You are on page 1of 517

Solar and Infrared

Radiation Measurements
ENERGY AND THE ENVIRONMENT
SERIES EDITOR
Abbas Ghassemi
New Mexico State University

PUBLISHED TITLES

Hydroelectric Energy: Renewable Energy and the Environment


Bikash Pandey and Ajoy Karki

Geologic Fundamentals of Geothermal Energy


David R. Boden

Introduction to Bioenergy
Vaughn Nelson and Kenneth Starcher

Introduction to Renewable Energy, Second Edition


Vaughn Nelson and Kenneth Starcher

Environmental Impacts of Renewable Energy


Frank R. Spellman

Geothermal Energy: Renewable Energy and the Environment,


Second Edition
William E. Glassley

Energy Resources: Availability, Management, and Environmental Impacts


Kenneth J. Skipka and Louis Theodore

Finance Policy for Renewable Energy and a Sustainable Environment


Michael Curley

Wind Energy: Renewable Energy and the Environment, Second Edition


Vaughn Nelson

Solar Radiation: Practical Modeling for Renewable Energy Applications


Daryl R. Myers

Solar and Infrared Radiation Measurements, Second Edition


Frank Vignola, Joseph Michalsky, and Thomas Stoffel

Forest-Based Biomass Energy: Concepts and Applications


Frank Spellman

Solar Energy: Renewable Energy and the Environment


Robert Foster, Majid Ghassemi, Alma Cota,
Jeanette Moore, and Vaughn Nelson
Solar and Infrared
Radiation Measurements
Second Edition

Frank Vignola
Joseph Michalsky
Thomas Stoffel
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742

© 2020 by Taylor & Francis Group, LLC


CRC Press is an imprint of Taylor & Francis Group, an Informa business

No claim to original U.S. Government works

Printed on acid-free paper

International Standard Book Number-13: 978-1-138-09629-5 (Hardback)

This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been made
to publish reliable data and information, but the author and publisher cannot assume responsibility for the validity of all
materials or the consequences of their use. The authors and publishers have attempted to trace the copyright holders of all
material reproduced in this publication and apologize to copyright holders if permission to publish in this form has not been
obtained. If any copyright material has not been acknowledged please write and let us know so we may rectify in any future
reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized
in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying,
microfilming, and recording, or in any information storage or retrieval system, without written permission from the
publishers.

For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://www.
copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-
8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For organizations that
have been granted a photocopy license by the CCC, a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for
identification and explanation without intent to infringe.

Library of Congress Cataloging‑in‑Publication Data

Names: Vignola, Frank, 1945- author. | Michalsky, Joseph, author. | Stoffel,


Thomas, author.
Title: Solar and infrared radiation measurements / Frank Vignola, Joseph
Michalsky, Thomas Stoffel.
Description: Second edition. | Boca Raton : CRC Press, Taylor & Francis
Group, 2020.
Identifiers: LCCN 2019013997 | ISBN 9781138096295 (hardback : alk. paper)
Subjects: LCSH: Solar radiation--Measurement--Textbooks. | Earth
temperature--Measurement--Textbooks. | Radiometers.
Classification: LCC QC912 .V56 2020 | DDC 523.7/2--dc23
LC record available at https://lccn.loc.gov/2019013997

Visit the Taylor & Francis Web site at


http://www.taylorandfrancis.com

and the CRC Press Web site at


http://www.crcpress.com
Dedication

We dedicate this book to our wives, Mary Lou Vignola,


Randi Michalsky, and Ann Stoffel, who provided encouragement
and support while we worked on the manuscript.
Contents
Preface to the Second Edition .......................................................................................................... xv
Preface to the First Edition ............................................................................................................xvii
Acknowledgments ...........................................................................................................................xix
Authors ............................................................................................................................................xxi

Chapter 1 Measuring Solar and Infrared Radiation .....................................................................1


Questions ......................................................................................................................5
References ....................................................................................................................5

Chapter 2 Definitions and Terminology........................................................................................7


2.1 Introduction ....................................................................................................... 7
2.2 The Sun.............................................................................................................. 7
2.3 Extraterrestrial Radiation ..................................................................................8
2.4 Solar Coordinates .............................................................................................. 9
2.5 Zenith, Azimuth, and Hour Angles ................................................................. 14
2.6 Solar, Universal, and Local Standard Time .................................................... 15
2.7 Solar Position Calculation Example ................................................................ 16
2.8 Sunrise and Sunset Times ...............................................................................20
2.9 Global, Direct Normal, and Diffuse Irradiance ..............................................20
2.10 Solar Radiation on Tilted Surfaces.................................................................. 23
2.11 The Spectral Nature of Solar Radiation .......................................................... 23
2.12 Fundamentals of Thermodynamics and Heat Transfer ...................................26
2.12.1 Conduction.........................................................................................26
2.12.2 Convection ......................................................................................... 27
2.12.3 Radiative Heat Transfer .....................................................................28
2.13 Photodiodes and Solar Cell Characteristics ....................................................28
2.14 Models ............................................................................................................. 30
Questions .................................................................................................................... 30
References .................................................................................................................. 30

Chapter 3 Historic Milestones in Solar and Infrared Radiation Measurement .......................... 33


3.1 Introduction ..................................................................................................... 33
3.2 Earliest Observations of the Sun and the Nature of Light............................... 33
3.3 Nineteenth-Century Radiometers.................................................................... 37
3.3.1 Pouillet’s Pyrheliometer (1837) ......................................................... 37
3.3.2 Campbell-Stokes Sunshine Recorder (1853, 1879) ............................ 37
3.3.3 Ångström Electrical Compensation Pyrheliometer (1893) ................ 38
3.3.4 Callendar Pyranometer (1898) ...........................................................40
3.3.5 Ångström and Tulipan Pyrgeometers (1899) ..................................... 41
3.4 Operational Radiometers of the Twentieth Century........................................ 41
3.4.1 Abbot Silver-Disk Pyrheliometer (1906) ........................................... 41
3.4.2 Smithsonian Water-Flow Pyrheliometer (1910)................................. 43
3.4.3 Marvin Pyrheliometer (1910)............................................................. 43

vii
viii Contents

3.4.4 Ångström Pyranometer (1919) ...........................................................44


3.4.5 Kipp & Zonen Solarimeter (1924).....................................................44
3.4.6 Robitzsch Bimetallic Actinograph (1932).......................................... 45
3.4.7 Eppley 180° Pyrheliometer (1930) .....................................................46
3.4.8 Eppley Model PSP (1957) .................................................................. 48
3.4.9 Yanishevsky Pyranometer (1957) ...................................................... 49
3.4.10 Eppley Model NIP (1957) .................................................................. 51
3.4.11 Eppley Model Precision Infrared Radiometer (PIR) (1968) .............. 53
3.4.12 Primary Absolute Cavity Radiometer (PACRAD) (1969)................. 54
3.4.13 Eppley Model 8-48 (1969) ................................................................. 55
3.4.14 LI-COR Model LI-200SA (1971) ...................................................... 56
3.4.15 Rotating Shadowband Radiometer (1975) ......................................... 57
3.4.16 World Standard Group (1979)............................................................60
3.5 Recent Advances in Solar Measurements ....................................................... 61
3.5.1 Automatic Hickey–Frieden Cavity Radiometer ................................ 61
3.5.2 Total Irradiance Monitor (TIM) ........................................................ 63
3.5.3 Cryogenic Solar Absolute Radiometer-Measure the Integral
Transmittance (CSAR-MITRA) ........................................................64
3.6 Summary ......................................................................................................... 65
Questions .................................................................................................................... 65
References .................................................................................................................. 65

Chapter 4 Direct Normal Irradiance ........................................................................................... 69


4.1 Overview of Direct Normal Irradiance ........................................................... 69
4.2 Pyrheliometer Geometry ................................................................................. 72
4.3 Operational Thermopile Pyrheliometers ......................................................... 74
4.4 Absolute Cavity Radiometers .......................................................................... 77
4.5 Uncertainty Analysis for Pyrheliometer Calibration....................................... 78
4.6 Uncertainty Analysis for Operational Thermopile Pyrheliometers ................80
4.6.1 Window Transmittance, Receiver Absorptivity, and
Temperature Sensitivity ..................................................................... 81
4.6.2 Solar Zenith Angle Dependence ....................................................... 81
4.7 Uncertainty Analysis for Rotating Shadowband Radiometer Estimates
of Direct Normal Irradiance ............................................................................ 83
4.8 Direct Normal Irradiance Models ...................................................................84
4.8.1 Ground-Based Modeling ...................................................................84
4.8.2 Satellite Model Estimates ..................................................................84
4.9 Historical and Current Surface-Measured Direct Normal
Irradiance Data ....................................................................................... 86
4.10 Current Issues Regarding Direct Normal Irradiance Measurements .............. 88
Questions .................................................................................................................... 89
References .................................................................................................................. 89

Chapter 5 Broadband Global Irradiance ..................................................................................... 91


5.1 Introduction to Global Horizontal Irradiance Measurements ......................... 91
5.2 Black-Disk Thermopile Pyranometers ............................................................97
5.2.1 Thermal Offsets............................................................................... 101
5.2.2 Nonlinearity..................................................................................... 103
5.2.3 Spectral Response............................................................................ 103
Contents ix

5.2.4 Angle of Incidence Response .............................................................. 104


5.2.5 Response Degradation......................................................................... 107
5.2.6 Temperature Dependence.................................................................... 108
5.2.7 Ice and Snow on Dome—Ventilators .................................................. 108
5.2.8 An Optical Anomaly ........................................................................... 109
5.3 Black-and-White Pyranometers ....................................................................... 111
5.3.1 Characteristics of Black-and-White Pyranometers ............................. 111
5.3.2 Lack of Thermal Offset....................................................................... 114
5.4 Photodiode-Based Pyranometers .................................................................... 115
5.4.1 Characterizing a Photodiode Pyranometer ......................................... 121
5.4.2 Removing Biases in Photodiode Pyranometer Measurements ........... 124
5.4.3 Reference Solar Cells .......................................................................... 130
5.5 Calibration of Pyranometers ........................................................................... 131
5.5.1 Shade–Unshade Calibration Method .................................................. 132
5.5.2 Summation Method Calibration.......................................................... 135
5.6 Pyranometer Calibration Uncertainties ........................................................... 137
5.6.1 Uncertainty Analysis Applied to Pyranometer Calibration ................ 139
5.6.2 An Example of the GUM Procedure Applied to the Calibration
Uncertainties of a Pyranometer .......................................................... 141
5.6.3 Importance of Understanding Limitations of Percent
Uncertainties ................................................................................. 144
Questions .................................................................................................................. 145
References ................................................................................................................ 146
Useful Links ............................................................................................................. 149

Chapter 6 Diffuse Irradiance .................................................................................................... 151


6.1 Introduction ..................................................................................................... 151
6.2 Atmospheric Scattering Concepts ................................................................... 151
6.3 Measuring Diffuse Irradiance ......................................................................... 153
6.3.1 Fixed Shadowband Measurements of Diffuse Irradiance................... 153
6.3.2 Calculated Diffuse Irradiance versus Shading Disk Diffuse.............. 155
6.3.3 Rotating Shadowband Diffuse Measurements .................................... 156
6.4 Calibration of Diffuse Pyranometers .............................................................. 157
6.5 Value of Accurate Diffuse Measurements ...................................................... 158
Questions .................................................................................................................. 159
References ................................................................................................................ 160

Chapter 7 Solar Spectral Measurements ................................................................................... 161


7.1 Introduction ..................................................................................................... 161
7.2 The Extraterrestrial Solar Spectrum ............................................................... 161
7.3 Atmospheric Interactions ................................................................................ 163
7.3.1 Rayleigh Scattering ............................................................................. 163
7.3.2 Aerosol Scattering and Absorption ..................................................... 163
7.3.3 Gas Absorption ................................................................................... 165
7.3.4 Transmission of the Atmosphere......................................................... 169
7.4 Broad Filter Radiometry.................................................................................. 170
7.4.1 Photometry .......................................................................................... 170
7.4.2 Photosynthetically Active Radiation (PAR) ........................................ 173
7.4.3 UVA and UVB .................................................................................... 174
x Contents

7.5 Narrow-Band Filter Radiometry .................................................................. 175


7.5.1 Aerosol Optical Depth ..................................................................... 175
7.5.2 Water Vapor ..................................................................................... 177
7.5.3 Sun Radiometers .............................................................................. 179
7.6 Spectrometry ................................................................................................ 181
7.6.1 Spectrometers................................................................................... 182
7.6.2 Spectral Models ............................................................................... 186
7.6.3 Discrete Spectral Measurements and Modeling Combined ............ 188
Questions ................................................................................................................. 188
References ............................................................................................................... 189

Chapter 8 Albedo..................................................................................................................... 193


8.1 Introduction .................................................................................................. 193
8.2 Broadband Albedo ....................................................................................... 193
8.3 Spectral Albedo ........................................................................................... 194
8.4 Bidirectional Reflectance Distribution Function ......................................... 199
8.5 Albedo Measurements .................................................................................200
8.5.1 Broadband Albedo ...........................................................................200
8.5.2 Spectral Albedo ............................................................................... 201
Questions .................................................................................................................202
References ...............................................................................................................202

Chapter 9 Measuring Solar Radiation on a Tilted Surface ..................................................... 205


9.1 Introduction .................................................................................................. 205
9.2 Effect of Tilt on Single Black Detector Pyranometers ................................206
9.3 Effect of Tilt on Black-and-White Pyranometers ........................................208
9.4 Effect of Tilt on Photodiode Pyranometers .................................................209
9.5 Recommendations for Tilted Irradiance Measurements ............................. 211
9.6 Modeling Photovoltaic System Performance with Data from
Photodiode Pyranometers ............................................................................ 212
Questions ................................................................................................................. 213
References ............................................................................................................... 214

Chapter 10 Shadowband Radiometers ....................................................................................... 215


10.1 Introduction .................................................................................................. 215
10.2 Introduction to the Rotating Shadowband Radiometer................................ 215
10.3 Rotating Shadowband Radiometer Using Silicon Detector ......................... 217
10.4 Multifilter Rotating Shadowband Radiometer ............................................. 223
10.5 SPN1 Sunshine Pyranometer ....................................................................... 225
Questions ................................................................................................................. 226
References ............................................................................................................... 227

Chapter 11 Infrared Measurements ........................................................................................... 229


11.1 Introduction .................................................................................................. 229
11.2 Pyrgeometers ............................................................................................... 230
11.3 Calibration ................................................................................................... 233
11.4 Improved Calibrations ................................................................................. 235
Contents xi

11.5 Other Pyrgeometer Manufacturers .............................................................. 236


11.6 Operational Considerations.......................................................................... 236
Questions ................................................................................................................. 239
References ............................................................................................................... 239

Chapter 12 Net Radiation Measurements .................................................................................. 241


12.1 Introduction .................................................................................................. 241
12.2 Single-Sensor (All-Wave) Net Radiometers ................................................. 242
12.3 Two-Sensor Net Radiometers ......................................................................244
12.4 Four-Sensor Net Radiometers ...................................................................... 245
12.5 Accuracy of Net Radiometers ......................................................................246
12.6 A Better Net Radiation Standard .................................................................246
12.7 Net Radiometer Sources .............................................................................. 247
Questions ................................................................................................................. 247
References ............................................................................................................... 247

Chapter 13 Radiometer Calibrations ......................................................................................... 249


13.1 Introduction .................................................................................................. 249
13.1.1 What Is Calibration?....................................................................... 249
13.1.2 Why Is Calibration Needed? .......................................................... 252
13.1.3 How Frequently Should a Radiometer Be Calibrated? .................. 253
13.2 Broadband Shortwave Radiometer Calibration ........................................... 255
13.3 Broadband Longwave Radiometer Calibration............................................ 257
13.4 Spectral Calibrations.................................................................................... 262
13.4.1 The Measurement Equation ........................................................... 262
13.4.2 Standard Lamps ............................................................................. 263
13.4.3 Langley Plots ..................................................................................264
Questions ................................................................................................................. 267
References ............................................................................................................... 267

Chapter 14 Ancillary Measurements......................................................................................... 271


14.1 Introduction .................................................................................................. 271
14.2 Ambient Temperature .................................................................................. 271
14.2.1 Types of Temperature Sensors ....................................................... 272
14.2.2 Response Times.............................................................................. 272
14.2.3 Measuring Temperature ................................................................. 272
14.3 Wind Speed and Wind Direction ................................................................. 274
14.3.1 Sensor Terminology ....................................................................... 275
14.3.2 Anemometer ................................................................................... 275
14.3.3 Cup Anemometers .......................................................................... 275
14.3.4 Propeller Anemometers .................................................................. 277
14.3.5 Sonic Anemometers ....................................................................... 278
14.3.6 Installing Anemometers ................................................................. 278
14.3.7 Wind Vanes .................................................................................... 278
14.4 Relative Humidity ........................................................................................ 281
14.5 Atmospheric Water Vapor............................................................................ 282
14.5.1 Using GPS Satellites to Measure Precipitable Water Vapor .......... 283
14.5.2 Installing a Global Positioning System Antenna ........................... 283
xii Contents

14.6 Pressure ......................................................................................................284


14.6.1 Aneroid Displacement Transducers .............................................284
14.6.2 Piezoresistive Barometers ............................................................284
14.6.3 Piezocapacitance Barometers ....................................................... 286
14.7 Sky-Imaging Systems ................................................................................ 286
14.7.1 Site Surveys .................................................................................. 286
14.7.2 Sky Conditions ............................................................................. 289
14.8 Circumsolar Instruments ........................................................................... 292
14.8.1 Sun and Aureole Measurement .................................................... 292
14.8.2 Attempts to Automatically and Continuously Monitor the
Circumsolar Irradiance ................................................................ 292
14.8.3 Use of the Sky Imager to Measure Circumsolar Irradiance ........ 294
14.8.4 Use of the Rotating Shadowband Radiometer.............................. 294
14.9 Recommended Minimum Accuracies for Operational Instruments ......... 295
Questions ................................................................................................................. 295
References ............................................................................................................... 296

Chapter 15 Solar Monitoring Station Best Practices ................................................................. 299


15.1 Introduction ................................................................................................ 299
15.2 Choosing a Site .......................................................................................... 299
15.3 Grounding and Shielding ........................................................................... 301
15.4 Data Logger and Communications ............................................................302
15.5 Measurement Interval ................................................................................ 303
15.6 Cleaning and Maintenance ........................................................................ 305
15.7 Record Keeping .........................................................................................307
15.8 Importance of Reviewing Data ..................................................................309
15.9 Quality Control of Data ............................................................................. 311
15.10 Field Calibrations ....................................................................................... 313
15.11 Physical Layout of a Solar-Monitoring Station .......................................... 315
Questions ................................................................................................................. 317
References ............................................................................................................... 317

Chapter 16 Solar Radiation Estimates Derived from Satellite Images and Auxiliary
Measurements ......................................................................................................... 319
16.1 Introduction ................................................................................................ 319
16.2 Geostationary Satellites ............................................................................. 320
16.3 Deriving Irradiance from Satellites ........................................................... 321
16.3.1 Physical Models............................................................................ 321
16.3.2 Empirical Models ......................................................................... 322
16.3.3 Global Irradiance ......................................................................... 322
16.3.4 Pixel-to-Cloud Index Conversion ................................................. 322
16.3.5 Cloud Index to Global Horizontal Irradiance Conversion ........... 323
16.3.6 Direct Irradiance .......................................................................... 323
16.3.7 Diffuse Irradiance ........................................................................ 324
16.3.8 Tilted Irradiance ........................................................................... 324
16.4 Status of Satellite Irradiance Models ......................................................... 324
16.5 Comments on Modeling and Measurement ............................................... 325
Questions ................................................................................................................. 326
References ............................................................................................................... 326
Contents xiii

Appendix A: Measurement Uncertainty Principles ................................................................. 329


Appendix B: Modeling Solar Radiation..................................................................................... 343
Appendix C: Sunshine Duration ................................................................................................ 363
Appendix D: Sun Path Charts .................................................................................................... 365
Appendix E: Solar Position Algorithms..................................................................................... 385
Appendix F: Useful Conversion Factors .................................................................................... 393
Appendix G: Sources for Equipment ......................................................................................... 397
Appendix H: BORCAL Report................................................................................................... 401
Appendix I: Failure Modes ......................................................................................................... 435
Appendix J: How to Build a Pyranometer with a Solar Cell or Photodiode.......................... 441
Appendix K: Content Required for a Comprehensive Datafile ............................................... 445
Appendix L: Solar Radiation Databases ................................................................................... 455
Answers to Chapter Questions ...................................................................................................469
Index .............................................................................................................................................. 485
Preface to the Second Edition
We are gratified by the positive response to the first edition of Solar and Infrared Radiation
Measurements and grateful for the opportunity to update our book with recent advances in this
second edition. Our goal for this edition is for it to continue to serve as a ready source of knowledge
for those interested in understanding radiometry for applications in renewable energy and the global
radiation budget.
In the process of revising some of the existing content, we have added new chapters and appen-
dices addressing the latest research and organizing key information concerning:

• Radiometer calibration
• Estimating solar irradiance at the surface from satellite observations
• Measurement uncertainty principles
• Building a simple pyranometer
• A comprehensive new data format
• Sources of measured irradiance data

We apologize to our readers for the errors in the first edition. We have done our best to correct previ-
ous errors as we worked to avoid errors in this new edition. We thank the editorial staff at Taylor &
Francis for creating the second edition from our draft manuscripts.

xv
Preface to the First Edition
The rather specialized field of solar and infrared radiation measurements has become increasingly
important due to the increased demands by the renewable energy and climate change research
communities for data with higher accuracy and increased temporal and spatial resolutions. Recent
advances in radiometry, measurement systems, and information dissemination have also increased
the need for refreshing the literature available for this topic.
The objective of this book is to provide the reader with an up-to-date review of the important
aspects of solar and infrared radiation measurements: radiometer design; equipment installation,
operation, maintenance, and calibration; data quality assessment parameters; and the knowledge
necessary to properly interpret and apply the measured data to a variety of topics. Each of the
authors has more than 40 years of experience with this subject, primarily as the result of developing
and operating multiple measurement stations, working with the industry to improve radiometry, and
conducting various research projects.
The book’s scope and subject matter have been designed to help a wide audience understand the
general subject and to serve as a technical reference. A student new to the field will benefit from the
review of terminology and the historical perspective for radiometry before addressing more detailed
topics in radiometry that we hope will be of interest to the more experienced reader.

xvii
Acknowledgments
The  authors thank the sponsors of their organizations for the support over the years that helped
build the experience and expertise that went into this book.
Sponsors of the University of Oregon Solar Radiation Monitoring Laboratory who should be
recognized for their support are the Bonneville Power Administration, Energy Trust of Oregon, and
the National Renewable Energy Laboratory. We also thank the University of Oregon and Luminate,
LLC, for encouraging our participation in this book.
In  addition to the primary sponsorship of the Global Monitoring Division’s Radiation Group
by the National Oceanic and Atmospheric Administration, considerable support comes from the
U.S. Department of Energy Office of Science with additional contributions from the National
Aeronautics and Space Administration, the U.S. Department of Agriculture through Colorado State
University, and the Environmental Protection Agency.
The DOE BER Atmospheric Radiation Measurement Program of the U.S. Department of Energy
Office of Science and the DOE Office of Energy Efficiency and Renewable Energy of the U.S.
Department of Energy provided support for the Sensing, Measurement, and Forecasting Group
at the National Renewable Energy Laboratory and its development and operation of the Solar
Radiation Research Laboratory.
Individuals who have generously given advice or aided with the figures during the writing of
this book include Peter Armstrong, John Augustine, Will Beuttell, Victor Cassella, Chris Cornwall,
Patrick Disterhoft, Robert Dolce, Ellsworth Dutton, Julian Gröbner, Chris Gueymard, Peter Harlan,
Bobby Hart, John Hickey, Rich Kessler, Gary Hodges, Ed Kern, Peter Kiedron, Tom Kirk, Jan
Kleissl, Vikki Kourkouliotis, Kathy Lantz, Fuding Lin, Tilden Meyers, Daryl Myers, Richard
Perez, Josh Peterson, Ibrahim Reda, Ruud Ringoir, Angie Skovira, and Steve Wilcox.

xix
Authors
Frank Vignola is the director of the University of Oregon Solar Energy Center and runs the Solar
Radiation Monitoring Laboratory. He received his Bachelor of Arts in physics from the University
of California–Berkeley in 1967 and his Doctor of Philosophy in physics from the University of
Oregon in 1975, with his thesis on elementary particle physics. He decided to apply his skills to
more practical applications and started working in solar energy in 1977 at the University of Oregon.
Dr. Vignola helped establish and manage the SRML solar radiation monitoring network of the Solar
Radiation Monitoring Laboratory that has created the longest-running high-quality solar radiation
data set in the United States. He organized and participated in many workshops on solar resource
assessment and has written and contributed to approximately 100 papers on solar resource assess-
ment. He  was technical chair of the 1994 and 2004 conferences of the American Solar Energy
Society and for several years chaired the Technical Review Committee for the American Solar
Energy Society that managed the publication of several white papers on solar energy. He is currently
an associate editor for solar resource assessment for the Solar Energy Journal. To help facilitate the
use of solar energy, he created and maintains a solar resource assessment website that is accessed
by over 150,000 users a year. In  addition, he has served on the boards of the American Solar
Energy Society and the Solar Energy Industries Association and is past president of the Oregon
Solar Energy Industries Association. Dr. Vignola has helped pass solar tax credit and net metering
legislation in Oregon and was author of Oregon’s law requiring 1.5 percent of the capital for solar
in new public buildings.

Joseph Michalsky received his Bachelor of Science in physics from Lamar University and Master
of Science and Doctor of Philosophy in physics from the University of Kentucky. He  began his
career with Battelle Memorial Institute that operates the Department of Energy’s Pacific Northwest
National Laboratory. Following that he was with the Atmospheric Sciences Research Center a part
of the State University of New York in Albany. He then joined the Global Monitoring Division within
the Earth System Research Laboratory in the Office of Oceanic and Atmospheric Research, a part
of the National Oceanic and Atmospheric Administration. He continued his research part-time for
a few years with the University of Colorado’s Cooperative Institute for Research in Environmental
Sciences. Now fully retired he continues to dabble in atmospheric research of his own choosing. His
early career focused on astronomical research before taking on problems in solar energy. His focus
in the major part of his career has been in the solar energy and atmospheric sciences. Dr. Michalsky
has over 100 refereed publications in astronomy, solar energy, and atmospheric science.

Thomas Stoffel received his Bachelor of Science in aerospace engineering from the University of
Colorado–Boulder and an Master of Science in meteorology from the University of Utah. He began
his professional career as an aerospace engineer at the U.S. Air Force Propulsion Laboratory simu-
lating gas turbine engine performance and infrared radiation signatures. He  returned to school
to pursue his interests in radiative transfer and atmospheric science. Upon graduation, he worked
at what is now  Earth System Research Laboratory at the National Oceanic and Atmospheric
Administration to analyze urban–rural differences in solar radiation. In 1978, he began his career
in renewable energy at the Solar Energy Research Institute (now the National Renewable Energy
Laboratory) operated for the U.S. Department of Energy in Golden, Colorado. Progressing from
his start as an associate scientist to retiring as a principal group manager, he and his team of sci-
entists and engineers were responsible for the development and dissemination of solar resource

xxi
xxii Authors

information for the advancement of solar technologies and climate change research. This  effort
included the development of the Solar Radiation Research Laboratory, which continues to provide
research-quality solar resource measurements, radiometer calibrations, and systems for data acqui-
sition and quality assessment (https://www.nrel.gov/grid/solar-resource/renewable-resource-data.
html). He has authored or contributed to more than 80 publications addressing various aspects of
solar and infrared radiation measurements.
1 Measuring Solar and
Infrared Radiation
If the observation of the amount of heat the sun sends the Earth is among the most important
and difficult in astronomical physics, then it also may be termed the fundamental problem
of meteorology—nearly all whose phenomena would become predictable given the original
quantity of this heat and what kind it is.
Samuel Pierpont Langley (1834–1906)

Solar radiation measurements are the basis for understanding Earth’s primary energy source.
Studying the sun and applying this understanding has improved life on Earth since the dawn of
history. Builders have taken advantage of their knowledge of the solar resource to incorporate day-
light and space heating into their structures since the time of Greek city states, if not earlier. Today,
solar-generating facilities that are 500 MW in size are in operation and larger facilities are under
construction. The global photovoltaic market reached recent annual growth rates of almost 50 per-
cent (IEA PVPS 2017). By 2016, the installation of facilities with 402,000 MW of peak solar-gener-
ating capacity globally had the equivalent annual power output of 70–80 coal power plants (REN21
2018). The world is fitfully preparing for the solar age when oil will be depleted and environmen-
tally benign sources of energy will be needed to meet global energy demands sustainably. Today
a wide variety of solar technologies are used, and new and improved solar technologies are being
explored in the lab and tested in the field. The sun’s energy is free and available to everyone, and the
technologies that turn solar irradiance into useable energy are dropping in price as the solar indus-
try grows and production increases. Given the right financial structure, many solar technologies can
find cost-effective markets. For  example, a  cost-effective project to replace kerosene lamps with
photovoltaic (PV) panels and batteries in the Dominican Republic was successful (Perlin 1999).
Since 1985, the Luz Solar Energy Generation Stations (SEGS) plants in California, with 354 MW of
generating capacity, have produced electricity from solar thermal–electric power facilities (Kearney
1989; CEC 2018). The future for the solar industry is bright as sustainable energy generating facili-
ties are deployed to meet the energy demands for the growing world population while reducing the
deleterious effects of greenhouse gas emissions.
Part of the fundamental infrastructure that supports the growing solar industry includes solar
resource assessment and forecasting. As in hydroelectric generation where stream flow data are
necessary to design and operate the facility, knowledge of the temporal and spatial behavior of
the solar resource is critical to the design and operation of solar electric-generating facilities. With
multimillion and even billion-dollar solar projects under development, the need for long-term,
high-quality solar data is imperative for reducing the financial risks and costs. Solar data are needed
in the planning and design of the facility as well as for its operation and integration into the energy
mix. Testing new technologies to characterize and validate the design improvements requires the
highest-quality solar irradiance measurements because a 1–2 percent increase in energy conversion
efficiency can translate into millions of dollars in cost reductions for the power plant design and
operation. Developers and financial institutions are interested in the economic viability of projects
and the ability of the facility to service its debt in even the cloudiest years. Better solar resource
information results in more accurate estimates of a project’s performance that reduces uncertainty
and risk for the investors. Lower risks translate into lower interest rates that result in lower costs to
the ratepayers and more assured profits to developers and owners.

