100% found this document useful (8 votes)
4K views350 pages

ASHRAE Design Guide For Natural Ventilation (2021)

Uploaded by

Hany Mohamed
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (8 votes)
4K views350 pages

ASHRAE Design Guide For Natural Ventilation (2021)

Uploaded by

Hany Mohamed
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 350

ASHRAE Design Guide for

Natural Ventilation
Peter Simmonds • Erin McConahey
ASHRAE Design Guide for
Natural Ventilation
ABOUT THE AUTHORS
Peter Simmonds, PhD, Fellow/Life Member ASHRAE, has been involved
in the design and operation of naturally ventilated buildings around the world
for more than 40 years. Dr. Simmonds has authored or coauthored more than
60 technical papers, articles, and books. He is the author of the ASHRAE
Design Guide for Tall, Supertall, and Megatall Building Systems (2015, 2020)
and a coauthor of ASHRAE/REHVA’s Active and Passive Beam Application
Design Guide (2014).
Peter received the Carter Bronze Medal for Best Paper/Presentation in
1993 from the Chartered Institution of Building Services Engineers (CIBSE)
and The Crosby Field Award (2019) and the John James International Award
(2020) from ASHRAE. He has also been awarded the ASHRAE Distinguished
Service and Exceptional Service awards.

Erin McConahey, PE, LEED AP (BD+C) HBDP, Fellow ASHRAE, is a


principal in mechanical engineering at Arup. With more than 25 years of
design leadership and project management experience on complex projects in
the built environment, Erin has been recognized with Fellow status by both
Arup and ASHRAE. She is a semiannual contributor to ASHRAE Journal’s
“Engineer’s Notebook” column. With degrees in mechanical engineering,
structural engineering, and ethical leadership, Erin’s strength lies in leading
teams to develop multidisciplinary solutions that balance technical needs with
long-term benefit.
ASHRAE Design Guide for
Natural Ventilation
Peter Simmonds • Erin McConahey

Peachtree Corners
ISBN 978-1-947192-54-6 (paperback)
ISBN 978-1-947192-55-3 (PDF)
© 2021 ASHRAE
180 Technology Parkway, Peachtree Corners, GA 30092 · www.ashrae.org
All rights reserved.
Cover design by Laura Haass. Cover photos from iStockphoto.com and Arup.
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 book 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 book be reproduced, stored in
a retrieval system, or transmitted in any way or by any means—electronic, photocopying, recording, or other—without
permission in writing from ASHRAE. Requests for permission should be submitted at www.ashrae.org/permissions.

Library of Congress Cataloging-in-Publication Data


Names: American Society of Heating, Refrigerating and Air-Conditioning
Engineers.
Title: ASHRAE design guide for natural ventilation
Other titles: Design guide for natural ventilation
Description: Atlanta, GA : ASHRAE, 2020. | Includes bibliographical
references. | Summary: “ASHRAE Design Guide for Natural Ventilation
assists owners, architects, engineers, facilities personnel, and
building design professionals in exploring the feasibility of natural
ventilation for their building project during the early phases of
design”-- Provided by publisher.
Identifiers: LCCN 2020012206 | ISBN 9781947192546 (paperback) | ISBN
9781947192553 (adobe pdf)
Subjects: LCSH: Natural ventilation.
Classification: LCC TH7674 .A84 2020 | DDC 697.9/2--dc23
LC record available at https://lccn.loc.gov/2020012206

ASHRAE STAFF
SPECIAL PUBLICATIONS Cindy Sheffield Michaels, Editor
James Madison Walker, Managing Editor of Standards
Lauren Ramsdell, Associate Editor
Mary Bolton, Assistant Editor
Michshell Phillips, Senior Editorial Coordinator
PUBLISHING SERVICES David Soltis, Group Manager of Electronic Products
and Publishing Services
Jayne Jackson, Publication Traffic Administrator
DIRECTOR OF PUBLICATIONS
AND EDUCATION Mark S. Owen

Updates and errata for this publication will be posted on the


ASHRAE website at www.ashrae.org/publicationupdates.
CONTENTS
Foreword ........................................................................................ ix
Preface............................................................................................ xi
Acknowledgments......................................................................... xiii
CHAPTER 1—INTRODUCTION .........................................................1
1.1 Purpose of the Guide ................................................... 1
1.2 Organization of the Guide ............................................ 2
1.3 How to Use the Guide .................................................. 3
1.4 When Not to Use Natural Ventilation ............................. 6
1.5 Perceived Benefits and Risks of Natural Ventilation ........ 11
1.6 Applicable Codes, References, and Standards .............. 23
1.7 References................................................................. 23
APPENDIX A1.1—Model Parameters for Office Examples ..... 27
A1.1.1 Heat Load to the Space .............................................. 28
A1.1.2 Ventilation Requirements............................................. 32
A1.1.3 References................................................................. 32
APPENDIX A1.2—Model Parameters for Classroom Examples.. 33
A1.2.1 Heat Load to the Space .............................................. 34
A1.2.2 Ventilation Requirements............................................. 37
A1.2.3 References................................................................. 38
CHAPTER 2—KEY CONCEPTS .........................................................39
2.1 Natural Ventilation versus Natural Conditioning............ 39
2.2 Mixed-Mode Ventilation.............................................. 39
2.3 How Does Natural Ventilation Work?........................... 43
2.4 Planning for Buoyancy-Driven Ventilation ..................... 45
2.5 Planning for Wind-Driven Ventilation............................ 49
2.6 Designing Successful Naturally Conditioned Spaces ...... 55
2.7 References................................................................. 63
APPENDIX A2.1—Equations for Simple
Buoyancy-Driven Ventilation ..................... 65
APPENDIX A2.2—Equations for Wind-Driven Ventilation...... 69
APPENDIX A2.3—Equations for Terrain-Corrected
Wind Ratios ................................................ 71
APPENDIX A2.4—Pressure Coefficients for Simple
Rectilinear Geometries............................... 73
A2.4.1 Low-Rise Buildings...................................................... 74

v
vi | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A2.4.2 High-Rise Buildings .................................................... 76


A2.4.3 Reference.................................................................. 77
APPENDIX A2.5—Interpolated Pressure Coefficients Tables .....79
CHAPTER 3—OUTDOOR CONDITION ASSESSMENT ..................... 93
3.1 Air Quality ................................................................ 94
3.2 Outdoor Noise Environment ....................................... 97
3.3 Wind Climate .......................................................... 100
3.4 Historical Weather Data ........................................... 107
3.5 Future Climate Trends .............................................. 119
3.6 References .............................................................. 124
APPENDIX A3.1—Outdoor Air Quality Assessment Example...129
APPENDIX A3.2—Outdoor Versus Indoor Noise Assessment
for a Naturally Ventilated Office............. 135
APPENDIX A3.3—Interpreting Wind Rose Data
Between Sites .......................................... 137
APPENDIX A3.4—Interpreting Wind Rose Data by
Outdoor Air Temperature ........................ 139
APPENDIX A3.5—Maps of Natural Conditioning Potential
by Season ................................................ 141
APPENDIX A3.6—Use of Climate Consultant for the
Adaptive Comfort Model ......................... 145
A3.6.1 Step 1. Produce a Bioclimatic Chart to
Represent all Strategies............................................. 145
A3.6.2 Step 2. Select Preferred Strategies to
Represent Natural Conditioning ................................ 146
A3.6.3 Step 3. Determine if a Heating System is Required ...... 147
A3.6.4 Step 4. Determine if a Backup Mechanical
Cooling System is Required....................................... 147
A3.6.5 Step 5. Determine if an Infiltration-Resistant
Envelope is Required................................................ 149
A3.6.6 Caveats when using Climate Consultant for the
Adaptive Comfort Model. ......................................... 149
A3.6.7 References .............................................................. 152
APPENDIX A3.7—The Presumption of Global Warming and
RCP Definition....................................... 153
CHAPTER 4—FAÇADE AND BUILDING
CONFIGURATION ANALYSIS ...........................................161
4.1 Balancing Daylight and Natural Ventilation Access in
Perimeter Zones....................................................... 162
4.2 Assessing Building Configuration for Natural Ventilation...164
4.3 Considerations for Placing Natural Ventilation Openings ... 167
4.4 Selecting Opening Types for Natural Ventilation ......... 171
4.5 Selecting Window Glazing for Reduction of
Conductive Heat Gains and Losses ........................... 179
4.6 Selecting Window Glazing for Solar Control............... 188
4.7 Limiting WWRs to Control Heat Gain ........................ 191
4.8 Configuring Shading and Blinds in
Natural Ventilation Schemes ..................................... 192
4.9 References .............................................................. 197
APPENDIX A4.1—Operable Element Types ......................... 199
CONTENTS | vii

APPENDIX A4.2—Estimating Infiltration .............................. 201


CHAPTER 5—NAVIGATING VENTILATION COMPLIANCE ............209
5.1 Introduction to Available Ventilation Compliance Paths.... 210
5.2 Wind and Thermal Buoyancy Formulas to
Determine Airflow through an Opening...................... 211
5.3 Prescriptive Design Requirements in
ANSI/ASHRAE Standard 62.1-2016 and
Older Versions......................................................... 212
5.4 ANSI/ASHRAE Standard 62.1 on Path A and Path B
Opening Sizing ........................................................ 213
5.5 Prescriptive Path on Opening Compliance per
California’s Title 24 Code......................................... 213
5.6 Engineered Natural Ventilation Systems...................... 214
5.7 Wind-only Window Sizing Procedure.......................... 216
5.8 Estimating Appropriate Size for Natural Ventilation
Openings for Minimum Outdoor Air Quantities on
Cold Days ............................................................... 217
5.9 Energy Modeling for Natural Ventilation in
ANSI/ASHRAE/IES Standard 90.1-2019..................... 217
5.10 Natural Ventilation Documentation for the
LEED® Rating System................................................ 218
5.11 References............................................................... 220
APPENDIX A5.1—Foundational Calculations for
Envelope Flow Models ............................. 223
APPENDIX A5.2—ASHRAE Standard 62.1-2016
Excerpt on Natural Ventilation................. 227
APPENDIX A5.3—Excerpts from ASHRAE Standard 62.1-2019
on Natural Ventilation ............................ 229
APPENDIX A5.4—Minimum Ventilation Requirements
Per ASHRAE Standard 62.1-2019 ............. 245
APPENDIX A5.5—Sample Results from an Engineered
Natural Ventilation System Analysis ........ 249
A5.5.1 New York City, NY, Results........................................ 250
A5.5.2 Reference ................................................................ 254
CHAPTER 6—Demonstrating Comfort Standard Compliance for
Natural Conditioning..............................................255
6.1 Confirming the Requirement to Meet a
Comfort Standard .................................................... 255
6.2 Identifying the Appropriate ASHRAE Standard 55
Comfort Compliance Methods .................................. 256
6.3 Limitations on the Use of the Adaptive Comfort Method... 261
6.4 Applying the Adaptive Comfort Zone Method ............. 263
6.5 Documentation of Compliance with the Standards ...... 275
6.6 Credit Documentation for LEED®
Thermal Comfort Credit............................................ 275
6.7 Rules of Thumb for Managing Expectations around
Comfort Results........................................................ 275
6.8 Cold Day Concerns in Natural Conditioning Schemes..... 279
6.9 References............................................................... 280
APPENDIX A6.1—Sample Results from Natural Conditioning
Analysis Using DesignBuilder ..................... 283
A6.1.1 Office Space............................................................ 283
A6.1.2 Classroom Space ..................................................... 283
A6.1.3 References............................................................... 288
viii | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

APPENDIX A6.2—Sample Results from Natural Conditioning


Analysis Using IESVE .................................. 289
A6.2.1 Office Space ........................................................... 290
A6.2.2 Classroom Space..................................................... 294
A6.2.3 Comparative Comfort .............................................. 295
A6.2.4 References .............................................................. 295
Appendix A.6.3—Engineered Natural Ventilation System
Analysis Sample Results........................... 297
A6.3.1 New York City, NY, Results ....................................... 297
CHAPTER 7—Controlling Natural Ventilation and
Natural Conditioning Systems................................ 307
7.1 Overview of Building Management Controls for
Natural Ventilation Systems....................................... 307
7.2 Sampling of Typical Control Sequences of Operation ......308
7.3 Natural Ventilation Sequences of Operation............... 310
7.4 Natural Ventilation Heating Mode
Enabled/Disabled Sequences of Operation................ 312
7.5 Natural Conditioning Sequences of Operation ........... 313
7.6 Mixed-Mode Changeover Sequence of Operation ...... 315
7.7 Selecting Actuators................................................... 319
7.8 References .............................................................. 324
7.9 Bibliography............................................................ 325
APPENDIX 7.1—Sensor Types Used for
Natural Ventilation Control........................ 327
A7.1.1 Outdoor Sensors Active in Sequences of Operations... 327
A7.1.2 Indoor Sensors Active in Sequences of Operations...... 331
A7.1.3 References .............................................................. 334
CASE STUDIES
CASE STUDY—A.P.C. Store, Hollywood, CA ...................................... 89
CASE STUDY—Fries House, Seattle, WA .......................................... 157
CASE STUDY—ICITY, Moscow, Russia ............................................. 205
CASE STUDY—Wilshire-Gayley, Los Angeles, CA.............................. 305

This guide is accompanied by supplemental appendices, which can be found at


ashrae.org/naturalventilation. These files provide both helpful excerpts from related
standards and results from studies for cities in a variety of climate zones. If the files or
information at the link are not accessible, please contact the publisher.
FOREWORD
The use of natural ventilation to provide fresh air and a comfortable indoor
environment goes back to when humans first constructed shelters for
themselves. We are aware that these ventilation flows can be produced by the
wind blowing through an opening or from cold air entering a warm room.
Consequently, it seems legitimate to ask why a design guide for natural
ventilation is needed—it hardly seems to be rocket science!
However, since the invention of air conditioning at the start of the 20th
century we have become accustomed to controlling our indoor environment
whatever the weather. We heat our buildings in winter and cool them in
summer and maintain comfortable indoor conditions year-round. However,
this comes at a cost, especially in regards to cooling a building. The associated
energy demand is huge, leading to brownouts on hot summer days when, for
example, more energy is used to cool the buildings in Los Angeles, California
than is used by the cars on the roads in that city. The associated greenhouse gas
emissions associated with this energy consumption are large, and the
contribution of the built environment to the current climate emergency is a
matter of huge and immediate concern.
In response to these concerns and also to the perception that conditioned
interior spaces are not, after all, that comfortable—too warm in winter and too
cold in summer—there is a renewed interest in the use of natural ventilation in
modern buildings. Now, however, given our experience with conditioned
interior spaces, we require naturally ventilated buildings to perform to high
standards in terms of indoor air quality. It should supply adequate amounts of
clean air for breathing, aid in the removal of indoor pollutants, and maintain a
comfortable indoor temperature throughout the seasons.
The provision of this high indoor air quality by natural ventilation in the
face of the variability of the external environment is a major challenge to
designers, architects, engineers, facilities personnel, and building owners.
Further, this new emphasis on natural ventilation rather than resorting to the
standard air-conditioning approach in the context of complex modern

ix
x | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

buildings means that these professionals need to acquire new skills in the
design and operation of these buildings. This design guide brings together our
current natural ventilation knowledge to enable designers to make informed
decisions about the potential use of natural ventilation, especially at the early
design stages when its feasibility is being assessed.
This guide provides a comprehensive overview of the issues associated
with the application of natural ventilation in modern buildings. It discusses the
pros and cons of natural ventilation and introduces the main concepts
associated with natural ventilation. It also describes the relevant properties of
the outdoor environment that must be considered, and the responses in terms of
façades, window types, and other openings. Since natural ventilation is more
variable than mechanical ventilation, the important issue of standards
compliance is discussed both in terms of ventilation and conditioning. Finally,
it is a common misconception to equate a naturally ventilated building to a
passive building, while in practice modern naturally ventilated buildings have
openings controlled by a building management system that responds to the
external environment and the interior conditions. This requires unconventional
control mechanisms—it is not possible just to set a thermostat on the wall. The
ramifications of this concept are considered in detail.
This guide is a very welcome addition to the ASHRAE design guides.
Written by two leading experts with many years of experience in the field,
ASHRAE Design Guide for Natural Ventilation clearly and logically lays out
the issues faced when designing a naturally ventilated building. Many worked
examples are provided along with case studies at particular locations that
provide excellent examples of best practices and the issues to consider when
implementing natural ventilation.
I am confident that this design guide will be influential in promoting the
design of energy-efficient, naturally ventilated buildings—for both new
buildings and those undergoing refurbishment. As a result, the building
industry will play its part in mitigating climate change and providing a
sustainable future.

Paul Linden, FRS, FRMetS


Director of Research
G.I. Taylor Professor Emeritus of Fluid Mechanics
Blasker Distinguished Professor Emeritus of Environmental Science and
Engineering
Department of Applied Mathematics and Theoretical Physics,
University of Cambridge, Cambridge, England, U.K.
Department of Mechanical and Aerospace Engineering (MAE),
University of California San Diego, San Diego, CA, USA
PREFACE
With the global initiative to lower energy use and carbon footprint, the idea
of more widely implementing natural ventilation in lieu of heating, ventilation,
and air conditioning (HVAC) is appealing to many design teams. ASHRAE
Standard 62.1, Ventilation for Acceptable Indoor Air Quality and ASHRAE
Standard 55, Thermal Environmental Conditions for Human Occupancy,
establish compliance criteria for minimum natural ventilation and adaptive
comfort, respectively. Design teams must determine whether the climate,
building geometry, and occupancy is well-suited for a natural ventilation
solution and then prove that their design meets the standards. As of 2020, best
practice guidance for natural ventilation is dispersed among many books,
guides, manuals, and standards, and often couched in heavily theoretical
descriptions. This guide provides a reference that outlines a step-by-step
process for determining the feasibility of applying natural ventilation in a
building, whether alone or in a mixed-mode or hybrid ventilation context, to
the natural ventilation design community and their associated partners.

xi
ACKNOWLEDGMENTS
This design guide was produced with an accumulation of experts and
expertise from around the world. From the U.S., we’d like to acknowledge
Stuart Dols of the National Institute of Standards and Technology (NIST),
James Lo of Drexel University, Alexandra Menchaca of Thornton
Thomassetti, Gail Brager from the University of California, Berkeley,
members of ASHRAE Standing Standard Project Committee 55, “Thermal
Environmental Conditions for Human Occupancy,” and Tom Hartman of the
Hartman Company. From the U.K., we had Malcolm Cook of Loughborough
University, Andrew White, and Michael Holmes of Arup. From Canada, we’d
like to acknowledge Duncan Phillips of RWDI; from Australia, we’d like to
acknowledge Richard de Dear of the University of Sydney; and from
Denmark, we’d like to acknowledge Bjarne Olesen, Ongun B. Kanzanci, and
Arsen Melikov of the Danish Technical University.
We would also like to acknowledge individuals and companies that
provided us with figures, including Andrew White of Arup OASYS, Gail
Brager, and Stefano Schiavon of Center for the Built Environment, Nicholas
Peake of the Chartered Institution of Building Services Engineers (CIBSE),
Leon Glicksman of Massachusetts Institute of Technology (MIT), Murray
Milne of the Regents of University of California, Julie Swatson of Lakes
Environmental, Kristi Lovette of CPP, Mark Owen of ASHRAE, Amara
Rozgus of Consulting-Specifying Engineer Magazine, Cole Roberts of
WeatherShift, Daryl Herzmann of Iowa State University, Dan Bacon of
RWDI, Sophie Schlingemann of Intergovernmental Panel on Climate Change
(IPCC), Selena Holmes of U.S. Green Building Council (USGBC), Liam
Buckley of IES, David Cocking from DesignBuilder, Jannick Roth from
WindowMaster, and Andrea Love from Payette.
We would also like to acknowledge the contributions of Arup and its
members for financial, advisory, and material support.
We are, of course, eternally grateful to Mary Bolton, Cindy Michaels,
Mark Owen, and W. Stephen Comstock of ASHRAE Special Publications.

xiii
INTRODUCTION
1.1 PURPOSE OF THE GUIDE
ASHRAE Design Guide for Natural Ventilation assists owners, architects,
engineers, facilities personnel, and building design professionals in exploring
the feasibility of natural ventilation for their project during the early phases of
design. While many in the industry can see the appeal of natural ventilation for
environmental stewardship concerns, some reticence and confusion over the
application and interpretation of current related standards exist. This may be
because that there is currently not an ASHRAE-endorsed methodology that
establishes a rigorous analytical path, and the inputs and equations necessary to
complete even a feasibility analysis are located in disparate standards and
guidance documents. This guide consolidates and organizes the key
information necessary to perform a feasibility analysis by excerpting content
when approved by the originating authors or by reference to information in the
public domain.
The guide presents a logical sequence of activities to support an evidence-
based decision to pursue natural ventilation and/or natural conditioning so that
design teams can ask and answer the right questions at the right time during the
early design process. So often design teams do not fully appreciate the entire
complexity of the code compliance process or the extraordinary reliance on
changing outdoor air conditions when first presenting natural ventilation as a
form of sustainability measure. This guide provides not only the ordering of
content from twenty independent codes, guides, and standards that influence
natural ventilation design, it also identifies for the reader several free or low-
cost tools for accessing weather data or performing early-stage analysis.
Combined with a demonstration of analysis results and their interpretation, the
guide offers a step-by-step road map through the critical decisions that a design
team should make before embarking on a whole building design reliant on
natural ventilation. The progress from conception to fruition for a natural
ventilation scheme is one of collaborative review of critical analysis results by
a design team.

1
2 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 1.1 Organization of the Design Guide


Introductory Evaluating Compliance with Application
Material Inputs Standards Concerns

Chapter 1, Chapter 3, Chapter 5, Chapter 7,


“Introduction” “Outdoor Condition “Navigating Ventilation “Controlling Natural
Assessment” Compliance” Ventilation and Natural
Conditioning Systems”

Chapter 2, Chapter 4, Chapter 6,


“Key Concepts” “Façade and Building “Demonstrating Comfort
Configuration Analysis” Standard Compliance for
Natural Conditioning”

Natural ventilation relies on a variable source of air movement and cooling


capacity. Owners and tenants must be aware not only of the average expected
results but also of the frequency of the extreme conditions for the current day
and in a future climate-changed condition. Natural ventilation as a comfort-
conditioning method will always be compared to HVAC schemes that are
easier to control from a predictability perspective. Thus, the social
acceptability of a natural ventilation scheme is as important to its success as its
technical feasibility. This guide intends to present a logical process of
minimum steps and analyses necessary to facilitate decision-making by key
stakeholders.

1.2 ORGANIZATION OF THE GUIDE


This guide lays out a recommended order of steps in the decision-making
process. Each chapter highlights key issues to discuss with stakeholders and
the appropriate level of analysis to be performed to inform key decisions at
each stage in the decision-making process (See Table 1.1).
Chapter 1 addresses the perceived benefits and risks of using natural
ventilation, a list of the relevant codes and standards related to the design of
these systems, and the key assumptions about the building’s use that should be
examined before pursuing natural ventilation.
Chapter 2 starts with a high-level discussion of terminology and concepts
to ensure that all readers have a common understanding of the physical
principles of natural ventilation. It also includes a discussion on the similarities
and differences between natural ventilation and natural conditioning, as these
systems are often conflated, which can cause confusion when reviewing
acceptability criteria. Essentially, this chapter specifically describes the core
beneficial function of outdoor air in each named mode.
Chapter 3 begins a discussion of how to conduct the deep feasibility
analysis with a multifactor assessment of outdoor conditions. It addresses air
quality, wind analysis, and noise, along with reviews of historical weather data
and future climate-adjusted projections.
CHAPTER 1 | 3

Chapter 4 reviews the key characteristics of the building envelope and their
impact on the success of a naturally ventilated building. It starts with solar
control but also addresses the cold-day phenomenon often overlooked in
discussions of natural ventilation. This chapter also addresses the selection of
opening types and the practical implications of installation and zoning, given
that ASHRAE Standard 55 requires personal control of the opening (ASHRAE
2017a).
Chapter 5 reviews the prescriptive natural ventilation compliance path for
meeting ASHRAE Standard 62.1 (ASHRAE 2019a) and California’s Title 24
(CEC 2018a, 2018b) requirements for minimum ventilation. It also describes
when a design team might want to use an engineered path to comply with
ASHRAE Standard 62.1. This chapter demonstrates how to use a few typical
bulk airflow modeling tools to demonstrate compliance.
Chapter 6 moves the conversation to the question of comfort under a
natural conditioning scheme governed by ASHRAE Standard 55 (2017a). It
describes the use of the adaptive comfort standard, including the elevated
speed option, and compares this model to other comfort indices. This chapter
provides a combined dynamic heat transfer and bulk airflow analysis method
that delivers a simultaneous output of expected indoor temperatures and indoor
ventilation rates.
Chapter 7 explores the control of openings, including the use of motorized
openings and automated feedback based on outdoor air conditions.
As noted in Figure 1.1, the design guide is organized in the order of
milestone gateway decisions that will help design teams to decide whether to
continue investing in exploring the feasibility of natural ventilation. For
example, if a design team cannot confirm that the outside environmental
conditions can support natural ventilation for a reasonable number of hours,
then it makes no sense to engage in building design to support it.
This guide expands on a rule-of-thumb list that originally appeared in
ASHRAE Journal (McConahey 2008). This “Top 10” list is partially replicated
in Table 1.2 with cross-references to sections in this design guide that have
more detailed information about how one might explore the question to be
asked. The authors hope that this design guide will allow design teams to act
with greater confidence in planning their analysis, dialogue, and decision-
making journey toward applying natural ventilation or natural conditioning.

1.3 HOW TO USE THE GUIDE


The best way to use the guide is to envision a sample project and to read
through the steps as if they were going to be applied to one’s own context. The
examples used throughout include office and classroom occupancies, as these
are the room types that are often seen as positive candidates for natural
ventilation.
Appendix A1.1 establishes the analysis parameters for a typical office
space—its façade condition and internal heat loads—as it is used for the
4 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

•Strict filtration requirements

Chapter 1
•Contaminant dilution concerns
•Special pressurization relationships
When Not to •Speech privacy concerns
Pursue Natural •Internal heat load demands
Ventilation

•Outdoor air quality

Chapter 3
•Outdoor noise environment
•Wind climate
Outdoor Air •Historical weather trends
Assessment •Future climate trends

•Building configuration assessment for natural ventilation

Chapter 4
•Openings placement and opening types
•Window glazing selection for thermal and solar performance
Façade and •Window-to-wall ratio for solar control
Building •Shading and blinds in natural ventilation schemes
Configuration

•Prescriptive (ASHRAE Standard 62.1-201 or California Title 24-2016)


•Path A from Addendum L for ASHRAE Standard 62.1-201

Chapter 5
•Path B from Addendum L for ASHRAE Standard 62.1-201 (based on CIBSE [2005])
•Department of the Navy prescriptive wind-driven method
Navigating •Energy modeling considerations from ASHRAE Standard 90.1-201's Appendix G
Natural •Natural Ventilation documentation in U.S. Green Building Council LEED® Rating System
Ventilation
Compliance

•Confirming the requirement to meet a comfort standard


•Identifying the appropriate ASHRAE Standard 55 comfort compliance method

Chapter 6
•Limitations on the use of the adaptive comfort method
•Applying the Adaptive Comfort Zone Method
Demonstrating (i.e., occupant-controlled natural conditioning)
Comfort •Backchecking ASHRAE Standard 62.1201 compliance
Standard •Credit documentation for the U.S. Green Building Council LEED® thermal comfort credit
Compliance

•Overview of building management controls for natural ventilation


Chapter 7

•Natural ventilation sequences of operations


•Natural conditioning sequences of operations
Natural •Mixed-mode changeover sequences of operations
Ventilation and •Selecting actuators
Natural
Conditioning
Control

FIGURE 1.1 Recommended feasibility analyses necessary to pass gateway milestones


when exploring the feasibility of natural ventilation.
CHAPTER 1 | 5

TABLE 1.2 Questions to Ascertain Feasibility of a Natural Ventilation Scheme


(McConahey 2008)
Question to Ask
Data to Review Relevant Section
(If the Answer is Yes, Move to Next Question)
Building Envelope Is the building envelope performance optimized to minimize solar heat Section 4.6, “Selecting Window
gain into the building? Target a maximum total solar load of 4 W/ft2 Glazing for Solar Control
(40W/m2) of sun patch floor area in a cooling condition.

Internal Heat Loads Is the total internal heat load minimized to less than 2 W/ft2 (20 W/m2) Section 1.4.5, “Internal Heat
for naturally conditioned space or within the cooling capacity of auxiliary Load Demands”
systems?

Weather Normals: In looking at the climate data’s monthly mean minimum and mean Section 3.4.2, “Review of Mean
maximum, are there at least six months where the monthly maximum is Maximum and Mean Minimum
Mean Maximum/
less than 80°F (26.6°C) but mean minimum is higher than 32°F (0°C)? Monthly Temperatures”
Mean Minimum

Frequency of In further looking at climate data, does the frequency of occurrence Section 3.4.1, “Natural
Occurrence psychrometric chart for occupied hours have more than 30% of the time Conditioning Potential by U.S.
Psychrometric Chart between 60°F to 80°F (15.5°C to 26.6°C) and less than 70% relative Climate Zone”
humidity?

Ambient Environment Is the surrounding environment suitable for direct intake of air from Section 3.1, “Air Quality”
and Possible Locations outdoors (i.e., there are no security concerns; the ambient environment is
of Openings sufficiently quiet; air quality meets ASHRAE Standard 62.1 requirements;
openings are not near street level, near highways, or industrial plants, or
at elevation of a neighbor’s discharge)?

Window Locations and Can the equivalent of 4% to 5% of the floor area as window opening Section 5.1, “Introduction to
Sizes, Accessibility area be found with direct access to the window by everyone within 20 ft Available Ventilation
(6 m)? Compliance Paths”

Wind Rose, Feasible Can one rely on wind-driven effects for cooling? Section 3.3, “Wind Climate”
Flow Paths: Inlet to
Is there a direct low-pressure airflow path from a low-level opening to a Section 4.2, “Assessing Building
Outlet Under All Wind
high-level opening within the space, and will it be preserved once Configuration for Natural
Conditions
furniture/tenant installation (TI) work is complete? Ventilation,” and Section 4.3,
“Considerations when Placing
Natural Ventilation Openings”

High Afternoon Does the climate have regular outdoor air temperatures over 80°F Section 3.4.4.1, “Climate
Temperatures (26.6°C)? If yes, review whether exposed thermal mass is possible. Consultant”
Section 3.4.5, “Frequency of
Occurrence Analysis by Building
a Spreadsheet”

Diurnal Range on Does the climate have a diurnal range that has nighttime temperatures Section 3.4.5, “Frequency of
Hot Days below 65°F (19.8°C) for at least 8 hours a night on the worst-case days? Occurrence Analysis by Building
a Spreadsheet”
If yes, move to multizone modeling of thermal mass and consider night
purge.

Dew-Point Throughout the year, do you have consistent outdoor air dew-points Section 3.4.5, “Frequency of
Temperatures throughout the year of less than 64°F (16.5°C)? If yes, move to multizone Occurrence Analysis by Building
Throughout Year modeling and consider a radiant cooling system. a Spreadsheet”
Chapter 7, “Natural Ventilation
and Natural Conditioning
Systems Control”
6 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

sample analyses in Chapters 5 and 6. Appendix A1.2 establishes equivalent


parameters for a typical classroom model.
In Chapters 5 and 6, an example of a calculation and its results are
presented with interpretive text to guide the reader through the analysis
process. The space and openings under consideration are part of the
calculations required to assess that natural ventilation will work, provide the
required ventilation air for occupants, and provide occupant comfort. See
Figure 1.1. Sample output results for New York City, NY are provided in
Appendices A5.5 and A6.3. Results for Scottsdale, AZ; Los Angeles, CA;
Naples, FL; and Seattle, WA, which represent a sampling of climate regions,
are available with this design guide at www.ashrae.org/naturalventilation.

1.4 WHEN NOT TO USE NATURAL VENTILATION


Before proceeding further in the guide, it is essential to ensure that there
are no fundamental obstacles to the use of natural ventilation. Five aspects
must be discussed with the owner regarding the indoor environment before
considering any natural ventilation scheme:

• Strict filtration requirements


• Contaminant dilution concerns
• Special pressurization relationships
• Speech privacy concerns
• Internal heat load demands

1.4.1 Strict Filtration Requirements


The use of natural ventilation typically precludes the use of filtration on the
incoming airstream. Therefore, if there are particular health- or process-related
reasons where media-based filtration of any type is required, then natural
ventilation should not be used. The following types of spaces may fall under
this category:

• Hospital or clinic areas using strict filtration or pressurization-based


infection control protocols
• Presence of populations with respiratory-health sensitivities
• Indoor areas requiring gas-phase filtration, such as art museums,
galleries, or conservation workshops where artifacts are not in glass
cabinets
• Areas requiring high-efficiency particulate air (HEPA) or Ultra Low
Particulate Air (ULPA) filtration, such as cleanrooms in electronics,
chip manufacturing, and biotechnology fields
• Areas requiring filtration for food service reasons (See local codes as
they may not allow natural ventilation due to dust and lack of
temperature control for food. Typically, minimum requirements
include installing insect screens and limiting the size of unscreened
openings used for food pass-through.
CHAPTER 1 | 7

TABLE 1.3 Minimum Requirements for Filtration per ASHRAE Standard 62.1-2019
(ASHRAE 2019)
Required
Condition ASHRAE Standard 62.1-2019 Section
Filtration

Upstream of all cooling coils or other devices MERV 8 Section 5.9


with wetted surfaces through which air is
supplied to an occupiable space

Site within area of nonattainment for PM10 MERV 8 Section 6.1.4.1a


particulate matter

Site within area of nonattainment of the PM2.5 MERV 11 Section 6.1.4.2a


particulate matter

Note that if the space is air-conditioned as opposed to naturally ventilated,


minimum requirements for filtration, defined by ASHRAE Standard 62.1
(ASHRAE 2019a), would be automatically required.
In addition, some local codes or voluntary standards have equal or higher
filtration requirement for green and healthy building targets, such as:

• U.S. Green Building Council’s Leadership in Energy and


Environmental Design® Green Building Rating System (USGBC
LEED New Construction [NC])v4 requires compliance with ASHRAE
Standard 62.1-2010 as a prerequisite and a MERV 13 filter to earn the
Enhanced IAQ Strategy point (USGBC 2013).
• The 2019 California Green Building Standards Code requires a
MERV 13 filter (ICC 2019).
• The WELL Building Standard requires a MERV 13 filter in the
ventilation system (IWBI 2016).

Before natural ventilation is pursued, a discussion with the owner early in


the process is necessary to identify any locations that require special filtration
and to set the general tolerance for avoiding filtration.

1.4.2 Contaminant Dilution Concerns


There are two means of diluting the effect of internally generated
contaminants: increasing the flow rate of outdoor air (presuming that it does
not also contain the contaminants of concern) or directly exhausting air from
the space to remove contaminants. Both methods are used in conjunction with
the Ventilation Rate Procedure detailed in ASHRAE Standard 62.1-2019 when
mechanical ventilation is pursued (ASHRAE 2019a).
When a natural ventilation scheme is pursued under ASHRAE
Standard 62.1, there is no stated requirement to match ventilation dilution
rates; however, there is a requirement to provide exhaust in a list of specialized
occupancy types under Table 6.2, “Minimum Exhaust Rates” in ASHRAE
8 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Standard 62.1. (See Website Appendix WA1.1 for an extended discussion of


special exhaust requirements.)
There are situations where indoor contaminants are created through
manufacturing, production, or research processes. Dilution alone does not
provide sufficient protection for the occupants, and specialized exhaust
systems are required for health and safety reasons. Specialized exhaust systems
are required in these spaces to keep the concentrations of the contaminants low
in the breathing zone and below the contaminants’ flammability limits. If dust
or fumes are created by the processes inside the space, it is strongly
recommended that natural ventilation is not used. See Website
Appendix WA1.1 for more guidance.

1.4.3 Special Pressurization Relationships


Discussing the variety of processes or fumes in the space with the owner
early in the process is necessary before a natural ventilation scheme is pursued.
By definition, the use of natural ventilation means that indoor spaces are
subject to fluctuating airflows and variable space-to-space and indoor-to-
outdoor pressure differentials due to temperature and wind effects. In many
types of occupancies, this is acceptable; however, it is a point of conversation
with the owner and tenant stakeholders to determine whether any special
pressurization relationships must be maintained for comfort or safety,
regardless of the ventilation scheme used.
1.4.3.1 Standard Practice for Odor Containment. It is a standard industry
practice that areas designated to require exhaust by Table 6.2 in ASHRAE
Standard 62.1 or other requirements (See Website Appendix WA1.1) should
be kept at a negative pressure with respect to the adjacent areas. A natural
ventilation scheme would have to be able to accommodate this requirement to
avoid backflow from the “dirty” (i.e., high air class) zone into the cleaner ones.
It is important to note that air class limitations over recirculation apply across
all forms of ventilation governed by ASHRAE Standard 62.1, even though
they are listed in Table 6.1, “Minimum Ventilation Rates in Breathing Zone”
under the Ventilation Rate Procedure. It is necessary for the design team to
identify the air class for all of the rooms in the building and to review the
airflow paths created by the natural ventilation pressures under both buoyancy
and wind extremes to determine if there is ever a flow pattern that would cause
high class air to move into a lower-class area and contaminate it. If this is the
case, it is important to discuss with the owner how the high-air class area
should be isolated and independently ventilated with dedicated exhausts.
1.4.3.2 Building Pressurization for Envelope Protection and Reduction of
Outdoor Contaminants. There are additional circumstances in
which the owner is concerned for the environmental health and safety of the
building’s occupants arising from external effects such as infiltration of
airborne contaminants.
Many buildings are intentionally overpressurized to ensure the outflow of
conditioned and filtered air. This tends to keep the fabric of the building
CHAPTER 1 | 9

envelope dry to avoid mold. If a mixed-mode or hybrid ventilation system is


being considered, attention to the detailing of seals at natural ventilation
openings and their frames is necessary to avoid a point of failure for moisture
transmission into the building fabric, which could then lead to mold.
Similarly, the practice of whole building pressurization also prevents the
intrusion of outdoor contaminants such as allergens, dust, smoke particles, and
other contaminants of concern. This is obviously not possible in a naturally
ventilated building. Thus, before pursuing natural ventilation, it is necessary to
discuss with the owner whether the building is likely to house sensitive parties
who would suffer from respiratory distress from exposure to typical outdoor
contaminants. If this is the case, there may be ways to zone the building to
accommodate certain areas as fully mechanically ventilated with filtration.
A special form of exposure involves those facilities that may be at credible
risk of attack using an airborne agent. The following list, excerpted from
ASHRAE Handbook—HVAC Applications (2019b), shows example project
types where the likelihood of airborne attack should be considered during a
discussion with the owner prior to pursuing a natural ventilation scheme.
(Italicized items in this list were added by authors.)

• Military and government command centers


• Significant landmarks
• Single-failure point operations or equipment
• Corporate headquarters or critical operations centers
• Transportation hubs, including airports
• Communications nodes
• Popular tourist attractions
• Sites open to the public
• Sites frequently targeted by protests and demonstrations
• Locations near significant potential hazards such as nuclear power
plants or chemical manufacturing facilities
• Federally operated facilities (domestic and international)
• Hospitals providing emergency care

These facilities often have specialty HVAC systems in place that are not
compatible with the increased risk of outdoor air exposure from a natural
ventilation scheme. The design team should confer directly with the owner’s
environmental health and safety and security teams to assess vulnerability and
risk.

1.4.4 Speech Privacy Concerns


The typical acoustic complaint in naturally ventilated spaces is not that
there is too much noise from outdoors but that there is noise from indoor-
adjacent spaces and a lack of masking noise from the HVAC system. If speech
privacy or noise containment within a single zone is a functional requirement
of the space, natural ventilation may not be the appropriate choice. Refer to
Chapter 3 for a full discussion of noise break-in from outdoors.
10 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

One key design requirement of most naturally ventilated spaces is that


there is an open airflow path from outdoor intake to outdoor exhaust. This is
not a substantial concern when single-sided ventilation in an enclosed room is
used, because the noise generated within the room remains within the space.
When cross-ventilation across multiple zones is used, horizontally adjacent
spaces are typically open to one another for bulk air transfer, and
consequentially, the noise transfer that comes with it. Furthermore, many
naturally conditioned spaces use exposed thermal mass to absorb heat in the
afternoon, releasing it in a nighttime purge mode in the evenings. This
approach serves to reduce the available amount of noise-absorptive material in
the room finishes, leading to reverberant “live” spaces that bounce sound
waves throughout the space. In a reverberant space, people can be disturbed by
the conversations of others who are not in their line of sight. Besides the
irritation of disruption to their concentration, this phenomenon also tends to
lead occupants to be concerned about their own speech privacy.
The second issue compounds the speech privacy concerns. In typical
HVAC installations, high-velocity air passes against the metal edges of a
diffuser, creating a hissing noise that serves as diffuse background noise
throughout the space that reduces speech intelligibility. In a natural ventilation
scheme, this diffuser noise is typically missing. People often complain of the
space being too quiet and are thus highly sensitized to people overhearing their
conversations. Therefore, it is a common solution to introduce an active white
noise sound-masking system to address these speech privacy concerns.
Discussing the acoustic performance of the space with the owner,
especially in regards to speech privacy, is key.

1.4.5 Internal Heat Load Demands


In a natural ventilation scheme, the outdoor air conditions are often so
close to indoor air conditions that the heat absorption capacity of the air is
severely limited. Therefore, controlling internal head loads is essential to the
success of a natural conditioning scheme. Table 1.3 shows the limitations on
absorbed cooling, as cross-referenced to air change rate, for a 5°F (3°C)
temperature difference between indoor and outdoor air. Refer to the more
detailed development of the calculations in Chapter 2.
A 100 ft2 (9.29 m2) office space at a maximum of 10 ACH capacity would
be limited to the components in Table 1.4. With light-emitting diode (LED)
lighting typically around 0.5 W/ft2 to 0.6 W/ft2 (5 W/m2 to 6 W/m2), solar
control becomes essential. See Chapter 4 for more information on the loads
contributed from solar heat gain.
This short exercise further demonstrates how critical it is to design the
selection and arrangements of openings to ensure increased air-change rates.
(See Chapters 4 through 7 for more details.) Had the same exercise been done
with only 2.5 ach in the space under the same outdoor air conditions, there
would not have been enough cooling capacity (177 Btu/h [52 W]) to cover
even the heat load of a single person (245 Btu/h [70 W]). It is therefore
CHAPTER 1 | 11

TABLE 1.4 Sensible Cooling Capacity per Degree of Indoor to Outdoor Temperature
Difference in Short Height Spaces for Natural Conditioning
Sensible Cooling Capacity Slope Sensible Cooling Capacity at
Air Change Rate, Assuming a 5.4 R (3 K) in a
9 ft (2.75 m) High Space, Developed*,
ach 100 ft2 (9.29 m2) Space,
Btu/h/ft2/R (W/m2/K) Btu/h (W)

2.5 0.328 (1.87) 177 (52)

5 0.700 (3.98) 378 (111)

10 1.40 (7.93) 756 (221)


* Slope as approximated from Figure 4.3 in CIBSE (2005)

necessary for the design team to assess reasonable estimates of internal heat
gains and to keep those as low as possible.

1.5 PERCEIVED BENEFITS AND RISKS OF


NATURAL VENTILATION
Often, natural ventilation or natural conditioning is proposed as an
alternative to conventional heating, ventilating, and air-conditioning (HVAC)
systems. Many concerns must be considered in the application of natural
ventilation to commercial and institutional buildings. A good layperson
summary of the typically identified obstacles to natural ventilation can be
found in Melton (2014).
Persuasive arguments have been made for the appropriate relaxation of
comfort standards for the individual results in energy savings which, if applied
at scale, could contribute to climate change mitigation in the state. The
executive summary of a study funded by the California Energy Commission
(CEC) on energy savings from retrofitting buildings toward natural ventilation
is a classic example of this significant extrapolation: “The new knowledge and
tools will significantly increase the use of natural ventilation in commercial
buildings in California, which will, in turn, provide energy savings and reduce
greenhouse gas emissions” (Linden et al. 2015).
While social consciousness and reputational benefits may factor into their
decision to pursue natural ventilation, clients will typically need evidence
specific to the scale of their own projects to make a credible argument for a
tangible benefit before invoking that support. They deserve to understand the
benefits and risks of natural ventilation as an asset-embedded investment, so
they can judge its efficacy against other priorities for funding.
One of the issues with jumping directly to the energy-versus-comfort
argument is that it typically does not match benefit and risk to the same
stakeholder. There are arguments within the definition of comfort itself that
must be resolved prior to addressing energy concerns. When the issues are
12 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 1.5 Build-Up of Heat Load Density for 100 ft2 (9.29 m2) Office Space with
10 ach at 5°F (3°C) Temperature Differential to Outdoor Temperature
Description Reference Heat Load, I-P (SI)
Total available heat absorption capacity, Table 1.4, “Sensible Cooling 756 Btu/h (221 W)
based on 10 ach limit for 9 ft (2.5 m) high Capacity per Degree of Indoor to
space at 5.4°F (2.5 K) temperature differential Outdoor Temperature Difference in
(room—outdoors) Short Height Spaces for Natural
Conditioning”

One person: seated very light work, high- Table 1, “Representative Rates at 245 Btu/h (70 W)
velocity air movement Which Heat and Moisture Are Given
Off by Human Beings in Different
States of Activity” in ASHRAE
Handbook—Fundamentals (2017b)

One laptop with docking station, average 15 Table 8A, “Recommended Heat Gain 61 W × 3.41 Btu/h/W =
min peak power consumption for Typical Desktop Computers” in 208 Btu/h (61 W)
ASHRAE Handbook—Fundamentals
(2017b)

Differential available for lighting and solar — 303 Btu/h (90 W)


30 Btu/h/ft2 (9.7 W/m2)
0.9 W/ft2 (9 W/m2)

TABLE 1.6 Example Scorecard to Summarize Viability of Risk-Benefit Analysis for


Pursuing Natural Ventilation as an Asset-Bound Investment
Convinced Manageable
Topic Perceived Benefit Perceived Risk
of Benefit? Risk?
Indoor conditions Environmental quality  Comfort 
Occupant discretion Device control  Situational constraint 
Cost of ownership Market value  Operational risk 

adjusted to match scale and target audience, three classic dilemmas typically
frame the start of the conversations:

• Indoor conditions: Environmental quality versus comfort—individual


experience
• Occupant discretion: Device control versus situational constraint—
individual experience
• Cost of ownership: Market value versus operational risk—owner
experience

The design team’s goal is to address the likelihood and extent of risk and to
demonstrate the benefits of the particular application. Within each discussion,
the client can test their capability and tolerance level to manage the risk against
CHAPTER 1 | 13

the attractiveness of the noted benefits. Each comparison can be evaluated


independently for viability, and then if not all priorities are shown to be
reasonably viable, the client can clearly see the choice at hand.

1.5.1 Indoor Conditions: Air Quality versus


Occupant Comfort
The first concern that is typically raised is the trade-off between
environmental quality and comfort. Because the goal of this design guide is to
help design teams predict the viability of natural ventilation in a specific
location based on discussion with a well-informed client, this section offers
only a handful of pertinent findings from the academic literature that help to
frame the most common of the concepts and arguments that arise when the
debate over natural ventilation ensues. For a full literature review on indoor
environmental quality and thermal comfort, refer to Al horr et al. (2016) and
Rupp et al. (2015).
There is a significant amount of academic research—much of it using field
studies—trying to better define the acceptability of indoor conditions in
naturally ventilated spaces. The challenges of in-situ real-time data gathering
have limited many natural ventilation field studies to relatively few test sites
and relatively small test subject sample sizes. Test protocols vary in regards to
what information is collected during field observations, so it is not always
possible to compare research studies to each other, nor to have enough
comprehensive data to compare different aspects of indoor environmental
quality to each other at the time of testing. Moreover, the local weather and
cultural climate will often have a significant influence on both the
physiological and social acceptability of natural ventilation in commercial and
institutional buildings.
Arguments for improved indoor environmental quality revolve around the
fact that natural ventilation tends to increase ventilation rates during warmer
months or high indoor heat loads as compared to enclosed mechanically
ventilated buildings, although this must be proven through simulation for each
unique circumstance related to room layouts and window configurations. This
statement does not necessarily hold true during cold weather when windows
would be closed by the occupants. Similarly, the benefit of increased air
movement is not necessarily guaranteed in a naturally ventilated space—as
evidenced by the relatively low air movement recorded at the naturally
ventilated Berkeley Civic Center in Zhang et al. (2007)—and the relatively
high number of occupants desiring more air movement when they are at
neutral or warmer air temperatures. This finding led to a hypothesis from de
Dear (2010) that turbulence from increased air movement is welcomed in
warm environments due to the relative depth of cold versus warm-sensing
thermoreceptors in the skin. The fact that the same air speed can be
experienced as negative when one is cold and positive when one is warm, is an
example of alliesthesia, which is “the phenomenon whereby a given stimulus
14 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

can induce either a pleasant or unpleasant sensation, depending on the


subject’s internal state” (de Dear 2010).
The environmental argument goes further to say that naturally ventilated
spaces then have the resultant reduction in indoor CO2 concentrations. This
was shown to be the case for offices (Almeida et al. 2017) but not for
classrooms (Mendell et al. 2013), so again zonal contaminant production and
dispersion analysis for the particular application is necessary. It is important to
note that CO2 concentrations are typically a proxy for human occupancy and
the acceptable dilution levels for associated body-related odors (ASHRAE
2017a), and do not reflect contaminants that may enter from the outdoors.
The benefits of increased outdoor air quantities often point to evidence of a
measured reduction in incidences of sick building syndrome in offices, as
noted in Seppänen and Fisk (2001), which concludes that there is a
“statistically significant increase in symptom prevalences in air-conditioned
buildings, with or without humidification, relative to symptom prevalences in
naturally-ventilated buildings” (Sepannen and Fisk 2001). This is countered by
other simulated findings (Dutton et al. 2013) that argue that exposure to
outdoor concentrations of ozone and particulate matter PM2.5 in naturally
ventilated buildings can cause adverse, costly health outcomes for a limited
number of sensitive individuals, outweighing the benefits of the savings
accrued from reduced sick building syndrome. The disparate research findings
point to the importance of understanding the local outdoor air quality (see
Chapter 3) since it will determine whether the indoor air quality can be
expected to meet a more stringent level of acceptable contamination. Reduced
respiratory comfort for the sensitive receivers is due to lack of filtration, so if
outdoor air quality is poorer than the U.S. Department of Labor’s Occupational
Safety and Health Administration’s (OSHA) exposure standards, it may be
beneficial to use a mixed-mode approach to at least provide a zone of filtered
air provision for sensitive tenants.
Some also argue that natural ventilation creates a better acoustic
environment to support concentration because HVAC noise is eliminated. This
is typically offset by a greater risk of outdoor noise break-in and the loss of
sound masking from the HVAC noise that leads to increased perceptions of
speech privacy. (See Chapter 3 for more information on this topic.) The
ambient noise of the particular site and anticipated indoor activities should
inform the client as to whether there is an appreciable benefit or risk to going
with natural ventilation.
The perception of reduced thermal comfort typically identifies significant
variability in air change rates, air velocities, indoor relative humidity, and
indoor temperature as a cause of discomfort, particularly around excessively
high indoor temperatures during hot periods. The core concern is reduced
productivity arising from reduced workplace comfort or less effective learning
outcomes in students. As noted previously in this section, regular
unpredictable variability of air movement is not necessarily detrimental to
comfort if alliesthesia balances disruption with delight. Variability in air
temperature and humidity occurs on a much slower and far more predictable
CHAPTER 1 | 15

basis, both throughout the day and seasonally. The adaptive comfort model is
well-suited to provide the countervailing argument related to an expanded
thermal comfort definition in naturally ventilated spaces.
People are not uncomfortable at the same threshold temperatures in
naturally ventilated spaces as they would be in air-conditioned facilities.
Leaman and Bordass (2007) have noted from field study results comparing
naturally ventilated buildings to mixed-mode and air-conditioned buildings
that, “forgiveness scores are higher for buildings with natural ventilation
incorporated in some way[...] people may be more likely to tolerate otherwise
excessively uncomfortable conditions in buildings with natural ventilation.”
Deuble and de Dear (2010) further measure the forgiveness factor in two green
academic office buildings on the same campus, identifying that forgiveness is
higher in a naturally ventilated building as compared to its mixed-mode
neighbor (which ranked equivalent to a standard air-conditioned building).
Brager and de Dear (2000) have articulated the ASHRAE Standard 55 adaptive
comfort model, discussed at length in this design guide, based on multiple field
studies. Hellwig et al. (2006) reconfirmed that the actual perception of
occupants in naturally ventilated buildings aligns with ASHRAE Standard 55
(ASHRAE 2017a) based on German office building field study test results.
Occupants adjust their expectations away from thermal stability toward
accepting a “floating” comfort temperature linked to their memory of the
recent outdoor conditions. This is the basis of the adaptive comfort standard
described in more detail in Chapter 6.
With the understanding of the broader aspects of the dilemma, design
teams can conduct outdoor air quality analysis (see Chapter 3) and local site
testing as a first prediction of the best-case condition of the indoor air quality
of a naturally ventilated space. More detailed analysis using bulk airflow
modeling tools or computational fluid dynamics (CFD) can model both the
production of indoor pollutants and their dilution given the air change rates
created by natural ventilation. Further analysis using dynamic heat transfer and
bulk airflow modeling tools can be integrated to also model internal heat
sources and solar inputs to simultaneously estimate indoor air and mean
radiant temperatures. (See Chapter 6.) The results of these types of analyses
can provide important predictive evidence to help a client decide whether the
risks of comfort variability for prospective tenants can be offset by the
attractiveness of improved environmental quality.

1.5.2 Occupant Discretion: Device Control versus


Situational Constraint
A key aspect of the adaptive comfort model is occupant control of the
opening along with their ability to adjust their own clothing in response to
indoor conditions. The client must understand the need for this discretionary
behavior to be available to the tenants for a natural ventilation scheme to prove
successful over the life of the building. With the caveats raised in the preceding
discussion, this section presents a handful of key research studies that speak to
the question of personal control in naturally ventilated spaces. As discussed
16 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Environmental
Comfort
Quality

Reduced Absenteeism
Increased exposure to
from Sick Building
allergens and particulates
Syndrome

Increased noise break-in


Quiet environment without
from outdoors,
HVAC noise
less speech privacy

Reduced comfort from


Reduced indoor high temperature and
CO2 levels yearlong temperature
variability
Variability of air
Increased air speed and movement, indoor air
ventilation rates temperature and relative
humidity

FIGURE 1.2 Balancing indoor conditions: Environmental quality versus comfort.

previously, these studies are presented to frame the sides of the typical
conversations that arise when natural ventilation is considered. For a literature
review of the topic of window control, refer to Ackerly et al. (2011) for
window use in mixed-mode buildings. Roetzel et al. (2010) also provide a
literature review related to occupant control of natural ventilation, which more
broadly addresses façade and window configuration and weather-related
drivers.
Brager et al. (2004) demonstrated that occupants with control over an
operable window (by virtue of proximity) unconsciously adjusted their neutral
temperature upwards by almost 2.7°F (1.5°C) as compared to their
counterparts deeper into a naturally ventilated building.
De Carvalho et al. (2013) propose, through regression analysis, that
university students in naturally ventilated spaces chose their clothing weight
consistent with their thermal memory of the near term past and adjust their
expectations around the comfort temperature accordingly. This is seen through
de Carvalho et al.’s following equations (2013):

T out.ref = 0.187  maxT day.x + 0.813  T day.x-1 (1.1)

T comf = 18.8 + 0.30  T out.ref (1.2)


CHAPTER 1 | 17

where Tout.ref is the outdoor reference temperature, Tcomf is the comfort


operative temperature, max Tday.x is the temperature registered for a day, and
Tday,x–1 is the previous day’s average outdoor temperature.
Since satisfaction is related to the minimized difference between
anticipated comfort temperature and actual space temperature, in naturally
ventilated spaces occupants must have full control over nearby windows as
well as their clothing adjustment.
Discussions with the client need to address the likelihood of any situational
constraints on this occupant discretionary adjustment behavior. Constraints
would include things like seating assignments based on hierarchy, the
perceived social value of air conditioning, cultural expectations around what
clothing is appropriate, and the permissibility to modify clothing throughout
the day. Morgan and de Dear (2003) showed that clothing insulation on “casual
Fridays” tracked more adaptively with outdoor seasonal conditions as
compared to the clo value recorded for the professional business attire worn on
the other weekdays.
Hellwig et al. (2006) identify that 77% of occupants in naturally ventilated
spaces are satisfied with indoor temperature, and 87% feel that they have
control over the indoor temperature and air movement. This is as compared to
50% and 36%, respectively, for air-conditioned buildings. Hummelgaard et
al. (2005) showed that, despite a higher indoor temperature and higher CO2
values, occupants of naturally ventilated spaces stated greater satisfaction than
the occupants of air-conditioned spaces. Khatami et al. (2014) showed that
thermally dissatisfied occupants took advantage of manual window controls to
alleviate their situation, whereas previously thermally satisfied occupants did
little to engage with the operable windows. It also demonstrated that automatic
controls may be necessary to achieve indoor air quality on mild days when
thermal comfort concerns do not drive occupants to interact with the windows.
(See Chapter 7 for more information on automated controls.)
To be successful at tapping into the subconscious and/or physiological
adaptation of thermal neutrality, occupants must have personal direct access to
an opening device, and it must be intuitively simple enough for them to
modulate it to achieve their own comfort. Ackerly et al. (2011) quote advice
from Bordass et al. (2007) regarding guidance on the design of manual
controls:

• Clarity of purpose
• Intuitive switching
• Labeling/annotation
• Ease of use
• Indication of system response
• Degree of fine control

Ackerly and Brager (2013) also discuss the types and effectiveness of
window signaling systems that have been used to prompt users to open or close
18 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

the windows based on outdoor conditions. D’Oca and Hong (2014) performed
data-mining to discern:

two typical window opening office user profiles, one mainly physical
environmental driven and one mainly contextual driven. The results
indicated that office users interact with windows principally driven by
thermal discomfort (indoor air temperature) but also behave
accordingly to daily routine (time of the day) and habits (arriving and
leaving time). (D’Oca and Hong 2014)

Furthermore, they found that:

users mainly driven by physical environmental parameters had less


impact on natural ventilation than users driven by contextual factors
and habits, opening windows for shorter periods of time, interacting
less frequently and usually preferring smaller openings. (D’Oca and
Hong 2014)

As design teams develop natural ventilation proposals, they should engage


the client in a discussion about any likely workplace cultural and situational
constraints on the control of windows or clothing. It is important to explore the
social environment around the question of who would have control of the
windows and how many people would be affected by a single person’s
preferences. If the client feels that there will be equitable access to the natural
ventilation opening, then the issue becomes one of device control design.
When designing the façade and any operable elements, the design team should
test the intuitiveness of operation and the transparency of control with people
outside the design team who may be less familiar with the application of
natural ventilation in modern commercial or institutional buildings. Mockups
of how the manually operable devices are accessed around desks, partitions, or
other furniture can show the ergonomics of access and whether there are any
physical impediments to engaging with the device.

1.5.3 Cost of Ownership


The cost of ownership is unique to a locality and its relative influence of
energy costs and maintenance costs versus construction costs. This section
discusses the key concerns to balance for the client who holds the capital asset.
The question of first cost is the easiest to address. In pure natural
ventilation schemes, the design team can typically expect a reduced first cost
for HVAC equipment. However, there is often an increased cost to the façade
for incorporating operable elements. Similarly, if hybrid ventilation (mixed-
mode) systems or backup ventilation or conditioning systems per ASHRAE
Standard 62.1 (2019a) and ASHRAE Standard 55 (2017a) are applied, then
first-cost savings may be completed negated. It is important during the early
design phases to quickly assess the first-cost impact of applying a natural
ventilation scheme to understand the true construction cost so that all other
CHAPTER 1 | 19

Device Situational
Control Context

Increased occupant Cultural expectations of air


satisfaction in naturally conditioning as amenity for
ventilated buildings professionals

Personal access and Intuitive Permissibilty of occupant to


Control of Windows adjusting clothing

Adjustment of comfort
Equitable access to operable
temperature due to
windows
Operable Windows

Workplace hierarchies that


Benefits of Perimeter Seating/
assign window seating based
Daylight Access
on status

FIGURE 1.3 Balancing occupant discretion: device control versus situational constraint.

benefit/risk considerations can be weighed in light of this investment or


potential savings.
Benefits to the client over the life of the project include potential
reductions in annual energy and equipment maintenance costs. If energy and
maintenance savings are passed onto tenants in green or triple net leasing
arrangements, this can also improve market value for potential lessees.
Similarly, natural ventilation can often be a sales differentiator by delivering
user-controlled fresh air access in markets typically dominated by sealed
environments. Additionally, a building’s overall reputation for environmental
stewardship may be attractive to tenants who have their own sustainability
goals or policies. If the leasing market recognizes and values these elements,
natural ventilation may help improve overall property value in the event the
client wishes to sell the asset after occupancy is achieved. The client will likely
have the best insight into the market’s appetite for these issues.
The risks and uncertainties that the owner may need to manage have more
to do with operations than initial construction cost. Because windows are open
the majority of the time, naturally ventilated buildings do not generally use
positive pressurization for envelope protection or to resist infiltration in winter
months. Coupled with natural ventilation’s low-pressure differentials, which
cannot support filtration, there may be increased janitorial costs as compared to
a sealed building with central air conditioning. Similarly, if occupants leave
windows open overnight, there may be increased security costs to monitor the
20 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Market Operational
Value Risk

Improved reputation for Stability of long term value,


environmental stewardship in light of climate change

Improved resilience for Unpredictability of daily


operation during extreme outdoor conditions—first
hot day events cost of backup systems

Direct access to fresh air as Tenants misunderstanding


differentiating feature building operations

Decreased annual energy


Increased janitorial and
and equipment
security operating costs
maintenance costs

FIGURE 1.4 Balancing cost of ownership: market value versus operational risks.

perimeter or close the windows. Operations staff would be best placed to


provide a cost per square foot estimate of energy, janitorial, and security costs
of comparable sealed buildings and to advise on the differential between
savings and increased expenditure for the second two elements. Combined
with estimated energy costs, an estimate of the annual total cost of operations
can be reviewed against the increased lease value arising from the market
value estimates to determine a total life-cycle cost analysis or return on
investment (ROI). However, there is also some risk of increased energy usage
if occupants do not understand how the building is supposed to work and how
they are supposed to work with it. Schakib-Ekbatan et al. (2015) demonstrate
that occupants in a German naturally ventilated building leave their windows
open too long in the winter for 10% to 25% of the time and operate the
windows counter to the original energy concept for 10% to 40% of the time in
the summer, particularly when outdoor air temperatures are higher than indoor
temperatures.
Lastly, there are concerns on the part of some owners regarding the long-
term stability of asset value in light of climate change. These concerns revolve
around the increasing atmospheric temperatures that come with lowered heat
absorption capacity on hot days, greater variability of outdoor conditions in
extreme weather events, and poor outdoor air quality seen in more frequent fire
CHAPTER 1 | 21

events in drought-stricken areas. Chapter 3 includes some information about


how the design team might estimate natural ventilation potential in future years
by applying some standard scientific climate shifting estimates to provide
information about this concern.

