

the side of additional insulation beyond that shown to have the lowest LCC; adding insulation later
is seldom feasible or cost-effective.
The result of this example will vary significantly for other sets of assumptions, with the economic assumptions being the most highly variable. It is instructive to examine a few variations to
the initial assumptions for their impact. The table below summarizes the results of several other sets
of calculations that were conducted with the same assumptions except as noted in the table (in I-P
units only).
Economic Assumed
Scenario System Life,
#
years

Heat Cost,
$/MMBtu

Interest
Rate,
%

Heat Cost
Inflation
Rate,
%

Lowest LCC
Insulation
Thickness,
in.

Total
LCC,
$/ft

Base

20

15

3.5

$258.30

20

10

3.0

$209.13

20

20

3.5

$304.38

20

15

3.0

$223.90

20

15

3.5

$287.19

20

2.0

$153.35

20

15

3.0

$207.32

20

15

3.0

$235.42

20

20

3.5

$275.12

30

15

3.5

$302.78

10

30

15

3.5

$246.93

As can be seen from examining the table above, the life-cycle cost analysis (LCCA) is very
sensitive to the economic parameters.
While not shown in the table, in all the scenarios evaluated, as was true for the base case,
because of the fact that both 4.0 and 4.5 in. (102 and 114 mm) options require the same size conduit, the lower heat losses of the 4.5 in. (114 mm) option more than pay for its incremental cost
over the 4 in. (102 mm) option. This result would likely hold for other similar cases; i.e., once the
additional cost of a large conduit has been included it will likely be most economical from a LCC
perspective to increase the insulation thickness to the maximum possible while preserving the minimum air space.
Excepting the one scenario with what is felt to be an artificially low energy cost (#5), while all
analyses show that 3 in. (76 mm) or more of insulation provide the lowest LCC, the variation in
results given modest changes in the input parameters illustrates the prudence of performing an
LCCA for the prevailing parameters of the installation.

CALCULATING TEMPERATURES OF SYSTEM COMPONENTS AND


SURROUNDING SOIL TEMPERATURES
One of the virtues of using the thermal resistance formulation for buried piping heat
transfer calculations is that it allows for temperatures to be calculated within the system
as part of the process of determining the heat loss. The process was illustrated in most of
the previous examples, where it was used to determine insulation mean temperatures.
This process may be used to calculate the temperatures at any interface between system
components that have their resistances included in the calculation.

6.42

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

Unfortunately, estimating temperatures in soils surrounding a buried heat distribution


system is a much more difficult endeavor. One possibility is numerical methods that may
approximate any physical problem and are able to include such factors as the effect of
temperature on thermal properties. However, for most analyses, numerical analyses are
probably not warranted due to their complexity of use and the limitations on the input
data. As has been noted, there are a number of commercial computer-aided design (CAD)
programs available for numerical heat transfer, but few are able to model the true complexities of heat transfer in soil systems, including moisture migration, freezing/thawing,
and the surface heat balance. Thus, the use of numerical methods is seldom justified given
the difficulty or inability to accurately model heat transfer in soils. This is especially true
when taking into consideration the uncertainty that normally exists in parameters such as
the soil thermal properties that, as noted previously, may only be known to within 25%
even when the soil type, moisture content, and density are well known and homogeneous,
which is seldom the case. Analytical methods that are able to provide approximate solutions are much simpler and often are more than adequate given the constraints discussed.
While analytical methods in many cases must approximate boundary conditions and
physical geometries, they offer simplicity and transparency that numerical methods do
not. The analytical methods of the most use are based on exact or closed-form solutions
available for simple geometries and boundary conditions. Several basic analytical methods are presented in the subsections that follow.

Line Source of Heat


The steady horizontal line source of heat is a useful solution for approximating the
impacts of a buried pipeline with heat loss that is uniform along its length. This would be
a reasonable approximation for the situation of a piping system that was undamaged or
had uniformly degraded insulation but no leaks of the carrier medium. It would also be a
reasonable approximation for a conduit-type system with an air space that was flooded
with water from a source other than a major carrier pipe leak. This is a steady-state solution; thus, it assumes that the heat source has existed for an infinite amount of time. A
convenient form of the solution for a single steady line source of heat is (Phetteplace
1998):
q - ln x 2 + y + d 2
T x y = T o + -------------------------------------------4k s x 2 + y d 2
where
T(x,y)
To
q
ks
d

=
=
=
=
=

(6.47)

soil temperature at some arbitrary point (x,y), F (C)


undisturbed soil temperature, F (C)
heat flow per unit length from the line source, Btu/hft (W/m)
thermal conductivity of the soil, Btu/hftF (W/mK)
effective burial depth of the line source, ft (m)

An effective burial depth is used in lieu of the actual burial depth in order to account
for the thermal resistance between the air and the grounds surface. As well as allowing
the inclusion of the thermal resistance to heat transfer, using this formulation allows the
use of air temperatures that are widely available as opposed to ground surface temperatures that are seldom known. This approximation recognizes that the ultimate heat sink
is the air at the grounds surface; thus, we need to include the thermal resistance resulting from the convective heat transfer coefficient at the grounds surface. The equivalent

6.43

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

soil thickness of this resistance was defined by Equation 6.10; thus, our effective burial
depth is:
k
d = d + -----s
h

(6.48)

where d is the actual burial depth of the line source, ft (m).


With these equations we can calculate the temperature at any arbitrary point in the
soil surrounding a single line source of heat. The rate of heat loss used in Equation 6.47
may be calculated using the equations supplied previously in the chapter.
In applying the above equations to a buried heat distribution system we are making
several assumptions beyond those already discussed in this chapter:
The operating temperature of the system and environmental temperatures are not
changing.
The soil can be treated as a uniform medium with respect to its thermal properties.
The piping system can be reasonably approximated as a line source of heat.
The appropriateness of each of these assumptions depends on the specifics of the particular case at hand. Often, it is possible to make assumptions that will ensure a conservative result and/or set limits by parametric analysis.

Spherical Source of Heat


A spherical source of heat is a useful construct for modeling a carrier pipe leak on a
heat distribution system or in a situation where the thermal insulation is missing or damaged in a localized area. It might also be a reasonable approximation for a manhole. To
obtain a solution useful for our purposes we begin with the solution for a spherical source
of heat in an infinite media (Carslaw and Jaeger 1959):
r as
a s T ss
-----------T rs = T o + -------------erfc

r
2 t
where
Trs =
Tss =
as =
r
=
t
=
erfc =

(6.49)

soil temperature at the point of interest, F (C)


temperature of spherical source, F (C)
radius of spherical source of heat, ft (m)
distance from center of spherical heat source to the point of interest, ft (m)
time since heat flow started, days
complementary error function (see Abramowitz and Stegun [1972]).

This solution assumes that the spherical heat source is buried in an infinite medium.
To obtain a solution useful for a heat source buried some distance below the grounds surface (i.e., in a semi-infinite medium) we can use the mirror image method of Ingersoll
et al. (1954). This method places a fictitious second heat source of the same size but at a
temperature of Tss at a distance of twice the effective burial depth from the actual spherical source. This creates, in effect, an isothermal boundary at a point that would represent
the effective ground surface, as desired. If we also include the approximation for the convective surface resistance of Equation 6.10 we arrive at the following result:

6.44

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

r as
a s T ss
-----------T rs = T o + -------------erfc
2 t
r
a s T ss
hss + hr + a s 2 + x 2 1 2 a s
-
- erfc ---------------------------------------------------------------------------- ----------------------------------------------------------------- hss + hr + a s 2 + x 2 1 2

2 t

(6.50)

where
hr = effective depth of the point of interest, ft (m)
hss = effective depth to top of spherical heat source, ft (m)
x
= horizontal separation distance between center of spherical source of heat and the
point of interest, ft (m)
The effective depth of both the point of interest and the spherical heat source would
be calculated using Equation 6.48 with their actual depths, the soil thermal conductivity,
and the convective heat transfer coefficient at the ground surface.

Superposition
For heat transfer problems it is sometimes possible to use a principal called superposition, which allows the solution to one problem to be superimposed onto another. This
means that we can construct many solutions from a fundamental set of results for simple
geometries. One example is the superposition of two line sources of heat (i.e.,
Equation 6.47) to represent supply and return pipes of a district heating system. Another
example is that we can superimpose the result from the spherical heat source just discussed onto the solution for one or more pipes as modeled by the line source equation, as
will be shown in later calculations.
Application of the fundamental concepts presented in the previous subsections is provided in the subsections that follow.

Spherical Heat Leak


The situation of a spherical heat leak can be used to model a point leak on a heat distribution system, such as a carrier pipe leak. For systems operating above 212F (100C),
there are two obvious assumptions that one can make with respect to what temperature to
use for the spherical source. The first assumption is to assume that it is at atmospheric
boiling, since the escaping steam or high-temperature water will be at atmospheric pressure as soon as it leaves the carrier pipe and conduit (if present) and would lose any superheat obtained from expansion very quickly. Another assumption would be to set the
spherical source temperature to the carrier medium temperature, which clearly establishes
the upper limit under any scenario. Calculations have been made using each temperature
to illustrate the relative effect of these assumptions.
Before presenting any results a word of caution is in order. The results given here are
specific to this particular leak geometry, burial depth, assumed soil thermal properties,
and assumed undisturbed soil temperature. Thus these results cannot be extrapolated
directly to other situations.
For the sample calculations the following assumptions have been made:
Leak radius = 6.3 in. (160 mm), which represents a 6 in. (150 mm) nominal pipe
size (NPS) pipe with 2 in. (51 mm) of insulation and a 1 in. (25 mm) air space
all at the assumed leak temperature
Soil physical properties are taken as 15% moisture content sand/gravel soil with
a dry density of 115 lbm/ft3 (1.84 g/cm3), which results in the following thermal
properties (by use of Equations 6.1, 6.5, and 6.8):

6.45

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Thermal conductivity = 1.43 Btu/hftF (2.47 W/mK)


Thermal diffusivity = 0.92 ft2/day (0.0856 m2/day)
Convective coefficient at ground surface = 1.63 Btu/hft2F (9.25 W/m2C)
Depth to top of heat distribution pipe exterior casing = 4 ft (1.22 m)
Undisturbed soil temperature = 55F (12.8C)
Point of interest is at same depth as centerline of pipe

With these assumptions Figure 6.8 presents some results for the assumption of atmospheric boiling for the spherical source temperature (Equation 6.50). Figure 6.9 represents the same set of assumptions of Figure 6.8, except the temperature of the spherical
heat leak is 338F (170 C), the temperature of 100 psig (6.9 bar) saturated steam.
By studying Figures 6.8 and 6.9 several observations can be made. The first is that at
distances of 6 ft (1.8 m) or greater from the leak, the temperatures are never more than
approximately 7.5F (4.2C) higher than the undisturbed ground temperature for the case
of the 212F (100C) leak. For the 338F (170C) leak, at distances of 6 ft (1.8 m) or
greater from the leak the temperatures are never more than approximately 13F (7.2C)
higher than the undisturbed ground temperature. It is also apparent that in both cases there
is very little impact at distances greater than 6 ft (1.8 m) until periods of two weeks or
more have elapsed since the leak first occurred. Note that the separation distance given in
both Figures 6.8 and 6.9 is a horizontal separation distance, i.e., the point of interest is at
the same depth as the centerline of the steam/hot-water pipe. Any vertical deviation from
that depth will result in lower temperatures being predicted at the same horizontal separation distance. Also note that the separation distance is taken from the centerline of the
steam/hot-water pipe, not its exterior, which is the assumed exterior boundary of the leak.
Another observation that can be made from Figures 6.8 and 6.9 is how rapidly the
temperature drops with horizontal separation distance. Points close to the leak source
very rapidly achieve high temperatures, as would be expected, while those more distant
may take several weeks or more to achieve temperatures near the highest expected at that

Figure 6.8 Temperature as a function of horizontal separation from a spherical heat leak at
212F (100C) after persistence times of 1, 3, 7, and 30 days and 10 years.

6.46

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

separation distance. This suggests that acting promptly to repair leaks may be an effective
means of controlling the potential for heat damage in some instances. While it cannot be
discerned from these figures, at small separation distances the highest temperatures actually occur after about 30 days, then they moderate a few degrees before starting to again
climb to the ultimate value seen at very long times of leak persistence.
It is also interesting to compare predicted temperatures generated by the two different
source temperature assumptions. Close to the leak source at, for example, a horizontal
separation of 2 ft (0.61 m) for the 212F (100C) source, a soil temperature of 89F
(32C) is predicted after the leak has persisted for 30 days. For the 338F (170C) source,
a soil temperature of 116F (47C) is predicted at the same location and time of leak persistence. Thus, the effect of the higher source temperature is diminished from 126F
(70C) to only 27F (15C) at a distance of only 2 ft (0.61 m) from the leak/source.

Spherical Heat Leak with Superimposed Parallel Line Source of Heat


The scenario of a spherical heat leak with a superimposed parallel line source of heat
can be taken to simulate the situation where we have a leaking pipeline as before, but the
heat losses from the remainder of the pipe are also having a significant impact. This could
result from a situation where the heat losses from a pipeline are high due to poor, damaged, and/or missing insulation. To estimate the temperatures that adjacent buried utilities
would be subject to, we superpose the spherical heat leak result on the result from a line
source of heat that models the pipeline. The strength of the line source must be known
first; this may be calculated using the standard equations for buried heat distribution systems as outlined previously in this chapter. Appropriate adjustments must be made in
these calculations to account for damaged insulation, for example, when such is the case.
The line source strength that is calculated is then used in Equation 6.47 to find the temperature at any point of interest. Once this temperature is found it is substituted into Equation 6.50 for the undisturbed ground temperature, thus superimposing the line source

Figure 6.9 Temperature as a function of horizontal separation from a spherical heat leak at
338F (170C) after persistence times of 1, 3, 7, and 30 days and 10 years.

6.47

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

result onto the spherical heat source solution. As an example, consider superposing the
temperature impact of a single pipeline encased in a concrete enclosure as described here:
6 in. (150 mm) NPS pipe with 2 in. (51 mm) of insulation and a 1 in. (25 mm)
air space
Pipeline operating temperature = 338F (170C)
Pipe insulation is corrugated asbestos, 4 plies/in. (6.4 mm/ply), with thermal
properties per Table 6.3
Concrete enclosure = 2 in. (51 mm) thick, and its thermal conductivity is
1.13 Btu/hftF (1.96 W/mK)
Burial depth = 4 ft (1.22 m) in soil with the same physical and thermal properties as that used for the previous calculation
This results in a heat loss (line source strength) of 150 Btu/hft (143 W/m). The
results from this line source of heat are then superimposed onto the 338F (170C) spherical leak of the previous calculation. It is assumed that the pipeline is in operation and has
achieved a steady state before the spherical leak is superimposed. Figure 6.10 provides
the temperature history of points at horizontal distances of 1, 2, and 4 ft (0.30, 0.61, and
1.22 m) from such a pipeline. In the case of the 1 ft (0.30 m) horizontal separation distance, the pipeline modeled as a line source of heat adds approximately 40F (22C) to
the temperature at that point. In the case of the 2 ft (0.61 m) horizontal separation distance, the pipeline heat source adds approximately 28F (16C) to the temperature. And
finally, in the case of the 4 ft (1.22 m) horizontal separation distance, the pipeline heat
source adds approximately 18F (10C) to the temperature. From Figure 6.10 it is clear
that the combined effects of the line source and the spherical heat leak result in a significantly higher temperature at the 1 ft (0.30 m) horizontal separation distance than at the
larger separation distances.
It bears repeating that all of the results given in this section are specific to this particular leak geometry, burial depth, assumed soil thermal properties, assumed undisturbed

Figure 6.10 Spherical heat leak of Figure 6.9 with the heat loss from the pipeline
of 150 Btu/hft (143 W/m) superposed.

6.48

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

soil temperature, and pipeline heat loss. Thus, the results cannot be extrapolated directly
to other situations.

Thermal Impacts on Utilities Adjacent to Buried Heat Distribution Systems


The methods outlined in the previous subsections may be used to assess the potential
for the thermal impacts of buried heat distribution pipes on adjacent buried structures/utilities. This may be a concern when, for example, burying nonmetallic chilled-water lines
adjacent to heat distribution pipelines. The other obvious area of interest is when utilities
adjacent to buried heat distribution systems allege resulting thermal damage. Many thermal damage claims are made based on antidotal evidence such as steam rising from the
ground, items being too hot to touch, etc. While such observations may have some use in
defining the likely scenario, they are of little use in a quantitative assessment of the potential for thermal damage. Methods such as those outlined in this chapter may be used to set
reasonable bounds on temperatures that might have existed at locations where thermal
damage has been alleged or potentially could be. Along with the data and assumptions
discussed in previous sections, the material properties of the utilities sustaining damage
will need to be known, as will their service temperature limits.
There are preventative steps that a district heating utility can take to reduce the possibly of causing thermal damage to adjacent utilities and hence subsequent claims. Clearly
a routine inspection program is the first stepthis should include both a survey of the
surface along the pipeline routing as well as manhole and/or steam tunnel inspections.
The frequency of the inspections should be based on the condition of the system; clearly
systems in very good condition will warrant less frequent inspections, as would systems
operating at lower temperatures. Higher operating temperature and close proximity of
adjacent utilities susceptible to damage will warrant more frequent inspections in order to
reduce the important parameter of leak persistence time before corrective action is taken.
Heavy rain events that increase the likelihood of manhole and/or system flooding may
predicate additional inspections.
For many reasons (deep burial, long runs between manholes, piping system type,
etc.), for some direct buried systems inspections from the surface and within manholes
will not provide a comprehensive assessment of the systems condition. Inspection from
the surface and/or from manholes may also be impractical for other reasons. In these
instances, an infrared survey of the system may be useful. In addition to the possibility of
locating leaks, an infrared survey can provide reasonable estimates of heat loss on a segment-by-segment basis using the techniques developed by Zinko et al. (1996),
Phetteplace et al. (1998), and Phetteplace (1998). In addition to providing a useful planning tool, such results may also be combined with line source theory to estimate temperature impacts on adjacent utilities.

REFERENCES
Abramowitz, M., and I.A. Stegun, eds. 1972. Handbook of Mathematical Functions with
Formulas, Graphs, and Mathematical Tables, Applied Mathematics Series 55. Washington, DC: National Bureau of Standards.
Andersland, O.B., and B. Ladanyi. 2004. Frozen Ground Engineering, Second Edition.
Hoboken, NJ: John Wiley and Sons.
ASCE. 1996. Cold Regions Utilities Monograph. D.W. Smith Technical Editor. Reston,
VA: American Society of Civil Engineers.
ASHRAE. 2009. ASHRAE HandbookFundamentals. Atlanta: ASHRAE.
ASHRAE. 2011. ASHRAE HandbookHVAC Applications. Atlanta: ASHRAE.

6.49

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

ASHRAE. 2012. ASHRAE HandbookHVAC Systems and Equipment. Atlanta:


ASHRAE.
ASTM. 2006. ASTM C1114-06, Standard Test Method for Steady-State Thermal Transmission Properties by Means of the Thin-Heater Apparatus. West Conshohocken, PA:
ASTM International.
ASTM. 2010a. ASTM C177-10, Standard Test Method for Steady-State Heat Flux Measurements and Thermal Transmission Properties by Means of the Guarded-Hot-Plate
Apparatus. West Conshohocken, PA: ASTM International.
ASTM. 2010b. ASTM C518-10, Standard Test Method for Steady-State Thermal Transmission Properties by Means of the Heat Flow Meter Apparatus. West Conshohocken, PA: ASTM International.
ASTM. 2010c. ASTM C335/C335M-10e1, Standard Test Method for Steady-State Heat
Transfer Properties of Pipe Insulation. West Conshohocken, PA: ASTM International.
ASTM. 2010d. ASTM C680-10, Standard Practice for Estimate of the Heat Gain or Loss
and the Surface Temperatures of Insulated Flat, Cylindrical, and Spherical Systems
by Use of Computer Programs. West Conshohocken, PA: ASTM International.
ASTM. 2012a. ASTM C552-12b, Standard Specification for Cellular Glass Thermal
Insulation. West Conshohocken, PA: ASTM International.
ASTM. 2012b. ASTM C547-12, Standard Specification for Mineral Fiber Pipe Insulation. West Conshohocken, PA: ASTM International.
ASTM. 2012c. ASTM C240-08(2012), Standard Test Methods of Testing Cellular Glass
Insulation Block. West Conshohocken, PA: ASTM International.
ASTM. 2013. ASTM C533-13, Standard Specification for Calcium Silicate Block and
Pipe Thermal Insulation. West Conshohocken, PA: ASTM International.
Carslaw, H.S., and J.C. Jaeger. 1959. Conduction of Heat in Solids. Oxford: Oxford University Press.
CEN. 2009. EN 253:2009, District heating pipes Preinsulated bonded pipe systems for
directly buried hot water networks Pipe assembly of steel service pipe, polyurethane thermal insulation, and outer casing of polyethylene. Brussels: European Committee for Standardization.
Chyu, M.-C., X. Zeng, and L. Ye. 1997a. Performance of fibrous glass insulation subjected to underground water attack. ASHRAE Transactions 103(1):303308.
Chyu, M.-C., X. Zeng, and L. Ye. 1997b. The effect of moisture content on the performance of polyurethane insulation on a district heating and cooling pipe. ASHRAE
Transactions 103(1):30917.
Chyu, M.-C., X. Zeng, and L. Ye. 1998a. Behavior of cellular glass insulation on a DHC
pipe subjected to underground water attack. ASHRAE Transactions 104(2):16167.
Chyu, M.-C., X. Zeng, and L. Ye. 1998b. Effect of underground water attack on the performance of mineral wool pipe insulation. ASHRAE Transactions 104(2):16875.
Farouki, O.T. 1981. Thermal properties of soils. CRREL Monograph 81-1, U.S. Army
Cold Regions Research and Engineering Laboratory, Hanover, NH.
Freitag, D.R., and T. McFadden. 1997. Introduction to Cold Regions Engineering. New
York: American Society of Civil Engineers Press.
Grober, H., S. Erk, and U. Grigull. 1961. Fundamentals of Heat Transfer. New York:
McGraw-Hill.
Ingersoll, L.R., O.J. Zobel, and A.C. Ingersoll. 1954. Heat Conduction with Engineering,
Geological, and Other Applications. Madison, WI: The University of Wisconsin
Press.

6.50

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

6 Heat Transfer Calculations for Piping Systems

Kavanaugh, S.P. 2000. Investigation of methods for determining soil and rock formation
thermal properties from short-term field tests. ASHRAE Research Project 1118 Final
Report, ASHRAE, Atlanta.
Kersten, M.S. 1949. Laboratory research for the determination of the thermal properties
of soils, Engineering Experiment Station, University of Minnesota, Contract Report
for the U.S. Army Corps of Engineers, St. Paul District.
King, R., and S. Crocker. 1967. Piping Handbook, 5th Ed. New York: McGraw-Hill Book
Co.
Lunardini, V.J. 1978. Theory of n-factors and correlation of data. Proceedings of the
Third International Conference on Permafrost, Edmonton, Alberta, National
Research Council of Canada publication no. 16529.
Lunardini, V.J. 1981. Heat transfer in cold climates. New York: Van Nostrand Reinhold.
McCabe, R.E., J.J. Bender, and K.R. Potter. 1995. Subsurface ground temperature
Implications for a district cooling system. ASHRAE Journal 37(12):4045.
Meyer, V. 2013. Personal communication.
NAIMA. 2012. 3E Plus Insulation Thickness Computer Program, Ver 4.1. Alexandria,
VA: North American Insulation Manufacturers Association.
Phetteplace, G. 1998. Heat loss determination for district heating systems using surface
temperature measurements. Report Number ET-ES 98-13, Department of Energy
Engineering, Energy Systems Section, Technical University of Denmark.
Phetteplace, G., and V. Meyer. 1990. Piping for thermal distribution systems. CRREL
Internal Report 1059, U.S. Army Cold Regions Research and Engineering Laboratory, Hanover, NH.
Phetteplace, G.E., W. Willey, and M.A. Novick. 1981. Losses from the Fort Wainwright
heat distribution system. CRREL SR 81-14, U.S. Army Cold Regions Research and
Engineering Laboratory, Hanover, NH.
Phetteplace, G.E., D. Carbee, and M. Kryska. 1991. Field measurement of heat losses
from three types of heat distribution systems. CRREL SR 631, U.S. Army Cold
Regions Research and Engineering Laboratory, Hanover, NH.
Phetteplace, G., G. Pedrick, and S.K. Monaghan. 1998. Condition assessment for buried
heat distribution systems using infrared thermography. Proceedings of the EightyNinth Annual Conference of the International District Energy Association LXXXVIX:21929.
Rohsenow, W.M. 1998. Handbook of heat transfer. McGraw-Hill, New York.
RSMeans. 2013. RSMeans Online. Online databases. www.rsmeansonline.com
Siegel, R., and J.R. Howell. 1981. Thermal Radiation Heat Transfer. New York: Hemisphere.
Strock, C., and R. Koral. 1965. Handbook of Air Conditioning, Heating and Ventilating,
2d Ed. New York: Industrial Press.
Zinko, K., J. Bjrklev, H. Bjurstrm, M. Borgstrm, B. Bhm, L. Koskelainen, and
G. Phetteplace. 1996. Quantitative heat loss determination by means of infrared thermographyThe TX model, District Heating and Cooling (DHC) Annex 4 project
report, International Energy Agency (IEA). Available from the IEA DHC operating
agent: Netherlands Agency for Energy and Environment (NOVEM), Sittard, Netherlands.