1
2 Solar and Infrared Radiation Measurements

The solar industry is not the only group interested in the solar resource. For decades, farm-
ers have used solar data to evaluate their irrigation requirements. Climate and atmospheric
scientists use solar resource data in their studies. Solar forecasts based on numerical weather
prediction and sky imagery are used now, although there is a need for considerable improve-
ment in this field (Haupt et al. 2018; Kurtz et al. 2017). Architects are using sunlight to reduce
energy demands on buildings and some buildings are constructed with a zero net energy bal-
ance (NRC 2011).
There  is a broad community of users of solar data and groups who supply this information.
Many of these groups are governmental agencies tasked to provide these data. Scientists are run-
ning solar measurement networks to study the climate and weather (ARM 2017; Augustine et al.
2000). Industry is prospecting for sites with “bankable” solar resource data to convince the finan-
cial community of proposed project viability. Private groups and individuals are making measure-
ments to evaluate the performance of solar systems. Some schools are integrating the solar resource
data into their science and engineering curricula (Tyukhov et al. 2008; Vignola and Savage 2002;
Vignola and Tyukhov 2007; Vignola et al. 2010). To accommodate the needs of this broad group of
interested parties with their diverse requirements, a wide variety of solar instruments is available.
The price and quality of these instruments vary widely and the more expensive equipment usually
provides more accurate results. This book provides an extensive survey of the solar instrumentation
that is available and attempts to provide general performance characteristics that can assist in the
selection of the appropriate instrument for the required task and provide information necessary to
assess the accuracy and quality of the data produced.
The goal of this instrumentation and measurements book is to pass on the experience that we
acquired over the years setting up solar monitoring sites, measuring the solar resource, analyzing
the data, and working with a wide variety of users in need of accurate solar resource information.
It takes a lot of effort and expense to build a quality solar radiation database. From our collective
experience we learned that proper documentation and archiving of the measured solar irradiance
and relevant meteorological data are critically important for the full and appropriate use of the data
when producing a reliable and defendable dataset. In addition, funding for solar resource assess-
ment historically comes in waves, and it takes considerable planning, ingenuity, and determination
to keep a measurement record intact to produce a long-term quality product that meets the needs of
today’s stakeholders and generations to come.
The main body of this book describes the solar irradiance sensors and how they function while
providing a description of the type of solar radiation they are designed to measure. Additional
information on solar resource assessment and related topics is contained in the appendices.
Chapter  2  describes the nature of solar radiation and provides the terminology and basic equa-
tions used in solar and infrared radiation measurements. A thorough understanding of this back-
ground information is needed to fully comprehend the information in the following chapters.
Chapter  3 is a history of the development of solar and infrared instrumentation. This  chapter is
useful for understanding the difficulty in obtaining accurate radiation measurements and shows
the considerable improvements made over the years. Some of these instruments are still in use
today, and understanding historical data requires knowledge of how these instruments operated and
performed. Those who want to use these older data should become familiar with the instruments
that provided the data and understand their limitations.
Chapters 4, 5, and 6 cover broadband solar radiation measurements, instrumentation, estimating
measurement uncertainty, and radiometer calibrations. Chapter  4 discusses instruments used to
measure direct normal irradiance (DNI) or “beam” irradiance (radiation coming directly from
the sun). Astronomical-based calculations tell us exactly where the sun is in the sky, and we can
calculate the coordinates and angles associated with the direct solar radiation to a high degree of
accuracy. This knowledge plus the fact that the highest energy density of solar radiation received
at the surface comes from DNI enhances the need to make sound DNI measurements. Chapter 5
discusses the instruments used to measure the total irradiance on a horizontal surface, which we
Measuring Solar and Infrared Radiation 3

refer to in this book as global horizontal irradiance (GHI). Due to their relative ease of use, GHI
irradiance sensors are by far the most common and GHI measurements are used for a wide variety
of applications (Sengupta et al. 2017). There are three types of instruments used to measure GHI,
and these are characterized and evaluated in Chapter  5. The  measurement of diffuse horizontal
irradiance (DHI) or “sky” irradiance is covered in Chapter 6. Although diffuse measurements use
the same type of instruments that are used to measure global irradiance, only a limited number of
these instruments are suitable for measuring diffuse irradiance. The reason for selecting certain
instruments for measuring DHI will be covered in more detail in Chapter 6.
The more challenging aspects of measuring and modeling the spectral distributions of solar radi-
ation are addressed in Chapter 7. Incident solar radiation is composed of different wavelengths of
light that interact with the changing atmosphere in complex ways that change the relative amounts
of ultraviolet, visible, and near infrared radiation reaching the surface. Many natural processes
from photosynthesis to human vision use different portions of the solar spectrum. In addition, the
performance of solar (more appropriately called photovoltaic or PV cells) cells is dependent on the
spectral composition of the incident solar radiation as well as the magnitude of the total incoming
irradiance. Instruments that measure the spectrum of incident solar radiation are expensive and
sometimes difficult to maintain. However, good spectral measurements are of great scientific value
as well as having practical applications. Relatively inexpensive instruments are available to measure
selected portions of the spectrum, for example that portion to which the eye is sensitive for daylight-
ing applications. Instruments that provide high-resolution spectral data over most of the spectrum
have considerably higher costs, but deliver a much more useful and flexible dataset.
Chapter 8 provides an overview of measuring the surface albedo (the ratio of reflected to incident
irradiance), which illustrates some of the complexities of how the Earth’s surface reacts to radiation
from the sun and sky. Albedo is an important consideration for models that utilize satellite-based
observations because the albedo determines the amount and nature of the surface-reflected radia-
tion. In addition, calculations of irradiance on tilted surfaces also depend on characterizing the sur-
face albedo. The amount of reflected global irradiance is dependent on the albedo of the surface in
front of the collector, and albedos can vary widely from 10 to 90 percent or more, depending on the
surface characteristics (e.g., black asphalt or fresh snow).
One can obtain information about radiation on a tilted surface by utilizing DNI, GHI, DHI,
and surface albedo measurements or measure the irradiance tilted surface directly. Chapter  9
addresses several key aspects of measuring total, or global, irradiance on a tilted surface—global
tilted irradiance (GTI). Most solar collectors are tilted toward the equator to better intercept the
incident solar radiation throughout the year. Some of the instruments suitable for GHI or DHI mea-
surements are unsuitable for GTI measurements. The problems associated with GTI measurements
are discussed, and the performance characteristics of sensors when tilted are described. This is an
important topic as solar system performance is typically gauged against measurements of tilted
surfaces. Chapter 10 introduces the unique characteristics of shadowband radiometers developed
to compute the DNI from measurements of GHI and DHI. The accuracy, advantages, and limita-
tions of this approach are discussed.
Measurement of infrared (IR) irradiance is covered in Chapter 11. The Earth’s surface and the
atmosphere emit IR radiation in proportion to their temperature and emissivity (i.e., fraction of
emitted radiation compared to a theoretical ideal blackbody). Special instruments are needed to
measure the incoming and outgoing IR irradiance because IR irradiance is a minor component of
the irradiance seen by broadband sensors used for pyranometers and pyrheliometers. Broadband
refers to sensors that respond to a wide range of wavelengths.
Instruments that measure net radiation are discussed in Chapter 12. Net radiation is the difference
between the incident solar and infrared radiation from the sun and sky minus the solar and infrared
irradiance reflected and emitted from the ground. When studying the Earth’s energy balance, which
drives weather and climate, good measurements of the net irradiance are essential. Net radiation is
measured many ways, and accurate measurements of net radiation are extremely difficult.
4 Solar and Infrared Radiation Measurements

Chapter 13 provides an introduction to the calibration of radiometers for measuring broadband


and spectral irradiance. Important concepts of calibration traceability to internationally recognized
reference measurement standards, accepted practices for performing radiometer calibrations, and
estimating calibration measurement uncertainties are presented for pyrheliometers, pyranometers,
pyrgeometers, photometers, and spectroradiometers.
Non-radiation meteorological measurements are covered in Chapter 14. Instruments that measure
temperature, pressure, relative humidity and precipitable water vapor, wind speed, and wind direc-
tion are discussed. These meteorological parameters can affect the performance of radiometers;
for example, both thermopile and photodiode radiometers are sensitive to the ambient temperature,
and a temperature correction can reduce the uncertainty of the measurements by reducing the tem-
perature bias in the measurement. Analysis of PV module and system performance is another use
of meteorological measurements. PV module performance mainly depends on the incident solar
radiation, but temperature affects power production.
Best practices for the operation and maintenance of solar monitoring stations are discussed in
Chapter 15 along with important factors to consider when siting, equipping, and constructing the
station. The  ability to produce quality solar and infrared radiation and relevant meteorological
parameters reliably is enhanced by a careful planning process. Even with the best instruments, a
well-designed solar monitoring station is essential to ensure the accuracy, validity, and complete-
ness of the data (Sengupta, 2017). Siting to minimize blockage by nearby obstructions and to facili-
tate maintenance of the station are two key factors. Good record-keeping and routine calibration
of the station instruments are also essential actions to ensure high-quality solar irradiance data. To
get full value from solar measurements, good design, adequate maintenance, and validation of the
incoming data are required. Since there are a limited number of quality ground-based solar moni-
toring sites and it takes many years to build a long-term solar radiation database, planners and devel-
opers heavily rely on satellite-derived solar radiation databases for their solar resource information.
Satellite-derived solar radiation databases are used extensively to survey locations for potential
solar electric facilities, and the number of agencies and companies that provide such services has
significantly increased. Chapter 16 discusses the characteristics, strengths, and weaknesses of solar
radiation values derived from satellite images and auxiliary measurements.
Appendices A to L cover various topics related to modeling solar irradiance, conversion factors,
and manufacturers’ information useful for solar monitoring. Elements of estimating measurement
uncertainty supporting the material in relevant chapters are found in Appendix A. Models used to
estimate solar radiation are discussed in Appendix B. These elements include correlations between
various irradiance components and models that are used to estimate irradiance on tilted surfaces.
Appendix B presents the most common models and gives an idea of how they work. These mod-
els also are useful in understanding why certain measurements are made and why they are useful.
Appendix C introduces the concept of measuring sunshine duration and its applications. Sunshine
duration was one of the first proxies for estimating solar irradiance, and these measurements are still
made in many parts of the world. Although superseded by actual measurements of solar irradiance,
it is included for completeness. Appendix D presents a series of sun path charts for various latitudes
that provides a quick reference to the sun path at a given site and illustrates the location of the sun at
any given time of the day and year. Appendix E discusses the algorithms used to calculate the posi-
tion of the sun. Knowing the solar position is always useful when analyzing solar irradiance data,
designing a solar monitoring station, or applying solar radiation data. Many different units have been
used over the years to measure solar radiation and so Appendix F contains a table of useful conver-
sion factors. This table provides information on how to convert from one unit to another. A list of
solar instrument manufacturers is given in Appendix G. This list is fairly extensive and includes the
major sources of solar instrumentation and provides a basis for searching for commercially available
instruments. A sample calibration report from the National Renewable Energy Laboratory is given
in Appendix H. This Broadband Outdoor Radiation Calibration (BORCAL) report illustrates what
is included in a comprehensive calibration of a solar sensor. Some helpful troubleshooting hints for
Measuring Solar and Infrared Radiation 5

identifying common problems in radiometry and how to address them are available in Appendix I.
For the educator or do-it-yourself person seeking a hands-on project, Appendix J includes instruc-
tions for building a simple pyranometer based on readily available components. A comprehensive
data archive format including key metadata elements is in Appendix K. Lastly, in Appendix L, a list
of selected sources of solar and infrared radiation measurements is presented.

QUESTIONS
1. Name two uses of solar radiation data.
2. Describe the difference between broadband and spectral radiation and their applications.
3. Why do you want to learn about solar measurements?
4. Who would want more accurate solar data, a farmer scheduling his irrigation or a banker
financing a solar concentrating power plant? Explain your answer.
5. To which chapter or appendix in this book would you go to find a way to calculate the sun’s
position at any time?

REFERENCES
Atmospheric Radiation Measurement (ARM). 2017. ARM Climate Research Facility Annual Report-2016,
DOE/SC-ARM-16-056. https://www.arm.gov/publications/annual-reports/docs/doe-sc-arm-16-056.pdf
(Accessed February 2, 2018).
Augustine, J. A., J. J. DeLuisi, and C. N. Long. 2000. SURFRAD—A national surface radiation budget net-
work for atmospheric research. Bulletin of the American Meteorological Society 81 (10): 2341–2357.
California Energy Commission (CEC). 2018. Solar energy projects in California. http://www.energy.ca.gov/
sitingcases/solar/ (Accessed May 25, 2018).
Haupt, S.E., B. Kosoviç, T. Jensen, J.K. Lazo, J.A. Lee, P.A. Jiménez, J. Cowie, et  al. 2018. Building the
Sun4Cast System: Improvements in solar power forecasting. Bulletin of the American Meteorological
Society 121–135. doi:10.1175/BAMS-D-16-0221.1.
International Energy Agency Photovoltaic Power Systems (IEA PVPS) Program. 2017. Trends 2017 in photo-
voltaic applications, executive summary, report IEA PVPS T1-32-2017. http://www.iea-pvps.org/index.
php?id=92&eID=dam_frontend_push&docID=4249 (Accessed February 6, 2018).
Kearney, D. 1989. Solar electric generating stations (SEGS). IEEE Power Engineering Review 9 (8): 4–8.
doi:10.1109/MPER.1989.4310850.
Kurtz, B., F. Mejia, and J. Kleissl. 2017. A  virtual sky imager testbed for solar energy forecasting, Solar
Energy 158:753–759.
National Research Council (NRC). 2011. Achieving High-Performance Federal Facilities: Strategies and
Approaches for Transformational Change. Washington, DC: National Academies Press.
Perlin, J. 1999. From Space to Earth: The Story of Solar Electricity. Cambridge, MA: Harvard University
Press.
REN21. 2018. Renewables 2018 global status report. Paris, France: REN21 Secretariat. http://www.ren21.net/
status-of-renewables/global-status-report/ (Accessed May 22, 2019). ISBN 978-3-9818911-3-3
Sengupta, M., A. Habte, C. Gueymard, S. Wilbert, and D. Renné. 2017. Best Practices Handbook for the
Collection and Use of Solar Resource Data for Solar Energy Applications. 2nd ed. Golden, CO:
Technical Report NREL/TP-5D00-68886.
Tyukhov, I., M. Schakhramanyan, D. Strebkov, S. Mazanov, and F. Vignola. 2008. Combined solar PV and
earth space monitoring technology for educational and research purposes. Proceedings of the 37th
ASES Annual Conference. San Diego, CA.
Vignola, F. and I. Tyukhov. 2007. Discussion of PV lab equipment and photovoltaic systems for teaching the
science of photovoltaics. Proceedings of the 36th ASES Annual Conference. Cleveland, OH.
Vignola, F. and J. Savage. 2002. How BPA/WSUN school PV effort worked at Elmira High School.
Proceedings of the Solar 2002. Reno, NV: American Solar Energy Society Conference.
Vignola, F., S. Miklavzina, S. Daniels, M. Toamina, S. Thorin, I. Tyukhov, and A. Tikhonov. 2010. Energizing
the next generation with photovoltaics. Proceedings of the Solar 2010. Phoenix, AZ: American Solar
Energy Society.
2 Definitions and Terminology

Everything should be made as simple as possible, but not simpler.


Albert Einstein (1879–1955)

2.1 INTRODUCTION
This  chapter introduces basic terminology and concepts needed to understand the measure-
ment of solar and infrared radiation. In addition, there are several information sources offering
various glossaries, standard units and symbols, definitions, and terminology commonly used in
meteorology, climatology, radiometry, solar energy conversion, and related fields (ASTM 2015;
Beckman et al. 1978; Duffie et al 1991; IEC 2018; ISO 1999; Paltridge and Platt, 1976; WMO 1992,
2017) that present similar information from different viewpoints. After a brief historical perspec-
tive on the importance of observing the Sun, a methodology is provided to calculate the position
of the Sun in the sky while illustrating and explaining the specialized vocabulary and parameters
employed in the calculation. To facilitate understanding the nature and characteristics of solar
radiation, the incoming solar radiation is separated into principal solar components that result
from the Sun’s radiation interacting with the Earth’s atmosphere. These components along with
the rational for this separation are described. Then the spectral nature of solar radiation and other
concepts relevant to instrument design and operation are briefly discussed.
The Sun has fascinated humans since the dawn of recorded history. Striving to understand the
heavens and the stars led some civilizations to worship spirits that cause the rebirth of the Sun
every morning and the setting of the Sun at night. Ancient astronomers created solar calendars to
track the changing path of the Sun across the sky. Primitive civilizations built monuments such as
Stonehenge to mark the equinoxes and to give clues to changing seasons.
This fascination was driven not by idle curiosity, but by real needs to know when to prepare for
planting and when to hold rituals to ensure a good harvest or to celebrate holy days. More modern
societies, such as the Greeks, had laws protecting access to the Sun. The Anasazi in the southwest-
ern United States designed their cliff-side dwellings to make optimal use of the solar resource to
make their living space more comfortable.
Before going into modern-day instruments used to measure incoming solar radiation, it is worth-
while to discuss a few basics concepts about the Sun, how it presents itself in the sky, and the termi-
nology commonly used to describe solar irradiance. In addition, concepts and fundamentals of heat
transfer and the fundamentals of photodiodes are discussed to serve as the basis for discussions of
solar radiation sensors in later chapters.

2.2 THE SUN


The Sun is considered by astronomers to be a yellow dwarf star somewhere in the middle of its life
cycle. Nuclear fusion deep within the solar interior powers the Sun as hydrogen combines to form
helium that releases an immense amount of energy. Gravity keeps the Sun from blowing apart, and
the nuclear fusion cycle has kept the Sun radiating energy into space at a fairly constant rate for bil-
lions of years. The effective temperature of the Sun is 5772 K (5499°C or 9930°F) (NASA 2018). In
other words, the radiation from the Sun is similar to, although not exactly as, a blackbody heated to
5772 K (Figure 2.1). The solar power over all wavelengths reaching the top of atmosphere (TOA) of
the Earth is called the total solar irradiance (TSI), which varies by about 0.1 percent in proportion

7
8 Solar and Infrared Radiation Measurements

Extraterrestrial (Top of Atmosphere (TOA))

FIGURE 2.1 Solar spectral distributions for a blackbody temperature of 5772 K at the top of the atmosphere
(TOA). ASTM Standard G173-03 for extraterrestrial solar radiation and direct normal plus circumsolar irradi-
ance at sea level illustrating the known atmospheric absorption bands.

to the changing number of sunspots over the 11-year solar cycle (de Wit et al. 2017; Fröhlich, 2006;
Kopp 2016). Recently, TSI replaced the term solar constant (I0), acknowledging the very slight
variation in the Sun’s output.

2.3 EXTRATERRESTRIAL RADIATION


Solar radiation incident just above the Earth’s atmosphere is called extraterrestrial radiation.
The currently accepted value of the TSI at a mean Sun-Earth distance of 149,598,106 kilometers
(km) (92,955,88 miles (mi)) is 1360.8 ± 0.5 watts/square meters (Wm–2), or 0.035 percent, averaged
over Solar Cycle 23 during 1996 to 2008 (Kopp and Lean 2011; Kopp 2016). Virtually, all of this
energy falls between 280 and 4,000 nanometers (nm).
The eccentricity of the Earth’s elliptical orbit is 0.0167 and results in the Earth being closest to
the Sun (perihelion) in early January and farthest from the Sun (aphelion) in early July. The Earth
and the Sun are 1.67 percent closer at perihelion and 1.67 percent farther apart at aphelion than their
mean separation. Because the Sun can be considered a point source of radiation due to its distance
from the Earth, the solar intensity is inversely proportional to the square of the distance from the
Sun. The extraterrestrial irradiance impinging on the Earth varies over a range of 6.7 percent as
the Earth orbits the Sun. The extraterrestrial irradiance outside the Earth’s atmosphere on a surface
normal to the Sun is

DNI 0 = TSI ⋅ ( Rav / R ) 2 Wm−2  (2.1)


Definitions and Terminology 9

where Rav is the mean Sun-Earth distance, and R is the actual Sun-Earth distance depending on the
day of the year. An approximate equation for (Rav /R)2 is

( Rav /R ) = 1.000110 + 0.034221 ⋅ cos ( β )


2

+ 0.001280 ⋅ sin ( β ) + 0.000719 ⋅ cos ( 2β ) (2.2)

+ 0.000077 ⋅ sin ( 2β )

where β = 2π(d–1)/365 radians, and d is the day of the year (Spencer 1971). For example, January 15
is the 15th day of the year, and February 15 is the 46th day of the year. Since there are 366 days in a
leap year, there is a slight difference in the calculation in leap years. To calculate the extraterrestrial
global horizontal irradiance (GHIo), extraterrestrial direct normal irradiance (DNIo) is multiplied
by the cosine of the solar zenith angle (see Equation 2.7 and Figure 2.7). Information on calculating
hourly, daily, or monthly average DNIo and GHIo are given in Appendix B.

2.4 SOLAR COORDINATES


The Earth’s axis is tilted approximately 23.44 degrees with respect to the plane of the Earth’s orbit
around the Sun. As the Earth moves around the Sun, its rotational axis is fixed if viewed from space
(Figure 2.2). In June, the orientation of the axis is such that the Northern Hemisphere is pointed
toward the Sun. In December, the Earth’s position is at the opposite end of the orbit around the Sun,
and the orientation of the Earth’s axis is such that the Northern Hemisphere is pointed away from
the Sun. During the spring and fall equinoxes, the Earth’s axis is perpendicular to a line drawn
between the Earth and the Sun. The mean geometric orbit of the Earth around the Sun defines the
ecliptic plane. For an eclipse to occur, the orbit of the new (solar) eclipse or full (lunar) eclipse moon
has to be very near to the ecliptic plane, hence the name.
From an Earth-centered point of view, the Sun moves among the other stars in the heavens
on the celestial sphere (see Figure 2.3). The observer on Earth is stationary, and the objects in
the heavens look as if they are on the celestial sphere rotating at a rate of just under 360 degrees
per day (360°/d) or, equivalently, 15 degrees per hour (15°/h). The  plane running through the

FIGURE 2.2 The Earth circles the Sun in an elliptical orbit. The tilt of the Earth’s axis with respect to the
orbital plane is about 23.44 degrees. As the Earth orbits the Sun, the Northern Hemisphere tilts toward the
Sun in June and away from the Sun in December. On average the Earth rotates just under once every 24 hours.
(Courtesy of Angie Skovira of Luminate, LLC.)
10 Solar and Infrared Radiation Measurements

FIGURE 2.3 Illustration of the celestial sphere. The celestial equator or the equatorial plane is the projection
of the Earth’s equator extended to space. The Earth’s rotation axis extended to space forms the north and south
celestial poles. Angular distances above or below the equatorial plane measured from the center of the Earth
are called declination. Angles (measured in time units) from the point where the equatorial plane and the
orbital plane of the Earth around the Sun cross are the equinoxes. 0 hours is assigned to the spring (or vernal)
equinox. The illustration indicates the first day of summer with the Sun’s declination at 23.44 degrees and a
right ascension of 6 hours. (Courtesy of Chris Cornwall.)

Earth’s equator is designated the celestial equator (or equatorial plane), and the (celestial) polar
axis, identical to Earth’s polar axis, is the axis of rotation of the celestial sphere around the
Earth. This plane and axis extend to infinity. Similar to the fact that any location on Earth can
be specified by the latitude and longitude of the location, the position of any star on the celestial
sphere—the Sun in this discussion—can be determined by specifying two parameters: the right
ascension and the declination. The difference is that the locations on Earth are fixed while the
location of the stars change over the day and over the year as the Earth spins on its axis and takes
its journey around the Sun. For the observer on Earth, the stars or Sun look like they are moving
in a plane that is parallel to the equatorial plane at a rate of about 15°/h.
The right ascension of the Sun changes from 0° to 360°, or 0h to 24h in astronomical terminol-
ogy, as a consequence of the fact that the Earth orbits the Sun once a year. The two points on the
equatorial plane where the plane of the Earth’s orbit around the Sun—the ecliptic—are coincident
with the equatorial plane and are called the equinoxes. At the equinoxes the declination is equal
to zero. The  right ascension of the Sun is zero at the vernal equinox. Angular measurements,
Definitions and Terminology 11

called right ascension, are made in the equatorial plane as measured, counterclockwise, from
the vernal equinox point. The  right ascension of the Sun is relative to the Earth’s orbit around
the Sun but should be considered from the perspective of the observer on Earth where the Sun
appears to rotate around the Earth. Historically, the angular position of the right ascension on the
equatorial plane has been measured in hours since the Earth rotates 360°/d (24 h) or 15°/h similar
to the Earth orbiting the Sun. The  yearly 365-day journey completes a 360 degree orbit that is
historically divided into 24 hours. The right ascension of the Sun is 6 hours (90°) from the vernal
equinox at the summer solstice, 12 hours (180°) at the autumnal equinox, and 18 hours (270°) at
the winter solstice. The position of the Sun on the celestial sphere is shown at the summer solstice
in Figure 2.3.
An angular measurement above or below the equatorial plane toward the celestial polar axis is a
measurement of the Sun’s declination. The Sun is above the equatorial plane in summer, below it in
winter, and exactly on the plane on the first day of spring and on the first day of autumn. An approx-
imate formula for the declination of the Sun in degrees is (Spencer, 1971):

0.006918 − 0.399912 ⋅ cos ( β ) + 0.070257 ⋅ sin ( β ) 


 
dec = 
 (
− 0.006758 ⋅ cos ( 2β ) + 0.000907 ⋅ sin ( 2β )  ⋅ 180 π
 ) (2.3)
 
 − 0.002697 ⋅ cos ( 3β ) + 0.00148 ⋅ sin ( 3β ) 

where β is defined as 2π(d–1)/365 radians, and d is the day of the year. For example, February 20 is
the 51st day of the year. Since there are 366 days in a leap year, there is a slight difference in the
calculation for leap years.
To an observer on the Earth at 40°N latitude, the celestial sphere is tilted as in Figure 2.4. Note
that the closer the observer is to the equator, the closer the equatorial plane is to the zenith, or
equivalently the east and west horizon. And on the spring equinox and the autumn equinox for an
observer at the equator, the Sun will pass directly overhead. Now the task is to relate the celestial
position of the Sun to a local coordinate system defined for an observer’s position on the Earth’s
surface. The local meridional plane is defined by three points: due south, due north, and the zenith.
The meridional plane runs through the Earth’s axis of rotation and is perpendicular to the local
east–west axis and is the plane defining the local longitude of the site. As the celestial sphere rotates
around the Earth, the time during the rotation when the local meridional plane bisects the Sun is
called solar noon and is the moment when the Sun has an hour angle value of zero. Solar hour angle
uses time units to measure how much time the Sun is from reaching solar noon or how much time
the Sun is past solar noon. As mentioned earlier, the celestial sphere rotates at about 15°/h, and an
hour angle of 1 hour represents a rotation of 15 degrees. The hour angle is positive after solar noon
and negative before solar noon. In Figure 2.4, it will take about 2 hours for the celestial sphere to
rotate before the Sun is at solar noon or equivalently reaches the local meridian. The hour angle of
the Sun in Figure 2.4 is, therefore, approximately –2 hours (see Equation 2.9 for computing solar
hour angle values).
Before these celestial coordinates can be related to a more convenient local coordinate system,
a time adjustment factor known as the equation of time (EoT) needs to be discussed. Refer again to
Figure 2.2. Let’s assume that the Earth orbits the Sun in a perfect circle. The Earth rotates counter-
clockwise on its axis at a virtually constant rate because of the conservation of angular momentum.
A solar day is defined by the time it takes for the Sun to appear at the same point in the sky. This is
the time to complete one rotation plus a little more; this extra rotation is because the Earth rotates
counterclockwise with respect to the North Pole and also moves counterclockwise 1/365.25th of the
way around its orbit meaning that the Sun appears to have moved in the heavens from the previous
12 Solar and Infrared Radiation Measurements

FIGURE 2.4 Relation of solar angles from the observer’s point of view as seen in the Northern Hemisphere
at 40°N. The north and south directions are in the plane of the page. The east and west directions are out of
and into the page, respectively. The zenith line is perpendicular to the Earth’s surface. The local meridional
plane is in the plane of the page and is defined by the south, north, and zenith points. The equatorial plane is
perpendicular to the polar axis about which the celestial sphere rotates. In the northern hemisphere, the solar
hour angle, ha, is the angle between the Sun’s current position and solar noon that is due south and measured
along the equatorial plane in hours (about–2 hours in this figure). (Courtesy of Chris Cornwall.)

day by this much. One day equals 24 hours or 1,440 minutes, and 1/365.25th of 1,440 minutes is just
under 4 minutes. Note that a sidereal day is defined by the time it takes the Earth to complete one
rotation, and it is about 4 minutes shorter than a solar day of 1,440 minutes.
In fact, the Earth’s orbit around the Sun is an elliptical path that has the Sun at one of the foci.
Kepler’s second law of planetary motion states that the area defined by the sweep of a line between
the Sun and the Earth will be the same for the same increment of time. Figure 2.5 illustrates this; for
the A and B wedges to be equal in area, the Earth must move more quickly when it is closer to the
Sun. If we defined the day by the time it takes the Sun to appear at the same point in the sky each
day, say, due south, then the day would have different lengths at different points in its orbit around
the Sun. The day would be longer when the Sun and Earth were closest (perihelion) compared with
when they are the farthest apart (aphelion). This is because the Earth must turn a complete revolu-
tion and a little more as it progresses in its orbit around the Sun, and this progression around the
Sun proceeds at a faster clip when the two bodies are closer together. Since a constant day length
is preferable, 24 hours is set for civil time. This set time of 24 hours will cause the Sun to be at a
Definitions and Terminology 13

FIGURE 2.5 Illustration of the elliptical orbit of the Earth around the Sun with the area labeled A equal in
size to the area labeled B. According to Kepler’s second law, the time to traverse the arc bounding A is equal
to the time it takes to traverse the arc bounding B. Therefore, the Earth has to travel faster when it is closer to
the Sun than when it is farther from the Sun.

different position relative to our noon clock each day. This apparent repositioning of the Sun with
respect to local time is defined by the EoT. Spencer’s (1971) approximation for the EoT in radians is

0.000075 + 0.001868 ⋅ cos ( β ) − 0.032077 ⋅ sin ( β ) 


EoT =   (2.4)
 − 0 . 014615 ⋅ cos ( )
2 β − 0 . 040849 ⋅ sin ( )
2 β

where β is defined as 2π(d–1)/365 radians, and d is the day of the year as defined after Equation 2.3.
To change to minutes from radians, multiply by 1440/2π = 229.183.
The calculation provides the number of minutes that the Sun is behind or ahead of the local meridian
at solar noon. For example, the local meridian at Denver, Colorado, is about 105°W, which is 7 hours west
of Greenwich, England (0° longitude). On January 1 Equation 2.4 produces a result of –2.90 minutes,
meaning that the Sun will appear due south at its highest point in the sky for that date 2.90 minutes after
noon local time. In Boulder, Colorado, 0.25 degrees west of Denver, it will take 1 minute more for the
Sun to reach Boulder’s local meridian since the Sun moves 1 degree in 4 minutes—that is, 24 hours is
equal to 1,440 minutes, so 1,440 minutes/360 degrees is equal to 4 minutes per degree. Figure 2.6 is
called the analemma that results from the EoT and the declination changes over a year. It represents the
Sun’s position over a year for a fixed location on Earth at the same time every day.
A more accurate (±0.01°) algorithm for calculating the Sun’s declination and EoT is given in
Michalsky (1988), and a FORTRAN program can be found in Appendix E. A “C” program, solpos.c,
based on this algorithm from the National Renewable Energy Laboratory (NREL) Renewable
Resource Data Center (RReDC) website https://www.nrel.gov/grid/solar-resource/solpos.html.
The solpos.c code is easier to understand but is less accurate than the solar position program
(spa.c) (Reda and Andreas 2008) (available at https://rredc.nrel.gov/solar/codesandalgorithms/spa).
The latter contains matrix tables that make the algorithm more difficult to follow but provides a
stated accuracy of ±0.0003° over the period –2000 to 6000. More information on the solar position
algorithms is found in Appendix E.
14 Solar and Infrared Radiation Measurements

FIGURE 2.6 Solar analemma. Plot of solar declination verses EoT. This plot has the general shape of the
solar position in the sky if a photograph was taken on each day of the year at the same time of day.