1.5.4 Weighing the Benefits and Risks of


Natural Ventilation
Once the client has agreed that indoor conditions under a natural
ventilation scheme would be acceptable, that the target tenants would have
enough personal freedom to adjust devices and clothing, and that the cost of
ownership is manageable, then the conversation of comfort versus energy use
can be further explored in greater detail beyond the risk-benefit analysis
scorecard shown in Table 1.6.
As noted in the introduction to this benefit versus risk discussion (Section 1.5),
the CEC funded a study to determine the energy savings benefits of applying
natural ventilation in California. Linden et al. (2015) performed a comprehensive
study of the “benefits and barriers to retrofitting California commercial buildings
with natural or mixed-mode ventilation for cooling.” This study found that energy
savings for retrofitting natural ventilation in conjunction with aggressive
conventional energy savings (such as power and lighting energy use reduction,
ceiling fans, external shades, increased insulation, and supply air temperature reset)
could reduce cooling energy by 44% in pre-2008 buildings, but acknowledged that
only 25% of that is due to the natural ventilation itself after the other measures are
applied (Linden et al. 2015). When examined by the California climate zone, the
aggressive measures and natural ventilation packages average about 5% to 6%
ROI for large and medium offices but struggle to reach 4% for most small offices
(Linden et al. 2015). Chen et al. (2017) have a similar simulation-based exercise
across four U.S. climate zones for small and medium offices, estimating the
potential for a 17% mean energy savings in Atlanta, GA, 19.7% in Seattle, WA,
43.8% in San Francisco, CA, and 42.9% in Los Angeles, CA. (See Chapter 3 for
more discussion of climate suitability for natural ventilation.)
While both of these studies normalize for compliance with either the
adaptive comfort or predicted percentage dissatisfied and predicted mean vote
(PPD-PMV) comfort standard, Brager and Borgeson (2010) explicitly provide
a comparison using a simulation of energy cost, adaptive and PMV comfort for
a well-performing mixed-mode building placed in a handful of California
climate zones to explore an annualized understanding of comfort range
exceedances as an important parallel factor when evaluating the potential
impact on buildings’ energy use. This study is unique in that it simulates and
reports comfort simultaneously with energy savings for each of the studied
climate zones. This direct comparative reporting from simulation models can
be applied at the single-building level to provide credible trade-off information
to help with decision making on the energy-versus-comfort question after
many of the other risks have already been deemed manageable.
22 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 1.7 Codes, References, Guides, and Standards that


Govern Natural Ventilation Design

Key Chapters/Clauses Pertaining to Natural


Standard/Manual Name
Ventilation and Natural Conditioning

ANSI/ASHRAE Standard 62.1-2019, Section 6.4 with backup mechanical ventilation systems design
Ventilation for Acceptable Indoor Air Quality per Section 6.2 and/or 6.3.
(ASHRAE 2019a)

ANSI/ASHRAE Standard 55-2017, Section 5.4


Thermal Environmental Conditions for Human
Occupancy
(ASHRAE 2017a)

2019 Building Energy Efficiency Standards for Section 120.1.(c).2


Residential and Nonresidential Buildings for the
(See Website Appendix WA5.2, Section WA5.2.2.)
2019 Building Energy Efficiency Standards: Title
24, Part 6
(CEC 2018a)

2018 International Mechanical Code Section 402


(ICC 2017)

2019 California Mechanical Code, California Section 402.2, essentially replicates ASHRAE Standard 62.1
Code of Regulations Title 24, Part 4 based on
(See Website Appendix WA5.2, Section WA5.2.1.)
2018 Uniform Mechanical Code®
(CBSC 2019)

LEED Reference Guide for Building Design and Minimum Energy Performance prerequisite (p. 357)
Construction, v4 Optimize Energy credit (relies on calculation in Minimum
Energy Performance prerequisite)
(USGBC 2013)
Minimum Indoor Air Quality Performance prerequisite
(pp. 605–22)
Enhanced Indoor Air Quality Strategies credit (pp.645–56)
Thermal Comfort credit (p. 695–710)
(Note that this standard references ASHRAE Standard 62.1-
2010 and ASHRAE Standard 90.1-2010)

CIBSE Applications Manual AM10: Entire document


Natural Ventilation in Non-Domestic Buildings
(CIBSE 2005)

CIBSE Applications Manual AM13:2000: Document addresses mixed-mode or hybrid ventilation, within
Mixed Mode Ventilation which natural ventilation is discussed as one of the modes
(CISBE 2000) available
CHAPTER 1 | 23

1.5.5 Communicating with Occupants about


Natural Ventilation
As clarified in literature noted previously, occupants are essential
participants in their own comfort in naturally ventilated and mixed-mode
spaces. Participation is at the active physical level of opening and closing
windows (D’Oca and Hong 2014) or adjustment of clothing (Morgan and de
Dear 2003), but also at the psycho-physiological level in the unconscious
resetting of neutral temperature (Brager et al. 2004) and increased “forgiveness
factor” as evidenced in self-reported comfort at temperatures in naturally
ventilated buildings that exceed the traditional PMV comfort range (Leaman
and Bordass 2007, Deuble and de Dear 2010).
Martin and Fitzsimmons (2000) argue that occupants who have previously
worked in a mechanically ventilated space may need training “as they are less
likely to make effective use of the vents and any adjustable solar shading.”
They recommend “a written ‘users guide’” that contains the following
elements (Martin and Fitzsimmons 2000):

• Briefing from the architect/building services designer regarding the


intended operation of the building
• The use of the control system, the blinds, and windows, operation of
night cooling, the use of natural and artificial lighting, operation of
thermostatic control valves
• The responsibilities of occupants in opening windows and trickle vents,
shutting blinds, turning off lights, along with the benefits of these
actions
• Details of the expected temperature conditions throughout the year

Preparing all incoming occupants to understand the building operations


and their role in optimizing energy and comfort performance is a crucial part of
the success of naturally ventilated and mixed-mode buildings.

1.6 APPLICABLE CODES, REFERENCES, AND STANDARDS


Table 1.8 provides a summary of applicable codes, standards, and
references used throughout this guide that governs the design of natural
ventilation schemes in building projects. The table also includes the section or
chapter within each reference document that is most applicable to natural
ventilation and natural conditioning.

1.7 REFERENCES
Ackerly, K., L. Baker, and G. Brager. 2011. Window use in mixed-mode
buildings: A literature review. Summary Report April 2011. Berkeley, CA:
Center for the Built Environment.
24 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Ackerly, K., and G. Brager. 2013. Window signalling systems: Control


strategies and occupant behavior. Building Research & Information 41(3):
342–60.
Al horr, Y., M. Arif, M. Katafygiotou, A. Mazroei, A. Kaushik, and E.
Elsarrag. 2016. Impact of indoor environmental quality on occupant well-
being and comfort: A review of the literature. International Journal of
Sutatainable Built Environment 5(1):1–11.
Almeida, R., P. Pinho, M. Pinto, and L. Lemos. 2017. Natural ventilation and
indoor air quality in educational buildings: experimental assessment and
improvement strategies. Energy Efficiency 10:839–54.
ASHRAE. 2017a. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2017b. Chapter 18, Nonresidential cooling and heating load
calculations. In ASHRAE handbook—Fundamentals. Peachtree Corners,
GA: ASHRAE.
ASHRAE. 2019a. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2019b. Chapter 61, HVAC security. In ASHRAE handbook—
Applications. Peachtree Corners, GA: ASHRAE.
Bordass, B., A. Leaman, and R. Bunn. 2007. Controls for end users: A guide
for good design and implementation (BCIA1/2007). Bracknell, U.K.:
BSRIA Ltd.
Brager, G., and R. de Dear. 2000. A standard for natural ventilation. ASHRAE
Journal 42(10): 21–23, 25.
Brager, G., and S. Borgeson. 2010. Comfort standards and variations in
exceedance for mixed-mode buildings. Building Research & Information
39(2):118–33.
Brager, G., G. Paliaga, and R. de Dear. 2004. Operable windows, personal
control, and occupant comfort. ASHRAE Transactions 110(2).
CBSC. 2019. 2019 California mechanical code California code of regulations,
Title 24, Part 4, California building standards commission, Based on 2018
Uniform Mechanical Code®. Sacramento, CA: California Building
Standards Commission and Ontario, CA: International Association of
Plumbing and Mechanical Officials.
CEC. 2018a. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
CEC. 2018b. 2019 Nonresidential Compliance Manual for the 2019 Building
Energy Efficiency Standards: Title 24, Part 6 and Associated
Administrative Regulations in Part 1. CEC-400-2018-018-CMF.
Sacramento, CA: California Energy Commission.
Chen, J., G. Augenbroe, and X. Song. 2017. Hybrid ventilation potential
investigation for small and medium office buildings in different us climates
CHAPTER 1 | 25

under uncertainties. Conference paper of the 34th International Symposium


on Automation and Robotics in Construction (ISARC 2017).
CIBSE. 2000. CIBSE Applications Manual AM13: 2000, Mixed mode
ventilation. London: The Chartered Institute of Building Services
Engineers.
CIBSE. 2005. CIBSE Applications Manual AM10: Natural ventilation in non-
domestic buildings. London: The Chartered Institute of Building Services
Engineers.
de Carvalho, P., M. da Silva, and J. Ramos. 2013. Influence of weather and
indoor climate on clothing of occupants in naturally ventilated school
buildings. Building and Environment 59: 38–46.
de Dear, R. 2010. Thermal comfort in natural ventilation—A
neurophysiological hypothesis. Proceedings of the 6th Windsor
Conference: Adapting to Change: New Thinking on Comfort Cumberland
Lodge, Windsor, U.K., April 9–11.
Deuble, M. and R. de Dear. 2010. Green occupants for green buildings: The
missing link? Proceedings of the 6th Windsor Conference: Adapting to
Change: New Thinking on Comfort Cumberland Lodge, Windsor, U.K.,
April 9–11
D’Oca, S. and T. Hong. 2014. A data-mining approach to discover patterns of
window opening and closing behavior in offices. Building and
Environment 82: 726–39.
Dutton, S., D. Banks, S. Brunswick, and W. Fisk. 2013. Natural ventilation in
California offices: estimated health effects and economic consequences.
Report number LBNL-6910E. https://www.osti.gov/biblio/1167556.
Hellwig, R., S. Brasche, and W. Bischof. 2006. Thermal comfort in offices—
natural ventilation versus air-conditioning. Extended abstract. Proceedings
of International Society of Indoor Air Quality and Climate’s Healthy
Buildings 2006, Lisbon, Portugal, June 4–8.
Hummelgaard, J., P. Juhl, K.O. Sæbjörnsson, G. Clausen, J. Toftum, and G.
Langkilde. 2005. Indoor air quality and occupant satisfaction in five
mechanically and four naturally ventilated open-plan office buildings.
Proceedings Indoor Air 2005.
ICC. 2017. 2018 International Mechanical Code®. Country Club Hills, IL:
International Code Council, Inc.
ICC. 2019. California green building standards code (CGBSC), Title 24, Part
11 (CALGreen). Washington, DC: International Code Council.
IWBI. 2016. WELL building standard®, v1. New York: Delos Living LLC.
Khatami, N., A. Hashemi, M. Cook, and S. Firth. 2014. Effects of manual and
automatic natural ventilation control strategies on thermal comfort, indoor
air quality and energy consumption. Proceedings of the ZEMCH 2014
conference, June 4–6, Londrina, Brazil.
Leaman, A., and B. Bordass. 2007. Are users more tolerant of “green”
buildings? Building Research & Information 35(6):662–73.
26 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Linden, P., E. Arens, and N. Daish. 2015. Natural ventilation for energy
savings in California commercial buildings. CEC-500-2016-039.
Sacramento, CA: California Energy Commission.
Martin, A., and J. Fitzsimmons. 2000. Making natural ventilation work.
BSRIA Guidance Note GN 7/2000. Bracknell, U.K.: BSRIA.
McConahey, E. 2008. Mixed mode ventilation: Finding the right mix.
ASHRAE Journal 50(9): 36-48.
Melton, P. 2014. Natural ventilation: The nine biggest obstacles and how
project teams are beating them. Environmental Building News 23(8):
August 2014. https://www.buildinggreen.com/feature/natural-ventilation-
nine-biggest-obstacles-and-how-project-teams-are-beating-them.
Mendell, M., E.A. Eliseeva, M.M. Davies, M. Spears, A. Lobscheid, W.J. Fisk,
and M.G. Apte. 2013. Association of classroom ventilation with reduced
illness absence: A prospective study in California elementary schools.
LBNL-6259E. Berkeley, CA: Lawrence Berkeley National Laboratory.
Morgan, C., and R. de Dear. 2003. Weather, clothing and thermal adaptation to
indoor climate. Climate Research 24:267–84.
Roetzel, A., A. Tsangrassoulis, U. Dietrich, and S. Busching. 2010. A review
of occupant control on natural ventilation. Renewable and Sustainable
Energy Reviews 14(2010):1001–13.
Rupp, R, N. Vasquez, and R. Lamberts. 2015. A review of human thermal
comfort in the built environment. Energy and Buildings 105 (2015):178–
205.
Seppänen, O., and W. Fisk. 2001. Association of ventilation system type with
SBS symptoms in office workers. Indoor Air 12(2):98–112.
Schakib-Ekbatan, K., F. Çakici, M. Schweiker, and A. Wagner. 2015. Does the
occupant behavior match the energy concept of the building?: Analysis of a
German naturally ventilated office building. Building and Environment 84
(2015):142–50.
USGBC. 2013. LEED® Reference Guide for Building Design and
Construction v4. Washington, DC: U.S. Green Building Council.
MODEL PARAMETERS
FOR OFFICE EXAMPLES
This appendix summarizes the façade and internal load assumptions
associated with the office occupancy example used throughout the guide.
Figure A1.1.1 shows the office model used for the examples in this guide.
The assumptions regarding the building envelope are as follows. The
model office space is 12 ft wide, 14 ft high, and 18 ft long (3.6 m × 4.2 m ×
5.5 m). For the analysis model, surface 2 is the external south-facing wall that
contains the operable window identified as surface 7 below. The yellow patch
shown in Figure A1.1.1 shows the sun path within the space for reference.

FIGURE A1.1.1 Office model for the examples in the guide.

27
28 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A1.1.1 Room Data for the Office Space


Length, Width, Height, Floor Area, Volume, Glazed Area,
ft (m) ft (m) ft (m) ft2 (m2) ft3 (m3) ft2 (m2)

18 (5.5) 12 (3.6) 14 (4.2) 216 (84.1) 3024 (84.1) 60 (5.5)

Natural Ventilation ASHRAE Standard 62.1 California’s Title 24


Openings, Heat Loss, Heat Gain, Ventilation Ventilation
Btu/h (W) Btu/h (W) Requirements, Requirements,
ft2 (m2) cfm (L/s) cfm (L/s)
10 (0.9) 1100 (–398) 7156 (2028) 23 (12) 32 (15.2)

TABLE A1.1.2 Heat Gain Calculations through the Room Envelope


Area, R-Value, U-Factor, Tilt, Orientation,
Surface Type
ft2 (m2) ft2F/Btu/h (m2·k/W) Btu/h·ft2F (W/m2·K) ° °

1 216 (19.8) 0.77 (0.3) 0.58 (3.29) 180 0 Int solid

2 98 (9.0) 16.73 (2.9) 0.06 (0.34) 90 180 Ext solid

3 252 (23.1) 16.87 (2.9) 0.06 (0.34) 90 90 Int solid

4 168 (15.4) 16.87 (2.9) 0.06 (0.34) 90 0 Int solid

5 252 (23.1) 16.87 (2.9) 0.06 (0.34) 90 270 Int solid

6 216 (19.8) 0.77 (0.3) 0.58 (3.29) 0 270 Int solid

7 70 (6.4) 2.68 (0.7) 0.27 (1.53) 90 180 Ext glass


Temp. on Other Side, Space Temp., ∆T Across Wall, Heat Gain,
Surface Boundary
°F (°C) °F (°C) °F (K) Btu/h (W)
1 Adjacent 74 (23.3) 74 (23.3) 0 0

2 Normal 98 (36.6) 74 (23.3) 24 (13.3) 141 (40.7)

3 Adjacent 74 (23.3) 74 (23.3) 0 0

4 Adjacent 74 (23.3) 74 (23.3) 0 0

5 Adjacent 74 (23.3) 74 (23.3) 0 0

6 Adjacent 74 (23.3) 74 (23.3) 0 0

7 Normal 98 (36.6) 74 (23.3) 24 (13.3) 454 (130.7)

Total 595 (171.3)

A1.1.1 HEAT LOAD TO THE SPACE


The first step in calculating the heat load to the space is to calculate the
heat gain to the space through the building envelope. For this example, the
outdoor dry-bulb temperature is 98°F (36.6°C) and the inside operating
APPENDIX A1.1 | 29

TABLE A1.1.3 Input Values for the Standard Analyses in the Guide
Outdoor Design Condition Office Winter Office Summer
Outdoor temperature, °F (°C) 13.5 (–10) 98 (36)

Façade Geometry

Ceiling height, ft (m) 14 (4.2) 14 (4.2)

Room length, ft (m) 12 (3.6) 12 (3.6)

Window height from sill, ft (m) 7 (1.8) 7 (1.8)

Sill height, ft (m) 3 (0.9) 3 (0.9)

Window width, ft (m) 10 (3) 10 (3)

Window-to-wall ratio,% 42 42

Window separation, ft (m) 2 (0.6) 2 (0.6)

Façade Performance

Window U-factor, Btu/ft2 h°F (W/m2·K) 0.28 (1.588) 0.28 (1.588)

Wall R-value, ft2hr°F/Btu (m2·K/W) 17.38 (0.6) 17.38 (0.6)

Indoor Conditions

Indoor temperature, °F(°C) 70 (20) 78 (26)

Relative humidity, % 20 50

Air speed, fpm (m/s) 30 (0.15) 30 (0.15)

Clothing, clo 1 0.85

Metabolic rate, met 1 1

Occupants 250 Btu/h ea (70 W ea) 250 Btu/h ea (70 W ea)

temperature is 74°F (23.3°C). The space heat gain can be calculated for each
surface from the classic U × A × T = Btu/h. (See Table A1.1.2.) (It is assumed
that the surrounding spaces are all operated at the same temperature and,
therefore, there are no heat gains or losses from the surroundings.)
The next step is to calculate the solar radiation to the space. This example
assumes a glazing area of 40%, in accordance with ASHRAE Standard 90.1
(2019a), of the exterior walls (70 ft2 [6.4 m2]), a solar heat gain coefficient
(SHGC) equal to 0.25, and a solar radiation to a west-facing window in June to
be 213 Btu/h/ft2 (674 W/m2). Therefore, the solar radiation load to the space is
70 × 0.25 × 213 = 3727 Btu/h (6.4 × 0.25 ×674 = 1082 W).
The percentage of surface 2 that is glazed will vary, depending on the
analysis, as numerous sensitivity analyses are provided. Similarly, the SHGC
30 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A1.1.4 Difference in Lighting and Plug Loads for the Guide’s Calculations
When References to a “Traditional” versus “High Efficiency”
Input Assumptions are Used
High Efficiency
Traditional Loads
Target Loads
Load Btu/h/ft 2
(W/m2) Btu/h/ft 2 (W/m2)
Lighting 4 (12) 2.04 (6)

Plug load 6.8 (20) 3.4 (10)

TABLE A1.1.5 Difference in Lighting and Plug Loads for Traditional and
High Efficiency Calculations per Space Floor Area
Traditional High Efficiency
Floor
Floor Area, Btu/h/ft2 Total, Area, Btu/h/ft2 Total,
Load
ft2 (m2) (W/m2) Btu/h (W) (W/m2) Btu/h (W)
ft2 (m2)

Lighting 216 (19.8) 4 (12) 864 (237.8) 216 (19.8) 2.04 (6) 441 (118.9)

Plug load 216 (19.8) 6.8 (20) 1469 (396.3) 216 (19.8) 3.4 (10) 734 (198.2)

2333 (634) 1175 (317)

TABLE A1.1.6 Load Breakdown for a Traditional Cooling Load and a


High Efficiency Cooling Load
Traditional Cooling Load, High Efficiency Cooling Load,
Btu/h (W) Btu/h (W)
Envelope load 595 (171.3) 595 (171.3)

Solar gain 3728 (1082) 3728 (1082)

Lighting 864 (237.8) 441 (118.9)

Plug loads 1469 (396) 734 (198.2)

Occupants 500 (140) 500 (140)

Total 7156 (2028) 5998 (1711)

of the glazing is typically stated as part of the analysis description within each
run. It is typically one of the following:

• SHGC = 0.25
• SHGC = 0.375
• SHGC = 0.5
APPENDIX A1.1 | 31

TABLE A1.1.7 Minimum Ventilation Calculations for Example Office


Occupancy- Area-Based
Criterion Based Minimum Minimum Calculation Results
Ventilation Ventilation
ASHRAE Standard 62.1 5 cfm/person 0.06 cfm/ft2 2p × 5 cfm/person+ 23 cfm (12 L/s)
(2.5 L/s/person) *
(0.3 L/s/m2) 0.06cfm/ft2 ×216 ft2
2p × 2.5 L/s/person+
0.3 L/s/m2 × 20 m2

California’s Title 24 15 cfm/person 0.15 cfm/ft2 Higher of: 32 cfm (15.2 L/s)
(7.0 L/s/person) 0.15 cfm/ft2 ×216 ft2
(0.76 L/s/m2)
= 32 cfm
or
2p × 15 cfm/person =
30 cfm

(0.76 L/s/m2·20m2 =
15.2 L/s
or
2p × 7.0 L/s/person =
14 L/s)

The assumptions regarding heating and cooling loads are listed in


Table A1.1.3.
Next, calculate the space’s internal loads, which include:

• Lighting
• Plug loads
• Occupants
For present-day designs with a high efficiency energy goal, the lighting and
plug loads are significantly reduced from what we call “traditional” load (i.e.,
industry-standard allowances based on historical data based on the system
performance of older technologies). Both traditional loads and high efficiency
loads are represented in Table A1.1.4 for the office occupancy.
As the space floor area is 216 ft2 (19.8 m2), the following lighting and plug
loads are calculated and shown in Table A1.1.5.
There are two occupants at 250 Btu/h pp (70 W pp), which provide a heat
load to the space of 500 Btu/h (146 W).
The naturally ventilated office is assumed to be possibly occupied from
7:00 a.m. through 7:00 p.m., five days a week. The analysis was carried out for
selective months.
32 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A1.1.2 VENTILATION REQUIREMENTS


Minimum ventilation air is required to comply with ASHRAE
Standard 62.1 (2019) and California’s Title 24 (CEC 2018a, 2018b) ventilation
requirements. Minimum ventilation air, as required under ASHRAE
Standard 62.1’s Table 6-1, “Minimum Ventilation Rates in Breathing Zone,”
for an office space, would be as noted in Table A1.1.7, assuming two people in
a 216 ft2 (20 m2) space. In ASHRAE Standard 62.1 (2019b), the area-based
result and the occupancy-based results are added, whereas in California’s Title
24 the higher of the two results is selected.
The occupant met rate is assumed to be 1.0 as described within the relevant
analysis, and occupant clo is assumed to be 0.85 clo to 1.0 clo as described
within the relevant analysis.

A1.1.3 REFERENCES
ASHRAE. 2010. ANSI/ASHRAE/IES Standard 90.1-2010, Energy standard
for buildings except low-rise residential buildings. Peachtree Corners, GA:
ASHRAE.
ASHRAE. 2019a. ANSI/ASHRAE/IES Standard 90.1-2019, Energy standard
for buildings except low-rise residential buildings. Peachtree Corners, GA:
ASHRAE.
ASHRAE. 2019b. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
CEC. 2018a. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
CEC. 2018b. 2019 Nonresidential Compliance Manual for the 2019 Building
Energy Efficiency Standards: Title 24, Part 6 and associated
Administrative Regulations in Part 1. CEC-400-2018-018-CMF.
Sacramento, CA: California Energy Commission.
MODEL PARAMETERS FOR
CLASSROOM EXAMPLES

This appendix summarizes the façade and internal load assumptions


associated with the classroom occupancy example used throughout the guide.
Figure A1.2.1 shows the classroom model used for the examples in this guide.
The model classroom space is 30 ft wide, 14 ft high, and 30 ft long (9.1 m
× 4.2 m × 9.1 m). The floor area is 900 ft2 (82.6 m2) and the room volume is
12,600 ft3 (351 m3).
The bottom operable window is 28 ft wide and 1 ft high (8.5 m × 0.3 m),
the middle-fixed window is 28 ft wide and 6 ft high (8.5 m × 1.8 m), and the
top operable window is 28 ft wide and 1 ft high (8.5 m × 0.3 m). The yellow
patch shown in Figure A1.2.1 shows the sun path within the space.

FIGURE A1.2.1 Rectangular classroom model for the examples in the guide.

33
34 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A1.2.1 Room Data for the Classroom Space


Width, Height Length, Floor Area, Volume, Glazed Area,
ft (m) ft (m) ft (m) ft2 (m2) ft3 (m3) ft2 (m2)

30 (9.1) 14 (4.2) 30 (9.1) 900 (82.6) 12,600 (351) 168 (15.4)

Natural Ventilation ASHRAE Standard 62.1 California’s Title 24


Openings, Heat Loss, Heat Gain, Ventilation Ventilation
Btu/h (W) Btu/h (W) Requirements, Requirements,
ft2 (m2) cfm (L/s) cfm (L/s)

56 (5.1) 2688 (–970) 27,847 (6090) 418 (204) 465 (217)

TABLE A1.2.2 Heat Gain Calculations through the Room Envelope


Area, R-Value, U-Factor, Tilt, Orientation,
Surface Type
ft2(m2) ft2F/Btu/h (m2·k/W) Btu/h ft2F (W/m2·k) ° °

1 900 (86.6) 0.77 (0.3) 0.58 (3.29) 180 90 Int solid

2 252 (23.1) 17.38 (2.9) 0.05 (0.34) 90 270 Ext solid

3 420 (38.5) 17.38 (2.9) 0.05 (0.34) 90 180 Int solid

4 420 (38.5) 17.38 (2.9) 0.05 (0.34) 90 90 Int solid

5 420 (38.5) 17.38 (2.9) 0.05 (0.34) 90 0 Int solid

6 900 (82.6) 4.44 (1.0) 0.18 (1.02) 0 0 Ext solid

7 168 (15.4) 2.61 (0.6) 0.28 (1.588) 90 270 Ext glass


Heat
Temp. on Other Side, Space Temp., ∆T Across Boundary,
Surface Boundary Gain,
°F (°C) °F (°C) °F(K)
Btu/h (W)
1 Adjacent 74 (23.3) 74 (23.3) 0 —

2 Normal 98 (36.6) 74 (23.3) 24 (13.3) 302 (104.5)

3 Adjacent 74 (23.3) 74 (23.3) 0 —

4 Adjacent 74 (23.3) 74 (23.3) 0 —

5 Adjacent 74 (23.3) 74 (23.3) 0 —

6 Adjacent 74 (23.3) 74 (23.3) 0 —

7 Normal 98 (36.6) 74 (23.3) 24 (13.3) 1129 (325.5)

Total 1431 (430)

A1.2.1 HEAT LOAD TO THE SPACE


The first step for determining heat load to the space is to calculate the heat
gain to the space through the building envelope. For this example, the outdoor
dry-bulb temperature is 98°F (36.6°C) and the inside operating temperature is
74°F (23.3°C). The space heat gain can be calculated for each surface from the
APPENDIX A1.2 | 35

TABLE A1.2.3 Input Values for the Standard Analyses in the Guide
Outdoor Design Condition Classroom Winter Classroom Summer
Outdoor temperature, °F (°C) 13.5 (–10) 98 (36)
Façade Geometry
Ceiling height, ft (m) 14 (4.2) 14 (4.2)

Room length, ft (m) 30 (9.1) 30 (9.1)

Window height from sill, ft (m) 8 (2.4) 8 (2.4)

Sill height, ft (m) 3 (0.9) 3 (0.9)

Window width, ft (m) 28 (8.5) 28 (8.5)

Window-to-wall ratio, % 53 53

Window separation, ft (m) 3 (0.9) 3 (0.9)


Façade Performance

Window U-factor, Btu/ft2·h·°F (W/m2·K) 0.28 (1.588) 0.28 (1.588)

Wall R-Value, ft2·h·°F/Btu (m2·K/W) 17.38 (0.6) 17.38 (0.6)

Indoor Conditions
Indoor temperature, °F (°C) 70 (20) 78 (26)

Relative humidity, % 20 50

Air speed, fpm (m/s) 30 (0.154) 30 (0.154)

Clothing, clo 1 0.85

Metabolic rate, met 1 1

Occupants 250 Btu/h per person (pp) 250 Btu/h pp (70 W pp)
(70 W pp)

classic equation, U × A × ∆T = Btu/h. (See Table A1.2.2). (It is assumed that


the surrounding spaces are all operated at the same temperatures and,
therefore, there are no heat gains or losses from the surroundings.)
The next step is to calculate the solar radiation to the space. For this
example, a glazing area of 40%, in accordance with ASHRAE Standard 90.1
(2019a) of the exterior walls (420 ft2 [42 m2]), a solar heat gain coefficient
(SHGC) of 0.25, and solar radiation to the south-facing window in June of
213 Btu/h·ft2 (674 W/m2) is assumed. Therefore, the solar radiation load to the
space is 168 × 0.25 × 213 = 8946 Btu/h (15.4 × 0.25 × 674 = 2597 W).
The percentage of surface 2 that is glazed will vary, depending on the
analysis, as numerous sensitivity analyses are provided. Similarly, the SHGC
of the glazing is typically stated as part of the analysis description within each
run. It is typically one of the following:
36 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A1.2.4 Difference in Lighting and Plug Loads for


Traditional and High Efficiency Calculations
Traditional High Efficiency,
Load Btu/h/ft2 (W/m2) Btu/h/ft2 (W/m2)
Lighting 4 (12) 2.04 (6)

Plug load 6.8 (20) 3.4 (10)

TABLE A1.2.5 Difference in Lighting and Plug Loads for Traditional and
High Efficiency Calculations per Space Floor Area
Traditional High Efficiency
Floor Total, Total,
Area, Btu/h·ft2 Floor Area, Btu/h·ft2
Load Btu/h Btu/h
(W/m2) ft2 (m2) (W/m2)
ft2 (m2) (W) (W)
Lighting 900 (82.6) 4 (14) 3600 900 (82.6) 2.04 (7.2) 1836
(330.6) (168.6)

Plug 900 (82.6) 6.8 (24) 6120 900 (82.6) 3.4 (12) 3060
load (562.0) (281.0)

9720 (893) 4896 (450)

TABLE A1.2.6 Comparison of Traditional Heat Gains and


High Efficiency Heat Gains for the Classroom
Traditional Cooling Load, High Efficiency Cooling Load,
Btu/h (W) Btu/h (W)
Envelope load 1431 (430) 1431 (430)

Solar gain 8946 (2597) 8946 (2597)

Lighting 3600 (331) 1836 (169)

Plug loads 6120 (562) 3060 (281)

Occupants 7750 (2170) 7750 (2170)

Total 27,847 (6090) 23,023 (5647)

• SHGC = 0.25
• SHGC = 0.375
• SHGC = 0.5

The assumptions regarding heating and cooling loads are listed in


Table A1.2.3.
Next, calculate the space internal loads, which include:
APPENDIX A1.2 | 37

TABLE A1.2.7 Minimum Ventilation Calculations for Example Classroom


Occupancy-Based Area-Based
Criterion Minimum Minimum Calculation Results
Ventilation Ventilation
ASHRAE Standard 62.1 10 cfm/person 0.12 cfm/ft2 31 p × 10 cfm/person+ 418 cfm
(5 L/s/person) 0.12 cfm/ft2 × 900 ft2 (204 L/s)
(0.6 L/s/m2)
(31 p × 5 L/s/person +
0.6L/s/m2×82.6m2)

California’s Title 24 15 cfm/person 0.15 cfm/ft2 Higher of: 465 cfm


(7.0 L/s/person) (217 L/s)
(0.76 L/s/m2) 0.15 cfm/ft2 × 900 ft2= 135 cfm

(0.76 L/s/m2 × 82.6m2 = 62.7 L/


s)
or
31 p × 15 cfm/person = 465 cfm
(31 P × 7.0 L/s/person = 217 L/s)

• Lighting
• Plug loads
• Occupants

For present-day designs with a high efficiency energy goal, the lighting and
plug loads are significantly reduced from what we call the “traditional” load.
As the space floor area is 900 ft2 (216 m2), the following lighting and plug
loads are calculated and shown in Table A1.2.4.
There are two occupants at 250 Btu/h per person (70 W per person), which
provides a heat load to the space of 500 Btu/h (2170 W).
The naturally ventilated classroom is assumed to be possibly occupied
from 7:00 a.m. through 7:00 p.m., five days a week. The analysis was carried
out for selective months.

A1.2.2 VENTILATION REQUIREMENTS


Minimum ventilation air is required to comply with ASHRAE
Standard 62.1 (ASHRAE 2019b) and California’s Title 24 (CEC 2018a,
2018b) ventilation requirements. The required minimum outdoor air supply
volumes are as follows. Minimum ventilation air as required under ASHRAE
Standard 62.1’s Table 6-1, “Minimum Ventilation Rates in Breathing Zone,”
for a classroom space would be as noted in Table A1.2.7, assuming 31 people
in a 900 ft2 (82.6 m2) space. In ASHRAE Standard 62.1, the area-based result
and the occupancy-based results are added, whereas, in California’s Title 24,
the higher of the two results is selected.
38 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Occupant met rate is assumed to be 1.0 as described within the relevant


analysis. Occupant clo is assumed to be 0.85 clo to 1.0 clo as described within
the occupant comfort analysis.

A1.2.3 REFERENCES
ASHRAE. 2019a. ANSI/ASHRAE/IES Standard 90.1-2019, Energy standard
for buildings except low-rise residential buildings. Peachtree Corners, GA:
ASHRAE.
ASHRAE. 2019b. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
CEC. 2018a. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
CEC. 2018b. 2019 Nonresidential Compliance Manual for the 2019 Building
Energy Efficiency Standards: Title 24, Part 6 and associated
Administrative Regulations in Part 1. CEC-400-2018-018-CMF.
Sacramento, CA: California Energy Commission.
KEY CONCEPTS
2.1 NATURAL VENTILATION VERSUS
NATURAL CONDITIONING
The term natural ventilation has been in use at least since 1899, when
Robert Boyle & Son, Ltd., reprinted a 21-page pamphlet called “Natural
Ventilation,” based on an article from The Building News, dated May 26, 1899
(Boyle 1899b), and then published a 92-page compendium of the prevailing
arguments of the day entitled “Natural and Artificial Methods of Ventilation”
(Boyle 1899a). At the time, the challenge was how to design “engineered”
natural ventilation systems that gave an equivalent ventilation performance as
one might achieve with mechanical systems. Over a century later, this remains
the challenge we attempt to tackle in this design guide.
Ventilation is the provision of outdoor air to a space. When people talk
about natural ventilation in normal parlance, they usually envisage opening a
window to bring fresh air into a space. Natural is used to differentiate the
driving force causing the air movement as compared to a mechanical means.
Most people will open a window to relieve stale air or to cool a room down.
According to the definitions established by ASHRAE, and for the purposes of
this guide, the former (i.e., opening a window to relieve stale air) is considered
natural ventilation, and the latter (i.e., cooling a space) is considered natural
conditioning. It should be noted that if a team is successful at providing natural
conditioning, it is highly likely that natural ventilation has also been
accomplished.
Chapters 5 and 6 of this guide discuss the design approaches associated
with showing compliance with ASHRAE Standard 55 and 62.1.

2.2 MIXED-MODE VENTILATION


Mixed-mode ventilation (also known as hybrid ventilation) typically uses
natural conditioning when outdoor air conditions allow while using a
mechanical cooling system for the remainder of the time. This design

39
40 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 2.1 Definitions of Natural Ventilation and Natural Conditioning


Natural Ventilation Natural Conditioning
Per ASHRAE Standard 62.1-2019 (ASHRAE 2019) Per ASHRAE Standard 55-2017 (ASHRAE 2017a),
the definition of natural ventilation is “ventilation the definition of occupant-controlled naturally
provided by thermal, wind, or diffusion effects conditioned spaces is “those spaces where the
through doors, windows, or other intentional thermal conditions of the space are regulated
openings in the building.” primarily by occupant-controlled openings in the
envelope.”
Essentially, direct air intake from the outdoors into
the occupied space to support only the indoor air Essentially, direct air intake from the outdoors into
quality requirements. the occupied space supports the absorption of heat
gains. This can occur via
• direct cooling of occupants
• removal of air heated from computers or lights
• cooling of surfaces that then causes indirect cooling
of occupants through radiative heat exchange

Two paths for proving adequate ventilation for The criterion for thermal comfort and the limitations
occupants are governed by ASHRAE Standard 62.1- on the use of the adaptive comfort standard are
2019 (ASHRAE 2019): prescriptive design or the established by ASHRAE Standard 55-2017 (ASHRAE
performance method of compliance. 2017a).

philosophy is a response to the fact that peak outdoor air conditions in many
places in the United States have insufficient cooling capacity to ensure human
comfort. (See Chapter 3 for a more in-depth discussion of this topic.)
Per CIBSE Applications Manual AM13, Mixed Mode Ventilation, mixed-
mode systems are divided into three categories (CIBSE 2000):

• Contingency Design: In this approach, the building fabric is designed


to accommodate natural ventilation and natural conditioning but space
is reserved to allow mechanical cooling and ventilation equipment to
be installed in the future without great disruption.
• Complementary Designs: these approaches use integrated designs that
allow a smooth transition between natural and mechanical ventilation
systems. There are two subcategories of this approach about when the
mechanical systems operate: concurrent operations and changeover
operations, as noted in Table 2.2.
• Zoned Design: this approach prescribes different strategies to different
spaces depending on the location and occupancy of areas within the
building.

During the early phases of deciding whether to use mixed-mode systems, it


is appropriate to have a discussion with the building owner to determine
whether there should be a capital expenditure to purchase little-used equipment
in advance of occupancy, or whether it would suffice to reserve space,
electrical capacity, and louvered access to outdoor air to add equipment in the
future based on observed demand. The other reason to discuss the contingency
design option is in recognition of the trends of global warming. It is very likely
CHAPTER 2 | 41

TABLE 2.2 Categories of Mixed-Mode Ventilation Systems


Complementary Design
Zoned Design
Concurrent Changeover
Mechanical cooling and natural Mechanical cooling and natural ventilation occur Mechanical cooling and natural
ventilation occur in in ventilation occur in
• Same space • Same space • Different spaces
• Same time • Different times • Same time

Graphics courtesy of the Center for the Built Environment (2008).

that buildings designed for natural conditioning may currently have reduced
numbers of hours when that system can provide comfort cooling. If the owner
anticipates holding the asset for 20 years or more, a contingency design, at a
minimum, is strongly recommended.
The following common recommendations are made when natural
ventilation is coupled with auxiliary mechanical cooling (McConahey 2011):

• Concurrent Mixed Mode Ventilation—In this scheme, mechanical


cooling and natural ventilation will serve the same space at the
same time. This type of configuration is most often seen
incorporating a form of radiant cooling or heating as supplementary
heat transfer in the space to assist with human comfort. Radiant
cooling has benefits over air-based cooling systems in the
concurrent modes because there is limited energy waste with
dilution of cooling air by the air from outdoors and the savings of
fan energy. With radiant cooling as a predominantly independent
heat transfer mechanism as compared to convective cooling
through “wind-on-skin,” the space can be allowed to have a dry-
bulb temperature rise up to 80°F (26.6°C) or higher, with radiant
heat absorption occurring due to an artificially lowered mean
radiant temperature, thereby reducing perceived temperature.
• Changeover Mixed-Mode Ventilation—In this scheme,
mechanical cooling and natural ventilation will serve the same
space, but at different times during the year. This type of
configuration is usually triggered by relying on human choice or
through outdoor temperature tracking, depending on the
sophistication of the system. In general, keeping these systems
simple, but providing feedback to the users about appropriate
interaction with the building façade is beneficial. Some schemes
42 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

employ a red-light/green-light system to remind users that the


outdoor air is sufficiently cool to open the windows, while others
use email or text messaging to provide similar information. It is
usually the case that window-switches are incorporated into
changeover schemes, especially when occupants are in charge of
opening and closing windows, so that energy savings can be
ensured. Similarly, motorized windows that open in response to
outdoor air temperature may be appropriate in transient areas with
shared occupancy (e.g., lobbies, atria, dining areas, conference
rooms, and so on), as areas with no permanent ownership tend to
lack active occupant participation with windows unless the base-
condition upon entering is considered to be inadequate. Air-based
cooling systems are appropriate for changeover mixed-mode
systems as the HVAC operates in an ON/OFF mode and there is less
risk of air-conditioning the outdoors as compared to other mixed-
mode ventilation types.
• Zoned Mixed-Mode Ventilation—In this scheme, mechanical
cooling is provided to certain spaces, while other spaces in the same
building experience natural ventilation. This scheme is often used
with deep floor plates, as natural ventilation has a limited
applicability only to the perimeter zones that have direct access to
window openings. Additionally, there are often high heat load areas
such as server rooms or high-density occupancy conference rooms
or classrooms that may not be able to achieve sufficient cooling just
based on outdoor air temperature. These schemes must take care to
avoid adverse effects on the energy-consuming interior space
HVAC systems due to pressure variability in the naturally
ventilated zones. Additionally, some schemes employ a
temperature reset algorithm for the interior HVAC systems in order
to avoid excessive differences in perceived temperature between
naturally ventilated zones and those zones with constant air-
conditioning.”

When mixed-mode systems are pursued, the onus is on the design team to
ensure that compliance with all governing standards is achieved under all
operating conditions. This typically involves at least two parallel calculation
methodologies and demonstration of compliance at the changeover
temperature conditions in the case of engineered natural ventilation or natural
conditioning systems.
For instance, concurrent systems are often applied when there are relatively
few hours (less than 5% to 10% of occupied hours) when natural conditioning
is not possible. Concurrent systems do not qualify as naturally conditioned
spaces under the definitions of ASHRAE Standard 55-2017 (ASHRAE 2017a).
They can be considered to be naturally ventilated under ASHRAE Standard
62.1-2019 (ASHRAE 2019), so long as a prescriptive approach was pursued,
or an engineered approach demonstrates that the addition of cooling to the
CHAPTER 2 | 43

interior does not adversely affect buoyancy-based driving forces for natural
ventilation. The area would technically be using a mechanical conditioning
system, and ANSI/ASHRAE/IES Standard 90.1 (ASHRAE 2019a) energy
modeling protocols would force the analyst to assume that the mechanical
system is running.See Website Appendix WA5.5 for more details. The
California Energy Code (CEC 2018a) makes similar assumptions and sets the
parameters of the equipment type and operation of the supplemental system.
See Website Appendix WA5.2, section WA5.2.4, for more details.
Similarly, in the case of the changeover system, in which windows are
expected to be closed for extended periods of time, the building’s HVAC
system would generally be expected to have mechanical systems provide
ventilation quantities comply with Table 6-1, “Minimum Ventilation Rates in
Breathing Zone” in ASHRAE Standard 62.1 (ASHRAE 2019b) or local codes.
If the windows remain operable by the occupants during the cooling mode, an
argument could be made to the authorities having jurisdiction (AHJ) that
natural ventilation remains intact and the system is actually a concurrent
system. This would likely be the case if a local recirculation type of fan coil
unit system were used during the cooling mode. The predictive control of
change-over systems for comfort control as soon as the mechanical cooling is
activated is of particular concern. Mixed-mode systems relying on radiant
panels or chilled beams may have to have their period of natural ventilation use
shortened if outdoor air dew points are high. See Chapter 7 for more details on
this topic.
Mixed-mode systems have the potential to reduce operational energy use
as compared to only running mechanical systems throughout the year;
however, the cost of adding operable elements can be significant and thus must
be considered in the return-on-investment (ROI) calculations.

2.3 HOW DOES NATURAL VENTILATION WORK?


Fluids only move when they experience a pressure difference between two
physical points along a path. Unlike a motorized fan that creates constant high
pressure, natural ventilation relies on the available pressure sources of wind or
the pressure difference that is created by differences in buoyancy between
indoor and outdoor air. Table 2.3 describes the working principles in their
purest independent form. It should be noted that in most cases both principles
are at work whenever wind-driven air movement is available.
Natural conditioning relies on air movement across the skin especially as
outdoor temperatures rise to target indoor conditions. Indoor air speeds, even
under the most favorable conditions, are only 30% to 40% of the free exterior
wind speed in cross-ventilated spaces, 5% to 15% of the free exterior wind
speed in rooms with multiple openings in one wall only, and only 3% to 5% in
rooms with one opening.
Though obvious, the following basic natural ventilation principles bear
repeating:
44 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 2.3 Buoyancy versus Wind-Driven Natural Ventilation:


Working Principles and Limitations

Buoyancy-Driven Natural Ventilation


Wind-Driven Natural Ventilation
(Stack Effect)

FIGURE 2.1 Single-sided ventilation with FIGURE 2.2 Cross ventilation with
double opening, showing single-story use of opposing side openings, showing single-story
buoyancy (CIBSE 2005). use of wind-driven ventilation (CIBSE 2005).

Working Principle Working Principle

• Indoor air warmed by heat sources rises to the ceiling and • The difference in static pressure between the high
is typically less dense than the outdoor ambient air. When pressure on the windward face of a building and the
high- and low-level openings create a full airflow path, negative pressure created on the far leeward side creates
then a hydrostatic pressure differential exists at the upper air movement if intake and outlet openings are available
opening between indoors and outdoors, allowing the and connected via a relatively open path.
heated air to flow outwards. Through conservation of mass
within the space, cool air is then pulled in through the
lower opening.

Limitations in Application Limitations in Application

• When outdoor air is significantly cooler than indoor air, it • Wind velocity increases with height as the number of
can “fall into” the space if the upper opening does not ground-based obstructions decreases. This should be
have sufficient control over opening size. taken into account when considering the strength of air
movement through occupied spaces.
• When multiple floors are attached to a common stack,
there is a risk that the top floor intake openings become the • Care must be taken to ensure that the design can
exhaust openings as the path of least resistance for the maintain an open path of airflow of sufficient openness to
leaving hot air. The design solutions that tend to alleviate relieve stuffiness and provide cooling for the whole floor.
this concern include stack roof terminations that extend at Because the smallest opening on either side sets the total
least 1.5-floor heights above the last floor served and airflow, the windward occupant is typically the one in
isolation of the top floor so that it can function as a closed control because they are most inclined to close the
system with a dedicated buoyancy relief path through window in response to their own comfort.
operable skylights and/or windows.
CHAPTER 2 | 45

1. Natural ventilation only works if there are at least two openings (or a very
tall opening that acts as a two-way opening).
2. The amount of airflow is controlled by the size of the smallest opening.
3. Air will always follow the path of least resistance.

The following sections describe in more detail some of the technical


aspects to consider when configuring a building to accommodate internal air
movement without mechanical assistance. This section does not detail the
physics formulas that govern these mechanisms, as most designers will use
software programs to do detailed calculations. Thus, the next three sections are
abstractions of technical findings that can be used as rules of thumb to see if a
building shape is conducive to accommodating natural ventilation.

2.4 PLANNING FOR BUOYANCY-DRIVEN VENTILATION


The working principle behind buoyancy (or the stack effect) is taught to
schoolchildren as “heat rises.” In historical designs prior to the introduction of
air-conditioning, the stack effect was widely used as a natural means to remove
heat from an indoor space. Semiburied indigenous architecture with low
outdoor air inlets and fire-driven stack effect ventilation can be found at Mesa
Verde National Park in Colorado. (See Figures 2.3 through 2.5.) Most
Victorian homes and Antebellum southern homes feature tall ceilings and tall
double-hung windows that create a high and low opening to allow for a single-

FIGURE 2.3 Outdoor air inlet and fire pit within a kiva. (Roof not present.)
Licensed under Creative Commons Attribution-Share Alike 3.0 Unported license.
46 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 2.4 Visitor descending ladder FIGURE 2.5 Reconstructed roof of kiva
over fire pit for access. Outgoing smoke passes shows access via ladder. The central outdoor
through this same opening. air intake opening is shared between
adjacent rooms.
Courtesy: Erin McConahey.
Courtesy: Erin McConahey.

story stack effect. Even Robert Boyle in 1899 demonstrated the stack effect for
classrooms in his advocacy for engineered natural engineering via the use of
roof ventilators (see Figure 2.6). These days the phenomenon is much more
associated with the unfortunate experience of overheated top floors of
buildings with atrium spaces.

2.4.1 The Math Behind the Working Principle of


Buoyancy-Driven Ventilation
While experience and intuition tell us that heat rises, the important thing to
note is that heat rises with respect to the cooler surrounding air. When trying to
design a whole building that relies on buoyancy, the air density differences in
question are between the inside of the building and the outside of the building.
CIBSE Applications Manual AM10, Natural Ventilation in Non-Domestic
Buildings (CIBSE 2005) demonstrates the relative pressure gradients quite
well.
As shown in Figures 2.7 and 2.8, the neutral plane is the elevation where
the pressures are equal. The pressure difference between indoors and outdoors
is what drives the airflow direction. The internal pressure gradient is relatively
stable under occupied loads, as it is driven by internal heat gains. The outdoor
air pressure gradient will have a variable slope, becoming steeper as the
outdoor air temperature rises. Therefore, as the outdoor air temperature starts
to approach the indoor air temperature, less and less driving force is available.
Buoyancy-driving forces are typically extremely small at the moment
when the greatest outdoor airflow through the building is required, namely
CHAPTER 2 | 47

FIGURE 2.6 Diagram of engineered natural ventilation scheme from Natural and Artificial
Methods of Ventilation.
(Boyle 1899a)

warm days at the limit of natural ventilation’s capacity. For short height stack
ventilation, the pressure difference is linear to the temperature differential
between indoors and outdoors, as noted in Table 2.4.
Thus, on a warm day, when the (indoor/outlet temperature minus outdoor/
inlet temperature) temperature differential may be only 3°F (1.66°C), a single-
story high/low window configuration can only develop 0.00096 in. w.g.
(0.24 Pa). For reference, a typical MERV 8 prefilter would have an initial
pressure drop of 0.3 in. w.g. (75 Pa). (As a side note, this is one of the reasons
why filters are rarely found in natural ventilation systems. Refer to Chapter 3
for more information on outdoor air quality checks prior to pursuing natural
ventilation.)
When the openings are of unequal size, the total amount of airflow through
each of the openings must be the same, and the pressure drop through the
opening must be equal. Diagrammatically, this means that the neutral plane
moves toward the area of the larger opening.
See Appendix A2.1 for more information about the equations used to
calculate the neutral plane and to predict the airflow through an opening in a
simple buoyancy-driven design.
48 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 2.7 Internal and external FIGURE 2.8 Superimposed pressure


pressure gradients for a cold exterior gradients, neutral plane and airflow
condition. direction.
(CIBSE 2005) (CIBSE 2005)

TABLE 2.4 Pressure Differential per Degree of Indoor to Outdoor Temperature


Difference in Short Height Spaces for Equal Sized Openings (CIBSE 2005)
Height Pressure Differential Developed
35 ft (10.5 m) 0.00094 in. w.g./°F (0.426 Pa/°C)

11.5 ft (3.5 m) 0.00032 in. w.g./°F (0.147 Pa/°C)

The whole reason to do an estimate on average high- and low-level


opening sizes and to approximate the neutral plane location under weather
extremes is to ensure the minimum ventilation rate for compliance purposes as
compared to the nominal requirements of the mechanical ventilation rules in
ASHRAE Standard 62.1 (ASHRAE 2019) or other prevailing local standards.
See Chapters 5 and 6 for more information on computer methods used to
accomplish this task.
It is important to note that the desired flow rates achieved through stack
effect alone on a still day can only be adjusted by manipulating the size of the
openings and/or adjusting the height between openings. Occupants will
eventually only have the ability to control the opening size, as the heights
between available openings have been preset by the designer. Thus, it is crucial
that designers interested in natural ventilation use these simple calculations at
the earliest phases in the design process to determine in aggregate what the
opening sizes must be and at what heights they should be placed within the
façade.
CHAPTER 2 | 49

2.4.2 Complex Geometries and the


Determination of Neutral Plane
As can be seen in the calculations in Appendix A2.1, even with the
simplicity of a two-opening system the hardest thing for the designer to
determine by hand is the height of the neutral plane. This is exacerbated by the
fact that the neutral plane moves based on opening size and the real-life
variability of the temperature differential between indoors and outdoors, so one
can only check extreme conditions to set the boundaries of likely performance
when checking ventilation quantities.
When complex multifloor or multizone geometries with multiple windows
are used, it is extremely difficult to estimate the height of the neutral plane by
observation or simple calculation. In this case, the use of a simple modeling
tool is recommended. One such tool is Cool Vent (http://coolvent.mit.edu/),
created by the Building Technology Research Group at Massachusetts Institute
of Technology (MIT) under the supervision of Professor Leon Glicksman
(MIT n.d.). The visualization output can show approximately where the
airflow direction reversal begins, as noted in Figures 2.9 and 2.10. The
software can also estimate airflow rates (in m3/s [cfm]) as well as room
temperatures (visualized only in °C but reported in °C for the SI run or °F for
the I-P run in Hourly Report option).

2.5 PLANNING FOR WIND-DRIVEN VENTILATION


Wind-driven ventilation is a result of unequal surface pressures
experienced across a building envelope. As wind impinges on the obstruction
that is the building, a high-pressure zone is developed on the windward side, a
low-pressure zone is developed on the leeward side, and slight negative
pressure is developed on all the planes that are parallel to the wind’s path of
travel. This is consistent with Bernoulli’s Principle, which is an understanding
of the conservation of energy in fluid dynamics: when velocity increases,
pressure (i.e., potential energy) decreases, and vice versa. As noted in CIBSE
(2005), the distribution of pressure over a surface depends on the wind speed,
the wind direction, and the shape of the building.
The working principle behind wind-driven ventilation is readily
experienced whenever we are in the wind canyons of a city with tall buildings,
or even just in a room with a couple of open windows on opposing walls. It is
easy to discern how high pressure is generated on the windward side, and how
this “pushes” air through the space, just as a fan does in a ducted supply
system. But whereas a typical electric fan in an HVAC system can generate
high static pressure to push air through several obstructions (e.g., coils, filters,
duct silencers, dampers, and so on), the natural phenomenon of wind-driven
ventilation has a pressure generation capacity that is limited by air speed. Thus,
the occupants must have the ability to dynamically control the opening size to
limit or increase the amount of air that comes into the building based on the
externally imposed pressure from the wind. In low-height buildings, the path
between intake and leaving openings must also remain unobstructed due to
low-pressure availability.
50 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 2.9 Results visualization (I-P) from CoolVent (MIT n.d.), demonstrating flow
reversal between levels 3 and 4 and establishing the approximate location of the neutral pressure
level (NPL), flow rate through openings, and temperature.

FIGURE 2.10 Results visualization (SI) from CoolVent (MIT n.d.), demonstrating flow
reversal between levels 3 and 4 and establishing the approximate location of the NPL, flow rate
through openings, and temperature.
CHAPTER 2 | 51

FIGURE 2.11 Wind pressures acting on the building.


(CIBSE 2005)

2.5.1 The Math Behind the Working Principle of


Wind-Driven Ventilation
The math behind the working principle of wind-driven ventilation is
straightforward, but as with the neutral plane in buoyancy-driven ventilation,
two variables require a bit of work to get good input data. The first is how to
determine appropriate wind speed, and the second is how to determine
appropriate pressure coefficients. See Appendix A2.2 for more information on
the equations used to calculate flow through openings of a simple wind-driven
scheme.
2.5.1.1 Establishing Wind Speed. To calculate airflow in wind-driven
ventilation, it is necessary to apply a wind speed. As a rule of thumb, it is
possible to test the preliminary viability of natural ventilation schemes using
4.5 mph to 9 mph, or rather, 400 fpm to 800 fpm (2 m/s to 4 m/s) as
representative of gentle outdoor air movement. If a design cannot sustain target
indoor conditions under this outdoor velocity, it may be necessary to rethink
the size and distribution of openings.
The primary problem with establishing wind speed is that sometimes there
is either not a weather station close to the property or that the weather station is
at an airport and thus unrepresentative of the conditions on site. Airports and
naval facilities tend to have historical wind data, so if an airport is nearby, a
design team can obtain historical data through the sources included in Chapter
3 or can select a typical meteorological year from https://energyplus.net/
weather for that station. If there is no nearby airport, the design team can
review the data available from the U.S. Department of Energy’s (DOE) Wind
Energy Technologies Office (WETO), and in particular, a map published by
WETO titled “U.S. Average Annual Wind Speed at 30 m” at https://
windexchange.energy.gov/maps-data/325. (See Figure 2.12.)
52 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 2.12 United States annual average wind speed at 30 m.


(DOE n.d.)

Once the annual wind speed Uannual is estimated, then the design values for
hourly meteorological wind speed Umet at 100 ft (30 m) can be calculated
using Table 2.5.
Thus, the design Umet under extreme wind strength is 12.5 times that which
is available at poor wind speeds. As the original Uannual is typically measured
at an airport, the value of wind speed at the building site and height is further
adjusted by the application of terrain category, using the correction factors in
Table 2.6. (See Appendix A2.3 for the full details on these calculations).
This combination of steps allows the designer to have a reasonable starting
approximation of the wind speed at the site for the appropriate elevation where
natural ventilation openings may exist.

2.5.1.2 Establishing Pressure Coefficients. The next most difficult task in


the math behind wind-driven ventilation is determining the surface pressure
coefficients. Pressure coefficients are dimensionless numbers representing the
relative pressure strength experienced around the building envelope.
Chapter 24 of ASHRAE Handbook—Fundamentals emphasizes that “accurate
CHAPTER 2 | 53

TABLE 2.5 Estimating Hourly Meteorological Wind Speed (Umet) from


Annual Average Wind Speed (Uannual) (ASHRAE 2017b)
Percentage of Hourly Values that
Wind Speed Ratio (Umet/Uannual)
Exceed Umet

90% (design for poor wind strength) 0.2±0.1

1% (design for extreme wind strength) 2.5±0.4

TABLE 2.6 Terrain-Corrected Wind Speed Ratio,


Assuming Umet at Terrain Category 3
Terrain-Corrected Wind Speed Ratio
Height of Category 1— Category 2— Category 3— Category 4—
opening, ft (m) City Center Urban/Suburban Open Country Flat
40 (12.2) 0.48 0.75 1.03 1.19

75 (22.9) 0.59 0.86 1.12 1.27

150 (45.7) 0.74 1.01 1.24 1.36

300 (91.4) 0.93 1.17 1.36 1.46


Step 1. Select terrain category of site and the approximate height of the opening to determine terrain-corrected wind
ratio.
Step 2. Wind speed at height (UH) = hourly meteorological wind speed Umet times terrain-corrected wind ratio.
Authors’ note: For opening heights above 300 ft (91m), refer to Chapter 3 for more details on performing a detailed
wind study to determine wind speed.

determination of Cp can be obtained only from wind tunnel model tests of the
specific site and building or full-scale tests” (ASHRAE 2017b). The
development of a computational fluid dynamics (CFD) analysis can also
estimate pressure coefficients under a variety of simulated wind directions,
depending on the discretization of the mesh in the vicinity of the proposed
openings. In some cases, CFD can provide sufficient information to replace
physical wind tunnel tests for the purposes of examining the feasibility of a
natural ventilation scheme. See Chapter 3 for more information related to
obtaining pressure coefficients via wind tunnel and CFD.
For any given flow path, the driving force is contingent on the difference in
pressure coefficients between inlet and outlet. Typical wind-driven flow paths
create pressure differences (and thus pressure coefficient differences) between

• The windward side and the leeward side


• The windward side and the roof
• The windward side to parallel sides

At an early design phase for buildings of simple geometry, it may be


possible to take a simplified view of opening distribution using information
from Chapter 24, “Airflow Around Buildings” ASHRAE Handbook—
Fundamentals (ASHRAE 2017b). Simple sketches show the variability of
54 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 2.13 Example of computational fluid dynamics (CFD) model of exterior wind
regime.

pressure coefficients across the face of a building, considering the edge effects
at corners and roofs and the orientation of the wind direction. The chapter also
provides an estimate of surface averaged wall pressure coefficients for typical
buildings with simple rectilinear geometries. These figures are reproduced in
Appendix A2.4 for reference. Additional tables providing interpolated
information from data published as part of the AIVC Pressure Workshop can
be found in Liddament (1986).
Typically, for single-floor cross-ventilation one would review the
differential between windward and leeward side at 0° and 180° respectively as
the best-case scenario for wind-driven forces. It is important to note that for a
wall-mounted single-sided single-story ventilation scheme, the difference in
pressure coefficients is zero on any face if both the high and low openings are
within the central zone of the façade and uninfluenced by edge effects.

2.5.1.3 Buoyancy + Wind. Buoyancy- and wind-driven ventilation will


very rarely be isolated from one another, except on still days. The forces due to
wind will generally predominate over buoyancy but introduce a high level of
variability. It is strongly recommended that computer simulation methods are
used to determine performance across the breadth of outdoor temperatures and
wind speed/direction that would be experienced in a typical meteorological
year. See Chapter 5 for sample results of this type of analysis.
CHAPTER 2 | 55

TABLE 2.7 Comparison of Ventilation Rates from Reference Standards for


Two Typical Occupancies
Derived from Table 6.1 in Derived from Table 4-12 in
ASHRAE Standard 62.1 California’s Title 24
(ASHRAE 2019) (CEC 2018b)
Area-Based Ventilation Rate 0.06 cfm/ft2 (0.3 L/s/m2) 0.15 cfm/ft2 (0.76 L/s/m2)
0.4 ach 1.0 ach

Classroom Age 9+: Combined 13 cfm/person @35 p/1000 ft2 0.38 cfm/ft2 (1.93 L/s/m2)
Outdoor Air Rate (Area + People) (6.7 L/s/person)
2.5 ach
3.0 ach

Office: Combined Outdoor Air Rate 17 cfm/person @ 5p/1000 ft2 0.15 cfm/ft2 (0.76 L/s/m2)
(Area + People) (8.5 L/s/person)
1.0 ach
0.57 ach

2.6 DESIGNING SUCCESSFUL


NATURALLY CONDITIONED SPACES
As previously noted, establishing a continuous flow path is a prerequisite
for designing for either a naturally ventilated or a naturally conditioned space.
As compared to the prescriptive nature of natural ventilation design for the
basic goal of air quality, naturally conditioned spaces tend to require design
teams and occupants to have some intuitive understanding of a subset of
physics and human physiology to support the additional goal of thermal
comfort. Additionally, because natural ventilation solutions are not prevalent
in the architecture of the last half-century, the acceptability of a proposed
natural ventilation scheme requires navigating occupants’ expectations and
rebuilding the intuition of architects and engineers regarding the dependency
of indoor environmental control on outdoor conditions.