6.51

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

7
Thermal Storage

INTRODUCTION
This chapter considers thermal storage of hot water in tanks for application in hotwater district heating systems. The decision as to whether or not it is feasible to install
hot-water storage in a district heating system is more an economic decision than a technical decision. The chapter focuses on design in respect to sizing of the hot-water or thermal storage tanks. It does not cover the construction details of the tanks in depth. At the
conclusion of the chapter, example installations, predominantly from Denmark, are provided.

What is Thermal Storage?


Thermal storage is traditionally considered to be short-term storage of water-based
thermal energy. The water is stored in large insulated tanks, either colocated with the primary energy source, which could be a cogeneration combined heat and power (CHP)
plant or an energy-from-waste plant, or in some instances integrated into the heat distribution network. The water content in the tank by weight should generally be constant and
independent of energy content.
Thermal storage or heat storage tanks have been used in district heating systems for
several decades. More recently a different type of heat storage, currently known as seasonal thermal storage, has been investigated, allowing heat to be stored for weeks or
months; a few such storage systems are already in operation. The store can be designed as
a pond or a pool with a recommended water depth of not less than 30 ft (10 m) and constructed with an insulated cover. The particulars of such thermal storage systems are not
addressed in this chapter, but the principles are briefly described in the last section.

The Purpose of Thermal Storage


A thermal storage system makes it possible to create a time delay between heat production and heat consumption. The purpose of such a time delay is mainly of economic
nature and is related to the fact that the cost of heat production may vary with time. By
introducing a storage tank in the district heating system it is possible to produce heat at a
time when the heat production price is low and then use this low-cost heat at a time when
the production cost is high.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

This advantage is present especially where district heating systems are supplied by
CHP plants. This means that the heat production cost is not only related to the fuel cost
and other operational costs but also to the selling price of electricity.
In a deregulated or liberalized power market, consumers are likely to buy their power
from various suppliers, depending on who has the lowest price. Hence the selling price
for power will depend more and more on supply and demand. The prices are further
affected by the introduction of wind power and hydropower, which due to their nature of
relying on fluctuating natural resources create additional fluctuations in prices. Thus,
real-time pricing of electricity with the value changing during the day, even hour by hour,
will become more prevalent. As the selling price of electricity reflects the heat production
cost, heat storage tanks are mainly used in order to optimize power production.
The production of heat from CHP is mainly related to power production in two ways:
Back-pressure steam turbine production. The ratio between electric and heat
production is fixed; an increase in electricity production will result in an
increase in heat production. Typical production equipment is back-pressure
steam turbines or piston engine installations.
Extraction/condensing steam turbine production. An increase in the heat production will decrease the electric production. Typical production equipment is
extraction steam turbines.
How the heat storage tank will be operated depends on the type of production units in
the district heating network. See Chapters 2 and 3 for more information on heat production plants.
In addition to locating the thermal storage tanks at the main energy source, thermal
storage units can also be located in the district heating distribution network, e.g., as
decentralized heat storage tanks. The purpose of these tanks can be to save heat production costs, but the decentralized location of the tanks indicates that the purpose also serves
to reduce the pipe diameters of the network and thereby the network investments. This is
something that needs careful consideration and monitoring, as the system operator could
end up in a situation where the storage tank is not charged and then not able to supply
heat to the consumers. A decentralized location of storage tanks may also be considered if
the space at the central production unit does not allow for large storage tanks.
As the installed wind turbine production capacity is increasing in many electrical
grids, a future important role for heat storage may be to use the periodic surplus power
production from the wind turbines. This may be done by direct heating of the water in the
thermal storage tanks by electrical power. However, it may be even better to use the surplus power for running heat pumps supplying heat to the district heating system and/or
the thermal storage tanks.

How Thermal Storage Works


The basic principle of thermal storage is that water at different temperatures can be
stored in the same tank without mixing, thereby allowing the amount of energy in the tank
to change without changing the amount of water. Hot water has a lower density than cold
water and will therefore remain in the upper part of the tank while the cold water will
remain in the lower part of the tank. In between the cold water at the bottom and the hot
water at the top will be a transition zone or separation layer with a temperature gradient
from the cold to the hot water. This transition zone is typically around 3 ft (1 m) high.
When charging the storage tank, hot supply water is supplied at the top of the tank
simultaneously with extraction of the same mass amount of cold water from the bottom of
the tank. The hot and cold water stay separated due to the difference in density as the tran-

7.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

7 Thermal Storage

sition zone moves down in the tank. When discharged, hot water is extracted from the top
of the tank with simultaneous supply of cold water at the bottom and the transition zone
moves up. The energy capacity of the thermal storage system is related to the amount of
cold water that may be exchanged with hot water and hot water exchanged with cold. The
water is led to and from the tank through plate diffusers in the top and bottom sections of
the tank.
Thermal storage in a heat distribution network can be considered either a consumer
unit (during charging of the tank) or a production unit (during discharge of the tank).

BENEFITS
The overall benefits of having thermal storage in the district heating system is the
saving of costs, mainly operational costs, but it also offers flexibility to the operation of
the system. The benefits vary depending on the heat production application, as discussed
in the following subsections.

CHP Plants
Two different types of CHP plants are considered in this design guide: back-pressure
plants and extraction plants for large central power stations. In addition, consideration
could be given to using thermal storage in connection with small gas CHP engines. For
large central power stations, the cogeneration principles are different but the reasons for
introducing thermal storage are in both cases based on economic assessment. The benefits
offered by heat storage for the two types of plants are therefore in many ways identical,
although the approaches may be slightly different.
Back-Pressure CHP Plant
The main functions of heat storage with this type of plant are
to supply heat to consumers and allow the CHP plant to produce the necessary
heat at times with the highest electricity pricesthis operation mode is especially important when the plant operates in a system with time-varying tariffs
and
to allow the CHP plant to produce electricity until the storage tank is fully
charged, in case of capacity problems in the electricity grid.
Extraction CHP Plant
The main functions of heat storage with this type of plant are the following:
The CHP plant is allowed to operate at optimal ratios of electricity and heat production.
During periods with low electricity prices, e.g., at night, the heat can be produced at low cost and stored. Then when the prices are high, e.g., in the morning, the heat can be supplied from the storage.
At times with high electricity prices, the heat production can be reduced or cut
off and the heat will be supplied from the thermal storage tank.

Heat Production Optimization


Operators distinguish between base-load and peak-load heat demand. The base-load
demand is normally covered by supply from high-priority production units such as CHP
plants with low production costs, whereas the peak-load demand is covered by more
expensive units such as oil-fired heat-only boilers, which have a lower priority in the production hierarchy. Independent of whether the heat production is by cogeneration CHP,
heat-only boilers, or other technologies, thermal storage can be used to optimize heat production in a number of ways.

7.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Thermal storage may be used to store excess base-load heat production during
hours in the day where there is a low heat demand. The heat can then be used
later when the demand increases. This makes it possible to avoid some of the
peak-load heat production and replace it with base-load generation.
Thermal storage in the heat supply system enables extraction CHP to be fully
exploited, as its operation can be optimized by utilizing thermal storage as a
short-term reserve capacity. Such a use is limited to the scheduled operation of
the storage, and the storage cannot at the same time serve emergency purposes
because it may be discharged at the time when an emergency occurs. Therefore,
the storage does not add production capacity to the system.
Thermal storage can be used for smoothing the variation in the requirement for
peak production. Often peak-load boilers have a given output or a minimum
working output that may not always correspond to the encountered peak-load
demand. In such cases the introduction of excessive peak-load production will
cause the base-load production to have to modulate down. Thermal storage can
be designed so that it can absorb these fluctuations in marginal peak capacity
requirements through modulation.
Thermal storage can allow a total shutdown of the plant during weekends when
electricity prices are low or when the operator wants to avoid the delivery or
handling of fuel. The benefit of total shutdown over, for instance, a weekend is
the avoided potential extra cost for weekend delivery, including maintenance
staff working weekends. Also, the plant may be located in an area where it
would be good to avoid heavy truck transport of biomass during weekends when
most people are at home.
Production Capacity Reduction and Leveling
Thermal storage use to replace expensive low-priority heat production with more
cost-effective high-priority production normally occurs on a daily basis (24 or 48 hours)
as opposed to, for example, on a seasonal basis. Therefore, it is important to charge and
discharge the storage in such a way that the daily heat production is leveled to avoid peakload production at low-priority production units. At the same time, the selection of heat
production units to charge the storage depends on the short-term heat demand and the
high-priority base-load production capacity.
When the maximum heat capacity demand over a short time period is lower than the
high-priority production capacity (base-load production), there is no reason to use the
heat stored in the tank. When the minimum heat demand requirements in a certain period
are higher than the high-priority production capacity (base-load production), there is no
high-priority capacity that can used for charging the tank once it has been discharged. The
tank should therefore be left discharged until high-priority production capacity is available. Between these two scenarios, thermal storage can replace low-priority capacity with
high-priority capacity and thus present an economic benefit.
The amount of low-priority capacity that can be replaced by high-priority capacity,
and hence tank size, depends on the annual and daily variations.

OPERATION
Before a thermal storage system is designed it is important to consider the operation
of the district heating network to which it will be connected. It is also important to consider the supply and return temperatures and the impact they will have on pressurized versus nonpressurized storage and the T that might be achieved.

7.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

7 Thermal Storage

Principles of Operation
The thermal storage tank generally maintains a constant mass of water, the change in
volume depending on the temperature of the stored water. If the maximum temperature is
lower than the boiling point of the water, the tank may be operated at atmospheric pressure. In practice, a minor positive (gage) pressure is maintained at the top of the tank and
a steam or nitrogen blanket there prevents the oxygen from the atmosphere from diffusing
into the water.
In large hot-water district heating transmission systems the maximum temperature is
often above 212F (100C) and typically up to 248F (120C). If water with that temperature were stored in a tank at atmospheric pressure, permanent boiling would occur in the
top section of the tank. In order to avoid this, pressurized tanks must be used.
Thermal Storage Directly Connected to the District Heating Network
As a consequence of changing temperatures in the district heating network, some
expansion and contraction of the water in the network will take place. Therefore, water
must continuously be drained from or supplied to the network in order to maintain the
required pressure level. If no thermal storage tank is connected to the network, the pressure level is often maintained by connecting a pressurized vessel with a gas volume.
Water from the network is drained into this vessel if the pressure gets too high, or water
may be supplied from this vessel if the pressure in the network gets too low.
If a thermal storage tank is hydraulically connected directly to the network, the tank
may be used for pressurization of the district heating network. The pressure at the connection point between the tank and the network will be determined by the level of the water
surface in the tank. This may be sufficient for smaller networks if the supplied area is relatively flat or if the storage tank is located at a high position in the network.
If thermal storage is used for pressurization of the network, the design of the tank
should allow extra volume for the expansion of the water due to charging of the tank
(replacement of cold water with hot water) and also extra volume for the expansion of the
water in the distribution network.
In the integrated thermal store system shown in Figure 7.1, the network pump is controlled in order to fulfill the consumption requirements by maintaining a proper differential
pressure in all parts of the network. The production pump is controlled in order to remove
the generated heat from the production unit. This is often done by regulating the pump
speed in order to maintain the required supply temperature from the production unit.
An important feature of this type of thermal storage installation is that the charge and
discharge of the storage is done without further control or control equipment. The charge
or discharge of the storage is instead controlled by small variations in differential pressure
between the hot and cold inlet flanges to the storage tank. This differential pressure is
generated by the production pump and the network pump. If the pressure at the hot flange
is higher than the pressure at the cold flange (positive differential pressure), the storage is
charged; if the differential pressure is negative, the storage is discharged.
It should be noted that in general only one heat storage installation of this type can be
installed in a district heating system because the tank controls the energy level in the
whole system. If there are two tanks, one needs to be hydraulically separated from the
network.
Thermal Storage with Pressure Separation from the District Heating Network
Alternatively to the concept where the network is pressurized by the thermal storage
tank, the storage system may be connected to the network in a way that separates the
static pressure in the tank from that in the network. Such a separation can be provided by
a pump and valve configuration as shown in Figure 7.2.

7.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 7.1 Thermal storage directly connected to a district heating network.

Figure 7.2 Thermal storage with pressure separation from the heat network.

7.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

7 Thermal Storage

In this configuration, the pressurization of the network is maintained by a separate


expansion system with pressurized vessels and makeup water supply. The flow to and
from the thermal storage will have to be controlled by special controllers. The flow at one
of the connections, for instance the hot connection at the top of the tank, is regulated in
order to maintain a constant mass (not volume!) of water in the tank. The flow at the other
connection is regulated in order to achieve the required charge or discharge.
The concept with pressure separation of the thermal storage is far more complex than
the concept where the network is pressurized by the thermal storage tank. However, pressure-separated thermal storage tanks may be required in situations where large elevation
differences exist in the heat distribution network. This configuration is also used when the
thermal storage system is required to be installed in a higher-pressure network, e.g., 16 or
25 bar (232 or 363 psi), as it is likely to be too expensive at these pressure levels to install
a thermal storage system connected to the network. An advantage of the pressure-separated thermal storage system is that multiple thermal storage installations may be implemented at different locations in the same district heating system.
The pressure separation between the thermal storage and the network could also be
accomplished by installing heat exchangers instead of the pump and valve configuration,
but this will result in loss of temperature, which will reduce the storage capacity and
make continuous operation of the tank more difficult.

Storage Tank Monitoring


Temperature measurement is achieved either through the use of a series of temperature transmitters mounted at regular intervals through the height of the store or through
the use of a multipoint thermometer inserted from the top of the tank (see Figure 7.3). The

Figure 7.3 General layout of a thermal storage system.

7.7

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

externally mounted transmitters should be complete within stainless steel pockets to


allow the sensors to be changed in the event of instrument failure without having to isolate the tank and drain the entire water contents. The multipoint thermometer should be
flexible to allow it to be removed without any height restrictions.
The upper surface level in a thermal storage tank will vary; it can be monitored
through the use of a level transmitter. A differential pressure transmitter can be used to
calculate the height of the water column. Level control is provided by level switches to
alarm for high and low levels in the storage tank.

ChargingDirectly Connected Heat Storage


Figure 7.4 illustrates the basic principle of charging a thermal store directly connected to the heat distribution network. Cold water returning from the heat network is
pumped through the heat source labeled HS1 in the figure using pump HP1. The flow rate
of HP1 is controlled by temperature transmitter TT, which is placed after the heat source.
In a charging scenario the supply pump SP1 delivers the required heat to the network,
which is controlled by a differential pressure sensor DP at the end of the network. Excess
heat is directed into the thermal storage system, displacing cooler water, which is returned
to the network.

DischargingDirectly Connected Heat Storage


Figure 7.5 illustrates the basic principle of discharging a thermal store that is directly
connected to the heat distribution network. When additional heat is required by the distribution network, the thermal storage tank can be called on to discharge its content to assist
the heat source. In this instance, returning cold water from the heat distribution network is
directed to both the thermal storage tank and the heat source. HP1 is controlled by the
temperature transmitter TT. The supply pump SP1 delivers the required heat to the network controlled based on the differential pressure measured by a sensor at the critical
consumer of the network.

Figure 7.4 Principles of pressurization using thermal storagecharging.

7.8

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

7 Thermal Storage

Charging/DischargingPressurized and Decentralized Tank


In the previous subsections, the interaction between the thermal storage tank and the
district heating network was presented. It was shown how connection to an atmospheric
tank at the production unit could be arranged in such a way that the charge and discharge
are carried out on basis of the pressure condition in the heat network without a specific
control mechanism. However, for the pressurized tank installation and for a decentralized
tank installation the charge and discharge must be controlled with an external control
algorithm.
Mass Flow Regulation
The control algorithm shall ensure the following:
that the mass flow through the storage tank is as required and
that the total mass of the water in the heat storage tank remains unchanged.
The mass flow through the storage tank may be specified in several ways. For
instance, the flow can be specified by the operator of the system in order to obtain the
required charge or discharge (in Btu/h [MW]) of the tank. The basis for specifying what
the flow should be at a certain time can be very complex.
The flow through the heat storage tank may also be determined in order to fulfill certain requirements in the district heating network. For a decentralized storage tank the flow
through the tank can be controlled in order to maintain a specified differential pressure in
the network at the location of the tank.
If the differential pressure in the network increases it is an indication of surplus in the
supply to the network, and in this case the charge flow should be increased. If the differential pressure decreases it is an indication of a supply deficit at the district heating consumers. Figure 7.6 shows the general principle of a mass regulation system.

Figure 7.5 Principles of pressurization using thermal storagedischarging.

7.9

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure 7.6 Mass flow control.

Capacity Regulation
As an alternative to the mass flow regulation concept, where the flow through the
thermal storage system is adjusted according to a mass balance, it is also possible that the
flow is regulated in order to maintain a specified water level in the pressurized vessel of
the expansion system.
The concept of this method is that if a net amount of water mass is flowing from the
heat storage tank to the network, then the water level in the pressurized vessel of the
expansion system will increase. This can be used as a signal to reduce the flow out of the
tank and to increase the flow into the tank.
Operated in this way, the heat storage tank becomes part of the pressurization system
for the heat distribution network. However, the pressurization and expansion system must
still be in operation in case the water level in the pressurized expansion vessel keeps
increasing. In this case, the pressurization system should drain water from the system, if
possible to a separate atmospheric water storage tank, for later reuse as makeup water.
Figure 7.7 shows the general principle of a capacity regulation system.

Water Quality
It is essential that the district heating system water is maintained at the highest quality
for the long-term protection of the thermal storage tank. Water can remain in the storage
tank for long periods of time, particularly when the tank is shut off from the remainder of
the system, so it is important that it is treated to prevent corrosion. Treatment of the district heating water with chemicals is a normal procedure in any good-quality district heating system.

7.10

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

7 Thermal Storage

Figure 7.7 Capacity control.

In the case of a thermal storage tank with surplus volume for expansion of the water,
it is necessary to displace the air in the top of the tank in order to protect the water from
contamination with oxygen and consequently corrosion. As protection, steam or nitrogen
may be used to displace the air. Steam is often chosen for large storage tanks at power
plants where steam is already generated for other purposes. The advantage of steam is
that it will not interfere with the district heating water in the tank. The disadvantage is that
it may condense at cold parts of the tank, thereby leading to a collapse of the steam cushion being used to displace the air. Nitrogen is often used in smaller tanks connected to
decentralized CHP plants. It may be produced on site by nitrogen generator equipment.
The disadvantage is that nitrogen may diffuse into the district heating water in the tank.
The advantage is that the nitrogen blanket is relatively simple to maintain.

SPECIFIC DESIGN ISSUES


Temperature
In systems with a supply temperature below 212F (100C), the thermal storage tank
will normally be designed as a nonpressurized tank; above 212F (100C) the thermal
storage tank must be designed as a pressure vessel. In systems where operation with supply temperatures above 212F (100C) is limited to a few hours in winter, it is still possible to have a nonpressurized tank, either by bypassing the thermal storage tank or by
boosting the temperature after the tank. In this case the heat storage installation should be
equipped with a shunt installation for reducing the water temperature to below 212F
(100C) before the water enters the storage tank.

7.11

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Thermal storage systems should be designed with as large a temperature difference as


possible. The district heating return water temperature will predetermine the cold temperature and the hot temperature will be limited to the primary energy source temperature
and/or whether the thermal storage tank is a pressurized or nonpressurized design. For
nonpressurized storage tanks the temperature difference should be 54F to 72F (30C to
40C) and for pressurized storage tanks 90F to 99F (50C to 55C).

Pressure
In systems with a static pressure up to around 85 psig (6 bar) and a supply temperature below 212F (100C), it is common to use nonpressurized tanks. Depending on local
conditions, nonpressurized tanks may also be used in systems with higher pressures, but it
should be noted that there is a direct relationship between the height of the tank and the
pressure in the district heating system at the point of connection. This will result in some
restrictions in the system layout that may be more critical in a system with a higher static
pressure level.
Nonpressurized thermal store tanks are normally designed according to a standard for
vertical oil or water tanks. Pressurized tanks are designed in line with a standard for pressure vessels.

Sizing
Tanks
Sizing of the stored energy content is a cost-benefit analysis that will be based both
on well-known parameters as well as those subject to uncertainty. A well-known parameter could be the capital investment in the tank; an uncertain parameter could be the electricity market price. The calculation of the active storage volume between the diffusers
should be based on the energy content and the difference between supply and return
design temperatures as well as the diffuser design. It should be taken into account that the
unusable transition zone or separation layer will typically have a height or thickness of
approximately 3 ft (1 m). The choice of temperatures is very important, as the temperatures have a direct impact on the size of the tank.
The ratio between the height and the diameter of the tank is an important parameter.
On one hand, the heat loss from the thermal storage to the surroundings should be minimized, which is accomplished by minimizing the surface of the tank in relation to the volume. On the other hand, the proportion of the tank volume taken up by the transition zone
should be minimized, which is accomplished by having a tall tank with a relatively small
diameter. Designers have traditionally preferred a height/diameter ratio above 1.5 in order
to minimize the volume of the inactive separation layer.
Diffusers
Thermal storage tanks should be equipped with diffusers for charging and discharging the tanks. The diffusers used for charging and discharging of the hot water should be
located at the top of the tank. The diffusers used for charging and discharging the cold
water should be located at the bottom of the tank. The diameters of the pipes connecting
to the diffusers should be calculated based on the nominal maximum volume flow rate of
the storage. The diffuser pipes should be insulated in order to avoid affecting the thermal
zones in the tank.
The diffusers should be designed with guiding plates so that the flow will be even and
the diffuser will operate with the least fluid turbulence possible. The velocity through the
diffuser should be kept as low as possible and not above approximately 0.12 ft/s (0.04 m/s),

7.12

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

7 Thermal Storage

but this should be assessed for each case individually. Poor velocity control can, in some
instances, render the thermal storage unusable.
The outside diameter of the diffusers should not be larger than around one third of the
tank diameter, as illustrated in Figure 7.8. In general, both the bottom diffuser and the top
diffuser are fixed in the storage tank, but designs may be found where the top diffuser is
floating on the water. This construction will result in a better utilization of the tank, but
the piping construction will become more complicated.
For practical reasons there will have to be some volume beneath the lower diffuser
and above the top diffuser (in the case of fixed diffusers). These volumes of water cannot
be used for storing energy; this is also the case with the transition zone or separation layer
between the cold and the hot water. Therefore, the volume of water in the tank will be
larger than the volume that may actively be used for energy storage.

Regulatory Requirements (Europe)


In Europe, thermal storage tanks should be designed according to EN 14015 (CEN
2004) for atmospheric tanks and the Pressure Equipment Directive (EU 1997) for pressurized tanks.
When calculating the required thickness of the shell, roof, and bottom plating, a corrosion allowance of (0.04 in.) 1 mm should be added to all plating. Furthermore, evidence
should be provided that the shell of the tanks is assessed against fatigue fractures based
upon the transition zone passing any level in the shell twice a day for 365 days per year
during a lifetime of 45 years, i.e., a total of 33,000 cycles. The tank design should allow
for corrosion-protective surface treatment including the sandblasting of the inside shell,
roof, and bottom surfaces and all quality assurance requirements specified in EN 14015
(CEN 2004).

Figure 7.8 Diffuser sizing example.

7.13

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

ECONOMICS
The purpose of thermal storage is mainly of an economic nature. Therefore, the size
of the tank must be based on an economic analysis.
In the economic analysis the saving potential due to the thermal storage tank(s) is
estimated in relation to the tank investment. There are several economical analysis tools
that can be used for such an analysis. In the following subsection the net present value
(NPV) method is briefly discussed in relation to heat storage, as this is the most commonly used method. For economic analysis methods, including the NPV method, see
Chapter 37 of ASHRAE HandbookHVAC Applications (2011).

Net Present Value


Savings related to heat production costs will occur as daily savings, which can be
summarized over a year, resulting in an annual saving. A saving obtained in the future
will not have the same value as a similar saving obtained at the present time. This is partly
because of the inflation that normally causes a decrease in the value at a specified rate
over the years and partly because there is the actual bank interest rate that will result in an
increase during the years at a certain rate.
The NPV method can be used to evaluate the benefit of the thermal storage system.
Of course the aim is to get as high an NPV as possible; a negative NPV indicates that the
investment in thermal storage is not economically feasible and that it will result in a deficit.
How large the NPV should be in order to opt for the implementation of a thermal
storage tank is not easy to determine, and it is generally up to the individual project and
its objectives. The acceptable NPV will certainly depend on the uncertainty in calculating
the value. The larger the uncertainty, the larger the NPV should be before it is accepted.
One of the main uncertainties in calculating the NPV is the assessment of the annual
savings. These will depend on the energy production costs, which again depend on the
future fuel prices, which are very difficult to predict with the required accuracy. Thus, it is
prudent to calculate a number of NPVs based on various scenarios for the development of
a general price level and a fuel price level.
The following information is required in order to accurately size a heat storage tank:
Daily load variations during the year through a number of years
The possible savings related to the production units due to the installation of the
storage tank; the possible savings will typically be related to an optimized power
production in relation to large changes in the selling price of electricity
Prices for the heat storage tank installation, including pumps, pipes, valves, etc.
The calculations can be made more or less complex. For simplicity, often only two or
three load situations are considered in the calculations. For a more complete analysis,
more load situations will have to be taken into account and preferably the analysis should
be based on an hour-by-hour load during the year.
Furthermore, if the thermal storage system is decentralized (not at the production
site) with the purpose of minimizing the investments in the district heating pipe system,
the dimension of the tank should be based on hydraulic calculations of the network. This
will, however, make the dimensioning of the tank more complex.