2.5 ZENITH, AZIMUTH, AND HOUR ANGLES


A coordinate system that can be used for defining positions with respect to an observer fixed on the
surface of the Earth is shown in Figure 2.7. The solar zenith angle is defined as the angle between
the zenith and the Sun. The cosine of the solar zenith angle with respect to the observer on the
surface of the Earth can be obtained by calculating the vector product of the normal to the surface
and the unit vector coming directly from the Sun. Consider a reference coordinate system with the x
and y axes defined in the equatorial plane with the east–west axis for the x axis and the intersection
with the local meridian plane for the y axis. The celestial polar axis is the z axis. The normal to the
surface at a location of latitude (lat) degrees is then

elv is the elevation angle, measured up from horizon


sza is the zenith angle, measured from vertical
az is the azimuth angle, measured clockwise from north

n = cos ( lat ) y + sin ( lat ) z (2.5)

The unit vector to the Sun is tilted from the equatorial plane by the declination (dec) with an hour
angle (ha) (see Equation 2.9). Projecting the Sun’s vector onto the three reference axes yields

s = cos ( dec ) sin( ha) x + cos ( dec ) cos ( ha ) y + sin ( dec ) z (2.6)

The dot product of the unit vectors is the cosine of the angle between the two vectors. The angle
between the normal to the surface and the Sun’s direction is the solar zenith angle (sza). Therefore,

cos( sza) = sin(lat ) ⋅ sin( dec) + cos(lat ) ⋅ cos( dec) ⋅ cos( ha) (2.7)
Definitions and Terminology 15

FIGURE 2.7 View from the local horizontal plane. The solar zenith angle, sza, is measured from the zenith
position to the Sun. The solar elevation, or altitude, angle, elv, is measured from the horizon to the Sun. The azi-
muth angle, az, is measured from due north moving eastward from 0° to 360°. (Courtesy of Chris Cornwall.)

Astronomical telescopes generally operate with their axes of rotation oriented parallel to the Earth’s
axis of rotation. If the telescope is pointed at the declination of the star and then moved in hour
angle to the right ascension of the star, a simple sidereal clock drive can keep the star in the field
of view of the telescope by rotating in one direction from east to west at the Earth’s rotation speed.
Some telescopes and most automated solar trackers do not use equatorial mounts for tracking the
Sun but use alt-azimuth trackers instead. Solar azimuth (az) is measured from the north clockwise
toward the east and varies from 0° to 360°. The solar azimuth is shown in Figure 2.7. Altitude or,
more commonly, elevation (elv) is measured from the horizon to the solar position (0° to 90°) at a
fixed azimuth. The solar zenith angle, sza, is the complement to elevation so sza = 90°–elv.
Unlike stars, the Sun’s right ascension is constantly changing over the year from 0 hours the first
day of spring to 6 hours on the first day of summer, and so on (see Figure 2.4). Using only universal
time (UT is explained in Section 2.6), the solar declination and right ascension can be calculated.
Then using the site’s latitude and longitude, the solar position can be calculated in terms of hour
angle and solar zenith angle (Equation 2.7). It can alternatively be calculated in terms of solar zenith
angle and azimuth angle. The hour angle and declination can be used to point an equatorial-mounted
instrument, and the zenith angle and azimuth angle can be used to point an alt-azimuth-mounted
solar device. These calculations are performed in solpos.c and spa.c, described in Section 2.4, with
accuracies of ±0.01° and ±0.0003°, respectively.

2.6 SOLAR, UNIVERSAL, AND LOCAL STANDARD TIME


Algorithms are easier to formulate and use in solar time, but it is important to relate the solar position
and descriptions to the local standard time (LST). LST is different from clock time because clock
time is typically advanced by 1 hour in the summer to provide daylight saving time. LST is equal to
clock time in the winter and is not changed to daylight saving time in the summer. LST is the same
across the entire time zone, whereas solar time is related to the position of the Sun with respect to
the observer with solar noon as the time when the Sun crosses the local meridian (i.e., true south
when the solar azimuth is 180°). The relationship between LST and solar time is dependent on the
16 Solar and Infrared Radiation Measurements

observer’s exact longitude and time zone. The Earth rotates just under once every 24 hours; hence,
the Sun appears to travel across the sky at a rotation rate of about 15°/h. Solar angles are calculated
using Greenwich mean time (GMT). However, the Earth’s rotation is very gradually slowing, and
atomic clocks are now used for time standards. Coordinated universal time would be abbreviated
CUT in English and TUC in French (temps universal coordonné); however, neither abbreviation is
used, and coordinated universal time, designated UTC, is the international time standard based on
an atomic clock. UTC uses leap seconds to adjust for the ever-so-gradual slowing of Earth’s rotation
and to coordinate with the mean solar time or GMT or universal time (UT). UT and UTC are always
within a second of each other but have the slight difference previously described. The Sun moves
about 0.004 degrees in 1 second, so this difference generally is ignored.
Local standard time is related to GMT through the time zone of a person’s location. In Eugene,
Oregon, the local time is 8 hours earlier than GMT. In Berlin, Germany, the local time is 1 hour
ahead of GMT. Every 15 degrees of longitude roughly defines a time zone. Therefore, if a person’s
location is within ±7.5 degrees longitude of a longitude that divides evenly by 15 degrees, that per-
son would be assumed to be within the time zone of that longitude; however, a location’s time zone
is sometimes modified from this strict interpretation by other factors, usually related to commerce.
In any case, correctly specifying the longitude, the UT and the latitude allows for calculation of
solar position very precisely using the programs solpos.c or spa.c that are described in Section 2.4.

2.7 SOLAR POSITION CALCULATION EXAMPLE


An example is provided to show how to calculate the Sun’s position at a given location. Assume
the location is Omaha, Nebraska. It is 9:30 a.m. Central Daylight Time (CDT) in the morning on
September 13 in a non-leap year, for example 2017. The latitude of the city center is about 41.25°N;
the longitude is 96.00°W. Omaha is in the Central Time (CT) zone (area that observes either Central
Standard Time (CST) or CDT) of the United States, which is 6 hours earlier than Greenwich mean
time or UT. Since it is still during daylight savings time, the local standard time is 8:30 a.m. September
13 is day 256 of the year. Using Equations 2.3 and 2.4 to calculate declination and the equation of
time, the declination of the Sun is found to be 4.11° and the EoT +3.90 minutes for this date. The solar
hour angle is obtained by calculating time before solar noon. Solar noon is determined by using
the EoT. On September  13, the Sun reaches solar noon +3.90  minutes before noon LST for the
time zone around the longitude of 90°. Omaha is 6 degrees farther west than the center longitude
(also called the standard meridian) 90° for the CT zone and the time it takes for the Sun to be at
the solar noon position has to be included. Therefore, solar noon on this day is 12:00–3.90 minutes
+ (6 degrees · 4 minutes/degree) minutes or 12 hours and 20.1 minutes (12:20:06). The hour angle
at 8:30 a.m. LST at Omaha on September 13 is 8 hours + 30 minutes – (12 hours + 20.1 minutes)
or –3 hours – 50.1 minutes.
The relationship between LST (lst) and solar time (st) is

ls = lst + EoT + (longob − long sm ) / 15 (2.8)

where longsm is the longitude of the standard meridian from which the clock time for the time zone
is assigned, and longob is the longitude of the location of the observer. The units for this equation are
hours. If EoT is calculated in minutes, then it needs to be divided by 60 minutes/hour to get the
proper units. The hour angle in degrees is found by

ha = 15 ⋅ ( st − 12 ) or st = ha15 + 12 (2.9)

When solar time is noon (12:00), the hour angle is zero.


Now an equatorial mount solar tracker can align with the Sun using the solar coordinates 4.11°
declination and an hour angle of –3.835 hours (or –57.525° or –1.0040 radians). For executing the
Definitions and Terminology 17

sine and cosine functions, the angular arguments to these functions must be used correctly. Most
often programming languages use radians for trigonometric calculations. To set an alt-azimuth
tracker using the solar azimuth and zenith angles, local coordinates need to be used.
With the hour angle, latitude, and declination, one can now  calculate the cosine of the solar
zenith angle (sza) using Equation 2.7. The lat is the latitude of central Omaha 41.25°, dec is the
declination of the Sun with respect to the Earth calculated from Equation 2.3, which is 4.11° in this
example, and ha is the hour angle at 8:30 central standard time on September 13 as determined
already. The result is a solar zenith angle of 63.26°. The azimuth angle can also be determined from
the zenith angle, the hour angle, and the declination of the Sun.

− cos ( dec ) ⋅ sin ( ha )


sin ( azm ) = (2.10)
(
cos π 2 − sza )
The arcsine has more than one solution, and therefore it is necessary to exercise the following logic
to obtain the right solution between 0° and 360°:

if sin ( dec ) − sin (π / 2 − sza ) ⋅ sin ( lat )  > 0

then if sin ( azm ) < 0  azm = azm + 360

else azm = 180 − azm

For  this case azimuth is found to be 109.57°. Alternately, running solpos as mentioned in
Section  2.4  yields a solar zenith angle of 63.56° and a solar azimuth of 110.06°. These latter
results are more accurate, but they are consistent with the lower accuracy approximations given by
Equations 2.3 and 2.4.
To view the motion of the Sun across the sky during different times of year, sun path charts
can be plotted. These sun path charts are two-dimensional plots of solar elevation versus solar
azimuth and several solar tools use them to evaluate how objects around a site will cast shadows
on instruments or solar energy systems. A sun path chart for latitude 35°N is shown in Figure 2.8.
The time lines are plotted in solar time, but it is also possible to plot them in LST. Sun path charts
can be created in polar coordinates (Figure 2.9). If a person was standing in the center of this
polar coordinate plot and looking toward the equator, the Sun would rise behind the person’s
back, to the north and east in this example, and set behind the person’s back to the north and west
for 6 months of the year. Three-dimensional sun path charts are also useful in showing the path
of the Sun across the sky from an observer’s perspective (see Figure 2.10). For locations above
the Arctic Circle or below the Antarctic Circle, the Sun will never rise or set during parts of the
year. Slightly north of the Arctic Circle, say latitude 70°, during parts of May and June, the Sun
never sets, and during parts of December and January the Sun never makes it above the horizon.
More sun path charts for every 10 degrees of latitude from –90° to 90° are shown in Appendix
D. Site-specific sun path charts can be created online for any location on Earth (http://solardata.
uoregon.edu/SunChartProgram.php).
Another way to examine the path of the Sun is to look at the shadow cast by a vertical pole
or tree during the day. The curve traced is much like that generated by the gnomon of a sundial
(see Figure 2.11). The curves trace the end of the shadow cast by the top of a 10-meter pole during
different times of year. Near sunrise and sunset, the shadows can be extremely long as the Sun is
very low in the sky. During the middle of the day, the Sun casts the shortest shadow of the day on a
horizontal surface. The website that generates the sundial plots can be found online (http://solardata.
uoregon.edu/SunDialProgram.html). The sun path charts and sundial plots are useful when plan-
ning and constructing a solar monitoring station because the effects of shadows cast by surrounding
18 Solar and Infrared Radiation Measurements

FIGURE 2.8 Sun path chart for latitude 35°N plotted in solar time. The azimuth angle varies from 60° at
sunrise to 300° at sunset on June 21 and varies from 120° at sunrise to 240° at sunset on December 21. At the
equinoxes, the solar zenith angle at sunrise is at 90°, due east, and 270°, due west, at sunset. At solar noon,
the solar elevation angle varies from 78.44° on June 21 to 31.59° on December 21. The solar elevation is 0° at
sunrise and sunset. The solar zenith angle is 90° minus the solar elevation.

FIGURE 2.9 Sun path chart for latitude 45°N in polar coordinates. Hour lines are plotted in local solar time.
As the Sun moves closer to the center of the plot, it is higher in the sky. If the zenith angle was zero, the Sun
would be directly overhead. For the Sun to be directly overhead, a person would have to be at a location on
Earth between 23.44° latitude and –23.44° latitude at the right time of year.
Definitions and Terminology 19

FIGURE 2.10 Three-dimensional plot of the Sun for latitude 23.44°N. At solar noon on the summer solstice,
the Sun is directly overhead, yielding a shadowless solar noon. The sun paths are for summer solstice, spring
and fall equinoxes, and winter solstice. (Courtesy of Chris Cornwall.)

FIGURE 2.11 Sundial plot for a latitude of 45°. The lines follow the end of the shadow from the tip of a
10-meter-tall pole located at the (0,0) coordinate. Near sunrise or sunset, the distance to the end of the shadow
will be extremely large. In the summer when the Sun is highest in the sky, the shadow will fall closest to the
pole or gnomon. In the winter, when the Sun is lowest in the sky, the shadow at solar noon will have the longest
noontime reach. The solar time marks on the chart are for reference.
20 Solar and Infrared Radiation Measurements

objects can be studied. Prudent use of the sun path programs can identify the location with minimal
shading and identify times when shading will affect the monitoring equipment. The importance of
site surveys before establishing a solar monitoring station is discussed in chapter 15 on setting up a
solar monitoring station.

2.8 SUNRISE AND SUNSET TIMES


The calculation of sunrise and sunset times provides an easy exercise to test understanding of the
information presented so far. Sunrise and sunset occur when the Sun is at the horizon, the solar
zenith angle is 90°, or π/2 radians, and hence the cosine of the solar zenith angle is zero.
Setting the cosine of the solar zenith angle to zero in Equation 2.7 results in

hasunrise , sunset = ± cos −1  − tan ( lat ) ⋅ tan ( dec )  (2.11)

where hasunrise is the sunrise hour angle (–), and hasunset is the sunset hour angle (+). For the example
given for Omaha, this results in hour angles of ±93.61°. Of course, the sunrise and sunset times
would need to be modified for the EoT, which is 3.90 minutes, and the fact that Omaha is 6 degrees
west of the middle of the time zone or 24 minutes later in solar time. So, the 6 hour and 14 minutes
before solar noon would be modified by having sunrise occur 24  minutes later because Omaha
is west of the 6-hour meridian and then modifying by the EoT adjustment so that sunrise occurs
3.90 minutes earlier; these modifications would result in rise occurring 5 hours and 54.34 minutes
before noon LST or 6:05:40 a.m. Sunset would occur 6 hours and 34.54 minutes after noon LST or
6:34:32 p.m. by similar reasoning.
With all of these corrections the sunrise and sunset hour angles still will not be exactly the
same value as the sunrise and sunset times that appear in the local paper. The sunrise reported
in the paper may be earlier, and the sunset times may be later. The reason for this difference is
primarily caused by refraction (bending of light by the atmosphere) that is, on average, approxi-
mated 0.7 degrees for the Sun on the horizon at sea level. Refraction causes the Sun to appear
slightly higher in the sky than simple geometrical calculations indicate. Another effect that gives
an additional minute of sunlight at sunrise and sunset is the fact that sunrise and sunset refer to
the rising of the leading edge of the Sun and setting of the trailing edge of the Sun. Since the
Sun is about 0.5 degrees in diameter, the edges are about 0.25 degrees from the center of the Sun
that is the reference point for the earlier calculations. The angular travel time for 0.25 degrees is
about 1 minute.

2.9 GLOBAL, DIRECT NORMAL, AND DIFFUSE IRRADIANCE


The nature of solar radiation reaching the Earth and the terminology commonly used to describe
the incident radiation components are fundamental to choosing the instruments used to measure
the solar resource. Near noon on a day without clouds, about 25 percent of the solar radiation inci-
dent outside the atmosphere is scattered and absorbed as it passes through the Earth’s atmosphere.
Therefore, roughly 1,000 Wm–2 of solar irradiance reaches the Earth’s surface without being sig-
nificantly attenuated. This “beam” radiation coming from the direction of the Sun is called direct
normal irradiance (DNI). A pyrheliometer mounted on a solar tracker is used to measure DNI.
Some of the scattered sunlight is reflected back into space, and some of it reaches the surface
of the Earth (see Figure 2.12). The scattered radiation reaching the Earth’s surface is called diffuse
horizontal irradiance (DHI). Some radiation is also scattered by the Earth’s surface and then
re-scattered by the atmosphere to the observer. This re-scatter is part of the DHI an observer sees.
A pyranometer shaded from the direct solar radiation is used to measure DHI. The total amount of
Definitions and Terminology 21

FIGURE 2.12 Components of solar radiation reaching the Earth’s surface. The extraterrestrial radiation is
the Sun’s irradiance as it impinges on the top of the atmosphere (TOA) of the Earth. All of the radiation radi-
ated from the Sun is un-scattered as it reaches the TOA of the Earth in its orbit around the Sun. As the sunlight
passes through the atmosphere some of it is scattered and some is absorbed. The DNI measured at the surface
is the portion of the sunlight that comes through the atmosphere with no scattering or absorption. Light scat-
tered by the molecules, aerosols, and clouds that strikes the observation area is called DHI. This DHI also
includes some irradiance that scattered from the Earth’s surface and then scattered again by the atmosphere
to the observation site. In snow-covered areas, this re-reflected radiation can be significant. The total GHI is
the sum of the DNI projected onto the horizontal surface plus all the diffuse radiation that makes it to the
observing instrument.

sunlight from the entire sky hemisphere on a horizontal surface is called global horizontal irradiance
(GHI). An unshaded pyranometer mounted on a stable horizontal platform is used to measure GHI.
Solar radiation is a flux of energy measured in power (usually watts) per unit area (usually square
meters). When the direction of the flux is normal to the surface, the amount of flux is greatest.
The amount of energy per unit area falls off approximately as the cosine of the angle between the
normal to the surface and the incident angle. In the case of DNI, the incident angle is the solar
zenith angle. A plot of GHI, DNI, and DHI on a clear day is shown in Figure 2.13.
The GHI is equal to the DNI times the cosine of the solar zenith angle (sza) plus the DHI:

GHI = DNI ⋅ cos ( sza ) + DHI (2.12)

This is an exact formula for instantaneous irradiance and becomes an approximation if the irradiance is
measured over time because the cosine of the solar zenith angle varies with time. During shorter time
intervals, the GHI can be larger than the extraterrestrial irradiance reaching the Earth’s orbit because of
clouds reflecting a good deal of incident irradiance can be reflected off the side of clouds onto the sens-
ing instrument. This effect is often referred to as the lensing effect of clouds (see Figure 2.14).
The amount of sunlight scattered from the surface, called ground-reflected irradiance, depends
on the albedo of the surface. The albedo is the ratio of the reflected or up-welling solar radiation
that is reflected back to the sky to the incident or down-welling GHI irradiance. The albedo can
vary from about 0.1  for dark surfaces to 0.2  for grasslands to greater than 0.8  for freshly fallen
snow. A perfectly reflective surface would have an albedo of 1.0. Because the atmosphere will scat-
ter some of the reflected light back to the surface, areas with high surface albedo will have a larger
diffuse component due to multiple surface reflections.
22 Solar and Infrared Radiation Measurements

FIGURE 2.13 Plot of the solar radiation components on a clear day. The value of DNI is influenced by the
amount of atmosphere through which it travels. During the middle of a clear day the air mass changes slowly
but changes rapidly in the morning and evening hours. On this clear day, the diffuse irradiance is only about
10 percent of the maximum global irradiance. The GHI is the DNI projected onto a horizontal surface plus the
DHI. Hence, on a clear day the GHI plotted against time has a typical bell shape or cosine shape.

FIGURE 2.14 The plot illustrates the lensing effect of clouds. GHI values can momentarily exceed the extra-
terrestrial irradiance when additional DNI is reflected by the side of clouds toward the sensor.
Definitions and Terminology 23

2.10 SOLAR RADIATION ON TILTED SURFACES


If the surface under study is tilted with respect to the horizontal, the total irradiance received is
the incident diffuse radiation on the tilted surface plus the DNI projected onto the tilted surface plus
ground-reflected irradiance that is incident on the tilted surface. The amount of direct radiation on a
horizontal surface (DrHI) can be calculated by multiplying DNI by the cosine of the solar zenith angle.
On a surface that is tilted by T degrees from the horizontal and rotated γ degrees from the north–south
axis (with north facing being zero and east facing being 90°), the direct component on the tilted surface
is determined by multiplying the DNI by the cosine of the incident angle to the tilted surface (θT).

cos (θT ) = sin ( dec ) ⋅ sin ( lat ) ⋅ cos (T ) + sin ( dec ) ⋅ cos ( lat ) ⋅ sin (T ) cos ( γ )
+ cos ( dec ) ⋅ cos ( lat ) ⋅ cos (T ) ⋅ cos ( ha )
(2.13)
− cos ( dec ) ⋅ sin ( lat ) ⋅ sin (T ) ⋅ cos ( γ ) ⋅ cos ( ha )
+ cos ( dec ) sin (T ) ⋅ sin ( γ ) ⋅ sin ( ha )

where the dec, declination, lat, latitude, and ha, hour angle were discussed earlier. An alternate
formula calculating the cosine of the incident angle to a tilted surface is

cos (θT ) = cos (T ) ⋅ cos ( sza ) + sin (T ) ⋅ sin ( sza ) ⋅ cos ( az − γ ) (2.14)

where az is the solar azimuthal angle. Irradiance on tilted surfaces is discussed in more detail in
Chapter 9.

2.11 THE SPECTRAL NATURE OF SOLAR RADIATION


The ultimate source of the Sun’s radiative energy is the fusion of hydrogen atoms in the Sun’s core.
This energy propagates to the surface through a series of absorption and re-emission processes
finally reaching the visible part of the Sun’s surface, the photosphere, after tens of thousands of
years. This  complex process of nuclear physics and radiative transfer produces energy over a
wide range of the electromagnetic spectrum. Conventional naming and wavelength regions of
the solar and infrared spectrum are shown in Table 2.1. The resulting spectral distribution of the

TABLE 2.1
Solar and Infrared Spectral Regions
Region Abbreviation Wavelength Range—λ (nm)
Ultraviolet C UVC 100 ≤ λ < 280
Ultraviolet B UVB 280 ≤ λ < 315
Ultraviolet A UVA 315 ≤ λ < 400
Visible VIS 380 ≤ λ < 760
Near infrared IR-A 760 ≤ λ < 1400
Middle infrared IR-B 1400 ≤ λ < 3000
Far infrared IR-C 3000 ≤ λ < 1000000

Source: CIE TC 2-17, Solar spectral irradiance, Commission Internationale


de l’Eclairage (International Commission on Illumination), CIE
085-1989, 48  p, http://www.cie.co.at/publications/solar-spectral-
irradiance, 1989 (Accessed May 25, 2019); ISO, Space Environ-
ment (Natural and Artificial): Process for Determining Solar
Irradiances. International Organization for Standardization,
ISO 21348:2007(E), 40 p, 2007.
24 Solar and Infrared Radiation Measurements

Extraterrestrial Solar Spectrum - [Wehrli]


2.5

Irradiance (Wm–2 per nm) 2

1.5

0.5

0
0 500 1000 1500 2000 2500 3000
Wavelength (nm)

FIGURE 2.15 Spectral distribution of the extraterrestrial solar irradiance reaching Earth. Most of the spec-
tral absorption lines result from the absorption of radiation generated in the photosphere as it passes through
the solar chromosphere. (After Gueymard, 2018 and Wehrli, 1985.)

Sun’s radiation at the top of the atmosphere (TOA) of the Earth when the Earth-Sun distance is
at its mean (149.6 million km) varies with the wavelength as shown in Figure 2.15. The majority
of solar irradiance is between 280 and 3,000 nm. The absorption of some solar radiation by the
atoms and ions in the chromosphere of the Sun results in absorption lines in the extraterrestrial
solar spectrum. The effect of these ions can be seen in the extraterrestrial solar spectrum shown
in Figure  2.15  as distinct reductions in the irradiance, especially for wavelengths below about
700 nm.
The spectral distribution of incident solar radiation at the top of the atmosphere spans a large range
of wavelengths from the ultraviolet to the infrared with the peak around 500 nm. Approximately
47 percent of the incident extraterrestrial solar radiation is in the visible wavelengths from 380 to
780 nm. The near-infrared portion of the spectrum with wavelengths greater than 780 nm accounts
for another 46 percent of the incident energy and the ultraviolet portion of the spectrum below
380 nm accounts for 7 percent of the extraterrestrial solar radiation.
As the sunlight passes through the atmosphere, a large portion of the ultraviolet radiation is
absorbed and scattered. Air molecules scatter the shorter wavelengths more strongly than the longer
wavelengths. On days with few clouds, the preferential scattering of blue light by air molecules is
the reason the sky appears blue. Water vapor and atmospheric aerosols further reduce the amount of
direct sunlight passing through the atmosphere. The information in Table 2.2 shows that on a clear
day approximately 75 percent of the extraterrestrial DNI (ETRn) passes through the atmosphere
without being scattered or absorbed. The magnitude of the scattering and absorbing are about equal
under these idealized conditions.
Many molecules selectively scatter or absorb radiation in specific wavelength regions, creating
large variations in the solar spectrum as measured at the Earth’s surface (Figure 2.16). The abun-
dance of the atmospheric constituents along the sunlight’s path determines the depth of these
absorption regions. Therefore, the solar spectrum depends on the length of the path through the
atmosphere and the density of the molecules along this path. Said another way, the atmosphere is
a highly variable filter affecting the amounts of solar irradiance reaching the surface at any given
time and place.
Definitions and Terminology 25

TABLE 2.2
Absorption and Scattering for Typical Clear-Sky Conditions
Percent of Total Passing
Factor Percent Absorbed Percent Scattered Through the Atmosphere
Ozone 2 0 2
Water vapor 6 1.5 8
Dry air 8 5 12
Aerosol 2 3.5 4
Total not absorbed or scattered 82 90 75

Source: Hu, C. and White, R.M., Solar Cells: From Basic to Advanced Systems, McGraw-Hill, New York, 1983.

FIGURE 2.16 Spectral distribution of the incident solar radiation as it reaches the Earth’s surface. This plot
is for a specific aerosol content of the atmosphere when the air mass is 1.5 when the instrument is tilted 37
degrees. The dashed line is the ETR radiation, the solid black line is the total irradiance on the tilted surface,
and the gray line is the DNI plus the circumsolar radiation. This plot shows the solar spectrum for the ASTM
(2012) reference solar angle and aerosol content that are used for specifying instrument performance as a
function of spectral wavelength.

Users of solar radiation data fall mainly into two groups. The broadband group is interested in
the total amount of energy available over all wavelengths, as in many solar thermal applications.
The spectral group is concerned with specific regions of the solar spectrum or the spectral distribu-
tion of the irradiance. For example, those who evaluate the performance of photovoltaic devices
may use detailed spectral measurements. Others such as architects are interested in lighting for the
human eye, which is sensitive only to the visible part of the spectrum. The eye’s response to various
wavelengths in daylight conditions is called the photopic response. More information on the phot-
opic response is presented in Chapter 7. Daylighting illuminance is the sum, over all wavelengths, of
the product of solar irradiance as a function of wavelength times the photopic response as a function
of wavelength. The SI unit for illuminance is the lux (lumen m–2).
26 Solar and Infrared Radiation Measurements

2.12 FUNDAMENTALS OF THERMODYNAMICS AND HEAT TRANSFER


Thermopile-based detectors are used in a wide variety of radiometers to measure incident solar
radiation. To characterize these instruments and to understand how these instruments operate, a
background on the fundamentals of thermodynamics and heat transfer is provided. Some instru-
ments use the photoelectric effect and the basics of photodiodes and solar cells are discussed in
Section 2.13.
The basis for all thermopile radiometers is the thermoelectric effect, the generation of voltage
from temperature differences at the junctions of two dissimilar metals. The thermoelectric effect was
discovered in 1821 by Thomas Seebeck, a German physicist. The Seebeck Effect predicts 41 micro-
volts (µV) from junctions of copper and constantan (a copper-nickel alloy) per Kelvin (K) (as deter-
mined by T2–T1  in Figure 2.17) at room temperature (Bunch and Hellemans 2004; Velmre 2007).
Thermopile-based detectors measure the heat flow from the hot junction to the cold junction in the
instrument. These instruments are designed to maximize the heat flow from the sensor receiver (hot
junction) through the thermopiles to the heat sink (cold junction) and minimize the other paths that
bypass the thermopiles. To comprehend the rationale for various design features and to characterize
the performance of the solar instruments it is necessary to understand how heat flows.
There are three ways that heat can flow:

• Conduction of heat through a material


• Convection of heat by liquid or gas from a surface
• Radiative transfer of heat

The  nature of conduction, convection, and radiation of heat are discussed in Sections  2.12.1 to
2.12.3, and some of the basic equations governing heat transfer are given.

2.12.1 ConduCtion
Conduction is the transfer of heat through a material as a result of molecular motion. The more heat
applied to a material, the faster the molecules that make up the material will vibrate. When heat is
applied to a material, the molecules where the heat is applied will vibrate faster. This vibration causes
molecules nearest this heat source to vibrate faster, and slowly this vibration spreads throughout the
material. For a slab, the rate of heat transfer depends on the area where the heat is applied, the type

FIGURE 2.17 Circuit diagram illustrating the thermocouple where two different metals at different tempera-
tures generate an electrical voltage when electrons flow from the hot junction (T2) toward the cold junction (T1).
Definitions and Terminology 27

and structure of the material, the temperature difference across the material, and the thickness of the
material. A general formula for heat transfer across a slab is given by

dQ dT
= −kA (2.15)
dt dx

where dQ/dt is the rate of heat flow (change in heat over the time interval dt), dT/dx is the tempera-
ture gradient (the change in temperature over a change in distance traveled in the material (dx)), A is
area, and k is the thermal conductivity of the material given in (Wm–1K–1). The thermal conductivi-
ties of various materials at room temperature are given in Table 2.3. The thermal conductivity varies
widely and materials can be chosen to increase or decrease the rate of heat flow. Textbooks such as
the one by Yener and Kakac (2008) cover thermal conductivity.

2.12.2  ConveCtion
Convection is the movement of heat by a gas or a liquid from one place to another. A more apt
description of the processes of concern is advection, taking heat away from a surface and then
having the gas or fluid move the heat by convection. An example of advection is wind blowing air
molecules across a surface with heat transferred from the surface to the air. Thus, the air carries
away the heat from the surface by convection. Newton’s law of cooling states that the rate of change
in temperature of an object is proportional to the temperature difference between the object and the
fluid or gas taking away, or adding, heat.

dQ
= hA(Ts − Tg ) (2.16)
dt

where Ts is the temperature of the surface, Tg is the temperature of the gas or liquid, and h is the
heat transfer coefficient in units of Wm–2K–1. The heat transfer coefficient varies with the type of
gas or liquid, the type and geometry of the surface, and how the fluid or gas flows over the surface.
The heat transfer coefficient of water is about 33 times that of air. Therefore, air with high levels
of humidity has the capability of taking away more heat than dry air. Relative humidity can be a
factor in the performance of solar instruments because air containing more moisture is better at
transferring heat. The type of flow is also important. Laminar flow is when the gas or fluid moves in
parallel layers and no currents or eddies are perpendicular to the flow. Turbulent flow is chaotic and
there is considerable mixing between layers. There is also a transitional flow that is neither laminar
nor fully turbulent. Turbulent flow is better at transferring heat from the surface than laminar flow.
Convection and advection can be complex, and many papers and textbooks discuss convective heat
flow (see, e.g., Kreith and Manglik 2018; Lienhard and Lienhard 2002).

TABLE 2.3
Conductivity at Room Temperature
Material Conductivity (Wm−1 K−1)
Brass 120.000
Copper 390.000
Steel 50.000
Styrofoam 0.033
Water 0.600
28 Solar and Infrared Radiation Measurements

2.12.3 Radiative Heat tRansfeR


All bodies radiate and absorb heat. This radiation is in the form of electromagnetic waves. These
waves can vary in wavelength from X-rays and ultraviolet waves to visible, infrared, and radio
waves. The energy in these waves is inversely proportional to the wavelength; the longer the wave-
length, the less energy the wave carries. In 1897, Jožef Stefan (1835–1893) and Ludwig Boltzmann
(1844–1906) developed the Stefan Boltzmann law stating that the energy emitted by a body is pro-
portional to the fourth power of the temperature of the body.