2.6.1 Designing for Thermal Comfort


The choice to pursue natural conditioning is much more difficult than just
pursuing natural ventilation. In most codes and standards, the rules around
natural ventilation are trying to achieve only equivalence with the minimum
amounts of air necessary to remove odors and indoor-generated pollutants for
human health. The prescriptive method of determining the opening size and
placement for natural ventilation in compliance with Section 6.4 of ASHRAE
Standard 62.1 (2019b) is discussed further in Chapter 5. Table 2.7 shows the
relative baseline and occupied ventilation requirements in ASHRAE Standard
62.1 (2019) and California’s Title 24 (CEC 2018b) for reference.
As a basis of comparison, a typical south-facing office or classroom might
require 6 to 10 ach of 55°F (12.8°C) air to absorb the heat from within the
56 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 2.14 Effect of the temperature difference between incoming air and room air
temperature and air change rate on the sensible cooling capacity of ventilation air (height of
space taken as 2.75 m [9 ft]).
(CIBSE 2005)

room. This large difference in supply air versus room air establishes the heat-
absorbing capacity of the airflow.
In the case of natural conditioning, the available cooling air has a
temperature with a much smaller difference between incoming air (from
outdoors) and the room temperature. Given that the heat absorption capacity of
the air is proportional to the temperature difference between indoors and
outdoors, as the outdoor temperatures rise, the required air change rate must be
increased to maintain equivalent comfort conditions under constant heat load.
This is demonstrated in Figure 2.14.
Take, for example, a space that has a sensible cooling demand of the same
3.2 Btu/h/ft2 (10 W/m2) discussed in Table 1.5, it is clear from Figure 2.14 that
as the temperature difference between outdoors and indoors decreases from
9°F (16.2 K) toward 4.5°F (8.1 K), the air change in the space must jump from
2.5 K to 5 K (4.5°F to 9°F) to accommodate the heat loads.
CIBSE (2005) recommends that natural ventilation openings be sized
based on a 5.4°F (3 K) differential. Table 2.8 calculates the sensible cooling
capacity slopes from Figure 2.14, along with the maximum cooling capacity
available for the CIBSE-recommended temperature differential. By
extrapolation, the example above would use 4.3 ach to address the 3.2 Btu/h/ft2
(10 W/m2) if the temperature differential between indoors and outdoors was
only per the CIBSE design criteria.
CHAPTER 2 | 57

TABLE 2.8 Sensible Cooling Capacity per Degree of Indoor to Outdoor


Temperature Difference in Short Height Spaces for Natural Conditioning
Air Change Rate, Sensible Cooling Capacity
Sensible Cooling Capacity
Assuming a 9 ft (2.75 m) at 5.4°F (3 K) in a
Slope Developed*
High Space 100 ft2 (9.29 m2) Space
2.5 0.328 Btuh/ft2/R (1.87 W/m2/K) 177 Btu/h (52 W)

5 0.700 Btuh/ft2/R (3.98 W/m2/K) 378 Btu/h (111 W)

10 1.40 Btuh/ft2/R (7.93 W/m2/K) 756 Btu/h (221 W)

* Slope as approximated from CIBSE (2005)

TABLE 2.9 Buildup of Heat Load Density for 100 ft2 (9.29 m2) Office Space with a
3.2 Btu/h/ft2 (10 W/m2) Sensible Cooling Available if CIBSE (2005) Guidelines on
Maximum Sensible Cooling Capacity are Applied
Description Reference Heat Load, I-P Heat Load, SI
Total Available — 3.2 Btu/h/ft2 10 W/m2

One person: seated very light ASHRAE Handbook— 245 Btu/h × 70 W ×


work, high velocity air movement Fundamentals (1-0.27)/100 ft2 =
(ASHRAE 2017b), (1-0.27)/9.29 m2 =
1.8 Btu/h/ft2
page 18.4, Table 1
5.5 W/m2

One tablet PC, average 15 min ASHRAE Handbook— 36 W × 36 W/9.29 m2 =


peak power consumption Fundamentals 3.41 Btu/h /100ft2=
(ASHRAE 2017b), 3.9 W/m2
1.2 Btu/h/ft2
page 18.12, Table 8A

Differential available for lighting — 0.2 Btu/h/ft2 0.6 W/m2


and solar loads
Or 0.060 W/ft2

For reference, a 100 ft2 (9.29 m2) office space with this heat load would be
limited to the components in Table 2.9. Observe that there is no spare capacity
available to accommodate solar heat gain or lighting at this heat load density
target. This short exercise shows the significant limitations on internal heat
gains if natural conditioning is the desired comfort cooling scheme.
The success of naturally conditioned spaces is heavily contingent on
outdoor air conditions and the control of heat gains within the occupied space.
Chapters 1 and 3 address how one can progress a feasibility study on these two
topics. As an example of a typical space, Figures 2.15 and 2.16 demonstrate
the internal loads of this guide’s standard two-person office model (as
described in Appendix A1.1), as modified to reflect the influence of solar heat
gain coefficient (SHGC).
58 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 2.15 Breakdown of cooling loads in the summer for different SHGC values for a
“traditional” design.

FIGURE 2.16 Breakdown of cooling loads in the summer for different SHGC values for a
high efficiency design.

When the climate in conjunction with the owner’s need for high
internal heat loads does not support natural conditioning throughout the
entire year, design teams might consider mixed-mode ventilation, as
discussed in Section 2.2, “Mixed-Mode Ventilation.” Failing that, it would
be clear that the internal and solar heat gains exceed the capacity of a
natural conditioning system to achieve indoor comfort, and the design team
should revert to a comfort cooling system type that can. As a rule of thumb,
CHAPTER 2 | 59

FIGURE 2.17 Rule of thumb comparison of heat absorbing capacities of various comfort-
conditioning systems.
© Arup.
Courtesy: Arup.

Figure 2.17 shows the typical extent of peak heat absorption based on
comfort cooling type.

2.6.2 Designing for Occupant Control


Before Willis Carrier invented air conditioning in 1902, all indoor spaces
were naturally ventilated and naturally conditioned. There have now been
multiple generations of people in the developed world who have come to
expect that air conditioning should hold spaces at a comfortable temperature
automatically without their involvement. In many parts of the world, air
conditioning as “comfort on demand” is seen as an indicator of upward social
mobility. Thus, as natural conditioning is reintroduced to modern architecture,
we need to acknowledge that further education is required, both on the part of
the design professional and for the user. The success of most naturally
conditioned spaces heavily relies on occupants controlling the façade openings
to open them at the right time.
People used a variety of methods to cool themselves off based on
understanding the human body’s physiological mechanisms for heat rejection.
(See Table 2.10.) Some of these means also have manifestations in built
construction, as noted in several excellent diagrams and case studies of
historical precedents for the design of naturally ventilated spaces by Passe and
Battaglia (2015).
One characteristic of all these historic schemes is that people were actively
engaged in a change action to achieve habitability associated with human
comfort. In the case of fixed features, as noted in Table 2.10, occupants would
60 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

activate the feature at the right time of the year. Where it was not appropriate to
build such complicated physical infrastructure, certain indigenous tribes would
seasonally relocate, while others would move from insulated winter houses to
more open summer houses. Seasonal changes in clothing were the norm.
Human physiology did not change when air conditioning was invented, but
the psychology of one’s agency concerning the environment did. Thus, it is
crucial the design team and the client and prospective tenants discuss what
active participation is needed for a natural conditioning scheme to successfully
deliver human comfort. The need for occupant participation is enshrined in the
name ASHRAE Standard 55-2017 (ASHRAE 2017a) gives to the adaptive
comfort standard: “Determining Acceptable Thermal Conditions in Occupant-
Controlled Naturally Conditioned Spaces.” Furthermore, Section 5.4.1 of the
standard suggests: “Representative occupants are free to adapt their clothing to
the indoor and/or outdoor thermal conditions within a range at least as wide as
0.5 to 1.0 clo” (ASHRAE 2017a). See Table 2.11 for representative clothing
ensembles that demonstrate this 0.5 to 1.0 range. A greater variety of clothing
types can be found in Website Appendix WA6.3, “Engineered Natural
Ventilation System Analysis Sample Results,” accessible at ashrae.org/
naturalventilation.
The following basic natural conditioning principles bear repeating:

• The air is unlikely to be cooler indoors than it is outdoors, except when


passive natural cooling measures are applied (See Table 2.10).
• Well-designed and well-managed naturally conditioned office and
classroom spaces typically have measurable indoor air temperatures as
low as 3°F to 6°F (1.6°C to 3.3°C) warmer than the external dry-bulb
temperature due to heat pickup in the room.
• Permanent occupants of naturally conditioned spaces should be
allowed to alter their clothing throughout the day and across all
seasons.
• Occupants of naturally conditioned spaces should have access to and
discretionary control over an operable element in the façade that can
modulate how much airflow comes in from the outdoors.

The prediction of the range of indoor temperatures under natural


conditioning is a crucial part of obtaining client approval to pursue a natural
conditioning scheme. Simulation tools can demonstrate how predicted
temperatures compare to the allowable human comfort range across the year
and throughout the day. As described in depth in Chapter 6, ASHRAE
Standard 55 (ASHRAE 2017a) governs how the comfort range is defined,
depending on the occupants’ ability to control localized air speed and/or their
ability to control the natural conditioning openings.
The operational success of a natural ventilation or natural conditioning
scheme relies on the occupants proactively working with the building
components and/or by adjusting personal clothing to accommodate the
fluctuations in internal temperatures due to variability in the outdoor
conditions. While it is possible to have automated natural ventilation devices,
CHAPTER 2 | 61

TABLE 2.10 Cooling Mechanisms Available in a World without Air Conditioning


Physiological
Examples How it Works
Mechanism

Convective Heat Cross ventilation See Section 2.5.


Transfer: Skin to Air—
Basic natural ventilation
Wind towers, wind Single-sided inlets rise above the roof as an obstruction to
drivers to cause
catchers the clear air movement, taking advantage of the wind to
movement of uncooled
drive air down directly into the occupied area or over a body
ambient air
of water, if active cooling is available

Fire-driven chimneys Heat source drives upward buoyancy of air, taking


advantage of “stack effect” (see Section 2.4). Typically, the
Solar-driven chimneys
intake air is ambient air condition but may also pass through
earth tunnels for precooling in some indigenous
constructions.

Convective Heat Earth tunnel intakes Air during the day passes across earthen walls or through a
Transfer: Skin to Air— battery of ~6 in. (150 mm) rocks to absorb cooling before
Rock store cooling
Passive natural cooling entry to the space. This takes advantage of nighttime air to
measures used in precool the thermal mass. The thermal mass then
conjunction with natural convectively cools the air before it enters the space.
ventilation drivers
Fountain courtyards Evaporating water into hot dry air will reduce its temperature
and increase subsequent heat absorption capacity. It will
Cool towers also increase its relative humidity, thereby reducing the
Underground canals capacity for evaporative heat loss (see below). Beneficial in
(e.g., Persian qanat) dry climates with excellent flow-through air patterns to
prevent stagnation of moist air. Typically used in conjunction
Soaked linens on window with another air-driving natural ventilation force.
openings

Combined Convective Hand Fans Under low activity, evaporative heat loss is fairly constant
and Evaporative Heat until ambient air temperature is about 86°F (30°C), above
Ceiling Fans
Loss: Skin to Air— which active sweating begins to assist with cooling for
Typically, a localized Mist on skin thermoregulation. Once air temperature rises above the
phenomenon requiring typical skin temperature of 93°F (34°C), convection is
an active air-moving actually heating the skin. But if the relative humidity of the air
element is low enough, the air movement assists with the evaporation
of sweat. As water turns to vapor, it absorbs heat from the
skin surface, thus cooling the occupant.

Radiative Heat Transfer: Cave dwellings Construction material is cooled at night or heated during the
Skin to Surface— day. Warmth or coolth is stored in the thermal mass.
Typically embedded in Thick concrete, adobe, or Internally facing surfaces become a radiative cooling or
brick walls
built construction heating surface interacting with the occupants typically 8 to
Buried rooms 12 hours later in a “discharge” cycle.

Below floor air plenums


or water passageways
62 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 2.11 Representative Clothing Ensembles Demonstrating the Range of


Clothing Adjustability Required for Natural Conditioning under
ASHRAE Standard 55’s Adaptive Comfort Standard
Ensemble Type Near 0.5 clo Near 1.0 clo
Typical men’s Light trousers and short-sleeve shirt, socks, Trousers, long-sleeved shirt, long-sleeved
ensemble shoes sweater, T-shirt, socks, shoes

Typical women’s Knee-length skirt, short-sleeved shirt, Knee-length skirt, long-sleeved shirt,
ensemble pantyhose, sandals half-slip, pantyhose, suit jacket, shoes
* Table adapted using information from ASHRAE (2017b).

ASHRAE Standard 55 (ASHRAE 2017a) requires occupant control if the


adaptive comfort model is to be used to expand the range of acceptable
temperatures. Additionally, the success relies somewhat on the occupants’
psychological acceptance of the trade-off between higher quantities of fresh air
and a wider range of indoor temperatures. Further description of the influences
captured within the adaptive comfort model is included in Chapter 7. It is often
heard that natural ventilation is a “passive” system because it does not involve
large motors for HVAC equipment, but in reality, it is the occupants who
become the “active” participant who adjusts the building fabric system to
attain their own comfort. The occupants also become the feedback sensors for
the control system. As occupants feel warm, they will take action to open a
window. As they start to cool, they will close the window. In a naturally
ventilated space, this control loop is done independently of the actions of the
building management system.
Reinforcing adaptive comfort research from Humphreys and Nicol (2002)
and Brager and de Dear (2000), post-occupancy findings by Leaman and
Bordass (2007) demonstrate that users in “green” buildings show a higher level
of tolerance (i.e., “forgiveness”) for out-of-range temperature conditions as
compared to ratings for similar situations within a fully air-conditioned
environment. They find that, “[i]deally, individuals should be able to deal with
the source of their own discomfort easily and quickly,” and when this is not the
case, the building’s performance is judged on its “effectiveness, speed and
demonstrable effect” of its corrective response (Leaman and Bordass 2007).
Lastly, Leaman and Bordass (2007) state that users are more forgiving in their
expectations when they understand how the building is supposed to work.
These are all critical elements to consider when designing natural conditioning
and mixed-mode systems. In particular, it is important to ensure that any
automated control of windows that happens concurrently with occupant-
controlled windows must be carefully organized in the sequence of operations
to avoid overriding the occupant’s instincts for manual control and adaptive
comfort that is present in simply controlled naturally ventilated spaces. See
Chapter 7 for more discussion on the topic of automated controls for natural
ventilation.
CHAPTER 2 | 63

2.6.3 Designing Based on Client Engagement


Natural conditioning is often seen as an atypical solution for offices and
classrooms in much of North America because few geographic locations can
support natural conditioning for 100% of daytime hours. The client must be
brought into the design team’s exploration of natural conditioning as soon as
the outdoor air conditions are deemed feasible.
There are multiple parties involved in taking the idea of natural ventilation
from concept to construction. CIBSE Applications Manual AM10, Natural
Ventilation in Non-Domestic Buildings (CIBSE 2005) clearly lays out the
principal responsibilities and interactions of the design team members, and the
recommendations can be found in the table within Website Appendix WA2.1,
“equations for Simple Buoyancy-Driven Ventilation,” accessible at ashrae.org/
naturalventilation.
Because a building is a substantial investment, the client should have the
right to determine whether a natural conditioning or a natural ventilation
scheme enhances the value or productivity of the investment. For instance, in
the case of a classroom, does the natural ventilation scheme improve student
learning outcomes? For an office, are the workers more productive? For a
speculative office or residential building, can a developer fetch a higher sales
rate, lease value, and/or occupancy rate because of this feature? If the benefits
are available, they must be weighed against the differential in the first cost
between operable windows as compared to a fixed façade with an HVAC
system.
Similarly, a review of operational costs should be conducted. Is there an
additional cost of security because of a manual window that might be left
open? Are there extra janitorial costs associated with the direct entry of air into
the space without dust filters? Is there a risk associated with lowered
productivity due to higher indoor air temperatures? Can energy cost savings be
accrued?
All these discussions must be grounded in a rigorous analysis as the owner
or developer must be aware of the benefits, risks, and trade-offs for pursuing a
natural ventilation scheme.

2.7 REFERENCES
ASHRAE. 2017a. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2017b. ASHRAE handbook—Fundamentals. Peachtree Corners,
GA: ASHRAE.
ASHRAE. 2019a. ANSI/ASHRAE/IES Standard 90.1-2019, Energy standard
for buildings except low-rise residential buildings. Peachtree Corners, GA:
ASHRAE.
ASHRAE. 2019b. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
Boyle. 1899a. Natural and artificial methods of ventilation. London: Robert
Boyle & Son, Ltd.
64 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Boyle. 1899b. Natural ventilation. London: Robert Boyle & Son, Ltd.
Brager, G. with Center for the Built Environment (CBE). 2008. Natural
ventilation & mixed-mode buildings. Presented at ASHRAE Golden Gate
Chapter Meeting, Oakland, CA, January 2008.
Brager, G. and R. de Dear. 2000. A standard for natural ventilation. ASHRAE
Journal 42(10):21–23, 25.
CEC. 2018a. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
CEC. 2018b. 2019 Nonresidential Compliance Manual for the 2019 Building
Energy Efficiency Standards: Title 24, Part 6 and associated
Administrative Regulations in Part 1. CEC-400-2018-018-CMF.
Sacramento, CA: California Energy Commission.
CIBSE. 2000. CIBSE applications manual AM13:2000, Mixed mode
ventilation. London: The Chartered Institute of Building Services
Engineers.
CIBSE. 2005. CIBSE applications manual AM10, Natural ventilation in non-
domestic buildings. London: The Chartered Institute of Building Services
Engineers.
DOE. n.d. U.S. average annual wind speed at 30 m. Washington, DC: Wind
Energy Technologies Office at the U.S. Department of Energy. https://
windexchange.energy.gov/maps-data/325.
Humphreys, M.A, and F. Nicol. 2002. Adaptive thermal comfort and
sustainable thermal standards for buildings. Energy and Buildings
34(6):563–72.
Leaman, A., and B. Bordass. 2007. Are users more tolerant of “green”
buildings? Building Research & Information 35(6): 662–73.
Liddament, M. 1986. Air infiltration calculation techniques: An application
guide. Coventry, U.K.: Air Infiltration and Ventilation Centre.
MIT. n.d. CoolVent, Version 1.0.6150.40690. Boston, MA: Building Technology
Research Group at Massachusetts Institute of Technology. http://
coolvent.mit.edu/.
McConahey, E. 2011. Natural ventilation: Who, what, when, where, why, and
how. Consulting Specifying Engineer Magazine November 2011: 24–29.
Passe, U., and F. Battaglia. 2015. Designing spaces for natural ventilation: An
architect’s guide. New York: Routledge.
EQUATIONS FOR SIMPLE
BUOYANCY-DRIVEN
VENTILATION
This appendix summarizes the equations used to calculate airflow through
natural ventilation openings in a simple buoyancy-driven system.
The height of the neutral plane is denoted as per Li et al. (2007):
2

h n = -------------2- h (A2.1.1)
1+
where
hn = height of neutral plane above the lowest edge of the lowest
opening, m (ft)
h = height between centerline of upper and lower openings, m (ft)
 = opening ratio,  = cduau/cdlal
CDU = discharge coefficient for upper opening per ASHRAE
Handbook—Fundamentals (ASHRAE 2017)
Unidirectional flow: typically 0.65
Bidirectional flow:
• I-P: 0.40 + |Ti–To|
• SI: 0.40 + 0.0045|Ti–To|
CDL = discharge coefficient for lower opening as per ASHRAE
Handbook—Fundamentals (ASHRAE 2017)
Unidirectional flow: typically 0.65
Bidirectional flow:
• I-P: 0.40 + |Ti–To|
• SI: 0.40 + 0.0045|Ti–To|
AU = free area of upper opening, m (ft)
AL = free area of lower opening, m (ft)

As noted above, increasing the size of a single opening adjusts the height of
the neutral plane due to the opening ratio. It also slightly increases the total
airflow as noted in Figure A2.1.2.

65
66 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A2.1.1 Neutral plane for equal sized openings.


(CIBSE 2005)

FIGURE A2.1.2 Increase in airflow by increasing area of one opening.


(ASHRAE 2017)

Mathematically, the flow can be estimated using Equation A2.1.2


(ASHRAE 2017):

Q = 60C D A  2 gH NPL  T i – T o   T i (A2.1.2 I-P)

where
Q = airflow rate, cfm
CD = discharge coefficient for opening (ASHRAE 2017)
APPENDIX A2.1 | 67

Unidirectional flow: typically 0.65


Bidirectional: 0.40 + |Ti–To|
A = free area of smallest opening as corrected by Figure A2.1.2, ft2
g = gravitational constant, 32.2 ft/s
HNPL = height from midpoint of lower opening to the neutral plane (NPL), ft
Ti = absolute indoor temperature, °R
To = absolute outdoor temperature, °R

Q = C D A  2 gH NPL  T i – T o   T i (A2.1.2 SI)

where
Q = airflow rate, m3/s
CD = discharge coefficient for opening (ASHRAE 2017)
Unidirectional flow: typically 0.65
Bidirectional flow: 0.40 + 0.0045|Ti–To|
A = free area of smallest opening as corrected by Figure A2.1.2, m2
g = gravitational constant, 9.81 m/s
HNPL = height from midpoint of lower opening to the NPL, m
Ti = absolute indoor temperature, K
To = absolute outdoor temperature, K

When outdoor temperatures are warmer than indoor temperatures (i.e., To >
Ti), the subscripts for each temperature variable shall be reversed (i.e., all
outdoor subscripts become indoor and vice versa).

A2.1.1 REFERENCES
ASHRAE. 2017. ASHRAE handbook—Fundamentals. Peachtree Corners, GA:
ASHRAE.
CIBSE. 2005. CIBSE applications manual AM10: Natural ventilation in non-
domestic buildings. London: Chartered Institute of Building Services
Engineers.
Li, R., A. Pitts, and Y. Li. 2007. Buoyancy-driven Natural Ventilation of a
Room with Large Openings. Proceedings: Buildings Simulation 2007:
984–91.
EQUATIONS FOR
WIND-DRIVEN
VENTILATION
This appendix summarizes the equations used to calculated airflow through
natural ventilation openings in a simple wind-driven system.
The section below excerpts relevant information from Chapter 16,
“Ventilation and Infiltration” of ASHRAE Handbook—Fundamentals
(ASHRAE 2017) to support a hand calculation to confirm natural ventilation
flow through an opening in a wind driven scheme. The governing equation
determining the flow through an opening is

Q = 88.0C v AU (A2.2.1 I-P)

Q = C v AU (A2.2.1 SI)

where
Q = airflow rate, cfm (m3/s)
Cv = effectiveness of openings (typically for windows, 0.5 to 0.6 for
perpendicular winds and 0.25 to 0.35 for diagonal winds
(See the following note on estimating pressure coefficients, and
Appendix A2.4 for coefficients in other configurations.)
A = the sum of free areas of the inlet openings based on assumed
airflow path, ft2 (m2)
U = wind speed at opening, mph (m/s) (see Appendix A2.3 for how to
determine speed, UH)

Note: when openings are near the edge of a wall in the downwind space,
the discharge coefficients increase to 0.7 and 0.8, with larger values for bigger
openings (10% to 20% of the wall area.) For openings similar in size to the
cross-section of the downstream space, discharge coefficients of 0.8 to 0.9 are
possible.
As noted in Section 2.4.2, “Complex Geometries and the Determination of
Neutral Plane,” when more complex geometries are being explored, a

69
70 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

computer analysis tool such as CoolVent by the Building Technology


Research Group at Massachusetts Institute of Technology (MIT) or another
bulk airflow modeling tool is strongly recommended over performing these
calculations by hand.

A2.2.1 REFERENCES
ASHRAE. 2017. Chapter 16, Ventilation and infiltration. In ASHRAE
handbook—Fundamentals. Peachtree Corners, GA: ASHRAE.
MIT. n.d. CoolVent, Version 1.0.6150.40690. Boston, MA: Building
Technology Research Group at Massachusetts Institute of Technology.
http://coolvent.mit.edu/.
EQUATIONS FOR
TERRAIN-CORRECTED
WIND RATIOS
Chapter 24, “Airflow Around Buildings” of ASHRAE Handbook—
Fundamentals (ASHRAE 2017) advises that the wind speed at the subject
building height UH must be adjusted to account for the differences in terrain
categories, based on Equation A2.3.1.
 met amet H a
U H = U met  -----------  ---- (A2.3.1)
 H met   

Table A2.3.1 is used to determine the necessary parameters to adjust height


from typical meteorological readings at 100 ft (30 m) to the height of a natural
ventilation opening.
TABLE A2.3.1 Atmospheric Boundary Layer Parameters (ASHRAE 2017b)
Terrain Layer Thickness
Description Exponent a
Category δ, ft (m)
1 Large city centers, in which at least 50% of buildings 0.33 1500 (460)
are higher than 80 ft (25 m), over a distance of at least
0.5 mi (0.8 km) or 10 times the height of the structure
upwind, whichever is greater

2 Urban and suburban areas, wooded areas, or other 0.22 1200 (370)
terrain with numerous closely spaced obstructions
having the size of single-family dwellings or larger,
over a distance of at least 0.5 mi (0.8 km) or 10 times
the height of the structure upwind, whichever is greater

3 Open terrain with scattered obstructions having heights 0.14 900 (270)
generally less than 30 ft (9 m), including flat open
country typical of meteorological station surroundings

4 Flat, unobstructed areas exposed to wind flowing over 0.10 700 (210)
water for at least 1 mi (1.6 km), over a distance of
1500 ft (460 m) or 10 times the height of the structure
inland, whichever is greater

71
72 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

To expedite the calculation process, Table A2.3.1 provides the terrain-


corrected wind speed ratio UH/Umet assuming that Umet is derived from a
Terrain 3 location. This allows the designer that has use the simplified equation

U H = U met  terrain-corrected wind speed ratio (A2.3.2)

A2.3.1 REFERENCE
ASHRAE. 2017. Chapter 24, “Airflow around buildings.” In ASHRAE
handbook—Fundamentals. Peachtree Corners, GA: ASHRAE.
PRESSURE COEFFICIENTS FOR
SIMPLE RECTILINEAR
GEOMETRIES
This appendix includes figures from Chapter 24, “Airflow Around
Buildings,” of ASHRAE Handbook—Fundamentals (ASHRAE 2017) for
simple rectilinear geometries using graphical methods. The data provided is
used in calculating the wind effects to buildings and openings.

73
74 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A2.4.1 LOW-RISE BUILDINGS

FIGURE A2.4.1 Local pressure coefficients for a low-rise building with varying wind direction.
(ASHRAE 2017)
APPENDIX A2.4 | 75

FIGURE A2.4.2 Surface-averaged wall pressure coefficients for low-rise buildings.


(ASHRAE 2017)

FIGURE A2.4.3 Local roof pressure coefficients for roof of a low-rise building.
(ASHRAE 2017)
76 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A2.4.2 HIGH-RISE BUILDINGS

FIGURE A2.4.4 Local pressure coefficients (Cpx100) for a high-rise building with varying
wind direction.
(ASHRAE 2017)

FIGURE A2.4.5 Surface-averaged wall pressure coefficients for high-rise buildings.


(ASHRAE 2017)
APPENDIX A2.4 | 77

FIGURE A2.4.6 Surface-averaged roof pressure coefficient for tall buildings.


(ASHRAE 2017)

A2.4.3 REFERENCE
ASHRAE. 2017. Chapter 24, “Airflow around buildings.” In ASHRAE
handbook—Fundamentals. Peachtree Corners, GA: ASHRAE.
INTERPOLATED PRESSURE
COEFFICIENTS TABLES
The data presented in this appendix are excerpted from Air Infiltration
Calculation Techniques: An Applications Guide (Liddament 1986). Note:
while the original section, equation, and figure numbering from reference
document has been retained, this passage has been formatted to conform to
this design guide.
The data presented in this section are based on the interpolation of
published material, much of which was presented at the AIVC’s Wind Pressure
Workshop (AIVC 1984). The values presented must only be regarded as
approximate and therefore, if more accurate design data are required, recourse
to specific wind tunnel or full-scale measurements will have to be considered.
The intention of these data sets is to provide the user with an indication of the
range of pressure coefficient values which might be anticipated for various
building orientations and for various degrees of shielding. The data presented
is as follows:
Tables 6.2.1 to 6.2.6 cover low-rise buildings of typically no more than
three stories. Three degrees of shielding are considered; these are:
1. Open countryside, no obstructions—exposed.
2. Rural surrounding, some obstructions—semi-sheltered.
3. Urban building surrounded on all sides by obstructions of similar
size—sheltered.

For each condition, the pressure coefficient is expressed as a single average


value for each face of the building, where the reference pressure height for
wind speed is taken as the building height.
Data presented for each face and for the roof surfaces, where roof pitch
angles of
• <10°
• 11–30°
• and >30°
are considered.

79
80 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A2.5.1 Vertical distribution of mean wind pressure coefficients for various
surrounding obstruction heights.
(Liddament 1986)

Tables 6.2.1 to 6.2.3 cover data for buildings of length to width ratios of
1:1 and Tables 6.2.4 cover data for length to width ratios of 2:1.
Data depicted in Figure 6.2.1 contain results by Bowen (1976) showing the
vertical dependency of pressure coefficient for tall buildings.
Caution: The material and data presented in this section are intended solely
as a guide to current knowledge on air infiltration and related topics.The
information contained herein does no supersede [sic] any advice or
requirement given in any national codes or regulations, neither is its suitability
for a particular application guaranteed. No responsibility can be accepted for
the use of data presented in this publication.
APPENDIX A2.5 | 81

FIGURE A2.5.2 Vertical distribution of mean wind pressure coefficients for various
surrounding obstruction heights.
(Liddament 1986)
82 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A2.5.1 Wind Pressure Coefficient Data (Liddament 1986)


Low–rise buildings (up to 3 storeys)
Length to width ratio: 1:1.
Shielding condition: exposed.
Wind speed reference level: building height.

Wind Angle
Location 0 45 90 135 180 225 270 315

Face 1 0.7 0.35 –0.5 –0.4 –0.2 –0.4 –0.5 0.35

Face 2 –0.2 –0.4 –0.5 0.35 0.7 0.35 –0.5 –0.4

Face 3 –0.5 0.35 0.7 0.35 –0.5 –0.4 –0.2 –0.4

Face 4 –0.5 –0.4 –0.2 –0.4 –0.5 0.35 0.7 0.35

“Weighted” mean –.05 –.03 –.05 –.03 –.05 –.03 –.05 –0.03

Front –0.8 –0.7 –0.6 –0.5 –0.4 –0.5 –0.6 00.7


Roof
Rear –0.4 –0.5 –0.6 –0.7 –0.8 –0.7 –0.6 –0.5
(< 10° pitch)
Average –0.6 –0.6 –0.6 –0.6 –0.6 –0.6 –0.6 –0.6

Front –0.4 –0.5 –0.6 –0.5 –0.4 –0.5 –0.6 –0.5


Roof
Rear –0.4 –0.5 –0.6 –0.5 –0.4 –0.5 –0.6 –0.5
(11–30° pitch)
Average –0.4 –0.5 –0.6 –0.5 –0.4 –0.5 –0.6 –0.5

Front 0.3 –0.4 –0.6 –0.4 –0.5 –0.4 –0.6 –0.4


Roof
Rear –0.5 –0.4 –0.6 –0.4 0.3 –0.4 –0.6 –0.4
(>30° pitch)
Average –0.1 –0.4 –0.6 –0.4 –0.1 –0.4 –0.6 –0.4
APPENDIX A2.5 | 83

TABLE A2.5.2 Wind Pressure Coefficient Data (Liddament 1986)


Low–rise buildings (up to 3 storeys).
Length to width ratio: 1:1
Shielding condition: surround by obstructions equivalent to
half the height of the building.
Wind speed reference level: building height.

Wind Angle
Location 0 45 90 135 180 225 270 315

Face 1 0.4 0.1 –0.3 –0.35 –0.2 –0.35 –0.3 –0.1

Face 2 –0.2 –0.35 –0.3 0.1 0.4 0.1 –0.3 –0.35

Face 3 –0.3 0.1 0.4 0.1 –0.3 –0.35 –0.2 –0.35

Face 4 –0.3 00.35 –0.2 –0.35 –0.3 0.1 0.4 0.1

“Weighted” mean –0.1 –0.13 –0.1 –.13 –0.1 –.13 –0.1 –.13

Front –0.6 –0.5 –0.4 –0.5 –0.6 –0.5 –0.4 –0.5


Roof
Rear –0.6 –0.5 –0.4 –0.5 –0.6 –0.5 –0.4 –0.5
(< 10° pitch)
Average –0.6 –0.5 –0.4 –0.5 –0.6 –0.5 –0.4 –0.5

Front –0.4 –0.5 –0.6 –0.5 –0.4 –0.5 –0.6 –0.5


Roof
Rear –0.4 –0.5 –0.6 –0.5 –0.4 –0.5 –0.6 –0.5
(11–30° pitch)
Average –0.4 –0.5 –0.6 –0.5 –0.4 –0.5 –0.6 –0.5

Front 0.3 –0.5 –0.6 –0.5 –0.5 –0.5 –0.6 –0.5


Roof
Rear –0.5 –0.5 –0.6 –0.5 0.3 –0.5 –0.6 –0.5
(>30° pitch)
Average –0.1 –0.5 –0.6 –0.5 –0.1 –0.5 –0.6 –0.5
84 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A2.5.3 Wind Pressure Coefficient Data (Liddament 1986)


Low–rise buildings (up to 3 storeys).
Length to width ratio: 1:1
Shielding condition: Surrounded by obstructions equal
to height of the building.
Wind speed reference level: building height.

Wind Angle
Location 0 45 90 135 180 225 270 315

Face 1 0.2 0.05 –0.25 –0.3 –0.25 –0.3 –0.25 0.05

Face 2 –0.25 –0.3 –0.25 0.05 0.2 0.05 –0.25 –0.3

Face 3 –0.25 0.05 0.2 0.05 –0.25 –0.3 –0.25 –0.3

Face 4 –0.25 –0.3 –0.25 –0.3 –0.25 0.05 0.2 0.05

“Weighted” mean –0.15 –0.15 –0.15 –0.15 –0.15 –0.15 –0.15 –0.15

Front –0.5 –0.5 –0.4 –0.5 –0.5 –0.5 –0.4 –0.5


Roof
Rear –0.5 –0.5 –0.4 –0.5 –0.5 –0.5 –0.4 –0.5
(< 10° pitch)
Average –0.5 –0.5 –0.4 –0.5 –0.5 –0.5 –0.4 –0.5

Front –0.3 –0.4 –0.5 –0.4 –0.3 –0.4 –0.5 –0.4


Roof
Rear –0.3 –0.4 –0.5 –0.4 –0.3 –0.4 –0.5 –0.4
(11–30° pitch)
Average –0.3 –0.4 –0.5 –0.4 –0.3 –0.4 –0.5 –0.4

Front 0.25 –0.3 –0.5 –0.3 –0.4 –0.3 –0.5 –0.3


Roof
Rear –0.4 –0.3 –0.5 –0.3 0.25 –0.3 –0.5 –0.3
(>30° pitch)
Average –0.08 –0.3 –0.5 –0.3 –0.08 –0.3 –0.5 –0.3
APPENDIX A2.5 | 85

TABLE A2.5.4 Wind Pressure Coefficient Data (Liddament 1986)


Low–rise buildings (up to 3 storeys)
Length to width ratio: 2:1
Shielding condition: Exposed.
Wind speed reference level: building height.

Wind Angle
Location 0 45 90 135 180 225 270 315

Face 1 0.5 0.25 –0.5 –0.8 –0.7 –0.8 –0.5 0.25

Face 2 –0.7 –0.8 –0.5 0.25 0.5 0.25 –0.5 –0.8

Face 3 –0.9 0.2 0.6 0.2 –0.9 –0.6 –0.35 –0.6

Face 4 –0.9 –0.6 –0.35 –0.6 –0.9 0.2 0.6 0.2

“Weighted” mean –0.48 –0.17 –0.06 –0.17 –0.48 –0.17 –0.06 –0.17

Front –0.7 –0.7 –0.8 –0.7 –0.7 –0.7 –0.8 –0.7


Roof
Rear –0.7 –0.7 –0.8 –0.7 –0.7 –0.7 –0.8 –0.7
(< 10° pitch)
Average –0.7 –0.7 –0.8 –0.7 –0.7 –0.7 –0.8 –0.7

Front –0.7 –0.7 –0.7 –0.6 –0.5 –0.6 –0.7 –0.7


Roof
Rear –0.5 –0.6 –0.7 –0.7 –0.7 –0.7 –0.7 –0.6
(11–30° pitch)
Average –0.6 –0.65 –0.7 –0.65 –0.6 –0.65 –0.7 –0.65

Front 0.25 0 –0.6 –0.9 –0.8 –0.9 –0.6 0

Roof (>30° pitch) Rear –0.8 –0.9 –0.6 0 0.25 0 –0.6 –0.9

Average –0.18 –0.45 –0.6 –0.45 –0.18 –0.45 –0.6 –0.45


86 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A2.5.5 Wind Pressure Coefficient Data (Liddament 1986)


Low–rise buildings (up to 3 storeys).
Length to width ratio: 2:1
Shielding condition: semi sheltered
Wind speed reference level: building height.

Wind Angle
Location 0 45 90 135 180 225 270 315

Face 1 0.5 0.25 –0.5 –0.8 –0.7 –0.8 –0.5 0.25

Face 2 –0.7 –0.8 –0.5 0.25 0.5 0.25 –0.5 –0.8

Face 3 –0.9 0.2 0.6 0.2 –0.9 –0.5 –0.35 –0.6

Face 4 –0.9 –0.6 –0.35 –0.6 –0.9 0.2 0.4 0.2

“Weighted” mean –0.48 –0.17 –0.06 –0.17 –0.48 –0.17 –0.06 –0.17

Front –0.7 –0.7 –0.8 –0.7 –0.7 –0.7 –0.8 –0.7


Roof
Rear –0.7 –0.7 –0.8 –0.7 –0.7 –0.7 –0.8 –0.7
(< 10° pitch)
Average –0.7 –0.7 –0.8 –0.7 –0.7 –0.7 –0.8 –0.7

Front –0.7 –0.7 –0.7 –0.6 –0.5 –0.6 –0.7 –0.7


Roof
Rear –0.5 –0.6 –0.7 –0.7 –0.7 –0.7 –0.7 –0.6
(11–30° pitch)
Average –0.6 –0.65 –0.7 –0.65 –0.6 –0.65 –0.7 –0.65

Front 0.25 0 –0.6 –0.9 –0.8 –0.9 –0.6 0


Roof
Rear –0.8 –0.9 –0.6 0 0.25 0 –0.6 –0.9
(>30° pitch)
Average –0.18 –0.45 –0.6 –0.45 –0.18 –0.45 –0.6 –0.45
APPENDIX A2.5 | 87

TABLE A2.5.6 Wind Pressure Coefficient Data (Liddament 1986)


Low–rise buildings (up to 3 storeys).
Length to width ratio: 2:1
Shielding condition: sheltered
Wind speed reference level: building height.

Wind Angle
Location 0 45 90 135 180 225 270 315

Face 1 0.06 –0.12 –0.2 –0.38 –0.3 –0.38 –0.2 –0.12

Face 2 –0.3 –0.38 –0.2 –0.12 0.06 –0.12 –0.2 –0.38

Face 3 –0.3 0.15 0.18 0.15 –0.3 –0.32 –0.2 –0.32

Face 4 –0.3 –0.32 –0.2 –0.32 –0.3 0.15 0.18 0.15

“Weighted” mean –0.24 –0.14 –0.07 –0.14 –0.24 –0.14 –0.07 –0.14

Front –0.49 –0.46 –0.41 –0.46 –0.49 –0.46 –0.41 –0.46


Roof
Rear –0.49 –0.46 –0.41 –0.46 –0.49 –0.46 –0.41 –0.46
(< 10° pitch)
Average –0.49 –0.46 –0.41 –0.46 –0.49 –0.46 –0.41 –0.46

Front –0.49 –0.46 –0.41 –0.46 –0.4 –0.46 –0.41 –0.46


Roof
Rear –0.4 –0.46 –0.41 –0.46 –0.49 –0.46 –0.41 –0.46
(11–30° pitch)
Average –0.45 –0.46 –0.41 –0.46 –0.45 –0.46 –0.41 –0.46

Front 0.06 –0.15 –0.23 –0.6 –0.42 –0.6 –0.23 –0.15


Roof
Rear –0.42 –0.6 –0.23 –0.15 –0.06 –0.15 –0.23 –0.6
(>30° pitch)
Average –0.18 –0.4 –0.23 –0.4 –0.18 –0.4 –0.23 –0.4

A2.5.1 REFERENCES
AIVC. 1984. Technical Note AIC 13.1, 1984 Wind Pressure Workshop
Proceedings. Bracknell, U.K.: Air Infiltration Centre.
Bowen, A.J.1976. A wind tunnel investigation using simple building models to
obtain mean surface wind pressure coefficients for air infiltration
purposes. Report LTR LA 20N. Ottawa, Canada: National Aeronautical
Establishment, National Research Council Canada.
Liddament, M. 1986. Air infiltration calculation techniques: An application
guide. Coventry, U.K.: Air Infiltration and Ventilation Centre.
CASE STUDY—
A.P.C. STORE,
HOLLYWOOD, CA

FIGURE CS1.1 Lengthwise section through the building,

FIGURE CS1.2 Building exterior and the low-level and high-level openings

• Architect: Warren Design


• Climate: Los Angeles, CA

The A.P.C. store on Melrose Avenue in Hollywood, CA, is designed to be


conditioned as naturally as possible. The store is a single-story unit with high-
vaulted ceilings. The main space, called the church room, has a high-vaulted
ceiling. The two spaces at opposite ends of the church room have high ceilings
with skylights for natural daylight. The use of clear glass to allow natural

89
90 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE CS1.3 Section through the main showroom.

TABLE CS1.1 Upper and Lower Limit Results of the Natural Ventilation Analysis
Mean Monthly 80% Upper 80% Lower Hours Lower Hours Above
Month
Temperature Limit Limit than Low Limit High Limit

January 56 77.86 65.26 60 6

February 56 77.86 65.26 88 8

March 58 78.48 65.88 67 0

April 60 79.1 66.5 50 4

May 63 80.03 67.43 79 1

June 64 80.34 67.74 19 0

July 68 81.58 68.98 0 0

August 68 81.58 68.98 0 0

September 68 81.58 68.98 0 15

October 65 80.65 68.05 21 0

November 61 79.41 66.81 15 1

December 61 79.41 66.81 60 0

daylight into the space was proposed; however, this is not recommended from
a thermal standpoint due to the heat gain from the skylights.
In addition to the two entry/exit doors at opposite ends of the building,
there is a courtyard for the employees through which outside air can flow.
This report only contains an analysis of the main store space, which
consists of Showroom1 and 2 and the church room. The glazing was a
CASE STUDY—A.P.C. STORE, HOLLYWOOD, CA | 91

monolithic Solarban 70 XL glass in the exterior. The lighting power


consumption is in the order of 1.6 W/ft2 (5.5 Btu/h/ft2).
The store manager would like to see as many occupants, but in reality,
the average time shoppers will remain in the store is about 15 minutes. So
other than the store opening, the average number of shoppers will be
between 10 to 15 minutes, and the heat gain from these occupants must be
taken into account when designing a conditioning system for the store.
The ventilation air can enter the space through the low-level openings and
will exit the space via high-level openings. Further detailed analysis will be
required to determine the size and operation of the openings.
Table 1 shows the mean monthly temperature for each month, the 80% and
90% upper and lower limits as well the number of hours the store is lower than
the minimum limit, then heating would be required or higher than the
maximum limit when the radiant floor would be used for cooling.
OUTDOOR CONDITION
ASSESSMENT
As previously noted, a key element to the success of a natural ventilation or
natural conditioning scheme is the suitability of the outdoor air condition.
There are five primary aspects of outdoor air conditions that should be
explored when considering the feasibility of a natural ventilation scheme:

• Air quality
• Outdoor noise environment
• Wind climate (or wind regime)
• Historical weather data
• Future climate trends

As published weather files are used in a number of these analyses, it is


beneficial to discuss how to source weather files. Modeling naturally ventilated
spaces is heavily reliant on having high-quality weather files representing the
anticipated outdoor air conditions. As a point of reference, users of energy
simulation software programs will select from a wide variety of weather data.
The EnergyPlus weather website (DOE n.d.) compiles statistically processed
weather files from a wide variety of sources. Before trying to acquire weather
data by other means, it is strongly recommended that design teams check if an
existing EnergyPlus Weather Format (EPW) file is within 20 mi to 30 mi
(30 km to 50 km) and within a few hundred feet (100 m) of the site’s elevation.
The National Renewable Energy Laboratory (NREL) generates the *.epw
files on a periodic basis as updates within the National Solar Radiation
Database. The most current version of the typical meteorological year data
(TMY3) is based on a period from 1991 to 2005. These files are postprocessed
versions of data from a multiyear period to represent the typical 8760-hour
typical meteorological conditions for the purposes of supporting energy
modeling simulations for evaluation of whole-building energy use (including
mechanical systems) across the year. They do not represent extreme conditions
or the aggregation of actual conditions across the analysis period. Because they
are generally accepted in the industry as representative of the climate of the

93
94 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

location, they are a good source of commonly accepted public domain data for
evaluating the feasibility of natural conditioning in comparison to traditional
mechanical systems that would have themselves been subject to the same
*.epw files when demonstrating energy compliance.
If an *.epw file is not available, the designer should avoid using single-year
measured or test reference year-type (TRY) weather data, because no single
year can represent typical long-term weather patterns. More information on
selecting weather data appropriate for energy simulation is described in
Crawley (1998).

3.1 AIR QUALITY


The first step in determining if natural ventilation is appropriate is to assess
the regional and local air quality. This is done to meet the requirement set in
Section 4 of ASHRAE Standard 62.1-2019 (ASHRAE 2019), which states that
the design team provides documentation of an outdoor air quality investigation
to the building owner.

3.1.1 Regional Air Quality Assessment


Documenting the regional air quality consists of reporting on the
concentrations of key pollutants as compared to the U.S. Environmental
Protection Agency’s (EPA) National Ambient Air Quality Standards
(NAAQS), which can be found online (EPA n.d.-c). As part of complying with
the Clean Air Act, the EPA has established maximum allowable concentrations
on six principal pollutants (also called “criteria” air pollutants), namely carbon
monoxide (CO), lead (Pb), sulfur dioxide (SO2), nitrogen dioxide (NO2),
ground-level ozone (O3), and particulate matter (PM) at two different sizes
(PM10 for 10 µm or less and PM2.5 for 2.5 µm or less). All U.S. states have
been downgraded to maintenance status, which means that repeated readings
have been so low that the federal government continuously monitors
concentrations but does not mandate any corrective action, for CO (in 2010)
and NO2 (in 1998).
Note that the World Health Organization (WHO) has established air
quality guideline values that differ from NAAQS’ values (WHO 2018).
The EPA maintains the Green Book online (EPA n.d.-d) as a means of
publicly identifying counties in which certain pollutant concentrations have a
status of maintenance (monitoring) or nonattainment (lack of compliance).
Figure 3.1 shows the EPA’s Multi-Pollutant Map of Counties Designated
“Nonattainment,” which is a good place to start a regional air quality
assessment (EPA n.d.-b), as it might quickly show the design team if there are
any pollutants of concern in the area. Should the county be designated as a
nonattainment zone, further investigation into each of the individual pollutants
is necessary, as current design values measured by the EPA may show that the
county has come into compliance since the status designation date.
When pursuing a natural ventilation scheme, it is particularly important to
share the regional air quality information with the building owner, especially if
CHAPTER 3 | 95

TABLE 3.1 Comparison of NAAQS Criteria Thresholds to the WHO Guidelines


Pollutant Averaging Time NAAQS level WHO Level

Carbon Monoxide (CO) 8 hours 9 ppm (10.3 mg/m3) N/A

Lead (Pb) Rolling 3-month average 0.15 µg/m3 N/A

Sulfur dioxide (SO2) 1 hour 75 ppb (197 µg/m3) 20 µg/m3

Nitrogen dioxide (NO2) 1 hour 100 ppb (188 µg/m3) 200 µg/m3

Ozone (O3) 8 hours 0.07 ppm (137 µg/m3) 100 µg/m3

Particle Pollution PM2.5 1 year 12 µg/m3 10 µg/m3

Particle Pollution PM2.5 24 hours 35 µg/m3 25 µg/m3

Particle Pollution PM10 24 hours 150 µg/m3 50 µg/m3

the air quality values are not attaining the primary standards as the primary
standards’ levels are set “to provide general public health including protecting
the health of ‘sensitive’ populations such as asthmatics, children, and the
elderly” (EPA n.d.-c). In particular, it is important to be transparent about the
presence of particulate matter. In a design involving mechanical ventilation,
Section 6.1.4 of ASHRAE Standard 62.1 (2019) requires a minimum of a
MERV 8 filter for any area with a nonattainment on PM10 and a MERV 11
filter for any nonattainment areas for PM2.5. A natural ventilation scheme has
no such protection. Building owners need to understand the likely exposure
levels so that alternate but equivalent work or learning spaces for sensitive
populations are considered (as necessary) during the design discussion. A
research paper from Lawrence Berkeley National Laboratory (LBNL)
estimated that the adverse health costs of retrofitting existing office buildings
to natural ventilation schemes would outweigh the savings resulting from
reducing sick building syndrome (SBS) symptoms (Chao 2013). For areas
outside of the U.S., the Global Health Observatory (GHO) (WHO n.d.) is a
useful resource that publishes maps related to ambient PM10 and PM2.5 to
address similar concerns.
Section 6.1.4.3 of ASHRAE Standard 62.1 (ASHRAE 2019) requires
ozone-cleaning devices for those areas exceeding a designated ozone
concentration of 0.100 ppm (195 µg/m3) for an eight-hour exposure, which is
currently found (per 2017 measurements) only for the Greater Los Angeles
Area Basin (EPA n.d.-e). Historical supporting data available through a
localized ozone exceedance report that can be generated at the EPA’s Outdoor
Air Quality Data website (EPA n.d.-a).
A case study of outdoor air quality assessment can be found in
Appendix A3.1.
96 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.1 Counties designated “Nonattainment” for the Clean Air Act’s NAAQS:
Multi-pollutant County map (EPA n.d.-b).
Courtesy: EPA

3.1.2 Local Air Quality Assessment


In addition to the regional analysis, local air quality assessments are
necessary. Section 4 of ASHRAE Standard 62.1 (ASHRAE 2019) defines
reporting requirements assessing the local sources of odors, irritants,
contaminants, plumes, and exhausts on the property or adjoining sites.
Photo documentation of discharge locations, as well as conversations with
adjacent buildings’ property owners, is useful for determining the types, likely
concentrations, and frequency of discharge of pollutants. In the future, this will
assist with any concentration analysis that is done in conjunction with the
outdoor wind environment. Where effluent is known to contain contaminants
of concern, it is recommended that an outdoor computational fluid dynamics
CHAPTER 3 | 97

(CFD) analysis or a wind tunnel study is performed to determine whether


unacceptable concentrations of pollutants will be present in the intake air at the
natural ventilation openings.
An element not explicitly covered by ASHRAE Standard 62.1 (ASHRAE
2019) is the proximity of the site to major freeways, highways, or other
multilane roads with frequent traffic. Multiple studies show that areas
downwind of these types of roadways experience an increased concentration of
black carbon, ultrafine particles, CO, and nitrous oxides created by gasoline-
burning and diesel-burning vehicles. Additionally, there is supporting evidence
of adverse respiratory health outcomes arising from prolonged exposure to
these increased levels of pollutants (Brugge et al. 2007).

3.2 OUTDOOR NOISE ENVIRONMENT


The outdoor noise environment is a key component to determining whether
to consider a natural ventilation scheme. BSRIA Guidance Note 7/2000
(Martin and Fitzsimmons 2000) advises that noise from traffic (e.g., vehicles,
rail, or air) is often cited as the reason that natural ventilation is not pursued.
The argument is that persistent and irritating noise in the outdoor environment
would create an unmitigated adverse indoor noise environment because natural
ventilation relies on a hole in the building envelope.
There is relatively little research on acceptable indoor noise environments
for naturally ventilated office or classroom spaces. The most recent summary
of information about noise break-in for naturally ventilated environments can
be found in the California Energy Commission (CEC) report Natural
Ventilation for Energy Savings in California Commercial Buildings (Linden et
al. 2015). The information in this section is heavily derived from the material
collected therein.

3.2.1 Method of Determining Adequate


Outdoor Noise Environment
During the discussion of natural ventilation’s feasibility, an important
question is how do you know if the prevailing outdoor noise environment is
actually going to lead to an unacceptable indoor environment? There are three
components to manage to judge the suitability of the outdoor environment:

1. the target indoor noise environment


2 an assessment of the acoustic mitigation provided by the opening type
proposed
3 an estimate of outdoor noise ambient environment to determine the
likely outdoor noise near the façade opening, defined as 3 ft to 6 ft (1 m
to 2 m) from the façade

The calculation of allowable outdoor noise is then quite simple:

allowable outdoor noise near the facade  target indoor noise environment + opening attenuation
98 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Appendix A3.2 contains the results of an experiment comparing the


acceptability of outdoor noise to indoor noise in a naturally ventilated office
space.

3.2.2 Target Indoor Noise Environment


The first inclination would be to make the indoor noise environment under
natural ventilation schemes to equal that required of an indoor space with
background noise arising from mechanical ventilation equipment. However,
researchers have argued in the past that this may be too onerous for naturally
ventilated environments, as people have a higher tolerance for noise break-in
when it comes through the very same window opening that also provides
desired ventilation air.
Linden et al. (2015) identified that there is relatively little research about
the influence of external noise break-in on the acceptability of indoor noise
environments created in naturally ventilated spaces. The following summarizes
the key findings isolated for naturally ventilated building studies:

• In 1974, the Building Research Establishment argued that 55 dBa is a


reasonable standard for auto traffic noise within a two-person office,
based on complaints arising at 60 dBa and the rising influence of
telephone usage (BRE 1974).
• Wilson (1992) suggested a maximum of 55 dBLAeq to 60 dBLAeq was
appropriate for naturally ventilated spaces based on speech
intelligibility criteria between individuals and in phone usage.
• Wilson et al. (1993) found that internal noise levels in the range of
51 dBLAeq to 55 dBLAeq arising from traffic noise caused an increased
number of complaints even when occupational noise levels were
59 dBLAeq to 62 dBLAeq.
• For a study of 12 naturally ventilated buildings and one mechanically
ventilated building in the U.K. and Pakistan, Dubiel et al. (1996) found
that respondents’ assessment of “just right” indoor noise levels
depended on the activity in which they were engaged. During the
summer in the U.K., the noise levels were as shown in Table 3.2.

Linden et al.’s (2015) report generally contends that an indoor target of


55 dBLAeq should be applied for naturally ventilated spaces, as compared to
the approximately 40 dBLAeq to 45 dBLAeq that would typically be found in
mechanically ventilated spaces under U.S., U.K., and Australian standards. It
is strongly recommended that the assumptions surrounding indoor noise
targets be thoroughly discussed with the owner prior to pursuing natural
ventilation to agree whether a more lenient 55 dBLAeq or a more typical
45 dBLAeq is to be used.
Sound privacy is the primary driver of dissatisfaction related to the indoor
acoustic environment, according to Jensen et al.’s (2005) analysis of responses
from 23,450 respondents in 142 buildings, the majority of which were
mechanically ventilated. The results of the meta-analysis showed that only
CHAPTER 3 | 99

TABLE 3.2 Respondents’ Nominal Noise Level Preferences in a


Naturally Ventilated Building in the U.K. Based on Activity Type*
Noise Level Deemed “Just Right” in a
Activity
Naturally Ventilated Space

Computer work 49 dBLAeq

Reading 56 dBLAeq

Meting, phoning, word processing/typing 57 dBLAeq to 58 dBLAeq

Talking/taking a break 59 dBLAeq to 60 dBLAeq

Writing and other individual activities 60 dBLAeq to 61 dBLAeq


*Data Source: Dubiel et al. (1996)

18% to 25% of respondents found the HVAC noise to be the specific cause of
their dissatisfaction, whereas 59% to 82% of respondents identified people
talking on the phone, people overhearing private conversation, and people
talking in surrounding offices as the primary cause of dissatisfaction.
In naturally ventilated spaces, mechanical ventilation is typically not
available to provide background noise, thus removing a cause of acoustical
dissatisfaction while simultaneously reducing speech privacy.
Appendix A3.2 describes a field study performed to compare occupant
dissatisfaction with indoor as compared to outdoor noise in a naturally
ventilated office space.
3.2.3 Attenuation of Natural Ventilation Apertures
Linden et al. (2015) identified that noise apertures provide a means of
attenuating outdoor noise before it enters buildings, with open windows
generally assumed to be a 10 dBLAeq to 15 dBLAeq reduction and integrated
double facade designs estimated to rise in attenuation to 20 dBLAeq to
25 dBLAeq. Ryan et al. (2011) noted a 10.7 dBLAeq reduction for road noise.
These values are generally consistent with a 2018 study regarding the
relative attenuation of windows based on their openness position, showing that
“the median outdoor–indoor sound level differences were of 10 dB(A) for
open, 16 dB(A) for tilted, and 28 dB(A) for closed windows” (Locher et al.
2018). This study also provides an overview of other similar studies that have
been performed based on notable transport-related noise sources at the German
Aerospace Center (DLR), as noted in Table 3.3.
For a facade with basic open windows, Linden et al.’s (2015) report
contends that outdoor noise can be as high as 65 dbLAeq to 70 dbLAeq, as
measured at 3 ft to 6 ft (1 m to 2 m) from the external façade, based on the
following formula:

allowable outdoor noise near the facade  target indoor noise environment + opening attenuation

65 dBLAeq to 70 dBLAeq  55 + 10 to 15 dBLAeq


100 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 3.3 Mean Differences In Sound Levels Outdoors Based on


Measured Results From Multiple Studies*
2010 Study 2010 Study 2006
Window 2010 2006 Road
2018 Study Freight Passenger Aircraft
Position Road Noise Noise
Train Train Noise

Open 10.0 11.3 11.9 11.6 13.4 10.0

Tilted 15.8 18.6 18.0 17.7 13.7 15.3

Closed 27.8 30.1 29.7 30.1 27.0 25.6


*Data source: Locher et al. (2018)

Additionally, natural ventilation openings can use acoustically attenuated


ventilator products that achieve an A-weighted sound reduction index (Rw) in
the range of 11 to 26 dB(A), comparable to the sound transmission class (STC)
rating as per ASTM E413 (2016).

3.2.4 Estimating the Outdoor Noise Environment


In the best-case scenario, an acoustic consultant would use a sound meter at
the site to do an ambient noise assessment across a 24-hour period. This
consultant would then calculate the estimated ambient noise level as the facade
opening closest to the most egregious noise sources, taking into account noise
reflecting off of the facade surface itself.
At the feasibility stage, this is not always possible. Therefore, the design
team may have to estimate likely outdoor noise sources and the environment
through external benchmarks. The following sources allow the design team to
estimate outdoor ambient conditions by matching the environment type closest
to that found at the site. IAC Acoustics (n.d.) offers comparative examples of
noise levels. Additionally, Chapter 1 of the HUD Noise Guidebook (HUD
2009) offers an alternate set of comparisons for benchmarking.

3.3 WIND CLIMATE


As noted in Chapter 2, wind is one of two primary mechanisms for
generating the pressure differential required to achieve natural ventilation.
Thus, it is important to understand the wind climate (sometimes called a wind
regime) to comprehend how wind-driven forces are likely to impinge on the
building’s natural ventilation opening under all anticipated wind directions.
For natural ventilation in high-rise buildings, an excellent collection of case
studies and design recommendations can be found in Wood and Salib (2013).
Detailed mathematical analysis of the concerns for natural ventilation in highly
dense urban cores can be found in Ghiaus and Allard (2005).
A wind rose is a quick snapshot of the general wind climate. It indicates the
frequency, velocity, and direction of wind in general at a weather station close
to the site (typically in a free field configuration, such as at an airfield). It is
CHAPTER 3 | 101

useful for understanding the generalized wind climate with regard to


directionality and strength.
The meteorological data used to generate wind roses does not take into
account any localized effects arising from geomorphic or man-made features
that may cause obstructions or turbulence in upstream or downstream wind
conditions. Chapter 24 of ASHRAE Handbook—Fundamentals (ASHRAE
2017b) provides an overview of the phenomenon of wind pressure on
buildings and how to calculate the time-averaged surface pressure based on
wind velocity pressure. Regarding the development of pressure coefficient Cp,
it strongly advises that an “[a]ccurate determination of Cp can be obtained only
from wind tunnel model tests of the specific site and building or full-scale
tests.” CFD analysis has progressed to the point that this simulation tool can
also accurately take into account the wind climate and predict pressure
coefficients on uniquely shaped buildings to account for the local wind
variability based on the building in the context of its particular surroundings
and obstructions.
This section will discuss how to obtain wind data to produce wind roses as
a step toward assessing the feasibility of wind-driven natural ventilation. It will
also discuss how wind tunnels and CFD are used to extrapolate surface
pressure coefficients for use later in the natural ventilation analysis process.

3.3.1 Historical Wind Rose Analysis


The first place to start when assessing the wind climate of a site is a wind
rose analysis. There are several free software programs available to generate
wind roses from *.epw or *.tmy data available from the EnergyPlus weather
data site (DOE n.d.) or other online data sources. The following sections show
output from three of these free online tools.
The appropriate wind source must be used. Refer to Appendix A3.3 to see
how widely wind rose data deviates, even within the same city. When a design
team is considering a wind-driven natural ventilation scheme, it is strongly
recommended that an on-site weather station be installed to log wind direction
and speed so the localized effects are well understood as a design parameter.
This will allow data to be parsed for the time of day as well as the concurrent
temperature of the air.
A temperature-segregated version of the wind rose data is also quite useful
for design teams considering wind-driven ventilation. As noted in
Appendix A3.3, December conditions can vary widely from an annualized
view. Appendix A3.4 contains a study of wind direction by temperature bin to
explore the incoming wind direction when the outdoor air temperature is above
77°F (25°C)—the point at which natural ventilation schemes tend to rely on
very high velocities and air change rates to support indoor comfort.
3.3.1.1 Climate Consultant. Climate Consultant (UCLA n.d.) is a tool
produced by the University of California, Los Angeles. The user can download
the *.epw (EnergyPlus weather file format) of choice to create what is referred
to as a wind wheel, in which the brown bars show the frequency of wind
102 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.2 Wind wheel for the Los Angeles International Airport using Climate
Consultant 6.0, Build 12.
Courtesy: University of California, Los Angeles’ Energy Tools Design Group,
Professor Murray Milne.
© Regents of the University of California.

direction and the orange through yellow bars indicate the velocity achieved.
(See Figure 3.2.) An automation function allows the monthly wind wheels to
be generated.

3.3.1.2 Iowa Environmental Mesonet. Iowa State University’s Iowa


Environmental Mesonet website (ISU n.d.) also automatically generates wind
roses from actual historical weather readings within weather station networks.
(See Figure 3.3.) Annual aggregated and monthly aggregated wind roses are
automatically generated to represent the period of record. To use, click select
from map, pull down to the state network, and then click switch network, then
select the relevant station. The monthly data allows for a clearer representation
of seasonal influences at this particular site during the winter months.

3.3.1.3 Lakes Environmental’s WRPLOT. WRPLOT View™–Freeware


(Lakes Environmental n.d.) is available for plotting wind data using historical
weather files. (See Figure 3.4 for a wind rose generated using WRPLOT.)

3.3.1.4 WebMET. A limited number of nationally provided wind information


in the appropriate formats can be found on the WebMET website (WebMET
n.d.) for the U.S. and Canada.
CHAPTER 3 | 103

FIGURE 3.3 Wind rose (annual data) as generated from the Iowa State University
Mesonet.
Courtesy: Iowa State University

3.3.2 Generating Pressure Coefficients


from Wind Studies
There are four methods for estimating pressure coefficients at natural
ventilation openings. In order of increasing cost and complexity, they are:

• Manual estimation based on Chapter 24 in ASHRAE Handbook—


Fundamentals (ASHRAE 2017b)
• Desk study by a wind consultant based on previous experience and
observation of the surrounding environment
• CFD analysis of the local wind climate
• Wind tunnel testing
3.3.2.1 Manual Estimation of Pressure Coefficients (Based on ASHRAE
Handbook—Fundamentals). Chapter 24 of ASHRAE Handbook—
Fundamentals (ASHRAE 2017b) discusses airflow around buildings.
Section 2, “Wind Pressure on Buildings” (ASHRAE 2017b) describes how to
modify meteorological wind speed to reflect the wind speed anticipated at the
height of the building and within the terrain category anticipated. This
adjustment methodology may be found in Table A2.3.1, “Atmospheric
Boundary Layer Parameters (ASHRAE 2017b),” in Appendix A2.3.
104 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.4 Wind rose from Lakes Environmental’s WRPLOT.


Courtesy: Lakes Environmental

The chapter also provides images representing typical surface coefficients


for low-rise and high-rise buildings based on research completed during the
1970s and 1980s. If openings are located away from the edges of the wall,
these values can be used as approximate rules of thumb inputs to other
calculations of bulk airflow analysis through the indoor space, with the
pressure coefficients acting as boundary conditions at the openings. A few
images are replicated in Appendices A2.4 and A2.5.