SEASONAL THERMAL STORAGE


A seasonal thermal storage scheme is designed to retain heat stored during the summer months for use during the fall or winter. The heat is typically produced by solar thermal panels, although other heat sources may be used either separately or in parallel. The

7.14

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

7 Thermal Storage

increased application of solar thermal heat production and wind power is the main reason
for the increased interest and development of seasonal thermal storage.
There are various types of seasonal thermal storage: borehole, aquifer, or pit storage.
Borehole storage has limitations as to the temperature at which the water can be stored, as
it is the soil itself that provides the medium insulating capacity for storing the heat. The
same applies for aquifers; here in addition groundwater penetration can be an issue.
Pit storage (also called pond storage or pool storage) can, on the other hand, be a natural part of larger district heating systems. Large-scale storage capacity, i.e.,
2.5 million ft3 (75,000 m3) or 3.5 million ft3 (100,000 m3), can be established under parking lots or sports facilities or in areas used for large infrastructure construction. Tests
show that hot water at 195F to 205F (90C to 95C) can be stored with little heat loss
for several months. The larger the pit store, the smaller the heat loss, and the temperature
loss is only around 2F (1C) a week or month, depending on size and design. With pit
storage the water is heated to the required temperature and pumped into the insulated
store until it is fully charged. The pit is normally 3050 ft (1015 m) deep, with sloped
sides. Due to the shallowness, there is no stratification layer.

EXAMPLES OF THERMAL STORAGE


Thermal storage comes in all sizes, depending on the size of the district heating
scheme and its production and demand. The following subsections discuss example
installations to give the reader an idea of the sizes and types of established installations.

Denmark
In Denmark thermal storage is used at almost all power plants with district heat production. Table 7.1 includes the major thermal storage systems in Denmark, both pressurized and atmospheric thermal storage systems. While the examples provided in Table 7.1
are nearly all CHP plants, it should be noted that district heating schemes not associated
with a power station may also have thermal storage. There are more than 400 district heating schemes in Denmark, but there are no central records of which plants have a thermal
storage system or their characteristics.
Table 7.1 Sample Thermal Storage Applications of Danish District Heating Systems
Project

Type

Volume,
ft3 (m3)

Slagelse CHP

Atmospheric

38,700 (3600)

Hillerd CHP

Atmospheric

172,000 (16000)

Esbjerg CHP

Atmospheric

592,000 (55000)

Helsingr CHP

Atmospheric

156,000 (14500)

Avedre CHP

Pressurized

2 @ 258,000 (2 @ 24000)

Madsned CHP

Atmospheric

53,800 (5000)

Silkeborg CHP

Atmospheric

2 @ 151,000 (2 @ 14000)

stkraft

Atmospheric

72,100 (6700)

Studstrup CHP

Pressurized

355,000 (33000)

Skrbk CHP

Pressurized

301,000 (28000)

Nordjylland CHP

Atmospheric

269,000 (25000)

Denmark Technical University CHP

Atmospheric

86,100 (8000)

Maribo CHP

Atmospheric

64,600 (6000)

Asns CHP

Atmospheric

215,000 (20000)

Amager CHP

Pressurized

258,000 (24000)

Fyn CHP

Atmospheric

807,000 (75000)

7.15

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

One of the larger thermal storage systems in use in Denmark is the system of the
Avedre CHP. Avedre Power Station is situated south of Copenhagen and consists of
two power station units, Unit 1 from 1990 and Unit 2 from 2001. The thermal storage
facility of the Avedre Power Station (Figure 7.9) consists of two identical pressurized
tanks, each with an energy content of around 3800 MMBtu (4000 GJ). The district heating network that it is connected to has a total annual heat consumption in excess of 2.6
107 MMBtu (28,000 TJ).
The overall production capacity of Avedre Power Stations Unit 1 and Unit 2 is
2764 MMBtu/h (810 MW) of electricity and 3120 MMBtu/h (915 MJ/s) of heat. Unit 1
primarily uses coal, while Unit 2 can use a wide variety of fuels: natural gas, oil, straw,
and wood pellets. Unit 2 has facilities consisting of several parts. These parts combined
make record-high use of the energy in the fuels. By simultaneously generating heat and
electricity, Avedre Power Stations Unit 2 uses as much as 94% of the energy in the fuels
and has an electrical efficiency of 49%, making the unit one of the most efficient in the
world.

Other Countries
The Netherlands are also known for using thermal storage in their district heating systems, as they have much of their production focused around CHP. The authors are aware
of two atmospheric tanks in Rotterdam, one 86,100 ft3 (8000 m3) tank is located at the
CHP plant and one 64,600 ft3 (6000 m3) tank is positioned remote in the network. In
Amsterdam there are three 16,100 ft3 (1500 m3) storage tanks pressurized to 58 psig
(4 bar) installed at the CHP plant.
In the UK the thermal storage tanks tend to be smaller as the district heating schemes
are smaller. Often thermal storage systems are installed within the energy center building
and may be as small as 110 ft3 (10 m3), or even smaller. However, here there is often
focus on storage of domestic hot water in small building-based installations rather than

Figure 7.9 Thermal storage tanks of the Avedre Power Station.

7.16

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

7 Thermal Storage

the large-scale storage as seen, for example, with district heating in Denmark. One of the
larger thermal storage systems known to the authors is being planned for the Shetlands
and is more than 32,300 ft3 (3000 m3) in size. Most, however, are less than 10,800 ft3
(1000 m3). In Pimlico, London, the oldest district heating scheme in the UK, there is a
26,900 ft3 (2500 m3) storage tank.
Sweden and Finland also have thermal storage in some of their district heating systems.

REFERENCES
ASHRAE. 2011. ASHRAE HandbookHVAC Applications. Atlanta: ASHRAE.
CEN. 2004. EN 14015:2004, Specification for the design and manufacture of site built,
vertical, cylindrical, flat-bottomed, above ground, welded, steel tanks for the storage
of liquids at ambient temperature and above. Brussels: European Committee for
Standardization.
EU. 1997. Pressure Equipment Directive. Directive 97/23/EC of the European Parliament
and of the Council. Brussels: European Union.

BIBLIOGRAPHY
Hyman, L.B. 2011. Sustainable Thermal Storage Systems: Planning, Design, and Operations. New York: McGraw Hill.

7.17

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8
Operation
and Maintenance
INTRODUCTION
This chapter is devoted to operation and maintenance (O&M) of district heating systems. Topics covered include:
Workplace safety
System security
System operation
Water treatment and filtration
Maintenance
The coverage here is by no means exhaustive but should provide an overview for the
district heating system operator as well as provide references to additional sources of
information.
The importance of O&M practices to the safety and success of a district heating system cannot be overemphasized. Even systems of superior quality in design, component
selection, and construction can become inefficient and unreliable and have their lifetimes
drastically shortened by poor O&M practices. In addition, and more importantly, poor
O&M practices can lead to unsafe facilities that can endanger not only the workforce but
the public at large.
To achieve a successful O&M program, the management of the district heating system must be committed. This is a necessary condition, but not sufficient in itself. Programs must be organized such that execution becomes part of the everyday routine. The
establishment of metrics can also be an important aspect of a successful O&M program.
For safety, metrics such as periods of time since injuries have been used. For maintenance, simply performing tasks on schedule may represent one end of the scale, whereas
at the other end of the scale more proactive management will use predictive maintenance
augmented by methods such as vibration measurement for major pieces of equipment, use
of corrosion coupons, and infrared thermography, for example. In addition, regular compilation and review of metrics such as overall plant heat rate may be very useful in identifying the effectiveness of both maintenance and operations practices.
The effective O&M of a district heating system should achieve the following goals:
High reliability of the district heating plant and distribution piping system
(including redundancy) with minimum forced downtime
High energy efficiency of thermal energy production and distribution

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Compliance with environmental requirements


Minimum cost of thermal energy
Sustainability
In order to accomplish the above outlined goals, the operating personnel should
implement and use the following procedures and practices:
Optimum dispatch and utilization of plant equipment (for example, minimum
operation of the equipment at part loads) utilizing a supervisory control and data
acquisition (SCADA) system described in Chapter 3 of this design guide
Most energy-efficient and environmentally sustainable plant operation
Preventive maintenance program utilizing a computerized maintenance management system (CMMS)
Workplace safety
Plant fire and security protection
Continuous training and education of the O&M personnel
Many codes and standards exist that offer guidance and accepted practicessome
mandatory. The applicability of these codes and standards may vary from one jurisdiction
to another. ASHRAE HandbookHVAC Systems and Equipment (2008) provides a comprehensive overview of codes and standards for HVAC systems with a number of these being
applicable to district heating systems. Other sources of guidance/standards/regulatory
requirements for the United States are cited in this chapter where appropriate. The intent is
not to provide a comprehensive compilation of guidance/standards/regulatory requirements
for the United States; it is the responsibility of the district heating owner/operator regardless
of jurisdiction to be cognizant of all applicable guidance/standards/regulatory requirements
impacting the design, construction, and operation of the system. Obviously the same
responsibility extends to district heating systems outside of the United States.

WORKPLACE SAFETY
In most jurisdictions agencies exist that are responsible by law for ensuring workplace safety. Within the U.S., the cognizant agency in most cases is Occupational Safety
and Health Administration (OSHA), which is part of the U.S. Department of Labor. The
Occupational Safety and Health Act of 1970 (U.S. Congress 1970) created OSHA to help
employers and employees reduce injuries, illnesses, and deaths on the job. Since then,
workplace fatalities have been cut by more than 60% and occupational injury and illness
rates have declined 40% despite the fact that U.S. employment has more than doubled
during that same period. However, significant hazards and unsafe conditions still exist in
U.S. workplaces. Each year almost 5200 Americans die from workplace injuries in the
private sector, perhaps as many as 50,000 employees die from illnesses in which workplace exposures were a contributing factor, and nearly 4.3 million people suffer nonfatal
workplace injuries and illnesses. The cost of occupational injuries and illnesses totals
more than $156 billion each year.
Thus, district heating providers, like all businesses that operate in environments that
have hazards, must have workplace safety programs in place. The experiences of one district heating provider in developing an effective workplace safety program are described
by Toth and Merrill (2009).

Requirements
In general, OSHA standards require that employers maintain conditions or adopt
practices reasonably necessary and appropriate to protect workers on the job, be familiar
with and comply with standards applicable to their establishments, and ensure that

8.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

employees have and use personal protective equipment when required for safety and
health.

Hazards
Aside from the expected hazards on district heating system construction/repair sites,
many workplace hazards are found in the central plants of district heating systems. Workplace hazards may also exist in distribution system valve vaults and even within the enduser building where the consumer interface is located. For example, OSHA issues standards for a wide variety of workplace hazards, including toxic substances, harmful physical agents, noise, electrical hazards, fall hazards, trenching hazards, hazardous waste,
infectious diseases, fire and explosion hazards, dangerous atmospheres, machine hazards,
and confined spaces, among others. In addition, where there are no specific OSHA standards, employers must comply with the Occupational Safety and Health Acts General
Duty Clause, Section 5(a)(1), which requires that each employer furnish... a place of
employment which [is] free from recognized hazards that are causing or are likely to
cause death or serious physical harm to his employees (U.S. Congress 1970). For more
information on OSHA and their guidelines and standards, refer to their comprehensive
website, www.osha.gov.

SYSTEM SECURITY
Beyond the normal security that is dictated to protect property from loss due to theft
or vandalism, appropriate security must also be provided for district heating system facilities that might put the public at risk. Examples include protection of fuel storage facilities, refrigerant storage, water treatment chemicals, large thermal storage tanks,
distribution system valve vaults, etc. Increased physical security of fixed facilities has
received much attention after the terrorist attacks of September 11, 2001, in the United
States. A number of sources of information are available for both the design of new facilities and the retrofit of existing facilities; see, for example, Whole Building Design
Guides Secure/Safe, www.wbdg.org/design/secure_safe.php (NIBS 2012).

SYSTEM OPERATION
An operator needs to look at information from a wide array of systems, not just the
distributed control system. There also are fewer personnel in district heating plants today,
so the operators scope of work has widened. One advantage of district heating plants is
easier implementation of automation because the major and ancillary equipment is consolidated in one location. Computerized automatic controls can significantly affect system performance. A SCADA system to control the district heating plant and overall
system performance should be used by the plant operator. Such a system allows a single
operator to monitor performance at many points in the plant.
Software to consider when designing, managing, and improving district heating plant
performance should include the following (DA 1989):
Automatic controls that can interface with other control software (e.g., equipment manufacturers unit-mounted controls)
Energy management system (EMS) control
Hydraulic modeling, as well as metering and monitoring of distribution systems
Computer-aided facility management (CAFM) for integrating other software
(e.g., record drawings, O&M manuals, asset database)
Computerized maintenance management system (CMMS)
Automation from other trades (e.g., fire alarm, life safety, medical gases, etc.)
Regulatory functions (e.g., federal, state, and local agencies, etc.)

8.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Automatic controls of a district heating system may include standard equipment manufacturer control logic along with optional, enhanced energy-efficiency control logic.
Beyond standard control technology, the following control points and strategies may be
needed for primary and ancillary equipment as well as the overall system:
Discharge temperature and/or pressure
Return distribution medium temperature
Head pressure for distribution medium
Stack temperature
Carbon monoxide (CO) and/or carbon dioxide (CO2) level
Differential pressures
Flow rate of distribution medium
Peak and hourly heat energy output
Flow rate of water and fuel(s)
Reset control of temperature and/or pressure
Night setback
Variable flow through equipment and/or system control
Variable-frequency drive (VFD) control
Thermal storage control
Heat recovery cycle
Computerized energy management and control systems provide an excellent means
of reducing utility costs associated with the operation of district heating systems. These
systems can incorporate advanced control strategies that respond to changing weather,
user conditions, and utility rates to minimize operating costs.

WATER TREATMENT AND FILITRATION


A proper water treatment program is paramount to ensure district heating system
component life, operating efficiency (including customer connections), and control of
potential public health threats. Water quality is controlled by a combination of chemical
treatments and mechanical filtration. Chemical treatment is used to control corrosion,
scaling, biofouling, and bio threats. Mechanical filtration is used to remove suspended
particles. The following discussion of water treatment issues for district heating systems
comes largely from ASHRAE HandbookHVAC Applications (2011), to which the reader
is referred for additional details.

Corrosion
Corrosion is the destruction of a metal or alloy by chemical or electrochemical reaction with its environment. In most instances, this reaction is electrochemical in nature,
much like that in an electric battery. For corrosion to occur, a corrosion cell consisting of
an anode, a cathode, an electrolyte, and an electrical connection must exist. Metal ions
dissolve into the electrolyte (water) at the anode. Electrically charged particles (electrons)
are left behind. These electrons flow through the metal to other points (cathodes) where
electron-consuming reactions occur. The result of this activity is the loss of metal and
often the formation of a deposit. Various types of corrosion exist: general, localized or
pitting, galvanic, etc. Refer to ASHRAE HandbookHVAC Applications (2011) for additional details on the types of corrosion and the factors that contribute to corrosion. For
corrosion issues specific to dissimilar metals in piping systems, refer to the ASHRAE
Journal article by W. Sperko (2009).

8.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

Corrosion Protection and Preventive Measures


Materials Selection
Any piece of equipment can be made of metals that are virtually corrosion-proof
under normal and typical operating conditions. However, economics usually dictate material choices. When selecting construction materials, the following factors should be considered:
Corrosion resistance of the metal in the operating environment
Corrosion products that may be formed and their effects on equipment operation
Ease of construction using a particular material
Design and fabrication limitations on corrosion potential
Economics of construction, operation, and maintenance during the projected life
of the equipment (i.e., consider that expenses may be minimized in the long run
by paying more for a corrosion-resistant material and avoiding regular maintenance)
Compatibility of chemical additives with materials in the system
Also note that use of dissimilar metals should be avoided; where dissimilar materials
are used, insulating gaskets and/or organic coatings must be used to prevent galvanic corrosion.
Protective Coatings
The operating environment plays a significant role in the selection of protective coatings. Even with a coating suited for that environment, the protective material depends on
the adhesion of the coating to the base material, which itself depends on the surface preparation and application technique.
Maintenance of Protective Coating
Defects in a coating are difficult to prevent. These defects can be either flaws introduced into the coating during application or mechanical damage sustained after application. Where access to coated surfaces is available, the surfaces should be regularly
inspected visually or using methods such as holiday detection. Where access is not available, inspection will normally be impossiblewhich emphasizes the need for highintegrity coating and quality control doing initial installation. In any event, in order to
maintain corrosion protection, defects must be repaired.
Cycles of Concentration
Some corrosion control may be achieved by optimizing the cycles of concentration
(the degree to which soluble mineral solids originating in the makeup water have
increased in the circulating water due to evaporation). Generally, adjustment of the blowdown rate and pH to produce a slightly scale-forming condition (see the next section,
Scale Control) will result in an optimum condition between excess corrosion and excess
scale.
Chemical Methods
Protective film-forming chemical inhibitors reduce or stop corrosion by interfering
with the corrosion mechanism. Inhibitors usually affect either the anode or the cathode.
There are three basic types of inhibitors:
Anodic corrosion inhibitors establish a protective film on the anode. Though
these inhibitors can be effective, they can be dangerousif insufficient anodic
inhibitor is present, the entire corrosion potential occurs at the unprotected
anode sites. This causes severe localized (or pitting) attack.

8.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Table 8.1 Typical Corrosion Inhibitors


Corrosion Inhibitor Type
Anodic

Mainly Cathodic

General

Molybdate

Bicarbonate

Soluble oils

Nitrite

Polyphosphate

Other organics, such as azole or carboxylate

Orthophosphate

Phosphonate

Silicate

Zinc
Polysilicate

Cathodic corrosion inhibitors form a protective film on the cathode. These


inhibitors reduce the corrosion rate in direct proportion to the reduction of the
cathodic area.
General corrosion inhibitors protect by forming a film on all metal surfaces
whether anodic or cathodic.
Table 8.1 lists typical corrosion inhibitors of these three types.
The most important factor in an effective corrosion inhibition program is the consistent control of both the corrosion inhibition chemicals and the key water characteristics.
No program will work without controlling these factors.
Cathodic Protection
There are two types of cathodic protection: sacrificial anode and impressed current.
Sacrificial anodes reduce galvanic attack by providing a metal (usually zinc, but sometimes magnesium) that is higher on the galvanic series (see ASHRAE [2011]) than either
of the two metals that are coupled together. The sacrificial anode thereby becomes anodic
to both metals and supplies electrons to these cathodic surfaces. Proper design and placement of these anodes is important and usually requires the services of a National Association of Corrosion Engineers (NACE) registered engineer. When properly used, these
anodes can reduce loss of steel from the tube sheet of exchangers using copper tubes. Sacrificial anodes have helped supplement chemical programs in many cooling water and
process water systems. Sacrificial anodes may also be used on the buried steel conduits of
heat distribution systems; see Chapter 4 for additional discussion.
Impressed current cathodic protection is a similar corrosion control technique that
reverses the corrosion cells normal current flow by impressing a stronger current of
opposite polarity. Direct current is applied to an anodeinert (platinum, graphite) or
expendable (aluminum, cast iron)reversing the galvanic flow and converting the steel
from a corroding anode to a protective cathode. The method is very effective in protecting
essential equipment such as elevated water storage tanks, steel tanks, and softeners.
Impressed current may also be used on the buried steel conduits of heat distribution systems; see Chapter 4 for additional discussion.

White Rust on Galvanized Steel Cooling Towers


White rust is a zinc corrosion product that forms on galvanized surfaces. It appears as
a white waxy or fluffy deposit composed of loosely adhering zinc carbonate. The loose
crystal structure allows continued access of the corrosive water to exposed zinc. Unusually rapid corrosion of galvanized steel, as evidenced by white rust, can affect galvanized
steel cooling towers under certain conditions. Before chromates in cooling tower water
were banned, the common treatment system consisted of chromates for corrosion control
and sulfuric acid for scale control. This control method generally has been replaced by
alkaline treatment involving scale inhibitors at a higher pH. Alkaline water chemistry is

8.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

naturally less corrosive to steel and copper but creates an environment where white rust
on galvanized steel can occur. Also, some scale prevention programs soften the water to
reduce hardness, rather than use acid to reduce alkalinity; the resulting soft water is corrosive to galvanized steel.
Prevention
White rust can be prevented by promoting the formation of a nonporous surface layer
of basic zinc carbonate. This barrier layer is formed during a process called passivation
and normally protects the galvanized steel for many years. Passivation is best accomplished by controlling pH during initial operation of the cooling tower. Control of the
cooling water pH in the range of 7 to 8 for 45 to 60 days usually allows passivation of a
galvanized surface to occur. In addition to pH control, operation with moderate hardness
levels of 100 to 300 ppm as calcium carbonate and alkalinity levels of 100 to 300 ppm as
calcium carbonate will promote passivation. Where pH control is not possible, certain
phosphate-based inhibitors may help protect galvanized steel. A water treatment specialist should be consulted for specific formulations.

SCALE CONTROL
Scale is a dense coating of predominantly inorganic material formed from the precipitation of water-soluble constituents. Some common scales are:
Calcium carbonate
Calcium phosphate
Magnesium salts
Silica
Iron oxides
The following principal factors determine whether or not water is scale forming:
Temperature
Alkalinity or acidity (pH)
Amount of scale-forming material present
Influence of other dissolved materials, which may or may not be scale forming
As any of these factors changes, scaling tendencies also change. Most salts become
more soluble as temperature increases. However, some salts, such as calcium carbonate,
become less soluble as temperature increases. Therefore, they often cause deposits at
higher temperatures.
A change in pH or alkalinity can greatly affect scale formation. For example, as pH or
alkalinity increases, calcium carbonate, the most common scale constituent in cooling
systems, decreases in solubility and deposits on surfaces. Some materials, such as silica,
are less soluble at lower alkalinities. When the amount of scale-forming material dissolved in water exceeds its saturation point, scale may result. In addition, other dissolved
solids may influence scale-forming tendencies. In general, a higher level of scale-forming
dissolved solids results in a greater chance for scale formation. Indices such as the Langelier Saturation Index (Langelier 1936) and the Ryznar Stability Index (Ryznar 1944) can
be useful tools to predict the calcium carbonate scaling tendency of water. These indices
are calculated using the pH, alkalinity, calcium hardness, temperature, and total dissolved
solids of the water and indicate whether the water will favor precipitating or dissolving of
calcium carbonate.
Methods used to control scale formation include the following:
Limit the concentration of scale-forming minerals by controlling cycles of concentration or by removing the minerals before they enter the system (see the

8.7

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

External Treatments subsection). A cycle of concentration is the ratio of makeup


rate to the sum of blowdown and drift rates. The cycles of concentration can be
monitored by calculating the ratio of chloride ion, which is highly soluble, in the
system water to that in the makeup water.
Make mechanical changes in the system to reduce the chances for scale formation. Increased water flow and heat exchangers with larger surface areas are
examples of such changes.
Feed acid to keep the common scale-forming minerals (e.g., calcium carbonate)
dissolved.
Treat with chemicals designed to prevent scale. There are two classes of chemical scale inhibitors: threshold inhibition chemicals and scale conditioners. These
work by the following mechanisms.
Threshold inhibition chemicals prevent scale formation by keeping the
scale-forming minerals in solution and not allowing a deposit to form.
Threshold inhibitors include organic phosphates, polyphosphates, and
polymeric compounds.
Scale conditioners modify the crystal structure of scale, creating a bulky,
transportable sludge instead of a hard deposit. Scale conditioners include
lignins, tannins, and polymeric compounds.

Nonchemical Methods
Equipment based on magnetic, electromagnetic, or electrostatic technology has been
used for scale control in boiler water, cooling water, and other process applications. Magnetic systems are designed to cause scale-forming minerals to precipitate in a low-temperature area away from heat exchanger surfaces, thus producing nonadherent particles (e.g.,
the aragonite form of calcium carbonate versus the hard, adherent calcite form). The precipitated particles can then be removed by blowdown, mechanical means, or physical
flushing. The objective of electrostatics is to prevent scale-forming reactions by imposing
a surface charge on dissolved ions that causes them to repel.
Results of side-by-side comparative tests with conventional water treatment have
been mixed. A Federal Technology Alert report regarding these technologies (DOE 1998)
stresses that the success of the application depends largely on the experience of the
installer. The report includes a discussion of the potential benefits achieved and the necessary precautions to consider when applying these systems. The Final Report for
ASHRAE Research Project RP-1155 (Cho 2002) studied physical water treatment. For
this study, a physical water treatment device was defined as a nonchemical method of
water treatment for scale prevention or mitigation. Bulk precipitation was proposed as the
mechanism of scale prevention. Three different devices described as permanent magnets,
a solenoid coil device, and a high-voltage electrode were tested under laboratory conditions. Fouling resistance data obtained in a heat transfer test section supported the benefit
of all three devices when configured in optimum conditions.

External Treatments
Minerals may also be removed by various external pretreatment methods such as
reverse osmosis and ion exchange. Zeolite softening, demineralization, and dealkalization
are examples of ion exchange processes.