E = εσ T 4 (2.17)

where E is the energy emitted, e is the emissivity of the material, σ is the Stefan Boltzmann constant
equal to 5.670367 × 10–8 Wm–2K–4, and T is the temperature in K. A blackbody is a perfect emitter
with an emissivity equal to 1. The spectral distribution of the radiation emitted is given by Planck’s law

2hc 2 1
I ( λ ,T ) = ⋅ hc (2.18)
λ5
e λ kT −1

where λ is the wavelength, in micrometers, for example, h is Planck’s constant, c is the speed
of light, k is the Boltzmann constant, and T is the temperature. Planck’s constant is equal to
6.626070040.10 –34 Js, the speed of light is equal to 2.99792.108 ms–1, and the Boltzmann constant
is 1.38064852 × 10 –23 m2 kg s–2 K–1. Equation 2.18 can be used to estimate the spectral distribution
of the solar radiation incident outside the Earth’s atmosphere. The spectral flux density from the
Planck spectral distribution has to be lowered by the square of the ratio of radius of the Sun divided
by the distance between the Earth and the Sun.
The wavelength of the maximum energy emitted by a blackbody, from Wien’s law, is

2898
λ= (2.19)
T

where T is temperature in K, and λ is in micrometers.


Another important property of an object is its absorptivity, an object’s ability to absorb incident
solar radiation. In  thermal equilibrium, Kirchhoff’s law states that the absorptivity of a body is
equal to its emissivity. Therefore, a good absorber is a good emitter. A perfect blackbody has an
emissivity of one and an absorptivity of one. In general, the duller and darker the material, the closer
the emissivity is to one, and the more reflective, the lower its emissivity. Highly polished silver
has an emissivity of about 0.02 and is a great reflector and a poor absorber. The wavelength of the
incoming radiation, its angle of incidence with respect to the surface, and the shape of the object
all affect the emissivity and, hence, the absorptivity of a material. Building a high-performance
radiometer requires careful attention to these properties during the design and characterization of
the instrument. Detailed explanations of radiative heat transfer can be found, for example, in Siegel
and Howell (2002) or Kreith and Manglik (2018).

2.13 PHOTODIODES AND SOLAR CELL CHARACTERISTICS


Because of their simplicity to manufacture, many solar radiometers are based on the silicon photo-
diode that responds to radiation using the photoelectric effect rather than the thermoelectric effect
just described. A rudimentary description of how photodiodes can be used to measure solar radia-
tion is provided, forming the basis for characterizing solar sensors using photodiode detectors in
subsequent chapters.
Definitions and Terminology 29

Silicon photodiodes are made from crystalline silicon that has been transformed into a
semiconductor by introducing atoms that do not have the same number of valence electrons as
silicon. Silicon has four valence electrons (quadrivalent) that form tight bonds with four neigh-
boring silicon atoms. Introduction of atoms with one fewer or one extra valence electron is called
doping. A photodiode consists of n- and p-type doped silicon sections that are brought together.
N-type doped silicon contains a number of atoms with one excess electron when all the electron
bond pairs are filled in the crystalline structure. Typically for n-type doping, approximately every
thousandth silicon atom is replaced by a phosphorous atom (pentavalent) with one extra valence
electron. P-type doped silicon contains a number of atoms with one less electron than is needed to
complete the electron bond pairs. Typically for p-type doping, every millionth atom is replaced by
a trivalent atom such as boron. This missing electron acts much like a positively charged electron
and is treated as such when modeling semiconductor materials. In the manufacturing process the
p-type silicon is the base material and then the cell is doped with n-type atoms. At the boundary
of n-type and p-type materials, the loosely bound electrons in the n-type silicon drift over to the
p-type side to fill “holes” that are created by doping. The energy of the pairing of the valence
electrons is stronger that the electromagnetic energy between an outer unpaired negatively charge
electron and the proton in the nucleus resulting in the drift of electrons from the n-type material
to the p-type material. This drift results in a shift of charge because the nucleus of the doped
atoms that were originally associated with the electrons or holes did not move. An electric field
is created by the difference between the number of positively charged protons in the doping atom
and the number of negatively charged electrons in the bonds. The  layer of silicon where this
shift of electrons and holes occurs is called the depletion layer as the number of free holes and
electrons are depleted. This electric field of about 0.5 V separates the electron–hole pairs that are
created when solar photons knock the electrons from their bound state in the depletion layer of
the doped silicon semiconductor. This electric field allows the movement of the electron and hole
to their respective collection areas. When the photodiode is exposed to radiation and connected in
a circuit, a current will flow. The current created by this photodiode is proportional to the irradi-
ance and can then be measured.
Mathematically it is easier to treat the properties of solar cells in momentum space. For semi-
conductor materials, electrons can be treated as living in energy bands. Conduction bands are areas
with excess electrons and valence bands are areas with excess “holes.” Without light the conduction
and valence bands reach equilibrium. Incident light can disturb this equilibrium and, in a silicon
solar cell, if the energy of the photon is greater than 1.1 eV (electron volt), electrons can jump into
the conduction band. A  full discussion of how semiconductors work is the subject of solid-state
physics texts such as Kittel (2005).
When light is incident on a photodiode, the resulting current is the sum of the photodiode cur-
rent, the shunt resistance current, and the photocurrent.

 eV 
I out = I ph − I s ⋅  e kT − 1 − I sh (2.20)
 
 

where Iout is the output current, Iph is the photocurrent generated by the incident light, Ish is the shunt
resistance current, and the middle term is the result of the diode current (Is is the reverse diode
current, e is the electron charge of 1.602 10 –19 coulombs, V is the applied voltage in volts, k is the
Boltzmann constant, and T is the absolute temperature of the photodiode in K). The shunt resistance
current is an alternate path for the current to flow. Cells are designed with a large shunt resistance
to minimize the loss. The diode current is the current that flows when there is no light incident on
the device and is sometimes called the dark current. The circuitry for photodiodes is designed in a
manner that minimizes the dark current.
30 Solar and Infrared Radiation Measurements

2.14 MODELS
There  are a wide variety of instruments for measuring solar irradiance, but historically only a
limited quantity of measured solar data are available. To compensate for this limitation, many
models have been created to estimate solar irradiance values from meteorological observations at
the surface and from satellite-based images of the Earth’s atmosphere and surface. Many challenges
exist for comparing point measurements from ground stations vs. spatially averaged data from
satellite-based observations. In addition, relationships between the various solar components have
been developed to estimate and optimize the potential performance of energy conversion systems.
The discussion of solar radiation models is outside the scope of this book, but an overview is given
in Appendix B. These modeled estimates are often inadequate substitutes for measuring solar and
infrared irradiance when accurate information is required to validate a system’s field performance.
Today the financing and operation of large-scale solar electric power plants is driving the demand
for ever-increasing accuracy in solar radiation measurements.

QUESTIONS (Use equations in this chapter for calculations)


1. If the Sun was a perfect blackbody, then what would be its temperature?
2. Calculate the solar declination for September 23 to nearest hundredth of a degree.
3. What is the average extraterrestrial irradiance at the mean Earth–Sun distance?
4. What is the minimum solar zenith angle for a selected location to the nearest degree?
5. What is a selected LST to the nearest minute when it is solar noon on March 21  in
Greenwich, England?
6. What is the local standard time at 12:00 p.m. CDT on May 21 in Chicago?
7. What is the geometric solar zenith angle at sunrise to the nearest degree on April 20?
8. What is the relationship between DNI, DHI, and GHI?
9. How would you calculate DNI from measured GHI and DHI?
10. Calculate the cosine of the incident solar angle on a surface tilted 30 degrees from the
horizontal facing due west on June 21 at 2:00 p.m. near Denver at 40°N latitude and 105°W
longitude to two decimal places.
11. How much energy is radiated from a blackbody at room temperature to the nearest Wm−2?
12. How much energy is absorbed by a blackbody at room temperature to the nearest Wm−2?
13. Using a plot of the extraterrestrial solar spectrum, what is the peak wavelength of extrater-
restrial solar radiation?
14. Why are modeled solar values used in resource assessment instead of measured data?

REFERENCES
ASTM. 2015. Standard Terminology of Solar Energy Conversion, E772-15, ASTM International. West
Conshohocken, PA: American Society for Testing and Materials.
ASTM. 2012. Standard Tables for Reference Solar Spectral Irradiances: Direct Normal and Hemispherical
on 37° Tilted Surface, G173–03, ASTM Standards (E-891) and (E-892). West Conshohocken, PA:
American Society for Testing and Materials. https://www.astm.org/Standards/G173.htm (Accessed
May 19, 2018).
Beckman, W. A., J. W. Bugler, P. I. Cooper, J. A. Duffie, R. V. Dukle, P. E. Glaser, T. Horigome et al. 1978.
Units and symbols in solar energy. Solar Energy 21: 65–68. doi:10.1016/0038-092X(78)90118-4.
Bunch, B., and A. Hellemans. 2004. The History of Science and Technology: A Browser’s Guide to the Great
Discoveries, Inventions, and People Who Made Them from the Dawn of Time to Today. Boston, MA:
Houghton Mifflin.
Commission Internationale de l’Eclairage (CIE) 1989. Solar spectral irradiance. Commission Internationale
de l’Eclairage (International Commission on Illumination), TC 2-17, CIE 085-1989. http://www.cie.
co.at/publications/solar-spectral-irradiance (Accessed May 25, 2019).
Definitions and Terminology 31

de Wit, T. D., G. Kopp, C. Fröhlich, and M. Schöll. 2017. Methodology to create a new total solar irradiance
record: Making a composite out of multiple data records. Geophysical Research Letters 44 (3): 1196–1203.
Duffie, J. A., and W. A. Beckman. 1991. Solar Engineering of Thermal Processes. New York: John Wiley &
Sons.
Fröhlich, C. 2006. Solar irradiance variability since 1978. Space Science Reviews 90: 1–13.
Gueymard, C. A. 2018. Revised composite extraterrestrial spectrum based on recent solar irradiance observa-
tions. Solar Energy 169: 434–440. doi:10.1016/j.solener.2018.04.067.
Hu, C., and R. M. White. 1983. Solar Cells: From Basic to Advanced Systems. New York: McGraw-Hill.
International Electrotechnical Commission (IEC). 2018. International Electrotechnical Commission
Glossary, TC 82. Geneva, Switzerland: International Electrotechnical Commission.
International Organization for Standardization (ISO). 1999. Solar energy—Vocabulary, TC-180. ISO
9488:1999. https://www.iso.org/obp/ui/#iso:std:iso:9488:ed-1:v1:en (Accessed May 25, 2019).
International Organization for Standardization (ISO). 2007. Space environment (natural and artificial)—
Process for determining solar irradiances. ISO 21348:2007(E).https://www.iso.org/standard/39911.
html (Accessed June 5, 2019).
Kittel, C. 2005. Introduction to Solid State Physics. Hoboken, NJ: John Wiley & Sons.
Kopp, G. 2016. Magnitudes and timescales of total solar irradiance variability. Journal of Space Weather and
Space Climate 6 (27): A30. doi:10.1051/swsc/2016025.
Kopp, G., and J. L. Lean. 2011. A new, lower value of total solar irradiance: Evidence and climate significance.
Geophysical Research Letters 38: L01706. doi:10.1029/2010GL045777.
Kreith, F., and R. M. Manglik. 2018. Principles of Heat Transfer. 8th ed. Boston, MA: CL Engineering.
Lienhard IV, J. H., and J. H. Lienhard V. 2002. A Heat Transfer Textbook. Cambridge, MA: Phlogiston Press.
Michalsky, J. J. 1988. The  astronomical almanac’s algorithm for approximate solar position (1950–2050).
Solar Energy 40: 227–235.
National Aeronautics and Space Administration. 2018. NASA  sun fact sheet. https://nssdc.gsfc.nasa.gov/
planetary/factsheet/sunfact.html (Accessed May 18, 2018).
Paltridge, G. W., and C. M. R. Platt. 1976. Radiative Processes in Meteorology and Climatology. Amsterdam,
the Netherlands: Elsevier Scientific Publishing Company.
Reda, I., and A. Andreas. 2008. Solar position algorithm for solar radiation applications (revised). NREL
Report No. TP-560–34302.
Siegel, R., and J. R. Howell. 2002. Thermal Radiation Heat Transfer. New York: Taylor & Francis Group.
Spencer, J. W. 1971. Fourier series representation of the position of the sun. Search 2: 172.
Velmre, E. 2007. Thomas Johann Seebeck (1770–1831). Proceeding of the Estonian Academy of Sciences,
Engineering 13 (4): 276–282.
Watt Engineering, Ltd. 1978. On the Nature and Distribution of Solar Radiation. U. S. Department of Energy.
HCP/T2552-01. UC-59, 62, 63A. Supt. of Docs., U.S. Govt. Print. Off., 1978. 256 p.
Wehrli, C. 1985. Extraterrestrial Solar Spectrum. Publication no. 615. Davos Dorf, Switzerland: Physikalisch-
Meteorologisches Observatorium Davos + World Radiation Center (PMOD/WRC).
World Meteorological Organization. 1992. International Meteorological Vocabulary. WMO/OMM/
BMO – 182. Geneva, Switzerland: Secretariat of the World Meteorological Organization, 802 p.
World Meteorological Organization. 2017. WMO Guide to Meteorological Instruments and Methods of
Observation (the CIMO Guide), WMO-No. 8. Geneva, Switzerland: World Meteorological Organization.
https://www.wmo.int/pages/prog/www/IMOP/CIMO-Guide.html.
Yener, Y., and S. Kakac. 2008. Heat Conduction. Boca Raton, FL: CRC Press.
3 Historic Milestones
in Solar and Infrared
Radiation Measurement
To measure is to know.
Lord Kelvin (1824–1907)

3.1 INTRODUCTION
The  ability to measure solar and infrared radiation has evolved through centuries of scientific
discovery and engineering applications. Historically, measurements have played a critical role in the
development of accepted physical laws. Today, solar and infrared radiation measurements are par-
ticularly important for advancing atmospheric physics, developing renewable energy technologies,
and understanding climate change. The goal of improving data accuracy is common to the develop-
ment of instrumentation for measuring solar and infrared radiation. This chapter provides a selective
account of important solar and infrared radiation observations and instrumentation developments.

3.2 EARLIEST OBSERVATIONS OF THE SUN AND THE NATURE OF LIGHT


Sunspots and their cyclic behavior have long been of interest in scientific observation. The invention
of the telescope in the early seventeenth century provided the capability to determine that the dark
spots seen on the face of the Sun, noticed as early as the fourth century B.C.E., were part of the Sun
(Eddy 1976). In addition to viewing the number of sunspots, the early telescopes provided improve-
ments to observations of the faculae (bright areas on the surface of the Sun that typically occur before
a sunspot) size, shape, location, and most of the other observational detail recorded today. In 1848, a
number of European observatories began regular observations using a standard scheme developed by
Rudolf Wolf (Waldmeier 1961). Combined with future efforts to standardize measurement practices,
the work started by Rudolf Wolf produced the most reliable historical record of sunspot numbers in the
nineteenth century. Interest in sunspots continues as an area of recent research (Kopp 2016).
In 1666, Sir Isaac Newton discovered the spectral distribution of sunlight using glass prisms to
display the visible spectrum in a darkened room. He also developed the corpuscular theory of light
as a means of explaining propagation in a straight line and the property of specular reflection from
mirror-like surfaces. Other discoveries in the seventeenth century include the law of refraction by
Willebord Snell and the description of light diffraction, discovered independently by Francesco
Grimaldi and Robert Hooke (Coulson 1975). The  electromagnetic character of visible radiation
would not be discovered for another two centuries.
The measurement of direct solar radiation, the amount of “beam” radiation from the direction of
the sun, is essential for understanding the total solar irradiance (TSI) output from the Sun. Formerly
known as the solar constant, determining the mean value of TSI and its slight variability continues
to be the subject of research in astronomy, atmospheric physics, climate change, renewable energy
technology development, and other applications (Dudok de Wit et al. 2017; Kopp and Lean 2011).
In 1837, Claude Servais Mathias Pouillet produced the first successful research on TSI measure-
ments and introduced the term pyrheliometer from the Greek fire, sun, and measure (Crova 1876).

33
34 Solar and Infrared Radiation Measurements

In 1913, Charles Greeley Abbot and Loyal Blaine Aldrich at the Smithsonian Institution developed
the first radiometer designed for measuring global horizontal irradiance (GHI) (downwelling total
hemispheric radiation from the sun and sky). Using the Greek words for fire, above, and a measure,
Abbot and his colleagues named the pyranometer for its ability to measure solar heat from above.1
Intended to serve as a standard instrument for solar radiation measurements, details of the design
were given by Abbot and Aldrich (1916).
Other notable advances in understanding and measuring optical radiation are listed in Table 3.1,
and more detailed descriptions of selected instruments are presented in this chapter.

TABLE 3.1
Highlights of Optical Radiation Measurement
Year Event
1611–1612 David and Johannes Fabricius, Christoph Scheiner, and Galileo Galilei independently document the
existence of sunspots and the Sun’s rotation on its axis.
1666 Sir Isaac Newton discovers the spectral nature of visible sunlight using glass prisms.
1800 William Hershel discovers infrared radiation using thermometers and a prism to measure spectral
irradiance of sunlight beyond the visible range.
1825 Herschel develops an actinometer as the first instrument to transform information about cooling rate into
solar radiation measurements.
1837 First use of the term pyrheliometer. Claude Pouillet invents his pyrheliometer and measures the solar
constant as 1.76 cal cm−2 min−1 (1227 Wm−2) within about 10 percent of the currently accepted value of
1,361 Wm−2.
1879 Campbell–Stokes burning sunshine recorder developed. The instrument remains in use to provide daily
records of percent possible bright sunshine. These data are often used to estimate the equivalent global
irradiation at the measurement site.
1884 Boltzmann derives the Stefan-Boltzmann law to describe the relationship between a body’s temperature (T)
and its emitted radiant energy (W): W = εσT4.
1893 Knut Ångström invents the electrical compensation pyrheliometer.
1898 H. S. Callendar invents the Callendar pyranometer using four platinum wire grids wound on strips of mica
to form electrical resistance thermometers. Two square grids were painted black and two were naturally
reflective platinum.
1900 The Smithsonian Institution in Washington, DC, begins measuring global horizontal radiation.
1901 The U.S. Weather Bureau begins first measurements of solar radiation by deploying three Ångström’s
electric compensation pyrheliometers in Washington, D.C., Baltimore, MD, and Providence, RI, to
measure direct normal irradiance.
1904 Samuel P. Langley attempts correlation between weather and solar radiation.
1905 Knut Ångström’s electrical compensation pyrheliometer is adopted as a measurement standard at the
Meteorological Conference, Innsbruck and by the Solar Physicist Union, Oxford.
1909 The U.S. Weather Bureau begins routine measurements of solar GHI using the Callendar pyranometer in
Washington, D.C., with later installations in Madison, WI (1911) and Lincoln, NE (1915).
1910 Smithsonian water-flow pyrheliometer is perfected into the primary standard.
Abbot develops the secondary standard silver-disk pyrheliometer.
The Marvin pyrheliometer is invented.
Measurements of direct normal irradiance commence on Mt. Fuji, Japan, and at Madison, WI; Lincoln,
NE; and Santa Fe, NM.
1911 The U.S. Weather Bureau begins its first routine measurement of direct normal irradiance with the
deployment of Marvin pyrheliometers at Madison, WI; Lincoln, NE; and Santa Fe, NM.
(Continued)

1 In 1961, in the second edition of the “Guide to Meteorological Instrument and Observing Practices, (World Meterological
Organization (WMO), 1965)” the World Meteorological Organization first recommended the use of the term pyranom-
eter for hemispherical measurements to replace the more commonly used terminology 180º pyrheliometer.
Historic Milestones in Solar and Infrared Radiation Measurement 35

TABLE 3.1 (Continued)


Highlights of Optical Radiation Measurement

Year Event
1912 Measurements in Algeria show 20 percent reduction of sunlight due to volcanic ash from the Mt. Katmai
eruption in Alaska.
1913 First revision of Smithsonian Pyrheliometric Scale, based on water-flow pyrheliometer.
Abbot develops balloon-borne pyranometer carried to 45,000 ft. for solar constant measurements. Mean
value of the solar constant given as 1.933 cal cm−2 min−1 (1348 Wm−2) within about 1 percent of the
currently accepted value of 1,360.8 ± 0.5 Wm−2.
1915 In April, the Monthly Weather Review begins regular publication of measured solar irradiance data from the
U.S. Weather Bureau.
1916 C. G. Abbot and L. B. Aldrich at the Smithsonian Institution develop a pyranometer for measuring global
horizontal irradiance.
1919 A. K. Ångström invents electrical compensation pyrgeometer for measuring infrared irradiance.
1923 The Kimball-Hobbs pyranometer (later adopted by Eppley for their “light bulb” pyranometer) is invented.
1924 Deployment of the first Moll-Gorczynski pyranometer, later known as the Kipp & Zonen solarimeter.
1927 Shulgin develops double water-flow pyrheliometer to improve measurement accuracy.
1930 The Eppley Laboratory, Inc. commercializes the Kimball-Hobbs 180° pyrheliometer and it is later given
the nickname, “Eppley light bulb pyranometer.”
1932 The Smithsonian Pyrheliometric Scale is revised based on double water-flow pyrheliometer (adopted as the
standard by Smithsonian Institution).
Robitzsch bimetallic pyranometer (actinograph) provides self-contained recording mechanism, but
not recommended by the Radiation Commission of the International Association of Meteorology for any
measurements except daily total irradiation.
1950–1951 The U.S. Weather Bureau increases the number of pyranometer measurement stations to 78 using the
Eppley 180° pyrheliometer.
1952 WMO’s Subcommission on Actinometry recommends the silver-disk pyrheliometer for measuring direct
normal irradiance.
1954 U.S. Weather Bureau changes method of calibrating Eppley pyranometers from outdoor to an indoor
integrating sphere and artificial lighting.
1955 The International Geophysical Year (IGY) was established by the International Council of Scientific
Unions to observe geophysical phenomena and to secure data from all parts of the world; to conduct this
effort on a coordinated basis by fields, and in space and time, so that results could be collated in a
meaningful manner (World Data Center A 1958). This global scientific effort marked the beginning of the
space race, prompted the development of world data centers, and advanced solar monitoring and weather
forecasting capabilities.
1956 International Radiation Conference in Davos, Switzerland, recommends new measurement scale,
“International Pyrheliometric Scale (IPS).”
Yanishevsky pyranometer is developed for measuring global and diffuse irradiance and surface albedo in
the USSR.
1957 The International Pyrheliometric Scale 1956 is established. Ångström Scale is increased by 1.5 percent,
and 1913 Smithsonian Scale is decreased by 2.0 percent for compliance.
1957 The Eppley Laboratory, Inc. introduces Model PSP and the Model NIP.
1968 First measurements of extraterrestrial irradiance (total solar irradiance) are reported from flights of the
USAF/NASA X-15 by the Eppley Laboratory, Inc. and the Jet Propulsion Laboratory (TSI = 1360 Wm−2).
1969 The Eppley Laboratory, Inc. introduces model 8-48 “black-and-white” pyranometer. The light bulb
pyranometer is discontinued.
1969 The PACRAD is reported by J. M. Kendall, Sr. at JPL. The design is based on electrical heating
substitution equivalent to radiation heating of the cavity receiver.
1970 The WRC conducts the Third IPCs where the first electrically self-calibrating absolute cavity radiometer
participates.
(Continued)
36 Solar and Infrared Radiation Measurements

TABLE 3.1 (Continued)


Highlights of Optical Radiation Measurement

Year Event
1971 Lambda Instruments Corporation (now LI-COR) develops a pyranometer with a photodiode detector
(Model LI-200).
1975 DIAL radiometer concept is developed for low-cost automated field measurements of atmospheric
turbidity. The design formed the basis for future commercial systems known as rotating shadowband
pyranometer, rotating shadowband radiometer, thermopile shadowband radiometer, and multifilter
rotating shadowband radiometer.
1977 PACRAD III becomes the reference radiometer for the IPC.
The University of Oregon in Eugene begins the measurement of DNI and other solar irradiance elements to
establish what is now the longest continuous record of solar resources available from anywhere in the
United States.
1978 NIMBUS satellite radiometers begin measurement record of total solar irradiance using the HF absolute
cavity radiometer.
1979 WRR is established by the WRC as the internationally recognized measurement standard for DNI (above
700 Wm−2).
The U.S. National Weather Service develops and tests solar radiation forecasts for 1 to 3 days ahead using
measurements from three agricultural stations in South Carolina (Jensenius and Carter 1979).
The National Oceanic and Atmospheric Administration (NOAA) proposes a new system to provide daily
solar energy forecasts 1 and 2 days in advance for the contiguous U.S. based on data from 34 solar
measurement stations in the NOAA Solar Radiation Network (Jensenius and Cotton 1981).
1990 The Atmospheric Radiation Measurement (ARM) Program was conceived by the U.S. Department of
Energy to produce improved measurements of broadband and spectral solar and infrared irradiance,
aerosol and cloud properties, and other parameters critical for developing improved predictions of climate
change (Turner and Ellingson 2017).
1994 MFRSR is commercialized as a ground-based instrument for measuring optical depth and solar radiation
components using seven independent photodiode detectors. (Harrison et al. 1994)
2003 TIM begins measuring TSI from space as part of NASA’s Earth-observing system SORCE. With an
estimated absolute accuracy of 350 ppm (0.035%), the TIM can detect relative changes in solar irradiance
to less than 10 ppm/yr (0.001%/yr), allowing determination of possible long-term variations in solar
output (Kopp et al. 2005).
2004 WISG of Pyrgeometers is established by the WRC in Davos, Switzerland, as an interim reference for
atmospheric downwelling infrared radiation measurements (WMO 2006).
2010 Variable condition pyrheliometer comparisons completed at the NREL (Michalsky et al. 2011).
2012 PREMOS/PICARD launched to measure TSI (Fröhlich 2012)
Kipp & Zonen introduces pyrheliometers and pyranometers with RS-485 Modbus® digital output for
connecting directly to Supervisory Control and Data Acquisition (SCADA) systems to meet the needs of
utility-scale solar power plants for monitoring nearly real-time solar and meteorological conditions.
2014 California Independent System Operator (CAISO) specifies a minimum of one solar/meteorological
measurement station for solar powered generation stations greater than one megawatt, increasing the
demands for more cost-effective radiometers.
2016 Solar resource forecasting methods available for utility-scale solar power plants (Haupt et al. 2018)
2017 International Electrotechnical Commission (IEC) releases International Standard IEC 61724-1:2017,
Photovoltaic system performance monitoring – Guidelines for measurement, data exchange and analysis,
defining pyranometer “accuracy classes” for PV testing conformity.
2018 The Total and Spectral Solar Irradiance Sensor (TSIS-1) joins the NASA SORCE orbiting space mission
(NASA 2018).

Source: Coulson, K. L., Solar and Terrestrial Radiation, Methods and Measurements, Academic Press, New York, 1975.
Historic Milestones in Solar and Infrared Radiation Measurement 37

3.3 NINETEENTH-CENTURY RADIOMETERS


3.3.1 Pouillet’s PyRHeliometeR (1837)
As previously introduced, Pouillet’s pyrheliometer was the first to be called a pyrheliometer and
was designed as a calorimeter for measuring solar energy. The radiation receiver was a water-filled
cylindrical vessel with a blackened absorbing surface aligned perpendicular (normal) to the sun as
shown in Figure 3.1. The other surfaces of the cylindrical vessel were silvered to reduce the heat
exchange between the receiver and the atmosphere. The receiver fluid temperature was measured
with a thermometer in direct contact. The cylinder was rotated about the solar alignment axis to
provide mixing of the calorimeter fluid (later changed to mercury). Measurements of direct normal
irradiance (DNI) were based on alternating 5-minute periods of aligning the blackened receiver
surface normal to and away from the sun. The corresponding temperature rise and fall, the heat
capacity of the calorimeter, the total heat received per unit of surface, and unit of time could then
be computed. From his clear-sky measurements, Pouillet estimated the TSI to be 1227 watts/square
meters (Wm−2) (1.76  calories/square centimeters per minute (cal cm−2  min−1)) (Coulson 1975) or
about 10 percent lower than the currently accepted and most probable value of 1360.8 ± 0.5 Wm−2
(Kopp and Lean 2011).

3.3.2 CamPbell-stokes sunsHine ReCoRdeR (1853, 1879)


Although the Campbell-Stokes Sunshine Recorder is not a radiometer, this device and some later
variants remain in use today for measuring bright sunshine duration. These observations then can
be used to estimate daily amounts of solar irradiation (Iqbal 1983). J. F. Campbell invented his

FIGURE 3.1 Pouillet’s pyrheliometer–1837. DNI measurements were based on the temperature changes of
the water-filled blackened cylindrical receiver resulting from alternating exposure to the sun and the sky at
5 minute intervals. The end of the thermometer (T) was inserted into the receiver (A) and positioned for read-
ing (B) ahead of the alignment disc (C). (From Fröhlich, C., Metrologia, 28, 111–115, 1991.)
38 Solar and Infrared Radiation Measurements

burning sunshine recorder in 1853 by using a glass ball filled with water. This spherical lens was
placed inside a white stone bowl (later mahogany wood) with a 4-inch diameter surface painted with
an oil paint or varnish and engraved with hour lines. During periods when the solar disk was free
of clouds or dense haze, the concentrated solar radiation melted the paint (burned the mahogany).
The total length of the bright sunshine burns was compared with the total day length to compute
percent-possible bright sunshine. Later, a solid glass sphere replaced the water-filled ball to allow
operation in freezing conditions. In  1879, Sir G. G. Stokes significantly modified the Campbell
sunshine recorder to assume its present form (Figure 3.2). The solid glass sphere is supported by a
metal structure that holds specially treated paper strips with precise hourly and sub-hourly mark-
ings. The strips are available in three different lengths for the different periods: summer, winter,
and spring/fall. Daily, a single card is inserted and precisely registered in position with the time
markings using a metal pin. The treated paper is stable over a range of humidity levels, including
effects of precipitation, allowing the paper to burn when direct normal irradiance levels exceed
210 Wm−2. In 1962, the Commission for Instruments and Methods of Observation (CIMO) of the
World Meteorological Organization (WMO) adopted the Campbell–Stokes sunshine recorder as a
standard of reference (WMO 2008). In 1966, the analysis of bright sunshine data from 233 mea-
surement stations and solar radiation measurements from an additional 668 stations were used to
produce the first set of monthly world maps of the daily mean total solar radiation on a horizontal
surface (Löf et al. 1966). Routine measurements of bright sunshine duration using the Campbell-
Stokes instrument continue to the present day. These data are included in the records maintained by
the World Radiation Data Center (http://wrdc.mgo.rssi.ru/).