3.3.2.2 Desk Study by Wind Consultant. It is often beneficial to hire a


wind consultant to complete a low-cost desk study to explore whether wind-
driven natural ventilation is a viable year-round source of a pressure
differential to drive natural ventilation and/or natural conditioning.
Wind consultants typically have a much greater depth of knowledge of
detailed wind history information than what is found online. They also have a
great deal of professional judgment surrounding the air distribution and
emission dispersion patterns of various types of exhausts that may adversely
affect the outdoor air quality surrounding the building. As natural ventilation
schemes rely on untreated outdoor air to come into a human-occupied space, it
is necessary to understand both how air moves around a building as well as the
hyper-localized outdoor air quality at openings. This outdoor air measurement
at the building is in addition to the preliminary regional and local air quality
assessment mentioned in Section 3.1, “Air Quality,” of this guide.
CHAPTER 3 | 105

FIGURE 3.5 CFD sample showing outdoor wind passing around building obstructions.

Because of the depth of hourly data available to a wind consultant, it is


often also possible to request temperature-binned wind roses, similar to those
that were previously noted. This allows the design team to determine wind
strengths at the time when air velocity is most needed. The design team can use
this information to calculate likely pressure coefficients between the windward
and leeward sides of the building as the generator of air movement within.
Website Appendix WA3.1, “Sample RFP Text for an Environmental Wind
Desk Study,” accessible at ashrae.org/naturalventilation, includes sample text
for a request for proposal (RFP) for an environmental wind desk study scope of
work that is appropriate to this early level assessment of wind-based
opportunity and risk.

3.3.2.3 Computational Fluid Dynamics Analysis of the Local Wind


Climate. Computational fluid dynamics (CFD) is a tool that is used
to simulate indoor or outdoor airflow, temperature, and concentrations. It can
be used to represent any type of fluid, gas, or liquid. It is essentially a tool that
iteratively solves the Navier-Stokes equations to represent a steady-state
condition as a snapshot of an instance in time.
CFD can be used to inform performance-based design and is a powerful
visualization tool. (See Figures 3.5 and 3.6.) The reasons to use CFD in lieu of
wind tunnel testing are primarily because of its reduced cost and greater
malleability of the model (since it is not being physically built). Current CFD
analysis tools have historically been used to predict airflow patterns within
106 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.6 CFD sample showing air entering the building in a single-sided natural
ventilation scheme.

well-bounded indoor volumes. More recently, they have been successfully


benchmarked against wind tunnel test results to create a methodology for using
CFD to represent outdoor wind climates.
When exploring natural ventilation, the CFD analyst must understand the
wider wind climate around the building. Only a professional well experienced
in CFD modeling of exterior environments should be used because of the
following:

• the need to apply expertise in determining cell size based on proximity


to the building,
• the proper evaluation and representation of boundary conditions, and
• the correct interpretation of results.

Typically, the scope of a CFD analysis for natural ventilation is divided


into two pieces: that which assesses the airflow around the building to
determine pressure coefficients (covered in this chapter) and that which
assesses airflow inside of the building (see Section 6.4.2, “Use of Bulk Airflow
and Dynamic Thermal Simulation Modeling”). This is due to the very different
cell sizes to accommodate the level of detail needed in outputs for each case.
Website Appendix WA3.2, “Sample RFP Text for a Computational Fluid
Dynamics Study,” accessible at ashrae.org/naturalventilation, includes sample
text of an RFP for a CFD scope of work for an early assessment of the potential
for wind-driven natural ventilation.
CHAPTER 3 | 107

3.3.2.4 Environmental Wind Tunnel Study for Natural Ventilation


Performance. Wind tunnel testing has historically been performed
on structures of complex shape to assist structural engineers in determining
maximum wind forces (both positive and suction pressures). A physical model
is built to include an array of pressure taps that are installed around the model
to physically measure pressures under wind conditions. The model is placed on
a turntable within a wind tunnel to allow a variety of wind directions to be
tested. (See Figure 3.7.)
Wind tunnel testing requires that the testing facility model the area within
the vicinity of the building at the appropriate level of three-dimensional (3-D)
detail to demonstrate how adjacent features might affect the airflow pattern
hitting the building. The upwind profile (off the turntable and nearer to the fan)
is created to simulate a level of roughness representative of the urban,
suburban, or open conditions of the site. There is a significant amount of
mathematical complexity associated with scaling measured results from the
physical model to extrapolate them to the actual predicted values of pressure
and anticipated pressure coefficients at natural ventilation openings. The
design team should request post-processed results in the final report to avoid
any confusion in applying the reported pressure coefficients in CFD modeling.
Unlike structural wind tunnel testing, environmental wind tunnel testing
does not primarily rely on wind climate using maximum design speeds with
high-year return periods to meet ASCE 7-10 safety requirements (ASCE
2010). Instead, environmental testing uses the mean and gust wind speeds
anticipated to be parallel to those that might be found in the typical
meteorological year file.
There is a need to design natural conditioning systems in particular for
reasonably conservative mean air speeds if the wind-driven pressure
differential is the primary means of air movement inside of the building. As
such, it is often useful to take the data determined in the aforementioned desk
study and spend more time on wind tunnel cases in the directions of the high-
temperature hours.
Appendix WA3.3, “Sample RFP Text for a Wind Tunnel Study, accessible
at ashrae.org/naturalventilation, includes sample text of an RFP for a wind
tunnel study of the specific building geometry at the site in conjunction with a
representation of surrounding obstructions.

3.4 HISTORICAL WEATHER DATA


If the results of the outdoor air quality and noise environment tests are
adequate to support natural ventilation, then the design team may consider
whether to explore the capacity of the outdoor environment to support natural
conditioning. This requires additional analysis to explore feasibility. There are
four primary means of assessing national conditioning potential during the
early stages of a project:

• Estimating natural ventilation potential by U.S. climate zone


• Reviewing mean maximum and mean minimum monthly temperatures
108 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.7 Wind tunnel with test sample.


Courtesy: Cermak Peterka Petersen (CPP).

• Performing a frequency of occurrence analysis using software tools


• Performing frequency of occurrence analysis by building a spreadsheet

3.4.1 Natural Conditioning Potential by U.S. Climate Zone


The analysis that led to the creation of the ASHRAE climate zone map
placed boundaries around geographic areas having similar climatological
performance. The climate of the U.S. is broadly partitioned latitudinally, then
by humidity performance into moist, dry, and marine climates.
If the project is located in the continental U.S., the maps in Appendix A3.5
and the summary charts in Figures 3.9 and 3.10 provide some quick rules of
thumb to assist in determining the number of hours when natural conditioning
is feasible. Note that the data represented is based on McConahey (2011), and
thus representative of historical weather data as of that time. The analysis
behind the maps and graphs is based on the assumption that daytime outdoor
air conditions would nominally be under 70% relative humidity (rh) and in the
temperature range of 60°F to 80°F (15.6°C to 26.7°C) to absorb internal heat
gains while maintaining target indoor adaptive comfort criteria conditions with
increased air movement. Similar to the work done in the ASHRAE Advanced
Energy Design Guides (ASHRAE n.d.), this analysis designates a
CHAPTER 3 | 109

FIGURE 3.8 ASHRAE climate zone map.

representational city within each climate zone. The graph in Figure 3.9 shows
the comparison of annual natural conditioning potential for those 15 ASHRAE
climate zone representational cities.
The analysis in Figure 3.9 demonstrates that the use of natural conditioning
alone is unlikely to achieve comfort conditions throughout the year. Heating is
obviously needed when temperatures fall. In some cases, it would be wise to
pursue a mixed-mode approach to provide auxiliary cooling on the warmest of
days. It is important to discuss with the owner how many hours the occupants
would be able to use natural conditioning. As noted in Chapter 1, the team must
weigh occupant productivity across the year against the cost of installing operable
elements in façade systems. One might question whether it is worth the difficulties
to provide operable windows if they can only be used for 18% of daytime hours, as
in Fairbanks, AK. Whereas it would seem to be an obvious choice to consider
natural conditioning schemes for Los Angeles, CA. Emmerich et al. (2003)
prepared an analysis of ten California cities to determine climate suitability for
direct daytime ventilation as well as nighttime ventilation when needed. Emmerich
et al. (2011) performed a similar analysis of four representational cities to test
climate suitability of natural ventilation and similarly found that Los Angeles
ranked at 99.4% effective direct cooling across a range of combined internal heat
gains of 0.9 W/ft2 to 7.4 W/ft2 (10 W/m2 to 80 W/m2).
110 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.9 Annual natural conditioning potential for 15 U.S. cities based on 60°F to 80°F
(15.6°C to 26.7°C) less than 70% rh limit for 7:00 a.m. to 7:00 p.m.
(McConahey 2011)

To know how best to handle the mixed-mode combination of a mechanical


conditioning system and a natural conditioning system, it is necessary to
further examine the data to understand seasonal climate behaviors, as noted in
Figure 3.10. Viewing the data in this way immediately highlights issues like
the summer humidity and/or temperature extremes in the southern latitudes
and the extreme heating demands in the northern latitudes during the winter.
The other way to visualize the seasonal data is by looking at the natural
conditioning potential percentages against the backdrop of a national isotherm
map. Seasonal maps showing this type of analysis can be found in
Appendix A3.5.
In summary, a quick overview of the graphs and the maps in
Appendix A3.5 quickly allows a design team to have a first pass estimate of
CHAPTER 3 | 111

FIGURE 3.10 Seasonal natural conditioning potential for 15 U.S. cities based on 60°F to
80°F (15.6°C to 26.7°C), less than 70% rh limit for 7:00 a.m. to 7:00 p.m.
(McConahey 2011)

natural conditioning potential without calculations. This is only meant for


initial guidance, with the expectation that a more in-depth analysis would be
performed to justify the natural conditioning potential of the particular site of
the building under design.

3.4.2 Review of Mean Maximum and


Mean Minimum Monthly Temperatures
If the climate zone seems to have a sufficient number of natural
conditioning hours available, then the design team should look at the mean
maximum and mean minimum monthly temperatures. These can typically be
found on the National Oceanic and Atmospheric Administration’s (NOAA)
112 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.11 Mean monthly, mean monthly maximum, and mean monthly minimum
temperatures and precipitation for Los Angeles, CA.
(NOAA n.d.-b)

1981–2010 U.S. Climate Normals website (NOAA n.d.-a). First, select the
“Search Tool,” then “Monthly Normals.” Then, the data can be downloaded or
represented as a plot of mean monthly maximum and mean monthly minimum
temperatures, as shown in Figure 3.11.
This data is useful to compare to the limits set by Section 5.4.1 of
ASHRAE Standard 55, which are limited to “the prevailing mean outdoor
temperature [being] greater than 10°C (50°F) and less than 33.5°C (92.3°F)”
(ASHRAE 2017a). If the mean monthly temperatures meet the criteria, further
weather analysis is necessary to determine how many hours of applicability are
available. (See Section 3.4.3 for a frequency of occurrence analysis.)
For months that do not fall into the monthly prevailing mean temperature
criteria range of ASHRAE Standard 55, a mechanical conditioning system will
likely be required during noncompliant hours of occupancy. If a change-over
mixed-mode/hybrid ventilation approach is pursued, the mean monthly
maximum and minimum temperatures are typically used to determine the
number of months in which occupants could be reasonably expected to use the
natural ventilation option. It is reasonable to assume that occupants need at
least four to six months in which they could choose to use windows, otherwise,
the operable window amenity is so continuously underutilized that people may
feel unaccustomed to the natural ventilation option, and thus under
incentivized to open the window. In terms of weather analysis, the eligible
months for some level of natural ventilation in a mixed-mode system
corresponds to those months in which the outdoor historic temperatures fall
simultaneously into the range of a mean monthly maximum of less than 80°F
(26.7°C) and a mean monthly minimum of greater than 32°F (0°C).
CHAPTER 3 | 113

3.4.3 Frequency of Occurrence Analysis Overview


The primary tool used to test climate capacity is a frequency of occurrence
analysis of the typical meteorological year. This data is typically found in a
*.epw or *.tmy format from the EnergyPlus weather data site (DOE n.d.) or
other online data sources.
Frequency of occurrence analysis can be done in a spreadsheet but is most
frequently done as a graphic representation using a psychrometric chart. The
goal is to count the number of hours that fall into temperature range bins that
would lead to a successful natural conditioning system. This information is
crucial to determining whether the outdoor climate can support a natural
conditioning scheme throughout the entire year, or whether a mixed-mode
approach is advisable.

3.4.4 Frequency of Occurrence Analysis


using Software Tools
These tools covered in this subsection are useful for generating a quick
view of psychrometric data from typical meteorological year weather files.
They generally report out in percentage of hours in compliance with the target
comfort criteria set by the user.
Generally speaking, if the total hours not addressed by natural conditioning
exceed 5% or the client’s stated allowable exceedance annually, the design
team and the client should discuss whether an air-conditioning system will be
needed. Similarly, if the total hours below 60°F (15.6°C) exceed 10%, then a
heating system and an insulative infiltration-resistive envelope should be put in
place. It is reasonable to expect that the users will consider natural ventilation
to be a viable option only if they believe that they could use it for at least ~25%
to 30% of their time in the building. Otherwise, the occupants’ habitual lack of
engagement with manual window operations translates to a reduced incentive
to actively track outdoor air temperatures as a proxy for available outdoor air-
based cooling.
3.4.4.1 Climate Consultant. Produced by the University of California, Los
Angeles, Climate Consultant (UCLA n.d.) is a tool that allows a user to
download their chosen *.epw file to examine a year’s worth of hourly data
superimposed onto a psychrometric chart. Each dot on the chart represents the
temperature and humidity of each of the 8760 hours per year. Different design
strategies are represented by specific zones on this chart. The percentage of
hours that fall into each of the 16 different design strategy zones gives a
relative idea of the most effective passive heating or passive cooling strategies.
This lets the user quickly visualize how many hours would fall into different
design strategies that could be used to justify the use of natural conditioning.
See Appendix A3.6 for more information on the use of Climate Consultant
specific to the limitations of use related to its representation of the adaptive
comfort model. Bearing its limitations in mind, it is recommended that Climate
Consultant is used only as a quick feasibility checking tool. A spreadsheet
(discussed in more detail in Section 3.4.5, “Frequency of Occurrence Analysis
114 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

The frequency of occurrence


psychrometric chart will automatically
estimate the number of hours within each
passive or active design strategy based on
criteria set by the user. The display show
enumerates hours for both the ASHRAE
Standard 55 Analytical Comfort Model
and the ASHRAE Standard 55 Adaptive
Comfort Model, the former also with the
elevated air speed allowance. The dots
represent outdoor air conditions.

FIGURE 3.12 Frequency of occurrence psychrometric chart in °F using Climate


Consultant 6.0, Build 12.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California.
CHAPTER 3 | 115

The frequency of occurrence


psychrometric chart will automatically
estimate the number of hours within each
passive or active design strategy based on
criteria set by the user. The display show
enumerates hours for both the ASHRAE
Standard 55 Analytical Comfort Model
and the ASHRAE Standard 55 Adaptive
Comfort Model, the former also with the
elevated air speed allowance. The dots
represent outdoor air conditions.

FIGURE 3.13 Frequency of occurrence psychrometric chart in °C for Los Angeles, CA,
8:00 a.m. to 5:00 p.m., using Climate Consultant 6.0, Build 12.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California.
116 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 3.4 Column Headings of Weather Data in *.epw Format


Dry-Bulb Dew Point Relative
Data
Year Month Day Hour Minute Temperature, Temperature, Humidity,
Source
°F (°C) °F (°C) %

by Building a Spreadsheet”) allows for a more malleable assessment of the


heat pickup based on design input as the project progresses.
It should be noted that the tool can also advise on the percentage of time
when thermal mass might be used to provide passive midday radiant cooling if
the design can leave thermal mass (i.e., concrete or other phase-change
materials) within the occupied spaces. Typically, if there are a large number of
hours when the afternoon temperatures exceed 80°F (26.7°C), exposed thermal
mass in the ceiling, walls, or floor could be beneficial so long as the outside air
temperature drops significantly overnight and can be allowed into the building
for a night purge cycle for cooling the slabs. Typically, a dynamic heat transfer
and bulk airflow model similar to those used in Chapter 6 are necessary to
confirm the effectiveness of thermal mass to provide auxiliary passive cooling.
3.4.4.2 Oasys CLIDAT. The Oasys software suite also has a very basic
frequency of occurrence psychrometric analysis tool that plots conditions out
of an *.epw file. It does not automatically generate any design strategy
recommendations, but it does give the user the ability to set a boundary box for
counting hours within the file. The box’s limited options only include selecting
upper and lower bounds for relative humidity and dry bulb.

3.4.5 Frequency of Occurrence Analysis by


Building a Spreadsheet
The most straightforward way to handle manipulating the data into a spreadsheet
is to strip the top eight lines of the *.epw file in a text editor. Then, the comma-
delimited file can easily be opened in a spreadsheet tool. The columns of interest are
within the first nine columns, which are noted in Table 3.4.
Because Section 5.4, “Determining Acceptable Thermal Conditions in
Occupant-Controlled Naturally Conditioned Spaces” of ASHRAE Standard 55
(ASHRAE 2017a) uses adaptive comfort standards that do not take humidity
into account, the “Dew-Point Temperature” and “Relative Humidity” columns
are not essential. That said, it is strongly recommended that a humidity
assessment be reviewed with the owner, as high humidity hours (above 65% rh
to 70% rh) that technically fall into the bounds of the standard are very likely
to cause complaints due to occupant discomfort. Also, the “Minute” and “Data
Source” columns generally do not contain useful information for the
psychrometric counting exercise. It is, however, useful to count the number of
lines of data to ensure it equals 8760. If not, check the “Minute” column to
ensure that the file contains only using one line per hour, especially if the data
did not come from the EnergyPlus Weather site (DOE n.d.).
If a spreadsheet is used, then the user must mathematically set the
temperature ranges that are defined in ASHRAE Standard 55 (2017a) as
CHAPTER 3 | 117

FIGURE 3.14 Frequency of occurrence psychrometric chart in °F for Los Angeles, CA, 8:00
a.m. to 5:00 p.m., using Oasys CLIDAT.
Courtesy: Arup.

FIGURE 3.15 Frequency of occurrence psychrometric chart in °C for Los Angeles, CA, 8:00
a.m. to 5:00 p.m., using OASYS CLIDAT.
Courtesy: Arup.
118 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.16 Modified outdoor air in °F (including heat pickup) versus 80% acceptability
limits. Compliance count for 8:00 a.m. to 5:00 p.m. Data from Los Angeles, CA’s *.epw file.

comfort zone boundaries for the adaptive comfort standard. See Chapter 6 for a
spreadsheet that provides this information when analyzing indoor resultant
temperatures against ASHRAE Standard 55 criteria. Early in the process, when
the design team has not yet done an analysis to predict internal conditions, then
it is generally necessary to take outdoor air temperatures and add an internal
heat gain allowance of 4°F to 6°F (2°C to 3°C), which is known as the heat
pickup temperature. This heat pickup temperature is added to account for the
heat that may be absorbed into the incoming outside airstream before the air
touches human skin. This is obviously not a replacement for a dynamic heat
transfer and bulk airflow modeling analysis but is a reasonable approximation
of what a well-performing naturally conditioned space should experience. (See
Chapter 6 for more information on how this heat gain allowance can be
estimated.) The modified outdoor air temperatures can be compared to the
allowance ASHRAE Standard 55’s adaptive comfort temperature ranges by
month or by day, as noted in Chapter 6.
Lastly, it should be noted that ASHRAE Standard 55’s (ASHRAE 2017a)
definition of adaptive comfort allows heating to be used, so any counts related
to being too cold would be easily addressed by the design of that system.
With just an *.epw file and a user assumption regarding heat pickup of 6°F
(3°C), it is possible to generate a visual representation of likely compliance
against a flat mean 80% acceptability limit within the adaptive comfort model.
CHAPTER 3 | 119

FIGURE 3.17 Modified outdoor air in °C (including heat pickup) versus 80% acceptability
limits. Compliance count for 8:00 a.m. to 5:00 p.m. Data from Los Angeles, CA’s *.epw file.

3.5 FUTURE CLIMATE TRENDS


Natural conditioning is heavily dependent on outdoor air conditions being
cool enough to absorb indoor heat while remaining below skin temperature
after heat pickup. If this is not the case, then the natural ventilation airflows are
heating the individual. In those high-temperature cases, natural conditioning
can continue to support some level of heat rejection from the skin, but only
when relying on the occupants sweating and a localized increase in air speed
from fans to help the sweat evaporate. The ability of the air to absorb moisture
is contingent on relative humidity, so high temperature/high humidity
environments are not particularly successful at supporting natural
conditioning.
Scientific research has been documenting a systemic rise in global
atmospheric temperature, which automatically brings with it the air’s higher
capacity to hold moisture, proving the accuracy of the oft-heard adage “the wet
will get wetter, and the dry drier.” See Appendix A3.7 for more information on
the background related to the presumption of global warming and a simplified
description of the Representative Concentration Pathways (RCP).
As the climate changes in a warming direction, some northern geographies
will likely will benefit, as more winter and shoulder season temperatures will
shift into the adaptive comfort range. In other, mostly southern climates, the
number of hours that are considered too hot is likely to increase. Future climate
120 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.18 Climate-shifted mean monthly temperatures in °F from the WeatherShift™


*.epw file for Los Angeles, CA, for RCP8.5 emission scenarios and 50% percentile warming
percentile. Data excerpted from WeatherShift (n.d.) on June 2, 2018.

trends must be presented to the owner with an evaluation of whether the


climate might prevent a natural conditioning scheme from effectively working.
The *.epw files currently available online are typically based on historical
data from 1991 to 2005, which effectively is the zero axis line on the graph. If
the design team is working with clients expecting a building to have a 50-year
life, it is important to note to the clients that the anticipated global warming
change under a scenario of the continued increase in carbon emissions
throughout the 21st century (RCP8.5) would be in the range of 4.5°F to 5.4°F
(2.5°C to 3°C). While no one can predict which path society will take
concerning mitigation of climate change, owners and clients should be made
aware of the potential temperature rise associated with the range of currently
modeled global warming scenarios within the global scientific community.
To explore the impacts of global warming during the feasibility stage, there
is free data online from climate-shifted *.epw files that account for scenarios
RCP8.5 and RCP4.5 for a limited number of geographic locations as a public
service. As noted in Aijazi and Brager (2018), future weather data can be
derived from several sources in the U.K., including a spreadsheet-based tool
called CCWorldWeatherGenerator, using coarse general circulation model
data, or the tool called CCWeatherGen for greater localization in the U.K.
(ECCD n.d.).
Aijazi and Brager (2018) also refer to another online tool using free data
called WeatherShift™ (WeatherShift n.d.). (Disclosure: One of the authors is
an employee of one of the firms that developed the climate-shifting
algorithms.) Full climate-shifted *.epw files can be purchased through IES
(n.d.) for a nominal fee.
Using the latter tool’s online visualization capabilities, the presented data
demonstrate that, as expected, local average monthly temperatures will
continue to rise as the global temperature warms. Figures 3.18 to 3.21 graph
CHAPTER 3 | 121

FIGURE 3.19 Climate-shifted mean monthly temperatures in °C from the WeatherShift™


*.epw file for Los Angeles, CA, for RCP8.5 emission scenarios and 50% percentile warming
percentile. Data excerpted from WeatherShift (n.d.) on June 2, 2018.

FIGURE 3.20 Climate-shifted average monthly dew point from the WeatherShift™ *.epw file
for Los Angeles, CA, for RCP8.5 emission scenarios and 50% percentile warming percentile.
Data excerpted from WeatherShift (n.d.) on June 2, 2018.

the worst-case scenario of RCP8.5, which presumes unmitigated rising carbon


emissions.
A check on average monthly dew point and average daily relative humidity
shows that despite increases in dew point, there are no significant changes
anticipated in relative humidity as the increased air temperature is capable of
holding increased amounts of moisture.
Taking the 50-year design life problem for natural conditioning, a review
of the data shows that there is a systematic uplift of the mean monthly
temperature as noted in Table 3.5.
If the uplift in the mean is applied to the previous modified outdoor air and
the newly shifted adaptive comfort boundaries are imposed, it is clear that
122 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 3.21 Climate-shifted average relative humidity from the WeatherShift™ *.epw file
for Los Angeles, CA, for RCP8.5 emission scenarios and 50% percentile warming percentile.
Data excerpted from WeatherShift (n.d.) on June 2, 2018.

TABLE 3.5 Uplift of Mean Temperature between Present and 2065


Under RCP8.5, 50% Percentile
2065 Minus
Jan Feb Mar Apr May Jun Jul Aug Sept Oct Nov Dec
Present

Uplift of mean, °F 5.9 7.2 6.4 7.6 6.7 8.6 6.4 5.8 5.0 4.7 4.8 7.9

Uplift of mean, °C 3.6 3.8 4.0 4.2 4.4 5.0 4.0 3.2 3.2 2.4 3.2 2.6

more hours fall into the too hot range. Figures 3.22 and Figure 3.23 can be
directly compared to the graphs in Chapter 7, where actual internal heat gains
are modeled for indoor spaces to estimated likely indoor temperatures.
For ease of reference, Table 3.6 shows the shift in compliance status that is
represented in Figures 3.22 and 3.23. Cold hours drop, and overheating hours
increase.
It is important to recall that Los Angeles, CA, currently has one of the most
favorable climates for supporting natural conditioning, with a coastal
environment and onshore wind conditions. Even this somewhat ideal location
experienced an increase in the number of hours calculated to be too hot from
1% presently to 6% in 2065. That said, there was a net increase in good and
good with fan hours from 80% presently to 89% in the future, primarily by too
COLD hours moving upwards in temperature.
It is impossible to pinpoint the actual increase in global temperature or the
localized ramifications of this warming phenomenon. The actual global
warming increase by 2065 and beyond will be due to an unpredictable
combination of ongoing emissions processing activity in the global ecosystem
and the emission-creating (or reducing) actions of the presently living
population. Engineers should not be held liable for overheating that occurs in
CHAPTER 3 | 123

FIGURE 3.22 2065 RCP8.5, 50% percentile modified outdoor air in °F (including heat
pickup) versus 80% acceptability limits, compliance count for 8:00 a.m. to 5:00 p.m. Data from
the Los Angeles, CA *.epw file.

FIGURE 3.23 2065 RCP8.5, 50% percentile modified outdoor air in °C (including heat
pickup) versus 80% acceptability limits, compliance count for 8:00 a.m. to 5:00 p.m. Data from
the Los Angeles, CA *.epw file.
124 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 3.6 Comparison of Adaptive Comfort Compliance Status for the Present
Compared to 2065 under RCP8.5, 50% percentile. Compliance Count for
8:00 a.m. to 5:00 p.m. Data from Los Angeles, CA’s *.epw File
Compliance Present, Present, 2065 RCP8.5, 2065 RCP8.5, Hours
Status Hours % of hours Hours % of Hours Difference

TOO COLD 610 19% 187 6% –423

GOOD 2533 77% 2284 70% 249

GOOD with fan 95 3% 611 19% –516

TOO HOT 47 1% 203 6% –156

Total 3285 100% 3285 100%

the future if a thorough discussion of available climate-shifted data is


discussed with the owner to raise awareness of long-term risks and benefits
when deciding on pursuing natural conditioning.

3.6 REFERENCES
Aijazi, A., and G. Brager. 2018. Understanding climate change impacts on
building energy use. ASHRAE Journal 60(10):24–32.
ASCE. 2010. ASCE/SEI 7-10, Minimum design loads for buildings and other
structures. Reston, VA: American Society of Civil Engineers.
ASHRAE. n.d. Advanced energy design guides. https://www.ashrae.org/
technical-resources/aedgs.
ASHRAE. 2017a. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2017b. ASHRAE handbook—Fundamentals. Peachtree Corners,
GA: ASHRAE.
ASHRAE. 2019. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
ASTM. 2016. ASTM E413-16, Classification for rating sound insulation.
West Conshohocken, PA: ASTM International.
BRE. 1974. BRE DIG 162, Traffic noise and overheating in offices. Bracknell,
U.K.: Building Research Establishment.
Brugge, D., J. Durat, and C. Rioux. 2007. Near-highway pollutants in motor
vehicle exhaust: A review of epidemiologic evidence of cardiac and
pulmonary health risks. Environmental Health 6:23.
Chao, J. 2013. Berkeley Lab indoor air roundup: Natural ventilation comes
with health risks, and more. Berkeley, CA: Berkeley Lab. https://
newscenter.lbl.gov/2013/09/26/indoor-air-roundup-natural-ventilation-
comes-with-health-risks-and-more/.
Crawley, D. 1998. Which weather data should you use for energy simulations
of commercial buildings? ASHRAE Transactions 104(2):498–515.
Peachtree Corners, GA: ASHRAE.
CHAPTER 3 | 125

DOE. n.d. EnergyPlus weather data. Washington, DC: U.S Department of


Energy’s Building Technologies Office (BTO). https://energyplus.net/
weather.
Dubiel, J., M. Wilson, and F. Nicol. 1996. Decibels and discomfort—An
investigation of noise tolerance in offices. Proceedings of the CIBSE/
ASHRAE Joint National Conference, Harrogate, U.K., September 29–
October 1, 1996, 2: 184–91.
ECCD. n.d. Climate change world weather file generator for world-wide
weather data—CCWorldWeatherGen. http://www.energy.soton.ac.uk/
ccworldweathergen/.
Emmerich, S., A. Persily, W. Dols, and J. Axley. 2003. Impact of natural
ventilation strategies and design issues for California applications,
including input to ASHRAE Standard 62 and California Title 24. NISTIR
7062. Gaithersburg, MD: National Institute of Standards and Technology.
Emmerich, S., B. Polidoro, and J. Axley. 2011. Impact of adaptive thermal
comfort on climatic suitability of natural ventilation in office buildings.
Energy and Buildings 43: 2101–107.
EPA. n.d.-a. Air data—Ozone exceedances. Washington, DC: U.S.
Environmental Protection Agency. https://www.epa.gov/outdoor-air-
quality-data/air-data-ozone-exceedances.
EPA. n.d.-b. Counties designated as nonattainment for Clean Air Act’s
National Ambient Air Quality Standards (NAAQS). Washington, DC: U.S.
Environmental Protection Agency. https://www3.epa.gov/airquality/
greenbook/mapnpoll.html.
EPA. n.d.-c. NAAQS table. Washington, DC: U.S. Environmental Protection
Agency. Retrieved from https://www.epa.gov/criteria-air-pollutants/naaqs-
table on January 21, 2019.
EPA. n.d.-d. Nonattainment areas for criteria pollutants (Green Book).
Washington, DC: U.S. Environmental Protection Agency. https://
www.epa.gov/green-book.
EPA. n.d.-e. 8-Hour ozone (2008) designated area/state information with
design values. Washington, DC: U.S. Environmental Protection Agency.
https://www3.epa.gov/airquality/greenbook/hbtcw.html.
Ghiaus, C., and F. Allard. 2005. Natural Ventilation in the urban environment:
Assessment and design. London: Earthscan.
HUD. 2009. Chapter 1, “Basic overview of the environmental noise problem.”
In HUD Noise Guidebook. Washington, DC: U.S. Department of Housing
and Urban Development. https://www.hudexchange.info/onecpd/assets/
File/Noise-Guidebook-Chapter-1.pdf.
IAC Acoustics. n.d. Comparative examples of noise levels. http://
www.industrialnoisecontrol.com/comparative-noise-examples.htm.
IES. n.d. WeatherShift data files: Future weather data. http://www.iesve.com/
support/weatherfiles/weathershift.
ISU. n.d. Iowa State University: Iowa Environmental Mesonet: IEM Site
Information. Ames, IA: Iowa State University. http://
mesonet.agron.iastate.edu/sites/locate.php?network=WFO.
126 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Jensen, K.L., E. Arens, and L. Zagreus. 2005. Acoustical quality in office


workstations as assessed by occupancy surveys. Proceedings of Indoor Air
2005, Beijing, China, September 4-9, 2005: 2401–405.
Lakes Environmental. n.d. WRPLOT View™ - Freeware: Wind rose plots for
meteorological data. https://www.weblakes.com/products/wrplot/
index.html.
Linden, P., E. Arens, and N. Daish. 2015. Natural ventilation for energy
savings in California commercial buildings. CEC-500-2016-039.
Sacramento, CA: California Energy Commission. https://
www.energy.ca.gov/2016publications/CEC-500-2016-039/CEC-500-
2016-039.pdf.
Locher, B., A. Piquerez, and M. Habermacher. 2018. Differences between
outdoor and indoor sound levels for open, tilted, and closed windows.
International Journal of Environmental Research and Public Health 15(1).
https://www.researchgate.net/publication/
322582803_Differences_between_Outdoor_and_Indoor_Sound_Levels_
for_Open_Tilted_and_Closed_Windows.
Martin, A., and J. Fitzsimmons. 2000. Making natural ventilation work: BSRIA
Guidance Note GN 7/2000. Bracknell, UK: BSRIA.
McConahey, E. 2011. Natural ventilation: Who, what, when, where, why, and
how. Consulting Specifying Engineer Magazine November 2011: 24–29.
https://www.csemag.com/articles/natural-ventilation-who-what-when-
where-why-and-how/.
NOAA. n.d-a. 1981–2010 U.S. Climate normals. Retrieved from https://
www.ncdc.noaa.gov/data-access/land-based-station-data/land-based-
datasets/climate-normals/1981-2010-normals-data on June 26, 2019.
NOAA. n.d.-b. 1981–2010 U.S. Climate normals: Data tools. Retrieved from
https://www.ncdc.noaa.gov/cdo-web/datatools/normals on June 26, 2019.
Oasys Software. n.d. UNIPAC software for students and institutions. https://
www.oasys-software.com/education/. London: Arup. BEANS v.16.2 suite of
software, specifically CLIDAT v.16.2 and SUNPOS v.16.2.
Ryan, M., M. Lanchester, and S. Pugh. 2011. Noise reduction through facades
with open windows. Proceedings of Acoustics 2011, Gold Coast, Australia,
November 2–4, Paper No. 37.
UCLA. n.d. Climate consultant download page. Los Angeles: University of
California, Los Angeles’s Department of Architecture and Urban Design.
http://www.energy-design-tools.aud.ucla.edu/climate-consultant/request-
climate-consultant.php.
WeatherShift. n.d. WeatherShift™, V.2.0. http://www.weather-shift.com/heat.
WebMet.com. n.d. The Meteorological Resource Center. http://
www.webmet.com/.
WHO. 2018. Ambient (outdoor) air quality and health. Geneva: World Health
Organization. https://www.who.int/news-room/fact-sheets/detail/ambient-
(outdoor)-air-quality-and-health.
CHAPTER 3 | 127

WHO. n.d. Global Health Observatory (GHO) data: Exposure to ambient air
pollution. Geneva: World Health Organization. https://www.who.int/gho/
phe/outdoor_air_pollution/exposure/en/.
Wilson, M. 1992. A review of acoustic problems in passive solar design.
Proceedings of EuroNoise’92, Imperial College, London, September 14-
18, 1999, pp. 901–08.
Wilson, M., F. Nicol, and R. Singh. 1993. Measurements of background
noise levels in naturally ventilated buildings, associated with thermal
comfort studies: Initial results. Proceedings of the Institute of Acoustics
15(8): 283–95.
Wood, A., and R. Salib. 2013. Natural ventilation in high-rise office buildings:
An output of the Council on Tall Building and Urban Habitat
Sustainability Working Group. London: Routledge.
OUTDOOR AIR QUALITY
ASSESSMENT EXAMPLE

This appendix shows an example of an outdoor air quality assessment for


the Los Angeles county area. As shown in the multipollutant map in
Figure 3.1, this area has five pollutants in nonattainment status. A historical
trends analysis presented to a building owner based on data regarding regional
air quality was downloaded from the EPA’s Air Trends website (EPA n.d.).
Graphing the information for the relevant weather station (31080 for Los
Angeles/Long Beach/Anaheim, CA) against its equivalent national standard
immediately demonstrates the current state of noncompliance based on
measured data. These show historical trends towards improving air quality in
the area. Nitrogen dioxide (NO2), PM2.5, and PM10 have trended downward
since nonattainment status was originally determined, and they have now
dipped below the national standards threshold. Ozone (O3) continues to hover
just above the national standard for the eight-hour exposure limits.
The presentation to the owner reassured them that for the primary concerns
related to PM2.5 and PM10, outdoor air quality was adequate for natural
ventilation. The discussion around ozone acknowledged that this problem is an
issue throughout the Los Angeles Basin, where they have multiple other
facilities. None of their other facilities use the ozone air-cleaning devices
recommended by ANSI/ASHRAE Standard 62.1-2019 (ASHRAE 2019).
Ozone-related filtration is not required by any law within the state of
California. Thus, the owner determined that having natural ventilation would
pose no greater risk to the occupants than a normally filtered mechanical
ventilation system in California. The owner allowed the investigation into the
feasibility of natural ventilation to proceed to the next step.
Simultaneously, the owner then commissioned an external consultant to
complete on-site outdoor air quality testing, with the results as noted in
Table A3.1.1.
These measurements confirmed that ozone and PM2.5 continued to track
high locally, likely due to the busy urban nature of the site, construction nearby,
and the proximity of a freeway and railway line within 3000 ft (914.4 m) of the

129
130 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A3.1.1 Documentation of CO against NAAQS for the Los Angeles/Long Beach/
Anaheim, CA region from 2000 to 2016. Source data from the EPA (n.d.).
Courtesy: Arup

FIGURE A3.1.2 Documentation of lead (Pb) against NAAQS for the Los Angeles/Long
Beach/Anaheim, CA region from 2000 to 2016. Source data from the EPA (n.d.).
Courtesy: Arup
APPENDIX A3.1 | 131

FIGURE A3.1.3 Documentation of nitrogen dioxide (NO2) against NAAQS for the Los
Angeles/Long Beach/Anaheim, CA region from 2000 to 2016. Source data from the EPA (n.d.).
Courtesy: Arup

FIGURE A3.1.4 Documentation of PM10 against NAAQS for the Los Angeles/Long Beach/
Anaheim, CA region from 2000 to 2016. Source data from the EPA (n.d.).
Courtesy: Arup
132 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A3.1.5 Documentation of PM2.5 against NAAQS for the Los Angeles/Long
Beach/Anaheim, CA region from 2000 to 2016. Source data from the EPA (n.d.).
Courtesy: Arup

FIGURE A3.1.6 Documentation of ozone against NAAQS for the Los Angeles/Long Beach/
Anaheim, CA region from 2000 to 2016. Source data from the EPA (n.d.).
Courtesy: Arup
APPENDIX A3.1 | 133

FIGURE A3.1.7 Documentation of sulfur dioxide (SO2) against NAAQS for the Los
Angeles/Long Beach/Anaheim, CA region from 2000 to 2016. Source data from the EPA (n.d.).
Courtesy: Arup

TABLE A3.1.1 Comparison of On-Site Measured Outdoor Air Quality


Against NAAQS Standards
Four-Hour (9 a.m. to 1 p.m.)
Pollutant Averaging Time NAAQS Level
On-Site Measurement
Carbon Monoxide (CO) 8 hours 9 ppm (10.3 mg/m3) <1.0 ppm to 8 ppm

Lead (Pb) Rolling three-month 0.15 µg/m3 <1.0 µg/m3


average

Nitrogen Dioxide (NO2) 1 hour 100 ppb (188 µg/m3) <0.087 ppm (87 ppb)

Ozone (O3) 8 hours 0.07 ppm (137 µg/m3) 0.31 ppm

Particle Pollution PM2.5 1 year 12 µg/m3 N/A

Particle Pollution PM2.5 24 hours 35 µg/m3 38 µg/m3 to 52 µg/m3

Particle Pollution PM10 24 hours 150 µg/m3 Not measured

Sulfur Dioxide (SO2) 1 hour 75 ppb (197 µg/m3) <0.021 ppm = 21 ppb
134 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A3.1.2 NAAQS Level Definitions


Primary/ Averaging
Pollutant Level Form
Secondary Time
Carbon monoxide Primary 8 hours 9 ppm (10.3 mg/m3) Not to be exceeded more than
(CO) once per year
1 hour 35 ppm

Lead (Pb) Primary and Rolling 3-month Not to be exceeded


secondary average

Nitrogen Dioxide Primary 1 hour 100 ppb 99th percentile of 1-hour daily
(NO2) maximum concentrations,
averaged over 3 years

Ozone (O3) Primary and 8 hours 0.070 ppm Annual fourth-highest daily
secondary maximum 8-hour
concentration, averaged over 3
years

Particle Pollution Primary 1 year 12.0 µg/m3 Annual mean, averaged over 3
PM2.5 years

Secondary 1 year 15.0 µg/m3 Annual mean, averaged over 3


years

Particle Pollution Primary and 24 hours 35 µg/m3 98th percentile, averaged over
PM10 secondary 3 years

Primary and 24 hours 150 µg/m3 Not to be exceeded more than


secondary once per year on average over
3 years

Sulfur Dioxide Primary 1 hour 75 ppb 99th percentile of 1-hour daily


(SO2) maximum concentrations,
averaged over 3 years

Secondary 3 hours 0.5 ppm Not to be exceeded more than


once per year

site. Ultimately, the owner opted for a handful of sliding doors opening to
outdoor terraces for certain multioccupant classroom areas to have the option of
natural ventilation, but otherwise, the design team provided filtration for the
mechanical ventilation systems serving the entire building.
Table A3.1.2 shows the NAAQS level definitions for reference.

A3.1.1 REFERENCES
ASHRAE. 2019. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
EPA. n.d. Air quality—Cities and counties. Washington, DC: U.S.
Environmental Protection Agency. https://www.epa.gov/air-trends/air-
quality-cities-and-counties.
OUTDOOR VERSUS INDOOR
NOISE ASSESSMENT FOR A
NATURALLY VENTILATED OFFICE
This appendix discusses a field study performed to compare occupant
dissatisfaction with indoor noise as compared to outdoor noise in a naturally
ventilated office space. The goal of the study was to test the concern around
speech privacy reduction in naturally ventilated spaces.
In 2012, a field study was performed at the David Brower Center, an office
building situated along a busy street in Berkeley, CA. The measured indoor
noise level with the windows open was 50 to 58 dBLAeq in the private offices
and 51 to 52 dBLAeq in the conference room. This occurred while the outdoor
noise level was 64 dBLAeq during the 9:00 a.m. to 11:00 a.m. traffic period.
The study did not control for internal noise generation.
The post-occupancy survey results for this building showed that 84% to
86% of occupants identified people talking on the phone, people overhearing
their conversations, and people talking in neighboring areas as the primary
cause of their dissatisfaction, with an additional 60% complaining of excessive
echoing of voices or other sounds. Only 31% of respondents attributed
acoustical dissatisfaction to outdoor traffic noise, and only 24% attributed it to
other outdoor noise. These findings were consistent with the overall findings
from the occupant satisfaction database referenced in Jensen et al. (2005) and
continuously augmented and maintained by the Center for the Built
Environment at the University of California, Berkeley (CBE n.d.).
The researchers highlight that the traditional limits in the U.S., U.K., and
Australia for background noise deriving from mechanical ventilation
equipment are on the order of approximately 40 to 45 dBLAeq. Although the
indoor window-open noise conditions of 50 to 58 dBLAeq were clearly higher
than the traditionally allowable limits, the respondents complained far more
about the speech privacy issues as compared to the outdoor noise break-in
issues. The indoor noise conditions are consistent with the available
recommendations and research regarding the target indoor environment as
summarized in Section 3.2.2, “Target Indoor Noise Environment” in
Chapter 3.

135
136 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A3.2.1 REFERENCES
CBE. n.d. Occupant survey toolkit. Berkeley, CA: Center for the Built
Environment at the University of California, Berkeley. https://
cbe.berkeley.edu/resources/occupant-survey/.
Jensen, K.L., E. Arens, and L. Zagreus. 2005. Acoustical quality in office
workstations as assessed by occupancy surveys. Proceedings of Indoor Air
2005, Beijing, China, September 4-9, 2005: 2401–405.
INTERPRETING
WIND ROSE DATA
BETWEEN SITES
This appendix demonstrates the importance of selecting a source of wind
data that is as close to the wind climate conditions of a project site as possible.
When interpreting wind rose data, it is important to acknowledge that most
sites are not that close to a nationally monitored weather station that produces
wind data. Especially when a design team is considering a wind-driven natural
ventilation scheme, it is strongly recommended that an on-site weather station
be installed to log wind direction and speed so that the localized effects are
well understood as a design parameter. This will allow for data to be parsed for
time of day as well as concurrent temperature of the air.
As an example of the risks of using wind rose data alone, Figure A3.3.1
shows two weather stations that are only 30 mi (50 km) apart: Los Angeles
International Airport (which is on the coast) and the Burbank-Glendale-
Pasadena Airport (which is in an inland valley). Both experience the reversal
of wind direction during the winter months due to a phenomenon known as the
Santa Ana winds, but the actual wind direction that occurs during those periods
is heavily influenced by the surrounding geography, as shown in the
comparison of results in Figure A3.3.1.

137
138 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

(a) (b)

(c) (d)

FIGURE A3.3.1 Comparison of wind rose data at two airports in Los Angeles, CA, only
30 mi (50 km) apart: (a) historical data from the Burbank-Glendale-Pasadena Airport (all
year), (b) historical data from the Los Angeles International Airport (all year), (c) historical
data from Burbank-Glendale-Pasadena Airport (December only), and (d) historical data from
the Los Angeles International Airport (December only).
Courtesy: Iowa State University Mesonet
INTERPRETING
WIND ROSE DATA BY
OUTDOOR AIR TEMPERATURE
This appendix demonstrates that it is possible to preprocess weather files to
determine the wind direction and speed for those hours when natural
ventilation is most likely to fail (i.e., on warm days). The wind climate source
data files must be manipulated to isolate by temperature; however, both the
*.epw and system advisor model (*.sam) files used by the wind rose software
contain outdoor air weather conditions within them.
As an example of this type of analysis, the wind roses in Figure A3.4.1
demonstrate how a temperature-dependent wind rose can assist in orienting a
building to catch wind-driven air movement when the outdoor air temperatures
are warm, and thus when natural ventilation from buoyancy alone is likely to
fail. The results in Figure A3.4.1 are from an early feasibility analysis
examining weather data from the San Francisco International Airport for 1948
to 1999 (RWDI 2001). This analysis helped to confirm the northwest-facing
long façade of the San Francisco Federal Office Building.

139
140 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

(a) (b)

FIGURE A3.4.1 Comparison of wind rose under (a) cold versus (b) warm temperatures to
inform an orientation of the building that would optimally capture wind pressure on hot days.
Courtesy: RWDI

A3.4.1 REFERENCE
RWDI. 2001. Wind climate study, Final Report. New Federal Office Building,
San Francisco, California. Guelph, Ontario, Canada: RWDI.
MAPS OF
NATURAL CONDITIONING
POTENTIAL BY SEASON
The maps in Figures A3.5.1 through A3.5.4 are derived from an older
visualization of mean daily average temperatures from the National Oceanic
and Atmospheric Administration (NOAA). Similar visualization data can be
found under the “Divisional Mapping” tab on the NOAA National Centers for
Environmental Information website (NOAA n.d.). The data available is by
single month in a single year unless one clicks on the 1901 to 2000 mean. Of
greater use is broad-scale city-level mapping of monthly means from 1981-
2010, which is also available from this same website. A download of the local
design city will be necessary to generate ANSI/ASHRAE Standard 55’s
adaptive comfort model 80% acceptability limits if an *.epw file is not
available.
What can be readily seen is that the mean daily average temperatures need
to be of the order of at least 50°F to 60°F (27.7°C to 33.3°C) mean temperature
to support any significant level of natural conditioning. This is consistent with
the ANSI/ASHRAE Standard 55 lower threshold limit that states that natural
conditioning shall not be applied when outdoor prevailing mean temperature is
below 50°F (10°C).

141
142 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A3.5.1 Map of January historical weather isotherm against seasonal natural
conditioning potential for 16 U.S. cities based on 60°F to 80°F (15.6°C to 26.7°C), less than
70% rh limit for 7:00 a.m. to 7:00 p.m. (McConahey 2011).
Courtesy: CSE Magazine

FIGURE A3.5.2 Map of April historical weather isotherm against seasonal natural
conditioning potential for 16 U.S. cities based on 60°F to 80°F (15.6°C to 26.7°C), less than
70% rh limit for 7:00 a.m. to 7:00 p.m. (McConahey 2011).
Courtesy: CSE Magazine
APPENDIX A3.5 | 143

FIGURE A3.5.3 Map of July historical weather isotherm against seasonal natural
conditioning potential for 16 U.S. cities based on 60°F to 80°F (15.6°C to 26.7°C), less than
70% rh limit for 7:00 a.m. to 7:00 p.m. (McConahey 2011).
Courtesy: CSE Magazine

FIGURE A3.5.4 Map of October historical weather isotherm against seasonal natural
conditioning potential for 16 U.S. cities based on 60°F to 80°F (15.6°C to 26.7°C), less than
70% rh limit for 7:00 a.m. to 7:00 p.m. (McConahey 2011).
Courtesy: CSE Magazine
144 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A3.5.1 REFERENCES
ASHRAE. 2017. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
McConahey, E. 2011. Natural ventilation: Who, what, when, where, why, and
how. Consulting Specifying Engineer Magazine November 2011: 24–29.
https://www.csemag.com/articles/natural-ventilation-who-what-when-
where-why-and-how/.
NOAA. n.d. Climate at a glance: Divisional mapping. https://www.ncdc.noaa.gov/
cag/divisional/mapping/110/tavg/201704/1/mean.
USE OF
CLIMATE CONSULTANT FOR THE
ADAPTIVE COMFORT MODEL
Climate Consultant is a free online tool available from the University of
California, Los Angeles, that has an analysis module that evaluates the
applicability of natural conditioning for a site (UCLA n.d.).
This appendix presents a method for determining the most suitable natural
cooling strategy for a particular site. The method also recommends whether a
tight insulative envelope and a backup mechanical system are required.

A3.6.1 STEP 1. PRODUCE A BIOCLIMATIC CHART TO


REPRESENT ALL STRATEGIES
Plot the range of temperatures and humidities on the bioclimatic chart in
the Climate Consultant program.
In the following figures in this appendix, the criteria within Climate
Consultant Version 6.0, Build 12, are changed from the defaults for this
analysis:
• Select Comfort Model: Select “Adaptive Comfort Standard per
ASHRAE Standard 55-2010” (Note: it may be necessary to run in
ASHRAE Standard 55 Comfort Model first and then come back to
change to the Adaptive Comfort Model.)
• Criteria #7—Adaptive Comfort Model Using Natural Ventilation:
Change to “% Acceptability Limits to be 80%” to match ASHRAE 55-
2016 criteria.
• Criteria #8—Fan-Forced Ventilation: Change to 236 fpm (1.2 m/s) to
match ASHRAE Standard 55-2016 allowance.
• Criteria #9—Internal Heat Gain Zone: Change to 60°F (15.6°C) for
balance point temperature below which heating is needed. This input
should be changed to reflect the temperature differential from the free
heating that is anticipated to arise from the baseload of internal heat
load density. The tool unfortunately requires this estimate to be
performed outside of the software.

145
146 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A3.6.1 Frequency of occurrence psychrometric chart in °F, using Climate


Consultant 6.0, Build 12, for all active strategies active.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California.

When the chart generates, all strategies are active.


The results immediately show the applicability of natural conditioning
using the Adaptive Comfort Model for 36% of the time, the reliance on internal
heat gain for free warming for 42% of the time, the availability of passive solar
direct heat gain for 30% of the time if orientation allows, and the heating
required for 13% of the time.

A3.6.2 STEP 2. SELECT PREFERRED STRATEGIES TO


REPRESENT NATURAL CONDITIONING
The user can click on the preferred strategy name in the legend itself to
select the strategies associated with natural conditioning. For the following
case, the most minimal approach is taken, not relying on optimum solar
orientation for free heating, as this would occur only on one side of the
building. Sun shading of windows is also removed, pending actual building
design. Thus, the following strategies are selected as the minimum viable
selection for natural ventilation without careful façade design:

• Criteria #1—Comfort—ASHRAE Standard 55 Model


• Criteria #7—Adaptive Comfort Ventilation
• Criteria #8—Fan-Forced Ventilation Cooling
• Criteria #9—Internal Heat Gain
• Criteria #16—Heating
APPENDIX A3.6 | 147

FIGURE A3.6.2 Frequency of occurrence psychrometric chart in °C, using Climate


Consultant 6.0, Build 12, for all strategies active.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California.

This combination results in 99.2% of daytime hours between 8 a.m. and


5 p.m. falling into the comfort zone.
It can be seen that reliance on heating has jumped from 13.3% of hours to
18.8% of hours because of the loss of the strategy for passive solar direct gain.

A3.6.3 STEP 3. DETERMINE IF A HEATING SYSTEM


IS REQUIRED
Because the set point for internal heat gains has been set at 60°F (15.6°C),
heating is operational below that set point.
In the case of this run, only the daytime hours were previously being
explored, and already 18.8% of the time heating was necessary. Doing the
same run but for all hours in the day results in 37.1% of hours requiring
heating.

A3.6.4 STEP 4. DETERMINE IF A BACKUP MECHANICAL


COOLING SYSTEM IS REQUIRED
Under the climate circumstances in the run above, it is likely that a backup
air-conditioning system would not be required as there is less than 1% of hours
that are not covered by the passive strategies.
Doing the same exercise at a location 30 mi. (50 km) inland at the Burbank
Airport would result in a need for backup air-conditioning as noted in the
148 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

The frequency of occurrence psychrometric


chart will automatically estimate the
number of hours within each passive or
active design strategy based on criteria set
by the user. The display show enumerates
hours for both the ASHRAE Standard 55
analytical comfort model and the ASHRAE
Standard 55 adaptive comfort model, the
former also with the elevated air speed
allowance. The dots represent outdoor air
conditions.

FIGURE A3.6.3 Frequency of occurrence psychrometric chart in °F for Los Angeles, CA,
8:00 a.m. to 5:00 p.m., using Climate Consultant 6.0, Build 12.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California.
APPENDIX A3.6 | 149

following section. This is because there are 17.5% of hours are deemed not
comfortable by the natural conditioning approach alone. As a rule of thumb, if
the total hours above the boundary exceed 5% or the client’s stated allowable
exceedance annually, the design team should discuss with the client whether an
air-conditioning system will be needed.

A3.6.5 STEP 5. DETERMINE IF AN INFILTRATION-RESISTANT


ENVELOPE IS REQUIRED
If an air-conditioning system or a heating system is used, then an
infiltration-resistant and insulative envelope is required to meet the energy
codes. It is strongly recommended that an insulative envelope capable of
holding infiltration to less than 0.5 air changes per hour (ach) is used. This
would have implications for the seal quality on the operable windows that are
specified and would tend to prohibit the use of permanently open holes or
jalousie windows.
If more than 10% of hours are below 67°F (19°C) but above 60°F (15.6°C)
and a heating system is not used, it is still recommended that an infiltration-
resistant envelope is used to limit uncontrolled drafts. It is strongly
recommended that an insulative envelope capable of holding infiltration to less
than 1.0 ach is used.

A3.6.6 CAVEATS WHEN USING CLIMATE CONSULTANT FOR


THE ADAPTIVE COMFORT MODEL
There are two caveats for using the Climate Consultant representation of
the Adaptive Comfort standard. The first warning is that it takes an overly
generous understanding of the temperature boundaries. If one were to take the
mean monthly temperature from the Weather Data Summary (Under Criteria),
and apply the ASHRAE Standard 55-2017 (ASHRAE 2017) adaptive comfort
standard to generate the upper and lower 80% acceptability limits, the results
would be as noted in Table A3.6.1. It should be noted that the temperature
differential and the midpoint of range between the upper 80% and lower 80%
acceptability limits change based on the mean monthly temperature. However,
the Climate Consultant software draws the boundary lines for the adaptive
comfort design strategy to represent the year’s lowest lower bound and the
year’s highest upper bound. This technically overrepresents the number of
annual hours where the adaptive comfort model applies.
The second caveat is that Climate Consultant additionally represents
unmodified outdoor air conditions superimposed on indoor comfort
boundaries. While it shows indoor heat gain as a design strategy, the Adaptive
Comfort Model boundaries are not adjusted to accommodate the same. This
means that the upper end of the representation of the adaptive comfort range is
likely too high by the value of the anticipated heat pickup temperature
differential as previously noted (typically ~4°F to 6°F [2°C to 3°C]).
150 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

The frequency of occurrence psychrometric


chart will automatically estimate the
number of hours within each passive or
active design strategy based on criteria set
by the user. The display show enumerates
hours for both the ASHRAE Standard 55
analytical comfort model and the ASHRAE
Standard 55 adaptive comfort model, the
former also with the elevated air speed
allowance. The dots represent outdoor air
conditions.

FIGURE A3.6.4 Frequency of occurrence psychrometric chart in °C for Los Angeles, CA,
8:00 a.m. to 5:00 p.m., using Climate Consultant 6.0, Build 12.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California.
APPENDIX A3.6 | 151

FIGURE A3.6.5 Frequency of occurrence psychrometric chart in °F, using Climate


Consultant 6.0, Build 12, for Los Angeles, CA, all day.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California

FIGURE A3.6.6 Frequency of occurrence psychrometric chart in °C, using Climate


Consultant 6.0, Build 12, for Los Angeles, CA, all day.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California
152 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A3.6.1 Calculation of Upper and Lower 80% Acceptability Limits Associated with the
Weather File used in the Climate Consultant Results
Temperature,
Jan Feb Mar Apr May Jun Jul Aug Sept Oct Nov Dec
°F (°C)
Mean Monthly 57 (14) 58 (14) 58 (15) 60 (16) 63 (17) 65 (18) 68 (20) 68 (20) 68 (20) 65 (18) 61 (16) 58 (14)

Upper 78 (26) 78 (26) 78 (26) 79 (26) 80 (27) 80 (27) 81 (27) 82 (28) 81 (27) 80 (27) 79 (26) 78 (26)

Lower 65 (19) 66 (19) 66 (19) 67 (19) 67 (20) 68 (20) 69 (20) 69 (21) 69 (20) 68 (20) 67 (19) 66 (19)

It is therefore recommended that Climate Consultant be used only as a


quick feasibility checking tool, bearing its limitations in mind. A spreadsheet
would allow for a more malleable assessment of the heat pickup based on
design input as the project progresses. See Chapter 3 for more information on
spreadsheet tools.

A3.6.7 REFERENCES
ASHRAE. 2017. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
UCLA. n.d. Climate consultant, Version 6.0, Build 12. Los Angeles:
University of California, Los Angeles’s Department of Architecture and
Urban Design. http://www.energy-design-tools.aud.ucla.edu/climate-
consultant/request-climate-consultant.php.
THE PRESUMPTION OF
GLOBAL WARMING AND
RCP DEFINITION
As noted by the Nobel-Prize-winning Intergovernmental Panel on Climate
Change (IPCC), “Scientific evidence for warming of the climate system is
unequivocal (NASA n.d.-a). There is consensus among climate scientists that
the temperature of the global climate is rising as evidenced by historical
measurements from multiple international scientific bodies. These
measurements show that global temperature has risen almost (1.8°F) 1°C since
1980.
At the time of publication, 97% of climate science abstracts include a
position on anthropomorphic global warming, supporting the idea that current
warming trends are likely caused by human activity (Cook et al. 2013).
While work is underway in some nations to contain future greenhouse gas
(GHG) emissions, the prevailing operational understanding is that the climate
will continue to warm at least until 2050 due to the persistence of the
substantial levels of GHG emissions that have been discharged since the mid-
20th century. The IPPC’s Fifth Assessment Report (AR5) projects future
climate conditions based on four carbon mitigation scenarios, which together
map the range of likely outcomes, and thus projected global temperatures, as
noted in Table A3.7.1 and Figure A3.7.2.
More details on climate change and its projected impact on buildings can
be found in Aijzi and Brager (2018). Aijzi and Brager do not specifically speak
about natural ventilation concerns but are an excellent overview of climate
modeling and the creation of future weather files. While the article’s interest is
in developing localized weather files for predicting future energy use, these
same climate-shifted *.epw files are used in modeling the feasibility of natural
ventilation and natural conditioning as well.
Even in the most optimistic scenario (RCP2.6), atmospheric temperatures
would still be likely to rise for the first two decades of the life of a 50-year-old
building built in 2020. As current carbon emission trends do not appear to be
following this path, in the next best-case scenario of RCP4.5, the same
building would experience rising outdoor air temperatures for its first four
decades. These are serious concerns to be discussed with any owner
considering implementing natural ventilation in new buildings.

153
154 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A3.7.1 Temperature data from four international science institutions (NASA’s
Goddard Institute for Space Studies, NOAA National Climatic Data Center, the Met Office’s
Hadley Centre/Climatic Research Unit, and the Japanese Meteorological Agency). All show
rapid warming in the past few decades and that the last decade has been the warmest on record.
(NASA n.d.-b)

FIGURE A3.7.2 Global average surface temperature change for 21st century, taking the
extreme RCP scenarios (RCP8.5 and RCP2.6) (IPCC 2014). This figure assesses global
warming’s impact on the feasibility of natural conditioning in the future (IPCC 2014).
Courtesy: IPCC.
APPENDIX A3.7 | 155

TABLE A3.7.1 Description of Representative Concentration Pathways (RCPs)


Description of
Scenario Projected CO2 eq by
Atmospheric Carbon Greenhouse Gas Emissions
Name 2100
profile until 2100
RCP8.5 No foreseen decline High baseline emissions unmitigated ~1370 ppm

RCP6.0 Atmospheric carbon peaks Medium emissions baseline; ~850 ppm


around 2080 then declines High mitigation

RCP4.5 Atmospheric carbon peaks Medium-low mitigation ~650 ppm


around 2040 then declines

RCP2.6 Atmospheric carbon peaks Very low with negative carbon ~490 ppm peak before
around 2020 then declines emissions by ~2070 2100, then dropping
Table derived from data in Table 2, Table 4, and Figure 11 of van Vuuren et al. (2011).

A3.7.1 REFERENCES
Aijazi, A., and G. Brager. 2018. Understanding climate change impacts on
building energy use. ASHRAE Journal 60(10):24–32.
Cook, J., D. Nuccitelli, S.A. Green, M. Richardson, B. Winkler, R. Painting, R.
Way, P. Jacobs, and A. Skuce. 2013. Quantifying the consensus on
anthropogenic global warming in the scientific literature. Environmental
Research Letters 8(2). DOI:10.1088/1748-9326/8/2/024024.
IPCC. 2014. Figure SPM.6(a). In Climate change 2014: Synthesis report.
Contribution of working groups I, II and III to the Fifth Assessment Report
of the Intergovernmental Panel on Climate Change. Core Writing Team:
Pachauri, R.K., and L. Meyer. (eds.)]. IPCC, Geneva, Switzerland.
NASA. n.d.-a. Climate change: How do we know? Washington, DC: National
Aeronautics and Space Administration. https://climate.nasa.gov/evidence/.
NASA. n.d.-b. Scientific consensus: Earth’s climate is warming. Washington,
DC: National Aeronautics and Space Administration. https://
climate.nasa.gov/scientific-consensus/.
van Vuuren, D.P., J. Edmonds, M. Kainuma, K. Riahi, A. Thomson, K.
Hibbard, G.C. Hurtt, T. Kram, V. Krey, J.-F. Lamarque, T. Masui, M.
Meinshausen, N. Nakicenovic, S.J. Smith, and S.K. Rose. 2011. The
representative concentration pathways: An overview. Climatic Change
109: 5. https://doi.org/10.1007/s10584-011-0148-z.
CASE STUDY—
FRIES HOUSE, SEATTLE, WA

FIGURE CS2.1 A section through the residence.