BIOLOGICAL GROWTH CONTROL


Biological growth (algae, bacteria, and fungi) can interfere with a cooling operation
due to fouling or corrosion and may present a health hazard if present in aerosols pro-

8.8

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

duced by the equipment. Heating equipment operates above normal biological limits and
therefore has fewer microbial problems. When considering biological growth in a cooling
system, it is important to distinguish between free-living planktonic organisms and sessile
(attached) organisms. Sessile organisms cause the majority of the problems, though they
may have entered and multiplied as planktonic organisms.
Biological fouling can be caused by a wide variety of organisms that produce biofilm
and slime masses. Slimes can be formed by bacteria, algae, yeasts, or molds and frequently consist of a mixture of these organisms combined with organic and inorganic
debris. Organisms such as barnacles and mussels may cause fouling when river, estuarine,
or seawater is used. Biological fouling can significantly reduce the efficiency of cooling
by reducing heat transfer, increasing back pressure on recirculation pumps, disrupting
flow patterns over cooling media, plugging heat exchangers, and blocking distribution
systems. In extreme cases the additional mass of slime has caused the cooling media to
collapse.
Microorganisms can dramatically enhance, accelerate or, in some cases, initiate localized corrosion (pitting). Microorganisms can influence localized corrosion directly by
their metabolism or indirectly by the deposits they form. Indirect influence may not be
mediated by simply killing the microorganisms; deposit removal is usually necessary.
Direct influence can be substantially mediated by inhibiting microorganism metabolism.
Algae use energy from the sun to convert bicarbonate or CO2 into biomass. Masses of
algae can block piping, distribution holes, and nozzles. A distribution deck cover, which
drastically reduces the sunlight reaching the algae, is one of the most cost-effective control devices for a cooling tower. Biocides are also used to assist in the control of algae.
Algae can also provide nutrients for other microorganisms in the cooling system,
increasing the biomass in the water. Bacteria can grow in systems even when nutrient levels are relatively low. Yeasts and fungi grow much more slowly than bacteria and find it
difficult to compete in bulk waters for the available food. Fungi thrive in partially wetted
and high-humidity areas such as the cooling media. Wood-destroying species of fungi can
be a major concern for wooden cooling towers, as the fungi consume the cellulose and/or
lignin in the wood, reducing its structural integrity.
Most waters contain organisms capable of producing biological slime, but optimal
conditions for growth are poorly understood. Equipment near nutrient sources or that has
process leaks acting as a food source is particularly susceptible to slime formation. Even a
thin layer of biofilm significantly reduces heat transfer rates in heat exchangers.

Control Measures
Eliminating sunlight from wetted surfaces such as distribution troughs, cooling
media, and sumps significantly reduces algae growth. Eliminating dead legs and low-flow
areas in the piping and the cooling loop reduces biological growth in those areas. Careful
selection of materials of construction can remove nutrient sources and environmental
niches for growth. Maintaining a high-quality makeup water supply with low bacteria
counts also helps minimize biological growth. Equipment should also be designed with
adequate access for inspection, sampling, and manual cleaning.
Sometimes the effective control of slime and algae requires a combination of
mechanical and chemical treatments. For example, when a system already contains a considerable accumulation of slime, a preliminary mechanical cleaning makes the subsequent application of a biocidal chemical more effective in killing the growth and more
effective in preventing further growth. A buildup of scale deposits, corrosion product, and
sediment in a cooling system also reduces the effectiveness of chemical biocides. Routine

8.9

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

manual cleaning of cooling towers, including the use of high-level chlorination and a biodispersant (surfactant), helps control Legionella bacteria as well as other microorganisms.
Microbiocides
Chemical biocides used to control biological growth in cooling systems fall into two
broad categories: oxidizing and nonoxidizing biocides.
Oxidizing Biocides
Oxidizing biocides (chlorine, chlorine-yielding compounds, bromine, bromochlorodimethylhydantoin [BCDMH or BCD], ozone, iodine, and chlorine dioxide) are among
the most effective microbiocidal chemicals. However, they are not always appropriate for
control in cooling systems with a high organic loading. In wooden cooling towers, excessive concentrations of oxidizing biocides can cause delignification, and overdosing of
oxidizing biocides may cause corrosion of metallic components. In systems large enough
to justify the cost of equipment to control feeding of oxidizing biocides accurately, the
application may be safe and economical. The most effective use of oxidizing biocides is
to maintain a constant low-level residual in the system. However, if halogen-based oxidizing biocides are fed intermittently (slug dosed), a pH near 7 is advantageous because, at
this neutral pH, halogens are present as the hypohalous acid form (HOR, where R represents the halogen) over the hypohalous ion form (OR). The effectiveness of this shock
feeding is enhanced due to the faster killing action of hypohalous acid over that of hypohalous ion. The residual biocide concentration should be tested, using a field test kit, on a
routine basis. Most halogenation programs can benefit from the use of dispersants or surfactants (chlorine helpers) to break up microbiological masses.
Chlorine has been the oxidizing biocide of choice for many years, either as chlorine
gas or in the liquid form as sodium hypochlorite. Other forms of chlorine, such as powders or pellets, are also available. Use of chlorine gas is declining due to the health and
safety concerns involved in handling this material and in part due to environmental pressures concerning the formation of chloramines and trihalomethane.
Bromine is produced either by the reaction of sodium hypochlorite with sodium bromide on site or by release from pellets. Bromine has certain advantages over chlorine: it is
less volatile, and bromamines break down more rapidly than chloramines in the environment. Also, when slug feeding biocide in high pH systems, hypobromous acid may have
an advantage because its dissociation constant is lower than that of chlorine. This effect is
less important when biocides are fed continuously.
Ozone has several advantages compared to chlorine: it does not produce chloramines
or trihalomethane, it breaks down into nontoxic compounds rapidly in the environment, it
controls biofilm better, and it requires significantly less chemical handling. The use of
ozone-generating equipment in an enclosed space, however, requires care be taken to protect operators from the toxic gas. Also, research has shown that ozone is only marginally
effective as a scale and corrosion inhibitor (Gan et al. 1996; Nasrazadani and Chao 1996).
Water conditions should be reviewed to determine the need for scale and corrosion
inhibitors and then, as with all oxidizing biocides, inhibitor chemicals should be carefully
selected to ensure compatibility. To maximize the biocidal performance of ozone, the
injection equipment should be designed to provide adequate contact of the ozone with the
circulating water. In larger systems, care should be taken to ensure that the ozone is not
depleted before the water has circulated through the entire system.
Iodine is provided in pelletized form, often from a rechargeable cartridge. Iodine is a
relatively expensive chemical for use in cooling towers and is probably only suitable for
use in smaller systems.

8.10

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

Nonoxidizing Biocides
When selecting a nonoxidizing microbiocide, the pH of the circulating water and the
chemical compatibility with the corrosion and/or scale inhibitor product must be considered. The following list, while not exhaustive, identifies some of these nonoxidizing biocides.
Quaternary ammonium compounds
Methylene bis(thiocyanate) (MBT)
Isothiazolones
Thiadiazine thione
Dithiocarbamates
Decyl thioethanamine (DTEA)
Glutaraldehyde
Dodecylguanidine
Benzotriazole
Tetrakis(hydroxymethyl)phosphonium sulfate (THPS)
Dibromo-nitrilopropionamide (DBNPA)
Bromo-nitropropane-diol
Bromo-nitrostyrene (BNS)
Proprietary blends
The manner in which nonoxidizing biocides are fed is important. Sometimes the continuous feeding of low dosages is neither effective nor economical. Slug feeding large
concentrations to achieve a toxic level of the chemical in the water for a sufficient time to
kill the organisms can show better results. The water blowdown rate and biocide hydrolysis (chemical degradation) rate affect the required dosage. The hydrolysis rate of the biocide is affected by the type of biocide, along with the temperature and pH of the system
water. Dosage rates are proportional to system volume; dosage concentrations should be
sufficient to ensure that the contact time of the biocide is long enough to obtain a high kill
rate of microorganisms before the minimum inhibitory concentration of the biocide is
reached. The period between nonoxidizing biocide additions should be based on the system half life, with sequential additions timed to prevent regrowth of bacteria in the water.
Handling Microbiocides
All microbiocides must be handled with care to ensure personal safety. In the United
States, cooling water microbiocides are approved and regulated through the U.S. Environmental Protection Agency (EPA) and, by law, must be handled in accordance with labeled
instructions. Maintenance staff handling the biocides should read the material safety data
sheets and be provided with all the appropriate safety equipment to handle the substances.
Automatic feed systems should be used to minimize or eliminate the handling of biocides
by maintenance personnel.
Other Biocides
Ultraviolet irradiation deactivates microorganisms as the water passes through a
quartz tube. The intensity of the light and thorough contact with the water are critical in
obtaining a satisfactory kill of microorganisms. Suspended solids in the water or deposits
on the quartz tube significantly reduce the effectiveness of this treatment method. Therefore, a filter is often installed upstream of the lamp to minimize these problems. Because
ultraviolet light leaves no residual material in the water, sessile organisms and organisms
that do not pass the light source are not affected by the ultraviolet treatment. Ultraviolet
irradiation may be effective on humidifiers and air washers where the application of biocidal chemicals is unacceptable and where 100% of the recirculating water passes the

8.11

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

lamp. Ultraviolet irradiation is less effective where all the microorganisms cannot be
exposed to the treatment, such as in cooling towers. Ultraviolet lamps require replacement after approximately every 8000 hours of operation.
Metallic ions, namely copper and silver, effectively control microbial populations
under very specific circumstances. Either singularly or in combination, copper and silver
ions are released into the water via electrochemical means to generate 1 to 2 ppm of copper and/or 0.5 to 1.0 ppm of silver. The ions assist in the control of bacterial populations
in the presence of a free chlorine residual of at least 0.2 ppm. Copper, in particular, effectively controls algae.
Liu et al. (1994) reported control of Legionella pneumophila bacteria in a hospital
hot-water supply using copper-silver ionization. In this case, Legionella colonization
decreased significantly when copper and silver concentrations exceeded 0.4 and
0.04 ppm, respectively. Also, residual disinfection prevented Legionella colonization for
two months after the copper-silver unit was inactivated.
Significant limitations exist in the use of copper and silver ion for cooling systems.
Many states are restricting the discharge of these ions to surface waters, and if the pH of
the system water rises above 7.8, the efficacy of the treatment is significantly reduced.
Systems that have steel or aluminum heat exchangers should not be treated by this
method, as the potential for the deposition of the copper ion and subsequent galvanic corrosion is significant.

Legionnaires Disease
Like other living things, Legionella pneumophila, the bacterium that causes Legionnaires disease (legionellosis), requires moisture for survival. Legionella bacteria are
widely distributed in natural water systems and are present in many drinking-water supplies. The Legionella bacterium is a very small bar-shaped cell measuring 0.000079 by
0.000012 in. (2 by 0.3 m). Systems with potable water between 80F and 120F (26.7C
and 48.9C), cooling towers, certain types of humidifiers, evaporative condensers, whirlpools and spas, and the various components of air conditioners are considered to be
amplifiers of the number of the bacteria in the water. The bacteria are killed in a matter of
minutes when exposed to temperatures above 140F (60C). Due to the small cell size of
the bacteria, legionellosis can be acquired by inhalation of Legionella organisms in aerosols. Aerosols can be produced by cooling towers, evaporative condensers, decorative
fountains, showers, and misters. It has been reported that the aerosol from cooling towers
can be transmitted over a distance of up to 2 mi (3.2 km). If air inlet ducts of nearby air
conditioners draw the aerosol from contaminated cooling towers into a building, the air
distribution system of that building can transmit the disease. When an outbreak of
Legionnaires disease occurs, cooling towers are often the suspected source. However,
other water systems may produce an aerosol and should not be neglected. Amplification
of Legionella within protozoans has been demonstrated, and Legionella bacteria are
thought to be protected from biocides while growing intracellularly. Amplification of
Legionella bacteria in biofilm and slime masses has also been shown by a number of
researchers. Microbial control programs should consider the effectiveness of these products against slimes as part of a Legionella control program.
Prevention and Control
The Legionella count required to cause illness has not been firmly established
because many factors are involved, including (1) the virulence and number of Legionella
in the air, (2) the rate at which the aerosol dries, (3) the wind direction, and (4) susceptibility to the disease of the person breathing the air. The organism is often found at sites

8.12

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

not associated with an outbreak of the disease. It has been shown that it is feasible to
operate cooling systems with Legionella bacteria below the limit of detection and that the
only method to prove that a system is operating at these levels is to specifically test for
Legionella bacteria rather than to infer from total bacteria count measurements.
Periodic monitoring of circulating water for total bacteria count and Legionella count
can be accomplished using culture methods. Monitoring system cleanliness and using a
microbial control agent that has proven efficacy or is generally regarded as effective in
controlling Legionella populations are also important. Other measures to decrease risk
include optimizing cooling tower design to minimize drift, eliminating dead legs or lowflow areas, selecting materials that do not promote the growth of Legionella, and locating
the tower so that drift is not injected into air handlers. ASHRAEs Position Document on
Legionellosis (2012) has further information on this topic.

SUSPENDED SOLIDS AND DEPOSITATION CONTROL


Where a cooling tower is used in a district heating/cooling system, it acts as a great
air filter. Any airborne debris that is drawn into the tower will make its way to the basin or
sump and accumulate. Since most particles in condenser water from cooling towers are
smaller than 0.0004 in. (10 m) and organic, they are great food sources for microorganisms and bacteria and could also lead to scaling of heat transfer surfaces. Hence, condenser water filtration is extremely important, and effective filtration down to 0.00001 in.
(0.25 m) will assist in filtering out Legionella cells. There are several types of filtration
equipment available for use; these are described in the section that follows.

Mechanical Filtration
Strainers, filters, and separators may be used to reduce suspended solids to an acceptably low level. Generally, if the screen is 200 mesh, equivalent to about 0.003 in.
(0.076 mm), it is called a strainer; if it is finer than 200 mesh, it is called a filter. A summary of the particle size ranges for the various filter technologies is provided in Table 8.2.
In specifying filtration systems, third-party testing by a qualified university or private
test agency should be requested. The test report documentation should include a description of methods, piping diagrams, performance data, and certification.
Strainers
A strainer is a closed vessel with a cleanable screen designed to remove and retain foreign particles down to 0.001 in. (0.025 mm) diameter from various flowing fluids. Strainers extract material that is not wanted in the fluid and allow saving the extracted product if
it is valuable. Strainers are available as single-basket or duplex units and manual or automatic cleaning units and may be made of cast iron, bronze, stainless steel, copper-nickel
alloys, or plastic. Magnetic inserts are available where microscopic iron or steel particles
are present in the fluid.
Table 8.2 Summary of Filter Technology Particle Size Ranges
Filter Technology
(Listed in Order of Effectiveness)

Typical Range of Smallest Particle Filtration Level,


microns

Cartridge filter

0.01 to +100

Sand filter (including hybrid design)

0.25 to 40

Bag filter

1.0 to +100

Centrifugal separators

5 to 75

Disc filter

15 to 25

Automatic self-cleaning strainer

15 to 50

8.13

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Cartridge Filters
Cartridge filters are typically used as final filters to remove nearly all suspended particles from about 0.004 in. (0.1 mm) down to 0.00004 in. (0.001 mm) or less. Cartridge
filters are typically disposable (i.e., once plugged, they must be replaced). The frequency
of replacement, and thus the economical feasibility of their use, depends on the concentration of suspended solids in the fluid, the size of the smallest particles to be removed,
and the removal efficiency of the cartridge filter selected.
In general, cartridge filters are favored in systems where contamination levels are less
than 0.01% by mass (<100 ppm). They are available in many different materials of construction and configurations. Filter media materials include yams, felts, papers, nonwoven materials, resin-bonded fabric, woven wire cloths, sintered metal, and ceramic
structures. The standard configuration is a cylinder with an overall length of approximately l0 in. (250 mm), an outside diameter of approximately 2.5 to 2.75 in. (6.25 to
70 mm), and an inside diameter of about 1 to 1.5 in. (25 to 38 mm), where the filtered
fluid collects in the perforated internal core. Overall lengths from 4 to 40 in. (100 to
1000 mm) are readily available.
Cartridges made of yams, resin-bonded fibers, or melt-blown fibers normally have a
structure that increases in density toward the center. These depth-type filters capture particles throughout the total media thickness. Thin media, such as pleated paper (membrane
types), have a narrow pore size distribution design to capture particles at or near the surface of the filter. Surface-type filters can normally handle higher flow rates and provide
higher removal efficiency than equivalent depth filters.
Cartridge filters are rated according to manufacturers guidelines. Surface-type filters
have an absolute rating, while depth-type filters have a nominal rating that reflects their
general classification function. Higher-efficiency melt-blown depth filters are available
with absolute ratings as needed.
Sand Filters
A downflow filter is used to remove suspended solids from a water stream. The
degree of suspended-solids removal depends on the combinations and grades of the
medium being used in the vessel. During the filtration mode, water enters the top of the
filter vessel. After passing through a flow impingement plate, it enters the quiescent
(calm) freeboard area above the medium.
In multimedia downflow vessels, various grain sizes and types of media are used to
filter the water. This design increases the suspended-solids holding capacity of the system, which in turn increases the backwashing interval. Multimedia vessels might also be
used for low-suspended-solids applications, where chemical additives are required. In the
multimedia vessel, the fluid enters the top layer of anthracite media, which has an effective size of 0.04 in. (1 mm). This relatively coarse layer removes the larger suspended particles, a substantial portion of the smaller particles, and small quantities of free oil. Flow
continues down through the next layer of fine garnet material, which has an effective size
of 0.012 in. (0.3 mm). A more finely divided range of suspended solids is removed in this
polishing layer. The fluid continues into the final layer, a coarse garnet material that has
an effective size of 0.08 in. (2.03 mm). Contained in this layer is the header/lateral assembly that collects the filtered water.
When the vessel has retained enough suspended solids to develop a substantial pressure drop, the unit must be backwashed either manually or automatically by reversing the
direction of flow. This operation removes the accumulated solids out through the top of
the vessel.

8.14

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

Bag-Type Filters
Bag-type filters are composed of a bag of mesh or felt (through which the filtered
media must pass) supported by a removable perforated metal basket, placed in a closed
housing with an inlet and outlet. The housing is a welded, tubular pressure vessel with a
hinged cover on top for access to the bag and basket. Housings are made of carbon steel
or stainless steel. The inlet can be in the cover, in the side (above the bag), or in the bottom (and internally piped to the bag). The side inlet is the simplest type. In any case, the
liquid enters the top of the bag. The outlet is located at the bottom of the side (below the
bag). Pipe connections can be threaded or flanged. Single-basket housings can handle up
to 220 gpm (13.9 L/s), multibaskets up to 3500 gpm (221 L/s).
The support basket is usually constructed of 304 stainless steel perforated with 1/8 in.
(3.2 mm) holes. (Heavy wire mesh baskets also exist.) The baskets can be lined with fine
wire mesh and used by themselves as strainers, without adding a filter bag. Some manufacturers offer a second, inner basket (and bag) that fits inside the primary basket. This
provides for two-stage filtering: first a coarse filtering stage, then a finer one. The benefits
are longer service time and possible elimination of a second housing to accomplish the
same function.
The filter bags are made of many materials (cotton, nylon, polypropylene, and polyester) with a range of ratings from 0.00004 to 0.033 in. (0.001 to 0.84 mm). Most common are felted materials because of their depth-filtering quality, which provides high dirtloading capability, and their fine pores. Mesh bags are generally coarser but are reusable
and, therefore, less costly. The bags have a metal ring sewn into their opening; this holds
the bag open and seats it on top of the basket rim.
In operation, the liquid enters the bag from above, flows out through the basket, and
exits the housing cleaned of particulates of a desired size. The contaminant is trapped
inside the bag, making it easy to remove without spilling any downstream.
Centrifugal-Gravity Separators
In this type of separator, liquids/ solids enter the unit tangentially, which sets up a circular flow. Liquids/solids are drawn through tangential slots and accelerated into the separation chamber. Centrifugal action tosses the particles heavier than the liquid to the
perimeter of the separation chamber. Solids gently drop along the perimeter and into the
separators quiescent collection chamber. Solids-free liquid is drawn into the separators
vortex (low-pressure area) and up through the separators outlet. Solids are either purged
periodically or continuously bled from the separator by either a manual or automatic
valve system.
Special Methods
Localized areas frequently can be protected by special methods. Thus, pump-packing
glands or mechanical shaft seals can be protected by freshwater makeup or by circulating
water from the pump casing through a cyclone separator or filter into the lubricating
chamber.
In smaller equipment, a good dirt-control measure is to install backflush connections
and shutoff valves on all condensers and heat exchangers so that accumulated settled dirt
can be removed by backflushing with makeup water or detergent solutions. These connections can also be used for acid cleaning to remove calcium carbonate scale.

SELECTION OF WATER TREATMENT METHOD


As discussed in the previous sections, many methods are available to prevent or correct corrosion, scaling, and biofouling. The selection of the proper water treatment
method, and the chemicals and equipment necessary to apply that method, depends on

8.15

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

many factors. The chemical characteristics of the water, which change with the operation
of the equipment, are important. Other factors contributing to the selection of proper
water treatment are:
Economics
Chemistry control mechanisms
Dynamics of the operating system
Design of major components (e.g., the cooling tower or boiler)
Number of operators available
Training and qualifications of personnel
Preventive maintenance program
The following subsections offer general water treatment guidance for the types of
systems found in district heating systems.

Once-Through Systems (Seawater or Surface-Water Cooling)


In many jurisdictions, use of public water bodies for once-through cooling systems is
no longer permitted or it is very difficult to obtain required permits. Where permitted,
economics is an overriding concern in treating water for once-through systems (in which
a very large volume of water passes through the system only once). Protection can be
obtained with relatively little treatment per unit mass of water because the water does not
change significantly in composition while passing through the equipment. However, the
quantity of water to be treated is usually so large that any treatment other than simple filtration or the addition of a few parts per million of a polyphosphate, silicate, or other
inexpensive chemical may not be practical or affordable. Intermittent treatment with
polyelectrolytes can help maintain clean conditions when the cooling water is sedimentladen. In such systems, it is generally less expensive to invest more in corrosion-resistant
construction materials than to attempt to treat the water.

Open Recirculating Systems (Cooling Towers)


In an open recirculating system with chemical treatment, more chemical must be
present because the water composition changes significantly by evaporation. Corrosive
and scaling constituents are concentrated. However, treatment chemicals also concentrate
by evaporation; therefore, after the initial dosage, only moderate dosages maintain the
higher level of treatment needed. The selection of a water treatment program for an open
recirculating system depends on the following major factors:
Economics
Water quality
Performance criteria (e.g., corrosion rate, bacteria count, etc.)
System metallurgy
Available staffing
Automation capabilities
Environmental requirements
Water treatment supplier (some technologies are superior to others in terms of
economics, ease of use, safety, and impact on the environment)
An open recirculating system is typically treated with a scale inhibitor, a corrosion
inhibitor, an oxidizing biocide, a nonoxidizing biocide, and possibly a dispersant. The
exact treatment program depends on the previously stated conditions.
A water treatment control scheme for a cooling tower might include:
Chemistry and cycles of concentration control using a conductivity controller
Alkalinity control using automatic injection of sulfuric acid based on pH

8.16

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

Scale control using contacting water meters, proportional feed, or traced control
technology
Oxidizing biocide control using an oxidation-reduction potential (ORP) controller
Nonoxidizing biocide control using timers and pump systems
For cooling tower condenser water systems that will be shut down and then restarted,
ASHRAE HandbookHVAC Applications (2011) provides recommended procedures.

Closed Recirculating Systems (District Heating Distribution Systems)


In a closed recirculating system, water composition remains fairly constant with very
little loss of either water or treatment chemical. Closed systems are often defined as those
requiring less than 5% makeup per year. The need for water treatment in such systems is
often ignored based on the rationalization that the total amount of scale from the water
initially filling the system would be insufficient to interfere significantly with heat transfer, and that corrosion would not be serious. However, in systems where leakage losses
are common, corrosion products can accumulate sufficiently to foul heat transfer surfaces
or deteriorate the interior of the piping system. Therefore, all systems should be adequately treated to control corrosion. Systems with high makeup rates should be treated to
control scale as well.
The selection of a treatment program for closed systems should consider the following factors:
Economics
System metallurgy
Operating conditions
Makeup rate
System size
Possible treatment technologies include the following:
Buffered nitrite
Molybdate
Silicates
Polyphosphates
Oxygen scavengers
Organic blends
Before new systems are treated, they must be cleaned and flushed. Grease, oil, construction dust, dirt, and mill scale are always present in varying degrees and must be
removed from the metallic surfaces to ensure adequate heat transfer and to reduce the
opportunity for localized corrosion. Detergent cleaners with organic dispersants are available for proper cleaning and preparation of new closed systems.

Medium- and High-Temperature Hot-Water Systems


Medium-temperature hot-water district heating systems (250F to 350F [120C to
175C) and high-temperature hot-water district heating systems (above 350F [175C])
require careful consideration of treatment for corrosion and deposit control. Makeup
water for such systems should be demineralized or softened to prevent scale deposits. For
corrosion control, deaerators and oxygen scavengers such as sodium sulfite can be added
to remove dissolved oxygen.

8.17

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

European Practice in Closed Distribution Systems


In the practice of low-temperature hot-water district heating, Europeans have established practices for water treatment in the distribution systems. These distribution systems
are of high integrity and have very low rates of leakage/makeup. The European practice,
in particular that of the Nordic counties, relies less on corrosion inhibitors than do the
North American practices. Makeup water is filtered, demineralized or softened, and
deaerated. Sodium hydroxide is then added to raised the pH to 9.510. Subsequently, the
corrosion rate and concentrations are closely monitored, normally by the operating staff
rather than by the supplier of the corrosion inhibitor chemicals as is the normal practice in
North America. An International Energy Agency (IEA) report (1996) compares the North
American approach, which relies more heavily on corrosion inhibitors, to the European
approach in a case study that concludes that while both provide adequate corrosion protection, the European approach will likely be less costly.