3.3.3 ÅngstRöm eleCtRiCal ComPensation PyRHeliometeR (1893)


Unlike previous designs based on calorimetry, the Ångström electrical compensation pyrheliom-
eter was the first to equate electrical power with solar irradiance. Developed by Knut Ångström
in the late nineteenth century, this pyrheliometer was recognized for its reliability and accuracy
(Ångström 1893, 1899). With only minor design developments based mostly on improved electrical

FIGURE  3.2 Campbell–Stokes sunshine recorder continues to provide measurements of bright sunshine
duration at several locations around the world. Sensitized recording paper is available in three lengths cor-
responding to the season of measurement.
Historic Milestones in Solar and Infrared Radiation Measurement 39

FIGURE 3.3 Ångström electrical compensation pyrheliometer continues to serve as a reference instrument


for some countries. This unit was operated during the Eleventh IPC at the PMOD, WRC in 2010.

measurement devices, the Ångström pyrheliometer remains in service today as a primary reference
instrument for many solar measurement laboratories (Finsterle 2016). The design uses two strips of
Manganin2 foil painted on one side with Parson’s optical black lacquer. The strips are mounted side
by side at the base of a collimator tube and aligned with two rectangular apertures at the front of the
instrument (Figure 3.3). A thermojunction is attached to the back of each receiver strip. Electrical
leads from the strips and the thermojunctions are connected to the instrument controls. The operator
controls a reversible shutter at the front of the collimator to alternately shade and unshade each of
the strips at 90-second intervals. In operation, the shaded strip is heated by a known electrical cur-
rent to the same temperature as the exposed strip. The rate of energy absorption by the exposed strip
(solar heating) is thermoelectrically compensated by the rate of energy supplied electrically to the
shaded strip (electrical heating). The thermojunctions attached to the back of each strip are used to
monitor their temperature equivalence using a galvanometer or digital null detector. Observations of
solar irradiance consist of a series of shade–unshade measurements in which each strip is alternately
exposed and shaded. The DNI is determined from the equivalence of radiant and electrical power:

DNI ⋅ α ⋅ b = C ⋅ r ⋅ i 2 (3.1)

where DNI is the direct normal irradiance (Wm−2), α is the absorptance of the receiver strip per
unit length, b is the width of the receiver strip [cm], C is a constant based on measurement units
(9,995.175), r is the resistance of the receiver strip per unit length (ohms/cm), i is the electric current
through the shaded strip (amperes)
In practice, the instrument manufacturer supplied a constant K defined as

K = C ⋅ r / (α ⋅ b ) (3.2)

2 Manganin is an alloy of typically 86 percent copper, 12 percent manganese, and 2 percent nickel.
40 Solar and Infrared Radiation Measurements

to simplify the irradiance determination

DNI = K ⋅ i 2 (3.3)

The instrument functioned as the European absolute reference beginning in 1905 and ended in 1956
with the adoption of the International Pyrheliometric Scale (IPS) that resolved the differences with
the Smithsonian scale that was established in 1913 (Fröhlich 1991).

3.3.4 CallendaR PyRanometeR (1898)


The Callendar Pyranometer was invented in 1898 by H. L. Callendar. When introduced, this
pyranometer was called a bolometric sunshine receiver. An improved version was available in 1905
(Callendar and Fowler 1906). The  instrument used four platinum wire grids wound on strips of
mica to form an electrical resistance thermometer. As shown in Figure 3.4, one pair of grids was
painted black to enhance the absorption of solar radiation, and the other pair was left as reflect-
ing platinum metal. The checkerboard receiver had an area of about 33.6 cm2 and was mounted
in a protective evacuated glass bulb with an outside diameter of about 9.1  cm (Coulson 1975).

FIGURE 3.4 The Callendar pyranometer was a successful all-weather design based on platinum resistance
thermometry. One of the blackened receiver sections has lifted from the detector surface.
Historic Milestones in Solar and Infrared Radiation Measurement 41

The temperature difference between the pairs of grids when exposed to sunlight was proportional
to the incoming solar radiation. The  popularity of this early pyranometer was due in large part
to the integrated data-recording mechanism. A  self-adjusting Wheatstone bridge with a pen and
recording drum produced an ink trace of the temperature difference in the receiver pairs with time.
A  calibration factor was supplied with each instrument to convert the recorded data into global
irradiance. The Callendar pyranometer enjoyed considerable popularity during the first two decades
of the twentieth century. An example of this pyranometer was used at the University of Cambridge
School of Botany for more than 50 years and agreed within 1 percent (more specifically, the cor-
relation coefficient was 0.99) when compared with the British Meteorological Office pyranometer
(Beaubien et al. 1998). The U.S. Weather Bureau used this instrument from 1908 until the last unit
was removed from service (Hand 1941).

3.3.5 ÅngstRöm and tuliPan PyRgeometeRs (1899)


Some of the first reported measurements of nocturnal radiation, infrared (longwave) downwelling
irradiance, were collected by two radiometers designed by Knut Ångström as part of a program
by the Smithsonian Institution to select four world observatories for measuring solar radiation
(Dorno 1923). The measurements were used to determine the overall cloud climatology as part of
the search for locations suited for solar observatories. Measurement comparisons of the Tulipan
and the Ångström pyrgeometers were reported at the Physical Meteorological Observatory in
Davos, Switzerland, from November 1920 to October 1921. The Tulipan design was based on “the
over-distillation of ether” using a blackened hollow sphere as an absolute blackbody (Dorno 1920).
The other Ångström design, which would later be called a pyrgeometer, consisted of four Manganin
strips, of which two were blackened and two were polished. The blackened strips radiated to the
atmosphere while the polished strips were shielded. The electrical power required to equalize the
temperature of the four strips was assumed to equate to the longwave irradiance. Based on 16 nights
of operation, the Tulipan pyrgeometer had a persistent high bias of 40 percent compared with the
reference Ångström pyrgeometer.

3.4 OPERATIONAL RADIOMETERS OF THE TWENTIETH CENTURY


3.4.1 abbot silveR-disk PyRHeliometeR (1906)
At the beginning of the twentieth century, continuing investigations to determine the solar constant
(now referred to as the total solar irradiance, or TSI) suggested the need for establishing a mea-
surement standard for solar irradiance. Seeking improved measurement accuracy, Charles Greeley
Abbot developed his pyrheliometer based on the principles of calorimetry. Designed as a mea-
surement reference, the instrument was disseminated in quantity by the Smithsonian Institution
(more than 100 copies were constructed and distributed internationally). As shown in Figure 3.5,
a blackened silver disk (38 millimeters (mm) in diameter and 7 mm thick) is positioned near the
base of a collimating tube limiting the instrument field of view to 5.7 degree full angle. In the final
design, a mercury-in-glass thermometer was inserted into a radial chamber bored into the edge of
the disk. The disk was filled with mercury for rapid heat transfer from the blackened receiver to the
thermometer. The mercury was sealed in the disk using a thin steel liner, surrounded by cord and
shellacked, and wax was used to seal the thermometer stem (Coulson 1975). Three fine steel wires
held the disk inside a copper box. This assembly then was enclosed in a wooden box to provide some
thermal insulation from changes in ambient temperature.
Operation of the instrument required properly aligning the receiver disk with the Sun, alter-
nately shading and unshading the receiver, accurately keeping time to 1/5th second, and reading
the thermometer to a few hundredths of a degree Celsius. Summarizing the procedure originally
42 Solar and Infrared Radiation Measurements

specified by Abbot (1911), it is apparent the instrument design was best suited for measurements
under very clear and stable atmospheric conditions:

Time (seconds) Operator Action


0 Align instrument with the sun, open instrument cover, close shutter
80 Read thermometer
180 Read thermometer, open shutter, check alignment
200 Read thermometer
300 Read thermometer, close shutter
320 Read thermometer
420 Read thermometer, open shutter

Note: Continue the readings in above order as desired.

Determining the direct normal irradiance required the correction of all temperature readings for
air, stem, and bulb temperatures, and multiplying the corrected temperature differences by a cali-
bration constant for the specific instrument used for the measurements. The calibration constant
was obtained by direct comparison with the Smithsonian water-flow pyrheliometer traceable to the
Smithsonian 1913 Scale of Radiation.
The importance of the silver-disk pyrheliometer to the accurate measurement of solar radia-
tion continued for several decades. As late as 1952, the WMO recommended the instrument for
measuring DNI.3

FIGURE  3.5 Design elements of the Abbot silver-disk pyrheliometer used for measuring direct normal
irradiance: Thermometer (T) measures the disk (D) temperature located in the base (B) of the instrument
separated by a collimating tube (C) defining the field of view that can be shuttered (S) from the DNI.

3 Meeting of the Subcommission on Actinometry, Brussels, 1952.


Historic Milestones in Solar and Infrared Radiation Measurement 43

3.4.2 smitHsonian WateR-floW PyRHeliometeR (1910)


Also developed by C. G. Abbot at the Smithsonian Institution, this absolute pyrheliometer, the
Smithsonian Water-Flow Pyrheliometer, was based on the measurement principles of calorim-
etry. As shown in Figure  3.6, the water-flow pyrheliometer measures DNI entering tube (BB),
passing through a precision aperture (C), and absorbed by a blackened chamber (AA) with a
cone at the base of the chamber to enhance absorption of the radiation. The hollow cylinder (K)
insulates the receiving chamber from changes in ambient air temperature. A spiral watercourse
encloses the receiving chamber. Distilled water at constant temperature is supplied to an inlet
tube (E) where the water enters the chamber wall at a constant rate and passes through a spiral
ivory piece (D1) and exits by a similar ivory spiral (D2). A measured length of “hair-like” plati-
num wire is threaded through each of the ivory spiral channels such that the water bathes the
platinum wires, forcing them to the temperature of the water entering and exiting the instrument.
When the instrument is exposed to DNI, the resulting temperature difference measured at D1 and
D2 and the known aperture area are used to compute solar irradiance. To confirm the water-flow
temperature measurements, a coil of wire (H) is used to electrically heat the receiving chamber
with a measured current.
These electrical substitution measurements were used to confirm the equivalence of solar and
electrical power performance of the instrument. Abbot’s design incorporated key features in use
today by the pyrheliometers defining the World Radiometric Reference: (1) cavity enhancement of
the receiver to increase photon absorption, (2) insulating “muffler” to limit heat exchange between
the internal receiver and exterior surfaces of the instrument, and (3) electrical substitution by means
of a wire-wound heater at the receiver for comparing measurements of solar-induced heating with
a known amount of electrical heating. Although the instrument did not become widely used, it was
the basis for the Smithsonian 1913 Scale of Radiation for solar radiation measurements (Abbot and
Aldrich 1932). The instrument’s function as the American absolute reference ended in 1956 with
the adoption of the IPS in 1956, which resolved the differences with the Ångström scale developed
in 1905 (Fröhlich 1991).

3.4.3 maRvin PyRHeliometeR (1910)


Prior to serving as chief of the U.S. Weather Bureau from 1913 to 1934, C. F. Marvin was developing
improved instruments for measuring bright sunshine duration and a pyrheliometer for measuring
solar DNI. Similar in many respects to the Abbot silver-disk pyrheliometer, the Marvin pyrheliome-
ter used a resistance wire instead of a mercury-in-glass thermometer for measuring the temperature

FIGURE 3.6 Smithsonian water-flow pyrheliometer measures DNI by monitoring the temperature rise of
water between points D1  and D2. Abbot’s design introduced the concepts of cavity enhancement for absorb-
ing solar radiation and electrical power substitution used in modern pyrheliometers. (From Abbot, C.G.,
The  Sun and the Welfare of Man, in Smithsonian Series, Vol.  2, C. G. Abbot ed., The  Series Publishers,
New York, 1929.)
44 Solar and Infrared Radiation Measurements

of the blackened silver-disk receiver (Foote 1918). An electrically operated shutter mounted in front
of the collimating tube aperture provided for automated shade and exposure of the receiver to solar
radiation. Due to its simpler operation, the Marvin pyrheliometer replaced the Ångström pyrheliom-
eter in 1914 as the instrument used by the U.S. Weather Bureau (Covert 1925).

3.4.4 ÅngstRöm PyRanometeR (1919)


Anders Knutsson Ångström, Swedish physicist and meteorologist and son of Knut Ångström, used
his father’s principle of electrical compensation developed for the Ångström pyrheliometer as the
basis for his pyranometer nearly two decades later. As shown in the schematic diagram in Figure 3.7,
the Ångström pyranometer detector consisted of two white strips and two black strips mounted on
an insulating framework at the end of a nickel-plated cylinder. A glass hemisphere was mounted to
a metal disk to protect the detector from the weather. Thermojunctions were attached to the backs
of the strips in good thermal contact and electrically isolated. The operation of the pyranometer was
based on maintaining the temperature of the white strips by supplying a measured electrical heating
to match the amount of radiant energy absorbed by the black strips (Coulson 1975). In spite of some
attractive design attributes, the Ångström pyranometer was never put into widespread use.

3.4.5 kiPP & Zonen solaRimeteR (1924)


A significant advance in the practical measurement of solar radiation is attributed to Moll (1923),
who, as a physics professor at the University of Utrecht, designed the Moll thermopile. His design,

FIGURE 3.7 Schematic diagram of the Ångström electrical compensation pyranometer with white thermo-
pile strips (A), black thermopile strips (B), glass hemisphere (G) mounted in a supporting disk (D), desiccant
chamber (C), adjusting screws (S), and spirit level (L). (From Coulson, K.L., Solar and Terrestrial Radiation,
Methods and Measurements, Academic Press, New York, 1975.)
Historic Milestones in Solar and Infrared Radiation Measurement 45

FIGURE 3.8 Concept of the Moll thermopile using thermojunctions of thin (0.005 mm) blackened strips of
Manganin and constantan joined at A and soldered at points B and C to copper posts P. The mounting posts
are fastened to a thick brass plate about 2 cm in diameter and coated with lacquer to provide electrical isolation
and good thermal conductivity for the active junctions.

depicted in Figure 3.8, addressed the measurement performance characteristics still sought in mod-
ern pyranometers and pyrheliometers:

• High level of responsivity (detector output signal produced per unit of incident solar
irradiance)
• Rapid response to changes in solar irradiance (instrument time constant)
• Linearity over the range of expected levels of solar irradiance (consistent responsivity)
• Uniform response to radiation within the solar spectrum (generally 0.3–3.0 micrometers
(µm))
• Minimum influence of ambient temperature on the detector output

In  1924, the Moll thermopile was used by Dr. Ladislas Gorczynski, director of the Polish
Meteorological Institute, as the basis for both a pyrheliometer and pyranometer. The  resulting
Moll–Gorczynski pyranometer was presented to Kipp & Zonen, a scientific instrument company in
Delft, the Netherlands, and by 1927, the company was offering “Solarimeters and Pyrheliometers”
in their sales catalog.4 The handmade Moll thermopiles continued to serve as the basis of Kipp &
Zonen radiometers until 1989 when the designs for models CM 1, CM 3, CM 5, and CM 6B were
based on electroplated polyamide foil thermopiles.

3.4.6 RobitZsCH bimetalliC aCtinogRaPH (1932)


Based on a simple thermo-mechanical design, this portable and self-contained recording pyranom-
eter, the Robitzsch bimetallic actinograph, was popular for monitoring solar radiation at remote
locations and contributed to the measured data used for estimating the world distribution of GHI
(Löf et al. 1966). Since its introduction in 1932, the instrument design has been refined and remains
commercially available today. Designed by Max Robitzsch (1887–1952), a German meteorologist
and professor at the Meteorological Observatory at Lindenberg, the instrument uses a blackened

4 A brief history of Kipp & Zonen is available from http://www.kippzonen.com.


46 Solar and Infrared Radiation Measurements

bimetallic strip about 8.5 × 1.5 cm mounted under a protective glass dome about 10 cm in diameter.
The differing thermal expansion properties of the materials used to form the bimetallic strip create
a simple bimetallic thermometer that responds proportionally to changes in solar irradiance. Using
an ink pen and a mechanical linkage attached to the free end of the bimetallic strip, the Robitzsch
design records the receiver deflections on chart paper wrapped on a clock-driven drum. The drum
rotation can be made in 24 hours or in 7 days to produce records of daily irradiation. The bimetallic
receiver, mechanical linkage, and recording drum were housed in a weather-resistant metal case
(Figure 3.9). Prior to the 1970s, the measurement uncertainties introduced by a variety of unde-
sirable response characteristics resulted in daily total irradiation within ±10 percent of the best
performing instruments (Coulson 1975). More recent design improvements have addressed these
shortcomings:

• Calibration of the rectangular receiver varied greatly with solar azimuth and elevation.
• Pen zero-point depended on the varying thermal exchanges with the day/night longwave
atmospheric conditions; receiver deflections varied with ambient temperature.
• Time response was relatively slow at about 15 minutes due to the large thermal mass of
the receiver.
• There were optical distortions caused by the glass dome.

3.4.7 ePPley 180° PyRHeliometeR (1930)


To address the needs of agriculture and to advance the meteorological observing capabilities in the
United States in the early 1920s, the Weather Bureau and the National Bureau of Standards devel-
oped a new instrument for measuring total hemispheric solar irradiance on a horizontal surface
(Kimball and Hobbs 1923). Marvin suggested the black and white annular ring concept for the ther-
mopile detector used in the “180° pyrheliometer” design as shown in Figure 3.10. Eppley Laboratory,

FIGURE 3.9 The Robitzsch bimetallic actinograph was a popular instrument for measuring daily amounts
of solar irradiance from about 1932 to 1970 at meteorological stations around the world.
Historic Milestones in Solar and Infrared Radiation Measurement 47

FIGURE 3.10 The Eppley Laboratory, Inc., model 50 “light bulb” pyranometer used from 1951 to 1975 by
the National Weather Service in their SOLRAD network of 61 stations. (Courtesy of The Eppley Laboratory,
Newport, RI.)

Inc. introduced the first pyranometer based on the Weather Bureau thermopile at a meeting of
the American Meteorological Society in May 1930 and continued to produce the radiometer with
improvements for more than 30 years (Coulson 1975). The National Weather Service (formerly the
U.S. Weather Bureau) used variants of this pyranometer from 1951 to 1975 in their SOLRAD net-
work (NREL, 1992).
The  thermopile detector consisted of gold-palladium and platinum-rhodium alloys with 10
(early model) or 50 (later model) thermojunctions in good thermal contact but electrically insu-
lated from the concentric rings of the annular design. Measuring about 29 mm in diameter, the
detector was made from two thin silver rings and a small inside disk. The  outer ring and the
small disk were coated with magnesium oxide, a material with high reflectance in the spectral
range of incident solar radiation. Depending on the year of manufacture, the inner ring was
coated with either lamp black or Parson’s optical black lacquer to provide for low reflectance
at solar wavelengths (SOLMET 1979). The  glass bulb was about 76  mm in diameter, made of
soda-lime glass about 0.6 mm thick, and filled with dry air at atmospheric pressure at the time
of assembly (Coulson 1975). This glass envelope shape was the basis for the common name for
this pyranometer as the Eppley Light Bulb. Instrument performance was affected by the degrada-
tion of the soda-lime glass transmittance, the degradation of black detector paint’s absorptivity
48 Solar and Infrared Radiation Measurements

with exposure, the lack of detector temperature compensation (except for later models), and the
optical distortions associated with the blown-glass envelope. The Parson’s optical black lacquer’s
absorptivity deteriorated significantly under exposure to sunlight, turning from black to gray to
green. The last production version of this pyranometer was named model 50 to reflect the number
of thermojunctions in the detector. In  response to the energy crisis of the 1970s, considerable
effort was applied to rehabilitating the historical measurement records based on this pyranometer
at the National Climatic Data Center to reduce the effects of these instrument performance char-
acteristics on data uncertainty (SOLMET 1978). The resulting data archive provides the original
measured value with additional corresponding “Engineering Correction” data that account for
known instrumentation deficiencies:

1. Calibration changes
2. Solar radiation measurement scale differences
3. Midscale chart recorder setting
4. Pyranometer detector paint degradation
5. Calibration “cross-match problem” due to the uses of Parson’s black and lamp black paints
among the network instruments and calibration references
6. Temperature response characteristics

In total, the amount of adjustment to these hourly solar irradiance data could exceed 20 percent,
depending on the pyranometers and chart recorders deployed over time at a SOLMET station.

3.4.8 ePPley model PsP (1957)


The first model of the precision spectral pyranometer (PSP) developed by The Eppley Laboratory,
Inc. was introduced in 1957 (Marchgraber and Drummond 1960). The PSP has been widely accepted
for deployment on land, sea, and aircraft platforms (Figure 3.11). The advanced design improve-
ments of the PSP at the time it was introduced included:

• Electrical compensation of the detector response to ambient temperature changes


between –20°C and +40°C
• Precision optics protecting a smaller thermopile detector than previously used in black-
and-white designs resulting in better angular response than previously available
• A single Parson’s optical black painted receiver to eliminate spectral reflectance and dura-
bility issues associated with the white magnesium oxide coating previously used in black-
and-white designs
• Improved heat transfer characteristics of the detector, domes, and instrument body
• Faster time response for measuring rapid changes of irradiance (1 second for 1/e)
• Rugged design for improved long-term operation
• Replaceable desiccant for maintaining dry air within the body
• Replaceable outer hemispheres for optional color filter measurements of spectral bands

The wire-wound copper–constantan thermopile produces about 10 microvolts (µV) per Wm−2 and


is mounted under two precision-ground Schott glass hemispheres as shown in Figure  3.12. The
double-dome design addresses the need to minimize the convective heat losses between the thermo-
pile and the surrounding environment resulting in thermal offsets proportional to the net infrared
radiation at the time of measurement. The  white sunshade protects the instrument’s cast bronze
body from thermal gradients caused by differential solar heating. An optional ventilator is avail-
able for providing forced air over the outer hemisphere to reduce the effects of dew, snow, dust, and
other potential contaminants from accumulating on the outer dome, hence improving the irradiance
measurements (Figure 3.13).
Historic Milestones in Solar and Infrared Radiation Measurement 49

FIGURE  3.11 The  Eppley Laboratory, model precision spectral pyranometer. (Courtesy of The  Eppley
Laboratory, Newport, RI.)

FIGURE 3.12 (a) Wire-wound F3 thermopile developed by The Eppley Laboratory for use in their model
precision spectral pyranometer. (b) Final assembly of an Eppley F3  thermopile ready for installation in a
model PSP. (Courtesy of The Eppley Laboratory, Newport, RI.)

3.4.9 yanisHevsky PyRanometeR (1957)


The  Yanishevsky (now Yanishevsky-Savinov) series of pyranometers served as the principal
instrument for solar irradiance measurements in the former Soviet Union and continues that tra-
dition in Russia and Central Asia today. The thermopile-type detector used in early models was
constructed in a square checkerboard pattern of alternate black and white squares and rectangles
(Figure 3.14). Later models have a radial pattern of alternate black and white segments. The lat-
est version pyranometer detector, Yanishevsky-Savinov M115-M, has alternating black and white
squares forming a planar detector surface. The thermocouples are composed of alternating strips of
Manganin and constantan. The hot junctions are blackened with carbon soot, and cold junctions are
50 Solar and Infrared Radiation Measurements

FIGURE 3.13 Two model PSPs installed in Eppley ventilators to reduce effects of outer hemisphere con-
tamination by dew, frost, snow, dust, and other sources. PSP with clear WG295 outer dome in foreground and
PSP with RG780 outer dome in background. (Courtesy of The Eppley Laboratory, Newport, RI.)

FIGURE 3.14 The Yanishevsky pyranometer served the needs of the former Soviet Union for solar irradi-
ance measurements. The black-and-white thermopile detector measures 3 × 3 cm.
Historic Milestones in Solar and Infrared Radiation Measurement 51

whitened with magnesium oxide. A single glass hemisphere protects the black-and-white receiver.
Early models also had an auxiliary opaque metal hemisphere for providing a measurement of zero
offsets. The measurement performance of the M115-M has been compared with other pyranometers
deployed as part of the International Polar Year (Ivanov et al. 2008). Results of the 24-day compari-
son of hourly irradiances measured by the Norwegian Polar Institute’s reference CMP 11 pyranom-
eter show agreement within accepted measurement uncertainty limits at 6.3  ±  5.6  Wm−2  over a
measurement range of 0–500 Wm−2 (Ivanov et al. 2008).

3.4.10 ePPley model niP (1957)


The normal incidence pyrheliometer (NIP) designed and manufactured by The Eppley Laboratory,
Inc. measures the solar “beam” DNI. As shown in Figure  3.15, the blackened wire-wound

FIGURE 3.15 (a) The Eppley Laboratory model normal incidence pyrheliometer mounted on an automatic
solar tracker for measuring DNI. (Courtesy of The Eppley Laboratory, Newport, RI) (b) NIP design elements
shown in cross section.
52 Solar and Infrared Radiation Measurements

copper–constantan thermopile receiver is at the base of a view-limiting tube providing the


one-to-ten geometry for the aperture-opening-to-tube-length responsible for the NIP’s 5.7 degree
field of view. In 1956, consistent with the existing solar tracker alignment capabilities and the need
to provide instrument design standards, the International Radiation Commission established pyr-
heliometer design recommendations that included this viewing geometry (Bolle 2008). Over the
years, the double-walled tube has been constructed with an outer wall made of chrome-plated brass,
white-painted brass, or stainless steel with the inner wall made of brass and fitted with light-baffle
diaphragms to limit stray light from reaching the thermopile receiver. The  viewing end of the
tube is fitted with a quartz window of 1 mm thickness (currently Infrasil II material). The sealed
double-walled construction is filled with dry air at atmospheric pressure at the time of manufacture.
This design improves measurement stability, especially during windy conditions. The pyrheliom-
eter has two flanges. The rear flange has three holes previously used for mounting on older solar
trackers. The front flange has a sighting diopter arrangement for alignment with the target on the
rear flange (Figure 3.16).
Recommended as one of the instruments for use during the 1957–1958 International Geophysical
Year (IGY), the NIP became a popular instrument for measuring DNI with these performance
characteristics:

• Sensitivity: Approximately 8 µV/Wm−2


• Impedance: Approximately 200 ohms
• Temperature dependence: ±1 percent over ambient temperature range  –20°C to +40°C
(Compensation can be supplied over other temperature ranges)
• Linearity: ±0.5 percent from 0 to 1,400 Wm−2
• Response time: 1 second (1/e signal)
• Mechanical vibration: Tested up to 20 g’s without damage
• Calibration: Referenced to Eppley primary standard group of pyrheliometers
• Size: 11 inches long
• Weight: 5 pounds

After 2015, a new model sNIP (“short NIP”) became available to address the needs for improved
measurement performance. Notably, the sNIP has a 5 degree field of view, faster time response,
reduced thermal conduction and convection effects, and meets the performance specifications of an
International Organization of Standards (ISO) Class A and a WMO High Quality Pyrheliometer.

FIGURE  3.16 Diopter target used on the rear flange of a NIP to determine alignment with solar disc.
The detector is fully illuminated when the sun spot image is fully within the white area of the target.
Historic Milestones in Solar and Infrared Radiation Measurement 53

3.4.11 ePPley model PReCision infRaRed RadiometeR (PiR) (1968)


The Eppley Laboratory, Inc., model precision infrared radiometer (PIR), is designed to measure atmo-
spheric or terrestrial infrared radiation over the wavelength range of 3.4–50 µm (Drummond 1970).
This pyrgeometer is similar in construction to the model PSP pyranometer with a single black wire-
wound copper–constantan thermopile receiver mounted in a cast bronze body that serves as the cold
junction. A protective dome surrounds the thermopile hot junctions and selectively transmits the infra-
red radiation to the receiver (Figure 3.17). A temperature compensation circuit nominally adjusts the
thermopile output for changes in ambient temperature between –20°C and +40°C (other temperature
compensation ranges are available). Early models of the PIR used a dome made of thallium bromide-
iodide (KRS-5) that was replaced in the 1980s with a silicon hemisphere coated with tellurium and zinc
selenide to improve thermal conductance and wavelength response of the instrument. As described in
more detail in Chapter 10, the PIR receiver responds to these sources of infrared radiation:

• Incoming atmospheric or terrestrial (depending on mounting orientation)


• Effective instrument case temperature
• Effective hemispheric dome temperature

Therefore, the PIR design includes thermistors in contact with the hemispherical “solar blind” filter
and the pyrgeometer case to provide temperature data coincident with the thermopile voltage gener-
ated during irradiance measurements. The irradiance can be computed from the calibration coef-
ficients determined for the thermopile, case emittance, and dome correction factor (see Chapter 11).
Historically, these calibration coefficients have been primarily determined from measurements
made with laboratory blackbody calibrators. Since 2003, pyrgeometer calibrations are traceable
to the interim World Infrared Standard Group (WISG) developed by the World Radiation Center
(WRC; Marty et al. 2003; Gröbner and Los 2007). The WISG is a group of pyrgeometers that form
a reference standard that includes two Eppley PIRs.

FIGURE 3.17 The Eppley Laboratory model PIR used for total hemispheric longwave radiation measure-
ments (foreground) with model PSP (background). (Courtesy of The Eppley Laboratory, Newport, RI.)
54 Solar and Infrared Radiation Measurements

3.4.12 PRimaRy absolute Cavity RadiometeR (PaCRad) (1969)


This  radiometer is historically significant for several reasons. Developed at the Jet Propulsion
Laboratory (JPL) to support the National Aeronautics and Space Administration (NASA) Deep
Space Network Energy Program, the PACRAD was needed to improve the measurement of simu-
lated solar output in space chambers. Nearing retirement, J. M. Kendall, Sr. was tasked with design-
ing a new radiometer after the thermal control system of at least one space probe did not function
as predicted by space simulator testing. The simulator design problem was linked to the inability
to accurately measure the simulated solar flux in the space chamber. The resulting JPL standard
total-radiation absolute radiometer designed by Kendall was used to measure the absolute total
radiation intensity with an absolute accuracy of 0.5 percent and to experimentally determine the
Stefan-Boltzmann constant to within 0.5 percent of the theoretical value (Kendall 1968). About this
time, the scientific community needed an absolute solar irradiance measurement reference. A series
of pyrheliometer measurement comparisons had produced solar irradiance measurement differ-
ences approaching 5 percent among the Ångström Scale of 1905, the Smithsonian Scale of 1913,
and the IPS of 1956 (Latimer 1973). Clearly, an improved international reference for solar radiation
measurement was needed.
Working from first principles of measurement, Kendall applied several new design concepts to
develop the PACRAD instrument (Figure 3.18). First, he introduced the cavity-shaped receptor to
enhance the detector absorptivity. Earlier pyrheliometer designs used flat plate detectors (i.e., disks
or rectangular plates) that were prone to reflecting photons regardless of surface coating and col-
limator design. Second, the PACRAD had a built-in electrical calibrating heater to provide cavity

FIGURE  3.18 The primary absolute cavity radiometer became the basis for the electrically self-calibrat-
ing absolute cavity radiometers used to define the WRR. (From Kendall, J.M., Primary Absolute Cavity
Radiometer, Technical Report 32-1396, Pasadena, CA, 1969.)
Historic Milestones in Solar and Infrared Radiation Measurement 55

heating accurately and equivalent to radiation heating. As a result, PACRAD irradiance measure-
ments were based on the ability to make accurate and traceable electrical power measurements
(i.e., voltage and resistance) and accurate dimensional measurements (i.e., aperture area) to produce
solar irradiance data in Wm–2. A third critical element in Kendall’s design was the attention to heat
transfer within the instrument. The materials selection, component design and placement, includ-
ing the compensation cavity, were based on the need to absorb the incoming solar radiation that is
converted to heating of the cavity receiver. The heat flows through a metallic thermal resistor to the
massive heat sink to produce a temperature difference of a fraction of a Kelvin. The temperature
difference is measured by a thermopile that is electrically calibrated as part of the measurement
procedure. Key to the PACRAD design and operation was the thorough characterization of the
effects of the various types of heat transfer within the radiometer and the resulting correction fac-
tors for all heat transfers that are not equivalent to the heating of the cavity by electrical substitution.
Three models of the PACRAD were built and compared with Eppley Ångström pyrheliometers
(Kendall 1969). Solar measurements taken at Table Mountain, California, on April 23, 1969, with
the three PACRAD models agreed with one another within 0.11 percent and the Ångström pyr-
heliometers agreed with each other within 0.18 percent. However, solar irradiance measurements
from the pyrheliometers were 2.3 percent lower than the PACRADs. This discrepancy was later
resolved with the establishment of the World Radiometric Reference (WRR) based on a series of
International Pyrheliometer Comparisons (IPCs) conducted by the WRC in Davos, Switzerland
(Fröhlich 1991).
During IPC-III in 1970, Kendall’s third model, PACRAD III, was observed to be very stable
and was the only absolute cavity radiometer in the comparison. By 1975, five other absolute cavity
designs had been developed and participated in IPC-IV (Fröhlich 1991; Zerlaut 1989):

• Active cavity radiometer (ACR) by Willson (1973)


• Eppley-Kendall PACRAD-type (EPAC) by the Eppley Laboratory, Inc. and J. Kendall
(Drummond and Hickey 1968)
• Model PMO developed by the Physikalisch-Meteorologisches Observatorium
• Model PVS by Yu. A. Skliarov at the Saratov University in the former Soviet Union
• Model TMI, PACRAD-type, by Technical Measurements, Inc.