• Architect: GB Design
• Climate: Seattle, WA

The building operates with a low energy, intelligent mixed-mode


ventilation strategy. Where seasonally possible, it will be naturally ventilated
in the office, bedrooms, living room, dining room, and winter garden, using
motorized opening vents in the façades. The building management system
maximizes free cooling, avoiding the use of air conditioning where possible.
The air-conditioning cooling is primarily taken from the surroundings via the
ground source heat pump. During hot days, it takes the heat from the building
and puts it back into the ground, keeping the building cool and returning the
energy to the ground for reuse later. During cold days, it takes heat from the
ground and puts it into the building to keep it warm.

WINTER GARDEN
The winter garden is a space that requires a critical conditioning system to
facilitate its many uses. The space will function as a greenhouse for plant
growth and at the same time be habitable. To conserve energy, the

157
158 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE CS2.2 Configuration of the winter garden.

FIGURE CS2.3 Section through the winter garden.


CASE STUDY—FRIES HOUSE, SEATTLE, WA | 159

conditioning system should conserve the least amount of energy, but also
provide comfortable conditions for space occupants throughout the year. The
space is mostly glazed; the glazing should be specifically selected to enhance
plant growth. Special attention will be paid to the surface temperatures of the
glass and how these will influence occupant comfort.
The choice of glazing is very important, as the glazing must be
multifunctional. Presently, we suggest using Solarban 70XL glass for this
space as it has excellent U-factor and SHGC properties. This glass is also very
suitable for plant growth.
A radiant floor could be used to maintain minimum space temperatures,
and this would also compensate for the glass surface temperatures.
Natural ventilation would be utilized as much as possible to maintain
adaptive temperatures in the space. During peak cooling days, ventilation must
be used to ensure the space does not overheat. Ceiling fans could also be used.
The winter garden would be controlled by a home automation system.
Occupancy sensors would determine whether the space was occupied or not. If
unoccupied the operable exterior doors would be closed, the lighting would be
switched off, and the space temperature would be reset to a minimum.
If occupied, the occupancy sensors would signal to the radiant floor to
maintain minimum temperature and, if required, operate the lighting to provide
the preset lighting levels. The next scenario would be to instigate a natural
ventilation strategy by opening the exterior doors. If further air movement is
required, then the ceiling fans will be operated. The control scenarios could be
overridden by secondary level controls such as closing the exterior doors if it
were raining.
FAÇADE AND BUILDING
CONFIGURATION ANALYSIS
The design of the façade and building envelope is a crucial element to the
success of a natural ventilation system. The operable façade requires the well-
considered multidisciplinary balancing of light, heat, and air transmission. The
envelope allows daylight to enter the building, mitigates solar transmission and
conductive heat transfer, and introduces air indoors in a controllable manner.
Whereas air-conditioned areas can rely on cooling air to absorb excessive
solar heat, naturally ventilated spaces do not have this same latitude for heat
absorption. The rule-of-thumb graphic in Figure 4.1 shows a nominal
comparison of the heat absorption capacity of various HVAC systems. A
natural ventilation system can only exist when all internal heat gains are kept
to a minimum. Thus, the façade must act as a mitigator of solar heat gain while
still allowing daylight into the space.

FIGURE 4.1 Rule-of-thumb comparison of the heat absorbing capacities of various comfort-
conditioning systems.
Courtesy: Arup.
© Arup

161
162 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

4.1 BALANCING DAYLIGHT AND NATURAL VENTILATION


ACCESS IN PERIMETER ZONES
As with HVAC zoning, the perimeter zone of a building is typically
defined as the areas within approximately 15 ft (4.6 m) of the façade.
Depending on the room configuration, this can be in the range of 10 ft to 20 ft
(3.0 m to 6.1 m). These areas are subject to the fluctuations of outdoor
temperature and the sun’s intensity in both solar heat gain and glare.
There are a few rules of thumb to consider when assessing the façade in the
context of natural ventilation:

1. The prescriptive distance between the occupant and the operable ele-
ment providing the natural ventilation intake opening and the size of
the opening itself, based on the applicable standard (See Table 4.1).
2. The size of window glazing in a side-lit room such that it is likely to
achieve a 5% daylight factor. This factor is equivalent to the brightness
one can work by when the sun’s rays fall on the external wall in
question. Per Littlefair (1988), the rule of thumb formula that can be
used is as follows:
2
DFxAx  1 – R 
W = -------------------------------------- (4.1)
MxxT
where
W = total glazed area of window in single wall, ft2 (m2)
DF = daylight factor, %
A = total area of all the room surfaces (ceiling, floor, walls, and
windows), ft2 (m2)
R = area-weighted average reflectance of the room surfaces (as-built
from component reflectance, typically light-colored wall = 0.4
to 0.6, carpet = 0.15 to 0.30, white ceiling = 0.7 to 0.8)
M = maintenance factor, typically 0.7 in an urban setting for vertical
glazing to account for dust over time
θ = angle of visible sky measured from vertical (parallel to glass)
down to first visual obstruction due to adjacent buildings
T = light transmission (taking into account any fixed shading)
3. Many concept rules of thumb note that useful daylight penetration for
diffuse light is typically
• 2.5 × window head height with no shading device
• 2 × window head height if shading device is used
Thus, in the typical office model described in Appendix A1.1, one
would require the following minimum parameters:

• 8.6 ft2 (0.8 m2) openings to comply with ANSI/ASHRAE


Standard 62.1-2016’s (ASHRAE 2016) prescriptive requirements
for natural ventilation
CHAPTER 4 | 163

TABLE 4.1 Prescriptive Requirements for Natural Ventilation Openings


Maximum Distance Minimum Size of
Standard
to Operable Element Openings
ASHRAE Standard 62.1-2016 (ASHRAE 2016b) 25 ft (7.6m) 4% of served area

ANSI/ASHRAE Standard 62.1-2019 (ASHRAE 2019b) Calculated based on room Calculated per
geometry per Section 6.4.1.6, “Location
Sections 6.4.1.2 to 6.4.1.5 and Size of Openings”

California’s Title 24-2016 (CEC 2018) 25 ft (7.6 m) 4% of served area

• 8.6 ft2 (0.8 m2) of openings to comply with ASHRAE Standard


62.1-2019’s (ASHRAE 2019b) Path A for a 10 ft × 1 ft (3 m × 0.3
m) natural ventilation opening
• 8.6 ft2 (0.8 m2) of openings to comply with California’s Title 24
(CEC 2018)
• Minimum window head height of 9 ft (3 m) to daylight the space if
external shading
• Minimum of 65 ft2 (6 m2) of glazing to achieve a daylight factor of
2% if the visible sky angle is 60°, visible light transmittance is 0.6,
and maintenance factor of 0.7 is used

The daylight factor of 2% is used, as this is typically recommended for


computer use and paper tasks associated with the office and classroom spaces
under consideration in this guide. Thus, it is important to understand that the
size of glass necessary to support daylighting for higher daylight factors may
be far larger than the size of the window opening necessary to support natural
ventilation. Furthermore, depending on the window or natural ventilation
opening type, the opening caused by a window is not equal to the glazing area.
The focus of this guide is not primarily on daylighting but the natural
ventilation scheme. The interplay between daylighting and thermal comfort in
the selection of glazed elements and shading configuration is a crucial
multidisciplinary dialogue and iterative analysis exercise. It is strongly
recommended that a daylighting analyst be part of the design team during the
period when façade configuration is under consideration.
The key to a successful natural ventilation scheme is arranging the building
to have as many occupied spaces with direct access to an operable element as
possible. It is often the case in office hierarchies that people with status have
enclosed offices along the perimeter of the building, thereby cutting off access
to natural ventilation for all internal zones. Thus, it is often more beneficial to
have open-plan offices along the perimeter, where multiple people can access a
single operable element. Shared stewardship of the window position should be
negotiated in the local workgroup.
There are cases when outdoor air is brought in over the top of exterior
offices with ceiling lids; however, this is typically suitable for ventilation and
night purge, and will not necessarily support a natural conditioning scheme if
the movement of air is not passing across the occupants’ skin surface.
164 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

When multiple perimeter zones sharing a common exhaust are proposed


for large-scale, building-wide natural ventilation schemes (such as in atrium
buildings), it is strongly recommended that a computational fluid dynamics
(CFD) study is used to determine the relative effect of the different zones on
other zones. For instance, in a case of cross ventilation, a decision by the
windward person wanting cooling air may adversely affect the comfort on the
downstream side of the building without those colleagues having any control
over what is happening.
Lastly, there is a natural ventilation typology of double-loaded corridors
that was formerly popular in school design. Classrooms had single-sided
ventilation openings but also high-level openings into a common corridor to
help to support cross ventilation. Many of these corridors were insufficiently
open to the outdoors to act as an independent exhaust path. Many of these
schemes have reverted to single-sided ventilation only (or air conditioning) to
prevent noise transmission from the adjacent spaces. Understanding the flow
path of the natural ventilation air through a building must be balanced with the
functional needs of the space, particularly concerning acoustics and pollutant
control.

4.2 ASSESSING BUILDING CONFIGURATION


FOR NATURAL VENTILATION
As described in Chapter 2, the primary mechanisms driving natural
ventilation are the stack effect or cross-ventilation. The first analysis of the
proposed building form is to determine whether either of these mechanisms is
available. It is typically useful to analyze a cross-section of the proposed
building to determine whether viable airflow paths are present. The classic
typologies for natural ventilation include:
• the atrium building
• single-zone cross-ventilation
• multizone cross-ventilation
• single-zone, single-sided ventilation
Figures 4.2 through 4.7 show some examples of these schemes at an early
concept stage, demonstrating the likely or desired airflow paths based on the
understanding of the mechanisms of natural ventilation.
It is possible to integrate multiple typologies in a single building; however, this
can add complexity to the interaction between zones and tends to require the
performance-based analysis discussed in Chapter 5. Figure 4.7 shows how each
zone type must maintain the integrity of its independent airflow path. The
classroom is a single-zone cross-ventilation with stack effect backup typology. The
upper store and music rooms use a balanced inlet, outlet terminal that functions for
multiple wind directions to supply and exhaust from a space. Although not shown
in the figure, the two rooms on the right could have each been a single-sided
ventilation scheme. The atrium could have had its own natural ventilation scheme
using doors parallel to the page and operable skylights.
As can be seen from Figures 4.2 through 4.7, as soon as a building
configuration is at a concept stage, it is quite simple to use an understanding of
airflow physics to determine the viability of natural ventilation. Refer to Chapters 5
and 6 for more information on establishing minimum opening sizes to inform the
CHAPTER 4 | 165

FIGURE 4.2 Example of atrium configuration for stack-driven natural ventilation.


Courtesy: Arup.
© Arup

FIGURE 4.3 Example of a simple single-zone cross-ventilation with stack effect.

Courtesy: Arup.
© Arup
166 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 4.4 Example of multizone cross-ventilation with availability of secondary stack


effect at clerestory.
Courtesy: Arup.
© Arup

FIGURE 4.5 Cross-ventilation to dedicated stack per floor, including upper windows for
night purge function.
Courtesy: Arup.
© Arup
CHAPTER 4 | 167

FIGURE 4.6 Single-sided, single-zone natural ventilation using stack effect.


Courtesy: Arup.
© Arup

FIGURE 4.7 Multiple airflow path types combined in a single building.


Courtesy: Arup.
© Arup

selection of window types and sizes for compliance with natural ventilation
requirements.

4.3 CONSIDERATIONS FOR PLACING


NATURAL VENTILATION OPENINGS
As wind blows onto and around buildings it creates regions in which the
static pressure is above or below that of the undisturbed airstream. (See
Chapter 2 for more details.) Positive pressure on the windward side forces air
168 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

into the building, and negative pressure on the leeward side pulls it out of the
building. Pressures on the other sides are negative or positive depending on the
wind incidence angle and the building shape. The rate of interior airflow is
determined by the magnitude of the pressure difference across the building and
the resistance to airflow of the openings. The size, shape, type, and location of
the openings, especially the inlets, determine the velocity and pattern of
internal airflow. See Chapter 5 for more information on performance-based
methods of determining airflow through openings.
When designing and placing windows and openings for ventilation, the
following factors must be considered:
• Predominant external wind and directions when the winds occur
• Construction of the building envelope and landscaping, which may
hinder or facilitate natural ventilation of the interior spaces
• Inlet location and type. These have the largest effect on the airflow
pattern through the space and the assurance of cooler air movement
across the occupied zone.
• Outlet location and type. These have little effect on indoor airflow
patterns.
• Conditions upstream of the inlet. If possible, the airflow approaching
the building inlet should not pass closely over a large hot surface (such
as a sunbaked asphalt parking lot), which will heat the incoming air.
• Window functionality. Separating the light-admitting, viewing, and
ventilating purposes of windows may be advantageous.
• Window orientation with respect to solar heat gain. For minimum solar
gain, openings should be located primarily on the polar-facing and
equator-facing sides of the building rather than the east and west sides,
and all openings shall be completely shaded between 8 a.m. and 6 p.m.
solar time during the cooling season to minimize heat gain.
• Desired air velocity and air distribution pattern. Elevation of incoming
air to cool the occupied zone is generally favorable, and changes in
indoor airflow direction tend to retard air speed.

4.3.1 Single-Sided Natural Ventilation


When openings are restricted to only one surface, ventilation will usually
be weak and is independent of the wind direction. Average internal wind speed
will not change significantly with increasing window size. One-sided
ventilation can be effective when two openings are placed on the windward
face of a building, the wind angle is oblique (20° to 70°), and the windows are
as far apart along the external face as possible (DOD 1990) if deflectors such
as wind-catching wing walls are used. If stack ventilation is used on still days
to achieve adequate fresh air supply, openings must be placed both low and
high within the zone.

4.3.2 Cross-Ventilation
Cross-ventilation provides the greatest interior velocities and the best
overall air distribution pattern. Openings in both positive and negative pressure
zones are required for cross-ventilation. For windows on adjacent walls, the
CHAPTER 4 | 169

overall room air distribution is best when the wind incidence angle is
perpendicular to the building face. For windows on opposite walls, oblique
wind incidence angles give 20% to 30% higher average velocities than
perpendicular winds (DOD 1990).
For wind-driven ventilation, outlet height has little influence on interior
airflow, but inlet height has a great effect on the airflow pattern in the room. If
stack ventilation is used for adequate fresh air supply on still days, openings
must be placed both low and high in the building.

4.3.3 Vertical Placement of Openings


There is usually an abrupt drop (up to 25%) in airspeed below the level of
the inlet sill (DOD 1990). For body cooling, openings within the occupied
height of the room are beneficial to place air speed directly across the skin.
Positive pressures built up on the windward face of the building can direct
the airflow up to the ceiling or down to the floor of the room. For instance, a
window located high on the wall directs airflow up to the ceiling because the
positive pressure built up on the building face is larger below the window than
above it. This is useful for night purge but may not be beneficial to occupant
comfort during the day, so occupant control of the openings is typically very
beneficial to achieve thermal comfort and to avoid uncomfortable or nuisance-
related wind speeds or direction.
Window type also affects vertical placement since different window types
produce different airflow patterns. See Appendix A4.1 for more information.

4.3.4 Window Shape and Size


Window inlet shape is the most important factor in determining the
efficiency of wind-based cooling. A long, horizontally shaped opening is the
best at capturing and admitting winds for more angles of wind incidence. A
horizontal window performs better than both square and vertical windows in
perpendicular winds and improves its effectiveness in winds with a 45°
incidence angle.
Square and vertical shapes give peak performance in perpendicular winds.
If the wind incidence angle is confined to a narrow band and openings can be
placed perpendicularly to the wind, then square openings will also work
effectively. However, if the wind incidence angle varies, then horizontal
openings will work more effectively under a greater variety of conditions and
should be used. Tall openings are less effective than horizontal or square
shapes for all wind incidences. They do, however, act as natural two-way
openings on still days for supporting stack ventilation without additional high-
level window openings.
The effect of window size depends on whether openings are cross-
ventilating. The following key principles can act as a rule of thumb for the
initial placement of openings:
• Single-sided ventilation: If openings are on one surface only, size has
little effect on airflow. In cross-ventilated rooms, airflow is determined
mainly by the area of the smallest openings; average indoor velocity
and number of air changes are highest when the outlet area is 1.25
times the inlet area.
170 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

• Cross-ventilation: Ventilation is more efficient for a greater number of


incidence angles when inlets are larger than the outlets. If concentrated
flow in a restricted area of the room is desired, the inlets may be sized
smaller than the outlets and placed immediately adjacent to the
occupied space to be ventilated. In general, use the largest area of
openings possible with an inlet area equal to or slightly less than the
outlet area.

Determining the effective area for a window type must be done carefully to
ensure a sound basis for subsequent calculations of volume flow rate.
Ventilation capacity is related to the way the window opens and the
surrounding head, sill, and jamb details. With top-hung windows, for example,
the triangular opening on each side of the open window is significant. Window
sills, reveals, insect screens, and internal and external blinds all have a major
impact on the final effective area that is achieved. Ensure that the required
effective area strategy is carried through into detailed design via clear
specifications in the design documents given to the contractor.

4.3.5 Insect Screens


Insect screening decreases the ventilation effectiveness of openings. The
amount of decrease in velocity varies with screen type and the incident wind
direction and velocity. Decreases in velocity are greater with lower wind
speeds and oblique winds and can be as high as 60%. (See Table 4.2.)
Because insect screening lowers the effectiveness of the ventilation
openings, its presence must be factored in when sizing windows. In the
window sizing procedure, a porosity factor is used to lower the opening’s
ventilation effectiveness when screens are used.
Screens should be in all areas where insects, rodents, or birds could prove
to be an annoyance or damage the contents of a room. Unless the specific
requirements of the local environment dictate otherwise, 14 mesh wire screens
should be used (Baker et al. 1986). This allows greater interior airflow than
higher density mesh and should prevent most insects from entering the
building. It is possible to eliminate screens on the upper floors in high-rise
buildings (above four stories) if the designer and building owner mutually
agree to its acceptability based on the reduced flying insect activity at higher
elevations.
When a building is adjacent to a highway, parking lot, or other dusty area,
screens may assist in reducing the infiltration of windborne dust, dirt, and other
debris. Using screens for this purpose, however, must not interfere with the
requirements for adequate ventilation. Screens should be maintained regularly.

4.3.6 Openings in Use


The use of opening windows tends to cause increased levels of user
satisfaction only when local occupant control is part of the strategy, as required
by the adaptive comfort model described in Section 5.4, “Determining
Acceptable Thermal Conditions in Occupant-Controlled Naturally
Conditioned Spaces” in ANSI/ASHRAE Standard 55-2017 (ASHRAE 2017a).
However, the use of the same opening for winter and summer ventilation
may be unsatisfactory in certain situations. The range of openness must be
CHAPTER 4 | 171

TABLE 4.2 Reduction in Wind Velocity Due to Insect Screens as a


Function of Incidence Angle (Baker et al. 1986)
Perpendicular Incidence Angle 67.5° Incidence Angle
Outside Velocity, Inside Velocity Reduction, Inside Velocity Reduction,
m/s (fpm) m/s (fpm) % m/s (fpm) %
0.75 (150) 0.49 (98) 35 4 (80) 47

1.23 (250) 0.87 (178) 29 0.75 (153) 39

2.5 (500) 1.33 (267) 47 1 (200) 60

3.3 (650) 1.79 (353) 47 1.33 (262) 60

3.8 (750) 2.64 (520) 31 2.23 (438) 42

Average 38 50

carefully considered. Many awning window mechanisms are designed for a


maximum of three locked positions. If winter drafts are a concern, many
smaller low-level openings under manual operation and larger consolidated
high-level openings under automatic control may be more appropriate. The
provision of automatically controlled high-level window openings can control
overall flow during the winter and can be available for night purge ventilation
while giving occupants a greater degree of daytime control. Any control
mechanisms (whether manual or automatic operation) should be both
adjustable and robust. Integrating window actuators or blind actuators requires
careful consideration. See more information in Chapter 7 for control of
motorized actuators.
It can be argued that open windows are only likely to be a security problem
at the ground- or first-floor levels of a building. Restricting the length of the
throw of the operating mechanism or actuator arms may be sufficient in many
situations. A facility to lock manually windows in a secure position should be
provided.

4.4 SELECTING OPENING TYPES FOR


NATURAL VENTILATION
After an airflow path is determined to be viable, the next step is to
determine what types of operable elements are desired, as each type may have
different effective flow areas and pressure characteristics, which are important
to consider when performing the ventilation compliance calculations in
Chapter 5.
Having defined the ventilation strategy, the vent opening type needs to be
chosen. Natural ventilation openings can be sorted into three broad types:
• Operable windows, skylights, and doors
• Operable roof monitors consisting of dampers and/or louvers
• Façade-integrated background trickle ventilators
172 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

The major advantage of operable windows and skylights is that they are
familiar to occupants and hinged versions can be made to shut tight relatively
easily. In comparison with louvers, dampers, and trickle ventilators, windows
generally have shorter crack length, more effective seals that are easier to
provide, higher closing forces, and greater ease of placement around the
perimeter due to the benefit of simultaneous daylight functions. Sealing
windows is usually achieved by expanded rubber gaskets. In aluminum-framed
windows, the tolerance achieved is usually very fine and good sealing is
achieved. However, with some steel-framed and timber windows, racking or
twisting of the frame can occur, particularly with large windows, resulting in
poor sealing at the window and the window frame interface with the
surrounding building.
Motorized dampers are widely used in mechanical ventilation systems to
modulate or shut off airflow. However, when used for natural ventilation, they
have a major disadvantage in that they do not shut as tightly as most windows
as designed. This is due to
• longer crack lengths,
• difficulties with rotating seals,
• problems with mechanical strength and closing forces, and
• poor insulation performance, with associated condensation problems.

If motorized dampers are used, they should be integrated into operable


louver elements that address rain shedding water control and should
incorporate blade seals as opposed to just metal-on-metal closure.
Trickle ventilators are designed to provide a minimum outdoor air
ventilation rate when the larger windows in the space are closed, typically
under high wind, wind-driven rain, and cold conditions. There is typically a
small counterweighted backdraft damper that modulates in response to wind
pressure to hold an established flow performance versus pressure differential
from the manufacturer that can be used to match the necessary natural
ventilation opening size. (See Figure 4.8.)
The mounting configurations for trickle vents are typically over the
windows attached to the top edge of the window, as this avoids cold air coming
into the space at a low level and adversely causing drafts and pools of cool air
at the feet. Alternatively, the window wall configuration shown in Figure 4.9
has an installed trickle vent below a finned tube convector to immediately heat
the incoming air at the glass line to offset drafts coming down the cold inner
face of the glazed surface as well as from any frame elements that are cold due
to infiltration or thermal bridging. This figure also demonstrates how the
trickle vents function as an optional ventilation path as compared to the
manually operable windows shown in Figure 4.9.

4.4.1 Window Types


Appendix A4.1 contains a reference table that demonstrates the most prevalent
types of windows and their benefits and drawbacks. Windows vary widely in terms
of their effective “throw,” geometry, and robustness. Care must be taken to ensure
that the effective free area is achieved. Most window- and door-opening
mechanisms were originally designed for manual operation. Windows used for
CHAPTER 4 | 173

(a) (b) (c)


FIGURE 4.8 Sample trickle vent appearance from (a) inside and damper operation from
outside, (b) open and (c) closed.
Courtesy: Erin McConahey

FIGURE 4.9 Installation of trickle vent below the finned tube convector in window wall
mullion system prior to fascia enclosure.
Courtesy: Erin McConahey

automatic control may require adaptation to accept motorized actuators and may
need strengthening to accommodate forces applied at different places and in
directions that may not be parallel to that in which the window opens. In particular,
windows and their actuators should be designed to accept wind forces anticipated
at the building.
Figures 4.10 through 4.23 demonstrate a variety of window types for
reference.
Wind-admitting devices that exclude solar light and heat may be achieved
by a combination of glazed and opaque elements designed to decouple
ventilation from access to views and the provision of daylight, both of which
174 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 4.11 Offset vertical pivot window.


Courtesy: Arup.
© Cody Andresen, Arup

FIGURE 4.10 Horizontal pivot


window.
Courtesy: Arup.
© Arup FIGURE 4.12 Bottom-hung window, placed for
stack effect and night purge of ceiling mass.
Courtesy: Arup.
© Arup
CHAPTER 4 | 175

FIGURE 4.13 Manual top-hung or awning type


window.
Courtesy: Arup.
© Arup

FIGURE 4.14 Top- and bottom-hung


windows with blinds for single-sided
ventilation.
Courtesy: Arup.
© Arup
FIGURE 4.15 Bank of motorized awning
windows with fall protection for high-rise application.
Courtesy: Arup.
© Arup
176 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 4.17 Awning windows in skylight


configuration.
Courtesy: Arup.
© Arup

FIGURE 4.16 Awning windows in a


high-rise building with limitation on clear
opening depth to prevent falling objects.
Courtesy: Arup.
© Arup

FIGURE 4.18 Small manual side-hung


casement window.
Courtesy: Arup.
© Arup
CHAPTER 4 | 177

FIGURE 4.19 Motorized side-hung casement window.


Courtesy: Arup.
© Simon Doling

FIGURE 4.20 Tilt and turn window.


Courtesy: Arup.
© Giles Rocholl Photography.
178 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 4.21 Unsealed jalousie-type window.


Courtesy: Arup.
© Hufton Crow_L_G

FIGURE 4.22 Louver integration FIGURE 4.23 Sliding window.


independent of view window.
Courtesy: Erin McConahey.
Courtesy: Arup.
© Graham Young
CHAPTER 4 | 179

may require physical shading at times when incoming air may be beneficial.
Alternatively, a combination wall consisting of independent elements to be put
in place by the user based on outdoor and indoor conditions might consist of:
• A sliding glass panel that provides view and light while eliminating air,
dust, insects, and rain.
• A sliding panel of opaque louvers for providing ventilation air while
protecting from the sun and light rains. (Insulated opaque panels may
also reduce the outward flow of heat in winter or at night when
ventilation for cooling is not desired.)
• A sliding panel of insect screening for providing air while eliminating
insects.

As an alternative, fixed opaque louvers may be used on the lower part of a


window wall with operable windows above for ventilation, light, and view. In
warm-humid climates such as the tropics, it is important to admit wind for
cooling while preventing the admittance of wind-driven rain. Koenigsberger et
al. (1959) report that only M-shaped fixed louvers satisfy the requirement of
keeping the rain out and allowing the breeze to enter without deflecting it
upward away from the body in the living space. The M-shaped louvers reduce
the velocity of the wind by 25% to 50% percent, with the larger reductions
occurring at higher wind speeds. The velocity reductions are equivalent to or
less than those of other louver types.

4.4.2 Roof Monitor Types


In atrium-types buildings or large spaces with multiple zones sharing
common exhaust paths, roof monitors are often used. Figures 4.24 through
4.28 show the variety and creativity that can be applied when determining the
shape of these types of rather large devices. The devices typically have some
type of operable motorized damper that responds to the air temperatures in the
space and the stack to open and close. Some also have automatic rotation to
turn the exhaust opening to always be on the leeward side of the device or
other wind-driven devices to cause suction or supply.

4.5 SELECTING WINDOW GLAZING FOR REDUCTION OF


CONDUCTIVE HEAT GAINS AND LOSSES
Natural ventilation systems designers often stress the failure modes of the
warm season because of the loss of outdoor air cooling capacity to maintain
thermal comfort. The heating period also requires careful design to both
control the incoming ventilation as per the trickle vents noted above, but in
terms of controlling heat loss just as one would do in an air-conditioned
building. It would be a shame for a naturally conditioned building to have
higher energy costs than an air-conditioned and heated equivalent. The
requirements for energy efficiency in basic building envelope performance
under ANSI/ASHRAE/IES Standard 90.1-2019 (ASHRAE 2019a) or
California’s Title 24 (CEC 2018) are not waived just because natural
ventilation is applied.
180 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 4.24 Simple louvered roof outlet opening with independent clerestory window.
Courtesy: Arup.
© Arup

FIGURE 4.25 Spinning turbine creates FIGURE 4.26 Rain-protected four-sided


suction pressure for wind-driven exhaust. opening acts as windcatcher.
Courtesy: Arup. Courtesy: Arup.
© Arup © Brett Boardman Photography
CHAPTER 4 | 181

FIGURE 4.27 Combined flue terminal roof devices combine natural ventilation supply and
exhaust within the same unit.
Courtesy: Arup.
© Peter Hyatt

FIGURE 4.28 Rotating roof exhaust outlet with integrated weather vane to ensure that
exhaust opening always points toward the leeward side of the device.
Courtesy: Arup.
© Andrew Holt
182 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 4.29 Heat loss through glazing for different WWR and U-factors.

During the heating season, the component of the building envelope most
vulnerable to unwanted heat loss by radiation and convection is glazed areas.
Insulation can substantially reduce both heat loss through glazed components
at night and undesirable solar heat admission during the day.
To illustrate the effect of the window-to-wall ratio (WWR) and the glazing
U-factor, this analysis in Figure 4.29 compares different WWR and U-factors,
using the 40% WWR as presented in ANSI/ASHRAE Standard 90.1-2019
(ASHRAE 2019a) as compared to the 65% WWR that is typically seen in
floor-to-ceiling glazing configurations.
As the WWR and U-factor increase, the heat loss increases. For energy-
efficient designs, it is recommended to maintain both WWR and U-factors as
low as possible, especially in heating-dominant, naturally ventilated
environments.

4.5.1 Estimating Heat Loss for Sizing Heating Systems


Once the cold incoming airflows from infiltration and natural ventilation
openings are determined, then it is possible to perform a heat balance
calculation, as one would on any other space. The heat load consists only of
• conductive heat losses through the building envelope and
• worst-case incoming cold air from infiltration and natural ventilation
openings combined times the safety factor allowance agreed with the
owner/client.

For the office space defined in Appendix A1.1, the comparison of heat loss
for the stated U-factor of 0.27 Btu/h ft2 ·°F for different locations is shown in
Figure 4.30. The spaces have identical dimensions and constructions, but due
CHAPTER 4 | 183

FIGURE 4.30 Heat loss from the space for the different locations noted for the space
configuration as defined in Appendix A1.1.

to the outdoor air conditions, heat loss is driven higher in cold climates. When
calculating the heat loss of a space, all internal loads are assumed to be off and
the solar radiation to the space is ignored, as the maximum heat loss occurs at
night when the temperatures are lowest.
In practice, solar radiation as an external load to the space could influence
space conditions during daytime hours, as could the internal loads that are
present in various combinations. The internal space gains influence the internal
temperature which in turn will influence any buoyancy-driven natural
ventilation on cold days.
Taking the coldest climate of New York, Figure 4.30 shows three
scenarios:
• The space fully occupied with all lighting and plug loads working
• The space fully occupied with no lighting on and the plug loads
working
• The space fully occupied with the lighting on and plug loads off

The red column shows the steady-state heat loss from the space, the orange
column shows the lighting load in the space, the gray column shows the plug
loads to the space, and the yellow column shows the occupant load to the
space. The blue column represents the total internal heat gain, which would be
184 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 4.31 The heat loss from the space and different operation of the internal loads for
New York City, NY.

required to be removed through natural ventilation or mechanical cooling.


Results for other cities can be found in Website Appendix WA4.1, “Heat Loss
Balance Comparisons,” accessible at ashrae.org/naturalventilation.
Even in this coldest climate, with a high performance insulative envelope,
there is enough baseload of heating during occupied hours to accommodate
much of the heat loss. Thus, sizing of the heating system must accommodate
the requirements during unoccupied hours inclusive of infiltration through the
window cracks under high wind, and for the period when the minimum
outdoor air openings (i.e., windows cracked open or trickle vents are open) are
activated. Total heating capacity still needs to be sized to account for the
cooling air that is brought in through these less controlled paths. See
Appendix A4.2 for more information on how to estimate infiltration.
When sizing the heating system, the design team needs to make a judgment
about how much poor occupant behavior may take place during the cold
periods. The heating system cannot be sized to accommodate windows opened
to their full natural conditioning size of openness without significant amounts
of wasted heated air escaping to the outdoors. It is typically the case that up to
a 40% safety factor on heating system sizing can be used to accommodate for
start-up. This can be available to overcome some anomalous behavior that
CHAPTER 4 | 185

FIGURE 4.32 Inside surface temperature of glazing with a SHGC = 0.25 for New York City,
NY, for both summer and winter.
Graph derived from www.payette.com/building-science/glazing-and-winter-comfort-tool.

allows more than minimum outdoor air ventilation quantities into the building
during the coldest hours. Social pressures and advisory messaging may be
necessary to preserve the heat available by keeping the larger windows closed
during these periods.

4.5.2 Radiant Impacts on Comfort


During the cold season, when there is significantly less air movement in the
space, the effects of the cold inner surface of glazing must be considered by the
design team. During the warm periods, the difference between the inside
surface temperature of the glass and the indoor air is likely to be small if low
emissivity (low-E) coatings are used. This is because the incoming air
temperature tracks identically with outdoor air temperature for the natural
conditioning scheme. In the winter case, the internal heat gains help support an
elevated internal air temperature as compared to the glass surface temperature.
This can lead to occupants experiencing radiant asymmetry.
Figure 4.32 shows an estimate of the anticipated inside surface
temperatures under a typical winter and a summer day for the high-
performance glazing considered in other analyses in this chapter.
186 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Occupants will experience radiant heat transfer with the glass because its
surface temperature is less than skin temperature. Because the winter glass
temperature at fenestration is significantly depressed in temperature as
compared to other indoor surfaces, the occupants lose heat to the glass
radiatively, and the cold surface cools the air in contact with it, causing
downdraft. Both of these phenomena may cause local thermal discomfort for
occupants close to surfaces of extreme temperature difference.
ANSI/ASHRAE Standard 55 (ASHRAE 2017a) and Chapter 9 in ASHRAE
Handbook—Fundamentals (ASHRAE 2017b) denote that the comfort of a
seated or standing occupant should be constrained to radiant asymmetry less
than the values noted in Table 4.3. Figure 4.33 also shows how the thermal
comfort estimate of the percentage of people dissatisfied (PPD) must be
increased to account for this radiant temperature asymmetry.
For these reasons, it is quite often the case that perimeter convectors are
placed at the perimeter to do three things simultaneously:
• Create a warm upward stream of heated air that directly offsets the heat
loss through the glazing and heats the lower inner surface of the
glazing.
• Place heated air in the occupant location most vulnerable to heat loss
from radiant heat transfer with the window.
• Disrupt downdrafts by providing heated air to mix with the falling
cooled airstream (see Section 4.5.3, “Cold Downdrafts in Naturally
Ventilated Spaces.”)

Additionally, radiant floors near the perimeter may be beneficial in regards


to counteracting radiant heat losses if there is a direct line of sight from the
floor to the glazed surface. If radiant floors are used, the heating output is
limited by the allowable surface temperature of the floor. Section 5.3.4.4 of
ANSI/ASHRAE Standard 55 (ASHRAE 2017a) limits the surface temperature
to 66.2°F to 84.2°F (19°C to 29°C) if people are seated with their feet in
contact with the floor. Radiative heat transfer from the floor is minimally
effective in offsetting convective heat losses within the space but can have a
secondary effect on downdraft by heating the glazed surface.
Using a dynamic thermal analysis tool can also generally provide the
internal surface temperature of glazed elements if the appropriate thermal
performance of the components has been used. These tools are used for setting
natural ventilation opening sizes in an engineered system approach, as
described in Chapter 5. Some of these tools also automatically calculate for
comfort. If so, the tools will calculate the increase in PPD, otherwise, the
radiant asymmetry effects can be applied to the PPD value manually. This is an
important step when reviewing the total comfort during the colder periods
when either natural conditioning or natural ventilation is pursued.

4.5.3 Cold Downdrafts in Naturally Ventilated Spaces


Cold downdraft, the second factor creating discomfort during the heating
season, primarily originates from occupant exposure to cold air currents falling
parallel to the cold window and spreading to a horizontal surface. Draft is
unwanted local cooling of the body caused by air movement. It is most
CHAPTER 4 | 187

TABLE 4.3 Allowable Radiant Temperature Asymmetry (ASHRAE 2017b)


Radiant Temperature Asymmetry °C (°F)

Ceiling Warmer than Ceiling Cooler than Wall Warmer than Wall Cooler than
Floor Floor Air Air

<5 (9.0) <14 (25.2) <23 (41.4) <10 (18.0)

FIGURE 4.33 Local thermal discomfort caused by radiant asymmetry as per Chapter 9 of
ASHRAE Handbook—Fundamentals.
(ASHRAE 2017b)

prevalent when the whole-body thermal sensation is cool (below neutral).


Draft sensation depends on air speed, air temperature, activity, and clothing.
Sensitivity to draft is greatest where the skin is not covered by clothing,
especially the head region comprising the head, neck, and shoulders and the
leg region comprising the ankles, feet, and legs.
Cold convective air currents, formed by warm room air hitting the cold
window surface, create discomfort at the occupant’s feet and ankles. The
strength of these currents depends on the height of the windowpane, as well as
the interior temperature of the glass. In the case of natural ventilation, there are
two other sources of potential draft arising from ventilation air entering the
building:
• wind-driven infiltration through sealed window cracks
• wind- or stack-driven ventilation air intentionally coming through the
designed minimum outdoor air openings

In these cases, the air tends to fall to the floor immediately upon entering
the building, so the perimeter convectors are a key mitigating factor in
188 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

preventing ankle level draft for low-level minimum outdoor air openings. If
trickle vents are installed at the head of the window, the minimum outdoor air
quantities will exacerbate the velocity of the downdraft already present from
the cooled surface of the glazing. If inner surface temperatures or large
amounts of minimum outdoor air are necessary (as in the case of classrooms
and conference rooms), it is recommended that a localized CFD analysis is
performed to confirm the sizing of the heating elements in the perimeter
convectors and determine whether the convectors can be passive or should
incorporate fans to encourage mixing. Alternatively, overhead recirculation
type heaters can be used in extreme cases of draft, often observed as under
desk heaters in offices.

4.5.4 Condensation without Offsetting Air Movement


Another aspect of cold weather design for natural ventilation schemes
revolves around thermally broken mullions and frames. Some may incorrectly
believe that thermal breaks are not necessary when there are already cracks or
leakage through the envelope due to natural ventilation. Façade elements
should be thermally broken to avoid unforeseen areas of condensation, just as
they would be in a sealed building.
It is strongly recommended that the internal estimated temperature and
absolute humidity be entered into a façade condensation analysis tool, such as
THERM by Lawrence Berkeley National Laboratory (LBNL n.d.). If the
results show a risk of condensation, then additional work with the façade
consultant may be necessary to adjust thermal conductivity and detailing for
this installation. Otherwise, a localized source of heating air with mechanical
air movement may be required.

4.6 SELECTING WINDOW GLAZING


FOR SOLAR CONTROL
As noted in both Chapters 1 and 2, controlling internal heat gain must
include an understanding of what solar transmission is imposed on the space.
Natural conditioning has significant limitations on heat absorption capability
as the outdoor air temperature approaches the indoor temperature. Controlling
solar gains during these periods becomes essential to the success of a natural
conditioning scheme.
Refer to the typical traditional office space in Appendix A1.1 with a south-
facing window in June to see the comparison of the effects of different solar
heat gain coefficients (SHGCs) on the total heat load in the space. Recall that
this is a 40% WWR to match both ANSI/ASHRAE/IES Standard 90.1
(ASHRAE 2019a) and California’s Title 24’s (CEC 2018) prescriptive
maximum.
As compared to the recommendation of McConahey (2008) to limit solar
heat gain to 4 W/ft2 (43 W/m2) the 40% WWR for an SHGC of approximately
0.4 or less can help achieve the purpose. In hotter areas with high solar
intensity during afternoon hours, an effective SHGC that is further reduced due
to external shading will be beneficial for indoor comfort.
Each glazing type provides differing amounts of resistance to solar heat
gain, with clear single pane glass typically transmitting 80% of its solar heat,
CHAPTER 4 | 189

TABLE 4.4 Comparison of SHGC Effect on Solar Heat to be Absorbed for the
Office Space Defined in Appendix A1.1
Solar Influence % of
SHGC
Total Heat Load Solar Btu/h·ft2 Solar, W/ft2 (W/m2)

0.5 52% 16 4.8 (52)

0.375 43% 12 3.6 (39)

0.25 34% 8.3 2.4 (26)

FIGURE 4.34 Breakdown of cooling loads in the summer for different SHGC values for a
“traditional” design.

or rather an SHGC of 0.80. Reflective and absorbing glazing types can reduce
cooling loads 15% to 30% below that of clear glass with some reduction in
transmitted light. Single pane (body-tinted) heat-absorbing glazing is less
effective than reflective glazing because it absorbs the solar heat into the glass,
thereby increasing the heat convected and radiated into the internal space.
Special solar control coatings and films laminated between panes can reduce
transmitted solar heat gain even further.
190 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 4.35 Sankey diagram of an insulated glazing unit with a SHGC of 0.25.

Better performance can generally be obtained with the heat-rejection


element in place as the exterior panel of a double-glazed window. This often
allows a protected inner film or coating to function as high-performance solar
control, as noted in Figure 4.35. This figure shows that for the 100% of solar
heat that impinges on the outer pane, 19% comes into the space as direct solar
transmission and 6% comes in as longwave warming radiation from the inner
pane’s heated surface. The rest of the heat is either reflected from the initial
surface itself or absorbed by the outer pane and reradiated mostly to the
outdoors due to a spectrally selective, low-E coating on one of the inner
surfaces. This is the physics behind most high-performance solar control
glazing on the market. Further reduction in solar heat gain can be achieved by
adding frit to the external pane; however, this can sometimes obstruct views,
reduce daylight, and even further add to solar heat absorption within the
external pane.
Because typically lower SHGC comes with lower visible light
transmittance (VLT), the design team must weigh the relative benefit of
daylighting versus solar control, specifically in the case of natural
conditioning. For example, in the case of air-conditioned spaces, one can
CHAPTER 4 | 191

FIGURE 4.36 Solar radiation load to a space in W/ft2 (W/m2) floor area for different
SHGCs and different glazing areas (40% and 65%) for both summer and winter for Los Angeles,
CA.

compare energy saved from HVAC energy reduction to energy used for
electrical lighting. In those cases, clearer glazing with effective exterior
shading may be justified over expensive high-performance glazing itself via a
life-cycle cost analysis. In the case of natural conditioning, however, the
benchmark is not energy but comfort. There are no other means to absorb
indoor heat except through outdoor air. Once the internal heat loads from
functional necessity are defined (including lighting load when daylighting is
not available), then this tells the design team how much solar heat may enter
the space and keep it within the acceptable comfort range. It is strongly
recommended that paired daylight and thermal comfort analysis be performed
for a typical perimeter space in each orientation to determine the glazing
performance best suited to the building.

4.7 LIMITING WWRS TO CONTROL HEAT GAIN


As previously noted, both ANSI/ASHRAE/IES Standard 90.1-2019
(ASHRAE 2019a) and California’s Title 24 (CEC 2018) have a prescriptive
baseline of 40% WWR. To test the relative effect of a WWR in conjunction
with an SHGC, the example in Figure 4.36 assumes a base case glazing area of
40% WWR and a glazing area of 65% WWR typical of most present-day
192 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

designs that use floor-to-ceiling glass. Results for both summer and winter for
Los Angeles, CA, are shown. The total winter solar radiation is higher than the
total summer solar radiation due to the sun being lower and therefore
penetrating the space further with higher solar intensity. Results for other cities
can be found in Website Appendix WA4.2, “Solar Radiation Load
Comparisons,” accessible at ashrae.org/naturalventilation.
Following the rule of thumb in McConahey (2008) of solar load to the
space being less than 4 W/ft2 (43 W/m2) for naturally conditioned spaces,
one can easily see that none of the 65% WWR options meet this criterion for
this solar climate. Furthermore, the 40% WWR and SHGC of 0.5 also do not
meet this requirement for the summertime. In the winter condition, solar heat
gain can often be welcomed as free passive heating, therefore the limitation
on 4 W/ft2 (43 W/m2) in that season does not have to apply.

4.8 CONFIGURING SHADING AND BLINDS IN


NATURAL VENTILATION SCHEMES
Solar control and shading can be provided by a wide range of building
components including:
• Landscape features such as mature trees or hedgerows
• Exterior elements such as overhangs or vertical fins
• Horizontal reflecting surfaces called light shelves
• Interior glare-control devices such as Venetian blinds or adjustable
louvers

External shading devices and interior blinds both have two separate but
interlinked functions: controlling light intensity for visual comfort (glare) and
controlling solar heat intensity for thermal comfort. In the context of natural
conditioning, it is important to understand these as separate phenomena. Blinds
are sometimes also used for privacy, and exterior shading devices often offer
the opportunity for architectural differentiation. However, this discussion will
revolve around the impact of these devices on natural ventilation and natural
conditioning schemes only.
The fundamental challenge when designing shading is always the tension
between solar harvesting of daylighting versus solar control. Solar harvesting
in the form of light shelves or interpane deflectors produces better daylight
penetration but runs the risk of glare and higher thermal discomfort in a
naturally conditioned space. Solar shading does the inverse: poor daylight
penetration, reduced blind use, and reduced heat gain to the space. (See
Figure 4.37.)
From a natural conditioning perspective, the following design
considerations must be balanced.
1. External shading devices in warm, sunny climates are designed to pre-
vent a portion of the incoming solar heat from falling onto the external
pane of glass, thereby limiting excess solar gain in the space in the
expectation that it can remain within allowable indoor adaptive comfort
conditions.
CHAPTER 4 | 193

FIGURE 4.37 Difference between solar harvesting and solar shading as concepts.
Courtesy: Arup.
© Arup

(a) (b)

FIGURE 4.38 Different shading devices: (a) reflective staggered overhangs and (b) parallel
fins.

2. In cold, temperate climates, low winter sun may be welcome for


passive solar heating but can create a problem in terms of controlling
glare and reducing contrast ratios.
3. Blinds do not keep any heat out of the space, but they can provide glare
control and shading for an occupant who would have otherwise been
sitting in a hot sun patch.
4. For productivity and learning concerns, nearly all climates require a
means to control and diffuse daylight to provide natural illumination of
internal spaces.
5. Exterior shading devices are particularly effective at allowing clearer
(high VLT) glass while still offering solar control. They do not negate
the need for internal blinds for glare control under low-angle sun.
6. Blinds typically must travel along with the glazing of operable
elements to prevent fluttering noise or tearing the blinds’ fabric
material.
194 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 4.39 Sun shading chart with shading occluding above 60° above the horizon,
south-facing window in northern hemisphere, using Climate Consultant 6.0, Build 12. Burbank,
CA, June 21 to December 21.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California

FIGURE 4.40 Sun shading chart with overhang occluding above 60° above the horizon,
west-facing window in northern hemisphere, using Climate Consultant 6.0, Build 12. Burbank,
CA, June 21 to December 21.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California
CHAPTER 4 | 195

The following general advice is offered:


1. Use fixed overhangs to control high angle direct beam solar radiation.
Indirect (diffuse) radiation should be controlled by other measures,
such as high-performance glazing.
2. Do not worry about shading north pole or south pole facing glass in
nontropical latitudes since it receives very little direct solar gain. In the
tropical latitudes, disregard this rule of thumb.
3. Remember that shading effects daylighting; consider both
simultaneously. For example, an external light shelf bounces natural
light deep into a room through high windows while shading lower
windows. It may be desirable for an internal light shelf to incorporate
blinds from the light shelf downwards to provide glare control for the
occupant closest to the window while allow daylighting for the
occupants toward the interior.
4. Carefully consider the durability of shading devices. Over time,
operable shading devices can require a considerable amount of
maintenance and repair.
5. When relying on landscape elements for shading, consider the cost of
landscape maintenance and upkeep on life-cycle cost.
6. When relying on fixed external elements, consider the operations and
maintenance cost of how birds and other vermin may interact with
these elements.

Designing effective shading devices will depend on the solar orientation of


a building façade. For example, simple, fixed overhangs are very effective at
shading north- or south-pole-facing windows (depending on the hemisphere)
in the summer when sun angles are high. The optimal length of an overhang
depends on the size of the window and the relative importance of allowing or
preventing heat from entering the building for the range of solar angles. An
overhang sized to fully shade a south-facing window in summer will also
shade the window in spring when some solar heat may be desirable.
In certain climates, the same overhang may be as effective if applied to the
east and west faces, as in Los Angeles, CA. The Climate Consultant tool
allows overhang and fin simulation superimposed on the sun path diagram, as
viewed from inside the window (UCLA n.d.). In this tool, the shading
exclusion can be dynamically set and the “bearing angle” (i.e., the orientation
of the façade) to determine the shading effectiveness.
For about 20% of the time during the June to December period, both the
south-facing and west-facing windows are exposed to prevailing solar intensity
(for an occlusion down to 50° degrees).
As demonstrated in Table 4.5 and Figure 4.41, the use of fins shades the
window far better. In this climate, the overhangs are approximately as effective
at the south and west but given that peak outdoor air conditions exist
simultaneously with the sun on the west face in the afternoon, better shading is
desirable. Thus, for these orientations in hot climates, it may be more
appropriate to apply a lower WWR on that face of the building, apply a dense
array of vertical fins or perforated metal parallel-to-glass shading devices, or
196 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 4.5 Comparison of Exposed versus Total Sun Hours for Burbank, CA, for a
Half Year for an Overhang versus Fin Configuration
Exposed Hours > Total Sun Hours >
Configuration % Exposed
68°F (20°C) 68°F (20°C)
Overhang shading above 586 1709 34%
60° above horizon: south
facing

Overhang shading above 660 1709 39%


60° above horizon: west
facing

Fins shading beyond 60° 337 1709 20%


horizontal view range: west
facing

FIGURE 4.41 Sun shading chart with vertical fins occluding beyond a 60-degree horizontal
view angle between fins, west-facing window in northern hemisphere, using Climate Consultant
6.0, Build 12. Burbank, CA, June 21 to December 21.
Courtesy: UCLA’s Energy Tools Design Group.
© Regents of the University of California

apply operable shading devices that deploy each afternoon to become a


protective solar barrier.
In the northern hemisphere, these general rules of thumb apply:

1. A roof overhang is the simplest and most maintenance-free exterior


shading device. They are most effective on the south side, but can also
CHAPTER 4 | 197

be used on the southwest-, southeast-, and north-facing façades. On


east- or west-facing walls, overhangs must be very deep to be effective.
2. Careful detailing horizontal exterior shades will maintain the
ventilation efficiency of the openings.
3. Vertical fins or wingwalls are the most appropriate shading devices for
east- and west-facing openings, which receive sun at low angles, and
for southeast and southwest openings in combination with horizontal
shading. Wingwalls can also increase the ventilation potential of the
building by diverting air into the openings if they are properly
designed.
4. Operable exterior shutters, rolldown shades, and blinds can provide
effective shading on any façade. They are most useful on east and west
openings, which are difficult to shade with overhangs or vertical
shading devices.

The thermal performance of closed exterior shutters depends on how well


the heat absorbed by the shade is dissipated to the outdoor air. For this reason,
light-colored reflective shutters are preferred in hot climates. For naturally
ventilated buildings, the specification of such devices should be treated with
care since air movement to the building interiors is reduced when the shutters
or shades are in their closed position, even if the shutter or shade is perforated.
Site obstructions such as buildings and trees may provide effective
building or window shading. Analysis of the site using a sun path diagram is
recommended to determine when such shading occurs.

4.9 REFERENCES
ASHRAE. 2016. ANSI/ASHRAE Standard 62.1-2016, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2017a. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2017b. ASHRAE handbook—Fundamentals. Peachtree Corners,
GA: ASHRAE.
ASHRAE. 2019a. ANSI/ASHRAE/IES Standard 90.1-2019, Energy standard
for buildings except low-rise residential buildings. Peachtree Corners, GA:
ASHRAE.
ASHRAE. 2019b. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
Baker, P.H., R.D. Heap, and S. Sharples. 1986. Airflow through perforated
screens at small pressure differences. Building Services Engineering
Research and Technology 7(2):96–97.
CEC. 2018a. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
198 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

DOD. 1990. Military handbook 1011/2: Cooling buildings by natural


ventilation. MIL-HDBK-1011/2. Alexandria, VA: U.S. Department of
Defense’s Department of the Navy.
Koenigsberger, O., J.S. Millar, and J. Costopolous. 1959. Window and
Ventilator Openings in Warm and Humid Climates. Architectural Science
Review 2(2):82–96.
Littlefair, P.J. 1988. Average daylight factor: A simple basis for daylight design.
Building Research Establishment (BRE) Information Paper IP 15/88. Watford,
U.K.: Building Research Establishment.
LBNL. n.d. THERM. Berkeley, CA: Lawrence Berkeley National Laboratory.
https://windows.lbl.gov/software/therm.
McConahey, E. 2008. Mixed mode ventilation: Finding the right mix.
ASHRAE Journal 50(9):36–48.
OPERABLE ELEMENT TYPES
This appendix presents a compilation of operable element types, as
originally presented in Table 3.2, “Main Window Types and Actuator
Options” in CIBSE Applications Manual AM10, Natural Ventilation in Non-
Domestic Buildings (CIBSE 2005).

199
200 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A4.1.1 Main window types and actuator options (CIBSE 2005).
Courtesy: CIBSE

A4.1.1 REFERENCE
CIBSE. 2005. CIBSE applications manual AM10, Natural ventilation in non-
domestic buildings. London: The Chartered Institute of Building Services
Engineers.
ESTIMATING INFILTRATION
This appendix provides useful information for estimating infiltration for
natural ventilation.
It is accepted that infiltration will occur in all buildings. This is air leakage that
comes into the building through cracks in the building envelope, whether in opaque
systems or at window edges. The National Fenestration Rating Council (NFRC)
has established a testing protocol called ANSI/NFRC 400 (NFRC 2017). The
testing is done under 1.57 lb/ft2 (75 Pa) of differential pressure for residential and
commercial products and allowed to be done under 6.24 lb/ft2 (300 Pa) if a Heavy
Commercial (HC) or Architectural Window (AW) rating under the North
American Fenestration Standards (NAFS) is required.
It should be noted that both Sections 5.4.3 and 5.8.3 of ASHRAE Standard
90.1 (ASHRAE 2019) and Section 110.6(a)1 of the California Energy Code (CEC
2018) limit the amount of air leakage as well. The limitations as defined by
ASHRAE (2019) and CEC (2018) are noted in Tables A4.2.1 and A4.2.2.
Once the tested air leakage and test pressure are known, the designer can
estimate the equivalent crack area. This crack area can then be used in the orifice
equation (Equation A4.2.1) to determine the worst-case air leakage based on wind
conditions and anticipated developed pressure differentials on each façade at each
floor. See Chapter 3 for more information on estimating pressure coefficients from
wind and at openings. The following calculations are derived from the orifice
equation represented in ASHRAE Handbook—HVAC Applications (ASHRAE
2015).

V = 776CA 2p
---------- (A4.2.1 I-P)

V = CA 2p
---------- (A4.2.1 SI)

where
V = volumetric flow, cfm (m3/s)
C = flow coefficient, typically 0.65

201
202 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A4.2.1 Allowable Air Leakage Based on 1.57 lb/ft2 (75 Pa) Test Protocol
(Unless Otherwise Noted) (ASHRAE 2019)
Allowable Air Leakage,
Fenestration Type
cfm/ft2 (L/s/m2)
Glazed swinging entrance doors, glazed power-operated sliding entrance 1.0 (5.1)
doors, glazed power-operated folding entrance doors, and revolving doors

Curtain wall and storefront glazing 0.06 (0.3)

Unit skylights having condensation weepage openings 0.3 (1.5)

Nonswinging doors intended for vehicular access and material transportation 1.3 (6.6)

Other opaque nonswinging doors, glazed sectional garage doors, and upward 0.4 (2.0)
acting glazed non-swinging doors

All other products tested at a pressure of at least 1.57 lb/ft2 (75 Pa) in 0.2 (1.0)
accordance with AAMA/WDMA/CSA 101/I.S.2/A440 or NFRC 400

All other products tested at a pressure of at least 6.24 psf (300 Pa) in 0.3 (1.5) for 6.24 lb/ft2
accordance with AAMA/WDMA/CSA 101/I.S/A440 (300 Pa) testing protocol

TABLE A4.2.2 Maximum Allowable Air Leakage Requirements under


Test Protocol of NFRC 400 or ASTM E283 at a Pressure Differential of
1.57 lb/ft2 (75 Pa) (CEC 2018)
Class Type Rate

Windows (cfm/ft2) of window area All 0.3

Residential doors (cfm/ft2) of door area Swinging, sliding 0.3

All other doors (cfm/ft2) of door area Sliding, swinging (single door) 0.3

Swinging (double door) 1.0

A = flow area (or leakage area), ft2 (m2)


∆p = pressure difference across path, in. of water (Pa)
r = gas density in path, lb/ft3 (kg/m3)

The I-P version of this equation can typically be reduced to

Q  cfm infiltration  = 2610  A crack  p (A4.2.2)

The design team may estimate crack area as

A crack = wall area  relevant A c  A w ratio (A4.2.3)

where
Acrack = wall area × relevant Ac /Aw ratio
APPENDIX A4.2| 203

TABLE A4.2.3 Crack versus Wall Area Ratios by


Wall Construction Tightness (ASHRAE 2015)
Exterior Building Wall Construction Tightness
Tight Average Loose Very Loose
Ac/Aw ratio 5.0 × 10–5 1.7 × 10–4 3.5 × 10–4 1.2 × 10–3

Ac/Aw ratio = area of the crack divided by the area of the wall using Table 2,
“Typical Flow Areas of Walls and Floors of Commercial
Buildings” from ASHRAE Handbook—HVAC Applications
(ASHRAE 2015), excerpted in Table A4.2.3, for an estimate of
exterior building wall construction tightness

A4.2.1 REFERENCES
ASHRAE. 2015. Chapter 53, Fire and smoke control.In ASHRAE handbook—
HVAC applications. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2019. ANSI/ASHRAE/IES Standard 90.1-2019, Energy standard
for buildings except low-rise residential buildings. Peachtree Corners, GA:
ASHRAE.
CEC. 2018. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
NFRC. 2017. ANSI/NFRC 400, Procedure for determining fenestration
product air leakage. Greenbelt, MD: National Fenestration Rating Council.
CASE STUDY—
ICITY,
MOSCOW, RUSSIA

FIGURE CS3.1 Rendering of ICITY.

• Architect: Helmut Jahn


• Climate: Moscow, Russia

The ICITY City Project is located in Moscow, Russia, and lies just to the
northeast of Moscow City, the new high-rise district in Moscow. The design
brief provided by the MR Group requests a Class A office development with a
total above-grade area of 175,100 m2 (1,751,000 ft2) (excluding podium
parking and mechanical). The project is to have two towers: Tower 1 has a
height limit of 126.9 m (418 ft), and Tower 2 is to be designed to fulfill the

205
206 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE CS3.2 Prevailing mean for nine months.

outstanding buildable area with a desired height in the range of 236 m (778 ft).
The location of the two towers was generally established by an insolation
study, which positioned the tall tower at the northern end of the site and the
lower tower at the site’s southern half.
The project site is 2.06 hectares and is surrounded by elevated roads, or
“flyovers.” The towers will stand on an elevated podium or stylobate that has
its top level above the level of the flyovers so that the first office levels are
18 m to 20 m (60 ft to 66 ft) above the ground level. The podium will contain
parking, retail, and ground-level entry, as well as delivery and waste removal
functions. The top podium level will be landscaped and provide outdoor
relaxation space for the building occupants.

OFFICE
The primary use is to be office space located in the towers. The desired
lease span (i.e., distance from the central core to the windows) is 10 m to 12 m
(33 ft to 39 ft), with 14 m to 15 m (46 ft to 49 ft) still acceptable. The ideal
floor plate size for leasing is 2500 m2 (25,000 ft2) (2200 m2 [22,000 ft2]
minimum), and for sale, a floor plate size of 2000 m2 to 2200 m2 (20,000 ft2 to
22,000 ft2) is preferred. The desired floor to floor height is 4.2 m (14 ft), and
with this, a clear office ceiling height of 3 m (10 ft) will be achieved.
CASE STUDY—ICITY, MOSCOW, RUSSIA | 207

Leasing divisions should allow for 1 to 4 units per floor with areas in the
range of 300 m2 to 500 m2. Ideally, floor plates could be sold as whole units,
but smaller unit sales should be considered (150 m2 to 200 m2). The office
layouts indicated in the architectural concept design demonstrate these size
requirements.

NATURAL VENTILATION STRATEGY


The two towers are naturally ventilated through the façade. The typical
façade module is 2.8 m (9.3 ft) wide and consists of a 2100 mm (83 in.) fixed
glass panel and a 700 mm (27 in.) hinged window panel. The fixed glass panel
is double glazed and features high-performance and heat-reflecting properties.
The narrow operable window features an exterior perforated, stainless steel
panel that protects against the sun, wind, and rain. This panel also contains
soundproofing elements to minimize the noise from the surrounding highway
and ring road. The interior side of the hinged window consists of a translucent
printed screen pattern applied to heat-insulating double glazing.
In total there are 60 openings around the perimeter of a typical floor.
The side-hinged windows can be electronically opened inward on each
floor to provide individually controlled natural ventilation. When the windows
are open, fresh air is drawn into the offices through the fixed perforated panels.
A control panel in each office can be used to individually control lighting,
ventilation, temperature, and the operability of the windows and blinds. At
night, the building management system (BMS) can electronically open the
hinged windows to provide night cooling and modulate temperature
fluctuations within the building.
There is a natural ventilation system designed for the towers. Around the
perimeter of the towers, there are specifically designed openings that can be
both manually and automatically opened or closed. When utilizing the natural
ventilation system, there is a possibility of using only natural ventilation for
1500 hours during the occupied periods. This will also reduce the building’s
energy consumption.
From Figure 1 we can see that March, October, and November have a
majority of hours that are categorized as “TOO COLD,” and therefore, natural
ventilation should not be considered during these months, as well as for
December, January, and February. September has more than 90 hours when it
is categorized as “TOO COLD,” and we would not recommend utilizing
natural ventilation during this month.
NAVIGATING
VENTILATION COMPLIANCE
Navigating a natural ventilation scheme through compliance with codes and
standards can be tricky, as most of these documents are written with mechanical
ventilation systems in mind. For human health and well-being, ventilation (i.e., the
provision of outdoor air to occupied spaces) is governed by codes that either define
the requirements explicitly or reference ANSI/ASHRAE Standard 62.1 (ASHRAE
2019) in whole or in part.
In addition to ventilation compliance, energy use in most jurisdictions is also
governed by a code, which is typically either ANSI/ASHRAE/IES Standard 90.1
(ASHRAE 2019a) or a state-generated code such as California’s Title 24, Part 6
(CEC 2018). These energy codes will often prescribe how natural ventilation is to
be modeled in performance methods of compliance.
The voluntary U.S. Green Building Council’s (USGBC) LEED® v4 rating
System (USGBC 2013) recognizes natural ventilation compliance based on
diagrams informed by CIBSE Applications Manual AM10, Natural Ventilation in
Non Domestic Buildings (2005) and CIBSE Applications Manual AM13, Mixed
Mode Ventilation Systems (2000) application manuals for the “Minimum Indoor
Air Quality Performance” prerequisite and the “Enhanced Indoor Air Quality
Strategies” credit. In conjunction with proving these credits, LEED recognizes that
modifications to the energy calculations performed under ANSI/ASHRAE/IES
Standard 90.1 (2019a) and California’s Title 24, Part 6 (CEC 2018) may be
required for documenting the “Minimum Energy Performance” prerequisite and
the “Optimize Energy” credit.
This chapter will discuss the application of the various ventilation compliance
methods and any associated energy code calculations that require attention or
modification to ensure consistency for submission to LEED.
As a preemptive word of warning, complying with the natural ventilation
requirements described in this chapter does not imply any likelihood of compliance
with the comfort standard, ANSI/ASHRAE Standard 55 (ASHRAE 2017a). At
high outdoor air temperatures, the amount of air necessary to provide a heat-
absorption function is far higher than the amount of air required to meet the
minimum ventilation requirements. The design team should never assume that
complying with ANSI/ASHRAE Standard 62.1 (ASHRAE 2019b) will result in a
comfortable space. See Chapter 6 for more information on how to size openings

209
210 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

for natural conditioning and demonstrate compliance with ANSI/ASHRAE


Standard 55 (ASHRAE 2017a).