Steam Systems
Many treatment methods are available for steam systems and their boilers; the
method selected depends on the following:
Makeup water quality
Makeup water quantity (or percentage condensate return)
Pretreatment equipment
Boiler operating conditions
Steam purity requirements
Economics
Pretreatment for makeup water could consist of the following elements:
Water softeners (for removal of calcium and magnesium hardness)
Deaerators (for removal of dissolved gases, especially oxygen and CO2)
Dealkalizers (to remove alkalinity in systems with high makeup rates and high
alkalinity makeup water)
Demineralizers (to remove almost all the hardness, alkalinity, and solids,
depending on the application)
Reverse osmosis (which can remove up to 99% of minerals and dissolved solids
from the water leaving almost pure H2O)
Contaminants that are not removed mechanically by pretreatment equipment must be
treated chemically. Once in the feedwater, dissolved gases, hardness, and dissolved minerals must be treated to prevent deposition and corrosion. ASME and American Boiler
Manufacturers Association (ABMA) have published recommended water chemistry limits for boiler water (ASME 1994; ABMA 2005). Treatment of the boiler water is often
determined by the end use of the steam. Treatment regimens may include reduction of
alkalinity; hardness removal; silica removal; oxygen reduction; and the feed of scale and
corrosion inhibitors, oxygen scavengers, condensate treatment chemicals, and antifoams.
These regimens are described in the following subsections.
Prevention of Scale
After the feedwater is pretreated, scale is controlled with phosphates, acrylates, polymers, chelates, and/or coagulation programs. Chelates, polymers, and acrylates work by
binding the hardness, thereby preventing precipitation and scale formation. Phosphates
and coagulation programs work in combination with sludge conditioners (tannins, lignins, starches, and synthetic polymers) to produce a softened precipitate that is removed
by blowdown of the boiler.

8.18

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

Prevention of Corrosion and Oxygen Pitting


While boilers can corrode as the result of low boiler water pH or misuse of certain
chemicals, corrosion is primarily caused by oxygen. After mechanical deaeration, boiler
feedwater must be treated chemically. Effective deaeration removes most of the dissolved
oxygen (most properly operating deaerators can remove oxygen down to 0.007 ppm).
Oxygen scavengers such as catalyzed sodium sulfite and proprietary organic oxygen
scavengers should then be fed to react with the residual oxygen in the feedwater after
deaeration. Oxygen scavengers will provide added protection not only to the boiler but to
the steam and condensate system as well. Oxygen at levels as low as 0.005 ppm can cause
oxygen pitting in the steam and condensate system if not chemically reduced by oxygen
scavengers.
Most of the corrosion damage to boilers and associated equipment occurs during idle
periods. The corrosion is caused by the exposure of wet metal to oxygen in the air or
water. For this reason, special precautions must be taken to prevent corrosion while boilers are out of service, via either wet boiler lay-up or dry boiler lay-up:
Wet boiler lay-up is a method of storing boilers full of water so that they can be
returned to service. It involves adding extra chemicals (usually something to
increase alkalinity, an oxygen scavenger, and a dispersant) to the boiler water.
The water level is raised in the idle boiler to eliminate air spaces, and the boiler
is kept completely full of treated water. Superheaters require special protection.
Nitrogen gas can also be used on airtight boilers to maintain a positive pressure
on the boiler, thereby preventing oxygen in-leakage.
Dry boiler lay-up is usually used for longer boiler outages. It involves draining,
cleaning, and drying the boiler. A material that absorbs moisture, such as
hydrated lime or silica gel, is placed in trays inside the boiler. The boiler is then
sealed carefully to keep out air. Periodic inspection and replacement of the drying chemical are required during long storage periods.

Steam Distribution and Condensate Collection Systems


Two problems associated with steam and condensate systems include general corrosion and pitting corrosion. In order to prevent condensate corrosion, the systems must be
protected from acidic conditions, which lead to general corrosion, and oxygen, which
leads to pitting corrosion. Protection can be by mechanical or chemical means or a combination of both. The following subsections describe methods commonly used for condensate system protection.
Protection from General Corrosion
Mechanical Protection
Reduce alkalinity from boiler feedwater to minimize the amount of CO2 in the system. CO2 reacts with water (condensate) and forms carbonic acid that corrodes condensate system metal. Alkalinity can be reduced by dealkalization, demineralization, and
reverse osmosis. Note that in many cases mechanical reduction of alkalinity is not needed
due to low alkalinity makeup water and/or feedwater.
Chemical Protection
Use volatile amines, such as morpholine, diethylaminoethanol (DEAE), or cyclohexylamine, to neutralize carbonic acid and keep the condensate pH between 8.0 and 9.0.

8.19

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Protection from Pitting Corrosion


Mechanical Protection
Reduce oxygen from all boiler feedwater to prevent oxygen carryover to the steam
and condensate system (via mechanical methods and chemical oxygen scavengers).
Chemical Protection
Either (1) feed filming amines, such as octadecylamine, to the steam to form a thin,
hydrophobic film on the condensate system surfaces or (2) feed a volatile oxygen scavenger to the steam to scavenge the oxygen.
The need for chemical treatment can be reduced by designing and maintaining tight
return systems so that the condensate is returned to the boiler and less makeup is required
in the boiler feedwater. The greater the amount of makeup, the more the system requires
increased chemical treatment.

MAINTENANCE
A smart maintenance program is of prime importance to reliability and optimum performance of a district heating system. Unplanned downtime is one of the primary enemies
of the reliable supply of thermal energy. District heating maintenance requirements will
vary widely from one district heating system to another. Aside from the type of equipment used (fire-tube vs. water-tube boilers, water spread limiting vs. drainable, dryable,
testable piping, etc.), other factors such as the operating environment; the operating temperature; and the age of the plant, distribution system, and terminal equipment will have
major impacts on the required maintenance. For major pieces of equipment such as boilers, chillers, and pumps, some guidance will be available from the equipment manufacturer. For field-erected portions of the central plant, the distribution system, and the
consumer interconnect, it is usually necessary to develop a maintenance program based
on the methods of construction, the components used, and the age of the system. For
items such as filters, maintenance requirements will vary widely depending on site-specific conditions and may also vary significantly over the life of the system, with special
precautions being needed during and following the commissioning stage.
While directed mainly at in-building systems, Chapter 39 of ASHRAE Handbook
HVAC Applications (2011) contains much useful information on development and execution of an O&M program, including the life-cycle-cost implications, elements of successful O&M programs, O&M plan creation and implementation, setting O&M goals,
choosing an O&M strategy, automated fault detection, documentation, and staffing.
There are three basic maintenance strategies for mechanical systems: run to failure,
preventative maintenance, and condition-based maintenance (ASHRAE 2011).
Run to failure is a strategy applied when the cost of maintenance or repair may
exceed the cost of replacement or losses in the event of failure. Only minimum
maintenance such as cleaning or filter change is performed. The equipment may
or may not be monitored for proper operation, depending on the consequences
of failure. For example, a window air conditioner may be run although it is
vibrating and making noise, then replaced rather than repaired.
Preventive maintenance classifies resources allotted to ensure proper operation
of a system or equipment under the maintenance program. Durability, reliability,
efficiency, and safety are the principal objectives. Preventive maintenance can
be defined as the systemic and periodic inspection and servicing required to
keep equipment in proper operating condition. It means fixing things before they
break and thus keeping equipment in continuous service or ready for service.
The reliability of plant equipment depends largely upon its maintenance, and the

8.20

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

cost of operation in a well-maintained plant is consistently lower than in a


poorly maintained one. In addition, proper preventive maintenance results in
improved working conditions and better worker morale.
Condition-based maintenance uses manual and automated inspection and monitoring to establish the current condition of equipment. It also uses condition and
performance indices to optimize repair intervals.
When considering the basic maintenance strategies it is important to keep in mind
that forced outages for repair or replacement of equipment parts that have failed in service
can be very costly. Through the application of proper operating procedures and careful
inspection, it is possible to increase the length of time over which the components of the
district heating system can be operated before any repairs are required. This, in turn, will
prolong the useful life of the equipment and minimize forced maintenance. The principal
causes of forced outages and excessive maintenance are
sustained and frequent overloading of the equipment;
operating with improper flow, temperature, and pressure conditions;
fouling of heat transfer surfaces;
inadequate water conditioning;
improper lubrication; and
failure to test/inspect components such as valve vault sump pumps or drainage
systems.
Most maintenance programs combine more than one of the three basic approaches.
Predictive maintenance, which is a type of condition-based maintenance, attempts to further refine the maintenance program by supplementing projections that are normally statistically based with nondestructive testing such as infrared imaging, temperature
measurement, and vibration measurement and analysis. Some of these techniques may be
used as part of an automated fault detection system. Additional detail on maintenance
management is contained in ASHRAE HandbookHVAC Applications (2011).
In district heating systems where potentially hazardous conditions could be the result
of maintenance strategy, risk assessment should also be a component of the maintenance
program. Risk-based inspection procedures have long been used in the power, petroleum,
and petrochemical industries, and recently the process has been codified by ASME in a
standard (ASME 2012).
Due to elevated temperatures and pressures, the distribution portion of a district heating system may place additional responsibility on the district heating provider due to the
potential exposure of the public at large as well as of adjacent properties of others. Potential exposures may be addressed by hydraulic analysis, including failure modes (as discussed in the Chapter 4), or by thermal analyses (as discussed in Chapter 6).
The ability to performance maintenance will be significantly impacted by the design
of each aspect of the district heating system. Favorable maintainability is defined as the
ability to maintain the system easily, safely, and in a cost-effective manner. Maintainability should be a design objective and a maintainability review should be part of the design
process.
The plant and distribution system managers have the ultimate responsibility for
scheduling and performing preventive maintenance. They are responsible for all inspection and servicing required for the safety of the plant, distribution system, and consumer
interconnection. They will assign other operating or maintenance personnel the responsibilities for maintenance of specific pieces of equipment, as required by the preventive
maintenance record system. Some items listed for daily inspection by an assigned indi-

8.21

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

vidual also require hourly inspections by the operating personnel. These hourly inspections do not relieve the assigned operator of his/her daily responsibility to inspect,
service, and record the equipment conditions.
Inspection is the first step in a preventive maintenance program. The early detection
of a problem can greatly reduce the amount of damage, simplify maintenance, and prolong equipment life. The key to effective inspection is a complete understanding of the
equipments operating characteristics. The operator should know the condition, sound,
temperature, pressure, speed, vibration, and performance characteristics of each piece of
equipment in the plant, distribution system, and consumer interconnection, particularly
those for which he/she is assigned responsibility. Any change in normal characteristics
should be immediately reported, investigated, and corrected.
The supervisory control and data acquisition (SCADA) software integrated with the
computerized maintenance management system (CMMS) allows monitoring, diagnostics,
predictive maintenance, and process control to manage district heating system assets in an
optimum mode. Integrating asset management in the SCADA system allows close cooperation among operations, engineering, and maintenance personnel in order to obtain the
most economical system performance. An improved interface with the data collection
component of the SCADA software allows maintenance personnel to react more quickly
and with better information when equipment begins to behave less than optimally. Desirable functionality for the CMMS is outlined in the following section.

CMMS Functionality
The CMMS is a tool that helps to improve the reliability, utilization, performance,
and operability of equipment while minimizing the cost of labor and materials. Technology advancements have enabled many innovations in CMMS functionality. Technology
provides a platform, architecture, and delivery mechanism for the CMMS software
(Berger 2011).
Web-Based Technology
One of the most important influences on the CMMS development has been the
changing computer and internet technologies. The Microsoft Windows operating system resulted in improvements to the equipment and user interface, followed by report
and graphic components to help district heating system operators extract management
information from the CMMS. In recent years, CMMS vendors have developed a Webarchitected product in response to requests from users. This pattern of advancement will
continue with mobile solutions and wireless telecommunications.
Open Architecture
Open architecture enables plant operators to access the CMMS application and data
from any type of computer or handheld device, running any operating system and using
only a browser, from any location. This constitutes a cost-effective solution for plant managers.
Automating a Best-Practice Workflow
The CMMS allows controlling the district heating system maintenance process,
which may consist of handling a work request ordering parts or performing a predictive
maintenance routine, for example. The software facilitates the best-practice workflow at
each step in the process. A workflow component captures the process flow electronically
using customizable business rules, allowing operators to see graphically what the flow
looks like and to determine the current status of items moving through the flow. Work-

8.22

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

flow also enables some parts of the process, such as approvals and notification, to be automated by routing data to a specific persons email, pager, telephone, or handheld device.
Monitoring Equipment Condition to Predict Failure
One of the most visible improvements to the CMMS during the past few years has
been the increased emphasis on reliability. In an effort to move to a more planned environment, operations, maintenance, and engineering departments are working together to
minimize the potential for equipment failure through use of tools such as condition monitoring. The more sophisticated CMMS packages allow trend analysis for a variety of
meters and combinations of triggers, as well as alarming operators, scheduling a preventive maintenance routine, or taking corrective action automatically when a condition is
reached. Some CMMS vendors offer advanced graphics capabilities similar to what is
available with plant automation systems, thereby providing a real-time display of equipment conditions.
Spare-Parts Management
CMMS vendors have been building the softwares e-functionality. This is in response
to plant manager demand for e-commerce, e-procurement, and e-marketplaces. Operators
are looking to reduce costs by automating many of the supply-chain functions. Advanced
features include standardized electronic catalogues, portals of suppliers, electronic quotations and purchase orders, and electronic payment upon receipt of parts. Some CMMS
vendors offer features to manage spare-parts inventory, correlating service level to costs,
automatic correction of lead times and reorder points, handling consignment, and tracking supplier performance.
Physical Plant Management
There has been a steady improvement in CMMS features and functions that deal with
asset management. Modern CMMS packages can record reliability-centered maintenance
data, track warranty information for each component of a plant, build a troubleshooting
database, manage contract labor and materials, and drop work orders from one day to
another on a graphical work schedule while in simulation mode. More sophisticated packages have features for tracking equipment throughout their life cyclesfrom engineering
design to operation, maintenance, and disposal.
Finally, in todays green world, district heating system operators look for CMMS
packages that tie in sustainability features and functions, such as a thermal mapping of
operations or carbon footprint calculators.
Diagnostics
SCADA systems will often provide many of the fundamental inputs needed for an
automated CMMS. The self-monitoring and diagnosis component of SCADA categorizes
internal diagnostics into four standard status signalsfailure, function check, out of specification, and maintenance required. Diagnostic systems include a standard and open
interface for reporting all alarm conditions and provide a means of categorizing alert conditions by severity (OBrien 2011). The technology facilitates routing of alerts to appropriate consoles based on operator-selectable severity categories. In other words, it sends
the right information to the right person at the right time without flooding the operator
with alarms that are irrelevant to his/her duties. It also provides recommended corrective
actions and detailed help, as well as an indication of the overall condition of the equipment.
System operators should not become overwhelmed by the amount of diagnostic data.
The diagnostic data should also be reliable, allowing the operator to take action at the
right time. The information that the operator sees should follow best practices and guide-

8.23

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

lines for alarm management. True alarms should require immediate action, and care
should be given not to overload the operator with too many alarms and alerts.
Thermal Imaging
Thermal imaging applications and equipment may be valuable tools in a maintenance
program. The use of infrared (IR) methods is growing as the cost of the technology is
being reduced, allowing expanded use (Kennedy 2012). IR cameras have proven their
worth for detecting heat where it is unwanted. A complementary technology is the IR
window, which gives a view of energized electrical equipment without opening doors or
panels, thereby reducing personnel risk and reducing inspection time.
The widespread acceptance of thermal imaging as a maintenance practice forced
more emphasis on employee safety during IR scans. The 2012 edition of NFPA 70E outlines procedures for IR scans (NFPA 2012). IR windows make NFPA 70E-compliant IR
scans safer, faster, and easier.
Document Management
Blueprints, manuals, and information can be managed by the CMMS in electronic
format. The CMMS represents a transformation from a manual-based system to an electronic equivalent for inputting work and purchase orders, as well as preparing reports
such as equipment cost history, budget variance, and schedule compliance (Berger 2011).
The CMMS has the ability to link external documents. For example, CMMS users might
want to attach engineering drawings to a work order, and documents can be linked to
CMMS master files, forms, and reports.
The more sophisticated CMMS applications offer a wide range of features and functions related to document management. The document management package may
include:
Equipment operating manuals
Spec sheets and installation manuals
Troubleshooting guides
Engineering drawings, blueprints, and photographs
Contracts and warranty documentation
Training manuals and help videos
For new equipment, gathering the right documents in the right format is straightforward once policies, procedures, and a CMMS with even minimal document management
capability are in place. The difficulties lie with older assets. Documentation might be hard
to find and time-consuming to convert from hard-to-read paper formats. One option of
dealing with such assets is to scan and organize the documents using internal resources.
Once the image is digitized, the CMMS document management functionality takes
over. An important feature is the ability to generate templates for different document
types to facilitate information storage and retrieval. Each document should have a userdefinable template that provides information such as document type, document name,
description, date created, reference number, version, owner, and any related documents.
Template information allows indexing, searching, sorting, and filtering by any field on the
template and all integrated within the CMMS.
Another key document management feature to look for is the ability to establish a
document hierarchy. This allows the plant operator to view and search for documents
using an expandable, indented tree structure.
One advanced feature available with CMMS packages is the ability to mark up documents and save them as new files. For example, lubrication points might be added to a
digitized drawing or a photo of the equipment, and the new document might be attached

8.24

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

8 Operation and Maintenance

to a preventive maintenance work order. Each time the preventive maintenance comes
due, the work order will include the edited drawing.
Many documents tied to specific equipment do not remain static throughout the
equipment life cycle. As documents are changed or manipulated, there must be an audit
trail or history as to what changed when. The management of change or version control is
a critical feature of document management systems.
CMMS Work Breakdown Structure
Modern CMMSs can track project scope and link to work management and coordinate these with the project schedule (Reeve 2012). A major project scope may include a
shutdown procedure, outage measures, significant maintenance, and/or new construction
and system restart. A project might also be a software implementation or upgrade. The
work breakdown structure (WBS) is a deliverable-oriented grouping of project components. The WBS should be stored in the CMMS database. WBS-style project cost tracking emphasizes cost management within the CMMS and links work orders to cost
accounts and cost accounts to schedule activities. Cost accounts contain the budget, actual
costs, and estimate-to-complete values. The CMMS uses work orders to track work,
whether corrective or preventive, associated with specific equipment. The work orders
can have estimates as well as actual costs. The overall benefit is enhanced cost control
through project cost tracking.
A WBS provides the best start to managing a project. The WBS can be connected to
schedule activities through the cost account as a many-to-one relationship. A project
schedule, however, contains substantial detail, including logic ties, resource availability
constraints, and durations. The scheduling also allows for critical path analysis and
resource leveling. Using percent-complete input from a schedule, it is possible to calculate earned value, which is a true measure of where you are in the project. It is also possible to develop a WBS within the schedule software.
Most equipment shutdowns require detailed scheduling and critical path management. Larger projects also benefit from a WBS connected to a schedule. Work selected
for the outage gets downloaded to the schedule. Once the project starts, daily progress is
captured against the schedule, whereas actual hours go against a CMMS work order to
feed the WBS. Schedule activities can be summarized by cost account providing progress
summary to the WBS.
Because the CMMS already collects the necessary databudgets, purchase orders,
actual costsit makes sense to capitalize on a single point of entry. Eliminating redundant data entry frees up labor hours that can be used to ensure the CMMS works correctly.
Outside of scheduling, the CMMS can meet 100% of shutdown management needs.
The software can be enhanced to allow one product, with integration, to collect the data
necessary for managing the project. The ideal design involves a CMMS, a WBS, and a
project schedule. Incorporating the three into an integrated WBS design optimizes the
plant management. This overall design enhances the ability to manage scope, duration,
and cost.
Project Management
One of the many purposes of project cost reporting is to know whether the project is
on schedule and on budget before project completion. This project cost report can be used
for project management. Transactional capture tells when, what, and why the scope was
altered and who approved it. The change order column shows these dollar changes. The
report gets its data from the CMMS, allowing forecasting if the project will exceed the
budget. WBS-style project cost tracking simplifies cost management.

8.25

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

REFERENCES
ABMA. 2005. Boiler water quality requirements and associated steam quality for industrial/commercial and institutional boilers, Publication 402. American Boiler Manufacturers Association, Arlington, VA.
ASHRAE. 2012. ASHRAE Position Document on Legionellosis, ASHRAE, Atlanta.
ASHRAE. 2011. ASHRAE HandbookHVAC Applications. Atlanta: ASHRAE.
ASHRAE. 2008. ASHRAE HandbookHVAC Systems and Equipment. Atlanta: ASHRAE.
ASME. 1994. Consensus on operating practices for the control of feedwater and boiler
water chemistry in modem industrial boilers. Research Committee on Water in Thermal Power Systems, Industrial Boiler Subcommittee, American Society of Mechanical Engineers, New York.
ASME. 2012. ASME PCC-3-2007 (R2012), Inspection Planning Using Risk-Based
Methods. New York: ASME.
Berger, D. 2011. Software branches out. Plant Services 32(11).
Cho, Y.I. 2002. Efficiency of physical water treatments in controlling calcium scale accumulation in recirculating open cooling water system. ASHRAE Research Project RP1155 Final Report, ASHRAE, Atlanta.
DA. 1989. Technical ManualCentral Boiler Plants. TM 5-650, Headquarters, U.S
Department of the Army, Washington, DC.
DOE. 1998. Non-chemical technologies for scale and hardness control. Federal Technology Alert DOE/EE-0162, U.S. Department of Energy, Washington, DC.
Gan, F., D.-T. Chin, and A. Meitz. 1996. Laboratory evaluation of ozone as a corrosion
inhibitor for carbon steel, copper, and galvanized steel in cooling water. ASHRAE
Transactions 102(1).
IEA. 1996. A review of European and North American water treatment practices,
1996:N8, International Energy Agency (IEA) District Heating and Cooling, Paris.
Kennedy, S. 2012. See the safety. Plant Services 32(1).
Langelier, W.F. 1936. The analytical control of anticorrosion water treatment. Journal of
the American Water Works Association 28:1500.
Liu, Z., J.E. Stout, L. Tedesco, M. Boldin, C. Hwang, W.F. Diven, and V.L. Yu. 1994.
Controlled evaluation of copper-silver ionization in eradicating Legionella pneumophila from a hospital water distribution system. The Journal of Infectious Diseases
169:91922.
Nasrazadani, S., and T.J. Chao. 1996. Laboratory evaluations of ozone as a scale inhibitor
for use in open recirculating cooling systems. ASHRAE Transactions 102(2).
NFPA. NFPA 70E: Standard for Electrical Safety in the Workplace, 2012 ed. Quincy,
MA: National Fire Protection Association.
NIBS. 2012. Secure/Safe. Whole Building Design Guide. Washington, DC: National
Institute of Building Sciences. www.wbdg.org/design/secure_safe.php
OBrien, L. 2011. Coming to America. Plant Services 32(11).
Reeve, J. 2012. Too much informationPart 1. Plant Services 32(1).
Ryznar, J.W. 1944. A new index for determining amount of calcium carbonate scale
formed by a water. Journal of the American Water Works Association 36:472.
Sperko, W. 2009. Dissimilar metals in heating and AC piping systems. ASHRAE Journal
51(4).
Toth, E., and T. Merrill. 2009. Maintaining a safety culture: Distribution system a focus
at NRG Thermal. District Energy, First Quarter 2009.
U.S. Congress. 1970. Occupational Safety and Health Act of 1970, 91st Congress
(December 29), United States Code 29, 15.

8.26

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix A
96-Hour Boiling
Water Test
The following test procedure grew out of the efforts of the Federal Agency Committee on Underground Heat Distribution Systems to ensure that insulations used in systems
would be able to withstand boiling events. The version of the test procedure provided in
this appendix consists of excerpts from Procedures for Establishing Acceptability of
Underground Heat-Distribution Conduit Systems, dated July 1, 1964, and published by
Department of The Army, Corps of Engineers; Department of The Navy, Bureau of Yards
& Docks; and Department of The Air Force, Directorate of CE. For this information to
integrate smoothly into this book, the headings and text/footnote numbering have been
modified from the original and the content has been edited to be consistent with the style
of this design guide.

TESTS OF COMPLETE SYSTEM


Apparatus and Specimens for Other than Wet Poured Systems
A test box or tank of exterior dimensions approximately 19 4 4 ft (5.8 1.2 1.2 m)
shall be used to enclose the specimen under test. A 20 ft (6.1 m) test specimen conforming to the conduit design on which contract approval is desired shall be used, including
joints, vents, end seals, waterproofing membranes and sealants, insulation, casing materials, and other components required to make a complete system, and using 4 in. (100 mm)
pipe for conveying the heat. The specimen shall be installed along the longitudinal centerline of the tank or box with the system extending through both ends of the box. The specimen shall be supported in the box in a level or horizontal position with the top of the
conduit at least 14 in. (0.36 mm) below the top of the box or tank. The test specimen shall
be provided with a field joint installed at the approximate center of the specimen. If the
type of conduit tested requires a reinforced monolithic concrete base slab, the test specimen shall include a construction joint in the 20 ft (6.1 m) long concrete base slab.
Watertight joints shall be provided between the surface of the conduit system and the
ends of the tank or box. A drain that is flush with the bottom of the insulation or the airspace inside the cavity, whichever is lower, shall be attached to the cover plate at one end
of the specimen. The drain shall be attached to the lower end of the specimen and shall
be fitted with a valve. Glass fiber insulation, 1/4 in. (6.4 mm) in thickness or equivalent,
should be wrapped around and secured to the conduit within the box or tank to simulate
the thermal resistance of a typical cover of damp earth. A surge tank equipped with a
gage glass shal1 be attached to the lower end of the specimen to indicate the water level

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

in the specimen during the boiling test. The bottom of the surge tank shall be connected
to the bottom of the insulation space in the conduit with nipples and a valve for restricting surges between the tank and the specimen, and the top of the surge tank shall be connected to the steam space in the insulation cavity with suitable copper pipe or tubing. A
cold-water feed line shall be connected to the insulation cavity at the bottom of one end
plate of the conduit specimen. The vents for the interior cavity shall be identical in size
and point of attachment to that which the manufacturer proposes to use in the field.
Thermocouples shall be affixed to the exterior surface of the top of the conduit 5 ft
(1.5 m) from each end of the casing specimen and shall be secured to the casing specimen for at least 3 in. (76 mm) from the junction point. A pressure gage or mercury
manometer, whichever is better suited to the pressures anticipated, shall be connected
directly to the top of insulation cavity of the specimen at the high end. The 4 in.
(100 mm) steam supply pipe should be several inches (millimetres) longer than the conduit test specimen and be fitted with a flange at each end. A thermocouple should be
brazed to each of the two exposed ends of the 4 in. (100 mm) steam pipe, and the thermocouple leads should be secured to the pipe for at least 3 in. (76 mm) from the junction.
The flanges on the two ends of the steam pipe shall be matched with two blank plates
each, with one pair of blank plates to be adapted for supplying high pressure steam from
an external source. For electric heating, two of the blank plates shall each be fitted with
an electric resistance hairpin heater and a top and bottom pipe nipple. The nipples should
be flush with the inside surface of the 4 in. (100 mm) pipe at the bottom and nearly so at
the top. One bottom nipple shall be connected to an inlet water line; the other bottom
nipple shall be connected to a valve for draining the pipe. One upper nipple shall be connected to a tee for a safety pressure relief valve and a valve for venting the pipe of air; the
other upper nipple should be connected to a tee for a pressure gage and an adjustable
pressure controlling unit. For steam heating from an outside source, one blank plate shall
be fitted with a steam inlet pipe and the other blank plate shall be fitted with a bottom
nipple for steam trap and drainage and a top nipple for a pressure gage.