Because of its performance in IPC-III, the PACRAD III was chosen as the reference instrument
for IPC-IV. The  weighted average of all instruments for all observations during IPC-IV was
1.0017 with a standard error of only 0.18 percent among the absolute cavity instruments (Zerlaut
1989). The subsequent establishment of the absolute WRR measurement scale is a result of the
excellent agreement among all of the absolute cavity instruments at IPC-IV and later comparisons.
In 1979, based on the results of IPC-IV, WMO adopted the WRR as the replacement for the IPS
of 1956 for the measurement of DNI. Thus, the PACRAD instrument served an important role in
establishing the WRR and providing for more accurate and consistent solar irradiance measure-
ments around the world.5

3.4.13 ePPley model 8-48 (1969)


Developed as a lower-cost replacement for the 180° pyrheliometer or Eppley Light Bulb pyranom-
eter described earlier, the Epply Model 8-48 offered improved measurement performance charac-
teristics with a more robust design than its predecessor. In place of a rather large and delicate glass
envelope surrounding the detector, a precision-ground optical Schott glass WG295  hemisphere

5 During this same period, other researchers were developing advanced pyrheliometer designs for solar irradiance
measurements from space. A summary of the results from Willson’s active cavity radiometer, NASA’s eclectic satellite
pyrheliometer, and others is available from Willson (1993).
56 Solar and Infrared Radiation Measurements

FIGURE  3.19 The  Eppley Laboratory, model 8–48  pyranometer exhibits minimal thermal offset due to
equal exposure of thermopile junctions to the effective sky temperature. (Courtesy of The Eppley Laboratory,
Newport, RI.)

protects the black-and-white segments of the differential thermopile detector from the elements
(Figure  3.19). The  black (hot) junctions were originally painted with Parson’s optical black and
the white (cold) junctions coated with barium sulfate. The relatively large receiver produced about
10 µV per Wm−2 and a time response of 5 seconds for a change in signal of 1/e. The challenge of
constructing a planar receiver from black and white segments resulted in angular response speci-
fications of ±2 percent for incidence angles 0–70° and ±5 percent for incidence angles 70–80°.
The balanced black-and-white receiver design, however, results in minimal thermal offset on the
order of 1 Wm−2 (Dutton et al. 2001) and negates the need for a protective sunshade over the body
and an inner dome to limit the convective heat losses to the surroundings. In the course of produc-
tion, these pyranometers have had bodies made from machined aluminum (early years) and painted
cast aluminum (current production). A  temperature compensation circuit nominally adjusts the
thermopile output for changes in ambient temperature between –20 and +40°C (other temperature
compensation ranges are available).

3.4.14 li-CoR model li-200sa (1971)


The origin of this now ubiquitous pyranometer was based on the need for a low-cost radiometer to
support the work of the Department of Soil and Water Sciences at the University of Wisconsin (Kerr
et al. 1967). The silicon photodiode detector used in the original design was mounted below a plastic
diffuser to provide adequate angular response and coupled with a purpose-built solid-state integrator
for recording daily sums of global irradiance. The prototype radiometer was deployed in a 17-station
mesoscale network in Wisconsin to study the temporal and spatial variations of solar radiation for
applications in bioclimatology (Kerr et al. 1968). Lambda Instruments Corporation produced the first
commercial version of the design in 1971. The company name was changed to LI-COR, Inc. in 1978.
As shown in Figure 3.20, the LI-200SA pyranometer is compact (2.38 cm diameter × 2.54 cm tall)
and lightweight (28  g). Because the detector is based on the photoelectric effect rather than the
thermoelectric effect used for thermopile radiometers, the LI-200SA has a very fast time response
Historic Milestones in Solar and Infrared Radiation Measurement 57

FIGURE  3.20 LI-COR, model LI-200SA  photodiode-based pyranometer (upper) and the instrument’s
solar spectral response (lower) between 400 and 1,100 nm compared with the spectra for extraterrestrial, sea
level, and blackbody radiation corresponding to a temperature of 5900 K. (From LI-COR Environmental,
Lincoln, NE. With permission.)

(10  microseconds (µsec)). The  silicon photodiode, however, does not  respond uniformly to all
wavelengths and does not respond at all to wavelengths greater than about 1,100 nanometers (nm)
(Figure 3.20). This limited spectral response increases the measurement uncertainty when compared
with thermopile-based detectors (King and Myers 1997). In 2015, an improved Model LI-200R was
introduced with design improvements addressing the operation and maintenance of the instrument
(Thomas et al. 2015).

3.4.15 Rotating sHadoWband RadiometeR (1975)


Using a single LI-COR LI-200SA pyranometer to measure global and diffuse solar irradiance for
computing the coincident direct irradiance was the design basis for the DIAL radiometer (Wesely
1982). The original DIAL radiometer prototype got its name from two aspects of the instrument
design (Figure  3.21). When viewed from the axis of rotation, the shade assembly resembles the
face of a dial with a revolving position indicator formed by the shade band. Also, the instrument
name could be derived from its abilities to measure diffuse (D), estimate direct (I), and measure
58 Solar and Infrared Radiation Measurements

FIGURE  3.21 The  DIAL radiometer used a rotating shade (resembling a dial) to alternately shade and
unshade a fast-response pyranometer to measure diffuse (D) and global (AL) irradiances for estimating the
corresponding incident (I) direct normal irradiance component. (From Wesely, M.L., J. App. Meteorol., 21,
1982.)

global (AL) solar irradiance components. The design innovations addressed the needs for a low-cost
radiometer system that would be simple to install, operate, and maintain for purposes of estimating
the broadband solar irradiances at the Earth’s surface and the corresponding atmospheric turbidity.
The innovations included the following:

• No mechanical adjustment to the shading device


• Compact and lightweight mounting arrangement
• Commercially available pyranometer using a silicon photodiode mounted beneath a
cosine-correction diffuser
• Integrated assembly, suitable for field applications in remote areas

The DIAL radiometer concept was later commercialized for applications in solar resource assess-
ment and atmospheric science. Advanced commercial versions based on the DIAL radiometer con-
cept are now known as a rotating shadowband radiometer (RSR),6 rotating shadowband pyranometer
(RSP),7 rotating shadowband irradiometer (RSI)8 or multifilter rotating shadowband radiometer
(MFRSR).9 The original DIAL radiometer design used a LAMBDA pyranometer positioned under
a motorized shading band that completed one rotation about every 4 minutes. The band was ~0.9 cm

6 Irradiance, Inc. Model RSR2 (http://www.irradiance.com/)


7 Ascension Technology, Inc., Model RSP formed the basis for the RSR2
8 CSP Services GmbH, Model Twin-RSI (http://www.cspservices.de)
9 Yankee Environmental Systems, Inc., Model MFR-7 (http://www.yesinc.com/products/radvis.html)
Historic Milestones in Solar and Infrared Radiation Measurement 59

wide and ~6.0 cm away from the 0.8 cm diameter diffuser on the pyranometer. With a shadowband
rotation speed of about 0.25 revolutions per minute (rpm), the pyranometer detector was shaded
for about 0.7 seconds. This shading interval was sufficient for strip-chart recorders of the day to
respond and display the minimum signal from the fast-response photo-diode detector used in the
pyranometer. The  diffuse and global irradiance data were based on the minimum (shaded) and
maximum (unshaded) pyranometer signals, respectively.
Initial comparisons of vertically integrated estimates of atmospheric turbidity made with a DIAL
radiometer and two broadband thermopile pyranometers (The Eppley Laboratory, Inc. model 8-48—
one unshaded and one shaded) suggested single reading accuracy of ±30 percent for turbidity mea-
surements for solar zenith angles less than 65 degrees. When data from the DIAL radiometer were
corrected for angular response and excessive shading of diffuse sky radiance due to shade bandwidth,
averaged results compared within ±2 percent. While the Eppley 8-48 is a broadband pyranometer
with little spectral dependence compared with a photodiode-based pyranometer, this  comparison
missed the spectral response problem seen on clear days with rotating shadow band radiometers
because the spectral effect only becomes significant (>3%) for angles greater than 65 degrees.
Since the introduction of the rotating shadowband concept, considerable effort has been made to
characterize and improve the system design and data correction algorithms (Augustyn et al. 2004;
Harrison et al. 1994; King and Myers 1997; Michalsky et al. 1988; Rosenthal and Roberg 1994;
Stoffel et al. 1991; Vignola 2006; Wilcox and Myers 2008; Wilbert et al. 2015, 2016). Specifically,
these response characteristics of an RSR10 continue to be studied:

• Spectral response: From the outset of DIAL radiometer design, the need to provide fast
response to changing irradiance levels has been met by the use of a pyranometer with a
silicon photodiode detector. The time response for such a detector is on the order of 10 µsec
compared with 1–5 seconds (or longer) for thermal detectors. Because the photodiode
detector responds non-uniformly to the solar radiation between about 300 and 1,100 nm,
various correction schemes have been developed to simulate the corresponding broadband
measurements. Because a silicon photodiode is less responsive near the 400 nm spectral
region (i.e., blue), so-called spectral mismatch errors of up to 30 percent can occur for mea-
surements of clear-sky diffuse irradiance (Vignola 1999). The advent of small thermopile
detectors has provided the opportunity for fast-acting and broadband spectral response
when used in an RSR (https://midcdmz.nrel.gov/apps/sitehome.pl?site=TSR).
• Temperature response: The  electrical current produced by the photoelectric effect in a
photodiode detector increases with increasing temperature. Corrections to the measured
solar irradiance have been made with coincident measurements of air temperature near
the RSR and from detector temperature measurements from a thermistor placed near
the photodiode detector. RSR irradiance measurement uncertainties due to temperature
response can be on the order of 3 percent for a 40°C temperature change (i.e., from winter
to summer) (King and Myers 1997; Wilbert et al. 2015).
• Angular response: Computing accurate direct normal irradiance from RSR measure-
ments of global and diffuse components requires adequate design and characterization
of the pyranometer for angular or “cosine” response. Typically, commercially available
pyranometers have adequate angular response characteristics between incidence angles
of 0 and 60° as measured normal to the detector surface. Pyranometer angular response
is commonly undefined for incidence angles greater than 85 degrees. Accounting for the
specific angular response characteristics of a pyranometer can reduce the systematic biases
in direct normal irradiance estimated from RSR measurements by a few percent for solar
zenith angles less than about 70 degrees.

10 Here, an RSR will refer to all such instruments estimating DNI from measurements of GHI and DNI using a rotating
shadowband.
60 Solar and Infrared Radiation Measurements

FIGURE 3.22 Multifilter rotating shadowband radiometer detector arrangement as positioned beneath the
diffuser.

Coming full circle, the original intent of the DIAL radiometer design has been fulfilled by the
MFRSR with its capabilities to measure solar spectral irradiance at selected wavelengths allowing
for the retrieval of atmospheric optical depths of aerosols, water vapor, and ozone (Harrison and
Michalsky 1994). The MFRSR uses six independent interference filter-silicon photodiode combina-
tions, mounted in a temperature-controlled enclosure, to detect spectral irradiance using six nar-
rowband filters (10-nm wide). One unfiltered channel is used to obtain broadband solar irradiance
estimates (Figure 3.22). As a result, the MFRSR design provides for the spectral measurement of
diffuse and global irradiance and estimates of direct normal irradiance.

3.4.16 WoRld standaRd gRouP (1979)


The Physikalisch-Meteorologisches Observatorium Davos (PMOD) established the World Radiation
Center (WRC) to guarantee international homogeneity of solar radiation measurements by maintain-
ing the World Standard Group (WSG) of electrically self-calibrating absolute cavity radiometers.
The WSG defines the World Radiometeric Reference (WRR) as a detector-based measurement refer-
ence (Fröhlich 1978). During the most recent IPC-XII, the WSG comprises four well-characterized
instruments: PMO-2, PMO-5, CROM-2L, and PACRAD-III (Figure 3.23). The WSG serves as the
reference measurements for the IPC held every 5 years at the WRC (WMO 2016).
Historic Milestones in Solar and Infrared Radiation Measurement 61

FIGURE 3.23 The WRR is determined by a group of absolute cavity radiometers named the WSG. As of
2019, the WSG is composed of six instruments: PMO-2, PMO-5, CROM-2L, PACRAD-III, TMI-67814, and
HF-18748.11

3.5 RECENT ADVANCES IN SOLAR MEASUREMENTS


3.5.1 automatiC HiCkey–fRieden Cavity RadiometeR
The Eppley Laboratory, Inc. model automatic Hickey–Frieden (AHF) absolute cavity pyrheliom-
eter is the automated version of the model HF radiometer that first participated in an IPC in 1980
(Zerlaut 1989). Similar to the PACRAD and in collaboration with JPL, the HF cavity radiometer
was developed to serve as a reference standard for the measurement of DNI (Drumond et al. 1968).
Until recently, HF-18748 has been one of the WSG absolute cavity radiometers defining the WRR
at the WRC in Davos, Switzerland (WMO 2016).
An electrically self-calibrating instrument, the sensor consists of a balanced cavity receiver pair
attached to a circular wire-wound and plated thermopile (Figure 3.24). The blackened cavity receiv-
ers are fitted with heater windings that allow for absolute operation using the electrical substitution
method, which relates radiant power to electrical power in SI units. The forward cavity views the
DNI through a precision aperture immediately in front of the cavity receiver and behind a view-
limiting aperture. The precision aperture is made of Invar, an alloy of 63.8 percent iron, 36 per-
cent nickel, and 0.2 percent carbon with a very low coefficient of thermal expansion. The area is

11 Due to recently suspected instrument instabilities, TMI-67814 and HF18748 were not used to realize the WRR for the
latest International Pyrheliometer Comparisons (IPC). After inspection and repair, the two radiometers could be evalu-
ated during the inter-IPC period for continued participation in the WSG (WMO 2016).
62 Solar and Infrared Radiation Measurements

FIGURE 3.24 The Eppley Laboratory, model AHF electrically self-calibrating absolute cavity radiometer.
(Top: Courtesy of The Eppley Laboratory, Newport, RI; bottom: From Reda, I., NREL Report No. TP-463-
20619, 1996.)

nominally 50 mm2 and is measured for each unit. The rear receiver views an ambient temperature
blackbody. The  HF radiometer element with a baffled view-limiting tube and blackbody are fit-
ted into an outer tube that acts as the enclosure for the instrument and reduces thermal gradients
in the instrument due to wind and rapid changes to ambient temperature. The model AHF has an
automatic shutter attached to the outer tube for simplifying the self-calibration and measurement
controls.
The  operation of the cavity radiometer and the measurements of the required calibra-
tion parameters are performed using digital electronics (AHF and later versions of the HF).
Historic Milestones in Solar and Infrared Radiation Measurement 63

The control functions include setting of the calibration heater power level, activation of the
calibration heater, selection of the signals to be measured, and control of the digital multi-
meter measurement functions and ranges. The  measured parameters include the thermopile
signal, the heater voltage, and the heater current that is measured as the voltage drop across a
10-ohm precision resistor. The instrument temperature may also be measured using an inter-
nally mounted thermistor. The  digital multi-meter resolution of 100  nanovolts (nV) allows
for the measurement of a thermopile signal change equivalent in radiation to approximately
0.1 Wm−2 .
Although these are absolute devices, the radiometers are compared with the Eppley reference
cavity radiometers that have participated in multiple IPCs and other intercomparisons to provide
direct calibration traceability to the WRR.

3.5.2 total iRRadianCe monitoR (tim)


The  Laboratory for Atmospheric and Space Physics at the University of Colorado at Boulder
designed, built, and operates the solar total irradiance monitor (TIM) as one of the instruments
developed for NASA as part of the SOlar Radiation and Climate Experiment (SORCE) launched
in January 2003 (Kopp et al. 2003). The instrument continues the 25-year record of space-based
measurements of TSI that began in 1979 with the Earth radiation budget (ERB) instrument and
continues with the active cavity radiometer irradiance monitor (ACRIM) and variability of solar
irradiance and gravity oscillations (VIRGO) radiometers. With an estimated absolute accuracy of
350 parts per million (ppm) (0.035%), the TIM can detect relative changes in solar irradiance to less
than 10 parts per million per year (ppm/yr) (0.001%/yr), allowing determination of possible long-
term variations in solar output. The TIM has four electrical substitution cavity radiometers (ESRs)
to provide for measurement redundancy and for monitoring sensor degradation via duty cycling.
The cavity interiors are etched nickel phosphorus (NiP) producing solar reflectances in the range
of 10−4. Each ESR can be independently shuttered from sunlight. The electrical power needed to
maintain fixed cavity temperature during shutter cycles determines the radiative power absorbed by
the cavity due to the entering solar power. The ESRs are balanced as pairs, one acting as a thermal
reference while the other is actively driven to the temperature of the corresponding reference cavity.
The use of phase-sensitive detection provides for the high measurement precision of the instrument
(Gundlach et al. 1996).
Comparing TSI measurements during 2008  from the TIM (∼1361  Wm−2) coincident with
those from ACRIM III (∼1363  Wm−2) and VIRGO (∼1366  Wm−2) suggests a negative bias of
about 5 Wm−2 for the TIM (https://www.pmodwrc.ch/en/research-development/solar-physics/tsi-
composite/). The TSI measurements resulted in a workshop hosted by the National Institute of
Standards and Technology (NIST) and NASA in 2005 to discuss instrument design, measurement
uncertainties, and stabilities in detail (Butler et al. 2008). Notable among the findings was the
relative placement of the view-limiting and precision apertures relative to the cavity receiver in
the instrument designs (see Figure 3.25). The TIM positions the smaller precision aperture at the
front of the instrument so that only the light that is intended for measurement enters the cavity.
All other cavity radiometers place the view-limiting aperture at the front, allowing excess light
into the instrument interior that can lead to erroneously high irradiance measurements from scat-
tered light.
The TIM value of total solar irradiance is 1360.8 ± 0.5 Wm−2 during the 2008 solar minimum
period (Kopp and Lean 2011). The discrepancy between the canonical minimum value of TSI of
1365.4 ± 1.3 Wm−2 and the TIM value has been under investigation. The presently accepted value is
1360.8 Wm−2 value and is the value adopted for use in this book (Dudok de Wit et al. 2017; Fröhlich
2012; Kopp 2016).
64 Solar and Infrared Radiation Measurements

FIGURE  3.25 The  relative position of the view-limiting and precision apertures is different between the
designs of absolute cavity radiometers (left) and the total irradiance monitor (right). The optical layout of the
TIM places the small precision aperture at the front of the instrument so that only the solar radiation to be mea-
sured enters the cavity detector. With the precision aperture near the cavity detector (left), excess radiation is
allowed into the instrument interior that can lead to erroneously high irradiance measurements from scatter
(indicated by the arrows). (From Kopp, G. and Lean, J.L., Geophys. Res. Letters Frontier Article, 38, 201, 2011.)

3.5.3 CRyogeniC solaR absolute RadiometeR-measuRe tHe


integRal tRansmittanCe (CsaR-mitRa)
The WRC in Davos, Switzerland, and the National Physical Laboratory (NPL) in Middlesex, United
Kingdom, jointly developed the Cryogenic Solar Absolute Radiometer (CSAR) to provide direct
traceability of DNI measurements to the International System of Units (SI), reduce the overall uncer-
tainty of DNI measurements to ±0.01 percent, and become the new WRR for DNI measurements,
replacing the current WSG of electrically self-calibrating absolute cavity radiometers (Walter et al.
2017). At cryogenic temperatures, the copper cavity sensors used in the CSAR become thermally
super-conductive, limiting the errors introduced by the electrical substitution principle. However,
operating a terrestrial-based cryogenic cavity radiometer requires an entrance window. Therefore,
a companion instrument to Measure the Integral Transmittance of Windows (MITRA) is needed to
determine the attenuation of a window identical to the entrance window of the CSAR to correct the
irradiance values for window reflectance and absorptance losses. The solar irradiance measurements
performed with the CSAR-MITRA system support previous findings of a ~0.3 percent difference
between the SI radiometric scale represented by CSAR and the currently used WRR scale repre-
sented by the WSG.
Historic Milestones in Solar and Infrared Radiation Measurement 65

3.6 SUMMARY
The measurement of radiant energy remains a difficult and challenging science in spite of the fact
that fundamental radiometry has a long history of steady improvements. Beginning with the need to
satisfy basic curiosities of physicists and astronomers, the present interests in solar energy conver-
sion and global climate change continue to motivate the development of more accurate, reliable, and
economical radiometers for ground- and satellite-based observations. The current internationally
accepted measurement reference for solar irradiance, the WRR, is a detector-based standard with an
estimated measurement uncertainty of ±0.3 percent and a precision of ±0.1 percent. Improvement
of the WRR appears imminent as scientific and engineering advances are applied to the measure-
ment of solar energy.

QUESTIONS
1. Who is credited with developing the first pyrheliometer for measuring direct normal
irradiance?
2. What thermopile design facilitated the development of the modern pyranometer?
3. When was the WRR established as the international measurement standard for solar
irradiance?
4. What is the currently accepted value of the total solar irradiance?
5. Name the response characteristic contributing most to the estimated measurement uncer-
tainty of a photodiode-based pyranometer.
6. Name the type of radiometer designed to measure atmospheric or terrestrial infrared radia-
tion over the wavelength range of 3.4–50 µm.
7. In what year did the U.S. Weather Bureau begin the routine measurement of solar radiation
using a pyranometer?
8. What is the principal advantage of operating a radiometer at cryogenic temperatures?

REFERENCES
Abbot, C. G. 1911. The silver disk pyrheliometer. Smithsonian Miscellaneous Collection 56: 7.
Abbot, C. G. 1929. In: C. G. Abbot, ed., The Sun and the Welfare of Man. The Smithsonian Series Vol. 2.
New York: The Series Publishers.
Abbot, C. G., and L. B. Aldrich. 1916. The  pyranometer—An instrument for measuring sky radiation.
Smithsonian Miscellaneous Collection 66 (7).
Abbot, C. G., and L. B. Aldrich. 1932. An improved water-flow pyrheliometer and the standard scale of solar
radiation. Smithsonian Miscellaneous Collection 3182 (87): 8.
Ångström, K. 1893. Eine elegtrische Kompensationsmethode zur quantitativen Bestmmung strahlender
Warme. Nova Acta Regiae Societatis Scientiarum Upsala 3 (16) (Translated in Physical Review 1, 1894).
Ångström, K. 1899. The absolute determination of the radiation of heat with the electric compensation pyrhe-
liometer, with examples of the application of this instrument. Astrophysical Journal 9: 332–346.
Augustyn, J., T. Geer, T. Stoffel, R. Kessler, E. Kern, R. Little, F. Vignola, and B. Boyson. 2004. Update
of algorithm to correct direct normal irradiance measurements made with a rotating shadow band
pyranometer. Proceedings of Solar-,July 10–14, 2004. Portland, OR: American Solar Energy Society.
Beaubien, D. J., A. Bisberg, and A. F. Beaubien. 1998. Investigations in pyranometer design. Journal of
Atmospheric and Oceanic Technology 15: 677–686.
Bolle, H.-J. 2008. International radiation commissions 1896 to 2008: Research into atmospheric radiation
from IMO to IAMAS. International Association of Meteorology and Atmospheric Sciences, IAMAS
Publication Series No. 1. http://www.irc-iamas.org/resources/ (Accessed May 28, 2019).
Butler, J., B. C. Johnson, J. P. Rice, E. L. Shirley, and R. A. Barnes. 2008. Sources of Differences in On-Orbital
Total Solar Irradiance Measurements and Description of a Proposed Laboratory Intercomparison.
Journal of Research of the National Institute of Standards and Technology 113: 187–203.
Callendar, H. L., and A. Fowler. 1906. The horizontal bolometer. Proceedings of the Royal Society Series A.
77: 15–16.
66 Solar and Infrared Radiation Measurements

Coulson, K. L. 1975. Solar and Terrestrial Radiation: Methods and Measurements. New York: Academic Press.
Covert, R. N. 1925. Meteorological instruments and apparatus employed in the U.S. Weather Bureau. Journal
of the Optical Society of America 10: 299–425.
Crova, M. A. 1876. New pyrheliometer. In  J. D. Dana, B. Silliman, and E. S. Dana, eds., American
Journal of Science and Arts Volume XI, Third Series (January–June). New Haven, CN: Editors,
pp. 220–221.
Dorno, C. 1920. Suggestions concerning Dr. C. G. Abbot’s program for four world observatories for the obser-
vation of extraterrestrial solar radiation. Monthly Weather Review 48 (6).
Dorno, C. 1923. Progress in radiation measurements. Monthly Weather Review 50 (10).
Drummond, A. J. 1970. A survey of the important developments in thermal radiometry. Volume 14 of Advances
in Geophysics. New York: Academic Press.
Drummond, A. J., and J. R. Hickey. 1968. The Eppley-JPL solar constant measurement program. Solar Energy
12:217–232.
Drummond, A. J., W. J. Scholes, J. J. H. Brown, and R. E. Nelson. 1968. Radiation including Satellite
Techniques. Proceedings of the WMO/IUGG Symposium Held in Bergen, August  1968, No.  248.
Bergen, Norway: WMO.
Dudok de Wit, T., G. Kopp, C. Fröhlich, and M. Schöll. 2017. Methodology to create a new total solar irradi-
ance record: Making a composite out of multiple data records. Geophysical Research Letters 44 (3):
1196–1203. doi:10.1002/2016GL071866.
Dutton, E. G., J. J. Michalsky, T. Stoffel, B. W. Forgan, J. Hickey, D. W. Nelson, T. L. Alberta, and I. Reda.
2001. Measurement of broadband diffuse solar irradiance using current commercial instrumenta-
tion with a correction for thermal offset errors. Journal of Atmospheric and Oceanic Technology 18:
297–314.
Eddy, J. A. 1976. The maunder minimum. Science 192: 1189–1202. doi:10.1126/science.192.4245.1189
Finsterle, W. 2016. WMO international pyrheliometer comparison, IPC-XII, September 28–October 16, 2015,
Davos, Switzerland, Final Report. World Meteorological Organization, IOM Report No. 124. Geneva,
Switzerland: WMO.
Foote, P. D. 1918. Some characteristics of the Marvin pyrheliometer. Scientific Papers of the Bureau of
Standards, No. 323. Department of Commerce. Issued June 28, 1918. Washington, DC: Government
Printing Office.
Fröhlich, C. 1978. World Radiometric Reference, World Meteorological Organization, Commission for
Instruments and Methods of Observation, Final Report, WMO 490, 108-112.
Fröhlich, C. 1991. History of solar radiometry and the world radiometric reference. Metrologia 28: 111–115.
Fröhlich, C. 2012. Total solar irradiance observations. Surveys in Geophysics, 33:453–473. doi:10.1007/
s10712-011-9168-5.
Gröbner J., and A. Los. 2007. Laboratory calibration of pyrgeometers with known spectral responsivities.
Applied Optics 46: 7419–7425.
Gundlach, J. H., E. G. Adelberger, B. R. Heckel, and H. E. Swanson. 1996. New technique for measuring
Newton’s constant G. Physical Review D 54: R1256–R1259.
Hand, I. F. 1941. A summary of total solar and sky radiation measurements in the United States. Monthly
Weather Review 69: 95–125.
Harrison, L., and J. Michalsky. 1994. Objective algorithms for the retrieval of optical depths from ground-
based measurements. Applied Optics 33: 5126–5132.
Harrison, L., J. Michalsky, and J. Berndt. 1994. Automated multifilter rotating shadow-band radiometer: An
instrument for optical depth and radiation measurements. Applied Optics 33: 5118–5125.
Haupt, S. E., B. Kosovic, T. Jensen, J. K. Lazo, J. A. Lee, P. A. Jiménez, J. Cowie, et al. 2018. Building the
Sun4Cast® system, improvements in solar power forecasting. Bulletin of the American Meteorological
Society 99: 121–135.
Iqbal, M. 1983. An Introduction to Solar Radiation. New York: Academic Press.
Ivanov, B., P. Svyaschennikov, J.-B. Orbaek, C.-P. Nillsen, N. Ivanov, V. Timachev, A. Semenov, and
S. Kaschin. 2008. Intercomparison and analysis of radiation data obtained by Russian and Norwegian
standard radiation sensors on example of Barentsburg and Ny-Ålesund research stations. Journal of the
CNR’s Network of Polar Research Polarnet Technical Report, Scientific and Technical Report Series.
Cambridge, UK: CNR.
Jensenius Jr., J. S., and G. F. Cotton. 1981. The  development and testing of automated solar energy fore-
casts based on the model output statistics (MOS) technique. In Proceedings of the First Workshop on
Terrestrial Solar Resource Forecasting and on the Use of Satellites for Terrestrial Solar Resource
Assessment, pp. 22–29. Boulder, CO: American Solar Energy Society.
Historic Milestones in Solar and Infrared Radiation Measurement 67