5.1 INTRODUCTION TO AVAILABLE


VENTILATION COMPLIANCE PATHS
There are various paths available to demonstrate a natural ventilation scheme’s
compliance with the requirements of either ANSI/ASHRAE Standard 62.1
(ASHRAE 2019b) or California’s Title 24 (CEC 2018), as noted in Table 5.1.
Most of the prescriptive paths are designed for single-room natural ventilation
solutions. If multizone natural ventilation schemes (such as atrium buildings) are
proposed, then only the engineered path approach that uses computer simulation is
appropriate.
The simplified calculation of airflow through a single opening in a single zone
can be found in Appendix A2.1 for buoyancy-driven ventilation and in
Appendix A2.2 for wind-driven ventilation. These derive from ASHRAE’s
simplification of building physics calculations found in CIBSE (2005). The more
complex foundational equations for flow through an opening can be found in
CIBSE Applications Manual AM10, Natural Ventilation in Non Domestic
Buildings (2005).
It should be noted that under all circumstances governed by ANSI/ASHRAE
Standard 62.1 (ASHRAE 2019b) providing a mechanical ventilation system is a
requirement, even if it can be proved that an engineered natural ventilation system
can achieve minimum outdoor air quantities required under Table 6-1, “Minimum
Ventilation Rates in Breathing Zone,” in ANSI/ASHRAE Standard 62.1 (2019).
This requirement can be found in Section 6.1.3 and Section 6.4 of the standard
(ASHRAE 2019). The only exceptions that would allow for the removal of the
mechanical ventilation system are, per ASHRAE Standard 62.1, “where natural
ventilation openings that comply with the requirements of Section 6.4 are
permanently open or have controls that prevent the openings from being closed
during periods of expected occupancy” or where “the zone is not being served by
heating or cooling equipment” (ASHRAE 2019).
The design of the mechanical ventilation system must comply with all the
relevant clauses within ASHRAE Standard 62.1 (ASHRAE 2019) outside of
the prescriptive natural ventilation opening placement requirements of
Section 6.4. The required mechanical ventilation system is not a full cooling/
heating system, but it is the backup system that ensures that ventilation of the
occupants is provided even under scenarios, such as when

• still conditions outdoors are insufficient to drive buoyancy


• outdoor wind conditions are so strong that occupants must close the
windows
• a motorized device used within the natural ventilation scheme is not
operational
• changes in the function of the space require that operable windows are no
longer feasible
• outdoor air quality conditions are so poor that the windows must be closed
for health reasons
CHAPTER 5 | 211

TABLE 5.1 List of Ventilation Standards Compliance Paths for Natural Ventilation
Method Path Summary Basis of Path
Prescriptive single zone ANSI/ASHRAE Standard 62.1-2016 4% total opening method Historical rule of thumb
(ASHRAE 2016)

Prescriptive single zone Section 6.4, “Natural Ventilation Two opening, flow adjusted Lookup table for two
Procedure” of ASHRAE Standard 62.1- percentage method vertically spaced openings
2019 (ASHRAE 2019b): Path A for buoyancy-driven
ventilation. Compliance is
Natural ventilation systems shall be
based on preanalysis of
designed in accordance with this
generic spaces.
section and shall include mechanical
ventilation systems designed in
accordance with Section 6.2,
Section 6.3, or both.

Prescriptive multizone Section 6.4, “Natural Ventilation Iterative calculation method Buoyancy-driven, wind-
single cell buildings Procedure” of ASHRAE Standard 62.1- based on Section 4.3 of driven, buoyancy-driven +
2019 (ASHRAE 2019b): Path B CIBSE AM10 (2005) wind-driven ventilation
subpaths available

Engineered system ASHRAE Standard 62.1-2019 Performance-based Computer-based dynamic


performance multizone (ASHRAE 2019b) calculation to show heat transfer and bulk
The exception to Section 6.4, “Natural equivalency to prescriptive airflow modeling tools for
Ventilation Procedure,” states that “an minimum ventilation rate simulation of building
engineered natural ventilation system, requirements in physics related to natural
where approved by the authority Table 6.1, “Minimum ventilation airflow: refer to
having jurisdiction, need not meet the Ventilation Rates in Breathing Section 6.4.2,
requirements of Section 6.4” (ASHRAE Zone” “Engineered System
2019b). Compliance Path” for the
minimum required steps
and documentation.

Prescriptive single zone California’s Title 24, Part 6 4% total opening method Historical rules of thumb
(Energy Code) and California’s Title
See Website Appendix
24, Part 4 (Mechanical Code)
WA5.2, Sections WA5.2.1
and WA5.2.2.

Prescriptive single zone Military Handbook: Cooling Buildings Wind-driven method Wind-driven ventilation
by Natural Ventilation (DOD 1990) only
See Website Appendix
WA5.2, Sections WA5.2.1
and WA5.2.2.

5.2 WIND AND THERMAL BUOYANCY FORMULAS TO


DETERMINE AIRFLOW THROUGH AN OPENING
The two driving forces producing natural ventilation in a building are wind
pressure and thermal buoyancy (i.e., the stack effect). For wind-driven
ventilation for space with at least two openings, assuming inlet and outlet areas
are equal, use the simplified wind-driven ventilation equations Equations
212 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A2.2.1 or A2.2.2 from Appendix 2.2, “Equations for Wind-Driven


Ventilation.” For a space with only a single opening, apply Cv = 0.02 to these
same calculations to reflect the significantly reduced effectiveness of the
opening. For thermal buoyancy-driven ventilation, if there is no significant
internal resistance due to a partitioned interior, and assuming indoor and
outdoor temperatures are close to 80°F (26.7°C) and inlet and outlet openings
are equal, the flow due to stack effect is determined by Equations A2.1.1 and
A2.1.2 in Appendix 2.1, “Equations for Simple Buoyancy-Driven
Ventilation.”

5.3 PRESCRIPTIVE DESIGN REQUIREMENTS IN


ANSI/ASHRAE STANDARD 62.1-2016 AND
OLDER VERSIONS
ASHRAE Standard 62.1-2019 defines natural ventilation as
“ventilation provided by thermal, wind, or diffusion effects through doors,
windows, or other intentional openings in the building” (ASHRAE 2019b).
Outdoor air intakes (including openings that are required as part of a
natural ventilation system) shall be located such that the shortest distance
from the intake to any specific potential outdoor contaminant source shall
be equal to or greater than the separation distance listed in Table 5-1, “Air
Intake Minimum Separation Distance” of ANSI/ASHRAE Standard 62.1,
or the calculation method in the standard’s Normative Appendix B,
“Separation of Exhaust Outlets and Outdoor Air Intakes” (ASHRAE
2019b).
The prescriptive procedure is presented in Section 6.4, “Natural
Ventilation Procedure” (ASHRAE 2019b). Outdoor air is provided through
openings to the outdoors where the openable area is defined as the net-free
unobstructed area through an opening. Use of natural ventilation shall be
permitted for any zone or portion of a zone either solely or in conjunction
with mechanical ventilation systems per Section 6.4. Refer to Section 6.4
of ANSI/ASHRAE Standard 62.1 for additional information (ASHRAE
2019b).
A key element of the standard is that any zone designed for natural
ventilation shall be provided with a mechanical ventilation system unless
the locations and sizes of the openings of the prescriptive path remain
“permanently open or have controls to prevent them from being closed
during periods of expected occupancy” or if “the zone is not served by any
heating or cooling equipment” (ASHRAE 2019b). Any deviation from the
prescriptive path sizing requirements will void these exceptions.
The only other exception is to provide an engineered natural ventilation
system. See the following section for more information on this path.
The prescriptive path in older versions of ANSI/ASHRAE
Standard 62.1 required the open area to equal 4% of the floor area and to be
spread across certain geometries in a certain pattern. See ASHRAE
Standard 62.1 (2016a) for more details.
CHAPTER 5 | 213

5.4 ANSI/ASHRAE STANDARD 62.1 ON


PATH A AND PATH B OPENING SIZING
Updates made to ANSI/ASHRAE Standard 62.1-2019 (ASHRAE 2019b)
provide more prescriptive regulation on the sizing of openings, which
considers that the historical rule of thumb of 4% of floor area does not appear
to have recorded scientific basis.
Two paths for sizing the openings are discussed in Section 6.4, “Natural
Ventilation Procedure,” known as Path A and Path B, respectively (ASHRAE
2019b). These are limited to single-zone natural ventilation schemes only. Both
paths refer to the minimum ventilation flow rates required under Table 6-1,
“Minimum Ventilation Rates in Breathing Zone” as the criterion that must be
met through either path.
Path A uses a set of lookup tables and figures to represent preprocessed
analysis of typical room types. The designer enters the room’s required
minimum ventilation into the calculation methodology, and the calculation
returns a minimum opening size and height differential. This path sizes
openings on a buoyancy-driven flow only but allows that wind-driven flow is
possible so long as minimum flow rate can be guaranteed. Path A uses the term
dense to represent high-occupancy density, such as in a classroom, and not
dense to represent lower occupancy density, such as in an office building. See
Appendix A5.3 for additional information.
Path B essentially uses Section 4.3 of CIBSE AM10 (CIBSE 2005) to
perform single cell envelope flow model calculations to demonstrate how
opening sizes and locations meet the minimum flow rate from Table 6-1,
“Minimum Ventilation Rates in Breathing Zone” of ASHRAE Standard 62.1
(ASHRAE 2019b). See Website Appendix WA5.1, “CIBSE AM10 Excerpt on
Envelope Flow Models for Natural Ventilation,” accessible at ashrae.org/
naturalventilation, for further information on single-zone models that support
thermal buoyancy- and simple wind-driven natural ventilation. There is a
CIBSE natural ventilation tool is a Microsoft®-Excel®-based spreadsheet that
was developed as a supplement to CIBSE Applications Manual AM10 (CIBSE
2005). Though this spreadsheet calculates airflows for various scenarios
including buoyancy, wind and combined, it does so generically. The CIBSE
tool uses predefined formulas that disregard opening location, meteorological
data, and zone size and volume. Thus, it is appropriate for testing of feasibility
when the user can bring professional judgment to the input parameters and
interpretation of results.

5.5 PRESCRIPTIVE PATH ON OPENING COMPLIANCE


PER CALIFORNIA’S TITLE 24 CODE
California’s Title 24, Part 4 (CBSC 2019) and California’s Title 24, Part 6
(CEC 2018) govern natural ventilation. See Sections WA5.2.1 and WA5.2.2 of
Website Appendix WA5.2. Both provide a prescriptive method, but
California’s Title 24, Part 4 (CBSC 2019) is based on ANSI/ASHRAE
214 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Standard 62.1-2016 (ASHRAE 2016), and its 4% openings at 25 ft (7.6 m)


access distance rule.
The California Building Energy Efficiency Code (CEC 2018) requires that
mechanical heating and cooling systems have the heating and cooling shut off
when windows are opened. See Section WA5.2.3 in Website
Appendix WA5.2, “Excerpts on Natural Ventilation from California’s Title
24,” accessible at ashrae.org/naturalventilation. Mechanical ventilation must
continue to be provided even though natural ventilation may be able to provide
the minimum outdoor air quantities of 15 cfm/person or 0.15 cfm/ft2 (7.5 l/s or
0.76 l/s·m2). For more detailed information, refer to the California Energy
Code (CEC 2018). See Section WA5.2.3 in Website Appendix WA5.2,
“Excerpts on Natural Ventilation from California’s Title 24,” accessible at
ashrae.org/naturalventilation.
Lastly, the California Energy Code requires that energy performance
software and the user inputs provide backup direct expansion (DX)/ gas-fired
supplementary heating and cooling equipment whenever natural ventilation
systems result in unmet load hours of 150 hours or more in cooling or heating.
The alternate compliance manual lists the required performance of the
supplemental unit and requires that the software raised an exceptional
calculation method in use whenever natural ventilation is proposed in the
model. For more detailed information, refer to the California Energy Code
(CEC 2018). See Section WA5.2.4 in Website Appendix WA5.2, “Excerpts on
Natural Ventilation from California’s Title 24.”

5.6 ENGINEERED NATURAL VENTILATION SYSTEMS


The minimum ventilation requirements of ANSI/ASHRAE Standard 62.1
(ASHRAE 2019b) and California’s Title 24, Part 6 (CEC 2018) are determined
from their respective minimum ventilation air rates. The calculated airflow
derived through an engineered natural ventilation system approach is
compared to these minimum values. If the calculated airflow is lower than the
minimum compliance airflow, then the openings need to be adjusted until
minimum compliance values are achieved.
These dynamic path and bulk airflow models use zones to represent
volumes of air that are connected via flow resistances to form a network. Each
zone is assumed to be well-mixed, containing single-point evaluations of
temperature, pollutant concentrations, and hydrostatic pressure at a point in
space. Flow resistances represent airflow paths between zones within the same
occupied space, between adjacent spaces through openings, or between a space
and the outside air through openings. Airflow rates are calculated between
zones and through openings. These models are particularly suitable for annual
simulations and coupling with dynamic thermal models.
These models differ from analytical models because they solve a set of
simultaneous, nonlinear equations that ensure that the entire airflow network is
in balance (i.e., that the net mass flow into and out of any one zone is zero).
This requires the use of a computer program.
CHAPTER 5 | 215

The following steps can be taken to dynamically model natural ventilation


in iteration to size the openings for compliance:

1. Construct a scale model of the space in the simulation tool.


2. Locate the correct geographical locations of the space and the correct
weather file.
3. Build the correct sized openings together with the correct Cp value for the
openings in the model based on one of the prescriptive methods of sizing
openings.
4. Select either buoyancy-driven, wind-driven, or buoyancy-and-wind-driven
analysis.
5. Run at least a representative date in each season across a full diurnal cycle,
inclusive of summer and winter design conditions, as a representative view
of performance across the year.
6. Review results to determine whether the design achieves the minimum
compliance with the reference standard.
7. Use the results to adjust the designs, such as size and position of openings,
until compliance is achieved with adequate window opening sizes.
8. Provide results of the dynamic analysis which show temperatures across
the floor plane, space humidity levels across the floor plane, the number of
air changes in the space, and space surface temperatures.
9. The results of an engineered dynamic path analysis provide clear guidance
to the designer on how to prove a naturally ventilated system provides ade-
quate ventilation, space temperatures, and occupant comfort as well as a
surface condensation analysis.

For the analysis results found in Appendix A5.5, refer to the standard
model classroom and office parameters in Appendix A1.1 and Appendix A1.2,
respectively.
Figure 5.1 shows a single month’s output to allow discussion of
observations. In this case, it should be noted that the windows close each
evening and open in the morning upon occupancy. So, the zero flow periods
occur simultaneously with a lack of occupancy. The two horizontal lines are
superimposed on the results from the dynamic heat transfer and bulk airflow
model, which simulates the drivers of natural ventilation based on the loads in
the space and the outdoor air conditions.
Figure 5.1 shows that the flow is likely to meet Table 6-1 “Minimum
Ventilation Rates,” in ASHRAE Standard 62.1-2019 (ASHRAE 2019b) but
has a few days in which California’s Title 24, Part 6’s (CEC 2018) minimum
flow rates are not achieved. The codes would require that the minimum
ventilation be provided for all occupied hours, so this month would fail for this
set of opening sizes. Another iteration with a slightly large inlet opening would
need to be performed before submission to the AHJ under the engineered path
method.
216 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 5.1 Natural ventilation flow in CFM versus ANSI/ASHRAE Standard 62.1
(ASHRAE 2019) and California’s Title 24, Part 6 (CEC 2018) minimum ventilation rate
requirements for New York City, NY, in June for wind-driven ventilation.

More results for five cities with different climatic conditions can be found
in Website Appendix WA5.3, “Sample Results from an Engineered Natural
Ventilation System Analysis.”
Based on the minimum ventilation air requirements, the user can see
whether the minimum ventilation air is provided through natural ventilation to
the space. A primary assessment of natural ventilation air to a space can be
provided by a prescriptive model, or for better accuracy, a dynamic path and
bulk airflow movement tool.
If detailed information is required regarding air movement within the
space, then CFD analysis should be used. However, a CFD model requires
boundary conditions, and these are provided from bulk airflow or combined
dynamic heat transfer and bulk air models.

5.7 WIND-ONLY WINDOW SIZING PROCEDURE


The Florida Solar Energy Center developed a wind-driven ventilation
procedure that has been adopted by the U.S. Department of the Navy (DOD
1990). The handbook provides guidance and criteria for the design of buildings
to be totally or partially cooled by natural ventilation. It also describes several
natural criteria, design criteria for natural ventilation and for zoned or seasonal
CHAPTER 5 | 217

occupant and maintenance manuals, and guidelines for wind tunnel testing.
Appendices include forms and overlays for the designer’s use and describe the
fundamental principles of comfort related to airflow, a methodology for
climate analysis, prediction, and evaluation. Refer to Website
Appendix WA5.4, “Florida Solar Energy Center’s Wind-Only Window
Sizing,” accessible at ashrae.org/naturalventilation.

5.8 ESTIMATING APPROPRIATE SIZE FOR


NATURAL VENTILATION OPENINGS FOR
MINIMUM OUTDOOR AIR QUANTITIES ON COLD DAYS
This chapter has demonstrated how dynamic thermal analysis could be
used to size natural ventilation openings. At this point, it is useful for the
design team to use an iterative process with these tools to determine the
minimum amount of outdoor air that can reasonably be expected from
infiltration as a baseline, presuming that no other mechanical ventilation or
exhaust systems are operating on the floor plate. See Appendix A4.2 for a
means of estimating infiltration.
Once the lowest infiltration airflow is determined, it should be compared to
the nominal requirements under ANSI/ASHRAE Standard 62.1 (ASHRAE
2019b), as noted in Table 6-1, “Minimum Ventilation Rates in Breathing
Zone” to see if these rates are already accommodated. If not, then subsequent
iterations to increase window size can be used to determine the target range of
user operability on the natural ventilation openings during the times when
outdoor air conditions are at or below 50°F (10°C). This should then inform
the range of opening sizes needed and their location (high versus low). This
information should then be shared with the client to sign off on the minimum
outdoor air equivalency approach because this will form the basis of the
heating calculations. The owner may wish to agree to a safety factor to the
calculations to account for user preference for increased airflows.
The same minimum outdoor air opening range should also be shared with
the window/fenestration designer or vendor to explore the controllability of the
windows. Often, the ability to control the opening size is limited when the
window is near the closed position due to gaskets, preset positions, and
configurations of locking mechanisms.
If the natural ventilation or natural conditioning openings are determined to
be too large for manual controllability under low flow targets, it is often the
case in colder climates that “trickle ventilators” or window ventilators are
used. See Chapter 4 for more detail about these devices.

5.9 ENERGY MODELING FOR NATURAL VENTILATION IN


ANSI/ASHRAE/IES STANDARD 90.1-2019
When natural ventilation or natural conditioning is proposed as a solution,
it is often done as a means of energy savings. When a design team is trying to
show compliance with ventilation and comfort, it also needs to ensure that
218 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

compliance with energy code or energy performance criteria is also


documented. This section addresses the nuances of managing a natural
ventilation model through the process outlined in Appendix G of ANSI/
ASHRAE/IES Standard 90.1-2019 (ASHRAE 2019a), the governing standard
for several jurisdictions for demonstrating performance-based compliance with
an energy code as well as LEED® (USGBC 2013) compliance for Minimum
Energy Performance as noted below.
Standard 90.1 User’s Manual (ASHRAE 2017b) makes allowances for
natural ventilation under the performance rating system if the building does not
have a mechanical cooling system, but under the Energy Cost Budget Method,
both the proposed and design models are assumed to be air-conditioned. It
should be noted that LEED (USGBC 2013) only recognizes the performance
rating system, which has very specific requirements. See Website
Appendix WA5.5, “Energy Modeling for Natural Ventilation per ASHRAE
Standard 90.1-2019,” accessible at ashrae.org/naturalventilation, for more
details.
The California Energy Code (CEC 2018a) makes similar assumptions and
sets the parameters of the equipment type and operation of the supplemental
system. See Section WA5.2.4 in Website Appendix WA 5.2, “Excerpts on
Natural Ventilation from California’s Title 24,” accessible at ashrae.org/
naturalventilation, for more details.

5.10 NATURAL VENTILATION DOCUMENTATION FOR THE


LEED® RATING SYSTEM
The LEED rating system (USGBC 2013) will accept the results of the
ASHRAE Standard 90.1-2010 as the basis of calculation for compliance with
the Energy Performance prerequisite called Minimum Energy Performance
and the credit called Optimized Energy Performance. This is contingent on the
documentation showing full model compliance with ASHRAE Standard 90.1
Appendix G, “Performance Rating Method” with interpretation as allowed by
ASHRAE’s Standard 90.1 User’s Manual (ASHRAE 2017b). LEED
Reference Guide for Building Design and Construction (USGBC 2013)
identifies that a natural ventilation scheme itself does not require the
submission of an exceptional calculation method; however, a substantial
amount of backup documentation is required, as previously noted.
Note that the “Minimum Indoor Air Quality Performance” prerequisite
requires additional documentation on engineered natural ventilation systems,
depending on who the AHJ is. This is covered under LEED® Reference Guide
for Building Design and Construction’s (USGBC 2013) guidance on the AHJ
exception. Generally, the code enforced by the AHJ needs to be proven to be
equal to or more stringent than ASHRAE Standard 62.1-2010 (ASHRAE
2010) in effect under the credit definition, a description of the engineered
system must be provided to demonstrate that the intent of ASHRAE Standard
62.1-2010 (ASHRAE 2010) is being met, and a plan for approvals with the
AHJ (or evidence of approvals) must be provided. Note that the AHJ may also
CHAPTER 5 | 219

require parallel documentation to demonstrate how the project complies with


the most recent ventilation standard required in the jurisdiction.
LEED also requires that these naturally ventilated and mixed-mode spaces
have some form of confirming measurement under the minimum indoor air
quality performance prerequisite. The following is an excerpt from the LEED
Building Design and Construction Reference Guide (USGBC 2013):
One of the three must be provided:
1. Provide a direct exhaust airflow measurement device capable of
measuring the exhaust airflow. This device must measure the
exhaust airflow with an accuracy of ±10% of the design minimum
exhaust airflow rate. An alarm must indicate when airflow values
vary by 15% or more from the exhaust airflow set point.
2. Provide automatic indication devices on all-natural ventilation
openings intended to meet the minimum opening requirements. An
alarm must indicate when any one of the openings is closed during
occupied hours.
3. Monitor carbon dioxide (CO2) concentrations within each thermal
zone. CO2 monitors must be between 3 ft and 6 ft (900 mm and
1800 mm) above the floor and within the thermal zone. CO2 moni-
tors must have an audible or visual indicator or alert the building
automation system if the sensed CO2 concentration exceeds the set
point by more than 10%. Calculate appropriate CO2 set points using
the methods in Appendix C of ASHRAE Standard 62.1-2010.

Per the LEED® Reference Guide for Building Design and Construction
(USGBC 2013), for naturally ventilated spaces, the LEED® Enhanced Indoor
Air Quality Performance credit provides:

• One point for use of the appropriate strategies in Section 2.4 of CIBSE
Applications Manual AM10, Natural Ventilation in Non-Domestic
Buildings (CIBSE 2005). See Website Appendix WA5.6, “CIBSE
AM10’S Flowchart for LEED® Documentation for Natural Ventilation,”
accessible at ashrae.org/naturalventilation, for more detail.
• An additional point for any one of the following three options:
a. Design the project to reduce exterior contaminant pollutants by
using CFD or tracer gas analysis to prove that outdoor air intakes
experience lower than threshold concentrations on all of the
contaminants governed by the National Ambient Air Quality
Standards (NAAQS)
b. Provide an analysis of special indoor contaminants, produce a
material handling plan for containment and a sensor/alarming
system to monitor indoor air quality for that pollutant
c. Provide CIBSE AM10 (CIBSE 2005), Section 4, “Design
Calculations,” to predict that room-by-room airflows will provide
effective natural ventilation, with submission of data for every
220 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

room. See California’s Title 24 (CEC 2018) and CIBSE AM10


(CIBSE 2005) for additional information.

Mixed-mode spaces can achieve two points under the “Enhanced Indoor
Air Quality Performance” credit by (USGBC 2013):

• One point for use of the appropriate strategies in CIBSE Applications


Manual 13-2000, Mixed Mode Ventilation (CIBSE 2000), Section 2.1
and Figure 2.1 to perform contingency design, complementary design,
and zone design review in iteration to optimize the strategy and provide
documentation of compliance with intent (USGBC 2013). See Website
Appendix WA5.7, “CIBSE AM13’s Strategy Selection for Mixed-
Mode Ventilation,” accessible at ashrae.org/naturalventilation, for an
excerpt of this strategy selection approach and figure for the reader’s
reference.
• An additional point for any one of the following four options (USGBC
2013):
a. Design the project to reduce exterior contaminant pollutants by
using CFD or tracer gas analysis to prove that outdoor air intakes
experience lower than threshold concentrations on all of the
contaminants governed by the National Ambient Air Quality
Standards (NAAQS).
b. Provide an analysis of special indoor contaminants, produce a
material handling plan for containment and a sensor/alarming
system to monitor indoor air quality for that pollutant.
c. Comply with CIBSE AM10 (CIBSE 2005), Section 4, “Design
Calculations,” to predict that room-by-room airflows will provide
effective natural ventilation, with submission of data for every
room. See Website Appendix WA5.1, “CIBSE AM10 Excerpt on
Envelope Flow Models for Natural Ventilation,” accessible at
ashrae.org/naturalventilation, for more information.
d. Provide increased ventilation at 30% above minimum requirements
within the prerequisite documentation for the mechanical
ventilation system.

5.11 REFERENCES
ASHRAE. 2010. ANSI/ASHRAE/IES Standard 90.1-2010, Energy standard
for buildings except low-rise residential buildings. Peachtree Corners, GA:
ASHRAE.
ASHRAE. 2016. ANSI/ASHRAE Standard 62.1-2016, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2017a. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2017b. Standard 90.1 user’s manual. Peachtree Corners, GA:
ASHRAE.
CHAPTER 5 | 221

ASHRAE. 2019a. ANSI/ASHRAE/IES Standard 90.1-2019, Energy standard


for buildings except low-rise residential buildings. Peachtree Corners, GA:
ASHRAE.
ASHRAE. 2019b. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
CBSC. 2019. 2019 California Mechanical Code: California Code of
Regulations, Title 24, Part 4. Based on 2018 Uniform Mechanical Code®.
Ontario, CA: International Association of Plumbing and Mechanical
Officials.
CEC. 2018. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
CIBSE. 2000. CIBSE applications manual AM13: Mixed mode ventilation.
London: The Chartered Institute of Building Services Engineers.
CIBSE. 2005. CIBSE applications manual AM10: Natural ventilation in non-
domestic buildings. London: The Chartered Institute of Building Services
Engineers.
DOD. 1990. Military handbook 1011/2: Cooling buildings by natural
ventilation. MIL-HDBK-1011/2. Alexandria, VA: U.S. Department of
Defense’s Department of the Navy.
USGBC. 2013. LEED® reference guide for building design and construction,
LEED version v4. Washington, DC: U.S. Green Building Council.
FOUNDATIONAL CALCULATIONS FOR
ENVELOPE FLOW MODELS
This appendix presents the foundational calculations for natural airflow
through an opening as documented in CIBSE Applications Manual AM10:
Natural Ventilation in Non-Domestic Buildings (CIBSE 2005). The following
excerpt is replicated verbatim with permission from CIBSE. Note: the original
section and equation numbering from the reference document are retained.

4.2.1 Envelope Flow Models: Single-Cell


Envelope flow models solve the equations that govern the flow of air
through openings in the envelope of a building. They rely on assumptions
about the internal density (temperature) distribution.

4.2.1.1 Basic Equations for an Opening


The basic equations for a single- cell can be illustrated by considering an
opening in the envelope. It is common practice to express the relationship
between the flow rate through an opening and a pressure difference across it by
means of the discharge coefficient and a specified geometric area:

2 p
qi = C di A i --------------i (4.1)

where i identifies the opening, q is the flow rate through the opening (m3 s–1),
Cdi is the discharge coefficient (—), Ai is the area of the opening (m2), pi is
the pressure difference (Pa) and  is the air density (kg·m–3).
The discharge coefficient is defined and measured under still-air conditions
with uniform density with the flow being generated by a fan. When the flow is
generated by a density difference (i.e. in the absence of any wind effects), it
has been shown that the pressure drops in equation 4.1 is given by:

pi = P E0 – P I0 –  0 gz i (4.2)

223
224 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

where PE0 and PI0 are the external and internal hydrostatic pressures
respectively at ground level (Pa), 0 is the density difference at ground level
(kg·m–3), g is the gravitational force per unit mass (m·s–2), and zi is the height
of opening i above ground level (m).
The density difference is defined by:

 0 =  E –  I (4.3)

where E and I are the densities of the external and internal air respectively
(kg·m–3).
Equation 4.2 is quite general. It applies to an opening aligned in any
direction, irrespective of whether the flow is inward or outward, and whether
the inside temperature is higher or lower than the outside temperature.
However, it is important to note that the height of the opening, zi, is the height
where the flow leaves the opening, i.e. the height of the outlet. This follows
from the outlet boundary condition, namely that the pressure at the outlet is
determined by the pressure of the surroundings, and the assumption that the
temperature of the air remains unchanged as it flows through the opening. This
means that for a long opening in the vertical direction, such as a chimney, zi
will change with the flow direction. For most openings, such as windows and
air vents, zi will not change with flow direction.
If the density of the internal air varies with height equation 4.2 becomes:
zi
 
p i = P E0 – P I0 –  E gz i + g   I dz
 (4.4)
 
 0 
Equation 4.4 allows non-uniform density distributions (as can occur with
atria and chimneys, for example) to be dealt with.

4.2.1.2 Effect of Wind


When the wind blows the external pressure around the opening is changed
by the pressure due the wind, pwi, which adds to the hydrostatic pressure and
the pressure drop equation becomes:

p i = P E0 – P I0 –  0 gz i + P wi (4.5)

where pwi is the pressure of the wind at the point on the envelope at which
opening i lies (Pa).
It is common practice to obtain pwi from wind tunnel tests where wind
pressures are quoted in the form of the pressure coefficient:
P wi – P ref
C pi = -----------------------
2
(4.6)
0.5U
where pref is a reference pressure (Pa) and U is the wind speed (m·s–1).
APPENDIX A5.1 | 225

Thus, the pressure difference across an opening whose inlet or outlet is


situated in the external flow can be written as:

2
p i = P E0 – P I0 – p ref –  0 gz i + 0.5U C pi (4.7)

The treatment of wind effect involves several assumptions. For example,


the size of the opening is assumed to be small enough for the value of pwi
obtained from a point measurement on a wind tunnel model to be appropriate.
This is reasonable for small openings such as air vents but may be less so
for open windows. An associated assumption is that the external wind flow
around the opening does not affect the discharge coefficient of the opening.
This is more reasonable when the outlet of the opening is exposed to the wind
(i.e. outward flow from the space) than when the inlet is exposed. None of
these assumptions is needed for the treatment of ventilation due to density

4.2.1.3 Conservation of Mass for the Envelope


The principle of mass conservation applied to the fixed volume defined by
the envelope gives a relationship between the flow rates of the openings:

 i q i = 0 (4.8)

i.e. the net mass flow into the building is equal to zero. In the context of design
procedures, it is an acceptable approximation to ignore the differences between
densities, so the equation can be simplified to qi = 0.

4.2.1.4 Solution of the Equations


An envelope flow model thus consists of the following three basic
equations. For the envelope:

q i = 0 (4.9)

and for each opening:

2 p
q i = C di A i S i --------------i (4.10)
0

and:

2
p i = p 0 – p 0 gz i + 0.5 0 U C pi (4.11)

Note: in equation 4.10, it is important to include the sign of the pressure


difference, which has been introduced into equation 4.10 by the symbol ‘S’;
226 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

the convention used here is that S = + 1 for flow entering the space and S = –1
for flow leaving the space.
Equation 4.11 is the same as equation 4.7 except that (PE0 – PI0 –Pref )
has been denoted by p0 for brevity. Although the bracket encloses three
terms, it is only the internal pressure PI0 that is a variable; PE0 and pref are
constant for a given analysis.
For the general case of N openings there are 2 N + 1 equations. The
unknowns are qi, Pi and Po. The equations can be solved by determining the
value of Po at which the continuity equation is satisfied. In physical terms,
when the ventilation pattern of a building is changed from one steady state to
another (e.g. by opening a window), the internal pressure adjusts until the
flows through the openings are such that the continuity equation is again
satisfied. Mathematically this adjustment is done by an iterative procedure.
When the equations are solved in this way it is referred to as an implicit
method.
The equations can be solved directly (i.e. without the need for iteration) by
specifying the value of Po and the values of qi (magnitude and direction) to
find the values of Ai which will give that particular flow pattern. When solved
in this way, the model is referred to as an explicit method. This method is
particularly useful in the initial design stages for sizing openings, such that the
openings give the required flow rates under a specified design condition.
Worked examples are given in section 4.3. Using these areas, the implicit
procedure can then be used for off- design calculation of flow rates.
The explicit procedure is easy to use for openings with a constant discharge
coefficient, i.e. one that does not vary with flow rate (Reynolds number). Most
purpose-provided openings fall into this category and manual calculations are
not difficult, as illustrated by the worked examples in section 4.3. The explicit
procedure is not appropriate for adventitious openings (see section 4.4.4)
because such openings cannot be sized in the same way as air vents and
windows.

A5.1.1 REFERENCE
CIBSE. 2005. CIBSE applications manual AM10: Natural ventilation in non-
domestic buildings. London: The Chartered Institute of Building Services
Engineers.
ASHRAE STANDARD 62.1-2016
EXCERPT ON
NATURAL VENTILATION
This appendix presents an excerpt of portions of ASHRAE Standard 62.1-
2016 (ASHRAE 2016) that apply to natural ventilation, reproduced here
verbatim with ASHRAE’s permission. Note: the original section numbering
from the reference document is retained.
6.4 Natural Ventilation Procedure. Natural ventilation systems shall be
designed in accordance with this section and shall include mechanical ventila-
tion systems designed in accordance with Section 6.2, Section 6.3, or both.
Exceptions:
1. An engineered natural ventilation system, where approved by the author-
ity having jurisdiction, need not meet the requirements of Section 6.4.
2. The mechanical ventilation systems shall not be required where
a. natural ventilation openings that comply with the requirements of
Section 6.4 are permanently open or have controls that prevent the
openings from being closed during periods of expected occupancy or
b. the zone is not served by heating or cooling equipment.
6.4.1 Floor Area to Be Ventilated. Spaces, or portions of spaces, to be natu-
rally ventilated shall be located within a distance based on the ceiling height, as
determined by Sections 6.4.1.1, 6.4.1.2, or 6.4.1.3, from operable wall open-
ings that meet the requirements of Section 6.4.2. For spaces with ceilings that
are not parallel to the floor, the ceiling height shall be determined in accor-
dance with Section 6.4.1.4.
6.4.1.1 Single Side Opening. For spaces with operable openings on one
side of the space, the maximum distance from the operable openings shall be
not more than 2H, where H is the ceiling height.
6.4.1.2 Double Side Opening. For spaces with operable openings on two
opposite sides of the space, the maximum distance from the operable openings
shall be not more than 5H, where H is the ceiling height.
6.4.1.3 Corner Openings. For spaces with operable openings on two adja-
cent sides of a space, the maximum distance from the operable openings shall

227
228 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

be not more than 5H along a line drawn between the two openings that are far-
thest apart. Floor area outside that line shall comply with Section 6.4.1.1.
6.4.1.4 Ceiling Height. The ceiling height (H) to be used in Sections
6.4.1.1 through 6.4.1.3 shall be the minimum ceiling height in the space.
Exception: For ceilings that are increasing in height as distance from the
openings is increased, the ceiling height shall be determined as the aver-
age height of the ceiling within 6 m (20 ft) from the operable openings.
6.4.2 Location and Size of Openings. Spaces or portions of spaces to be
naturally ventilated shall be permanently open to operable wall openings
directly to the outdoors. The openable area shall be not less than 4% of the net
occupiable floor area. Where openings are covered with louvers or otherwise
obstructed, openable area shall be based on the net free unobstructed area
through the opening. Where interior rooms, or portions of rooms, without
direct openings to the outdoors are ventilated through adjoining rooms, the
opening between rooms shall be permanently unobstructed and have a free area
of not less than 8% of the area of the interior room or less than 25 ft2 (2.3 m2).
6.4.3 Control and Accessibility. The means to open required operable open-
ings shall be readily accessible to building occupants whenever the space is
occupied. Controls shall be designed to coordinate operation of the natural and
mechanical ventilation systems.

A5.2.1 REFERENCE
ASHRAE. 2016. ANSI/ASHRAE Standard 62.1-2016, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
EXCERPTS FROM
ASHRAE STANDARD 62.1-2019
ON NATURAL VENTILATION
This appendix presents an excerpt of portions of ASHRAE Standard 62.1-
2019 (ASHRAE 2019) that apply to natural ventilation, reproduced here
nearly verbatim with ASHRAE’s permission. Note: the original section
numbering from the reference document is retained.

3. DEFINITIONS
3.1 Terminology

air:

exhaust air: air removed from a space and discharged to outside the build-
ing by means of mechanical or natural ventilation systems.
outdoor air: ambient air and ambient air that enters a building through a
ventilation system, through intentional openings for natural ventilation, or
by infiltration.
supply air: air delivered by mechanical or natural ventilation to a space
and composed of any combination of outdoor air, recirculated air, or trans-
fer air.
natural ventilation: ventilation provided by thermal, wind, or diffusion effects
through doors, windows, or other intentional openings in the building.

5.5.1 Location. Outdoor air intakes (including openings that are required as
part of a natural ventilation system) shall be located such that the shortest dis-
tance from the intake to any specific potential outdoor contaminant source
listed in Table 5-1 shall be equal to or greater than
a. the separation distance in Table 5-1 or
b. the calculation methods in Normative Appendix B
and shall comply with all other requirements of this section.

229
230 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

5.19.5 Recirculation. Air-handling and natural ventilation systems shall not


recirculate or transfer air from an ETS area to an ETS-free area.

6. PROCEDURES
6.1 General. The Ventilation Rate Procedure, the IAQ Procedure, the Natural
Ventilation Procedure, or a combination thereof shall be used to meet the
requirements of this section. In addition, the requirements for exhaust ventila-
tion in Section 6.5 shall be met regardless of the method used to determine
minimum outdoor airflow rates.
6.1.1 Natural Ventilation Procedure. The prescriptive or engineered sys-
tem design procedure presented in Section 6.4, in which outdoor air is provided
through openings to the outdoors, shall be permitted to be used for any zone or
portion of a zone in conjunction with mechanical ventilation systems in accor-
dance with Section 6.4.
6.2 Natural Ventilation Procedure. Natural ventilation systems shall comply
with the requirements of either Section 6.4.1 or 6.4.2. Designers shall provide
interior air barriers, insulation, or other means that separate naturally ventilated
spaces from mechanically cooled spaces to prevent high-dew-point outdoor air
from coming into contact with mechanically cooled surfaces.
6.2.1 Prescriptive Compliance Path. Any zone designed for natural ventila-
tion shall include a mechanical ventilation system designed in accordance with
Section 6.2, Section 6.3, or both.
Exceptions to 6.4.1:
1. Zones in buildings that have all of the following:
a. Natural ventilation openings that comply with the requirements of
Section 6.4.1.
b. Controls that prevent the natural ventilation openings from being closed
during periods of expected occupancy, or natural ventilation openings
that are permanently open.
2. Zones that are not served by heating or cooling equipment.
6.2.1.1 Ceiling Height. For ceilings that are parallel to the floor, the ceiling
height (H) to be used in Sections 6.4.1.3 through 6.4.1.5 shall be the minimum
ceiling height in the zone.
For zones wherein ceiling height increases as distance from the ventilation
increases, the ceiling height shall be the average height of the ceiling deter-
mined over a distance not greater than 6 m (20 ft) from the openings.
6.2.1.2 Floor Area to be Ventilated. The naturally ventilated area in zones
or portions of zones shall extend from the openings to a distance determined by
Sections 6.4.1.3, 6.4.1.4, or 6.4.1.5. Openings shall meet the requirements of
Section 6.2.1.6. For zones where ceilings are not parallel to the floor, the ceiling
height shall be determined in accordance with Section 6.4.1.1.
6.2.1.3 Single Side Opening. For zones with openings on only one side of
the zone, the naturally ventilated area shall extend to a distance not greater that
two times the height of the ceiling from the openings.
APPENDIX A5.3 | 231

6.2.1.4 Double Side Opening. For zones with openings on two opposite
sides of the zone, the naturally ventilated area shall extend between the open-
ings separated by a distance not greater than five times the height of the ceiling.
6.2.1.5 Corner Openings. For zones with openings on two adjacent sides
of a zone, the naturally ventilated area shall extend to a distance not greater
than five times the height of the ceiling along a line drawn between the outside
edges of the two openings that are the farthest apart. Floor area outside that line
shall comply with Section 6.2.1.3 as a zone having openings on only one side
of the zone.
Informative Note: Floor area outside that line refers to the remaining area
of the zone that is not bounded by the walls that have the openings and the line
drawn between the openings.
6.2.1.6 Location and Size of Openings. Zones or portions of zones to be
naturally ventilated shall have a permanently open airflow path to openings
directly connected to the outdoors. The minimum flow rate to the zone shall be
determined in accordance with Section 6.2.1.1. This flow rate shall be used to
determine the required openable area of openings, accounting only for buoy-
ancy-driven flow. Wind-driven flow shall be used only where it can be demon-
strated that the minimum flow rate is provided during all occupied hours.
Openings shall be sized in accordance with Section 6.4.1.6.1 (Path A) or Sec-
tion 6.4.1.6.2 (Path B).
Informative Note: Permanently open airflow path refers to pathways that
would allow airflow unimpeded by partitions, walls, furnishings, etc.

Table 6-5 Minimum Openable Areas: Single Openings a

Total Openable Areas in Zone as a Percentage of Az


Vbz/Az  Vbz/Az 
([L/s]/m2) (cfm/ft2) HS/WS  0.1 0.1 < HS/WS  1 HS/WS > 1

1.0 0.2 4.0 2.9 2.5

2.0 0.4 6.9 5.0 4.4

3.0 0.6 9.5 6.9 6.0

4.0 0.8 12.0 8.7 7.6

5.5 1.1 15.5 11.2 9.8

where
Vbz = breathing zone outdoor airflow, per Table 6-1.
Az = zone floor area, the net occupiable floor area of the ventilation zone.
WS = aggregated width of all single outdoor openings located at the same elevation.
HS = vertical dimension of the single opening or the least vertical dimension of the openings where there are multiple openings.

a. Volumetric airflow rates used to estimate required openable area are based on the following:
• Dry-air density of 0.075 lbda/ft3 (1.2 kgda/m3) at a barometric pressure of 1 atm (101.3 kPa) and an air temperature of 70°F (21°C)
• Temperature difference between indoors and outdoors of 1.8°F (1°C)
• Gravity constant of 32.2 ft/s2 (9.81 m/s2)
• Window discharge coefficient of 0.6
232 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Table 6-6 Minimum Openable Areas: Two Vertically Spaced Openings a

Total Openable Areas in Zone as a Percentage of Az

Hvs  8.2 ft (2.5 m) 8.2 ft (2.5m) < Hvs  16.4 ft (5 m) 16.4 ft (5 m) < Hvs
Vbz/Az  Vbz/Az 
(L/s/m2) (cfm/ft2) As/Al  0.5 As/Al > 0.5 As/Al  0.5 As/Al > 0.5 As/Al  0.5 As/Al > 0.5

1.0 0.2 2.0 1.3 1.3 0.8 0.9 0.6

2.0 0.4 4.0 2.6 2.5 1.6 1.8 1.2

3.0 0.6 6.0 3.9 3.8 2.5 2.7 1.7

4.0 0.8 8.0 5.2 5.0 3.3 3.6 2.3

5.5 1.1 11.0 7.1 6.9 4.5 4.9 3.2

where
Vbz = breathing zone outdoor airflow, per Table 6-1.
Az = zone floor area, the net occupiable floor area of the ventilation zone.
Hvs = vertical separation between the center of the top and bottom openings’ free operable area; in case of multiple horizontally spaced pairs of openings, use shortest
distance encountered.
As = openable area of smallest opening (top or bottom); in case of multiple horizontally spaced pairs of top-and-bottom openings, use aggregated areas.
Al = openable area of largest opening (top or bottom); in case of multiple horizontally spaced pairs of top-and-bottom openings, use aggregated areas.

a. Volumetric airflow rates used to estimate required operable area are based on the following:
• Dry-air density of 0.075 lbda/ft3 (1.2 kgda/m3) at a barometric pressure of 1 atm (101.3 kPa) and an air temperature of 70°F (21°C)
• Temperature difference between indoors and outdoors of 1.8°F (1°C)
• Gravity constant of 32.2 ft/s2 (9.81 m/s2)
• Window discharge coefficient of 0.6

6.4.1.6.1 Sizing Openings—Path A. Where the zone is ventilated using


a single opening or multiple single openings located at the same elevation, the
openable area as a percent of the net occupiable floor area shall be greater
than or equal to the value indicated in Table 6-5. Where the zone is ventilated
using two openings located at different elevations or multiple pairs of such
openings, the openable area as a percent of the net occupiable floor area shall
be greater than or equal to the value indicated in Table 6-6.
Where openings are obstructed by louvers or screens, the openable area
shall be based on the net free area of the opening. Where interior zones, or por-
tions of zones, without direct openings to the outdoors are ventilated through
adjoining zones, the opening between zones shall be permanently unobstructed
and have a free area of not less than twice the percent of occupiable floor area
used to determine the opening size of adjacent exterior zones, or 25 ft2 (2.3
m2), whichever is greater.
Informative Note: Tables 6-5 and 6-6 are based solely on buoyancy-driven
flow and have not been created to address thermal comfort.
6.4.1.6.2 Sizing Openings—Path B. The required openable area for a
single zone shall be calculated using CIBSE AM10, Section 4.3.
6.4.2 Engineered System Compliance Path. For an engineered natural ven-
tilation system, the designer shall
a. determine hourly environmental conditions, including outdoor air dry-bulb
temperature; dew-point temperature; outdoor concentration of contami-
nants, including PM2.5, PM10, and ozone where data are available; wind
APPENDIX A5.3 | 233

speed and direction; and internal heat gains during expected hours of natural
ventilation operation.
b. determine the effect of pressure losses along natural ventilation airflow
paths on the resulting flow rates, including inlet openings, air transfer grills,
ventilation stacks, and outlet openings during representative conditions of
expected natural ventilation system use.
c. quantify natural ventilation airflow rates of identified airflow paths account-
ing for wind induced and thermally induced driving pressures during repre-
sentative conditions of expected natural ventilation system use.
d. design to provide outdoor air in quantities sufficient to result in acceptable
IAQ as established under Section 6.2.1.1 or 6.3 during representative condi-
tions of expected natural ventilation system use.
6.4.3 Control and Accessibility. The means to open required openings shall
be readily accessible to building occupants whenever the space is occupied.
Controls shall be designed to coordinate operation of the natural and mechani-
cal ventilation systems.
6.4.4 Documentation. Where the Natural Ventilation Procedure is used, the
designer shall document the values and calculations that demonstrate confor-
mance with the compliance path and the controls systems and sequences
required for operation of the natural ventilation system, including coordination
with mechanical ventilation systems. Where the Prescriptive Compliance Path
is used for buildings located in an area where the national standard for one or
more contaminants is exceeded, any design assumptions and calculations
related to the impact on IAQ shall be included in the design documents.
234 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

8. OPERATIONS AND MAINTENANCE


8.1 General

Table 8-1 Minimum Maintenance Activity and Frequency for Ventilation System Equipment
and Associated Components

Inspection/Maintenance Task Frequency a


a. Investigate system for water intrusion or accumulation. Rectify as necessary. As necessary
b. Verify that the space provided for routine maintenance and inspection of open cooling tower water systems, closed Monthly
cooling tower water systems, and evaporative condensers is unobstructed.
c. Open cooling tower water systems, closed cooling tower water systems, and evaporative condensers shall be treated Monthly
to limit the growth of microbiological contaminants, including legionella sp.
d. Verify that the space provided for routine maintenance and inspection of equipment and components is unobstructed. Quarterly
e. Check pressure drop and scheduled replacement date of filters and air-cleaning devices. Clean or replace as Quarterly
necessary to ensure proper operation.
f. Check ultraviolet lamp. Clean or replace as needed to ensure proper operation. Quarterly
g. Visually inspect dehumidification and humidification devices. Clean and maintain to limit fouling and microbial Quarterly
growth. Measure relative humidity and adjust system controls as necessary.
h. Maintain floor drains and trap primer located in air plenums or rooms that serve as air plenums to prevent transport Semiannually
of contaminants from the floor drain to the plenum.
i. Check ventilation and IAQ related control systems and devices for proper operation. Clean, lubricate, repair, adjust, Semiannually
or replace as needed to ensure proper operation.
j. Check P-traps in floor drains located in plenums or rooms that serve as air plenums. Prime as needed to ensure Semiannually
proper operation.
k. Check fan belt tension. Check for belt wear and replace if necessary to ensure proper operation. Check sheaves for Semiannually
evidence of improper alignment or evidence of wear and correct as needed.
l. Check variable-frequency drive for proper operation. Correct as needed. Semiannually
m. Check for proper operation of cooling or heating coil for damage or evidence of leaks. Clean, restore, or replace as Semiannually
required.
n. Visually inspect outdoor air intake louvers, bird screens, mist eliminators, and adjacent areas for cleanliness and Semiannually
integrity; clean as needed; remove all visible debris or visible biological material observed and repair physical
damage to louvers, screens, or mist eliminators if such damage impairs the item from providing the required outdoor
air entry.
o. Visually inspect natural ventilation openings and adjacent areas for cleanliness and integrity; clean as needed. Semiannually
Remove all visible debris or visible biological material observed and repair physical damage to louvers, and screens
if such damage impairs the item from providing the required outdoor air entry. Manual and/or automatic opening
apparatus shall be physically tested for proper operation and repaired or replaced as necessary.
p. Verify the operation of the outdoor air ventilation system and any dynamic minimum outdoor air controls. Annually
q. Check air filter fit and housing seal integrity. Correct as needed. Annually
r. Check control box for dirt, debris, and/or loose terminations. Clean and tighten as needed. Annually
s. Check motor contactor for pitting or other signs of damage. Repair or replace as needed. Annually
t. Check fan blades and fan housing. Clean, repair, or replace as needed to ensure proper operation. Annually
u. Check integrity of all panels on equipment. Replace fasteners as needed to ensure proper integrity and fit/finish of Annually
equipment.
v. Assess field serviceable bearings. Lubricate if necessary. Annually
a. Minimum frequencies may be increased or decreased if indicated in the O&M manual.
APPENDIX A5.3 | 235

Table 8-1 Minimum Maintenance Activity and Frequency for Ventilation System Equipment
and Associated Components (Continued)

Inspection/Maintenance Task Frequency a


w. Check drain pans, drain lines, and coils for biological growth. Check adjacent areas for evidence of unintended Annually
wetting. Repair and clean as needed.
x. Check for evidence of buildup or fouling on heat exchange surfaces. Restore as needed to ensure proper operation. Annually
y. Inspect unit for evidence of moisture carryover from cooling coils beyond the drain pan. Make corrections or repairs Annually
as necessary.
z. Check for proper damper operation. Clean, lubricate, repair, replace, or adjust as needed to ensure proper operation. Annually
aa. Visually inspect areas of moisture accumulation for biological growth. If present, clean or disinfect as needed. Annually
a. Minimum frequencies may be increased or decreased if indicated in the O&M manual.

8.2 Ventilation System Operation. Mechanical and natural ventilation sys-


tems shall be operated in a manner consistent with the O&M manual. Systems
shall be operated such that spaces are ventilated in accordance with Section 6
during periods of expected occupancy.
236 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

(This is a normative appendix and is part of the standard.)

NORMATIVE APPENDIX B
SEPARATION OF EXHAUST OUTLETS AND OUTDOOR AIR
INTAKES

B1. GENERAL
This appendix presents an alternative procedure for determining separation dis-
tance between outdoor air intakes and exhaust air and vent outlets. This analyt-
ical method can be used instead of Table 5-1.
Exhaust air and vent outlets, as defined in Table 5-1, shall be located no closer
to outdoor air intakes, or to operable windows, skylights, and doors, both those on
the subject property and those on adjacent properties, than the minimum separa-
tion distance (L) specified in this section. The distance (L) is defined as the short-
est “stretched string” distance measured from the closest point of the outlet
opening to the closest point of the outdoor air intake opening, or to the operable
window, skylight, or door opening, along a trajectory as if a string were
stretched between them.
B1.1 Application. Laboratory fume hood exhaust air outlets shall be in com-
pliance with NFPA 45 and ANSI/AIHA Z9.5. Nonlaboratory exhaust outlets
and outdoor air intakes or other openings shall be separated in accordance with
the following.
B1.2 Outdoor Air Intakes. The minimum separation distance between
exhaust air/vent outlets, as defined in Table 5-1, and outdoor air intakes to
mechanical ventilation systems, or to operable windows, skylights, and doors
that are required as part of natural ventilation systems, shall be equal to dis-
tance (L) determined in accordance with Section B2.
Separation distances do not apply when exhaust and outdoor air intake systems
are controlled such that they cannot operate simultaneously.
APPENDIX A5.3 | 237

(This appendix is not part of this standard. It is merely informative and


does not contain requirements necessary for conformance to the stan-
dard. It has not been processed according to the ANSI requirements for a
standard and may contain material that has not been subject to public
review or a consensus process. Unresolved objectors on informative
material are not offered the right to appeal at ASHRAE or ANSI.)

INFORMATIVE APPENDIX J
INFORMATION ON NATURAL VENTILATION

J1. OUTDOOR AIR QUALITY DATA


Outdoor air quality data may be considered valid if it is demonstrated that the
data are both physically representative and spatially representative.
Physically representative data accurately reflect the air quality conditions
at the monitoring station from which they are derived. Data are considered
physically representative if they are obtained from
a. reports of historical levels of air pollutants published by the relevant local,
regional, or federal entity with statutory responsibility for collecting and
reporting air quality information in accordance with applicable air quality
regulations, or
b. an on-site monitoring campaign that is verifiably comparable to local,
regional, or federal guidelines and methods for demonstration of compli-
ance with applicable air quality regulations.
Spatially representative data are collected from a monitoring site that may
differ from the proposed project location but is informative of the air quality
conditions at the proposed project location. Data may be considered spatially
representative if they are
a. the same as those used by the entity charged with demonstrating regulatory
compliance for the geographic region that includes the proposed project
location, or
b. derived from an on-site monitoring campaign that also meets the require-
ment stated by criteria (b) of this annotation.

J2. NATURAL VENTILATION RATE


When calculating the ventilation rate, specific path(s) of the intended airflow
passage must first be determined along with flow directions. There are two
driving forces for natural ventilation: buoyancy and wind. The two driving
forces can work cooperatively or competitively based on the environmental
conditions of wind speed, direction, indoor/outdoor air/surface temperatures,
as well as the intentional airflow path and mechanisms.
a. In the case of an engineered natural ventilation system that results in multi-
ple flow scenarios, each must be examined and considered separately.
238 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Table J-1 Ventilation Intensity Brackets

Bracket (L/s)·m2 cfm/ft2 Commonly Encountered Space Typologies Bracket


1 0.0 to 1.0 0.0 to 0.2 Office, living room, main entry lobby
2 1.0 to 2.0 0.2 to 0.4 Reception area, general manufacturing, kitchen, lobby
3 2.0 to 3.0 0.4 to 0.6 Classroom, daycare
4 3.0 to 4.0 0.6 to 0.8 Restaurant dining room, places of religious worship
5 4.0 to 5.5 0.6 to 1.1 Auditorium, health club/aerobics room, bar, gambling
Not addressed: Lecture Hall and spectator areas (6 [L/s])/m2) and disco/dance floors (10.3 [L/s]/m2)

b. Specific pressure-based calculation of natural ventilation flow rate is docu-


mented in ASHRAE Handbook—Fundamentals, Chapter 16, Section 6:
1. Buoyancy-induced airflow can be calculated following Equation 38.
2. Wind-driven airflow can be calculated following Equation 37.
3. The overall pressure (driven by both wind and stack effect) converted to
resulting pressure difference between openings can found in Equation
36.
For obtaining wind-driven pressure, several methods are available:
a. ASHRAE Handbook—Fundamentals, Chapter 24, provides a method to
convert wind speed and direction into pressure coefficients that can be used
to determine wind-driven pressure.
b. CIBSE AM10, Chapter 4, provides a method to account for wind-driven
ventilation and outlines specific challenges to it in Section 4.4.1.
c. If the building has undergone wind tunnel test for structural stress, the same
test can provide detailed pressure coefficients.
d. Outdoor airflow simulation (such as computational-fluid-dynamics-based
simulation) can be used to obtain the specific flow condition at the intended
openings.
For intended openings that are large, such as open atrium or open balcony,
and/or when the flow path is not well defined, such as when only single or sin-
gle-side openings are available, the pressure-based method can be invalid, and
outdoor-indoor linked simulation should be used.

J3. PRESCRIPTIVE PATH A CALCULATIONS


J3.1 Ventilation Intensity. Spaces have been defined by a ventilation inten-
sity, which represents the amount of flow rate needed per Equation 6-1, divided
by the floor area of the space. Its units are (L/s)/m2 of floor area or cfm/ft2 of
floor area.

V bz Rp  Pz + Ra  Az
Ventilation Intensity = -------- = ------------------------------------------ (J-1)
Az Az

The ventilation intensity brackets in Table J-1 are used.


APPENDIX A5.3 | 239

J3.2 Single Openings. The flow through a single sharp opening due to bidi-
rectional buoyancy-driven flow (Vbd_sharp) (see Etheridge and Sandberg [1996]
in Informative Appendix M) is expressed as follows:

T
V so_sharp = 0.21  A w  gH s --------- (J-2)
T ref

where
Aw = free unobstructed area of the window, or openable area
T = temperature difference between indoors and outdoors. Given the
conservative nature of a prescriptive path, a temperature difference of
1°C (1.8°F) is assumed for these calculations. In reality, this temperature
will depend on the internal gains in the space and will likely be higher
than 1°C (1.8°F), leading to higher airflows (and a smaller window area
requirement).
Hs = vertical dimension of the opening
g = gravity constant
Tref = reference temperature in Kelvin (or Rankine), typically equal to Tin, Tout
or an expected average. A reference temperature of 21°C (70°F, 294K)
was assumed for these calculations.
A safety factor is incorporated assuming that an awning window is used.
Awning (or top-hinged) windows are among the most common windows used
for natural ventilation, and, because of their uneven vertical area distribution,
are more inefficient than a sliding window (sharp opening) at driving flow. An
efficiency v of around 83% (value used in these calculations) when compared
to sliding windows is inferred from

V so = V so_sharp   w (J-3)

Assuming a height-to-width ratio for the window of RH/W (R = H/W), the


window area can be rewritten as

H S2
Aw = ------------ (J-4)
R H/W

The required openable area as a fraction of the zone’s floor area is therefore
calculated by equating the bidirectional buoyancy-driven flow through a single
awning opening (Vso) to the goal flow rate (Vbz) obtained from Table 6-1.

V so = V bz (J-5)

And solving for window area,


240 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

  4 5
Aw  V bz  1
------ =  ----------------------------------------------------------------------
-  -----  100 (J-6)
Az  T- Az
 0.21  0.83  R 4  g --------
H  W T ref
J3.3 Vertically Spaced Openings. The flow rate Vvs through vertically spaced
openings of areas As (the smallest sum of opening areas, either upper openings
or lower openings) and Al (the largest sum of opening areas, either upper open-
ings or lower openings) is obtained using the following equation:

T
V vs = A eff  C d  2gH --------- (J-7)
T ref
where
Aeff = effective window area, defined as

1 As Aw
A eff = ----------------------- = ------------------- = -------------------------------------------- (J-8)
1- ----- 1- 1
----- + 1 + R2 1 + R 2   1 + ---
A s2 A l2  R

Aw = is the total sum of all opening areas


Aw = As + At (J-9)

R = area ratio between As and Al

As
R = ----- (J-10)
Al

H is the shortest vertical distance between the center of the lowest open-
ings and that of the upper openings.
All other constants are the same as in the single opening scenario.
The required openable area as a fraction of the zone’s floor area is therefore
calculated by equating the flow through two sets of vertically spaces openings
Vvs to the goal flow rate Vbz obtained from Table 6-1.
V vs = V bz (J-11)

Solving for window area:

Aw V bz (J-12)
-  1 + R 2   1 + --1-  ----
1-
------ = ----------------------------------------  100
Az T  R  A z
C d  2gH ---------
T ref
APPENDIX A5.3 | 241

J4. CONTROL AND ACCESSIBILITY (MIXED-MODE VENTILATION)


Mixed-mode ventilation is a hybrid system used to maintain IAQ and internal
thermal temperatures year-round using both natural and mechanical ventilation
systems.
a. Natural ventilation systems use natural forces such as wind and thermal
buoyancy to ventilate and cool spaces.
b. Mechanical ventilation systems use mechanical systems with fans to supply
and exhaust air from a space, provide humidity control, and, if required, fil-
ter possible contaminants.
By preferentially using natural ventilation when outdoor air conditions are
suitable, energy costs and carbon emissions can be minimized. Sensors are
used to identify when natural ventilation is less effective at providing suitable
indoor temperatures, humidity levels, and contaminant levels, and indicate that
a transition to mechanical ventilation should occur. The transition between
modes can be manual or automatic, as dictated by the needs of the owner/occu-
pants. The use of each mode when appropriate will ensure year-round accept-
able IAQ.
242 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

(This appendix is not part of this standard. It is merely informative and


does not contain requirements necessary for conformance to the stan-
dard. It has not been processed according to the ANSI requirements for a
standard and may contain material that has not been subject to public
review or a consensus process. Unresolved objectors on informative
material are not offered the right to appeal at ASHRAE or ANSI.)

INFORMATIVE APPENDIX K
COMPLIANCE
This appendix contains compliance suggestions that are intended to assist
users and enforcement agencies in applying this standard.

J5. SECTION 6 NATURAL VENTILATION PROCEDURE


Natural ventilation systems shall follow either the prescriptive or the
engineered system compliance path.

For the prescriptive compliance path:


K Is a mechanical system compliant with either Section 6.2 or 6.3 included?
K If no, does design comply with Exceptions 1 or 2 of Section 6.2.1?
K Do maximum distances from openings comply with Sections 6.2.1.2,
6.2.1.3, or 6.2.1.4?
K Do opening sizes comply with the requirements of Section 6.4.2?
K Is net free area of openings specified?
K Are sill-to-head heights specified?
K Are aggregate widths specified?
K Are controls readily accessible?

For the engineered compliance path:


K Do the design documents provide evaluation of the following:
K Hourly environmental conditions, including, but not limited to, outdoor
air dry-bulb temperature; dew-point temperature; outdoor concentra-
tion of contaminants of concern (including but not limited to PM2.5,
PM10, and ozone), where data are available; wind speed and
direction; and internal heat gains during expected hours of natural ven-
tilation operation.
K The effect of pressure losses along airflow paths of natural ventilation
airflow on the resulting flow rates, including, but not limited to, inlet
vents, air transfer grills, ventilation stacks, and outlet vents.
K Qualification of natural ventilation airflow rates of identified airflow
paths accounting for wind and thermally induced driving pressures.
K Outdoor air is provided in sufficient quantities to ensure pollutants and
odors of indoor origin do not result in unacceptable IAQ as established
under Section 6.2.1.1 and/or 6.3.
APPENDIX A5.3 | 243

K Outdoor air introduced into the space through natural ventilation sys-
tem openings does not result in unacceptable IAQ according to Sec-
tions 6.1.4.1 through 6.1.4.4.
K Effective interior air barriers and insulation are provided that separate
naturally ventilated spaces from mechanically cooled spaces, ensuring
that high-dew-point outdoor air does not come into contact with
mechanically cooled surfaces.
K Are controls readily accessible?