Test Procedure for Other than Wet Poured Systems


The electric heaters should be attached to the 4 in. (100 mm) steam pipe first and the
pipe filled with water until it flows from the vent connected to the top of the pipe. With
the heaters on, the vent should remain open until all air is removed from the pipe. After
the air has been removed, the vent should be closed, and after a small pressure has been
developed in the steam pipe by the electric heaters, a little water should be drained from
the drain pipe to create a slight steam space at the top of the pipe. The 4 in. (100 mm) pipe
should be operated as a closed steam boiler at a steam pressure of about 125 lb/in.2 gage
(8.6 bar gage) and a temperature of 350F (177C) at the pipe thermocouple positions for
a period of 24 h at the same time ventilating the insulation cavity with forced ventilation.
At this point, all openings into the conduit system should be closed and the ends of the
conduit system projecting from the test box or tank should be insulated with a minimum
of 6 in. (150 mm) of loose insulation contained in open-topped boxes attached to the ends
of the test apparatus and large enough to provide the 6 in. (150 mm) of insulation for the
flanges.
The open top of the large 19 ft (5.8 m) box should be covered with a thin sheet material during the heat loss test to limit free convection around the conduit without allowing
excessive temperature rise inside the box. The ambient temperature about the test apparatus should be maintained between 70F and 80F (21C and 27C) during the time of test.

A.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix A 96-Hour Boiling Water Test

The electric energy input required to maintain a steam pipe temperature of 350F (177C)
should be recorded hourly and continued until a steady input is recorded for at least 4 h.
After the steady-state heat loss of the new system has been determined for the dry
condition, the end insulation and the flange plates holding the electric immersion heaters
should be removed and the fitted flange plates for steam heating should be attached to the
4 in. (100 mm) pipe. The insulation cavity in the conduit should be filled with water to
saturate all the insulation and, after two hours, drained to lower the water level to the onehalf full mark, as indicated by the sight glass. Steam should be admitted to the pipe and
the pressure gradually increased while carefully watching the pressure gage attached to
the insulation cavity. If the pressure in the insulation cavity exceeds 15 psig (1.03 bar) on
Class A systems* or 12 in. w.g. (30.5 cm w.g.) on Class B systems before a steam pressure of 125 psig (8.6 bar) is reached in the pipe, the venting should be considered inadequate and should be corrected before proceeding with the test. The boiling test should
continue for 192 h with a steam pressure of 125 psig (8.6 bar) in the pipe and with the
water in the insulation cavity maintained throughout the test at the mid-height of the conduit. Temperature and pressure readings on the steam pipe, the pressure in the insulation
cavity, and temperatures on the exterior of the casing should be read at regular intervals
throughout the test.
After the boiling test has continued for 192 h, the water supply to the cavity should be
closed off and the excess water drained from the cavity. The insulation should be dried for
a period of 48 h by maintaining a steam pressure of 125 psig (8.6 bar) and ventilating the
insulation in the conduit. The maximum rate of ventilation for Class A systems should be
limited to a rate that (a) results in a pressure in the air space of 15 psig (1.03 bar) at the
inlet end or (b) results in an average linear velocity of 1000 ft/min (100 mm/min) in the
air spaces, whichever is lower. The maximum rate for Class B systems should be limited
to a rate that (a) results in a pressure of 12 in. w.g. (30.5 cm w.g.) anywhere in the air
space or (b) results in an average linear velocity of 1000 ft/min (100 mm/min) in the air
space of the conduit, whichever is lower.
After the drying period, the flange plates equipped for steam supply should be
removed from the 4 in. (100 mm) pipe and the flange plates equipped with electric
immersion heaters should be reconnected. The 4 in. (100 mm) pipe should be operated as
a closed steam boiler at a temperature of 350F (177C) using the procedure previously
described and the heating should be continued until the metered electric energy input rate
has again attained a constant value for at least a 4 h period.
Class A Systems
After completing the second heat loss test and while the steam pressure in the pipe is
maintained at 125 psig (8.6 bar), a 15 psig (1.03 bar) air pressure should be applied to the
insulation cavity of the conduit. After bringing the cavity of the system up to test pressure, the pressure should remain constant for two consecutive hours without addition of
air to the cavity.
After completing the air pressure test, air pressure and steam pressure should be
reduced to zero and the exterior insulation added for the heat loss tests should be removed
from around the conduit for inspection of the casing, membranes, tapes, adhesives, and
sealants. Immediately following inspection of the casing and waterproofing materials, the
*

See Heating and Cooling Distribution Systems, UFC-3-430-01FA, U.S. Department of Defense,
July 25, 2003, available at www.wbdg.org/ccb/DOD/UFC/ufc_3_430_01fa.pdf.
The average air velocity is to be determined by dividing the cubic feet per minute (cubic metres per second) at atmospheric pressure by the cross-sectional area of the ventilated air passages in a plane at right
angles to the centerline of the pipe.

A.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

steam pipe and its covering insulation should be carefully removed from the conduit.
Representative circumferential samples not less than 1 in. (25.4 mm) wide and the full
thickness of the insulation should be taken at right angles to the pipe at the center and at
the two ends of the conduit as soon as possible after removal from the conduit. Each sample should be quickly placed in an airtight container for a moisture determination. The
samples should be weighed in the airtight containers on an accurate balance or scale, after
which the container should be opened and placed in an oven at 215F (102C) until each
sample shows a constant weight. The percent of water by weight of each sample should
be determined from the initial and final weights of the containers after appropriate corrections are made for the weight of the empty container. The average water content of the
samples as taken should not exceed 5% by weight.
A visual inspection of insulation covering the pipe should be made, and photographs
should be taken to support visual evidence. Any one or more of the following kinds of
damage to the insulation constitute cause for rejection:
Portions fallen from the pipe
Eccentricity of the insulation relative to its original position
Significant opening of the longitudinal or circumferential joints
Significant erosion of the insulation at the joints
Cracks or ruptures in the insulation revealing the pipe surfaces
Physical or chemica1 changes in the insulation that are likely to impair its function
Spalling or delamination of the insulation

A.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix B
Climatic Constants
EQUATIONS
The climatic constants Tms, As, and tlag required for various calculations in Chapter 6
have been computed for all 5564 weather stations, domestic and international, listed in
Chapter 14 of ASHRAE HandbookFundamentals (2009) and are available at http://
tc62.ashraetcs.org/pdf/ASHRAE_Climatic_Data.pdf. If, however, climatic constants are
desired for another set of data either observed or purposefully fabricated, they may be calculated using the equations supplied here. If a solution is desired that is representative of
an average year, then average monthly data from many years of record will be needed.
Given appropriate monthly mean temperatures, the equations used are:
N

Ti
i=1
T ms = --------------N

and

2
A s = ---- C 12 + C 22
N


1 C
t lag = ------ tan ------2 + ---
2
C 1 2
where
N

C1 =

i=1
N

C2 =

i=1

i
ti
Ti
N

=
=
=
=
=

2t
T i cos ----------i

2t
T i sin ----------i

index for time increment


Julian date for temperature Ti, for average monthly data: 15.5, 45, 74.5, etc.
temperature at Julian date ti
total number of time/temperature sets, i.e., 12 for monthly data
total length of period, i.e., 365 (days)

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

These equations assume that monthly mean temperatures will be used; however, other
frequencies may be used as long as the data are evenly spaced and the appropriate adjustments are made to N and ti. For example, if daily data were to be used N would become
365 and the index i would run from 1 to 365 for the 365 Ti values while the corresponding
ti values would simply be 1, 2, 3, 365. The data set must represent one complete year
and not more.

REFERENCE
ASHRAE. 2009. ASHRAE HandbookFundamentals. Atlanta: ASHRAE.

B.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C
Case Studies
CASE STUDY A: DISTRICT ENERGY/CHP SYSTEM IN JAMESTOWN, NEW YORK
Dr. Ishai Oliker, PE, Principal, Joseph Technology Corporation, Inc.
INTRODUCTION
This case study addresses the development, construction and operation of the Jamestown district energy (DE)/combined heat and power (CHP) system in Jamestown, New
York. The conception of the system through the initial feasibility studies is discussed, followed by the development and construction of the system through phased implementation
and the current status of operation. The planning aspects that contributed to the successful
development of this system are highlighted and the customer savings are cited.
The DE/CHP concept was a result of a preliminary feasibility study financed by the
New York State Energy Research & Development Authority (NYSERDA) in October of
1981. The study examined the potential of a DE/CHP system to economically supply the
downtown area of Jamestown with thermal energy from an existing municipal single-purpose electric generation station. The results of the study were promising, indicating the
technical and economic viability of the project. In light of the positive findings, a comprehensive second-phase study was contracted in order to develop the necessary information
for a final decision. The objectives of the second-phase study included an engineering reference design as a basis for the financial analysis, a marketing program, a final design, the
engineering bid and specifications, the basis for the financial instruments for project
financing, and a project implementation plan. This study was financed by NYSERDA and
the City of Jamestown. Based on the favorable results of the feasibility studies, the City of
Jamestown committed to building a pilot system during the summer of 1984 and expanding it to include selected buildings in the downtown core area during the following years.

SYSTEM DESCRIPTION
The Jamestown DE/CHP system consists of three major components: the central
energy plant, the transmission and distribution network, and the participating buildings or
customers. Each component is addressed individually in the following subsections.

Central Energy Plant


The existing 50 MW Carlson Generating Station was selected as the central energy
source for the Jamestown DE/CHP system. The power plant includes four coal-fired boil-

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

ers and two steam-driven turbine-generator units (Units 5 and 6), both with General Electric non-reheat turbines. Unit 6 was selected for CHP modification considering its larger
heat output and relative ease of retrofit to cogeneration. This turbine is a 25,000 kW, 3600
rpm, 15-stage single-flow condensing unit designed to operate at 850 psig (5865 kPa)
steam pressure, 900F (482.2C) temperature, and 3.5 in. Hg (8.9 cm Hg) condenser pressure; it has a rated throttle steam flow of 238,072 lb/h (107,966 kg/h). Extraction steam
for regenerative feedwater heating is taken from four extraction points. The turbine has
one blanked-off extraction point at the 11th stage. The feedwater heating cycle consists of
four closed and one open feedwater heaters with makeup to the cycle through the condenser hotwell. Steam is extracted from the blanked-off 11th stage turbine for use in a
new district heat exchanger. Higher loads are served with additional steam from the auxiliary steam header and used in the existing auxiliary heat exchanger that is arranged in
series with the new district heat exchanger (Figure C.1). Single-purpose efficiency for
Unit 6 was 26.9%. After unit conversion to CHP operation, the efficiency increased to
32.5%, a 19% improvement. The maximum electric load reduction of the turbine-generator is 1.24 MWe during district heating operation. This reduction corresponds to a conversion ratio of 5.6:1 (reduction of one unit in electric generation results in a 5.6 unit gain in
heat generation, which is a very good ratio for an existing single-purpose turbine). Modifications to the turbine were not required since the redistribution of extraction flows is
minimal and all existing feedwater heaters remain in service without modification.
During peak heat load operation, the return water temperature is 160F (85C) with a
district heating water supply of 250F (121.1C). The district heating water flow rate during peak-load conditions is 498,000 lb/h (225,843 kg/h). The maximum extraction flow

Figure C.1 Jamestown DE/CHP supply diagram.

C.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

available from the turbines 11th stage provides heating for 379,000 lb/h (171,877 kg/h)
of district circulating water to 223F (106.1C). At the maximum heat load conditions,
119,000 lb/h (53,967 kg/h) of district heating water bypass the district and auxiliary heat
exchangers. The 379,000 lb/h (171,877 kg/h) of effluent water from the district heat
exchanger is passed through the auxiliary heat exchanger. This increases its temperature
to above 250F (121.1C) so that when it is mixed with the 223F (106.1C), 119,000 lb/h
(53,967 kg/h) bypassed water it will produce a total flow rate of 498,000 lb/h
(225,843 kg/h) at 250F (121.1C).
The district heat exchanger operates throughout the year, providing hot water for both
space heating and domestic use. The auxiliary heat exchanger operates about one third of
the year. The maximum operating pressure of the district heat exchanger is about 20 psia
(138 kPa); the maximum operating pressure of the auxiliary heat exchanger is 60 psia
(414 kPa) with a maximum steam flow of 18,900 lb/h (8,571 kg/h).

Transmission and Distribution Network


The transmission and distribution network transports district heating water from the
central plant to the customers and back (Figure C.2). It is an underground two-pipe closed
system with a maximum operating pressure of 232 psi (1600 kPa) and pumps sized for a
total design discharge pressure of 140 psi (966 kPa). The piping is sized for a maximum
velocity of 8 ft/s (2.44 m/s), based on the peak-load supply and return temperatures of
250F and 160F (121.1C and 85C).
The prefabricated conduit system consists of a carbon steel carrier pipe, polyurethane
insulation, a polyethylene casing, and a leak detection system. The leak detection system
combines alarm and fault locator capabilities and is built into the conduit during manufacture to protect the system and facilitate service.
This piping system meets European Standard EN 253 (CEN 2009) with regard to all
major construction details, including standards for fittings (EN 448 [CEN 2009a]), preinsulated valves (EN 488 [CEN 2011]), field joint assemblies (EN 489 [CEN 2009c]),
and the design of the system (EN 13941 [CEN 2010]). In this system the carrier pipe,
insulation, and casing are bonded together to form a single unit. A significant portion of
the forces caused by thermal expansion are passed as shear forces to the mating components and ultimately to the soil; thus, the system is restrained to a degree by the surrounding soil and hence the amount of expansion is significantly reduced.
The system uses piping with a wall thickness similar to Schedule 10 and Schedule 20.
Due to reduced piping material, the system not only is less expensive but also develops
reduced expansion forces and thus requires simpler methods of expansion compensation.
The extensive experience in Jamestown indicated that the welding of the thin-wall carbon
steel piping (in order to pass the X-ray test) has to be performed in accordance with special procedures specified by qualified metallurgical engineers and performed by trained
certified welders. Once the welders were properly trained, they produced adequate quality
of the piping joins.
The district heating piping is installed in shallow trenches requiring minimal excavation and no shoring. The conduits are laid directly in the trenches on a sand bed. There is
at least 6 in. (152 mm) clearance between conduits and between each conduit and the
adjacent trench wall. A homogeneous layer of stone-free sand covers the conduits with a
surface pavement on top. A detailed description of the EN 253 (CEN 2009) piping systems is provided in Chapter 4.
The system development in Jamestown is described in a subsequent section of this
case study, were the three phases of implementation are individually addressed.

C.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure C.2 Transmission and distribution piping system.

Buildings
The building conversions to district heating depended largely on the existing individual heating systems. The design philosophy for the building retrofits was based on the following considerations:
A plate-type heat exchanger is used in each building to transfer heat from the
district heating water supply to the building distribution system. This is necessitated by the high temperatures and pressures in the district heating distribution
network.
The district heating water supply temperature varies according to outdoor temperature, from a maximum of 250F (121.1C) on the design-day (outdoor temperature 3F [16.1C]) to approximately 170F (76.7C) during the summer

C.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

months. Therefore, the building heating system distribution temperature must


also be reset based on the outdoor temperature.
Building systems operating temperatures were selected to optimize the size of
system components, producing maximum temperature differentials between
supply and return.
Conversion of two-pipe steam heating systems to district heating was the most prevalent building retrofit. A plate-and-frame heat exchanger replaced the existing boiler as a
heat source. Existing steam and condensate piping, wherever possible, were used to form
a closed water loop, with the installation of circulating pumps, an expansion tank, and an
air removal system. All steam traps were removed and air vents were installed at system
high points.
Conversion of gas-fired hot-air heating systems involve the installation of a new hotwater heating coil in the return air duct, along with an associated plate-type heat
exchanger and closed-loop hot-water circulating system. In instances where a significant
amount of outdoor air was used, a preheater hot-water coil was installed in the outdoor air
supply duct.
Existing hot-water heating systems were the simplest and most cost-effective to retrofit. In most cases, it merely required the installation of a plate-type heat exchanger.
The conversion and interconnection of all district heating customers was timely and
economical. This is mainly attributed to the extensive bid packages, which introduced in
detail the various concepts of individual heating system conversion to district heating, and
to the training and consultations furnished by the consulting engineer from Joseph Technology Corporation, Inc (JTC).

Retrofitting Electrically Heated Buildings to Hot Water


For many years Jamestown customers enjoyed very low electric rates (34 cents/
kWh).This resulted in the wide use of electric resistance systems for space heating and
domestic hot water. However, the electrically heated buildings imposed a substantial peak
demand on the electric utility system, limited the capacity of the local electric distribution
system, and resulted in high operating costs for the customers. Therefore, conversion of
electrically heated buildings to hot water became of prime importance.
The development of the district hot-water system offered an opportunity to convert
the electrically heated buildings to hot water and interconnect them to the district heating
system. The advantages of the hot-water system versus the electric-based system are as
follows:
1. Substantial reductions in electric demand for the electric utility and in operating
costs for the occupants.
2. Substantial reductions in heat losses of the building envelope. A major source of
inefficiency in electrically heated buildings is uneven heating. Typically electrically heated buildings have no individual room temperature controls. This
results in overheating of the apartments, with an associated increase in energy
consumption. A hot-water system is equipped with individual room temperature
control devices and operates with variable supply temperatures and flow rates
closely following the demand.
3. The circulating hot water provides a better heating medium than electricity. An
electric heating system does not have any thermal capacity in the heating elements, which results in indoor temperature fluctuation. Water retains its heat and
stays in the heating element. Hot water yields a more uniform indoor temperature and overall comfort to the occupants. Hot water also provides superior heat

C.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

storage characteristics. In a hot-water system, a large volume of water in constant circulation provides a continuous heat reservoir available to handle peak
demands or swings in load.
4. Substantial reductions in the release of organic dust (dust frying) on electricheated surfaces, which have much higher temperatures than hot-water heating
elements. In order to keep the indoor air quality (IAQ) acceptable, a much higher
ventilation rate is required with electric heating (about 30%); a hot-water system
has a lower ventilation rate.
5. Flexibility of hot-water piping distribution systems. Since a hot-water system
uses forced circulation, the piping lines can follow the contours of the building
(as compared to steam piping, which always must be properly pitched for condensate drainage).
6. Safety. Because the temperatures of hot-water systems are much lower, hot
water is less dangerous than electric heating with respect to fire potential in the
heated space.
During many years of development of district energy systems with associated conversion of various building internal heating systems to hot water, the engineering consultant
developed and demonstrated in the field a number of innovative cost-effective retrofit
methodologies that allowed retrofitting electrically heated buildings to hot water at reasonable capital cost and payback period (about 57 years). The conversion methodology
used a combination of equipment and installation techniques that minimized costs, caused
minimal inconvenience to tenants, and resulted in aesthetically acceptable configurations.
To reduce costs, stamped steel radiators are used in place of baseboard-mounted finned
tubes typical of conventional installations. The stamped steel radiators operate with a
36F (20C) temperature drop, compared to the 20F (11C) drop typically used for
finned tubes. This reduced the required water flow by a factor of two, which allowed the
use of smaller pipes. Connections to radiators are made using prefabricated tee fittings. A
small depression in the wall is required to accommodate the lower tee fitting at connections to radiators.
The district heating hot-water pipe is brought underground to the mechanical room of
the building. A heat exchanger is installed to isolate the district water from the building
water. The following conversion methodology was used. The installation of the radiators
and lines required just a few hours per apartment. Lines were surface mounted and run
under attractive wood-grain plastic housings that resemble baseboard molding. Penetrations through walls were limited to two 3/4 in. (19 mm) holes for the supply and return
lines and are concealed by the housings. Each pair of lines feeds up to six apartments. The
radiators have self-contained thermostatic controls that require no wiring.
Operational experience demonstrates that conversion of electrically heated systems to
hot-water operation offers a reduction in operating costs by two to three times (depending
on the ratio of the electricity to gas/oil cost), energy efficiency improvements in the range
of 30% to 40%, substantial improvement of comfort conditions for the occupants,
improvement of IAQ, and reduction of the ventilation rate.

SYSTEM DEVELOPMENT
The successful development of the Jamestown DE/CHP system was a result of the
strong support from the city, the Board of Public Utilities (BPU), and NYSERDA, as well
as the technical expertise and well-orchestrated effort of the consulting engineer. The
overall cooperation and strong community support for the project enabled local officials

C.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

to enthusiastically promote the system, to obtain financing, and to meet an ambitious construction schedule.
Parallel development of the three main system componentscentral plant, distribution network, and building retrofitswas necessitated by the inflexible schedule. Work
had to be completed by the end of the summer in order for the system to be operable during the start of the heating season.

Phased Implementation Philosophy


The objective of phased implementation is to develop the system in stages, spreading
the capital expenditures in incremental investments over the development period and
allowing the system to generate revenues to offset the capital investment. The DE/CHP
system in Jamestown was developed in three phases, starting with a pilot system in the
first phase, a core system in the second phase, and planned annual growth in the third
phase. The pilot project initiated the effort in 1984, which eventually was expanded to the
core system in 19851986, and it has been growing ever since.
The purpose of the pilot project was to impart valuable experience in construction
and operation to the local district heating officials prior to embarking on a larger venture.
The pilot system was also used as a marketing tool to attract skeptical customers. The
authorization for the pilot system development was given in June 1984, with actual construction commencing in August and operation in November of the same year. The pilot
system served four buildings.
The second phase of district heating development in Jamestown involved the retrofit
and interconnection of 15 additional buildings the next two years (19851986), after an
aggressive marketing campaign. An extensive transmission system was installed as part
of the second phase development, providing the foundation for future growth.
The third phase is ongoing and involves the planned annual growth of the system.
This phase capitalizes on the existing network and merely requires the retrofit and interconnection of new customers along this transmission line. In 2013 the Jamestown piping
system is about 12 miles long and services 80 customers.

Marketing
The installation of the pilot system in 1984 created a public awareness that, coupled
with the marketing activities, replaced the initial skepticism with enthusiasm for district
heating and its benefits. The marketing aspects of the DE/CHP development in Jamestown involved the combined efforts of the Mayors Office, the BPU, other city officials,
and the consulting engineer.
Numerous public and private meetings were scheduled with prospective core customers in order to educate them and discuss the advantages of district heating for their buildings. A marketing campaign through the media, newspapers/magazines, radio, and
television was used to establish a public consciousness and acceptance, offering evidence
through the operation of the pilot system. Brochures were prepared to complement this
effort. The marketing venture targeted a diverse customer base, including schools,
churches, hotels, and hospitals, as well as retail, office, residential, and industrial customers. An ad hoc committee was formed under the sponsorship of the Mayors Office and
the BPU to develop a complete community awareness and involvement program. The
committee consisted of representatives from the following groups:
Major customers and contractors
Jamestown Manufacturers Association
Department of Economic Development

C.7

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Department of Industrial Development


Department of Public Works
As part of this coordinated effort, the consulting engineer examined prospective customers and presented them with economic packages indicating conventional heating
costs, district heating costs, and anticipated savings. The benefits and advantages of district heating were reiterated. Once a potential customer expressed interest in participating
in district heating, the consulting engineer was responsible for the conversion of their
heating system to district heating.

Ownership
Centralized energy projects have exhibited a wide range of ownership structures,
which can be classified as either private or municipal. For this project the municipal alternative was selected based on the minimal impact of regulatory constraints. Jamestown
had a distinct advantage over most other localities that have instituted district heating systems: it already operated a municipal electric plant. This electric utility was experienced
in dealing with the regulatory environment and was attuned to the citys needs and procedures. The existing structure of the Jamestown BPU presented a unique opportunity for
the city to institute a district heating system that is fully responsive to the interest of the
city, with only limited additional procedural, administrative, and managerial costs. The
Jamestown BPU also has existing authority to use public right-of-ways, which allows for
cost reductions. Other important factors in the selection of municipal ownership include
the federal and state tax-exempt statuses and the customer acceptance and trust of municipalities over profit-oriented, private entities.

SYSTEM ECONOMICS
The positive economic analysis results served as the cornerstone for the development
of the Jamestown DE/CHP system. The economic analysis was performed from the viewpoint of municipal ownership, utilizing its distinct advantages. The analysis determined
the annual carrying charges for the system and the unit cost of district heat. The analysis
employed the required revenue approach to determine the necessary charges for district
heating sales. The method used was to develop the total system costs and compare these
costs with the total quantity of heat sold to determine the minimum required charge for
district heating.
The operating expenses for the district heating system were composed of replacement
electricity costs, pumping costs, operation and maintenance (O&M) personnel, O&M
materials, and steam costs. The replacement electricity cost is charged against the district
heating system to compensate for the reduction in electrical output caused by the district
heating steam extraction requirements.