Jensenius Jr., J. S., and G. M. Carter. 1979. Specialized Agricultural Weather Guidance for South Carolina.
TDL Office Note 79–15. Silver Spring, MD: National Weather Service, National Oceanic and
Atmospheric Administration, U.S. Department of Commerce.
Kendall, J. M. 1968. The  JPL Standard Total-Radiation Absolute Radiometer. Technical Report 32–1263
(May 15, 1968). Pasadena, CA: Jet Propulsion Laboratory.
Kendall, J. M. 1969. Primary Absolute Cavity Radiometer. Technical Report 32–1396 (July  15, 1969).
Pasadena, CA: Jet Propulsion Laboratory.
Kerr, J. P, G. W. Thurtell, and C. B. Tanner. 1967. An integrating pyranometer for climatological observer sta-
tions and mesoscale networks. Journal of Applied Meteorology 6:688–694.
Kerr, J. P, G. W. Thurtell, and C. B. Tanner. 1968. Mesoscale sampling of global radiation analysis of data
from Wisconsin. Monthly Weather Review 96:237–241.
Kimball, H. H., and H. E. Hobbs. 1923. A  new form of thermoelectric recording pyrheliometer. Monthly
Weather Review 51:239–242.
King, D. L., and D. R. Myers. 1997. Silicon-photodiode pyranometers: Operational characteristics, historical
experiences, and new calibration procedures. Paper Presented at the 26th IEEE Photovoltaic Specialists
Conference, September 29–October 3. Anaheim, CA: Institute of Electrical and Electronics Engineers.
Kopp, G. 2016. Magnitudes and timescales of total solar irradiance variability. Journal of Space Weather and
Space Climate 6 (27), Article 30.
Kopp, G., and J. L. Lean. 2011. A new, lower value of total solar irradiance: Evidence and climate significance.
Geophysical Research Letters Frontier Article 38: L01706. doi:10.1029/2010GL045777.
Kopp, G., G. Lawrence, and G. Rottman. 2003. The total irradiance monitor design and on-orbit functional-
ity. Proceedings of the SPIE Conference: Telescopes and Instrumentation for Solar Astrophysics, San
Diego, CA, August 3–8, 2003, Vol. 5171, paper 5171-4.
Kopp, G., G. Lawrence, and G. Rottman. 2005. The total irradiance monitor (TIM): Science Results. Solar
Physics. 230:129–139.
Latimer, J. R. 1973. On the Ångström and Smithsonian absolute pyrheliometric scales and the International
Pyrheliometric Scale 1956. Tellus 25:586–592.
Löf, G. O. G., J. A. Duffie, and C. O. Smith. 1966. World distribution of solar radiation. Solar Energy 10:27–37.
Marchgraber, R. M., and A. J. Drummond. 1960. A  Precision Radiometer for the Measurement of Total
Radiation in Selected Spectral Bands. Monograph No. 4, 10–12, Paris, France: International Union of
Geodesy and Geophysics.
Marty, C., R. Philipona, J. Delamere, E. G. Dutton, J. Michalsky, K. Stamnes, R. Storvold, T. Stoffel, S. A.
Clough, and E. J. Mlawer. 2003. Downward longwave irradiance uncertainty under arctic atmospheres:
Measurements and modeling. Journal of Geophysical Research 108. doi: 10.1029/2002JD002937
Michalsky, J. J., R. Perez, R. Stewart, B. A. LeBaron, and L. Harrison. 1988. Design and development of a
rotating shadowband radiometer solar radiation/daylight network. Solar Energy 41:577–581.
Michalsky, J., E. G. Dutton, D. Nelson, J. Wendell, S. Wilcox, A. Andreas, P. Gotseff, et al.. 2011. An extensive
comparison of commercial pyrheliometers under a wide range of routine observing conditions. Journal
of Atmospheric and Oceanic Technology 28:752–766.
Moll, W. J. H. 1923. A  thermopile for measuring radiation. Proceedings of the Physical Society London
Section B 35: 257–260.
National Aeronautics and Space Administration. 2018. Total and spectral solar irradiance sensor, TSIS-1.
https://www.nasa.gov/goddard/tsis-1 (Accessed March 8, 2018).
National Renewable Energy Laboratory. 1992. User’s manual—1961–1990 National Solar Radiation Data
Base. Version 1.0. NSRDB—Volume 1. NREL/TP-463-4859, National Renewable Energy Laboratory,
243 pp. Also available as TD-3282, Asheville, NC: National Climatic Data Center.
Reda, I. 1996. Calibration of a solar absolute cavity radiometer with traceability to the world radiometric refer-
ence. NREL Report No. TP-463-20619.
Rosenthal, A. L., and J. M. Roberg. 1994. Twelve month performance evaluation for the rotating shadowband
radiometer. Contractor Report, SAND94-1248. Albuquerque, New Mexico: Sandia National Laboratories.
SOLMET. 1978. User’s Manual—Hourly Solar Radiation—Surface Meteorological Observations, Vol.  1,
TD-9724. Asheville, NC: National Climatic Data Center.
SOLMET. 1979. Final Report—Hourly Solar Radiation—Surface Meteorological Observations, Vol.  2,
TD-9724. Asheville, NC: National Climatic Data Center.
Stoffel, T., C. Riordan, and J. Bigger. 1991. Joint EPRI/SERI Project to evaluate solar radiation measurement
systems for electric utility solar radiation resource assessment. The Conference Record of the Twenty-
Second IEEE Photovoltaic Specialists Conference, Vol. 2992, 533. New York: Institute of Electrical and
Electronics Engineers.
68 Solar and Infrared Radiation Measurements

Thomas, T., D. Johnson, P. Morgan, D. Heinicke, D. McDermitt, G. Burba, R. Peterson, J. Wurm, and
B.  Biggs. 2015. A  new silicon photovoltaic pyranometer for measuring solar irradiance in meteoro-
logical and solar resource applications. 2015  PV Solar Resource Workshop, February 24–27, 2015.
Golden, CO: National Renewable Energy Laboratory. https://www.nrel.gov/pv/pvmrw-20150227-fri.
html (Accessed May 28, 2019).
Turner, D. D., and R. G. Ellingson, eds. 2017. The Atmospheric Radiation Measurement Program, The First
20 Years. AMS Meteorological Monograph Vol. 57. Boston, MA: American Meteorological Society.
Vignola, F. 1999. Solar cell based pyranometers: Evaluation of the diffuse response. In Proceedings of the
28th ASES Annual Conference, 255–260. Portland, OR: ASES.
Vignola, F. 2006. Removing systematic errors from rotating shadowband pyranometer data. Proceedings of
the 35th ASES Annual Conference. Denver, CO: ASES.
Waldmeier, M. 1961. The Sunspot-Activity in the Years 1612–1960. Zurich, Switzerland: Schulthess.
Walter, B., R. Winkler, F. Graber, W. Finsterle, N. Fox, V. Li, and W. Schmutz. 2017. Direct solar irradiance
measurements with a cryogenic solar absolute radiometer. AIP Conference Proceedings 1810, 080007.
American Institute of Physics. doi: 10.1063/1.4975538.
Wesely, M. L. 1982. Simplified techniques to study components of solar radiation under haze and clouds.
Journal of Applied Meteorology 21:373–383.
Wilbert, S., N. Geuder, M. Schwandt, B. Kraas, W. Jessen, R. Meyer, and B. Nouri. 2015. Best practices
for solar irradiance measurements with rotating shadowband irradiometers. IEA SHC Task 46 report.
http://task46.iea-shc.org/data/sites/1/publications/INSRSI_IEA-Task46B1_BestPractices-RSI_150819.
pdf (Accessed May 28, 2019).
Wilbert, S., S. Kleindiek, B. Nouri, N. Geuder, A. Habte, M. Schwandt, and F. Vignola. 2016. Uncertainty of
rotating Shadowband irradiometers and Si-pyranometers including the spectral irradiance error. AIP
Conference Proceedings 1734, 150009. doi:10.1063/1.4949241.
Wilcox, S., and D. R. Myers. 2008. Evaluation of radiometers in full-time use at the national renewable energy
laboratory. Solar Radiation Research Laboratory NREL/TP-550-44627. https://www.osti.gov/scitech/
servlets/purl/946331 (Accessed May 28, 2019).
Willson, R. C. 1973. Active cavity radiometer. Applied Optics 12:810–817.
Willson, R. C. 1993. Solar irradiance. In  Atlas of Satellite Observations Related to Global Change, eds.
R. J. Gurney, J. L. Foster, and C. L. Parkinson. New York: Cambridge University Press.
World Data Center A. 1958. IGY General Report. Washington, DC: World Data Center A.
World Meteorological Organization. 2016. International Pyrheliometer Comparison, IPC-XII, 28 September–16
October 2015, IOM Report No. 124. Davos, Switzerland: WMO.
World Meteorological Organization. 1965. Measurement of radiation and sunshine. Guide to Meteorological
Instrument and Observing Practices, 2nd ed., WMO No. 8, TP3.
World Meteorological Organization. 2006. CIMO-XIV, Activity Report, December  7–14, 2006, WMO
No. 1019 (Part II).
World Meteorological Organization. 2008. Measurement of radiation and sunshine. Guide to Meteorological
Instruments and Observing Practices, 7th ed., WMO-No. 8. http://www.wmo.int/pages/prog/www/
IMOP/IMOP-home.html (Accessed May 28, 2019).
Zerlaut, G. 1989. Solar radiation instrumentation. In R.L. Hulstrom, ed., Solar Resources. Cambridge, MA:
MIT Press.
4 Direct Normal Irradiance

We receive from the sun perfectly enormous quantities of radiation.


Charles Greeley Abbot (1872–1973)

4.1 OVERVIEW OF DIRECT NORMAL IRRADIANCE


Solar radiation that arrives at the Earth’s surface having come directly from the Sun’s disk and strik-
ing a planar surface facing the Sun is defined as direct normal irradiance (DNI). Even if the sky
is clear, DNI is smaller than would be measured at the top of the atmosphere (TOA) of the Earth
because DNI has undergone scattering (by molecules and aerosols) and absorption (by gases and
aerosols) within the Earth’s atmosphere. If clouds are between the Sun and the observer, and they
are optically thick, then no DNI reaches the Earth’s surface. The global horizontal irradiance (GHI)
observed at the surface is a mix of DNI that reaches the Earth’s surface without being scattered or
absorbed and diffuse horizontal irradiance (DHI), the irradiance resulting from molecules, aero-
sols, and clouds scattering the DNI. This partitioning of solar irradiance is ever changing because
the atmosphere is not static. The governing equation is

GHI = DNI ⋅ cos ( sza ) + DHI (4.1)

where sza is the solar zenith angle, the angle between the Sun’s position and the local vertical.
This  partitioning during the day is illustrated in Figure  4.1. It  is easy to see that the solid dark
line GHI is the sum of the direct normal component on the horizontal in gray (the first term in
Equation 4.1) plus the dotted line DHI (the second term of the equation). This partitioning is con-
sistent throughout the day unless the Sun is completely blocked and the first term on the right-hand
side in Equation 4.1 goes to zero, leaving GHI equal to DHI between 16:00 and 17:00 and after
17:30 where the solid dark line and the dotted line coincide.
The Sun’s output at the TOA of the Earth, called the total solar irradiance (TSI), has been mea-
sured by a variety of space-based radiometers over several decades (Fröhlich 1991; Kopp and Lean
2011). The current TSI measurements, which began in 2003, are being made by the total irradi-
ance monitor (TIM) on the SOlar Radiation and Climate Experiment (SORCE) satellite (http://
lasp.colorado.edu/sorce). In December 2017, the National Aeronautics and Space Administration
(NASA) launched the Total and Spectral Solar Irradiance Sensor (TSIS-1) that is expected to be
deployed on the International Space Station for, at least, the next 5 years. This arrangement will
provide overlapping measurements of TSI with SORCE for at least 2 years. TIM measurements of
the TSI indicate stable solar output with the largest variation of around 0.3 percent, lasting just a few
days, but with typical variations of TSI under 0.1 percent (Kopp and Lean 2011). The Sun’s output
is so stable that it goes by the slight misnomer “solar constant.” Even with this almost constant solar
output, the irradiance value at the TOA of the Earth changes for the simple reason that the distance
between the Sun and the Earth changes as explained in Chapter 2. The Earth’s orbit around the Sun
causes a change in the amount of solar radiation received at the TOA of the Earth’s atmosphere by
a little over 6.7 percent between early January when TSI has its highest value (minimum Earth–Sun
distance) and early July when TSI has its lowest value (maximum Earth–Sun distance). This change
is entirely predictable from orbital mechanics and must be considered when computing the extrater-
restrial radiation (ETR).

69
70 Solar and Infrared Radiation Measurements

FIGURE 4.1 A demonstration of the partitioning of the GHI into the horizontal component of DNI and DHI
using actual data from the NREL SRRL. DNI can be higher than GHI when it is clear (dashed line). When it
is cloudy, GHI is equal to DHI (e.g., between 16:00 and 17:00).

The amount of direct solar radiation that reaches the Earth’s surface is more uncertain and dif-
ficult to forecast than the other components. DNI (“beam”) is the only form of sunlight that can be
used by concentrating solar energy conversion technologies such as concentrating solar power (CSP)
thermal systems, which produce steam to generate electricity, and concentrating photovoltaic (CPV)
systems, which generate electricity directly. Flat plate devices used for heating water and photovol-
taic modules that produce electricity, often found on the roofs of houses or businesses, can use any
incident solar radiation—direct sunlight, or sunlight that has been scattered by clouds, molecules,
and aerosols in the atmosphere, or sunlight reflected by the surface. DNI plays an important role for
solar energy conversion because it offers the highest energy density of all these irradiance compo-
nents. As shown in Figure 4.2, the area under the DNI time-series curve can represent the largest
amount of solar energy available for a day, depending on the daily minimum solar zenith angle. Flat
plate solar collectors can be mounted with a fixed, adjustable, or tracking mode for optimum energy
conversion. Under clear and partly cloudy sky conditions, DNI often is the predominant contributor
to the plane of array (POA) solar irradiance.
Measuring DNI is the most expensive measurement in the field of broadband solar and infrared
radiation. Although a good pyrheliometer for measuring DNI may be less expensive than a good
pyranometer used to measure GHI, the pyrheliometer must be accurately pointed at the Sun from
sunrise to sunset. Good automatic solar trackers can be 5–10  times the price of a thermopile-
type pyrheliometer. Less expensive, semi-automatic clock-drive trackers have been used histori-
cally, but they require regular adjustments to keep the pyrheliometer pointed at the sun. Because
of the expense of a fully automated tracker or the labor requirement to routinely align a clock-
drive tracker, there are relatively few continuous, long-term measurements of DNI. Globally, there
are orders of magnitude more pyranometer installations measuring GHI than those with pyrhe-
liometers measuring DNI (http://www.nrel.gov/rredc/pdfs/solar_inventory.pdf; Sengupta et  al.
2015). However, long-term DNI data are crucial for selecting a CSP or CPV site, for estimating
the expected annual output from a functioning plant, and for developing an optimum operating
strategy for the solar plant. Given the millions of dollars in capital costs to establish a large solar
power plant, it is surprising that a reliable pyrheliometer and tracker are not used more often in
prospecting for appropriate sites.
Direct Normal Irradiance 71

FIGURE 4.2 Demonstration of the daily integrated DNI far exceeding the daily integrated GHI on a very
clear day.

After operations begin, DNI data are useful to estimate the plant’s efficiency and to spot prob-
lems with plant operations. Although surface measurements are the most accurate, it is reasonable
to use DNI estimated from satellite images to make initial surveys of the direct solar resource
(e.g., Vignola et al. 2005). Satellites can prospect over a wide area, and analysis of these satellite-
based surveys can help narrow the search for the optimum CSP or CPV location in an area. When
anchored by accurate ground-based data from the same general area, systematic errors in the satel-
lite estimates can be accurately assessed and a more precise determination of the solar resource
can be made. Another advantage of satellite estimates is that geostationary satellites have been
deployed for decades, so even a year’s worth of ground-based data can be used to validate satellite-
estimated irradiance for the area of interest. Satellite-derived estimates then can be used to evaluate
the inter-annual variability and long-term trends of solar radiation at the site. Further, satellite-
based estimates in combination with ground-based measurements are increasingly used in short-
term forecasting of the solar resource. DNI data from both sources will be an indispensable aid in
future solar energy power plant operation, allowing plant operators to optimize the integration of
renewables into a utility’s electrical grid. As renewable energy systems increase their contribution
to the electrical energy mix in the United States, the transmission of solar-generated electricity from
geographically dispersed solar power generators needs to be managed with accurate solar forecast
information.
An alternative to pointing a pyrheliometer at the Sun is to calculate the DNI using measurements
made with a rotating shadowband radiometer (RSR). Chapter  10 describes the various types of
RSRs in detail, but we introduce this novel instrument here as another method to obtain estimates
of the DNI, although with some loss of accuracy compared with a pyrheliometer. Figure 3.21 in
Chapter 3 from Wesely (1982) shows a simple implementation of the RSR and is the first description
of such an instrument known to the authors. The pyranometer was a Lambda (now LI-COR, Inc.
model LI-200) silicon photodiode instrument with a very fast (~1 microsecond (µsec)) time response.
The shade band axis is pointed south in a Northern Hemisphere deployment, and it is rotated with
constant speed. When the band is beneath the pyranometer, a measurement can be made of the GHI,
and when the pyranometer diffuser is completely shaded from DNI by the band, a measurement can
be made of the DHI. During stable atmospheric conditions, the difference between the shaded and
72 Solar and Infrared Radiation Measurements

un-shaded pyranometer measurements is the horizontal component of the DNI. The DNI can then
be obtained by dividing the horizontal DNI by cos(sza) at the time of the measurement:

DNI = ( GHI − DHI ) / cos ( sza ) (4.2)

As presented in Chapter 5, pyranometers by design have more sources of measurement uncertainty


than do pyrheliometers. Based on many years of characterizing the measurement performance of
pyranometers, especially those with photodiode detectors, it is possible to apply various corrections
to RSR measurements to reduce the uncertainty of the computed DNI. As described in Chapter 10,
it is important to understand how RSR data have been corrected for various pyranometer response
characteristics (e.g., angular, spectral, temperature, linearity, etc.). Although the RSR instrument
type eliminates the need for an automatic solar tracker to acquire DNI data, estimating the measure-
ment uncertainty of RSR data requires specific knowledge of the corrections applied by the system
or as part of the post-measurement processing.

4.2 PYRHELIOMETER GEOMETRY


Figure 4.3 is a photograph of an Eppley model Normal Incidence Pyrheliometer (NIP) mounted on
an Eppley model SMT-3 automatic tracker. Figure 4.4 is a schematic of the optical geometry within a
pyrheliometer with a circular field of view like the NIP. Sunlight enters from the right and is incident
on the receiver behind the aperture stop on the left. The Gershun tube (not shown) that surrounds
and holds this assembly has its interior blackened. A window in front of the field stop (not shown) is
typically selected from a material that transmits direct sunlight uniformly over all solar wavelengths
to the interior of the instrument. The window serves to seal the instrument from the elements. Sets of
blackened baffles (three in this example) are positioned to minimize scattered light within the instru-
ment. The detector sits directly behind and fills the aperture stop (or precision aperture) at the bottom
of the tube that, along with the window and field stop at the top of the tube, defines the acceptance
angles for the instrument. Figure 4.4 illustrates the angles that are defined by the field and aperture

FIGURE 4.3 The Eppley Laboratory, Inc. Model NIP pyrheliometer mounted on a Model SMT-3 automatic
solar tracker; the insert is a blowup of the target on the rear flange of the NIP used for alignment (see text for
explanation of the numbering).
Direct Normal Irradiance 73

FIGURE 4.4 Schematic and definitions of the angles used to define the field of view of pyrheliometers.

stops. The angle designated z0 is called the opening angle and is referred to when designating the field
of view for a pyrheliometer (one-half of the total field of view). The limit angle zl is denoted by this
term because beyond this angle no sunlight is detectable (assuming the Sun is a point light source).
However, since the Sun has a finite diameter that subtends about 0.52 degrees at the TOA of the
Earth, solar radiation is detectable up to 0.26 degrees beyond the limit angle relative to the center of
the solar disk. The slope angle zs is the most important angle with regards to tracking accuracy. If the
center of the Sun is at this angular distance from the center of the aperture, vignetting occurs (radia-
tion from the full disk of the Sun does not strike the detector surface). Since the Sun has a maximum
radius of about 0.26 degrees, the Sun center must be closer to the aperture center than an angle equal
to the slope angle minus 0.26 degrees to avoid vignetting. The data in Table 4.1 are from Gueymard
(1998) and list these angles for a few common pyrheliometers.
As an aside, LI-COR, Inc. (1982) published an instruction manual for the LI-2020 solar tracker
that is no longer sold. However, in the manual they give a useful explanation of the alignment for
the Eppley NIP, which is summarized here. The  Eppley NIP has a top flange with a hole and a
bottom flange with a target (see Figure  4.3 inset; the gray dots will be explained momentarily).
The instrument is aligned when the Sun’s image is on the central black dot of the target as seen
in the inset. A  white ring that is surrounded by a black ring surrounds the central black dot.
If the Sun’s image center is on the edge of the central black dot, its misalignment is 0.33 degrees
(position 1); if the Sun’s image center is on the edge of the white ring, its misalignment is 1.00
degrees (position 2); and if the Sun’s image center is on the edge of the outer black ring, its misalign-
ment is 2.00 degrees (position 3). Therefore, for the Eppley NIP with a slope angle of 1.78° given the
Sun’s radius of 0.26°, the Sun’s image must not be centered farther than halfway between the white

TABLE 4.1
Optical Geometry (Defined in Figure 4.4) for a Few Pyrheliometers
Slope Angle Opening Angle Limit Angle
Pyrheliometer zs (°) zo (°) zl (°)
Eppley NIP 1.78 2.91 4.03
Eppley AHF 0.804 2.50 4.19
Kipp & Zonen CHP 1 1.0 2.5 4.0

Source: Gueymard, C., J. Appl. Meteorol., 37, 414–435, 1998.


74 Solar and Infrared Radiation Measurements

and black ring boundaries, which is an angular misalignment of around 1.5 degrees (position 4).
With manually adjusted clock-drive trackers, alignments should be made during clear-sky condi-
tions at least every third day, especially around the equinoxes when the solar declination is changing
about 0.3 degrees per day. Typically, properly installed automatic trackers have no problem keeping
the Sun’s image well within this tolerance.

4.3 OPERATIONAL THERMOPILE PYRHELIOMETERS


Although there were a few photodiode pyrheliometers manufactured, at this time they are no longer
made. Their main problem is that their sensitivity changes with atmospheric composition because
they have a non-uniform spectral response. As Rayleigh scattering, aerosol scattering and absorp-
tion, and molecular absorption alter the incoming solar spectrum, the pyrheliometer detector’s
responsivity changes. There is a marked spectral shift to the longer wavelengths as the Sun passes
through larger air masses. More and more of the blue light is scattered from the beam irradiance
(DNI) as the sunlight travels longer distances through the atmosphere. This change in the spectral
distribution of the DNI affects the responsivity of the photodiode sensor, making them unsuitable
for accurate and consistent DNI measurements.
The  most accurate pyrheliometers are electrically self-calibrating absolute cavity radiometers
(ACRs). ACRs are not  typically used in the field because they are an order of magnitude more
expensive than thermopile pyranometers; the frequent self-calibrations for maintaining accurate
DNI values interrupt the data stream; and in their optimum operating mode they are used without a
window, which means that they must be attended to ensure that they are not damaged by exposure
to weather and insects. Occasionally, ACRs are used with special windows to allow accurate DNI
values to be obtained under a variety of conditions, but in almost all cases thermopile pyrheliom-
eters are used for operational DNI measurements.
The geometry that is used for the typical thermopile pyrheliometer is described in Figure 4.4.
The thermopile detector is positioned at the bottom of the Gershun tube behind and thermally in
contact with a black absorbing surface that fills the aperture stop. The hot junction is exposed to the
direct Sun, and the cold junction is attached to a heat sink in the pyrheliometer. The temperature
difference creates a voltage that is proportional to the received direct solar irradiance. Voltages gen-
erated by the DNI on a clear day with little haze peak near 8 millivolts (mV) for typical commercial
pyrheliometers and irradiances near 1,000 watts per square meter (Wm−2).
Table 4.2 contains some operational characteristics of the type of pyrheliometers most often
used for long-term DNI measurements. This  information is taken from WMO (2017) and ISO
(2018). There are no World Meteorologic Organization (WMO) entries for ISO (2018) Class C
pyrheliometers. ISO (2018) added a Class AA  for reference instruments used to calibrate field
instruments. These Class AA pyrheliometers are very expensive and not designed for routine field
measurements. The very restrictive specifications for AA pyrheliometers are given in ISO (2018).
To evaluate most operational pyrheliometers, an extensive comparison was conducted over
10  months in 2008–2009, including all winter months and the hottest summer months at a mid-
latitude, mid-continent, high-elevation site (Michalsky et  al. 2011). All manufacturers of thermo-
pile pyrheliometers at the time were invited to send instruments and all with pyrheliometers on the
market at the time were represented. Since the fall 2008, newer models are available from some of
these same manufacturers (e.g., EKO Instruments Trading Co., Ltd., The Eppley Laboratory, Inc.,
Hukseflux Thermal Sensors B.V., and Kipp & Zonen Delft B.V. are known to have new instruments,
see Appendix F), but no evaluation of these pyrheliometers on the scale of this 2008/2009 study
has been scheduled. This comparison included more than the traditional calibration conditions that
require clear, stable skies with direct irradiances greater than 700 Wm−2. The conditions sampled
included calm, windy, cold, hot, and thin to partly cloudy conditions. All instruments were calibrated
over the 10-month study on 10 different occasions using un-windowed ACRs. Calibrations are dis-
cussed in Section 4.5 after the description of the cavity radiometer. Outside the 10 calibration periods,
Direct Normal Irradiance 75

TABLE 4.2
Operational Characteristics of Class A, B, and C (ISO Designation)/High and Good
Quality (WMO Designation) Pyrheliometers
Characteristic ISO/WMO Class A/High Quality Class B/Good Quality Class C
Response time (95% response) <10 sec/<15 sec <15 sec/<30 sec <20 sec/NA
Zero offset [(a) response to 5 K/hour change in 1 Wm−2/2 Wm−2 3 Wm−2/4 Wm−2 6 Wm−2/NA
ambient temperature; (b) complete zero offset from 2 Wm−2/NA 4 Wm−2/NA 7 Wm−2/NA
(a) and other sources]
Resolution (smallest detectable change in Wm−2) NA/0.51 Wm−2 NA/1 Wm−2 NA/NA
Stability (change per year) 0.5%/0.1% 1.0%/0.5% 2%/NA
Temperature response (maximum error due to 0.5%/1% 1.0%/2% 5%/NA
ambient temperature change within 50 K range;
ISO specifies −10 to 40°C relative to 20°C)
Non-linearity (Deviation from the responsivity at 0.2%/0.2% 0.5%/0.5% 2%/NA
500 Wm−2 within 100–1,100 Wm−2 range)
Spectral sensitivity (Deviation of the product of the 0.2%/0.5% 1%/1% 2%/NA
spectral absorptance and spectral transmittance
within 300–3,000 nm range for WMO specs)
Tilt response (Deviation from responsivity at 0° tilt 0.2%/0.2% 0.5%/0.5% 2%/NA
(horizontal) due to change between 0 and 90° tilt at
1,000 Wm−2

3 cavity radiometers with windows that allowed for near-continuous operations were the standard to
which the thermopile pyrheliometers were compared. All instruments during the study were stable in
the sense that their calibrations were repeatable to within the uncertainty of the calibration process.
The  variable conditions pyrheliometer comparison (VCPC), which was the title and acronym
attached to this study, revealed three levels of performance for the instruments in the comparison.
This comparison ignores two outliers that performed so poorly that further discussion of them was
not considered. One of these was a photodiode pyrheliometer that was not expected to perform well,
and the other was a thermopile instrument that was available, but it is uncertain that it was used
correctly because there was no instruction manual for operating this instrument. The best perform-
ers, as expected, were the ACRs with windows that transmitted all solar wavelengths. The ACRs
with windows are called all-weather cavity radiometers. The estimated 95 percent confidence level
uncertainty in the measurements made by cavity radiometers without a window is 0.45 percent. A 95
percent confidence level means that 95 percent of the time, the measurements will be made within
the uncertainty level expressed. The 95 percent confidence level uncertainty will be referred to as
95 percent uncertainty throughout this book. Transferring the calibration to all-weather (windowed)
cavities produced a total 95 percent uncertainty of only 0.5 percent or just slightly more than the
uncertainty of the cavities without the window. Several thermopile pyrheliometers performed with
a 95 percent uncertainty between 0.7 and 0.8 percent; these were designated to be in the second tier
of performance. One manufacturer supplied four versions of the same pyrheliometer model that
had 95 percent uncertainties between 1.0 and 1.7 percent. Surprisingly, the 95 percent uncertainties
determined in this experiment are better than those claimed by their respective manufacturers.
These results assume that calibrations using ACRs without windows are performed on a regular
basis, temperature corrections are applied, and, most importantly, instrument windows are cleaned
on a routine basis, at least twice weekly or more often in particularly dusty or dirty conditions.
The results for most instruments in the study are given in Table 4.3. Temperature corrections in the
fourth column refer to mathematical corrections suggested by the manufacturer that are in addition
to those achieved by circuit design.
76 Solar and Infrared Radiation Measurements

TABLE 4.3
Instrument Identification and Uncertainties for Clear-Sky Conditions from the Various
Conditions Pyrheliometer Comparison (VCPC)
Temperature
Manufacturer Model Serial No. Correction (Yes/No) 95%Uncertainty
Eppley All-weather cavity 31114 N ±0.5%
Eppley All-weather cavity 29219 N ±0.5%
Eppley All-weather cavity 32452 N ±0.5%
Kipp & Zonen CH 1 970147 N +0.70%
–0.60%
Kipp & Zonen CHP 1 80011 Y +0.70%
–0.60%
Kipp & Zonen CHP 1 80010 Y +0.70%
–0.60%
Kipp & Zonen CHP 1 80009 Y +0.70%
–0.60%
Kipp & Zonen CH 1 30346 Y ±0.7%
Kipp & Zonen CH 1 930039 N ±0.7%
Kipp & Zonen CH 1 30340 N ±0.7%
Kipp & Zonen CH 1 calcium fluoride window 30347 Y +0.70%
–0.80%
Middleton DN5 5094 N +0.70%
–0.80%
Middleton DN5 5027 N +0.70%
–0.80%
Middleton DN5 5029 N +0.70%
–0.80%
Hukseflux DR01 8029 Y +0.70%
–0.80%
Hukseflux DR01 8028 Y +0.70%
–0.80%
Hukseflux DR01 8027 Y +0.70%
–0.80%
Kipp & Zonen CHP 1 OPROTO1 N ±0.9%
Eppley NIP ventilated 28322 Y +1.00%
–0.70%
Eppley NIP new brass 34504 Y +1.10%
–0.90%
Eppley NIP new brass 34507 Y +1.10%
–0.90%
Eppley NIP new brass 34129 Y +1.10%
–0.90%
Huskeflux DR01P 8026 N +1.10%
–1.30%
Eppley NIP old brass 16229 Y ±1.2%
Eppley NIP old brass 16319 Y ±1.2%
Eppley NIP old brass 16521 Y ±1.2%
(Continued)
Direct Normal Irradiance 77

TABLE 4.3 (Continued)


Instrument Identification and Uncertainties for Clear-Sky Conditions from the Various
Conditions Pyrheliometer Comparison (VCPC)
Temperature
Manufacturer Model Serial No. Correction (Yes/No) 95%Uncertainty
Eppley NIP stainless steel 31139 Y +1.40%
–1.20%
Eppley NIP stainless steel 25791 Y +1.40%
–1.20%
Eppley NIP stainless steel 31144 Y +1.40%
–1.20%
Eppley NIP stainless steel 28260 Y +1.70%
–1.30%
Matrix MK-III (photodiode) 2457 N –
Cimel 183A7 501657 N –

Source: Michalsky, J. et al., J. Atmos. Oceanic Technol., 28, 752–766, 2011.

4.4 ABSOLUTE CAVITY RADIOMETERS


Before a discussion of pyrheliometer calibration, it is necessary to review the design and perfor-
mance characteristics of the ACR used to define the internationally recognized World Radiometric
Reference (WRR) (Fröhlich 1991). The concept of an ACR is based on the ability to measure elec-
trical power and equate the measurement to solar flux power density. That is, an ACR is considered
an electrically self-calibrating instrument. The electrical substitution can take place in an active
or passive mode of operation. In the active mode, the electrical substitution is performed as part
of the irradiance measurement where the goal is to maintain a balanced, or null, output signal by
adjusting the electrical power to match the solar flux. In the passive mode, irradiance measurements
are taken between electrical calibration events when the electrical power is equated to the solar flux
at the time. A very simple schematic of one type of ACR operating in the passive mode is shown in
Figure 4.5. Only the passive mode of operating an ACR will be described here.
Direct sunlight enters from the left in Figure 4.5 passing through a view-limiting aperture, or field
stop, and encounters the precision aperture with a very accurately measured opening area. Sunlight

FIGURE 4.5 Schematic of an ACR.