A5.3.1 REFERENCE
ASHRAE. 2019. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
MINIMUM VENTILATION
REQUIREMENTS PER
ASHRAE STANDARD 62.1-2019
This appendix provides a benchmark for minimum ventilation air
requirements when a performance-based approach to compliance is used. For
the application of this design guide, only areas appropriate for offices and
classrooms are excerpted from Table 6-1, “Minimum Ventilation Rates in
Breathing Zone” of ASHRAE Standard 62.1-2019 (ASHRAE 2019). Table 6-1
is replicated here with ASHRAE’s permission.
TABLE A5.4.1 Modified Excerpts from Table 6-1, “Minimum Ventilation Rates in
Breathing Zone” from ASHRAE Standard 62.1-2019
People Outdoor Air Area Outdoor Air Default Values
Rate, Rp Rate, Ra Occupant Density Air
Occupancy Category
cfm/ L/s/ #/1000 ft2or Class
person person cfm/ft2 L/s/m2
#/100 m2

Educational Facilities

Daycare (through age 4) 10 5 0.18 0.9 25 2

Daycare sickroom 10 5 0.18 0.9 25 3

Classrooms (ages 5-8) 10 5 0.12 0.6 25 1

Classrooms (age 9 plus) 10 5 0.12 0.6 35 1

Lecture classroom 7.5 3.8 0.06 0.3 65 1

Lecture hall (fixed seats) 7.5 3.8 0.06 0.3 150 1

Art classroom 10 5 0.18 0.9 20 2

Science laboratories 10 5 0.18 0.9 25 2

University/college 10 5 0.18 0.9 25 2


laboratories

245
246 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE A5.4.1 Modified Excerpts from Table 6-1, “Minimum Ventilation Rates in
Breathing Zone” from ASHRAE Standard 62.1-2019 (Continued)
People Outdoor Air Area Outdoor Air Default Values
Rate, Rp Rate, Ra Occupant Density Air
Occupancy Category
cfm/ L/s/ 2
#/1000 ft or Class
person person cfm/ft2 L/s/m2
#/100 m2

Wood/metal shop 10 5 0.18 0.9 20 2

Computer lab 10 5 0.12 0.6 25 1

Media center 10 5 0.12 0.6 25 1

(Public Assembly) Libraries 5 2.5 0.12 0.6 10 1

Music/theater/ dance 10 5 0.06 0.3 35 1

Multiuse assembly 7.5 3.8 0.06 0.3 100 1

Cafeteria/ fast-food 7.5 3.8 0.18 0.9 100 1


dining

Gym, sports arena (play 20 10 0.18 0.9 7 2


area)

Sports Spectator areas 7.5 3.8 0.06 0.3 150 1

Auditorium seating area 5 2.5 0.06 0.3 150 1

Stages/studios 10 5 0.06 0.3 70 1

Office Facilities

Office space 5 2.5 0.06 0.3 5 1

Telephone/data entry 5 2.5 0.06 0.3 60 1

(General) Break rooms 5 2.5 0.06 0.3 25 1

Coffee stations 5 2.5 0.06 0.3 20 1

(Office) Breakrooms 5 2.5 0.12 0.6 50 1

Conference/ meeting 5 2.5 0.06 0.3 50 1

Corridors — — 0.06 0.3 — 1

(Office) Main entry lobbies 5 2.5 0.06 0.3 10 1

Reception areas 5 2.5 0.06 0.3 30 1

(Public Assembly) Lobbies 5 2.5 0.06 0.3 150 1


APPENDIX A5.4 | 247

TABLE A5.4.1 Modified Excerpts from Table 6-1, “Minimum Ventilation Rates in
Breathing Zone” from ASHRAE Standard 62.1-2019 (Continued)
People Outdoor Air Area Outdoor Air Default Values
Rate, Rp Rate, Ra Occupant Density Air
Occupancy Category
cfm/ L/s/ 2
#/1000 ft or Class
person person cfm/ft2 L/s/m2
#/100 m2

Occupiable storage for 5 2.5 0.06 0.3 2 1


dry materials

Occupiable storage 5 2.5 0.12 0.6 2 2


rooms for liquids or gels

Shipping/ receiving 10 5 0.12 0.6 2 2

Sorting, packing, light 7.5 3.8 0.12 0.6 7 2


assembly

Health club/aerobics 20 10 0.06 0.3 40 2


room

Health club/weight room 20 10 0.06 0.3 10 2

A5.4.1 REFERENCE
ASHRAE. 2019. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
SAMPLE RESULTS
FROM AN ENGINEERED
NATURAL VENTILATION
SYSTEM ANALYSIS
This appendix presents the results of the analysis of the office space as
described in Appendix A1.1 using the software Oasys BEANS (Oasys n.d.), a
dynamic heat transfer and bulk airflow modeling tool, for New York City, NY.
Results for other cities can be found in Website Appendix WA5.3, accessible
at ashrae.org/naturalventilation.
For the wind-driven analysis, the upper and lower opening sizes were kept
constant for each month and each location. The difference in results is due to
the ensuing climate. For an optimum design, an iterative process of varying
opening sizes to determine an optimum opening size is recommended.

TABLE A5.5.1 Analysis Model Opening Sizes and the Schedule of Openness
Throughout a Typical Day
Upper Lower
Area, Area, Separation,
8 9 10 11 12 13
ft (m)
ft2 (m2) ft2(m2)

Buoyancy 2.561 (0.25) 2.561 (0.25) 0.427 (0.12) % opening 50 100 100 100 100 100
14 15 16 17 18

100 100 100 100 50

249
250 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A5.5.1 NEW YORK CITY, NY, RESULTS


A5.5.1.1 Buoyancy-Driven Ventilation

FIGURE A5.5.1 Natural-ventilation-created CFM versus CFM thresholds from ASHRAE


Standard 62.1 and California’s Title 24 in for New York City, NY, in March as a result of
buoyancy-driven ventilation.

FIGURE A5.5.2 Natural-ventilation-created CFM versus CFM thresholds from ASHRAE


Standard 62.1 and California’s Title 24, for New York City, NY, in June, as a result of buoyancy-
driven ventilation.
APPENDIX A5.5 | 251

FIGURE A5.5.3 Natural-ventilation-created CFM versus CFM thresholds from ASHRAE


Standard 62.1 and California’s Title 24 for New York City, NY, in September as a result of
buoyancy-driven ventilation.

FIGURE A5.5.4 Natural-ventilation-created CFM versus CFM thresholds from ASHRAE


Standard 62.1 and California’s Title 24 for New York City, NY, in December, as a result of
buoyancy-driven ventilation.
252 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A5.5.1.2 Wind-Driven Ventilation

FIGURE A5.5.5 Natural-ventilation-created CFM versus CFM thresholds from ASHRAE


Standard 62.1 and California’s Title 24 for New York City, NY, in March, as a result of wind-
driven ventilation.

FIGURE A5.5.6 Natural-ventilation-created CFM versus CFM thresholds from ASHRAE


Standard 62.1 and California’s Title 24, for New York City, NY, in June, as a result of wind-
driven ventilation.
APPENDIX A5.5 | 253

FIGURE A5.5.7 Natural-ventilation-created CFM versus CFM thresholds from ASHRAE


Standard 62.1 and California’s Title 24, for New York City, NY, in September, as a result of
wind-driven ventilation.

FIGURE A5.5.8 Natural-ventilation-created CFM versus CFM thresholds from ASHRAE


Standard 62.1 and California’s Title 24, for New York City, NY, in December, as a result of
wind-driven ventilation.
254 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A5.5.2 REFERENCE
Oasys Software. n.d. UNIPAC software for students and institutions. https://
www.oasys-software.com/education/. London: Arup. BEANS v.16.2 suite of
software, specifically CLIDAT v.16.2 and SUNPOS v.16.2.
DEMONSTRATING COMFORT
STANDARD COMPLIANCE FOR
NATURAL CONDITIONING
An acceptable comfort zone is prescribed by ANSI/ASHRAE Standard 55-
2017, Thermal Environmental Conditions for Human Occupancy (ASHRAE
2017a). Comfort is defined as the conditions under which eighty percent or
more of the building occupants will find an area thermally acceptable in still
air and shade conditions (ASHRAE 2017a).
When natural ventilation and natural conditioning are pursued, it is
important to ensure that the minimum outdoor air requirements (as discussed
in Chapter 5) are achievable under all conditions, including hours when the
backup mechanical ventilation system is operational. This is required per
Section 6.1.3 and Section 6.4 of ANSI/ASHRAE Standard 62.1 (ASHRAE
2019b). Complying with the natural ventilation requirements does not imply
any likelihood of compliance with the comfort standard. At high outdoor air
temperatures, the amount of air necessary to provide a heat-absorption function
is far higher than the amount of air required to meet the minimum ventilation
requirements. The design team should never assume that complying with
ASHRAE Standard 62.1 (ASHRAE 2019b) will result in a comfortable space.

6.1 CONFIRMING THE REQUIREMENT TO


MEET A COMFORT STANDARD
Neither the International Mechanical Code® (ICC 2017) nor the California
Mechanical Code (CBSC 2019) reference ASHRAE Standard 55 as a
requirement for human comfort. Thus, it would be up to the local jurisdiction
to require compliance with ASHRAE Standard 55 (2017a) as a comfort
standard.
California Title 24, Part 6 (CEC 2018a) does not explicitly address the
permissibility of natural conditioning as a space-conditioning option. The code
allows the design team to demonstrate that a backup heating and cooling
system does not have to be modeled if the indoor condition model of the
natural ventilation system can show that the space can be controlled to remain
within a prescribed set of temperatures, namely the space thermostat throttling

255
256 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

range of 2°F (1°C). By the very nature of naturally conditioned spaces, the
indoor temperatures will likely fluctuate far more than this. Thus, if natural
conditioning is pursued in California, it is likely that a modification will need
to be negotiated with the authority having jurisdiction (AHJ) to redefine the
comfort range. It is recommended that the comfort control range be widened to
at least that of the guidelines in ANSI/ASHRAE Standard 55 (ASHRAE
2017a) as an alternate compliance path by showing the AHJ the comfort
control parallels between a natural conditioning system and the code-required
design of space-conditioning systems. This relies on the preexisting code
explicitly referencing the need to size space-conditioning equipment “as a
rule” to meet Section 140.4 of California’s Title 24, which states:

Indoor design temperature and humidity conditions for general comfort


applications shall be determined in accordance with ASHRAE Standard 55
or the ASHRAE Handbook, Fundamentals Volume, Chapter 8, except that
winter humidification and summer dehumidification shall not be required.
(CEC 2018a)

The remainder of the chapter, therefore, will address only the ASHRAE
Standard 55 (ASHRAE 2017a) requirements.

6.2 IDENTIFYING THE APPROPRIATE ASHRAE STANDARD 55


COMFORT COMPLIANCE METHOD
ASHRAE Standard 55 (ASHRAE 2017a) offers four methods for defining
the comfort zone criterion to demonstrate compliance. The standard is not a
design guide; rather, it is the collection of minimum criteria to be met in a pass/
fail manner. The four methods of evaluating thermal comfort are not
necessarily applicable to every situation. The standard defines limits on the
range of environmental and personal factors that each method for evaluating
thermal comfort can consider. The standard sets tighter limits on the range of
factors when using the graphical method and is least restrictive when using the
adaptive method. It is the design team’s responsibility to ensure that they are
using the correct method based on the known and predicted conditions of the
space and the representative occupants within the space.
Figure 6.1 and Table 6.1 represent these four paths and their applicability.
(Note that the “Adaptive Method” or the “Adaptive Comfort Method” are
colloquial terms used for the method officially called “acceptable thermal
environments only for occupant-controlled naturally conditioned spaces”
under Section 5.4 of ASHRAE Standard 55-2017 (ASHRAE 2017a).
There will be times of the year when a natural conditioning system could
operate under outdoor air conditions that allow the space to be compliant with
the simplified Graphical Method or the computer-based prescriptive Analytical
Method. Under warm conditions, however, most naturally conditioned spaces
rely on the elevated air speed of both the Elevated Air Speed Method and the
Adaptive Method, along with possibly the extended allowable comfort range
CHAPTER 6 | 257

FIGURE 6.1 Methods of determining thermal comfort, as provided in the Standard 55 User’s
Manual.
(ASHRAE 2016)

TABLE 6.1 Limits on Thermal Comfort Factors that can be Used Within Each of the
Four Methods of Determining Thermal Comfort of the Representative Occupant
(ASHRAE 2017a)
Average Air Speed
with ASHRAE Humidity Comfort Zone
met clo Other Criteria
Standard 55 Ratio Method
Reference Clause
< 0.2 m/s (40 fpm) per <0.012 kg·H2O/ 1.0 – 1.3 met 0.5 – 1.0 clo — Graphic Comfort
Section 5.3.1.1 kg dry air Zone Method

< 0.2 m/s (40 fpm) per Any 1.0 – 2.0 met 0 – 1.5 clo — Analytical Comfort
Section 5.3.2.1 Zone Method

> 0.2 m/s (40 fpm) per Any 1.0 – 2.0 met 0 – 1.5 clo See Figure 6.2 for air Elevated Air Speed
Section 5.3.3.1 speed limits if occupants Comfort Zone Method
do not have control

Up to 1.2 m/s (236 fpm) Any 1.0 – 1.3 met Free to adjust • No cooling installed Occupant-Controlled
per Section 5.4.2.4 if within a range • Heat not operating Naturally Conditioned
occupant control provided at least as • Occupants control Spaces (also known as
wide as 0.5 – openings the Adaptive Method
1.0 clo • Prevailing mean out- or the Adaptive
door air Comfort Method)
temperature 50°F to
92.3°F (10°C to
33.5°C)
258 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 6.2 Metabolic Rates for Typical Tasks Applicable to Natural Conditioning
Activity Met Units Metabolic Rate
Seated, quiet 1.0 18 Btu/h·ft2 (60 W/m2)
Reading, seated
Writing

Typing 1.1. 20 Btu/h·ft2 (65 W/m2)

Standing, relaxed 1.2 22 Btu/h·ft2 (70 W/m2)


Filing, seated

Filing, standing 1.4 26 Btu/h·ft2 (80 W/m2)

Walking about 1.7 31 Btu/h·ft2 (100 W/m2)

Lifting/packing 2.1 39 Btu/h·ft2 (120 W/m2)


Data from Table 5.2.1.2, “Metabolic Rates for Typical Tasks” from ASHRAE Standard 55 (2017a)

under the Adaptive Method. It is important to note that the methods are means
of showing comfort compliance under a very narrowly defined set of
parameters. Each path has constraints on applicability for each of the available
comfort paths in ASHRAE Standard 55 (ASHRAE 2017a), primarily
surrounding variables such as clo, met, and air speed. Table 6.1 summarizes
those constraints.
Chapter 9, “Thermal Comfort” of ASHRAE Handbook—Fundamentals
(ASHRAE 2017b) and ASHRAE Standard 55 (ASHRAE 2017a) describe how
to estimate met and clo in greater detail.

• met: The level of activity of the occupant is associated with their


metabolic rate, which in turn affects the thermal conditions at which
they are likely to be comfortable. Table 7, “Typical Insulation and
Permeation Efficiency Values for Western Clothing Ensembles” in
Chapter 9 of ASHRAE Handbook—Fundamentals (ASHRAE 2017b)
and Table 5.2.1.2, “Metabolic Rates for Typical Tasks” of ASHRAE
Standard 55 (ASHRAE 2017a) list the typical met rates by activity
type. For this design guide, activities related to office occupancies are
noted in Table 6.2. Time-weighted averaging of activities is allowed
under Section 5.2.1.3, “Time-Weighted Averaging” (ASHRAE 2017a)
to show that the metabolic rates for the space fall into the allowable
limits on met values as noted in Table 6.1.

As previously noted, the applicability of the natural conditioning clause is


capped at 1.3 met, and the applicability of the standard itself is capped at 2.0 met.
For ranges above 1.3 met, refer to Website Appendix WA6.1, “Excerpts of
ASHRAE Standard 55 on Natural Ventilation,” accessible at ashrae.org/
naturalventilation, to determine the likelihood of heat stress index in the space if
natural conditioning is pursued.
CHAPTER 6 | 259

FIGURE 6.2 Comparison of maximum air velocities from ASHRAE Standard 55.
(ASHRAE 2017a)

• clo: Table 9, “Garment Insulation Values,” in Chapter 9 of ASHRAE


Handbook—Fundamentals (ASHRAE 2017b) describes how to estimate
the thermal insulation provided by clothing. See Informative Appendix J,
“Occupant-Controlled Naturally Conditioned Spaces” of ASHRAE
Standard 55 (ASHRAE 2017a) for an example of estimating the clo value
for a typical women’s and men’s ensemble. A greater variety of clothing
types can be found in Website Appendix WA6.3, “Garment Insulation
Values,” accessible at ashrae.org/naturalventilation.

In addition to met and clo, a third variable related to air velocity is important to
calculate or estimate because air movement influences the bodily heat balance by
affecting the rate of convective heat transfer between the skin and air and the rate
of bodily cooling through the evaporation of skin moisture. Increased air
movement can increase the range of temperature and humidity in which people
will feel comfortable, and thus, elevated air speeds are a key tool in achieving
comfort in most naturally conditioned spaces. Figure 6.2 compares maximum air
velocities as denoted in each path’s requirements.
260 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Under the Elevated Air Speed Comfort Zone Method, the following limits
from Section 5.3.3.4 of ASHRAE Standard 55 (2017a) on air velocity apply:

• Air speed should be greater than 0.2 m/s (40 fpm)

• For operative temperatures to below 23.0°C (73.4°F), the limit to average


air speed Va should be 0.2 m/s (30 fpm).
• For operative temperatures to between 72.5°F and 77.9°F (22.5°C and
25.5°C), the upper limit to average air speed Va is approximated by a curve
using the following equation:

2
V a = 31375.7 – 857.295t o + 5.86288  t o   fpm, °F  (6.1 I-P)

2
V a = 50.49 – 4.4047t o + 0.096425  t o   m/s, °C  (6.1 I-P)

• For operative temperatures to above 77.9°F (25.5°C), the upper limit to


average air speed Va should be 0.8 m/s (160 fpm).

This upper limit for the Elevated Air Speed Comfort Zone Method—an air
speed of 160 fpm (0.8 m/s)—is appropriate for most offices and commercial
spaces, because it corresponds to the point at which loose paper, hair, and other
light objects may be blown about. Note that when the occupants have control,
there is no upper limit so long as one of the following control mechanisms is
available (ASHRAE 2017a):

a. One means of control exists for every six occupants or fewer.


b. One means of control exists for every 84 m2 (900 ft2) or less.
c. In multioccupant spaces where groups gather for shared activities, such
as classrooms and conference rooms, at least one control shall be pro-
vided for each space, regardless of size. Multioccupant spaces that are
subdivided by movable walls shall have one control for each space sub-
division.

In the case of natural conditioning, the air velocity can rise to 1.2 m/s
(236 fpm) as per Table 5.4.2.4 in ASHRAE Standard 55 (ASHRAE 2017a).
The Adaptive Comfort Method, or Occupant-Controlled Naturally
Conditioned Spaces Method, also requires that occupants have control over the
window openings. Although there are no regulations on control mechanism
layout or rules on sharing cohort size for this path, the rules of the Elevated Air
Speed Comfort Zone Method may serve as a useful starting point, with
expansion to more openings or more distribution of openings depending on the
results of the dynamic thermal analysis and bulk airflow modeling exercise.
CHAPTER 6 | 261

6.3 LIMITATIONS ON THE USE OF THE


ADAPTIVE COMFORT METHOD
Appendix J, “Occupant-Controlled Naturally Conditioned Spaces,” in
ASHRAE Standard 55 (ASHRAE 2017a) advises:

For the purposes of ASHRAE Standard 55 Section 5.4, occupant-


controlled naturally conditioned spaces are those spaces where the
thermal conditions of the space are regulated primarily by the
occupants through opening and closing of openings in the building
envelope. Field experiments have shown that occupants’ thermal
responses in such spaces depend in part on the outdoor climate and may
differ from thermal responses in buildings with centralized HVAC
systems primarily because of the different thermal experiences,
changes in clothing, availability of control, and shifts in occupant
expectations. This optional method is intended for such spaces.
The space in question must be equipped with operable openings to
the outdoors and can be readily opened and adjusted by the occupants
of the space.

In addition to the requirement for a mechanical ventilation system as noted


at the beginning of this chapter, ASHRAE Standard 55 (ASHRAE 2017a)
Section 5.4.1, “Applicability,” imposes a strict set of limitations on the use of
natural conditioning related to the elimination of mechanical cooling, limits on
metabolic rates, the assurance of the occupants’ ability to modify their
clothing, and limitations on temperature ranges outdoors. The following
sections discuss each limitation separately so the design team can navigate a
path of compliance or request appropriate modifications from the AHJ.

6.3.1 No Mechanical Cooling System Installed


The limitation set by Section 5.4.1a of ASHRAE Standard 55 states
“[t]here is no mechanical cooling system (e.g., refrigerated air conditioning,
radiant cooling, or desiccant cooling) installed. No heating system is in
operation” (ASHRAE 2017a). This section would officially preclude the
ability to pursue any form of mixed-mode or hybrid ventilation within the
same zone where natural conditioning takes place. Thus, changeover and
concurrent systems are not allowed under this limitation. Only zoned mixed-
mode/hybrid ventilation is technically allowed. This corresponds to a design in
which certain areas of the building may be conditioned by mechanical cooling
equipment, and other areas are permanently conditioned only by natural
conditioning system.
As noted in Appendix A3.5, there are virtually no locations in the continental
U.S. that are likely to achieve indoor comfort conditions solely relying on natural
conditioning for the daytime hours of typical occupancy. Thus, it would be the
responsibility of the design team to negotiate with the owner and the AHJ on
whether the specific language around installation should be applied. A stringent
application of this clause would shut down any further pursuit of a natural
262 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

conditioning scheme. This would be unfortunate as there are examples around the
world that show successful engineered control of natural conditioning and
mechanical conditioning systems working in conjunction with one another to
achieve comfort throughout the year and significant energy savings.
Significant discussion with the client and the AHJ will be necessary to
agree on any means to navigate the use of mixed-mode or hybrid ventilation
schemes.

6.3.2 Limits on Metabolic Rates


The limitation set by Section 5.4.1b states that “[r]epresentative occupants
have metabolic rates ranging from 1.0 to 1.3 met.” This requirement derives
from the fact that the Adaptive Comfort Method of establishing the allowable
comfort zone has an empirical basis from the meta-analysis of office
occupancies under a natural conditioning scheme. Because the research basis
had limited met, so too must the standard for the present time. The met value
limitation is reasonable because the reliance on natural conditioning for human
comfort and well-being should acknowledge that high met activities in high-
temperature conditions can lead to heat stress. If met rates are anticipated to be
higher than that which is covered by the standard, then the design team should
check for the risk of Heat Stress Index (HSI) as per the procedure in Website
Appendix WA6.2, “Heat Stress Index Calculation Method,” and discuss with
the owner and the AHJ whether it would be allowed to have some minor heat
stress under extreme conditions. Otherwise, mechanical cooling should be
provided.

6.3.3 Adaptability of Clothing


The limitation set by Section 5.4.1c of ASHRAE Standard 55 states
“[r]epresentative occupants are free to adapt their clothing to the indoor and/or
outdoor thermal conditions within a range at least as wide as 0.5 to 1.0 clo”
(ASHRAE 2017a). This criterion specifically limits the use of natural
conditioning to those social and cultural environments that allow all occupants
the freedom to modify their clothing toward short-sleeve shirts and lightweight
clothing when the air temperature rises indoors. This may expose more skin
surfaces to the air to improve convective heat loss, and so this consequence of
freedom of clothing adjustment must be discussed with the owner and any
human resource and supervisory personnel within any tenant organization. If
full skin and/or head coverage is required throughout the year, it is
recommended that a full analysis of clo factor be completed on the ensembles
that are socially acceptable before pursuing a natural conditioning scheme. See
example calculation in ANSI/ASHRAE Standard 55’s Informative Appendix
J, “Occupant-Controlled Naturally Conditioned Spaces” (ASHRAE 2017a).
Excerpted garment insulation values from Tables 7, 8, and 9 from Chapter 9 of
ASHRAE Handbook—Fundamentals (ASHRAE 2017b) are included in
Website Appendix WA6.3, “Garment Insulation Values,” accessible at
ashrae.org/naturalventilation.
CHAPTER 6 | 263

6.3.4 Constraints on Outdoor Air Temperatures


The limitation set by Section 5.4.1d states, “[t]he prevailing mean outdoor
temperature is greater than 10°C (50°F) and less than 33.5°C (92.3°F).” This
criterion establishes the allowable outdoor air temperatures that are likely to
ensure acceptable indoor temperatures with occupant control over windows
and openings. When temperatures expand outside this range on the low end,
heating systems are likely to exceed allowable heat outputs as regulated by the
applicable energy code, and thus, occupants will feel overcooled. When
temperatures extend outside the range on the upper end, the outdoor air
temperature is essentially above the 33°C (92°F) mean skin temperature of a
resting human in comfortable surroundings (as per Equation 85 in Chapter 9,
“Thermal Comfort” of ASHRAE Handbook—Fundamentals [ASHRAE
2017b]), and thus, the outdoor air loses its ability to cool a body during the
time period when the air temperature is this high. The actual outdoor air
temperatures are likely to fail in compliance for indoor conditions at
temperatures lower than the upper limit on prevailing mean as noted in this
limitation.

6.4 APPLYING THE ADAPTIVE COMFORT ZONE METHOD


The Standard 55 User’s Manual (ASHRAE 2016) provides a worked
example of the application of the Adaptive Comfort Method, which is
replicated in Website Appendix WA6.5, “Worked Example from Standard 55
User’s Manual,” accessible at ashrae.org/naturalventilation. Noted that this
example used a flat mean approach for the analysis proposed, which was
applicable in older versions of the standard. The new approach using a
prevailing running mean is discussed in Section 6.4.3, “Postprocessing Results
from Simulation Modeling.”
The procedural steps followed in that example include:
1. Evaluate the applicability of the design against the requirement for opera-
ble elements and the four previously noted constraints.
2. Calculate the comfort zone thresholds for compliance.
3. Determine comfort using a dynamic thermal simulation.
4. Resolve comfort conditions if space is found to be noncompliant.

6.4.1 Acceptability Limits


The acceptability limits for this comfort zone method originated as a
graphical representation of empirical data (noted in Figure 5.4.2 in ASHRAE
Standard 55 [2017a]), but the information has since been resolved into the
equations noted in Section 5.4.2.2:

Upper 80% acceptability limit (°C) = 0.31t pma  out  + 21.3 (6.2 I-P)

Upper 80% acceptability limit (°C) = 0.31t pma  out  + 60.5 (6.3 I-P)
264 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Lower 80% acceptability limit (°C) = 0.31t pma  out  + 14.3 (6.4 I-P)

Lower 80% acceptability limit (°F) = 0.31t pma  out  + 47.9 (6.5 I-P)

As noted in Standard 55 User’s Manual (ASHRAE 2016):

It is important to note that only the 80% acceptability limit is used


when a user is to show compliance with the Adaptive Method. The
90% acceptability limits are for illustrative purposes only but can be
used on a voluntary basis if a designer or building owner chooses to
provide higher level of acceptability. For any jurisdictions requiring
compliance with Standard 55 though, compliance is only required at
the 80% acceptability limits.

6.4.2 Use of Bulk Airflow and Dynamic Thermal


Simulation Modeling
When the design team must demonstrate the compliance of a proposed
scheme with ASHRAE Standard 55’s requirements, it is typical that a bulk
airflow and dynamic thermal simulation model will be required, at a minimum,
to understand the complexities of how a space performs under natural
ventilation and natural conditioning.
Bulk airflow analytical modeling involves solving simple (linear)
equations by hand (or spreadsheet) to give more detail about aspects of the
ventilation flow, e.g., stack pressure at a certain height or likely stratification
height. This type of modeling alone will address natural ventilation flows
under limited simple conditions, as noted in Chapter 5, but is not sufficiently
sophisticated to address natural ventilation.
Combined dynamic heat transfer and bulk airflow modeling tools are much
more appropriate for natural conditioning. These types of models use zones to
represent volumes of air that are connected via flow resistances to form a
network. Each zone is assumed to be well-mixed, containing single point
evaluations of temperature, pollutant concentrations, and hydrostatic pressure
at an air point. Flow resistances represent airflow paths between zones within
the same occupied space, between adjacent spaces through openings, or
between a space and the outdoor air through openings. Airflow rates are
calculated between zones and through openings. These models are particularly
suitable for annual simulations and coupling with dynamic thermal models.
These models differ from analytical models because they solve a set of
simultaneous, nonlinear equations that ensure that the entire airflow network is
in balance (i.e., that the net mass flow into and out of any one zone is zero).
This requires the use of a computer program. The tools are useful for checking
that minimum ventilation air requirements of ASHRAE Standard 62.1
(ASHRAE 2019b) are met, while also meeting the Adaptive Comfort Method
requirements in ASHRAE Standard 55 (ASHRAE 2017a). The models tend to
run relatively quickly, and some of the software programs can also calculate
the percentage of people dissatisfied (PPD) dynamically from the results of
this modeling technique.
CHAPTER 6 | 265

TABLE 6.3 Natural Ventilation/Natural Conditioning Analysis Software Packages


Software Name Description
CONTAM This is a multizone indoor air quality and ventilation analysis program produced
(NIST n.d.) by the National Institute of Standards and Technology (NIST). For natural
ventilation analysis, it is often used in conjunction with other software packages
such as EnergyPlus™ or TRNSYS engines to represent solar and internal heat gain
drivers.

EnergyPlus™ EnergyPlus is a whole-building energy simulation program created by the U.S.


(DOE n.d.) Department of Energy (DOE). The engine has an embedded multizone airflow
network modeling capacity that makes it well suited for bulk airflow and dynamic
thermal analysis. As public domain software, it is often used as the analysis
engine by many other graphical user interfaces noted in this table.

DesignBuilder Developed by DesignBuilder Software Ltd. and built on an EnergyPlus™ engine,


(DesignBuilder Software n.d.) DesignBuilder has an integrated interface to support load, energy, daylighting,
multizone airflow, and CFD analysis. See sample output in Appendix A6.1, made
available by the vendor.

IES Virtual Environment IES Virtual Environment is developed by Integrated Environmental Solutions Ltd.
(IESVE) (IES n.d.-b) and built on an EnergyPlus™ engine. It has an integrated interface to support
load, energy, daylighting, multizone airflow, and CFD analysis. See sample
output in Appendix A6.2 made available by the vendor.

OpenStudio® This open-source software and development platform was developed by the
National Renewable Energy Laboratory (NREL) with the DOE in collaboration with
(Alliance for Sustainable Energy n.d.)
Argonne National Laboratory (ANL), Lawrence Berkeley National Laboratory
(LBNL), Oak Ridge National Laboratory (ORNL), and Pacific Northwest National
Laboratory (PNNL), with current management under the Alliance for Sustainable
Energy, LLC. It uses the EnergyPlusTM engine and Radiance for daylight analysis.

Simergy™ This software was originally developed in conjunction with the CEC, LBNL, U.S.
(Digital Alchemy n.d.) DOE, Hydro Quebec, and several private entities to create an interface tool to
support robust building information modeling interoperability to the EnergyPlusTM
simulation engine.

Sefaira This software is currently owned by Trimble, Inc. and marketed as an early-stage,
(Trimble n.d.) high-performance design tool that is easily integrated with its SketchUp tool. It
uses the EnergyPlus™ engine and Radiance for daylight analysis.

Oasys Room Module This software is managed by Oasys Ltd., a subsidiary of Arup Group (Disclosure:
(Oasys Software n.d.) one of the authors is an employee of this company). Its analysis engine for natural
ventilation is based on CIBSE Applications Manual AM10’s calculation
methodologies. See sample output in Website Appendix WA6.6, “Engineered
Natural Ventilation System Analysis Sample Results,” accessible at ashrae.org/
naturalventilation.

TRNSYS TRNSYS is a general transient systems simulation program created by the Solar
(UWM n.d.) Energy Laboratory of the University of Wisconsin, Madison to simulate
interconnected and mutually forcing thermal, electrical, and energy systems. For
natural ventilation analysis, it is often linked with CONTAM (NIST n.d.) for the
dynamic airflow analysis.
266 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Once a dynamic heat transfer bulk airflow model has been completed,
some design teams use computational fluid dynamics (CFD) software
programs to look at detailed airflow patterns. Computational fluid dynamics
(CFD) is a branch of physics that deals with the study of the mechanics of
fluid: liquid, plasmas, and gases and forces acting on them. CFD is based on
Navier–Stokes equations that describe how pressure, velocity, density, and
temperature of a moving fluid are related. It uses numerical methods,
mathematical modeling, and software tools to analyze and solve problems that
involve fluid flows and liquid and gas interactions with surfaces, as defined by
boundary conditions.
Do not forget that a CFD model requires boundary conditions, which are
provided from combined dynamic heat transfer and bulk air models, so CFD is
not a replacement for these simpler multizone network models. CFD
techniques divide the space or group of spaces into a mesh of small cells and
solve complex, nonlinear, differential equations in each one. This gives a
detailed picture of the flow pattern (i.e., temperature, air velocities, and
pollutant distribution) in each room. If detailed information is required to
visualize air movement within the space, then CFD analysis should be used,
but it should not be the first analysis pursued, and it does not provide
information that supports demonstrating compliance with ASHRAE
Standard 55 (ASHRAE 2017a). CFD is typically a snapshot of a moment in
time where all temperatures and flows are fixed.
The following section on how to postprocess results is based on having the
output from a dynamic heat transfer bulk airflow model such as can be found
from an analysis in several software packages. ASHRAE and the authors do
not endorse software packages but only offer the list in Table 6.3 as a
nonexhaustive list of a variety of packages for this type of exercise as of mid-
2019.

6.4.3 Postprocessing Results from


Simulation Modeling
Dry-bulb and operative temperatures of a space, as well as airflow through
the space, are calculated by a simulation program (or manually). The
temperatures and airflow are calculated for each hour of the month. The results
from a bulk air analysis are fed into a spreadsheet for postprocessing.
The results data is sorted into workdays and hours of occupancy, then
evaluated for compliance with the comfort thresholds.
For the analyses performed for this guide, the assumed external dry-bulb
temperature used to seed the start of the prevailing mean calculation used the
first 24 hours of the outdoor air dry-bulb temperature.
The postprocessing spreadsheet separates data into the following columns:

• Workday
• Occupied hour
• Ventilation cfm per square foot
• Ventilation cfm
CHAPTER 6 | 267

• ASHRAE Standard 62.1 Threshold cfm


• ASHRAE Standard 62.1 compliance True/False
• California’s Title 24 threshold compliance
• California’s Title 24 threshold cfm
• California’s Title 24 compliance True/False
• Operative Temperature (from simulation results)
• Flat Mean 80% Lower Limit: no air speed (see Section 6.4.1,
“Acceptability Limits”)
• Flat Mean 80% Upper Limit: no air speed (see Section 6.4.1,
“Acceptability Limits”)
• Comfort Status: Too Cold/Good/Too Hot
• Flat Mean 80% Upper Limit: with air speed – 236 fpm = Flat Mean 80%
Upper Limit: no air speed (from Figure 6.2) +4°F
• Comfort Status with 236 fpm elevated air speed: Too Cold/Good/Too Hot
• Assumed External DB to Seed, then real date below
• Mean for last 24 hours
• Running Mean for last 168 hours (i.e., 7 days)
• TPMA(out): this is = (1-0.7) × mean for last 24 hours + 0.7 × Running mean
for last 168 hours (i.e., 7 days)
• Prevailing Mean 80% Lower Limit = 0.31 × TPMA(out) + 47.9
• Prevailing Mean 80% Upper Limit: No air speed = 0.31 × TPMA(out)
+ 60.5
• Comfort Status against prevailing mean: Too Cold/Good/Too Hot
• Prevailing Mean 80% Upper Limit with air speed at 236 fpm: Too Cold/
Good/Too Hot = Prevailing Mean 80% Upper Limit: No air speed + 4°F
(2°C)

6.4.4 “Flat” Mean versus Prevailing Mean to


Set Temperature Limits
Section 5.4.2.1.3 of ASHRAE Standard 55 (ASHRAE 2017a) requires that
the compliance analysis use a prevailing mean daily temperature within the
acceptability equations in 5.4.2.2 as noted in Section 6.41, “Acceptability
Limits” in this guide. According to Appendix J, “Occupant-Controlled
Naturally Conditioned Spaces,” in ASHRAE Standard 55, this requirement
“represents the broader external climatic environment to which building
occupants have become physiologically, behaviorally, and psychologically
adapted” (ASHRAE 2017a).
Older versions of ASHRAE Standard 55 allowed the user to apply the
historical monthly mean temperature to set the full month’s upper and lower
limits, essentially using the information in the exception to 5.4.2.1.3, which
can be invoked when TMY3 weather data is not available at a site. For the
purposes of this chapter, this option is known as the “flat” mean temperature
(as compared to the “prevailing” running mean temperature).
268 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 6.3 Monthly 80% upper and lower temperature limits for natural ventilation
compliance for New York using the flat mean approach.

TABLE 6.4 How the 80% Upper and Lower Limits are Calculated for
New York City, NY, using the Flat Mean Approach
New York March June September December
Mean monthly temperature 42.7°F (5.9°C) 71.3°F (21.8°C) 68.4°F (20.2°C) 38.5°F (3.6°C)

Upper 80% acceptability limit 73.7°F (23.2°C) 82.6°F (28.1°C) 81.7°F (27.6°C) 72.4°F (22.5°C)

Lower 80% acceptability limit 61.1°F (16.2°C) 70.0°F (21.1°C) 69.1°F (20.6°C) 59.8°F (15.5°C)

When a design team is testing the quick feasibility of a scheme, it may be


easier to obtain and calculate the flat mean temperatures to see if the scheme is
in the realm of gross compliance.
Using New York, NY, as an example, the flat mean temperature can be
calculated to set the full month’s 80% acceptability limits as noted in Table 6.4
and Figure 6.3.
Using the data from March, the results show the comparison between the
flat mean and prevailing running mean methods for a month’s worth of data.
CHAPTER 6 | 269

FIGURE 6.4 ASHRAE Standard 55 (ASHRAE 2017a) adaptive comfort method


comparison of indoor operative temperature to flat mean and prevailing mean criteria:
Buoyancy-driven ventilation for New York City, NY, in June, alpha = 0.7 for prevailing mean.

As noted in Figure 6.4, it is clear that the prevailing mean fluctuates in


response to the outdoor air temperatures, and it runs sometimes lower and
sometimes higher than the flat mean.
More runs for different seasons for New York City, NY, can be found in
Appendix A6.3 and for different cities in Website Appendix WA6.6,
“Engineered Natural Ventilation System Analysis Sample Results,” accessible
at ashrae.org/naturalventilation.

6.4.5 Demonstrating Compliance with the


Adaptive Comfort Method
Further postprocessing bins the flat and prevailing mean results to show
how many hours in the month meet the comfort standard—with and without
elevated air speed—for the model’s current configuration of openings. The
primary observation shown in Figures 6.5 and 6.6 is that there is less than a 5%
difference in the estimate of the hours falling into the compliance zone,
whether without elevated speed or with it. The greatest percent differential is in
the TOO HOT category, but as the total number of noncompliant hours is so
low, this represents a total of 17 hours difference in the variability of using the
flat mean instead of the prevailing mean.
Several hours remain in the TOO HOT category, which would fall into the
definition of exceedance hours. Per the definitions in Section 3 of ASHRAE
Standard 55, exceedance hours are “the number of occupied hours within a
270 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 6.5 Binned Data Showing Compliance with ASHRAE Standard 55’s
Adaptive Comfort Method using Flat and Prevailing Means
No Speed, With Speed,
New York, June TOO COLD TOO HOT
GOOD GOOD
Flat mean — 155 198 43

Prevailing mean — 138 198 60

% Difference Between — 11% 0% 39%


Flat vs Prevailing
Means

FIGURE 6.5 ASHRAE Standard 55’s flat mean natural conditioning compliance hours for
an office in New York, NY, due to buoyancy-driven ventilation.

defined time period in which the environmental conditions in an occupied


space are outside of the comfort zone” (ASHRAE 2017a). The designer can
perform iterations to increase opening sizes to comply with the Adaptive
Comfort Method for 100% of occupied hours.
Alternatively, one could discuss with the owner and the AHJ what the
acceptable number of exceedance hours is. If one looks for parallels with
mechanical comfort conditioning systems, ASHRAE Standard 90.1 (ASHRAE
2019a) Section 11.5.2i and Section G3.1.2.3 require that unmet load hours not
exceed 300 in terms of equipment sizing and performance within the energy
CHAPTER 6 | 271

FIGURE 6.6 ASHRAE Standard 55’s prevailing mean natural conditioning compliance
hours for an office in New York, NY, due to buoyancy-driven ventilation.

model showing compliance for energy efficiency. See Website Appendix


WA5.5, “Energy Modeling for Natural Ventilation per ASHRAE Standard
90.1-2019,” accessible at ashrae.org/naturalventilation, for more details.
Similarly, the Nonresidential Alternative Calculation Method Reference
Manual for the 2016 Building Efficiency Standards (CEC 2018b) states that
the heating system and cooling system must each individually have less than
150 unmet load hours (UMLH) each to comply with the energy code modeling
protocol. See Section WA5.2.4 of Website Appendix WA5.2, “Excerpts on
Natural Ventilation from California’s Title 24,” for more details.
More runs for different seasons for New York City, NY, can be found in
Appendix A6.3, and for different cities in Website Appendix WA6.6,
accessible at ashrae.org/naturalventilation.

6.4.6 Checking Simultaneous


ASHRAE Standard 62.1 Compliance
The same modeling software has an output of ventilation caused by
outdoor air entering the space. This hourly data can be plotted and evaluated
for the number of total hours that natural ventilation complies with both
ASHRAE Standard 55 and 62.1 or with both ASHRAE Standard 55 and
California’s Title 24. Figure 6.7 demonstrates the visualization of this type of
analysis. The zero-hours are hours of nonoccupancy and can thus be
272 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 6.7 Natural ventilation-created CFM versus ASHRAE Standard 62.1 and
California Title 24 thresholds in New York City, NY, in June as a result of buoyancy-driven
ventilation.

FIGURE 6.8 PPD results for New York City, NY, in June as a result of wind-driven natural
ventilation.
(a) (b) (c)

FIGURE 6.9 Operative temperature compliance (a) with the upper acceptability limit with elevated air speed for June in New
York, NY, (b) with the upper acceptability limit without elevated air speed for June in New York, NY, and (c) with the lower
CHAPTER 6 | 273

acceptability limit for June in New York.


274 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 6.10 Comparison of thermal comfort conditions to match the upper and lower
80% acceptability limits associated with Adaptive Comfort Method for June for a naturally
ventilated office in New York City, NY.
(Hoyt et al. 2017)

disregarded. There are, however, a handful of hours that are noncompliant. An


iteration to increase opening sizes would have to be completed to comply.
More runs for different seasons for New York City, NY, in Appendix A6.3,
and for different cities in Website Appendix WA6.6, “Engineered Natural
Ventilation System Analysis Sample Results,” accessible at ashrae.org/
naturalventilation.

6.4.7 Checking Adaptive Comfort Method Against


PPD Calculations from the Analytical Method
It is possible to represent the PPD results from the dynamic thermal
analysis using the typical calculations associated with the Analytical Method
but applied on an hourly basis.
In many instances, design teams will submit data to LEED based on spot-
checking the compliance of extreme conditions (i.e., highest and lowest
condition in the month or year) using the ASHRAE Thermal Comfort Tool
created by the Center for the Built Environment (CBE) at the University of
California at Berkeley (Hoyt et al. 2017). This tool can apply any of the
comfort zone methods available within ASHRAE Standard 55.
CHAPTER 6 | 275

Figure 6.9 shows the output of the tool when the operative temperature
inputs are just less than the upper 80% acceptability limits for June in New
York City, NY, using the Adaptive Comfort Method within the CBE tool.
Working from the hypothesis that human sensations are equal in a
conditioned environment and an adaptive comfort environment, the parallel
results shown in Figure 6.10 use the Analytical Method to match the upper and
lower 80% acceptability limits. This demonstrates that with a little bit of extra
clothing in the lower limit case, the Analytical Method delivers PPD
equivalent to that of the Adaptive Comfort Method at the outer edges of the
acceptability limits.

6.5 DOCUMENTATION OF COMPLIANCE


WITH THE STANDARDS
Standard 55 User’s Manual (ASHRAE 2016) provides a documentation
checklist for each of the four methods of defining the comfort zone. The
column associated with natural conditioning has been excerpted in Table 6.6.

6.6 CREDIT DOCUMENTATION FOR


LEED® THERMAL COMFORT CREDIT
The LEED® v4 rating system refers to ASHRAE Standard 55-2010 as the
basis of documentation (USGBC 2013). The only documentation for the credit
using the naturally conditioned Adaptive Method is to calculate and then plot
the average monthly outdoor temperatures and design operative temperatures
onto ASHRAE Standard 55-2010’s Figure 5.3 (which is now Figure 5.4.2 in
ASHRAE Standard 55-2017). The example given in LEED®: Reference Guide
for Building Design and Construction, v4 (USGBC 2013) is shown in
Figure 6.11 for reference. Demonstration of compliance is that the worst-case
operative temperature for each month falls within the 80% acceptability limits.

6.7 RULES OF THUMB FOR MANAGING EXPECTATIONS


AROUND COMFORT RESULTS
As can be noted in the extensive discussions above, ASHRAE Standard 55
(ASHRAE 2017a) and dynamic thermal analysis provide a means for
estimating the number of hours in a month or year where natural conditioning
can provide sufficient cooling to the space. When discussing results with the
owner or client, it is important to convey three key facts to manage
expectations when considering natural conditioning.

• On hot days, the air temperature inside is greater than the air
temperature outdoors.

The indoor space typically has heat loads created within the space or
solar heat gains entering the space. The cooler air from the outdoors is the
only air available in a pure natural conditioning scheme to remove this
276 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE 6.11 Example of documentation required for the LEED® Thermal Comfort credit
when natural conditioning is used.
(USGBC 2013)

airborne heat. The only way for an individual to perceive lower operative
temperatures in the space is to introduce other methods of heat rejection
from the skin, namely through radiative heat transfer to cooled thermal
mass surfaces or increased localized velocity for additional convective
cooling using the same indoor air temperature. As noted in the variety of
analyses above, these two phenomena can be modeled in thermal analysis
software to estimate or increase the number of hours of acceptability.
• The heat-absorbing capacity of air reduces as the outdoor air
temperature gets closer to the target indoor temperature.
This is a result of physics—absolute heat absorption capacity is
proportional to the product of airflow and temperature differential between
supply and space.

Heat absorbed = 1.08  airflow   T outside – T inside  (6.6 I-P)

where heat absorbed is in Btu/h, airflow is in cfm, and temperatures are in


°F.
Therefore, as the outdoor air temperature rises to meet the target indoor
temperature, the indoor air changes must increase to accommodate. There
are, however, reasonable limitations on how many air changes can be
achieved via natural conditioning alone. Dynamic thermal analysis can
show the air changes anticipated for the maximum available window size,
considering the driving forces of wind and stack effect. See Chapters 2 and
5 for more information.
CHAPTER 6 | 277

TABLE 6.6 Documentation Required to Demonstrate Compliance with the


Adaptive Comfort Model using Table 6-A from Standard 55 User’s Manual (2016)
Clause
within Document Notes
Section 6.2
a Method

b, c Operative temperature to, including tolerance Note A

b Outdoor design condition

b Total indoor load

b Exceedance hours

c Average air speed, Va

c Clothing insulation, Icl or clo

c Metabolic rate, met

c Excluded occupants (see Section 5.2.1.4 and 5.2.2.3)

c Excluded spaces (see Section 2.3, 4.2, and 4.3)

d Narrative describing how local thermal discomfort will not be likely to exceed limits,
including inputs, methods, and results

e System capacity that meets load for each space and/or system Note 1

f Method and equipment for occupant control of elevated air speed, Va

g Method and data for determining surface temperatures (for calculation of radiant
asymmetry and other local comfort effects)

g Method and data for determining air speed, radiant temperature asymmetry, vertical Note 2
air temperature difference, surface temperatures, temperature variation with time

h Method and data for determining mean radiant temperature


Note 1: In occupant-controlled naturally conditioned spaces, there may be no mechanical cooling system.
Note 2: Fluctuations are under the control of occupants.
Note A: Author’s note: humidity is excluded as the Adaptive Comfort path does not regulate this.

Using the relationship between the outdoor air temperature and the
upper 80% acceptability limit, Table 6.6 calculates the available heat
absorption capacity, presuming an upper limit of 10 or 20 air changes per
hour (ach) in a 10 ft (3 m) high space. The table shows the limits on
internal heat loads due to constraints imposed by the temperature
differential.
The first thing to observe from this table is that, while ASHRAE
Standard 55 (ASHRAE 2017a) will allow the application of the natural
conditioning thermal comfort criteria up to 92.3°F (33.5°C), the outdoor air’s
278 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

TABLE 6.7 Heat Absorption Capacity of Natural Conditioning Based On


Outdoor Temperature, Assuming 10 ach Limit
Temperature Heat Absorption Capacity of Natural Conditioning
80% Upper Differential Air Movement Through The Space
Outdoor
Acceptability Between 10 ach 20 ach
Temperature,
Outdoors
°F (°C) Limit, °F (°C)* and Indoor
Btu/h/ft2 W/ft2 W/m2 Btu/h/ft2 W/ft2 W/m2
Limit, °F (°C)

50 (10) 76 (24.4) 26 (14.4) 46.8 13.7 148 93.6 27.4 295

55 (12.8) 77.6 (25.3) 22.6 (12.5) 40.6 11.9 128 81.2 23.8 256

60 (15.6) 79.1 (26.1) 19.1 (10.6) 34.4 10.1 108 68.8 20.1 217

65 (18.3) 80.7 (27.0) 15.7 (8.7) 28.2 8.3 89 56.3 16.5 178

70 (21.1) 82.2 (27.8) 12.2 (6.7) 22.0 6.4 69 43.9 12.9 139

75 (23.9) 83.8 (28.7) 8.8 (4.8) 15.8 4.6 50 31.5 9.2 99

80 (26.7) 85.3 (29.6) 5.3 (2.9) 9.5 2.8 30 19.1 5.6 60

85 (29.4) 86.9 (30.4) 1.8 (1.0) 3.3 1.0 11 6.7 2.0 21

90 (32.2) 88.4 (31.3) –1.6 (–0.93) –2.9 –0.8 –9 –5.8 –1.7 –18
* Assuming outdoor air temperature equals the prevailing mean

capacity to absorb heat from the internal space is effectively exhausted by the
time the outdoor air temperature reaches 85°F (29.4°C), even with the 20 ach
scenario.
Second, in the best-case scenario with a well-designed space, one might
achieve a temperature differential in the 2°F to 5°F (1.1°C to 2.8°C) at the
upper end of the viable outdoor air temperature range for extremely low heat
load conditions (e.g., lights off, low activity levels and occupancy density,
shades pulled). Clients and occupants should be made aware of the operational
consequences as well as the fact that the indoor temperatures start to approach
the 85°F to 87°F (29°C to 30°C) range that would typically be considered
uncomfortable.

• Lowest heat-absorbing capacity occurs simultaneously to highest heat


load.

As noted in Table 6.7, the limitations on outdoor temperature versus the


temperature differential sets the absorption heat capacity. McConahey (2008)
has argued for two rules of thumb regarding control of heat gains when
considering natural conditioning:

• Is the building envelope’s solar heat gain to the space less than 4 W/ft2
(13 W/m2) across the perimeter zone floor area?
• Are the internal heat loads less than 2 W/ft2 (6.3 W/m2) of floor area?
CHAPTER 6 | 279

The reason for these rules of thumb is to ensure that there is likely to be
sufficient cooling capacity of outdoor air under extreme conditions. The
designer and client can allocate the 6 W/ft2 (19W/m2) as they wish. See
Chapter 4 for more information on calculating solar heat gain through a façade
and chapter 18, “Nonresidential Cooling and Heating Load Calculations,” in
ASHRAE Handbook—Fundamentals (ASHRAE 2017b) for guidance on
estimating internal heat loads.
In essence, a 6 W/ft2 (19 W/m2) heat absorption demand would be possible
with an outdoor temperature of 70°F (21.1°C) for 10 ach or up to 79°F
(26.1°C) if 20 ach are allowed. Where outdoor air temperatures are
systematically higher than these values, it is unlikely that the space would be
able to be naturally conditioned for 100% of the time unless severe reductions
in heat gains are implemented during peak temperature periods. These early
steps are strongly recommended:

• Perform an early assessment of outdoor air temperature trends (including


future climate change shifts).
• Compare the peak afternoon temperatures to Table 6.6 to determine
available heat absorption capacity.
• Compare available heat absorption capacity to the client’s need for internal
heat gain.
• Back calculate the allowable conduction and solar heat transmission
through the building envelope and select window-to-wall ratio and glazing
thermal properties accordingly.
• Enter the entire space design (with internal heat loads and envelope
performance) into a dynamic thermal analysis software program to
determine the window size and height of space that is necessary to ensure
that the required air change rate is achieved under the peak outdoor
temperature condition.

6.8 COLD DAY CONCERNS IN NATURAL


CONDITIONING SCHEMES
Most analyses of acceptable comfort levels using the adaptive comfort
model tend to concern themselves with compliance with the 80% upper
acceptability limit in ASHRAE Standard 55 (ASHRAE 2017a). This is
because the adaptive comfort standard prohibits mechanical cooling in
Section 5.4.1 of the standard but only defines that heating may not be in
operation when one is under natural conditioning, and the upper limit condition
represents the primary point of failure in the system.
It is possible that the space will not remain above the lower acceptability
limit at the time, and thus, it is possible to have and employ heating when the
occupant-controlled natural conditioning mode is not in effect.
Assuming a scenario where natural conditioning is used during the
nonheating season, it is likely that the opening sizes also have to be sized to
meet or exceed the natural ventilation openings sizes as governed under
ASHRAE Standard 62.1 (ASHRAE 2019b). One can also take for granted that
280 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

the operable elements remain available to the occupants for use as natural
ventilation openings during cold periods. The challenge is one of controlling
opening size during the heating season to allow the user to open the window to
meet the desired minimum outdoor air quantity without causing excessive loss
of heating from the space.
Heating elements near the façade are often recommended to offset building
envelope heat loss through the closed façade, especially through glazed
elements. In the case of a natural ventilation system, the sizing of these heating
elements must also accommodate for heat loss from unwanted infiltration and
the outside air quantities coming through the natural ventilation openings.
See Chapter 4 for a discussion of estimating heat loss, sizing heating
systems, understanding radiant impacts on comfort, and condensation risks
that are associated with significant amounts of glazing.

6.9 REFERENCES
Alliance for Sustainable Energy. n.d. Open Studio®. https://
www.openstudio.net.
ASHRAE. 2016. Standard 55 user’s manual based on ANSI/ASHRAE
Standard 55-2013 thermal environmental conditions for human
occupancy. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2017a. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2017b. ASHRAE handbook—Fundamentals. Peachtree Corners,
GA: ASHRAE.
ASHRAE. 2019a. ANSI/ASHRAE Standard 90.1-2019, Energy standard for
buildings except low-rise residential buildings. Peachtree Corners, GA:
ASHRAE.
ASHRAE. 2019b. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
CEC. 2018a. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
CEC. 2018b. 2019 Nonresidential Compliance Manual for the 2019 Building
Energy Efficiency Standards: Title 24, Part 6 and Associated
Administrative Regulations in Part 1. CEC-400-2018-018-CMF.
Sacramento, CA: California Energy Commission.
CBSC. 2019. 2019 California mechanical code California code of regulations,
title 24, part 4, California building standards commission, Based on 2018
Uniform Mechanical Code®. Sacramento, CA: California Building
Standards Commission and Ontario, CA: International Association of
Plumbing and Mechanical Officials.
DesignBuilder Software Ltd. n.d. DesignBuilder software packages. https://
designbuilder.co.uk/software/product-overview.
CHAPTER 6 | 281

Digital Alchemy. n.d. Simergy™. Woodinville, WA: Digital Alchemy, Inc.


https://d-alchemy.com/products/simergy.
DOE. n.d. EnergyPlus™. Washington, D.C.: U.S Department of Energy.
Hoyt, T., S. Schiavon, A. Piccioli, T. Cheung, D. Moon, and K. Steinfeld.
2017. CBE thermal comfort tool. Berkeley, CA: Center for the Built
Environment, University of California Berkeley. http://
comfort.cbe.berkeley.edu/.
ICC. 2017. 2018 International mechanical code®. Country Club Hills, IL:
International Code Council, Inc.
IES. n.d. IES Virtual Environment (IESVE). Glasgow, U.K.: Integrated
Environmental Solutions Ltd. https://www.iesve.com/software.
McConahey, E. 2008. Finding the right mix. ASHRAE Journal 50(9): 36–48.
NIST. n.d. CONTAM. Gaithersburg, MD: National Institute of Standards and
Technology. https://www.nist.gov/services-resources/software/contam.
Oasys Software. n.d. BEANS v.16.2 suite of software, specifically BEANS
v.16.2 and ROOM v.16.2 modules. UNIPAC software for students and
institutions. https://www.oasys-software.com/education/. London: Arup.
Trimble, Inc. n.d. Sefaira. https://sefaira.com/.
UWM. n.d. TRNSYS. Madison, WI: The University of Wisconsin, Madison.
https://sel.me.wisc.edu/trnsys/.
USGBC. 2013. LEED® reference guide for building design and construction,
LEED version v4. Washington, DC: U.S. Green Building Council.
SAMPLE RESULTS FROM
NATURAL CONDITIONING
ANALYSIS USING
DESIGNBUILDER
This appendix presents the results of the analysis of the office and
classroom spaces as described in Appendix A1.1 and A1.2 using
DesignBuilder (DesignBuilder n.d.), an integrated software suite that includes
load analysis, energy modeling, dynamic heat transfer, and bulk airflow
modeling, and computational fluid dynamics (CFD).
An office and classroom were modeled in DesignBuilder to include details
of the thermal envelope, ventilation requirements in accordance with ANSI/
ASHRAE Standard 62.1 (2019), and internal heat gains from lighting,
occupants, and plug loads. The internal gains and ventilation requirements
varied according to the office and classroom activity types, but the fabric and
glazing performance was the same for both models. The location and TMY3
*.epw hourly weather file used in both cases were for New York City’s Central
Park (latitude 40.8 N).

A6.1.1 OFFICE SPACE


The small office space, with a large fixed window and a high-level
operable window, is shown in Figure A6.1.1. The space is identical to the
space defined in Appendix A1.1.

A6.1.2 CLASSROOM SPACE


The classroom space, with a large central fixed window with operable
windows above and below, is shown in Figures A6.1.4. The model itself can be
interrogated to visualize the bulk airflow movements that drive natural
ventilation.
The software calculates the volume of airflow moving in and out of each
opening. The software can also represent hourly or daily aggregated results to
check compliance with the ventilation standard. The CFD module can be used
to calculate Fanger PMV and PPD comfort metrics (Fanger 1982) using time-
step data imported from the EnergyPlus dynamic simulation (DOE n.d.)

283
284 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A6.1.1 Office space with two FIGURE A6.1.2 Quality assurance
occupants and associated information check on natural ventilation airflows at a
technology (IT). given time step for a short height two-way
opening.

FIGURE A6.1.3 Hourly results with mechanical ventilation (shown in red) with additional
natural ventilation (shown in blue) “free cooling” utilizing concurrent mixed-mode control.
APPENDIX A6.1 | 285

FIGURE A6.1.4 Classroom space model with 31 occupants and associated equipment.

FIGURE A6.1.5 Quality assurance check on natural ventilation airflows at a given time
step for two independent opening separated by a height.

FIGURE A6.1.6 Airflow in and out of individual openings at selected simulation periods.
286 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A6.1.7 Hourly zone infiltration air change rate via windows only.

FIGURE A6.1.8 Daily frequency zone infiltration air change rate via windows only.
APPENDIX A6.1 | 287

FIGURE A6.1.9 PMV/PPD results from the CFD analysis, demonstrating greater
discomfort near the glazed surfaces but relatively neutral within the depth of the room.

FIGURE A6.1.10 PMV/PPD output from the results of the CFD at the occupied zone.
288 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A6.1.3 REFERENCES
ASHRAE. 2019. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
DesignBuilder. n.d. DesignBuilder, Version V6.1.5.2. Stroud, U.K.:
DesignBuilder Software Ltd.
DOE. n.d. EnergyPlus, Version 8.9. Washington, D.C.: U.S. Department of
Energy’s Building Technologies Office.
Fanger, P.O. 1982. Thermal comfort: Analysis and applications in
environmental engineering. Malabar, FL: Robert E. Krieger Publishing
Co., Inc.
SAMPLE RESULTS FROM
NATURAL CONDITIONING
ANALYSIS USING IESVE
This appendix presents sample outputs and results of the analysis of the
office and classroom spaces as described in Appendix A1.1 and A1.2 under the
use of the software IESVE (IES n.d.), an integrated suite of software that
includes load analysis, energy modeling, dynamic heat transfer and bulk
airflow modeling, and CFD.
An office and classroom example was modeled in IESVE to include details
of the thermal envelope, ventilation requirements in accordance with
ASHRAE Standard 62.1 (ASHRAE 2019), and internal heat gains from
lighting, occupants, and plug loads. The internal gains and ventilation
requirements varied according to the office and classroom activity types, but
the fabric and glazing performance was the same for both models. The location
and TMY3.epw hourly weather file used in both cases were for New York
City’s Central Park (located at latitude 40.8 N).

FIGURE A6.2.1 Office space model.

289
290 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A6.2.2 Office model input screen where the user sets minimum ventilation
requirements as appropriate to the local compliance standard.

FIGURE A6.2.3 Quality assurance check on natural ventilation airflows at a given time-
step for a short-height, two-way opening and the air temperatures of the incoming and leaving
flows.

A6.2.1 OFFICE SPACE


The small office space, shown in Figure A6.2.1, has a large fixed window
and a high-level operable window.
The office is controlled to be naturally ventilated under the following
conditions:
• Room air temperature is greater than 72°F (22.2°C), and
• Outdoor air temperature is greater than 56°F (13.3°C), and
• Outdoor air temperature is less than room air temperature by 2°F (1°C).
APPENDIX A6.2 | 291

FIGURE A6.2.4 Office hourly results for allowable natural ventilation hours, with a
regression trend.

FIGURE A6.2.5 Graphical representation of the CFD results for the office.
292 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A6.2.6 Rendering of the classroom model.

FIGURE A6.2.7 Classroom model input screen where the user can set minimum
ventilation requirement as appropriate to the local compliance standard.

The office air change rate from operable windows is shown in


Figure A6.2.4 with a regression trend.
IESVE can also display natural ventilation microscopic airflow with a high
fenestration opening on the left. Particle tracking at the opening demonstrates a
natural stack effect. The CFD ranges for the classroom are showing:

• Air velocity: 0 to 1.10 ft/s (0 to 0.33 m/s)


• Air temperature: 69°F to 80°F (20.6°C to 26.7°C)
• Percentage people dissatisfied (PPD): 0 to 16.5 PPD
• Carbon dioxide (CO2) concentration: 0 to 1100 ppm
APPENDIX A6.2 | 293

FIGURE A6.2.8 Quality assurance check on natural ventilation airflows at a given time-
step for two windows separated by height and the air temperatures of the incoming and leaving
flows.