Rates
The charge for district heating was set at $7.00/MMBtu ($7.40/GJ) in 1984. This rate
was a direct result of the economic analysis using a detailed cash flow. It allowed the utility to pay back the loan acquired to construct the system and it allowed the customers to
experience energy savings. A rate of $6.00/MMBtu ($6.34/GJ) was instituted for large
users in 1990. Large user status is granted when the monthly consumption exceeds 300
MMBtu (317 GJ). The two rates remained constant until 1991, when a 10% increase was
approved by the BPU. In 2000 rates were $9.00/MMBtu ($9.50/GJ) for small users and
$7.50/MMBtu ($7.92/GJ) for large users. In 2012 the district heating rate was $14.00/
MMBtu ($14.80/GJ)about 50% less than what individual customers (not connected to
the system) are paying.

C.8

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

Financing
In the context of municipal ownership, the normal source of funding for a district
heating project is through the issuance of long-term revenue or general obligation bonds.
A long-term municipal bond offers a fixed interest rate over the life of the project. The
volatility and relatively high levels of interest rates on long-term obligations, at the time,
led to the development of a broader spectrum of tax-exempt alternatives, including shortterm and floating-rate longer-term instruments. Short-term tax-exempt alternatives
afforded the opportunity to take advantage of the substantially lower interest rates during
the construction period. However, short-term bonds were available if long-term bonds
were intended to be the ultimate debt. A short-term debt is considered to be any debt with
a maturity of less than one year.
The Phase I development of the Jamestown district heating system, involving the
institution of a pilot system, was financed with short-term bonds. The later phases were
financed with long-term bonds, including the refinancing of the first phase.

SYSTEM BENEFITS
The benefits derived from the implementation of a DE/CHP system are multifaceted.
CHP benefits include customer savings, environmental advantages, demand-side management (DSM) application, and potential for urban economic revitalization. Customer savings and environmental advantages are among the most important by-products of the DE/
CHP development in Jamestown, with the DSM application becoming increasingly recognized.

Customer Savings
During the twenty-five years of district heating system operation in Jamestown, customers have experienced a cumulative savings of $16 million from participating in this
DE/CHP system instead of operating their individual equipment. The savings rate
increases with any increase in the price of fuel. Customer savings are expected to rise in
the future as the system grows with minimal capital investment.

Environmental Advantages
DE/CHP is an energy conservation measure noted for increased thermal efficiency,
which is reflected in the energy savings of connected customers. Higher thermal efficiency corresponds to more useful energy output per given quantity of fuel input. In addition, centralization of load eliminates the sharp spikes in demand of individual buildings.
These load spikes result in the oversizing of equipment, which are selected for peak conditions and fail to provide optimum performance at other conditions. Load leveling permits the plant to operate at reduced peak and at longer sustained intervals, which
contribute to enhanced energy utilization. All this translates to lower emissions and consequently reduced environmental pollution.
Boilers used by individual customers cycle on and off, producing a loss in efficiency,
partially associated with the incomplete combustion of fuel. The continuous and efficient
operation of the CHP plant, however, reduces the carbon monoxide and hydrocarbon emissions, which are characteristics of incomplete combustion. To date, 130 individual boilers
have been shut down by system customers connected to the district heating system.

Demand-Side Management Application


The substantial increase of demand for electrical power over the past years, along
with the enormous costs associated with the addition of new capacity, have led to the
emergence of DSM programs to address the disparity of peak loads versus the base loads
as well as their seasonal variation. In general, these programs target demand reduction by

C.9

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

removing inefficiencies, a more economically viable option than the installation of new
capacity.
DE/CHP offers potential for DSM application by replacing electrical heating systems. Electrical systems are converted to hot water and interconnected with the district
heating system. This conversion to district service eliminates the electrical demand for
heating. The application of district heating is significant for the City of Jamestown considering the winter peaking characteristics of the electrical utility. To date, 500 Jamestown apartments have been successfully converted from electric heat to hot-water-based
district heating.

FUTURE EXPANSION
The Jamestown system has been expanded every year and in 2013 serves 80 downtown buildings. Additional expansions are planned and implemented every year. Since
2004, one satellite boiler plant supplies residential customers located far from downtown,
and the buildings of the Jamestown Community College (located two miles from downtown) are connected with piping to another satellite district heating system. A mile and a
half long pipeline connecting the downtown district and the industrial corridor is planned.
In 2002, the Jamestown BPU repowered the Carlson Generating Station with a General Electric LM6000 gas turbine, converting its steam-operated system to a more efficient and technologically advanced combined-cycle mode.
The LM6000 gas turbine coupled with a Deltak heat recovery steam generator
(HRSG) allowed an increase in the electrical generating capacity to 95 MW, increasing
the amount of waste heat available for district heating by 50% and placing less reliance on
the old coal boilers to produce steam, resulting in higher operating efficiencies and a
reduction in exhaust pollutants.
Since the combustion process of natural gas occurs at much higher temperatures than
those of normal boiler operation, it allows for the production of steam using the high-temperature exhaust gases from the gas turbine exit. This removes dependency on old, inefficient boilers to produce the steam and allows for much higher fuel utilization. Until 2002,
the plant operated with 32.5% efficiency. The addition of a gas turbine and HRSG
improved the overall efficiency to about 56% when district heating is included.
The HRSG also includes a coil to be used for district heating purposes. A portion of
the 170F (76.7C) condensate leaving turbine #6 enters the coil and is heated to approximately 375F (190.6C). This water then passes through a new heat exchanger and provides the district heating water with 25 MMBtu/h (26.4 GJ/h) of heat, or it passes through
a flash separator and turns into low-temperature steam to provide the deaerators heat and
decrease the amount of steam needed to be extracted from the turbines.

District Cooling System


In 1999 a small district chilled-water system was constructed in Jamestown. The system is powered by electrical chillers and supplies the city hall and hotels.

CONCLUSION
The successful development of a DE/CHP system in Jamestown, New York, is the
result of a well-coordinated effort, starting with the systems conception and continuing
into its operation and growth. The promising findings of the initial feasibility study were
pursued further in a second-phase detailed study, used to make the final decision. The system was implemented in stages, starting with a pilot project that was used as a marketing
vehicle, demonstrating the systems benefits, savings, and reliability. A coordinated effort
among the Mayors Office, the BPU, and the consulting engineer produced an effective

C.10

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

marketing campaign. Meetings and advertising created a public awareness, which eventually led to extensive community participation.
The consulting engineer met with the perspective customers, providing individualized
attention and marketing leverage. The firm produced comprehensive bid and specification
packages and trained the contractors in the retrofit process for the various heating systems. The design and planning of the consulting engineer produced a cost-effective development with phenomenally low installation and retrofit costs. The twenty-five year
operational record has proven the high reliability of the underground piping system,
which requires practically no maintenance. Some of the buildings domestic hot-water
plate-and-frame heat exchangers experienced plugging caused by deposits from untreated
building potable water; these heat exchangers were cleaned or replaced.
The system ownership and financing capitalized on the advantages offered by the
municipal avenue. The BPU promotes and operates the Jamestown DE/CHP system. This
enables the City of Jamestown to control the major sources of energy: electricity and
heating. The municipal control of these energy sources is used as an economic development tool by the city to attract new business.
The future of district heating in Jamestown appears very attractive considering the
planned development of a system with twice the capacity of the present load. Any system
growth beyond the existing customer base is expected to enhance the economic operation
of the system.

REFERENCES
CEN. 2009a. EN 253:2009, District heating pipes Preinsulated bonded pipe systems for
directly buried hot water networks Pipe assembly of steel service pipe, polyurethane thermal insulation and outer casing of polyethylene. Brussels: European Committee for Standardization.
CEN. 2009b. EN 448:2009, District heating pipes Preinsulated bonded pipe systems for
directly buried hot water networks Fitting assemblies of steel service pipes, polyurethane thermal insulation and outer casing of polyethylene. Brussels: European
Committee for Standardization.
CEN. 2009c. EN 489:2009, District heating pipes Preinsulated bonded pipe systems for
directly buried hot water networks Joint assembly for steel service pipes, polyurethane thermal insulation and outer casing of polyethylene. Brussels: European Committee for Standardization.
CEN. 2009. EN 13941:2009, Design and installation of preinsulated bonded pipe systems
for district heating. Brussels: European Committee for Standardization.
CEN. 2011. EN 488:2011, District heating pipes Preinsulated bonded pipe systems for
directly buried hot water networks Steel valve assembly for steel service pipes,
polyurethane thermal insulation and outer casing of polyethylene. Brussels: European Committee for Standardization.

C.11

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

CASE STUDY B: CROTCHED MOUNTAIN BIOMASS DISTRICT HEATING SYSTEM


Gary Phetteplace, PhD, PE, GWA Research LLC
INTRODUCTION
Within the range of district heating systems addressed by this design guide the
Crotched Mountain biomass district heating system is a small system. However, it is felt
that the inclusion of its case study is appropriate both because it illustrates a small-scale
district heating system and because it is one of the older of many small-scale biomass
facilities. Small-scale biomass heating, and in some cases cogeneration, systems have
proliferated in number recently in the Northern United States and Canada, not to mention
the many existing installations in Scandinavia and Northern Europe. The Biomass Energy
Resource Center (BERC) (www.biomasscenter.org), for example, provides a database
that indicates there are currently 174 biomass heating system facilities in United States
and Canada, of which 9 are cogeneration. Some of these facilities are stand-alone single
buildings without distribution piping, but many are multiple-building campuses and complexes, and in many cases the boiler plant is located in a separate building from the heat
load; thus, distribution piping is used and all of the elements of a district heating system
are present.

BACKGROUND ON THE CROTCHED MOUNTAIN FACILITY


The Crotched Mountain campus is located in Greenfield, New Hampshire, on a
1400 acre mountaintop site. The mission statement of Crotched Mountain is: Crotched
Mountain Foundation is dedicated to serving individuals with disabilities and their families, embracing personal choice and development, and building communities of mutual
support (CM 2013). The campus includes hospital, rehabilitation, educational, housing,
and recreational facilities. Due to its remote location, additional measures must be taken
to ensure the reliable supply of utilities at the site.
The facility has 13 buildings that are heated by the biomass district heating system,
which represents 86% of the heated floor space of the 28 buildings on the campus. The
connected buildings in 2012 totaled 320,000 ft2 (29,700 m2) and with future building
plans will be approximately 600,000 ft2 (55,800 m2) at system build-out. Prior to the retrofit, the buildings were heated primarily with fuel oil. Several buildings had interconnection of their heating and cooling systems prior to the biomass district heating project.
The biomass district heating system operates year round, satisfying not only space heating loads but also domestic hot-water generation, pool heating, and absorption air conditioning.

DETAILS OF THE RETROFIT


Fuel
The biomass fuel for the Crotched Mountain system is a form of wood chips called
bole chips. These are a high-quality wood chip that is manufactured from tree trunks
alone (including bark); no branches or foliage are included in the chipping process (see
the BERC publication Woodchip Heating Fuel Specifications in the Northeastern United
States (2011) for more information on this and other types of wood-chip fuels. This type
of fuel is preferred for small-scale systems as it permits unattended chip feeding with relatively few issues. At Crotched Mountain the received fuel is typically at a moisture content of 40%. The heat content of bole chips is typically 4785 Btu/lbm (11,130 kJ/kg) at
42% moisture content based on green weight, i.e., including the moisture (BERC 2011).
Crotched Mountain purchases its fuel from a single vendor, a local logging company,

C.12

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

under long-term (five-year) contracts. The use of a single vendor is a bit atypical for an
installation of this size, which more frequently uses a chip broker who works with multiple suppliers and customers and may be able to provide more reliable supply. Details on
fuel delivery and storage are discussed in the following subsections, as are the economics
and environmental aspects of burning the chip fuel.

Central Plant
The boiler plant is a single self-contained structure including two boilers, a backup
oil boiler, chip handling equipment, pumps, emissions control equipment, other ancillary
equipment, and chip storage (Figure C.3).
Boilers
There are two biomass-fueled boilers, one with a capacity of 4 MMBtu/h (1.2 MW)
and another with a capacity of 8 MMBtu/h (2.4 MW). The biomass boilers are a modular
design that basically uses a standard fossil-fuel-fired boiler as the heat exchanger.
Beneath the conventional boiler the biomass combustion system is placed (Figure C.4).
The boiler sections are horizontal fire-tube type with oversized 3 in. (76 mm) fire tubes
specifically for solid fuel use. The combustion system consists of a refractory-lined steel
combustion chamber, cast iron grates designed especially for the wood chips, and combustion air blowers (see Figure C.5). The Crotched Mountain central plant also has an oilfired boiler with a capacity of 4.9 MMBtu/h (1.4 MW) for backup purposes. In addition
there are other oil-fired boilers distributed on the facility that may be used for backup if
necessary.
Fuel Storage
The wood-chip fuel is stored in bunkers at the central plant (Figure C.6). These bunkers are arranged such that tractor-trailers with live floor unloading are able to back

Figure C.3 The Crotched Mountain central plant. Wood chips are delivered through the two
overhead doors into bunkers below this level.

C.13

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure C.4 The 8 MMBtu/h (2.4 MW) wood-chip-fired boiler. Note the wood combustor
under a conventional fossil-fuel-type boiler.

partially into the bunker and discharge their load. Because of the remote location of the
Crotched Mountain facilityit is essentially on top of a small mountain and in a climate
where snow and ice are commonadditional chip storage capacity was prudent. This
required a unique bridge arrangement within the chip bunker that allows the trailers to
back in farther and unload their chips into the larger bunker. Total bunker capacity is
approximately 150 tons (136 Mg) (about 5 to 6 tractor-trailer loads). This capacity will be
adequate for one week during winter weather once the system is at full build-out (twice
the load in 2012).
Fuel Handling
As with the handling of any solid fuel, the handling of the wood chips presents special challenges, especially in a small-scale unmanned plant such as the Crotched Mountain facility. The approach used at Crotched Mountain is a simple system of augers and
conveyers that has proven to be very reliable in this service, both at this facility and at
many others like it. At Crotched Mountain each of the boilers has its own fuel delivery
system. The fuel train consists of the following:
The chips are extracted from the fuel storage bunker by a traveling auger system
located at the bottom of the bunkers (see Figure C.7). The auger moves along the
bunker to ensure an adequate supply of chips by preventing bridging within the
chip mass.

C.14

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

Figure C.5 Cross section of biomass combustor.


Courtesy of Messersmith Manufacturing, Inc.

Conveyor belts carry the chips from the auger output to the hopper (see
Figure C.8).
Chips are stored in a metering bin at the boiler. The level within the metering bin
is regulated within limits by activating/deactivating the conveyor belts and traveling bunker augers (Figure C.8).
From the metering bins chips are fed into the combustor sections of the boilers
by metering augers.
Figure C.9 shows a cross section of a typical biomass-fueled plant that uses a chip
handling system similar to the one used at Crotched Mountain, although the plant shown
is not Crotched Mountain.
Emissions Controls
The Crotched Mountain biomass system uses two stages of stack gas treatment to
control particulates. The first stage is a cyclone separator and the second stage is a baghouse. The baghouse is equipped with automatic jet-pulse cleaning. The baghouse
increases the capture of small particulate matter (PM10) by a factor of approximately 10.
Because of the effectiveness of the emissions control system, the local regulatory authority is not requiring Crotched Mountain to submit to monthly emissions testing.

C.15

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure C.6 Chip storage bunker. Note the truck bridge allowing trucks
to back farther into the oversized bunker.

Figure C.7 The auger systems travel along the length of the chip bunkers
pulling chips onto the conveyor belts.

C.16

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

Figure C.8 Conveyor belts carry chips to the metering bins (center).

Ash Removal
Ash is removed from the boilers manually on a daily basis. The system operators estimate that plant maintenance, including ash removal, requires about two man-hours daily.
Control System
The operation of the plant is fully automated and the plant is normally unmanned. An
integrated control system controls the traveling augers, belt conveyors, metering augers,
and combustion air blowers.
System Pumping
Circulation through the distribution system and plant is provided by two 15 hp
(11.2 kW) pumps arranged in parallel and equipped with variable-speed drives (VSDs).
Circulating pressure is approximately 70 psig (4.8 bar), with the central plant located
approximately 100 ft (30 m) lower in elevation than the main campus of the facility.

Distribution System
The distribution system is direct buried at a depth of approximately 2.5 ft (0.76 m) on
average. The system is a tree-type structure without loops. The piping is Schedule 40 steel
with welded joints preinsulated with 2.5 in. (64 mm) of polyurethane foam insulation and
jacketed with polyethylene. The pipe sizes vary from a maximum of 8 in. (200 mm) to a
minimum of 1.5 in. (38 mm). The normal supply temperature is 190F to 200F (88C to
93C) and the T is normally 10F to 20F (5.6C to 11.1C), with 30F (16.7C) being
the design value.

Building Interconnection and Loads


Buildings are interconnected to the system with plate-and-frame heat exchangers. As
noted previously, the connected buildings in 2012 total 320,000 ft2 (29,700 m2) and with

C.17

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure C.9 Cross section of a typical biomass-fueled plant;


this is not the Crotched Mountain plant but is similar.
Courtesy of Messersmith Manufacturing, Inc.

future building plans will be approximately 600,000 ft2 (55,800 m2) at system build-out.
The average consumption of wood chips has been 3200 tons (2900 Mg) per year over the
4.5 years of its operation. This amounts to approximately 100 truckloads of chips per
year.

OPERATIONAL EXPERIENCE
The operational experience with the plant to date has been very favorable. As previously stated, the operators indicate that about two man-hours per day are required for routine maintenance, including ash removal. Shortly after start-up a problem was observed
with the distribution piping near the central plant that was attributed to water infiltration
due to improper field joint completion, and the issue was corrected. Aside from that, the
distribution system does not appear to have any significant heat losses, as it fails to melt
any snow on the ground above it despite the rather shallow burial depth.

ECONOMIC BENEFITS TO THE USER


The construction of the system was facilitated by two contracts, one for the boiler
plant of approximately $1.7 million and a second contract for the piping system and
building interconnection of approximately $1.4 million. In the first 4.5 years of operation
the biomass district heating system has saved $1.73 million in fuel costs. The current
price of the delivered biomass fuel is $55/ton ($61/Mg), which is equivalent to fuel oil at
$0.85/gal ($0.23/L) assuming 65% combustion efficiency for the wood biomass and 75%

C.18

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

for oil. Historically, wood chip prices have proven to be much less volatile than fossil fuel
prices; in fact, BERC (2013) states that for Vermont real prices for wood chips have actually declined over the last 20 years, failing to even keep pace with inflation.

ENVIRONMENTAL BENEFITS
The principal environmental benefit of the Crotched Mountain district heating system
is the use of a renewable biomass fuel that is nearly carbon neutral. The biomass fuel is
also renewable and highly sustainable, unlike fossil fuels. For Crotched Mountain the
avoided carbon dioxide emissions that would have occurred if oil were burned, as the
facility did in the past, total about 10,700 tons (9700 Mg) for the 4.5 years that the biomass plant has been in operation. In addition, the carbon footprint of the local harvesting
operations for the bole chips is much less than that for the production and delivery of fuel
oil, the majority of which is imported. The use of locally harvested biomass fuels also
carries with it far less potential for unwanted environmental impacts such as the Exxon
Valdez oil spill or the Deepwater Horizon explosion and oil spill.

SOCIETAL BENEFITS
There are significant benefits to the local area from the Crotched Mountain biomass
district heating plant, including increased and stabile employment for those in the local
forest products industry and the businesses that support them. In addition, the use of wood
chips provides a market for low-quality forest products. A stable market for low-grade
forest products benefits landowners, which in turn helps to promote responsible forest
stewardship and sustainable forestry practices. On a national level the use of locally harvested wood-chip fuel reduces the U.S. dependence on foreign oil, much of which comes
from politically unstable regions of the world.

CONTACT INFORMATION
For more information contact:
John Parisi
Director of Building Services
Crotched Mountain
One Verney Drive
Greenfield, NH 03047
603-547-3311 x2120
jparisi@CrotchedMountain.org

REFERENCES
BERC. 2011. Woodchip heating fuel specifications in the Northeastern United States.
Montpelier, VT: Biomass Energy Resource Center. www.biomasscenter.org/images/
stories/Woodchip_Heating_Fuel_Specs_electronic.pdf
BERC. 2013. Benefits for Schools and Communities. Montpelier, VT: Biomass Energy
Resource Center. www.biomasscenter.org/resources/fact-sheets/benefits-for-schoolsand-communities.html
CM. 2013. Mission and History, Crotched Mountain website. www.cmf.org/About-Us/
Mission-and-History

C.19

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

CASE STUDY C: DISTRICT HEATING CONVERSION


FROM STEAM TO HOT WATER AT SAVANNAH REGIONAL HOSPITAL
David W. Wade, PE, President RDA Engineering, Inc.
The following case study was first published in Proceedings of IDEA 86th Annual Conference (1995,
pp. 21725) and is reprinted with permission from International District Energy Association. For this
information to integrate smoothly into this book, the headings and figure numbers have been modified from
the original and the content has been edited to be consistent with the style of this design guide.

ABSTRACT
In 1992, the Georgia Department of Human Resources decided to replace the steam
and chilled-water distribution systems at the 16-building Savannah Regional Hospital in
Savannah, Georgia. A life-cycle cost (LCC) comparison indicated that conversion of the
system from high-pressure steam to low-temperature hot water was feasible and costeffective.
A comprehensive design that included a phased changeover of the district system and
modifications to building air-handling units was conducted by RDA Engineering, Inc. A
construction budget of $1.6 million was allocated. Construction began in January of 1994
and was substantially completed in December of 1994.
This case study summarizes the LCC benefits of conversion from steam to low-temperature hot water, technical aspects of the design, construction, and preliminary results
of operating cost reductions.

INTRODUCTION
Savannah Regional Hospital was constructed circa 1969. The group of buildings that
make up the complex is located on an 85 acre campus approximately 2 miles south of the
city of Savannah, Georgia. The facility was originally intended to provide a full range of
hospital medical services, with expansions of the campus planned for later years. A central energy plant located at the perimeter of the property was constructed to provide highpressure steam and chilled-water utility services to the campus buildings.
Since the original construction date, two maintenance buildings and an Activities
Therapy building have been constructed. As of 1995, the facility was being used for drug
and alcohol abuse treatment and long-term care for certain patients. Full hospital services
are not provided and no expansion of the complex is expected in the states health-care
program.
In 1991 and 1992, problems occurred due to leaks in the twenty-two-year-old underground piping distribution system. As a result, the Department of Human Resources initiated a project to replace the underground piping from the central mechanical plant to the
sixteen buildings on the campus.
The modernization program has renovated building heating and air-conditioning systems, replaced underground piping, converted the central boiler plant to supply hot water,
and installed a new computerized control system. These significant changes in the building mechanical systems have resulted in more efficient operation and provide opportunities for revised maintenance and staffing to reduce operating costs.

ORIGINAL SYSTEM DESCRIPTION


The original steam distribution system was fed by two gas/oil-fired steam boilers in
the central mechanical plant. Each boiler was rated at 11,700,000 Btu/h at 125 psig
(3430 kW at 8.6 bar). During normal operation, one boiler was active and the other was
kept in reserve as a standby.

C.20

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

Steam was delivered to the buildings by an underground piping system extending


around the perimeter of the campus (Figure C.10). Branch take-offs from the steam and
condensate mains to each building were constructed within concrete vaults. The underground steam and condensate piping were installed together in a steel conduit totaling
more than 10,600 ft (325 m) in length.
Chilled-water underground distribution piping included approximately 9800 ft
(300 m) of steel pipe, which was insulated with cellular glass material and wrapped with
a water-resistant coating. The chilled-water piping was installed around the perimeter of
the campus, near the steam distribution piping. The central plant contained two chillers,
each originally designed for 450 tons (95 kW) of capacity. Normal operating procedure
was to run one chiller with the other unit on standby.
Within each building, the steam piping was connected to a pressure-reducing valve.
Low-pressure steam served steam heating coils located in one or more air-handling units
in each building. Steam was also used for instantaneous domestic water heaters. In the
kitchen, steam was provided to steam kettles, a three-compartment steamer, a coffeemaker, dishwasher booster heaters, and the heating coils in eight makeup air units.
Chilled-water piping in each mechanical room was connected to the underground
chilled-water supply and return distribution piping. In the original system, the chilledwater pump in the central mechanical plant served as the primary pump and smaller
pumps in each mechanical room operated as secondary pumps. Over the years, most secondary pumps were found to be unnecessary and were abandoned.

Figure C.10 Existing distribution systems prior to retrofit.

C.21

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

SYSTEM PROBLEMS
There were several problems with the original underground piping systems that made
the replacement project necessary. The primary problem resulted from leaks in the condensate piping. Since steam and condensate mains were installed in a common conduit,
when a leak occurred in a section of the condensate pipe it resulted in flooding of the surrounding conduit. This flooding caused a breakdown of the steam and condensate pipe
insulation and corrosion of the pipe casing. Large leaks resulted in flooding of the vaults
and other sections of the distribution system conduits.
The hospital staff also reported several incidents of leaks in the underground chilledwater piping distribution system. Investigations during repair work showed that groundwater entered imperfections in the external wrap of the chilled-water piping, broke down
the pipe insulation, and corroded the pipe from the outside. As with steam piping, each
time a leak related repair was made, a major shutdown of the cooling system occurred due
to a lack of loops and sectional valves.
In addition to the problems experienced in the distribution system, there were related
problems within the buildings. For instance, several of the heating coils in the building
air-handling units had developed leaks. Several steam-to-hot-water instantaneous heaters
needed replacement, and many of the building condensate return units were in need of
major repair or replacement. Repair/replacement projects for these deficiencies had been
requested by the hospital staff, and a coordinated approach to repair or replace the interrelated system components appeared desirable.