FIGURE A6.2.9 Classroom hourly results for allowable natural ventilation hours, with a
regression trend.
294 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A6.2.10 Graphical representation of the CFD results for the classroom with
high- and low-level openings.

A6.2.2 CLASSROOM SPACE


The classroom space with a large central fixed window with operable
windows above and below is shown in Figure A6.2.6.
The model itself can be interrogated to visualize the bulk airflow
movements that drive natural ventilation.
The classroom is set up to be naturally ventilated under the following
conditions:

• Room air temperature is greater than 72°F (22.2°C),


• Outdoor air temperature is greater than 56°F (13.3°C), and
• Outdoor air temperature is less than room air temperature by 2°F (1°C).

The classroom air change rate from operable windows is shown in


Figure A6.2.9 with a regression trend. Note the summertime drop in natural
ventilation is, in part, due to the school calendar.
The software can also display natural ventilation microscopic airflow with
high and low fenestration openings on the left. Particle tracking at the low
opening demonstrates a natural stack effect. The CFD ranges for the classroom
show:

• Air velocity: 0 to 1.10 ft/s (0 to 0.33 m/s)


• Air temperature: 69°F to 80°F (20.6°C to 26.7°C)
• Percentage people dissatisfied (PPD): 0 to 16.5 PPD
APPENDIX A6.2 | 295

FIGURE A6.2.11 PPD for the classroom (red) and office (blue) for ASHRAE Standard 55
compliance checks.

A6.2.3 COMPARATIVE COMFORT


IESVE can also estimate comfort for the occupants for different simulation
runs. See Figure A6.2.11.

A6.2.4 REFERENCES
ASHRAE. 2019. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
IES. n.d. IES Virtual Environment (IESVE). Glasgow, U.K.: Integrated
Environmental Solutions Ltd. https://www.iesve.com/software.
ENGINEERED
NATURAL VENTILATION SYSTEM
ANALYSIS SAMPLE RESULTS
This appendix presents the results of the analysis of the office space
described in Appendix A1.1 using the Oasys BEANS software (Oasys n.d.),
which is a dynamic heat transfer and bulk airflow modeling tool, for New York
City, NY. Results for other cities can be found in Website Appendix WA6.6,
accessible at ashrae.org/naturalventilation.

A6.3.1 NEW YORK CITY, NY, RESULTS

FIGURE A6.3.1 Monthly 80% upper and lower temperature limits for natural ventilation
compliance for New York City, NY, using the flat mean approach.

297
298 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

A6.3.1.1 Buoyancy-Driven Ventilation

FIGURE A6.3.2 Seasonal natural conditioning results for New York City, NY, due to
buoyancy-driven ventilation using a flat mean approach.

FIGURE A6.3.3 Seasonal natural conditioning results for New York City, NY, due to
buoyancy-driven ventilation using a prevailing mean approach.
APPENDIX A6.3 | 299

FIGURE A6.3.4 ASHRAE Standard 55 Adaptive Comfort comparison of indoor operative


temperature to flat mean and prevailing mean criteria: Buoyancy-driven natural ventilation for
New York City, NY, in March, alpha = 0.7 for prevailing mean.

FIGURE A6.3.5 ASHRAE Standard 55 Adaptive Comfort comparison of indoor operative


temperature to flat mean and prevailing mean criteria: Buoyancy-driven natural ventilation for
New York City, NY, in June, alpha = 0.7 for prevailing mean.
300 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A6.3.6 ASHRAE Standard 55 Adaptive Comfort comparison of indoor operative


temperature to flat mean and prevailing mean criteria: Buoyancy-driven natural ventilation for
New York City, NY, in September, alpha = 0.7 for prevailing mean.

FIGURE A6.3.7 ASHRAE Standard 55 Adaptive Comfort comparison of indoor operative


temperature to flat mean and prevailing mean criteria: Buoyancy-driven natural ventilation for
New York City, NY, in December, alpha = 0.7 for prevailing mean.
APPENDIX A6.3 | 301

A6.3.1.2 Wind-Driven Results

FIGURE A6.3.8 Seasonal natural conditioning results for New York City, NY, due to wind-
driven ventilation using a flat mean approach.

FIGURE A6.3.9 Seasonal natural conditioning results for New York City, NY, due to wind-
driven ventilation using a flat mean approach.
302 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE A6.3.10 ASHRAE Standard 55 Adaptive Comfort comparison of indoor operative


temperature to flat mean and prevailing mean criteria: Wind-driven ventilation in New York City,
NY, in March, alpha = 0.7 for prevailing mean.

FIGURE A6.3.11 ASHRAE Standard 55 Adaptive Comfort comparison of indoor operative


temperature to flat mean and prevailing mean criteria: Wind-driven ventilation in New York City,
NY, in June, alpha = 0.7 for prevailing mean.
APPENDIX A6.3 | 303

FIGURE A6.3.12 ASHRAE Standard 55 Adaptive Comfort comparison of indoor


operative temperature to flat mean and prevailing mean criteria: Wind-driven ventilation in
New York City, NY, in September, alpha = 0.7 for prevailing mean.

FIGURE A6.3.13 ASHRAE Standard 55 Adaptive Comfort comparison of indoor


operative temperature to flat mean and prevailing mean criteria: Wind-driven natural
ventilation in New York City, NY, December, alpha = 0.7 for prevailing mean.
CASE STUDY—
WILSHIRE-GAYLEY,
LOS ANGELES, CA

FIGURE CS4.1 Rendering of Wilshire-Gayley.

• Architect: Robert A.M. Stern Architects


• Project: Wilshire-Gayley, Los Angeles

The project is located at the intersection of Wilshire Boulevard and Gayley


Avenue in Los Angeles, CA. The project will consist of four subterranean
parking levels, four levels of mixed-use spaces, and 26 levels of residential
apartments totaling approximately 274,561 ft2 (27,456 m2).

305
306 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

FIGURE CS4.2 ASHRAE Standard 55 adaptive comfort comparison of indoor operative


temperature to flat mean and prevailing mean criteria: Wilshire-Gayley, January, alpha = 0.7
for prevailing mean.

FIGURE CS4.3 ASHRAE Standard 55 adaptive comfort comparison of indoor operative


temperature to flat mean and prevailing mean criteria: Wilshire-Gayley, September, alpha =
0.7 for prevailing mean.
CONTROLLING
NATURAL VENTILATION AND
NATURAL CONDITIONING SYSTEMS
7.1 OVERVIEW OF BUILDING MANAGEMENT CONTROLS FOR
NATURAL VENTILATION SYSTEMS
Both natural ventilation and natural conditioning systems require that the
occupant have control over the opening used to bring air into the building. The
simplest level of control, therefore, is merely manual control of the window.
Section 6.4.3 of ANSI/ASHRAE Standard 62.1 states that openings must be
“readily accessible” and that controls should “be designed to coordinate
operation of the natural and mechanical ventilation systems” (ASHRAE 2019).
Section 5.4, “Determining Acceptable Thermal Conditions in Occupant-
Controlled Naturally Conditioned Spaces” in ANSI/ASHRAE Standard 55
requires the elements to be “occupant controlled” with the definition including
“manually controlled or controlled through the use of electrical or mechanical
actuators under direct occupant control” (ASHRAE 2017).
When natural ventilation or natural conditioning is proposed, the client and
users might expect that the only means for obtaining cooling is from opening
and closing the windows, and adaptation is prevalent. This expectation is
challenged when a mixed-mode/hybrid ventilation scheme is applied, as
mechanical cooling is known to be available. Additional automated controls
are often required to accomplish these tasks. An excellent resource regarding
control of mixed-mode spaces, including case study examples, can be found in
Brager et al. (2007).
As natural ventilation and conditioning systems have come to be applied in
multioccupant, nonresidential spaces, there are an increasingly complex
number of control approaches available, which are designed to support energy-
efficiency measures, automated control of indoor temperature for comfort
conditioning, and concerns over indoor air quality.
This section describes several building automation control sequences for
natural ventilation and natural conditioning, including changeover approaches
to heating or cooling in a mixed-mode system and the typical sensors and
devices used to accomplish the desired performance. The descriptions are

307
308 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

necessarily generic and laid out for a single-zone approach with direct
occupant control. Multizone approaches that are integrated into multifloor uses
of atria are beyond the scope of this chapter, as these types of buildings have
complex indoor airflow patterns unique to the shape and internal loads of the
building.
The chapter ends with a discussion of considerations to make when
selecting window actuator hardware.

7.2 SAMPLING OF TYPICAL CONTROL


SEQUENCES OF OPERATION
Under all circumstances where more sophisticated control systems are used
to manipulate opening sizes, it is important to ensure that the control sequences
achieve their goals. In the case of natural ventilation, the goal is that verifiable
minimum indoor ventilation rates meet indoor air quality requirements
equivalent to ASHRAE Standard 62.1 (ASHRAE 2019). In the case of natural
conditioning, the former requirement remains, and the additional requirement
to achieve the agreed-upon comfort temperature requirements also applies.
Table 7.1 shows the most prevalent types of control sequences used and the
required or recommended sensors and hardware needed to accomplish the task.
The approach for this chapter is to provide a “toolbox” in which each sequence
is written at its most elemental level with the expectation that design teams will
understand the functional performance without confusion caused by conflating
too many interrelated sequences. It is possible to envision that more than one
sequence may exist for a given space, to represent differing control approaches
for different devices or for operating at different times.
Sensor feedback is necessary for the control of more complex systems.
Outdoor air temperature and relative humidity are measured values that are
typically meaningful in controlling changeover to mixed-mode systems. In
naturally ventilated spaces during colder weather, outdoor air temperature is
typically used to preemptively close motorized windows to their minimum
position or to keep them closed before occupants arrive. It is used primarily as
an override to the feedback coming from the indoor air temperature sensors in
purely naturally ventilated spaces where openings must leave control with the
occupants.
Indoor air temperature measurements are typically used as the primary
control input variable for temperature control. Carbon dioxide monitoring is
typically used as the primary control input variable for minimum ventilation
unless the prescriptive path of ASHRAE Standard 62.1 (ASHRAE 2019) or
California’s Title 24 (CEC 2018) is used.
External wind speed should influence the control strategy primarily as a
threshold for shutting the windows to maintain safety and reduce disturbance
from rain. Wind direction is a useful parameter for wind-driven ventilation
designs; however, the sensors tend to be overly sensitive, and thus the signals
must be postprocessed to slow down the response at the window actuator to
minimize disruption to the occupants and to improve actuator life. Rainfall
CHAPTER 7 | 309

TABLE 7.1 Recommended Sensors to Facilitate Control Sequences for Natural Ventilation,
Natural Conditioning, and Mixed-Mode/Hybrid Ventilation Systems
Outdoor Sensors Active Indoor Sensors Active Device Active
in SOO in SOO in SOO

Outdoor Air Bulb Temperature

Favorable Condition Indicator

Motorized Opening Actuator


Rain/Precipitation Presence

Indoor Relative Humidity

Zonal Ventilation Supply


Indoor Air Temperature

Window/Louver Closed
Outdoor Air Pollutants

Zonal Heating Device

Zonal Cooling Device


Wind Speed Outdoor

Surface Temperature
Control Sequence of Operations
Type

Relative Humidity

Indoor CO2 Level

& Position Status


Position Status
Wind Direction

Manual Action
Static Outdoor

Thermal Mass
Wind Speed

Air Pressure
Heating Natural Ventilation

Manual 1 2 X B

Manual button/switch for 1 2 X X B


motorized window

Automatic window E H E e H 1 2 XH B

Cold hours heating concurrent X X 1 T T X T

Cold hours heating changeover X X 1 X B

Manual X 1 2 X B
Natural Conditioning

Manual button/switch for X 1 2 X X B


motorized window

Manual with favorable condition X X 1 2 3 X B


indicator

Automatic window with X E H E e H X 1 2 XH B


temperature control
Night Purge Cooling

Natural conditioning with X 4 X 4 1 2 3 X B


hot hours cooling changeover

Automatic window for night time X X E H E e H X X 5 XH 5


purge of heat from thermal mass

X = sensor or action is participating in the SOOs


B = backup mechanical ventilation to be triggered if windows are not opened during occupied hours (plus alarm raised) per California’s Title 24 (CEC
2018). Backup mechanical ventilation indoor CO2 levels are not achieved
E = used as emergency closure override trigger to stay within window actuator strength, preserve indoor viable environment to protect occupants from
turbulence, dust, disturbance of papers, and wetting of indoor surfaces
e = optional emergency closure override trigger for air quality: sensor real-time capture, accuracy, and expense to be reviewed
H = special monitoring of wind direction and façade pressures for high-rise buildings to inform automated sequences for modulation of window open-
ings in response to strength of wind
T = Trickle ventilation either via permanently open trickle ventilators, infiltration, or manual windows left unseated by occupants
XH = special motorized window override for zonal smoke control if actuator is in a high-rise building
1 = indoor CO2 monitoring is required for LEED® credit compliance (USGBC 2013) or as required by California’s Title 24 (CEC 2018)
2 = window, louver, or door frame switch confirms “closed” status. Switches off indoor cooling per California’s Title 24 (CEC 2018) if not closed, but
back up ventilation must remain active. In LEED® (USGBC 2013), “closed” status triggers backup ventilation during occupied hours.
3 = optional red light/green light indicator (LED light) to act as an indicator of favorable outside air conditions so that occupants know when they
should open their window
4 = outdoor and indoor air relative humidity to be tracked to inform changeover conditions if chilled surfaces or chilled beams are the zonal cooling
method. This is to ensure that changeover occurs at a low enough dew point for the cooling system to avoid condensation.
5 = ideal window configuration for night purge cooling is high-level openings only, so that air passes across ceiling level thermal mass
310 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

measurements are typically used to determine that it is raining at a measurable


level; there is no modulation of response that is typically related to rainfall
with regard to controlling natural ventilation elements.
Appendix A7.1 includes a more detailed description of the sensor types and
their typical applications.
The devices listed in Table 7.1 are representative of what can be typically
found as being activated through sequences of operations. The reference to
zonal heating, cooling, or ventilation equipment in the far right three columns
are not specifying a particular type of equipment but only when they should be
triggered during a natural ventilation sequence. The design of such systems
and their operation is left to the design team based on traditional approaches to
thermal control.

7.3 NATURAL VENTILATION SEQUENCES OF OPERATION

7.3.1 Typical Manual Control and Manual Control with Button


Natural Ventilation Sequence of Operation
For the two LEED® “Minimum Air Quality Performance” prerequisite
options that do not require an exhaust fan, either of the following can be used
(USGBC 2013):

1. Window position status shall be monitored continuously during occupied


hours. When a window is closed, the zone shall go into alarm for ventila-
tion and the backup ventilation system shall be started.
2. Indoor CO2 shall be monitored continuously during occupied hours.
When the CO2 level rises higher than ambient outdoor CO2 plus either
an engineer-determined set point per ASHRAE Standard 62.1
(ASHRAE 2019) or 600 ppm per California’s Title 24 (CEC 2018), the
zone shall go into alarm for ventilation and the backup ventilation sys-
tem shall be started.

If a motorized device is used, then it is possible to use device or actuator


position feedback or end switches to meet the window position status
requirement. Note that buttons for motorized windows should be readily
accessible by the occupants within the directly ventilated space to comply with
the ventilation standards and codes.
If security is of concern and there is personnel capable of responding by
walking to the alarm zone to close windows, the following sequence may also
be of interest:

• After 30 minutes past the end of occupancy hours, window positions


for all manual windows shall alarm if the switches do not indicate a
“closed” position.
CHAPTER 7 | 311

7.3.2 Typical Automatic Window Natural Ventilation


Sequence of Operation
A typical automatic window natural ventilation sequence of operation
(SOO) is as follows:

1. At the start of occupancy hours, the automatic window shall be opened.


2. If a window should be open and it is not, then an alarm shall be raised and
the backup ventilation system shall be started.
3. After 30 minutes past the end of occupancy hours, the automatic window
shall be closed.
4. If a window should be closed and it is not, then an alarm shall be raised.
5. In the event of wind speeds exceeding an amount to be determined through
trial and error or design team wind modeling, close the windows.
6. In the event of wind speeds exceeding an amount to be determined through
trial and error or design team wind modeling and rain is sensed, close the
windows.
7. Optional: If outdoor PM2.5 pollutants rise above an amount to be set by the
owner based on sensor quality, close the windows.

For the two LEED® “Minimum Air Quality Performance” prerequisite


options that do not require an exhaust fan, either of the following can be used
(USGBC 2013):

1. Window position status shall be monitored continuously during occupied


hours. When a window is closed, the zone shall go into alarm for ventila-
tion and the backup ventilation system shall be started.
2. Indoor CO2 shall be monitored continuously during occupied hours.
When the CO2 level rises higher than ambient outdoor CO2 plus either
an engineer-determined set point per ASHRAE Standard 62.1
(ASHRAE 2019) or 600 ppm per California’s Title 24 (CEC 2018), the
zone shall go into alarm for ventilation and the backup ventilation sys-
tem shall be started.

If a motorized device is used, then it is possible to use device or actuator


position feedback or end switches to meet the window position status
requirement.
Note that the full automation of windows to the exclusion of manually
controlled windows would make a space ineligible for application of the
natural conditioning comfort range detailed in ASHRAE Standard 55
(ASHRAE 2017).
For high-rise buildings in which wind speeds are substantially increased
and may cause stress on the actuator motors and/or require openings to have
modulation more granular than just open/closed, it is sometimes necessary to
apply additional protective sequences based on advice from the device
manufacturer, such as:
312 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

1. When roof wind speeds rise above an amount to be determined through


trial and error or design team wind modeling and from directions to be
determined through trial and error or design team wind modeling, close the
windows.
2. When roof wind speeds rise above an amount to be determined through
trial and error or design team wind modeling and from directions to be
determined through trial and error or design team wind modeling, reduce
the range of opening to a high limit of 50% of full stroke.
3. When differential static pressure at a given elevation between opposing
faces of a building exceeds an amount to be determined through trial and
error or design team wind modeling, reduce the range of opening to a high
limit of 50% of full stroke.

7.4 NATURAL VENTILATION HEATING MODE


ENABLED/DISABLED SEQUENCES OF OPERATION

7.4.1 Typical Concurrent Heating Sequence of Operation


The following sequence is applicable if a perimeter finned tube convector
or radiator system is used.
1. When the indoor temperature falls below 68°F (20°C), activate the heating
element.
2. If automatic windows are present, close the windows to the minimum
“trickle” mode.
3. If dedicated trickle vents in the zone are available, close all automatic win-
dows and open the trickle vents.
4. If neither trickle vents nor trickle mode for automatic windows is available,
start the backup mechanical ventilation system.
5. (Typical): Modulate zone heating to hold a set point of 70.5°F (21°C) with
a 5°F (2.5°C) deadband.
6. (Specific to California): In California’s Title 24 (CEC 2018), Item #5 must
be overridden and placing the zonal control in a setback of 55°F (13°C)
with a 5°F (2.5°C) deadband instead is required whenever a window switch
is not closed. See Section WA5.2.3 of Website Appendix WA5.2, “Excerpts
on Natural Ventilation from California’s Title 24,” accessible at ashrae.org/
naturalventilation.

Leaving a heating mode is contingent on the warmth of the outdoor air


conditions:
1. Heating mode shall remain operational until outdoor air temperature
exceeds 68°F (20°C).
2. When the heating mode is terminated, revert systems to normal natural
ventilation operation.

For the LEED® “Minimum Air Quality Performance” prerequisite, this


option is likely most appropriate (USGBC 2013):
CHAPTER 7 | 313

• Indoor CO2 shall be monitored continuously during occupied hours.


When the CO2 level rises higher than ambient outdoor CO2 plus either
an engineer-determined set point per ASHRAE Standard 62.1
(ASHRAE 2019) or 600 ppm per California’s Title 24 (CEC 2018), the
zone shall go into alarm for ventilation and the backup ventilation
system shall be started.

7.4.2 Typical Changeover Heating Sequence of Operation


If an air-based heating system is used, it is presumed that this also is
coming from the mechanical backup ventilation system, in which case:
1. When the indoor temperature falls below 68°F (20°C), activate the heating
and backup ventilation system.
2. Modulate zone heating to hold a set point of 70.5°F (21°C) with a 5°F
(2.5°C) deadband.

Leaving a heating mode is contingent on the warmth of the outdoor air


conditions:
1. Heating mode shall remain operational until outdoor air temperature
exceeds 68°F (20°C).
2. When the heating mode is terminated, revert systems to normal natural
ventilation operation.

It is presumed that the monitoring of CO2 sensors during the backup


ventilation system is controlled by standard demand control ventilation
sequences, and this is not covered here as it is not related to natural ventilation.

7.5 NATURAL CONDITIONING SEQUENCES OF OPERATION

7.5.1 Typical Manual Control and Manual Control with


Button Natural Conditioning Sequence of Operation
The basic temperature monitoring sequence is as follows:
1. When the indoor temperature falls below 68°F (20°C), activate the zonal
heating mode (see item 1 in Sections 7.4.1 and 7.4.2 based on mixed-mode
heating type).
2. When the heating mode is terminated, revert systems to normal natural
ventilation mode.
3. Indoor temperature shall be monitored and an alarm raised when the indoor
temperature rises above the upper 80% acceptability limit, which is calcu-
lated, per ASHRAE Standard 55 (ASHRAE 2017) as

Upper 80% acceptability limit, F = 0.31 t pma(out) + 60.5

Upper 80% acceptability limit, C = 0.31 t pma(out) + 21.3


314 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

where tpma(out) is the simple arithmetic mean of all of the mean daily
outdoor air temperatures of no fewer than the last seven days and no
greater than the last 30 days, and the mean daily outdoor air temperature of
each day is the simple arithmetic mean of all the outdoor dry-bulb
temperature observations for a 24-hour period. See ASHRAE Standard 55
(ASHRAE 2017) for more details.
4. Indoor air temperature shall be monitored. The number of hours exceeding
the upper 80% acceptability limit shall be logged and stored per zone for at
least 365 days.

For the two LEED® “Minimum Air Quality Performance” prerequisite


options that do not require an exhaust fan, either of the following can be used
(USGBC 2013):

1. Window position status shall be monitored continuously during occupied


hours. When a window is closed, the zone shall go into alarm for ventila-
tion and the backup ventilation system shall be started.
2. Indoor CO2 shall be monitored continuously during occupied hours.
When the CO2 level rises higher than ambient outdoor CO2 plus either an
engineer-determined set point per ASHRAE Standard 62.1 (ASHRAE
2019) or 600 ppm per California’s Title 24 (CEC 2018), the zone shall go
into alarm for ventilation and the backup ventilation system shall be
started.

If a motorized device is used, then it is possible to use device or actuator


position feedback or end switches to meet the window position status
requirement. Note that buttons for motorized windows should be readily
accessible to the occupants within the directly ventilated space to comply with
the ventilation standards and codes.
If security is of concern and there is personnel capable of responding by
walking to the alarm zone to close windows, the following sequence may also
be of interest:

• After 30 minutes past the end of occupancy hours, window positions


for all manual windows shall alarm if the switches do not indicate a
“closed” position.

Within California’s Title 24 (CEC 2018), there is an energy-efficiency


requirement for concurrent heating:

7.5.2 Typical Manual Control with Favorable Condition Indicator


Natural Conditioning Sequence of Operation
All of the sequences noted in Section 7.5.1 apply, plus the following:

• The favorable conditions light shall shine when outdoor air conditions
are above 68°F (20°C) and below the upper 80% acceptability limit
CHAPTER 7 | 315

minus 3°F (1.7°C) to be determined by the design team based on


internal heat loads and anticipated solar gain and/or trial and error.

7.6 MIXED-MODE CHANGEOVER SEQUENCE OF OPERATION

7.6.1 Typical Automatic Window Operation with Temperature


Control Natural Conditioning Sequence of Operation
The basic temperature control sequence is as follows:
1. At the start of occupancy hours, the automatic window shall be opened.
• If a window should be open and it is not, then an alarm shall be
raised and the backup ventilation system shall be started.
2. At 30 minutes past the end of occupancy hours, the automatic window shall
be closed.
• If a window should be closed and it is not, then an alarm shall be
raised.
3. In the event of wind speeds exceeding an amount to be determined through
trial and error or design team wind modeling, close the windows.
4. In the event of wind speeds exceeding an amount to be determined through
trial and error or design team wind modeling and rain is sensed, close the
windows.
5. (Optional): If outdoor PM2.5 pollutants rise above an amount to be set by
the owner based on sensor quality, close the windows.
6. When the indoor temperature falls below 68°F (20°C), activate the zonal
heating mode (see item 1 in Sections 7.4.1 and 7.4.2 based on mixed-mode
heating type).
7. When the heating mode is terminated, revert systems to normal natural
conditioning mode.
8. A lower 80% acceptability limit shall be calculated, per ASHRAE
Standard 55 (ASHRAE 2017), as

lower 80% acceptability limit, F = 0.31 t pma(out) + 47.9

lower 80% acceptability limit, C = 0.31 t pma(out) + 14.3


where tpma(out) is the simple arithmetic mean of all of the mean daily
outdoor air temperatures of no fewer than the last seven days and no
greater than the last 30 days, and the mean daily outdoor air temperature of
each day is the simple arithmetic mean of all the outdoor dry-bulb
temperature observations for a 24-hour period. See ASHRAE Standard 55
(ASHRAE 2017) for more details.
9. If the room temperature is higher than the lower acceptability limit, allow
natural conditioning mode to start, otherwise, return to heating mode.
10. The upper 80% acceptability limit shall be calculated, per ASHRAE Stan-
dard 55 (ASHRAE 2017) as
316 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

Upper 80% acceptability limit, F = 0.31 t pma(out) + 60.5

Upper 80% acceptability limit, C = 0.31 t pma(out) + 21.3

11. If the space is in natural conditioning mode, modulate automatic windows


to control the room temperature to a set point of 0.5 × (lower 80% accept-
ability limit + upper 80% acceptability limit) with a deadband of 5°F
(2.5°C).
12. If the upper end of the deadband is exceeded for more than 30 minutes,
then the automatic windows shall operate to hold room temperature set
point at upper 80% acceptability limit minus 2.5°F (1.3°C) (i.e., half a
deadband).
13. Indoor air temperature shall be monitored. The number of hours exceeding
the upper 80% acceptability limit shall be logged and stored per zone for at
least 365 days. The number of hours exceeding the lower 80% acceptability
limit shall be logged and stored per zone for at least 365 days.

For the two LEED® “Minimum Air Quality Performance” prerequisite


options that do not require an exhaust fan, either of the following can be used
(USGBC 2013):
1. Window position status shall be monitored continuously during occupied
hours. When a window is closed, the zone shall go into alarm for ventila-
tion and the backup ventilation system shall be started.
2. Indoor CO2 shall be monitored continuously during occupied hours. When
the CO2 level rises higher than ambient outdoor CO2 plus either an engineer-
determined set point per ASHRAE Standard 62.1 (ASHRAE 2019) or 600
ppm per California’s Title 24 (CEC 2018), the zone shall go into alarm for
ventilation and the backup ventilation system shall be started.

If a motorized device is used, then it is possible to use device or actuator


position feedback or end switches to meet the window position status
requirement. Note that buttons for motorized windows should be readily
accessible to the occupants within the directly ventilated space to comply with
the ventilation standards and codes.
Note that the full automation of windows to the exclusion of manually
controlled windows would make a space ineligible for application of the
natural conditioning comfort range detailed in ASHRAE Standard 55
(ASHRAE 2017).
For high-rise buildings in which wind speeds are substantially increased
and may cause stress on the actuator motors and/or require openings to have
modulation more granular than just open/closed, it is sometimes necessary to
apply additional protective sequences based on advice from the device
manufacturer:
1. When roof wind speeds rise above an amount to be determined through
trial and error or design team wind modeling and from directions to be
CHAPTER 7 | 317

determined through trial and error or design team wind modeling, close the
windows.
2. When roof wind speeds rise above an amount to be determined through
trial and error or design team wind modeling and from directions to be
determined through trial and error or design team wind modeling, reduce
the range of opening to a high limit of 50% of full stroke.
3. When differential static pressure at a given elevation between opposing
faces of a building exceeds an amount to be determined through trial and
error or design team wind modeling, reduce the range of opening to a high
limit of 50% of full stroke.

7.6.2 Natural Conditioning with Hot Hours


Cooling Changeover
For operation during heating mode and natural conditioning mode, see
previous sections as applicable. In addition to the guidance in those sections,
apply the following:
1. (Typical): When the indoor temperature exceeds the upper 80% deadband
for more than 30 minutes, then the zonal backup cooling system shall be
activated and all automatic windows and trickle vents shall be set to
“trickle” condition. Modulate zone cooling to hold a set point of 74°F
(24°C) with a 5°F (2.5°C) deadband.
2. (Specific to California): Within California’s Title 24 (CEC 2018), if any
window (manual or automatic) within the zone remains open, the zonal
cooling control is required to be placed in a setback set point of 90°F
(23°C) with a 5°F (2.5°C) deadband instead the set point in Item #1 above.
See Section WA5.2.3 of Website Appendix WA5.2, “Excerpts on Natural
Ventilation from California’s Title 24,” accessible at ashrae.org/naturalven-
tilation.
3. (Optional): The favorable conditions light shall shine when outdoor air
conditions are above the lower 80% acceptability limit and below the upper
80% acceptability limit minus 3°F (1.7°C) to be determined by the design
team based on internal heat loads and anticipated solar gain and/or trial and
error.

If chilled radiant surfaces or chilled beams are the zonal cooling system,
then the following sequence is strongly recommended in addition to the above:
1. The outdoor dew point shall be calculated using outdoor temperature and
outdoor relative humidity.
2. The indoor dew point shall be calculated using indoor temperature and
indoor relative humidity.
3. When either the indoor dew point temperature rises above zonal service
chilled-water supply temperature minus 1.5°F (0.8°C), or the outdoor dew
point rises above zonal service chilled-water supply temperature minus
2.0°F (1.0°C)
a. the indicator of the favorable condition will advise occupants to
318 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

close manual windows


a. the automation system will close the automated windows and
trickle vents
b. the backup ventilation and cooling systems will start with maxi-
mum ventilation quantities and exhaust to purge the moist outdoor
air for a minimum of 30 minutes.
Note that concurrent cooling is not allowed by ASHRAE Standard 55
(ASHRAE 2017) and thus is not addressed here. Zonal mixed-mode/hybrid
approaches are also not covered, as the natural conditioning zones and the
conditioned zones are separate spaces that operate independently.

7.6.3 Automatic Window Control for Nighttime Purge


of Heat from Thermal Mass
This is a simplified sequence using only the current afternoon temperature
to determine whether to trigger the night purge.
1. The outdoor air temperature at 4:00 p.m. shall be stored.
2. At the end of occupancy hours, all automatic devices shall close to end the
natural conditioning sequence for the day.
3. If the outdoor air temperature at 4:00 p.m. was greater than 75°F (24°C), a
night purge mode shall be invoked as follows:
a. If the outdoor air temperature is below 75°F (24°C), then upper
(near the ceiling) automatic windows shall modulate to hold indoor
space temperature to 75°F (24°C) until 11:00 p.m.
a. Then if the outdoor air temperature is below 75°F (24°C), the upper
automatic windows shall open fully and remain open until the ceil-
ing thermal mass surface temperature sensors have reached 68°F
(20°C) or until outdoor air temperatures have reached a nighttime
minimum and started to rise above daytime set point temperature
minus 3°F (1.7°C).
a. The automatic windows will then shut.
a. If the slab has not reached 68°F (20°C) by the start of occupancy,
then the system will revert to normal natural conditioning mode.
For nominal protection of the indoor environment and the actuator
equipment, see the following sequence:
1. If a window should be open and it is not, then an alarm shall be raised and
the backup ventilation system shall be started.
2. If a window should be closed and it is not, then an alarm shall be raised.
3. In the event of wind speeds exceeding an amount to be determined through
trial and error or design team wind modeling, close the windows.
4. In the event of wind speeds exceeding an amount to be determined through
trial and error or design team wind modeling and rain is sensed, close the
windows.
5. (Optional): If outdoor PM2.5 pollutants rise above an amount to be set by
the owner based on sensor quality, close the windows (to avoid damp dust
on surfaces during night purge).
CHAPTER 7 | 319

For high-rise buildings in which wind speeds are substantially increased


and may cause stress on the actuator motors and/or require openings to have
modulation more granular than just open/closed, it is sometimes necessary to
apply additional protective sequences based on advice from the device
manufacturer:
1. When roof wind speeds rise above an amount to be determined through
trial and error or design team wind modeling and from directions to be
determined through trial and error or design team wind modeling, close the
windows.
2. When roof wind speeds rise above an amount to be determined through
trial and error or design team wind modeling and from directions to be
determined through trial and error or design team wind modeling, reduce
the range of opening to a high limit of 50% of full stroke.
3. When differential static pressure at a given elevation between opposing
faces of a building exceeds an amount to be determined through trial and
error or design team wind modeling, reduce the range of opening to a high
limit of 50% of full stroke.

7.7 SELECTING ACTUATORS


There are many considerations when specifying the actuator type to be
used, namely:
1. Control functionality requirements: on/off versus modulating
2. Actuator mechanism type
3. Actuator-linkage coordination

The following section will give an overview of the range of concerns to be


addressed by the design team. The primary questions are as follows:

• How will the system be controlled and managed?


• How will individual occupants be affected by and interact with the
system and its various components?

The various components and subsystems that need to be considered include


the following:

a. Windows, ventilators, and roof lights designed for manual control.


b. Actuators on openings requiring automated control
c. Dampers or motorized louvers more typically used in mechanical ven-
tilation systems
d. The building management system (BMS), which must be integrated
with the manual control with automation and monitoring effectively

For example, providing manual control of windows by the occupants at a


low level and automatic control of high-level vents or motorized relief louvers
may be an option that preserves line of sight for the occupants and still gives
320 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

them manual control. BMS software must anticipate manual interventions and
allow for these within its control logic.
When specifying an actuator, the design team must balance the appearance
with functionality. The following are key items to explore with potential
manufacturers and to test as appropriate for acceptability:

• Low noise during automated movement to minimize disruption


• Occupant access and safety features
• Maintenance and replacement concerns
• Precision of control prevent drafts and improve temperature control
• Minimized visual impact of wiring/cables on the windows
• Window size
• Visibility of actuator and linkage and aesthetic concerns
• Real-time indication to the building owner of faults with the equipment

When considering the use of actuated windows, especially in a high-rise


building, it is necessary to ensure that the ventilator is capable of fully opening
or closing in the time period required by the code to achieve smoke control
performance. Typically, smoke control requires any ventilation device to be
fully open or closed within 60 seconds. For use as a smoke vent, the ventilation
opening, actuator, linkage, and controls must be tested as a complete package
in a laboratory, so the early selection of devices and windows configuration is
essential if used as part of a smoke control system. Operable windows can have
a negative impact on the smoke control strategies in high-rise buildings, as
smoke control is based on passive containment and maintaining pressure
differences between building spaces to control the movement of smoke in the
event of a fire. Having multiple façade openings complicates the smoke control
design and controls scenario. This issue should be addressed at the beginning
of any proposal to pursue natural ventilation.

7.7.1 Actuator Control Functionality Requirements


Ventilators and windows can be controlled in several ways. These include
the following (in order of increasing sophistication):

• ON/OFF: The opening has two positions: fully open and fully closed.
This coarse method is best suited to buffer spaces and atria in which
close control of comfort conditions is not essential. It is also more
suitable for outlet positions at a high level or in locations remote from
individual workstations where there is a reduced risk of uncomfortable
drafts.
• Bank control: The space has several windows or ventilators in
different banks each with ON/OFF control. This allows the ventilation
rate to be varied by opening up banks of ventilators in succession. The
banks can be multiple identical ventilators or ventilators of different
capacities which can provide either background or rapid ventilation.
CHAPTER 7 | 321

• Stepped control, with each opening having a number of fixed


positions: This simplifies the controller/actuator interface but does not
offer such close control as a modulated system.
• Fully-modulating operation: The actuator position is infinitely
variable in response to a control signal. This method can potentially
give greater control but, in practice, conditions may vary more rapidly
than the control and vents can respond. The best strategy is to keep the
control algorithms simple. Vent position feedback is still important.

Many control strategies currently use ON/OFF (open/close) control of inlets


and particularly outlets to achieve an acceptable indoor environment. Better
control could be obtained if intermediate travel distances are possible;
however, this may increase the cost of the actuator.
The actuator should provide position feedback to the control panel to
enable it to operate at a very slow speed when in the automatic mode, which
can reduce noise and any potential impact or disturbance to the occupants. It
can also enable the motors to operate at a faster speed when activated by the
manual keypads, for example, to provide an immediate visual response to the
user, and at full speed in the event of an alarm signal for smoke clearance.
Feedback gives the BMS better input on the exact position of the window for
precision control of opening size. High precision actuators help ensure that
openings for night cooling remain within allowable security limits, and bring
the potential to reduce costs by removing separate trickle vents fitted to the
façades for minimum outdoor air provision requirements.
Most standard actuators have a factory-set speed of operation, which in
many instances can be noisy. As these settings cannot be easily adjusted for
quieter operation once installed, it is strongly recommended that the design
team require an operational mock-up of the actuator device on a window of the
proper configuration and weight to be tested and witnessed by the owner and
design team.
Most standard actuators have little or no controls intelligence, and they are
often operated in large poorly controlled “chunks” of opening distance
associated with heavier oscillations around comfort set points, hunting,
uncomfortable drafts, excessive energy consumption, and potentially shorter
actuator life. In addition, many actuators do not have any high wind pressure
safety overrides built into them, nor any seating end switches to avoid
overheating the motors when driven to closure. Many actuators are a floating
three-point control that increases opening size in a relative sense based on the
length of time that the opening action takes place. This is very different from
most traditional mechanical devices that can stroke to a known distance when
the direction is received.

7.7.2 Actuator Mechanism Type


The types of automatic actuators commonly used for natural ventilation,
together with their linkage options, typical applications, and comments on their
suitability, are as follows:
322 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

• Linear push-pull piston: A motor propels a push-rod forward. These


are most commonly pneumatic but electrohydraulic versions are also
available. Advantages include mechanical simplicity, robustness, fire
resistance of pneumatic units, and generation of large forces.
Disadvantages include large projecting cylinders and mechanical
damage to windows, and linkages and fixings which are not sufficiently
robust. Travel is typically 8 in. to 20 in. (200 mm to 500 mm) but
longer distances are possible with large cylinders. They are most
widely used for roof lights and high-level windows.
• Projecting chain drive push-pull: An electric motor drives a chain
over a sprocket wheel, providing linear motion to push out a window.
They are generally modest in size and mechanical strength, and have
limited travel of typically 6 in. to 8 in. (150 mm to 200 mm). A useful
feature is that the motion tends to be at right angles to the axis of the
actuator body, which can therefore be tucked away in the plane of a
window frame or recessed and concealed within it. Their compact size
and unobtrusive appearance make them best suited to smaller windows
such as inward and outward opening fanlights.
• Rack-and-pinion: A rotary electric motor drives a geared shaft, which
engages with one or more racks, providing linear motion with less
bulky projections than linear actuators. Typical travel distances are 20
in. (500 mm) but travel distances of 40 in. (1000 mm) or more are
possible. They are particularly useful for windows that require paired
actuators (one on each side). With a common pinion, the two racks
move together, which avoids twisting the window (as can happen, for
example, if one of a pair of linear actuators fails, sometimes breaking
the window). They are most commonly used for roof lights with
relatively light frames.
• Linear-sleeved cable or rod: Driven by a rack-and-pinion, worm gear,
or chain drive electric motor and allow linear motion to be transferred,
for example, to sliding sashes.
• Rotary: These are most commonly applied with dampers and louvers,
often rotating one of the shafts directly, with mechanical linkages to the
other louvers. Sometimes they also operate shafts connected to cranks
and lever arms to provide linear motion. There are two main types: one
with a bidirectional motor used for opening and closing, and one which
only motors in one direction and uses a spring to return, which can be
useful for fail-safe operations.
• Lead screw: These actuators form most linear actuators and are sub-
divisible into high and low power applications. For general applications
from 200 N to 2000 N, a low friction nut is driven along the lead screw
to provide motive force.
• Lever arm: These actuators have been used since the first use of natural
ventilation around 1900. Continued use is a testament to the robustness
of the gear. It is often regarded as visually intrusive and unnecessarily
cumbersome today (but is still being manufactured).
CHAPTER 7 | 323

• Gas struts with cables: A linear variant on the spring return motor is
where an automatic catch releases the window and a gas-filled strut
opens it or holds it open. The closure is then affected by pulling on a
cord, either manually or by using a small electric winch.

7.7.3 Actuator-Linkage Coordination


Many factors need to be considered when choosing an actuator and its
linkage:

• Location of vent and manner of opening: Linear actuators are most


commonly used for roof lights and chain drives for low-level windows,
where it is more important for actuators to be unobtrusive and for them
not to obstruct the interior. See CIBSE (2005) for a range of vent types.
• Weight and size of vent: The motor should be sized to support the
appropriate level of forces, including wind, snow, and ice loads. Large
windows may require two (or more) actuators. Where opening vents
are about 40 in. (1000 mm) wide or more, the window or curtain wall
supplier should determine whether two actuators are required to
address weight and window torque. This is dictated by the flexing of
the frame and the window’s capacity to achieve a full seal at the edges
with a single pull point.
• Angle of attack: Owing to geometrical restrictions, the actuator force
cannot always be applied exactly in the direction of motion and can
sometimes be nearly at right angles to it. This can lead to large shear
forces and moments on fixings.
• Travel and free area to be achieved: Chain actuators are usually
limited to 6 in. to 10 in. (150 mm to 250 mm), though some concealed
units can reach up to 20 in. (500 mm). The chains have limited
buckling strength, so they are most commonly used to open top or
bottom-hung windows by a small angle. Linear piston actuators
typically have travels up to 24 in. (600 mm), while rack and pinion
drives can go up to 40 in. (1000 mm) and more.
• Available space: Actuators tend to swing as a window opens so
articulated linkages are required. Actuator and linkage components
must clearly not clash with each other, or with the surrounding
structure, during their full range of travel.
• Travel speed: Low travel speed is desirable to reduce loadings on the
actuator and to reduce noise. However, where occupant control is used,
a visible response is also important, so people can see that their control
actions have had a result. If the actuator moves too slowly or is not
visible from the point of operation, then some visual feedback such as
an indicator light or display should be provided. For roof lights, rapid
closure is desirable to minimize the entry of rain. A two-speed
operation may be desirable; slow for normal operation and fast for
manual over-ride and emergencies.
324 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

• Linkages and fixings: Linkages and fixings must be carefully


considered for appropriateness, mechanical strength, durability, and
ease of installation and maintenance. The supplier of the ventilator
should be made responsible for the fixing and mechanical
commissioning of the actuator and linkage. This should include
adjustment of any end stops and limit switches which, if not carefully
done, can easily lead to mechanical damage to the actuator, linkages,
fixings, or ventilator

It is vital, however, that the link between actuator and ventilator is not
overlooked. The following questions should be considered:

• How do the actuator and link connect to the opening element and the
frame or building fabric?
• Are its connection points and fixings strong enough and are any
linkages properly aligned?
• Will the element it opens be able to close and seal effectively?
• How does it connect to the control system and supplies of pneumatic or
electric power?
• Will it stand up to the duty cycles required? Care needs to be taken so
that the automatic control system does not constantly exercise the
actuator, leading to very short service life.
• Can it be maintained safely and conveniently?
• Are there any safety overrides for occupant safety (i.e., to avoid
pinching fingers)?

Most actuators used for natural ventilation are low voltage DC, which
gives higher torque capacity and makes them electrically safer when using the
window profile for containment. There are line voltage actuators on the
market, but these are generally designed for heavy windows or intermittent use
as they are prone to overheating if used regularly. Adequate service space and
access must be provided to properly maintain the operable façade systems, the
motors, and facilitate the eventual replacement of components over the life of
the building.
Ideally, the majority of the concerns stated in this subsection can be best
addressed if the opening device, actuator, linkage, and controls are procured as
a fully integrated system. Even then it is important to check whether products
have been adequately developed: manufacturers sometimes bolt-on actuators
to standard natural ventilation products without sufficient thought. It is also
crucial that there is sufficient coordination of inputs and outputs of the local
window controller to be coordinated with the BMS high-level monitoring and
control direction supervision as the master controller.

7.8 REFERENCES
ASHRAE. 2017. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
CHAPTER 7 | 325

ASHRAE. 2019. ANSI/ASHRAE Standard 62.1-2019, Ventilation for


acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
Brager, G., S. Borgeson, and Y. Lee. 2007. Summary report: Control strategies
for mixed mode buildings. Berkeley, CA: University of California,
Berkeley’s Center for the Built Environment.
CEC. 2018. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
USGBC. 2013. LEED® reference guide for building design and construction,
v4. Washington, DC: U.S. Green Building Council.

7.9 BIBLIOGRAPHY
WindowMaster. 2018. eBook for Architects: Why and how to use natural
ventilation in your building design. Copenhagen: WindowMaster.
WindowMaster. 2018. NV Embedded®. An adaptable solution. Naturally
intelligent. Copenhagen: WindowMaster.
SENSOR TYPES USED FOR
NATURAL VENTILATION
CONTROL
This appendix describes the most often used sensor types to control natural
ventilation, natural conditioning, and mixed-mode, change-over automation
sequences.

A7.1.1 OUTDOOR SENSORS ACTIVE IN


SEQUENCES OF OPERATIONS

A7.1.1.1 Outdoor Temperature


Outdoor temperature is monitored to determine the likely heat absorption
capability of the incoming air. As noted in Chapter 2, the closer the
temperature gets to the target indoor temperature, the less cooling it can
provide, and thus more outdoor air must be introduced to the space.
Outdoor air temperature is typically monitored globally on the roof through
a weather station in a shaded position or shielded from the sun to ensure that
only global dry-bulb temperature is measured. Where the building is
particularly high or there are surface heat island effects during the summer
near the lower portions of the building, it is recommended that a few outdoor
air temperature sensors are dispersed along the height of the building to
monitor incoming air temperature. If floor-by-floor isolation of natural
ventilation is available, then occupants could be notified if there are heat island
effects on the lower floors. Stack ventilation control sequences that use
motorized dampers to manipulate the effective size of upper openings could
also benefit from monitoring temperature differential in elevation.
Changeover thresholds for moving between a natural conditioning mode
and a mechanical cooling/heating mode would be programmed into the
building automation system (BAS) based on the comfort standards agreed
upon with the client. It is likely that loop tuning and self-learning algorithms
may be necessary over the first year to watch the speed of outdoor air
temperature rises and drops. This will allow control sequences to be fine-tuned

327
328 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

to avoid overshoot, which is defined as leaving the windows open too long and
allowing the indoor temperatures to rise beyond the appropriate threshold.
Outdoor air temperature trends will also indicate whether a night purge of
thermally massive surfaces should be engaged. Typically, these sequences for
purely naturally ventilated spaces (as opposed to mixed-mode spaces) will
observe outdoor air temperature at 4:00 p.m. and if the temperature remains
above a designated value at that time, the automated windows will either
remain open overnight or else open a few hours early to do a heat purge from
concrete exposed surfaces. It is also possible to obtain temperature predictions
from the Internet; however, BAS machine learning based on historical data
comparing outdoor, indoor, and surface temperature may be a better predictor
of building performance on the next day. Simplified routines monitor the
temperature at 4:00 p.m. against a threshold temperature to decide to engage
the night purge mode. Night purge modes are typically controlled thereafter
based on indoor surface temperatures if thermal mass is engaged. See
Section A7.1.2.4, “Thermal Mass Surface Temperature,” which covers
monitoring surface temperature, for more information.
Lastly, outdoor air temperature is often used to set the water temperature
and thus heat output for perimeter finned tube convectors in what is known as a
compensated loop. This is because on cold days the conductive heat loss and
the heating for the natural ventilation air temperature boost are proportional to
the temperature differential between outdoor air and indoor target temperature.

A7.1.1.2 Outdoor Relative Humidity


Similar to the outdoor air temperature measurement, the outdoor relative
humidity measurement is typically made on the roof in a weather station and
transmitted through the BAS as representative of all incoming air. The outdoor
relative humidity is not used typically on an instantaneous basis but is used to
watch trends in outdoor air dew point so that windows can be closed before
excessive amounts of moisture have entered into the indoor space in a mixed-
mode system using cooling elements exposed within the room air volume.
The particular need for outdoor relative humidity sensors for radiant and
chilled-beam mixed-mode systems is because high outdoor dew point would
be detrimental to any mixed-mode changeover system whose mechanical
system was based on a dedicated outside air system (DOAS) approach for
humidity control. This is because DOAS systems typically rely on
dehumidifying no more than three air changes per hour of air, with localized
water-based cooling within the rooms. At changeover, because the DOAS can
take up to an hour to evacuate the air in the room fully, cooling from the
radiant panels is severely limited due to the risk of condensation. The local
cooling systems (often radiant panels or chilled beams) operate on a sensible-
cooling loop with an elevated supply water temperature around 60°F (15.6°C),
so outdoor air dew points at changeover should be below 55°F (12.8°C), to be
confirmed during commissioning with condensation sensors.
APPENDIX A7.1 | 329

A7.1.1.3 Wind Speed


Wind speed is typically measured on the roof via a weather station. It is
used as an indicator of the strength of the wind, which would naturally lead to
the increased static pressures experienced around the outer surfaces of the
building. Wind speed is also typically monitored through the BAS against a
threshold value set by the window or actuator manufacturer as part of their
warranty process. When the threshold is reached, the windows will close to
avoid uplift, warping, or other damage to the windows or their actuators.

A7.1.1.4 Wind Direction


Wind direction is typically measured on the roof via a weather station. It is
used as an indicator to determine which side of the building is windward if
automatic controls on natural ventilation openings are designed to respond by
closing windows against strong pressures on the windward side and opening
extra windows on the leeward side in the event of overheating.
It should be noted that wind direction fluctuates frequently. Control
sequences relying on this data point to make decisions should incorporate a
rolling time-averaging routine and should limit actuator position changes to a
15-minute interval so as to not disturb the occupants.

A7.1.1.5 Rain Sensors


Rainfall/precipitation is typically monitored at a roof via a weather station.
Typically, windows are constrained by the building code to have openings no
larger than 4 in. or 8 in. (100 mm or 200 mm). Where top-hung windows are
used, there is usually limited exposure to wind-driven precipitation entering
the building as the majority of the water will shed; however, as rainfall
becomes heavy, the droplets at the window edge can be entrained into the
incoming air. Determining what wind speed versus rainfall rates are allowed
while the windows stay open is often decided through trial and error. For a
natural ventilation system with motorized elements, a control sequence can be
included to allow the facilities manager to set a threshold value for wind speed
during confirmed rainfall beyond which the windows should close to their
minimum rain-shedding position (in some cases this may have to be fully
closed).

A7.1.1.6 Outdoor Pollutants


Chapter 3 addresses the early phase outdoor air quality analysis procedure
to determine that the site is acceptable for natural ventilation, including
assessment of the pollution sources from neighboring properties. By the time
that the building is operational, this is a decision of the past, and the facilities
personnel will need to monitor the components of new and existing effluent
sources that appear on adjacent properties. During operations, it is not cost-
effective to provide real-time monitoring in a building of most of the pollutants
controlled under the National Ambient Air Quality Standards (NAAQS) (EPA
n.d.). Therefore, on days with poor ambient air quality (arising, for example,
from smog or a fire), it is often left to the local facilities manager to determine
330 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

as to whether natural ventilation will be used, based on the news or by


obtaining data into the BAS from the Internet.
In addition, periodically, an observational survey of the building site and
its immediate surroundings should be conducted during hours in which the
building is expected to be normally occupied to identify local contaminants
from surrounding facilities that will be of concern if allowed to enter the
building. Generator, cooling tower, and boiler flue effluents, along with VOC
smells from garbage areas are typical culprits. Control over incoming delivery
vehicles to prevent idling in proximity to the building is also something that
the facilities manager would put in place to improve incoming air quality.
If there is a systematic concern over degrading air quality, local levels of
air pollutants may be monitored and used to demonstrate compliance with the
governing authority’s standards. Air pollutants should be monitored at a
minimum of two locations, on the windward and leeward side of the building
or building site, at a minimum of 10 ft (3 m) above ground level. Monitoring
should occur over periods representative of the intended natural ventilation
operation period (i.e., seasonality should be considered). Air quality
monitoring should include, at minimum, monitoring for PM10, PM2.5, and O3
using the Federal Reference Method (FRM) or Federal Equivalence Method
(FEM) monitoring devices. If the region is known to be out of compliance with
other outdoor air pollutants on the NAAQS list, those pollutants should be
monitored as well.
PM10 and PM2.5 are the levels most known to cause asthma and are the
target pollutants that typical HVAC filtration as required by ASHRAE
Standard 62.1 (ASHRAE 2019) would remove from mechanically ventilated
airstreams. Typically, these values are higher near busy roadways than what is
reported in the regional outdoor air quality reports. The values can also
increase due to nearby construction activity. As noted in Chapter 1, it is not
possible to traditionally filter a natural ventilation airstream due to the very
low pressure drops available. Therefore, it is incumbent on the owner to
periodically assess outdoor air quality for these two pollutants versus the
NAAQS and Occupational Safety and Health Administration (OSHA)
standards for indoor air quality. Laser light-scattering technology is available
to measure these particulate concentrations at the site either through portable
devices or fixed devices connected to the BAS system.
Pollen is the other pollutant that is typically larger than PM10 but still has
adverse health effects on sensitive parties. There may be cases when a localized
zone is converted away from the natural ventilation mode as an accommodation
for the affected party if a mixed-mode operation is available. Alternatively, the
individual could temporarily relocate to an area with filtered air.

A7.1.1.7 Static Outside Air Pressure at Façade


If a building relies on wind-driven ventilation and there are clear windward
and leeward sides of the building, it is also possible to install an array of static
pressure sensors and measure the pressure differential between opposite sides
to determine airflow direction as the wind impinges on the building.
APPENDIX A7.1 | 331

On tall and long faces that are impinged by the wind, it is often useful to
incorporate an array of static outdoor air pressure sensors outside the façade
(integrated into maintenance catwalks or external shading devices). These
devices give a gross mapping of the pressures around the building and are a
better indicator of the direction of airflow through the façade opening than the
global wind direction. Because static pressures are operating at very low
values, it is the differential between the windward and leeward sides at the
same floor level that is most important to monitor. These devices are very
sensitive to small fluctuations and the differentials can easily pass from
negative to positive due to gusts. Therefore, response-calming measures need
to be programmed into the BAS sequence if the differential static pressure is
used to open or close windows.

A7.1.2 INDOOR SENSORS ACTIVE IN


SEQUENCES OF OPERATIONS

A7.1.2.1 Indoor Temperature


Indoor temperature is monitored to provide feedback on internal
conditions. When natural conditioning begins to fail to hold acceptable indoor
temperatures, the BAS should automatically alarm so that action can be taken.
If the system is a fully automated mixed-mode system, then the BAS can
automatically close the natural ventilation openings and start the mechanical
cooling system. If the system is manually controlled, then the alarm would be
fielded by the facilities manager, who can send an email to occupants to close
their windows to improve their environmental temperatures. Acceptability
thresholds for indoor temperature would be programmed into the BAS based
on the comfort standards agreed upon with the client and the authority having
jurisdiction (AHJ). In the U.S., ANSI/ASHRAE Standard 55 (ASHRAE 2017)
is typically the reference standard if no other criterion is proposed.
Typically, indoor temperature sensors are placed away from solar gain and
with thermal buffering if mounted on a junction box cast into a thermally
massive wall or column. The goal of the sensor is to read air temperature in the
space as one would with a traditional air-conditioning system. Zoning for
naturally ventilated areas is complicated by the fact that air will find an outlet
that may not be in the line of sight to the person who originally opened the
window. Therefore, a spread of sensors is useful to get a temperature map of
the indoor environment to understand the aggregated effects of the windows
open at any given time.
In a naturally ventilated environment, the acceptable dry resultant
temperature range is less well defined and various approaches have been
suggested in the following subsection.

A7.1.2.2 Indoor Relative Humidity


Indoor relative humidity is not typically used as an indicator to directly
open or close windows as there are many humid parts of the world where
332 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

natural ventilation remains the predominant cooling method via increasing air
speed as noted under ASHRAE Standard 55 (ASHRAE 2017).
Indoor relative humidity is primarily monitored only if there are radiant
devices that are operating as part of a mixed-mode system as noted above. The
feedback is used in conjunction with the room temperature to calculate dew point
and to centrally control the sensible loop supply temperature or to adjust the local
cooling valve control sequence to mix the supply and return water streams to
achieve a higher supply water temperature going to the radiant devices. Both
approaches reduce the risk of condensation by manipulating water temperature
before the water touches the devices that are exposed to the room air.

A7.1.2.3 Indoor CO2 Monitoring


CO2 monitoring for demand control ventilation remains a requirement in
the energy code and LEED® guidance for high-density occupancies regardless
of ventilation system type. As a point of reference only, the energy codes
establish the CO2 level that triggers a reversion to increased air supply as a
parts-per-million (ppm) threshold above the measured or assumed ambient
CO2 level outdoors.
It should be acknowledged that all air that enters a naturally ventilated
space from outdoors has the CO2 concentration of the outdoors, so it is solely
the occupants that would be the source of high CO2 in the space. The indoor
CO2 measurement is used as a proxy for indoor air quality due to pollutants
coming from occupants. In the case of a natural ventilation scheme, the CO2
sensors should be confirming that the indoor CO2 levels are remaining low
enough to meet IAQ requirements. This would imply that the occupants are
opening their windows enough to ensure ventilation rates equivalent to a
mechanical ventilation scheme.
Since naturally ventilated buildings cannot provide a measurement of
airflow or a guaranteed consistent ventilation rate, it is necessary to
demonstrate that an equivalent level of air quality has been provided. This can
be done by showing that the IAQ achieved by the natural ventilation is
equivalent to the ventilation rate required under ANSI/ASHRAE Standard 62.1
(ASHRAE 2019) or California’s Title 24 (CEC 2018) during occupied hours.
One way of doing this is to use the CO2 level in the space as a proxy for
general IAQ levels as one does for demand control ventilation in mechanical
ventilation systems. Section 120.1(d), subsection 4c of California’s Title 24
(CEC 2018) requires control to 600 ppm above ambient, with the assumption
that ambient is 400 ppm. These would be reasonable thresholds for alarming
when windows are closed during occupied hours if allowed by the AHJ in lieu
of permanently open windows.
The LEED® v4 rating system (USGBC 2013) requires one of the following
for natural ventilation systems to comply with the “Minimum Indoor Air
Quality Performance” credit: direct exhaust measurement, indoor CO2
monitoring (see above), or window switch position and alarming.
APPENDIX A7.1 | 333

A7.1.2.4 Thermal Mass Surface Temperature


Embedded temperature sensors are typically installed either when a radiant
floor or ceiling system is used or when the manipulation of thermal mass is
completed during a night purge of absorbed heat. A temperature sensor sheath
is typically cast into the concrete along with a junction box, where the wiring
and transmitter are stored. The goal is typically to measure approximately 2 in.
to 3 in. (50 mm to 75 mm) below the surface as only approximately a 4 in.
(100 mm) depth is cooled during a swing of the diurnal cycle.
In the case of the night purge, the feedback from the surface temperature
sensor tells the BAS when to close the windows based on a sequence of
operations that sets a threshold of surface temperature to avoid dropping below
66°F (19°C) as per ASHRAE Standard 55 (ASHRAE 2017) requirements. If
the thermally massive or mechanically cooled surface is overhead, then the
temperatures can drop as far as comfort calculations related to thermal
asymmetry can allow.

A7.1.2.5 Window/Louver Closed Position Status


As noted within the description of the sequences, most of the standards
require some form of window-position monitoring. Typically, this is provided
through a switch in the frame for manually operated windows or doors and
either in the frame or the actuator for automatic devices. Typically, the
important information is that the window is “not closed,” as this would
designate a natural ventilation mode. When a building is designed to rely on
natural ventilation, the “closed” feedback during designated hours of
occupancy would raise an alarm that the space may not be achieving indoor air
quality, and is not in compliance with the permanent opening expectations of
ASHRAE Standard 62.1 (ASHRAE 2019). When a building has a mixed-mode
system, the “not closed” status should turn off the mechanical cooling system
and ramp down the air supply to meet mechanical ventilation requirements
only. Sophisticated controls sequences may wish to have more discretized
position feedback from the actuator for other reasons related to indoor velocity
control or the use of windows to automatically control for set point
temperatures.
As previously stated, the LEED® v4 rating system (USGBC 2013) requires
one of the following for natural ventilation systems to comply with the
“Minimum Indoor Air Quality Performance” credit: direct exhaust
measurement, indoor CO2 monitoring, or window switch position and
alarming. California’s Title 24 (California Energy Commission 2015a)
requires some form of interlocking controls to shut off mechanical systems
when the windows are open. See Section WA5.2.3 of Website Appendix
WA5.2, “Excerpts on Natural Ventilation from California’s Title 24,”
accessible at ashrae.org/naturalventilation.

A7.1.2.6 Favorable Condition Indicator


It was beneficial for some projects to have a lighted indicator near the
manual window that signals to the occupants when it is the ideal time to open
334 | ASHRAE DESIGN GUIDE FOR NATURAL VENTILATION

the window for natural conditioning. When used to notify the occupants that
the zone has moved out of heating mode, this indicator light is an aide in
conveying outdoor air temperature feedback when the occupant does not have
direct personal experience of airflow touching skin to make that decision
themselves. When the indicators are used to advise closing windows, it is
typically because of safety overrides (such as high wind or wind-driven rain)
or because the zone must go into its cooling mode within a mixed-mode
system.

A7.1.3 REFERENCES
ASHRAE. 2017. ANSI/ASHRAE Standard 55-2017, Thermal environmental
conditions for human occupancy. Peachtree Corners, GA: ASHRAE.
ASHRAE. 2019. ANSI/ASHRAE Standard 62.1-2019, Ventilation for
acceptable indoor air quality. Peachtree Corners, GA: ASHRAE.
CEC. 2018. 2019 Building Energy Efficiency Standards for Residential and
Nonresidential buildings for the 2019 Building Energy Efficiency
Standards: Title 24, Part 6, and Associated Administrative Regulations in
Part 1. CEC-400-2018-020-CMF. Sacramento, CA: California Energy
Commission.
EPA. n.d. NAAQS table. Washington, DC: U.S. Environmental Protection
Agency. Retrieved from https://www.epa.gov/criteria-air-pollutants/naaqs-
table on January 21, 2019.
USGBC. 2013. LEED® reference guide for building design and construction,
v4. Washington, DC: U.S. Green Building Council.
Expert Guidance on Designing and Implementing Natural Ventilation

ASHRAE Design Guide for Natural Ventilation provides a comprehensive overview of


the application of natural ventilation in modern buildings. In addition to a thorough discussion
of terminology and key concepts related to the physical principles of natural ventilation, this
guide also covers how to conduct a multifactor analysis to determine the feasibility of
implementing natural ventilation in a building. This guide also expands on currently available
natural-ventilation-related standards and codes from ASHRAE and other organizations, and
includes case studies of buildings from around the world that are excellent examples of best
practices for the implementation of natural ventilation, with worked examples included. This
guide is accompanied by supplemental appendices, which are available online in PDF
format. In addition to in-depth information about the design of natural ventilation projects,
they also provide more results of an analysis of natural ventilation systems for cities of
varying climates.

Written by two leading experts in the field, ASHRAE Design Guide for Natural Ventilation
assists engineers, architects, building owners, facilities personnel, and building design
professionals in exploring the feasibility of natural ventilation for their projects during the
early phases of design as a way to encourage designing energy-efficient naturally ventilated
buildings.

ISBN 978-1-947192-54-6 (paperback)


ISBN 978-1-947192-55-3 (PDF)
ISBN 9781947192546

180 Technology Parkway


Peachtree Corners, GA 30092 9 781947 192546
404-636-8400 (worldwide)
www.ashrae.org
Product code: 90306 06/21

You might also like