REPLACEMENT OPTIONS
The Department of Human Resources project called for replacement of underground
piping systems to 16 buildings at Savannah Regional Hospital. It was anticipated that
existing steam and condensate lines would be replaced by a similar system. The new system would be downsized somewhat since expansion of the campus was not anticipated.
The steam system would be replaced with separate steam and condensate lines installed
in a common trench. The lines would be installed along an alternate routing across the
campus to minimize interference with other utilities. Sectional and branch valves would
be installed in new concrete vaults. The project anticipated replacing the chilled-water
distribution piping mains with uninsulated ductile iron pipe.
Early in the planning process it was discovered that capital funding had been
requested for the following fiscal year to replace air-handling units and heating coils,
which were also considered major maintenance problems. A thorough survey of the facility revealed no process steam uses other than the kitchen requirements. After discussions
with the owner, a study was conducted to consider total renovation of the energy system
to a new hot-water district system rather than continue with high-pressure steam. Advantages of the district hot-water system approach include the following:
Costs and hazards involved with the maintenance of steam traps and valves
would be eliminated.
Maintenance and operation of heating controls would be greatly simplified.
Maintenance of underground distribution piping would be less costly.
Hot-water piping would be installed only 2 or 3 ft (less than 1 m) below grade,
making repairs less complex.
Energy savings would be available through temperature setback and variable
flow rates.
The installed cost of the alternate hot-water system would be less than that of a
replacement steam system.

C.22

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

Operation and maintenance of the boilers and auxiliary equipment would be


simpler and less expensive. The possibility of reducing boiler operator supervision was also considered.

CONSTRUCTION PROGRAM
In mid-1992, RDA Engineering, Inc., was selected to design a replacement steam and
chilled-water system for the Savannah Regional Hospital campus. The Department of
Human Resources had previously determined that steam and condensate system failures
along with leaks in the underground chilled-water piping presented a long-term problem
to be corrected by installation of new distribution systems to the major buildings on campus. RDAs initial assessment of the project was presented in a LCC evaluation in November of 1992.
One of the results of the LCC evaluation was to recognize the benefits of combining
the planned campus-wide air-handling unit replacement project with the replacement of
the steam and chilled-water distribution piping. The combination of these projects
allowed replacement of the steam system with a more efficient hot-water heating system
and the addition of state-of-the-art computerized controls.
Bids on the construction project were taken in October of 1993 with a total projected
cost of approximately $1.6 million. The work progressed throughout 1994 and was substantially complete in December of 1994.
The work includes replacement of air-handling units in each major building, cleaning
of duct systems, installation of computerized monitoring and control systems in each
building with telemetry to the central plant and maintenance office, replacement of
domestic hot-water heaters in each building, installation of variable-volume terminals for
limited air zoning within buildings, and balancing of air systems to then-current occupancy and uses. Throughout the campus, new hot- and chilled-water supply and return
piping was installed. A small steam generator was installed at the kitchen to provide
steam for cooking. At the central plant, the existing steam boilers were converted to provide hot water and auxiliary steam equipment was removed. The chilled-water pumps and
piping within the central plant were replaced with new variable-speed pumps for efficient
operation. Figure C.11 illustrates the new underground piping system.

REVISED CENTRAL PLANT REQUIREMENTS


The conversion project substantially reduced the requirement for 24-hour staffing at
the central boiler plant. Steam boilers were eliminated, automatic controls were installed
to operate boilers and chillers, and remote monitoring capability was implemented.
The installed heating system design uses circulating hot water to provide heating to
the hospitals buildings. Hot water at a temperature of approximately 150F to 200F
(65C to 93C) is pumped from the central boiler plant throughout the hospital complex.
Differential pressure sensors located in three remote buildings send an electronic signal
back to the central control system, which controls the speed of the pumps supplying hot
water. This reduces energy consumption during off-peak periods. Hot water leaves the
central plant at a temperature of approximately 200F (93C) during the coldest winter
conditions. An automatic temperature control valve located in the central plant resets the
supply temperature in relation to outdoor conditions to a low of approximately 150F
(65C) in the summer.
The project converted each of the steam boilers to hot-water boilers. This means that
the boilers are flooded with water and provide heat by warming circulating water from
approximately 140F to 200F (60C to 93C). No steam is produced. Without steam production, the necessity of blowdown, makeup water through a deaerator, and extensive

C.23

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

chemical treatment and the danger of high-pressure steam are eliminated. The hot-water
circulating system is completely closed and should have very little, if any, water makeup
requirement.
Each hot-water boiler has an automatic control system that maintains the discharge
temperature at approximately 200F (93C). In the event of an abnormal operating condition, the boiler will shut down automatically and a signal will be noted by the computer
control system. The computer control system will then energize the standby boiler for
backup operation. Boiler status can be checked remotely from a dial-up computer terminal.
Thermal expansion of water in the heating system is automatically accommodated in
two expansion tanks located in the central plant. In the event abnormal high pressure is
present, a pressure relief valve will relieve pressure at the central plant. Water makeup to
the system is provided through a water meter with a tie-in to the computer control system.
Greater than normal amounts of makeup water cause an alarm condition to be reported.
The boilers will continue to need routine maintenance and daily checks; however, the
absence of high-pressure steam and the new automatic control revisions will allow the
boilers to operate unattended during times other than the day shift. Constant operator
attendance is not required.
The project also modified the chilled-water system by installing a new variable-speed
pump in the central plant. This pump is controlled from differential pressure signals at the
three building locations. Existing chiller controls were left in place and an interface for
on/off operation was added by the new computer control system. Automatic chilled-water

Figure C.11 Post-retrofit distribution systems.

C.24

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix C Case Studies

temperature reset is accomplished by the computerized control system in normal operation. As with the boiler system, the chillers should operate normally under their own control systems, and in the event of an abnormal condition, the computer control system is
notified and an alarm is generated. The chillers should receive routine maintenance and
daily inspections; however, continuous operator attendance should not be necessary.

ENERGY USE EVALUATION


The hospital staff provided RDA with electrical and natural gas consumption histories for the period October 1990 through February 1995. Figure C.12 illustrates the
monthly consumption of natural gas.
RDA estimates that for the original steam system, approximately 48% of the energy
was used for space heating, 15% for water heating and kitchen requirements, and 37%
was consumed as line losses of the underground steam piping.
RDAs review of monthly energy consumption concluded that some energy waste
occurred in mild weather conditions due to simultaneous heating and cooling. Replacement of air-handling systems and state-of-the-art controls were expected to reduce both
electrical and gas use in this area.
High natural gas consumption for line losses was due to the high temperature (320F
[160C]) of the steam system and failed insulation on the underground piping system.
Boiler efficiency was also adversely affected by the high-pressure steam requirements
and low summer loads. RDA estimated the steam plants annual gas-to-steam efficiency
at 73%.
During the construction period, the steam system continued to be operated as changeover work was undertaken in each building. For the months of May through July, one
boiler was converted to hot water while the other provided steam to the distribution system. Beginning in September 1994, the steam system was decommissioned and only the
hot-water system operated.

Figure C.12 Savannah Regional Hospital monthly gas consumption, 19911994.

C.25

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Figure C.13 Savannah Regional Hospital monthly gas consumption,


three-year average and 19941995.

A review of natural gas consumption for the period from September 1994 through
February 1995 reveals a dramatic reduction in energy consumption compared to the average of the previous three years. Figure C.13 illustrates that the new hot-water system used
approximately 58% less natural gas for the six-month period than the average of previous
years. This significant reduction is attributed to the lower line losses, higher system efficiency, and better system control available with the hot-water system.

C.26

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix D
Terminology for
District Heating
Some of the following material was extracted from ASHRAE Terminology of Heating, Ventilation, Air
Conditioning, and Refrigeration, published in 1991 by ASHRAE; however, terminology that is specific to
district heating systems has been added. The reader is referred to ASHRAE Terminology of Heating, Ventilation, Air Conditioning, and Refrigeration for terms not included here.

A
Adjustable-frequency drive (AFD): electronic device that varies its output frequency to vary the rotating
speed of a motor given a fixed input frequency. Used with fans or pumps to vary the flow in the system as a
function of a maintained pressure.
Air eliminator (air vent): in a steam or water distribution system, a device that closes if either steam or
water is present in the vent body and opens when air or noncondensables reach it.
Authority (of a controller such as a control valve): ratio of effect on a manipulated variable of one input
signal as compared to that of another.

B
Blowdown:
1. discharge of water from a steam boiler or open recirculating system that contains high total dissolved solids. The addition of makeup water will reduce the concentration of dissolved solids to
minimize their precipitation.
2. in pressure-relief devices, the difference between actuation pressure of a pressure-relief valve and
reseating pressure, expressed as a percentage of set pressure or in pressure units.
Branch: in piping or conduit, another section of the same size or smaller, at an angle with the main.
Breeching: passage for conducting the products of combustion from a fuel-fired appliance to a vent or
chimney.

C
Carrier pipe: the pipe that carries the heating or cooling medium (steam, hot water, chilled water). Has
more stringent requirements than a service pipe in a domestic water system. For example, compliance with
ASME B31.1-2012, Power Piping (2012), is required.

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

Carrier pipe insulation: insulation that surrounds the carrier pipe. Usually mineral fiber, calcium silicate,
foam glass, polyurethane, or polyisocyanuate foam. May be more than one layer and more than one type of
material.
Casing: a thin-wall pipe that encapsulates the carrier pipe and carrier pipe insulation to prevent the insulation from getting wet. Usually steel, high-density polyethylene (HDPE), or fiberglass-reinforced plastic
(FRP).
Cellular glass insulation: an insulation material manufactured of glass and carbon to form a structure with
millions of hermetically sealed cells.
Central heating plant:
1. one serving all or most of the rooms in a building, as distinguished from individual room heaters.
2. one serving two or more buildings, also termed district heating or district cooling.
Chimney: one or more passageways, vertical or nearly so, for conveying flue gases to the outside atmosphere.
Cogeneration: sequential production of either electrical or mechanical power and useful thermal energy
(heating or cooling) from a single energy form. See also electric power generation: cogeneration.
Combined heat and power (CHP) system: system combining power production with the use of a lowerquality heat by-product of power generation for district heating.
Community energy system: centralized facility for generation and distribution of the heating and cooling
needs of a community, rather than individual heat or cold generators (i.e., furnaces and/or air conditioners)
at each residential, commercial, or institutional site.
Condensate: liquid formed by condensation of a vapor. In steam heating, water condensed from steam; in
air conditioning, water extracted from air, as by condensation on the cooling coil.
Conduit: a section of a heat/cooling distribution piping system consisting of two or more of the following:
carrier pipe, carrier pipe insulation, casing, external insulation, outer jacket.
Consumer interconnection: see consumer interface.
Consumer interface: the interface between the consumer of district heating or cooling and the district
heating and cooling utility; this equipment is normally located within the building being served. The consumer interface normally includes controls and may also include metering and heat exchanger(s) where
required by the installation specifics.

D
Design professional: individual responsible for the design and preparation of architectural or engineering
contract documents. Also see Engineer of Record (EOR).
District cooling: concept of providing and distributing, from a central plant, cooling to a surrounding area
(district) of tenants or clients (residences, commercial businesses, or institutional sites). Compare district
heating.
District heating: concept of providing and distributing, from a central plant, heating to a surrounding area
(district) of tenants or clients (residences, commercial businesses, or institutional sites). Compare district
cooling.
Geothermal district heating (GDH): use of geothermal energy either directly or passing through a
heat exchanger to a secondary loop, from which the secondary fluid is circulated through the distribution system.

D.2

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix D Terminology for District Heating

Low-temperature district heating: use of fluids at temperatures lower than 120C (50F) for district
heating supply; permitting use of energy sources at lower temperatures than are traditionally used, for
example, that of waste heat from electric power generation.
Waste-heat reclamation district heating: use of collected heat from an available process as a source
for a district heating system.
District heating system heating density: measure of heating demand per unit area. Customary units
are Btu 106/acre (kW/hectare).

E
Electric power generation:
Baseload generation: large, steady electric load of an electric system produced by the largest, most
efficient generating facilities within the system. Note: These facilities are operated to the greatest
extent possible to maximize system mechanical and thermal efficiency and to minimize system operating costs.
Cogeneration: any of several processes that either use waste heat from generation of electricity to satisfy thermal needs or process waste heat in the steam generation of electricity. See also cogeneration.
On-site generation: generation of any electrical energy on a customers property, with or without the
use of recoverable heat.
End plate: steel plate at the end of a conduit section that makes the conduit a watertight vessel. The carrier
pipe passes through this plate and is either welded to the plate or the plate is fitted with a gland seal that
allows the carrier pipe to move axially with respect to the plate.
Energy transfer station (ETS): See consumer interface.
Engineer of Record (EOR): the technical person who is legally responsible for the design of the project.
A person registered by a state or government to be qualified to make design and construction decisions for
the specific type of project being designed and constructed. A person with experience with several projects
of the same type. See also professional engineer (licensed engineer) and registered engineer.
Entry pit: a structure located immediately outside of a building or inside of a building that is being serviced by an underground heating or cooling distribution system, usually partially below grade. It has many
of the same features of a valve vault: adequate room for maintenance on appurtenances, the ability to keep
groundwater out, a positive drainage system, electrical power for electrical sump pumps and lights, effective ventilation to control temperature and humidity, and safety features for maintenance workers.
Expansion bend: a bend, usually a loop, put into a pipe run to relieve stresses induced by expansion and
contraction from temperature changes.
Expansion joint: device in a structure, a pipe run, etc., that can by linear compensation accept variation of
length from expansion or contraction due to temperature changes.

F
FRP: fiberglass-reinforced plastic, a material used primarily in the construction of pipes and tanks.
Field joint: location in a piping system where the carrier piping segments are joined together. In a preinsulated piping system the field joint includes joining both the carrier piping as well as insulating the joint
and providing a leaktight jacket across the joint.

D.3

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

G
Gland seal: adjustable seal welded to the end plate that allows the carrier pipe to move axially through the
plate without allowing water to enter the conduit annular air space. Intended to be a watertight, airtight
seal.
GRP: see FRP.

H
HDPE: high-density polyethylene, a common material used in piping systems. See also PEX.
Hydraulic decoupler: a cross-connection between supply and return piping, e.g., at the chilled-water plant
used to decouple the flow through the chillers from that in the distribution system.

I
Infrared CO2 meter: instrument for estimation of carbon dioxide (CO2) content, based on absorption of
infrared radiation.
Infrared imaging system: apparatus that converts the two-dimensional spatial variations in infrared radiance from any object surface into a two-dimensional thermal map of the same scene in which variations in
radiance are displayed in gradations of gray tones.

J
Jacket:
1. sealed space around a piece of equipment or a storage unit, through which a thermal medium can be
circulated.
2. integral covering, sometimes fabric reinforced, that is applied over insulation or the core, shield, or
armor of a cable to provide mechanical or environmental protection.
3. a thin-wall pipe or watertight plastic wrap on the outside of the insulation that is exposed to soil or
the weather. Metal jackets are typically aluminum or stainless steel. Plastic jackets are typically
polyvinylchloride, high-density polyethylene (HDPE), or fiberglass-reinforced plastic (FRP).

L
Leak plate: a flat circular ring on the outside of the conduit that is bonded to the conduit and to the valve
vault wall to prevent groundwater from entering the valve vault at the conduit wall penetration.

M
Mechanical joint: general form for gastight joints obtained by joining metal parts through a positive holding mechanical construction (such as flanged joint, screwed joint, or flared joint). In utility piping, may
consist of joints sealed by O-rings or lip-type seals, as distinguished from fused (e.g., welded) or cemented
joints.

O
Outer insulation: typically polyurethane or polyisocyanuate foamed in place on the outside of the casing,
usually covered with a jacket.

P
PEX: high-density polyethylene (HDPE) that has been cross-linked; a process that provides higher
strength at elevated temperatures. See also HDPE.
Pipe duct: tube or conduit for encasing pipe.

D.4

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix D Terminology for District Heating

Pressure dependent (PD): adjective; varying flow rate through a flow control device in response to
changes in pressure.
Pressure independent (PI): adjective; uniform flow rate through a flow control device unaffected by
changes in system pressure.
Product-integrated control (PIC): direct digital control (DDC) panel, factory-mounted and connected,
able to monitor, control, and diagnose the significant functions of the equipment of which it is a part.
Professional engineer (licensed engineer): designation reserved, usually by law, for a person professionally qualified and duly licensed to perform engineering services such as civil, electrical, mechanical, sanitary, and structural. See also Engineer of Record (EOR) and registered engineer.

R
Registered engineer: appropriately qualified and licensed professional engineer. See also professional
engineer and Engineer of Record (EOR).

S
Steam system, one-pipe: system in which the condensable vapor from the supply main passes into a heating unit and returns as condensate through the same supply main.
Steam trap: device for allowing the passage of condensate and preventing the passage of steam or for
allowing the passage of air as well as condensate.
Float and thermostatic steam trap: trap that relies on the density of water to raise a float-and-lever
mechanism to operate a valve head. It discharges condensate as it forms and enters the trap body. It
includes a thermostatic balance pressure or bimetallic air vent to allow free passage of air on start-up
and discharges noncondensable gases reaching the trap during operation.
Inverted bucket steam trap: trap in which the inlet is channeled to the bottom of the trap body so that
the condensate enters underneath the inverted bucket. A small hole in the top of the bucket helps discharge noncondensable gases and entrapped air. The outlet from the trap is at the top, and as long as the
trap is filled with steam, it floats in the condensate and keeps the outlet closed. The valve opens when
the trap fills with condensate, sinks, and discharges the condensate into the return line.
Liquid-expansion steam trap: trap with an adjustable bellows and steel valve head located downstream of the steam system. It operates on temperature rise, with the setpoint adjustable from the outside. It discharges at a fixed temperature for protection from freezing and for high-capacity venting
requirements.
Thermodynamic steam trap (disk trap): trap constructed with a cap containing a steel disk, which
fits against a flat seat. Condensate, discharging at close to saturation temperature, increases in velocity
and draws the disk down toward the seat, due to the lower pressure caused by the increased velocity
(the Bernoulli effect). Condensate discharging from high to low pressure flashes off and creates the
closing pressure above the disk within the cap. As this flash steam condenses, pressure is dissipated,
and the cycle repeats. The trap has limited air-venting capabilities.
Thermostatic balanced pressure steam trap: trap installed on the discharge side of a heating unit
and designed to pass air freely on start-up and condensate at a subcooled temperature but to prevent
steam vapor passing into the return. It can have a bellows or encapsulated metallic diaphragm containing a small quantity of volatile liquid. At the bottom of the diaphragm or bellows is attached a hardened, self-centering valve head operating on the pressure side of the valve seat. At ordinary
temperatures and atmospheric pressure, the valve is fully open to permit free passage of air and cold

D.5

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

District Heating Guide

condensate. The trap discharges at a fixed temperature below that of the steam saturation temperature
and closely follows the steam pressure/temperature curve.
Thermostatic bimetallic steam trap: trap installed where low-temperature discharge is required. It
incorporates a bimetallic element that, when heated, deflects and causes a downstream valve head to be
drawn up, closing the orifice. It discharges air and cold condensate freely on start-up. It is installed
only where a high waterlogging effect is tolerated.
Upright bucket steam trap (open-top bucket trap): trap with a bucket that is usually hinged so that
when condensate enters in sufficient quantity to fill the bucket and it sinks, the downward motion
opens a valve. Water is then discharged into the return line through a tube that extends almost to the
bottom of the bucket. After the discharge, the bucket returns to its normal position. Now seldom manufactured.
Storage cycle (thermal storage): complete charge and discharge of a thermal storage device.
Stratification: division into a series of layers, as with thermal gradients across a stream.

T
TCP/IP: transmission control protocol/internet protocol; a protocol suite developed by the U.S. Department of Defense to permit different types of computers to communicate and exchange information with
one another.
Thermal insulation: material or assembly of materials used to provide resistance to heat flow.
Blanket thermal insulation: relatively flat and flexible insulation in coherent form, furnished in units
of substantial area.
Block thermal insulation: rigid insulation preformed into rectangular units.
Board (slab) thermal insulation: semirigid insulation preformed into rectangular units having a
degree of suppleness particularly related to their geometrical dimensions.
Cellular elastomeric (cellular rubber) thermal insulation: insulation composed principally of natural or synthetic elastomers, or both, processed to form a flexible, semirigid or rigid foam having a predominately closed-cell structure.
Cellular polystyrene thermal insulation board: insulation composed of cellular polystyrene in the
form of boards, produced by heat and pressure (1) from expansion of foamable polystyrene beads
within a mold (beadboard) or (2) by in-situ foaming of molten polystyrene in an extrusion mode
(extruded board).
Cellular polyurethane thermal insulation: insulation composed principally of the catalyzed reaction
product of polyisicyanate and polyhydroxy compounds, processed usually with fluorocarbon gas to
form a rigid foam having a predominately closed-cell structure.
Fill thermal insulation (loose fill): insulation in granular, nodular, fibrous, powdery, or similar form
designed for installation by pouring, blowing, or hand placement. Examples are mineral or glass fiber,
cellulosic fiber, diatomaceous silica, perlite, silica aerogel, and vermiculite.
Foamed-in-place thermal insulation (foam-in-situ insulation): insulation formed by introducing
into prepared cavities a chemical component and a foaming agent that react to fill the space with a
foamed plastic.
Mineral fiber thermal insulation: insulation composed principally of fibers manufactured from rock,
slag, or glass, with or without binders.

D.6

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

Appendix D Terminology for District Heating

Perlite thermal insulation: insulation composed of natural perlite ore, a glassy volcanic rock
expanded by heat to form a cellular structure.
Reflective thermal insulation: insulation that reduces radiant heat transfer across spaces by use of
one or more surfaces of high reflectance and low emittance, for example, aluminum foil.
Thermal storage:
1. temporary storage of high- or low-temperature energy for later use.
2. accumulation of energy in a body or system in the form of sensible heat (temperature rise) or latent
heat (change of phase).
3. technology or systems of accumulating cooling or heating capacity for subsequent use.
Cool storage: technology or systems used to store cooling capacity.
Ice-on-coil thermal storage: container (tank) in which ice is formed on tubes or on pipes.
Heat storage: technology or systems used to store heating capacity.
Latent storage: use of a phase change of a medium for storing heating or cooling capacity.
Naturally stratified storage: thermal storage in which temperature stratification is achieved and
maintained by density differences alone and not by mechanical separators.
Sensible storage: use of a change in temperature of a medium for storing heating or cooling capacity.
Stratified storage: thermal storage vessel in which a thermocline exists.
Thermocline: layer of fluid in which the temperature and density gradient is greater than, and which separates, the cooler fluid below it and the warmer fluid above it.
Cooling tower ton: total heat-rejection capacity of a cooling tower; traditionally, 15,000 Btu/h (4.4 kW).
Note: This value is based on 25% compressor heat added to a ton of refrigeration.
Ton-day of refrigeration: heat removed by a ton of refrigeration operating for a day, 288,000 Btu (approximately 84.3 kW). It is a quantity approximately equal to the latent heat of fusion or melting of 1 ton
(2000 lb) of ice from and at 32F (0C).
Ton-hour: quantity of thermal energy in tons (12,000 Btu [3.52 kWh]) absorbed or rejected in one hour.
Treated sewage effluent: the end product of the sewage treatment process that may be used for applications such as condenser water in regions where fresh water is limited.

V
Valve vault: a valve vault is distinguished from a manhole because of extra features that help improve the
life expectancy of a heating or cooling distribution system. A valve vault has adequate room for maintenance on appurtenances, provisions to attempt to preclude the entrance of groundwater, a positive drainage
system, electrical power for electrical sump pumps and lights, effective ventilation to control temperature
and humidity, and safety features for maintenance workers.

D.7

2013 ASHRAE (www.ashrae.org). For personal use only. Additional reproduction, distribution, or
transmission in either print or digital form is not permitted without ASHRAE's prior written permission.

DISTRICT

Complete Design Guide for District Heating Systems

HEATING GUIDE

District Heating Guide provides design guidance for all major aspects of district heating systems,
including central heating plants, distribution systems, and consumer interconnections, for both
steam and hot-water systems. It draws on the expertise of an extremely diverse international
team with current involvement in the industry and hundreds of years of combined experience.
In addition to design guidance, this book also includes a chapter dedicated to planning, with
information on cogeneration of heat and electric power as well as the integration of thermal
storage into a district heating system. Guidance on operations and maintenance is also provided
to help operators ensure that systems function as intended. In addition, examples are presented
where appropriate, and three detailed case studies are included in an appendix.
This guide is a useful resource for both the inexperienced designer as well as those immersed
in the industry, such as consulting engineers with campus specialization, utility engineers, district
heating system operating engineers, central plant design engineers, and steam and hot-water
system designers. District Heating Guide fills a worldwide need for modern and complete design
guidance for district heating systems.

ISBN 978-1-936504-43-5

ASHRAE
1791 Tullie Circle
Atlanta, GA 30329-2305
404-636-8400 (worldwide)
www.ashrae.org

ASHRAE_DH_Guide_On-Template.indd 1

DISTRICT HEATING GUIDE

Comprehensive Reference
Planning & System Selection Central Plants Distribution Systems
Heat Transfer Calculations System O&M Consumer Interconnection

9 781936 50443 5

Product code: 90562

8/13

7/30/2013 9:58:51 AM

S-ar putea să vă placă și