78 Solar and Infrared Radiation Measurements

passing through the precision aperture impinges on the blackened cone where it is almost com-
pletely absorbed, thus heating the cone detector to which is attached the hot junction of a thermopile.
The cold junction of the thermopile is attached to the heat sink. The temperature differential pro-
duces a voltage in the thermopile that is precisely measured. The field stop then is covered to prevent
any sunlight from reaching the cone. The electric heater is activated to raise the temperature of the
cone to the same temperature and produce the same voltage in the thermopile as it had with the sun
striking it. The power to heat the cone electrically is divided by the area of the field stop to yield
power per unit area in milliwatts per square centimeter (mW cm−2) or, equivalently, Wm−2. In the pas-
sive mode of operation, this calibration is made before and after DNI measurements over a selected
interval (hourly or shorter) to account for the changing thermal environment that can occur during
the measurement period.
While the process seems simple, in fact, there are several small correction factors that must be
addressed to remove biases from the ACR measurements. The major factors (Reda 1996) are the
losses caused by reflection from the cone detector, which is nearly but not perfectly absorbing, and
the losses resulting from the heated cone cooling by emitting infrared radiation; the nonequivalence
of heating the cone with electrical power versus solar power; the scattered light from the precision
aperture; light diffraction into the cavity; the electrical losses in the wire leads that heat the cavity;
and, most importantly, the accuracy of the measurement of the area of the aperture.
The Physikalisch-Meteorologisches Observatorium Davos and World Radiation Center (PMOD/
WRC) developed and maintains the WRR. The WWR is realized by the World Standard Group
(WSG), which consists of several different manufacturers’ cavity radiometers donated by research
laboratories and manufacturers. In the twelfth IPC (IPC-XII), for example, four cavity radiometers
constituted the WSG (Finsterle 2016). Their average is defined to be the WRR. The  95 percent
uncertainty level for the WRR is 0.3 percent. Every 5 years an IPC is held in Davos, Switzerland;
the last was in 2015. The comparisons are conducted in early autumn at the PMOD. In 2015, at
IPC-XII, 111 participants calibrated 134  pyrheliometers, mostly cavity radiometers. The  partici-
pants then used their newly calibrated or recalibrated instruments to propagate the WRR factor to
those interested in tracing their measurements to the WRR.
In the United States, the National Renewable Energy Laboratory (NREL) of the U.S. Department
of Energy holds an NREL pyrheliometer comparison (NPC) every year except for IPC years.
The comparisons are in early autumn, as are the IPCs, when the sky conditions are most likely to
be cloudless. The standards used for the NPC include several cavity radiometers that are directly
traceable to the WRR. These comparisons are smaller and are mostly U.S. participants, but they are
not limited to the United States, and there are always a few international participants (Reda et al.
2017). NPC information is available at http://www.nrel.gov/aim/npc.html.

4.5 UNCERTAINTY ANALYSIS FOR PYRHELIOMETER CALIBRATION


The Joint Committee for Guides in Metrology (JCGM 2008) published the Guide to the Expression
of Uncertainty in Measurements (GUM) with minor changes from the 1995 version. This guide uni-
fies the approach to expressing uncertainty for the measurement sciences internationally. Two prac-
tical guides (Cook 2002; Taylor and Kuyatt 1994) are used in the discussion that follows to make the
GUM procedures of uncertainty analysis in radiometry somewhat more accessible. The application
of GUM to the calibration of a pyranometer explained in Reda et al. (2008) is a clear, published
example of the practical application of the GUM approach in radiometry.
A clear statement of the reliability of a measurement in terms of its uncertainty should accom-
pany every measurement. An important change in GUM from earlier descriptions of uncertainty
analyses was the separation of uncertainty associated with repeated measurements of the same
quantity that allows for a statistical treatment of the random behavior of the measurement (type A)
and the uncertainty from all other causes that cannot be treated statistically (type B). For example,
Direct Normal Irradiance 79

assume that a measurement of the DNI is made on a clear day near solar noon using a pyrheliometer.
The pyrheliometer outputs a voltage proportional to the solar radiation incident upon its detector.
Assume that the data logger recording the voltage output can sample every second and average
60 samples to get a 1-minute mean value and a standard deviation. This standard deviation is an
example of a type A uncertainty. The data logger has a stated uncertainty of how well the voltage
can be measured in each voltage range. That statement of the uncertainty for the voltage scale is
provided in the data logger manufacturer’s specifications and is, therefore, a type B uncertainty
since it was not obtained from a series of measurements.
A summary of the steps to follow in performing an uncertainty analysis using GUM include the
following:

1. Write down the measurement equation.


2. Calculate the standard uncertainty for each variable in the measurement equation either as
a type A or a type B uncertainty.
3. Calculate sensitivity factors associated with each variable.
4. Use the root sum of the squares of the standard uncertainties modified by sensitivity fac-
tors to determine the combined standard uncertainty.
5. Calculate the expanded uncertainty by multiplying the combined uncertainty by a cover-
age factor found from the student’s t-distribution to obtain the chosen standard level of
confidence, often 95 percent representing a coverage factor (k) approximately equal to 2.

The measurement equation that applies to the calibration of a pyrheliometer is

V
R= (4.3)
DNI

where R is the calibrated response to be determined in microvolts per watts per square meter
(µV/Wm−2), V is the voltage output of the thermopile in µvolts, and DNI is the reference DNI in
watts per square meter. The most accurate DNI results come from an electrically self-calibrating
ACR that is traceable to the WRR (see Section 4.4). For example, assume a measured 1-minute aver-
age voltage in full direct sunlight of 7,993 µV. The digital logging system used to read the voltage
has a resolution of 1 µV and has an auto-zero feature that ensures that there is no offset. The stan-
dard deviation of the 60 1-second samples that result in the 1-minute average of 7,993 µV is 111 µV.
The cavity radiometer, which was calibrated at an IPC, has a 95 percent uncertainty of 0.41 percent,
including the 0.3 percent uncertainty of the WRR. The  average cavity reading for the measure-
ments is 932.3 Wm−2 with a calculated standard deviation of 13.7 Wm−2 over the same minute that
the pyrheliometer samples were collected. Table 4.4 lists the standard uncertainties associated with
calibrating the pyrheliometer; both type A and type B uncertainties are listed that arise from both
the voltage measurement and the irradiance measurement.

TABLE 4.4
Derivation of Standard Uncertainty for Sample Pyrheliometer Calibration
Semirange (B)
Variable Value or Std Dev (A) Type Distribution DF Std Uncertainty
V resolution 7993 µV 0.5 µV B Rectangular 1000 0.29 µV
V measurements 7993 µV 111 µV A Normal 59 14.33 µV
DNI cavity 932.3 Wm−2 0.205% = 1.91 Wm−2 B Rectangular 1000 1.10 Wm−2
DNI measurements 932.3 Wm−2 13.7 Wm−2 A Normal 59 1.77 Wm−2
80 Solar and Infrared Radiation Measurements

The  resolution of the data logger on the scale used to read the voltage is 1  µV; this suggests
that the voltage can lie anywhere between –0.5 and +0.5 µV of the reading, implying a standard
uncertainty of 0.5/ 3 for a rectangular distribution. The standard sample deviation reported by the
data logger is divided by 60 to get the standard uncertainty of the mean, where 60 is the number
of samples in the 1-minute average. In a similar manner, the DNI standard uncertainties were cal-
culated. The root sum of the squares of the standard uncertainties of the voltage is 14.33 µV to two
decimal places implying that, for this case, the resolution of the meter does not add significantly
to the uncertainty. The square root of the sum of the squares of the standard deviations of the DNI
is 2.08 Wm−2. The sensitivity factors for V and DNI are calculated by taking partial derivatives of
Equation 4.3 with respect to V and DNI. Adding in quadrature, the combined standard uncertainty
of the responsivity (R) is
2
 ∂R   ∂R  ( ∆V )2 (V ⋅ ∆DNI)2
( ) ( )
2 2
2
uRC =  ∆V +   ∆DNI = + (4.4)
 ∂V   ∂DNI  DNI 2 DNI 4

Taking the square root yields a standard uncertainty uRC of 0.025, which implies that 68 percent of
the determinations of R would fall within ±0.025. Multiplying by a coverage factor of two, results
in a responsivity and expanded uncertainty (U E) of 8.574 ± 0.050 µV/(Wm−2) with 95 percent con-
fidence. The 95 percent confidence level implies that if this exercise were repeated, 95 of 100 times
the calibration would fall within this range of ±0.050 µV/(Wm−2).

4.6 UNCERTAINTY ANALYSIS FOR OPERATIONAL THERMOPILE


PYRHELIOMETERS
The previous section covered an uncertainty analysis for the calibration of the pyrheliometer that
will be deployed in the field. When the pyrheliometer is operated unattended in the field, other
factors affect the measurement and add uncertainty to those measurements. Some of the contribu-
tors to this added uncertainty are temperature dependence, nonlinearity, pointing accuracy, solar
zenith angle dependence, and the data logger voltage measurement, which may be of lower quality
than the calibration data logging system. A huge uncertainty that can easily overwhelm everything
described is that caused by dirty optics (i.e., the window used to seal the pyrheliometer). Dirty
optics result when instruments are not routinely maintained, with the cleaning schedule dependent
on the site and weather.
In Table 4.5 the listed uncertainties for estimating overall operational uncertainty are handled as
percent uncertainties to demonstrate a different approach from that used in Table 4.4.

TABLE 4.5
Estimating Standard Uncertainty for an Operational Pyrheliometer
Semirange (B)
Variable Value or Std Dev (A) Type Distribution DF Std Uncertainty
Calibration R 0.292% B Rectangular 1000 0.168%
Table 4.4
Nonlinearity 0.1% B Rectangular 1000 0.058%
Temperature dependence 0.5% B Rectangular 1000 0.289%
V resolution for 100 Wm−2 0.33% B Rectangular 1000 0.191%
Window transmittance 0.60% B Rectangular 1000 0.364%
V measurements 7993 µV 111 µV A Normal 59 0.179%
Combined standard uncertainty 0.562%
Direct Normal Irradiance 81

The combined standard uncertainty is calculated by taking the square root of the sum of the
squares of the standard uncertainties from the right-hand column of Table 4.5. Small uncertainties
that would not contribute significantly to the overall uncertainty have been intentionally neglected,
and some contributions to uncertainty may have been inadvertently forgotten, but all contributions
of consequence are included. The total standard uncertainty is given in Table 4.5, implying that with
the coverage factor of two the uncertainty of the field measurement with 95 percent confidence in
the result is 1.124 percent of the DNI measurement. This result is with the assumption of a clean
pyrheliometer window; remember that even a modest dust deposit on the window can reduce the
measured DNI and easily double or triple this uncertainty systematically.
An updated analysis of DNI uncertainties for operational solar instrumentation appears in
Vuilleumier et al. (2014). A similar, but specific analysis of the DNI uncertainties encountered in
the operation of a well-run Baseline Surface Radiation Network (BSRN) station suggests somewhat
larger expanded uncertainties than derived here (1.1% vs. 1.5%) for the Payerne station, which may
be partially explained by their less accurate data logger.

4.6.1 WindoW tRansmittanCe, ReCeiveR absoRPtivity, and temPeRatuRe sensitivity


The  largest uncertainties in Table  4.5 are those due to the window transmittance and the tem-
perature sensitivity of the instrument. If the window transmission were constant throughout the
solar spectrum, this term could be reduced or even neglected. Some windows have nearly uniform
transmittance, such as calcium fluoride (CaF2). However, CaF2 tends to scratch and is hygroscopic
(attracts and absorbs water). Sapphire is another excellent choice, but it is probably too expensive
to use on an operational pyrheliometer. Fused silica is the most popular choice for operational pyr-
heliometers because it transmits the ultraviolet well; however, the transmittance is not constant to
3,000 nanometers (nm) but shows some dips in transmission to low values near 2,800 nm. This issue
gives rise to the size of the transmission uncertainty used in Table 4.5. Another requirement for
spectral insensitivity is an absorbing paint for the hot junction of the thermopile that is wavelength
independent. Fortunately, there are several paints that are satisfactory in that they are nearly wave-
length independent and durable.
Temperature sensitivity in thermopile pyrheliometers usually is corrected to first order using
a temperature compensation circuit. Manufacturers supply the temperature-dependent correction
for any first-order mathematical temperature correction if there is no temperature compensation
circuitry, and if there is electronic temperature compensation then a second-order mathematical
temperature compensation may be suggested. Figures  9 and 10 in Michalsky et  al. (2011) illus-
trate the effects of correcting and not  correcting the temperature dependence of pyrheliometers.
Figure 4.6 is a plot of the difference in irradiance measurements between a very good thermopile
pyrheliometer and a cavity radiometer after all known temperature corrections have been made.
The cavity radiometer has no temperature dependence; therefore, the figure indicates that, although
most of the temperature dependence is removed, there is room for improvement in the thermopile
pyrheliometer at the higher ambient temperatures. Algorithms can be developed from data using
ambient air temperature as a proxy of the heatsink temperature (cold junction of the thermopile) for
the correction. The bottom line is that every appropriate attempt to correct for this effect will reduce
the bias and uncertainty of the direct normal irradiance measurement.

4.6.2 solaR ZenitH angle dePendenCe


This uncertainty is not included in Table 4.5 but is of significant concern in some models of pyr-
heliometers. This effect is not understood (Habte et al. 2016; Vignola and Lin 2010, 2011) but is
clearly illustrated, for example, in Figure 4.7, which gives data from a broadband outdoor radiom-
eter calibration (NREL 2010) at the Southern Great Plains Atmospheric Radiation Measurement
82 Solar and Infrared Radiation Measurements

FIGURE  4.6 The  temperature dependence of one very good pyrheliometer after all manufacturer’s
corrections applied.

FIGURE  4.7 Solar zenith angle dependence of the Eppley NIP 29008E6  pyrheliometer during a
2010 BORCAL. The morning and afternoon responsivities differ as a function of solar zenith angle.

Facility of the U.S. Department of Energy. The figure is a plot of the calibrated responsivity over
a day with a responsivity change of over 2.5 percent from 80 to 20° solar zenith angle during the
morning and almost no change from 20 to 80° in the afternoon. Sometimes this change is smaller,
and sometimes it is more symmetric about solar noon. This behavior is not uncommon for some
pyrheliometers, and it exceeds the uncertainty of any contribution listed in Table 4.5.
Direct Normal Irradiance 83

4.7 UNCERTAINTY ANALYSIS FOR ROTATING SHADOWBAND


RADIOMETER ESTIMATES OF DIRECT NORMAL IRRADIANCE
RSRs derive DNI by subtracting measured DHI from measured GHI and dividing by the cosine
of the calculated solar zenith angle at the time of measurement (Equation 4.2). Since the uncer-
tainty associated with the GHI is larger than the uncertainty associated with the DNI measured
by a pyrheliometer, the uncertainty of the calculated DNI, which includes the uncertainties of the
GHI and DHI, will necessarily be even larger. Further, there is a small uncertainty associated with
the solar position calculation, which is negligible with the best algorithms, and there is an addi-
tional uncertainty associated with the horizontal leveling of the pyranometer. Since most RSRs use
a photodiode-based pyranometer, these pyranometers exhibit a significant additional uncertainty
because they have a non-uniform spectral response to wavelengths between 400 and 1,100 nm and
they are not  responsive to wavelengths between 300 and 400  nm and beyond 1,100  nm. Other
significant measurement uncertainty issues are the temperature dependence of the silicon cell’s
response function and the angular (cosine) response of the photodiode pyranometer (see Chapter 5
for more details on photodiode pyranometers). Table  4.6 lists some of the major contributors to
uncertainty for an RSR that uses a photodiode pyranometer for the calculation of the direct irradi-
ance. The expanded uncertainty for a 95 percent confidence in the direct irradiance calculation is
3.173 percent. Certain factors have very little effect on the overall uncertainty of an instrument.
For  example, if we eliminate from Table  4.6 the uncertainty associated with nonlinearity, then
the combined expanded uncertainty decreases by only 0.002 percent. Therefore, it is important to
account for large contributors to uncertainty and minor contributors to uncertainty can safely be
ignored. It should be certain that the contribution is small, however, before neglecting it.
The assumptions made for Table 4.6 were that the calibration was derived for the calculated DNI
from the RSR. Had the photodiode pyranometer been calibrated for GHI, the uncertainty of the
DNI would have been significantly greater. Most RSRs use the GHI calibration factor. For example,
Michalsky et al. (2009) demonstrated that the calibration for the GHI and DNI are different by about
2.5 percent for the photodiode pyranometer they used. The major cause for this discrepancy is that
the spectral distribution of DNI and DHI are significantly different under clear skies. There would
be similar differences for another manufacturer’s photodiode pyranometers. In Table 4.6, the tem-
perature dependence is not corrected; there is about a 3 percent change in responsivity related to
changes in typical annual operating temperatures. If this temperature is measured and the correc-
tion is made, then this contribution to the uncertainty drops significantly to around 0.3 percent.
Similarly, if the cosine response is measured and a correction applied to the DNI calculation,

TABLE 4.6
One Estimate of the Standard Uncertainty of a Rotating Shadowband Radiometer
Semi-range (B)
Variable Value or Std Dev (A) Type Distribution DF Std Uncertainty
Calibration R 0.292% B Rectangular 1000 0.168%
Table 4.4
Nonlinearity 0.1% B Rectangular 1000 0.058%
Temperature dependence 1.5% B Rectangular 1000 0.866%
V resolution for 100 W m−2 0.33% B Rectangular 1000 0.191%
Spectral response 1.0% B Rectangular 1000 0.577%
Cosine response 2% B Rectangular 1000 1.155%
V 7993 µV 111 µV A Normal 59 0.179%
Combined standard uncertainty 1.587%
84 Solar and Infrared Radiation Measurements

this  large contribution to the uncertainty can be reduced to a much smaller number, perhaps as
low as 0.4 percent. In Table 4.6 it is assumed that there has been a correction applied to reduce the
uncertainty of the spectral correction to 1.0 percent. However, this contribution to the uncertainty
will dominate rotating shadowband radiometer measurement uncertainty if not corrected properly.
For more discussion of RSRs and photodiode pyranometers see Chapters 5 and 10.

4.8 DIRECT NORMAL IRRADIANCE MODELS


The  term model has different meanings. Two types of models are briefly discussed here. First,
we discuss models that are radiative transfer codes that require inputs describing the atmospheric
extinction sources in the path of the direct solar beam. The  second type of model uses satellite
images or atmospheric soundings as retrieved from satellite observations to estimate the direct solar
radiation that strikes the Earth’s surface.

4.8.1 gRound-based modeling


Clear-sky direct irradiance radiative transfer models perform very well if model inputs are accu-
rately specified. Michalsky et al. (2006) evaluated six direct irradiance models of varying complex-
ity. For the 30 cases studied, it was demonstrated that the estimated clear-sky DNI values differed
from measurements by less than 1 percent. The mean difference for the six models was 0.5 percent.
This mean difference implies that radiative transfer models are well developed, and, given the cor-
rect input parameters, clear-sky DNI values can be estimated to an accuracy of better than field
pyrheliometer measurements. So, why are models not  preferred over measurements? The  reason
is that good models need several accurately specified input parameters to yield correct broadband
irradiances. For example, aerosol optical depth (AOD) and some indication of its wavelength depen-
dence are needed to model clear-sky DNI. The wavelength dependence can be a simple Ångström
expression such as τ = βλ −α , which will be discussed in Chapter 7. The next important contributor
to direct beam extinction is precipitable water vapor, so some independent estimate of this quantity
is required, and an ozone column measurement or estimate, the third most important model input,
is needed for the model. Satellite-based estimates of global daily column ozone are available from
https://www.esrl.noaa.gov/gmd/grad/neubrew/SatO3DataTimeSeries.jsp. Precipitable water vapor
can be crudely estimated from surface relative humidity and pressure or, better yet, using mea-
surements from nearby radiosondes, if available. AODs are measured with sun radiometers, which
are not  widely available. Determining atmospheric AOD generally presents measurement chal-
lenges similar to monitoring the DNI. AOD data are, however, available from a growing network of
ground-based stations (https://aeronet.gsfc.nasa.gov/) and will be discussed in Chapter 7. The more
complex models of the spectral distribution of direct normal irradiance, which can be integrated to
calculate broadband solar irradiance, are discussed in Chapter 7.

4.8.2 satellite model estimates


Satellite estimates of surface solar irradiance use models that range in complexity from those based
on regressions of selected ground-based irradiance measurements versus satellite pixel brightness
(i.e., completely empirical) to those based on radiative transfer models using satellite-retrieved or
climatological information on cloud properties, aerosols, humidity, and temperature profiles.
Perez et al. (2002) published their paper on a well-known operational model for deriving sur-
face irradiance from satellite measurements that uses a hybrid approach. The technique uses the
pixel brightness as measured by a Geostationary Operational Environmental Satellite (GOES) but
considers the site’s elevation, a climatological estimate of the monthly average aerosol loading,
Direct Normal Irradiance 85

snow cover, and a specular reflectance correction factor. Clean Power Research uses this model
and its modifications for its SolarAnywhere service, which provides estimates of surface irradi-
ance (http://www.cleanpower.com/SolarAnywhere), including DNI. 3TIER (http://www.3tier.com/
en/products/) processes satellite data to generate surface irradiance using its own modified version
of the Perez et al. (2002) model.
A popular site for solar DNI information based on satellite retrievals of the input parameters
is the NASA  Surface meteorology and Solar Energy (SSE) website (http://power.larc.nasa.gov).
This is also known as the Prediction Of Worldwide Energy Resource (POWER) project website.
NASA’s SSE approach is a more physics-based approach in that it uses a radiative transfer code with
cloud optical property inputs from satellite retrievals and a chemical transport model to estimate
aerosol extinction to calculate surface solar DNI.
The uncertainties associated with satellite retrievals of DNI are about twice as large as those for
the retrievals of GHI from satellites. Satellite estimates of hourly DNI using the Perez et al. (2002)
model have biases of only a few percent and root mean square errors (RMSEs) in the 35–40 percent
range for hourly data. In Vignola et al. (2007), the RMSE and the mean bias error (MBE) of DNI
are calculated from hourly averages of frequently sampled ground-based (gndi) measurements and
one satellite image (sati) per hour. The MBE in percent is calculated using


i =n
( gndi − sati ) 1
MBE = 100 ⋅ i =1
⋅ (4.5)

∑ gndi  / n
i =n
n

 i =1 

where the sum is ideally formed over all the daylight hours in at least 1 year to determine a typical
MBE for a site. For the RMSE Vignola et al. (2007) use

1

∑ ( gndi − sati ) 
i =n 2 2
 1
RMSE = 100 ⋅  i =1
 ⋅ (4.6)
( n − 1)
∑ gndi  / n
i =n
  
   i =1 

where again the sum is ideally formed over all the daylight hours in at least 1 year to determine a
typical RMSE for a site.
As an aside, it may seem that a more straightforward way to calculate MBE in percentage would
be to use a formula such as

 i =n  gndi − sati 
MBE = 100 ⋅  ∑

 i =1  gndi
 / n
 
(4.7)

where the individual percentage differences are averaged; however, if a single gndi value were equal
to zero, MBE becomes undefined. If zeros were screened from the sum, small nonzero denomina-
tors would still produce large fractions that would have a large and unphysical influence on the
result. The same arguments apply to calculating the RMSE using

1
 i =n  gnd − sat 2  2
RMSE = 100 ⋅  ∑

 i =1 
i
gnd i
i
  /n
 
(4.8)
86 Solar and Infrared Radiation Measurements

The hourly RMSEs of 35–40 percent may seem high because MBEs and RMSEs are often quoted
for mean monthly daily totals, which are much smaller numbers. For example, Myers et al. (2005)
quote uncertainties around 14 percent for the sites in their study using the model of Perez et al.
(2002). The average monthly daily total (AMDT) is the total solar energy per square meter per day
summed over all days in the month divided by the number of days in the month. Every month is
likely to have direct sunlight; therefore, this calculation of MBE

 i =m ( gndi − sati ) 
MBE = 100 ⋅  ∑
 i =0 gndi
/m

(4.9)

is likely to give a valid and defined MBE. In this equation gndi and sati are AMDTs for one of the
m months in the sample size. Note this has the same form as Equation 4.7, but instead of summing
over individual hours the sum is over monthly averages of daily totals. The same argument applies
to the RMSE for AMDTs. Alternately, it would be as informative to calculate and compare MBE
and RMSE using equations of the form in Equations 4.5 and 4.6. It is useful to state just how MBE
and RMSE are calculated when quoting biases and uncertainties, and it should be required for
publications.

4.9 HISTORICAL AND CURRENT SURFACE-MEASURED


DIRECT NORMAL IRRADIANCE DATA
An excellent review of the historical solar irradiance data and accompanying metadata regarding
these can be found in Stoffel et al. (2010) and need not be repeated here. Years of calibrated and
well-maintained instrumentation that captures as complete a record of DNI data as possible are
critical in understanding the annual and inter-annual variability of radiation. Climate research and
central concentrating solar power generation require long records to understand the long-term vari-
ability of DNI.
In this section, only measurements of DNI with expectations of continuance current as of January
2018 will be discussed. Much of the focus, but not all, is on the United States and on measurements
made with first-class instruments. RSRs will be discussed because of their proliferation.
The BSRN was started by the World Climate Research Program in 1992 to measure solar and
infrared surface radiation with the highest-quality instruments and techniques available (Ohmura
et al. 1998). The data produced have been used for validating and improving satellite retrievals of
surface radiative fluxes, for comparisons to climate model calculations, and for monitoring subtle
long-term changes in the surface radiative environment. The database is maintained at the Alfred
Wegener Institute in Bremerhaven, Germany. Data are available from about 50 stations worldwide
at http://www.bsrn.awi.de/.
The World Radiation Data Center is in St. Petersburg, Russia. It was established by the WMO to
archive and publish solar radiation data. Most of the data from the more than 1,000 sites included
in the archive are GHI measurements representing daily total irradiation values. Some sites submit
hourly data, and some sites submit DNI.
In the United States, three small networks continue to provide moderate- to high-quality DNI
data: The University of Oregon Solar Radiation Monitoring Laboratory (UO SRML), the National
Oceanic and Atmospheric Administration (NOAA) SURFace RADiation (SURFRAD) network,
and the NOAA SOLar RADiation (SOLRAD; formerly ISIS) network.
The  UO SRML has run the Pacific Northwest (United States) Solar Radiation Data Network
since 1977. The annual average DNI from the three longest running stations is shown in Figure 4.8.
Currently, 15 stations operate in Idaho, Montana, Oregon, Utah, Washington, and Wyoming that
Direct Normal Irradiance 87

FIGURE 4.8 Hourly average DNI for Burns, Eugene, and Hermiston, Oregon, for the years from 1978 through
2016; monthly average hourly values are used to determine the annual average. The monthly average values
exclude hours or days with missing data or misaligned measurements.

measure DNI; six stations use pyrheliometers and eleven stations use RSRs. Eugene and Hermiston,
Oregon, have co-located RSRs and pyrheliometers to evaluate the accuracy of the RSRs. Starting
around 2015, data are taken at 1-minute intervals at most sites, although 5-minute and 15-minute
data are taken at a few stations. More information can be found at http://solardata.uoregon.edu/.
The NOAA SURFRAD network has operated since 1994. All SURFRAD sites are part of the
BSRN. As such the sites acquire high-quality direct solar radiation data as a subset of more exten-
sive radiation measurements. The locations of the seven sites, their data record, and much more can
be found at http://www.srrb.noaa.gov/surfrad/index.html.
The NOAA SOLRAD network is a remnant of the NOAA National Weather Service (NWS)
solar network that operated at maximum strength in the 1950s through 1980s. There are currently
seven NWS stations that continue to operate solar equipment that includes direct irradiance sen-
sors. The stations and the data can be found at https://www.esrl.noaa.gov/gmd/grad/solrad/index.
html.
There  are other medium- to high-quality direct solar irradiance measurements made in the
United States, but typically their operation is short-term; the NREL web page lists several of these
(http://midcdmz.nrel.gov). NREL’s high-quality data from the Solar Radiation Research Laboratory
(SRRL) in Golden, Colorado, dates from 1981 (see Figure  4.9). The  U.S. Department of Energy
operates the Atmospheric Radiation Measurement (ARM) site in northern Oklahoma (Stokes and
Schwartz 1994). High-quality and redundant measurements of solar direct irradiance have been
made since 1992 at the ARM central facility between Lamont and Billings, Oklahoma, and at nearly
20  extended facilities through central Oklahoma and into Kansas. The  extended facilities have
recently been reduced to nine sites nearer the central facility (https://www.arm.gov/capabilities/
instruments/sirs).
88 Solar and Infrared Radiation Measurements

FIGURE  4.9 Maximum observed DNI for Golden, Colorado, based on 1-minute (1999  onwards) and
5-minute (before 1999) measurements from August 1981 through December 2017. The data have not been
adjusted for Sun-Earth distance, showing the intra-annual and inter-annual variability. The effects of volcanic
eruption and local forest fires are illustrated by the local minima.

4.10 CURRENT ISSUES REGARDING DIRECT NORMAL


IRRADIANCE MEASUREMENTS
Several research topics need to be addressed in continuing research on DNI measurements. A few
are mentioned here.
Systems that focus light on objects (e.g., power towers) or pipes (parabolic troughs) have various
fields of view. For example, many just concentrate the sunlight from the solar disk and little beyond
that. Pyrheliometers have a standard field of view of 5 degrees. Some of the radiation detected by
these pyrheliometers comes from the aureole surrounding the sun and many concentrating systems
cannot focus this circumsolar irradiance. Therefore, DNI measurements overestimate the irradiance
capable of being focused by concentration solar systems. While models exist to estimate the cir-
cumsolar contribution to the DNI, there is considerable uncertainty in these estimates especially for
dusty environments, A brief discussion of circumsolar measurement history and ongoing research
can be found in Chapter 14.
The extensive pyrheliometer comparison performed in 2008 and 2009 demonstrated the uncer-
tainties that then current commercial instruments could achieve (Michalsky et al. 2011). Several
manufacturers have introduced new pyrheliometers and improved their designs in the last 10 years.
It  would be useful to conduct another comparison similar to this to access the current state of
pyrheliometry.
As discussed in Section 4.6.2, pyrheliometer response can change with the solar zenith angle of
the measurement. This response needs to be investigated for multiple samples, at least three, for
example, of all commercial pyrheliometers. Some means to correct for this dependence and obtain
better estimates of this effect on uncertainty in DNI measurements should be studied.
Direct Normal Irradiance 89

QUESTIONS
1. Write an equation  that relates GHI, DNI, and DHI? If GHI is 1,000  Wm−2, DHI is
100 Wm−2, and DNI is 1000 Wm−2, what is the solar zenith angle to two decimal places?
2. Name, at least, three sources of extinction that lead to the DNI detected at the Earth’s sur-
face being less than at TOA.
3. What is the main reason that the ETR from the Sun changes throughout the year? How
large can this change be? What is the Sun’s distance (approximately) in astronomical units
on February 28th (consult an earlier chapter in this book)?
4. Rank these DNI measurements in order of accuracy with best first: (1) thermopile pyrheli-
ometer, (2) GOES satellite retrieval, (3) ACR.
5. Name at least two sources of uncertainty when using an RSR to measure DNI that make
DNI measurements less certain than measurements made using a thermopile pyrheliometer.
6. How would one obtain a pyrheliometer calibration that is linked to the WRR?
7. What is a good reference for choosing a pyrheliometer with proven accuracy?
8. Did the DNI reach 700 Wm−2 in Eugene, Oregon, on January 1, 2009? How about January 2,
2009?
9. How many Dobson units of ozone were over Denver (40°, –105°) on March 30, 2017? What
satellite measured this?
10. What should accompany every measurement of DNI?

REFERENCES
Cook, R. R. 2002. Assessment of uncertainties of measurement for calibration and testing laboratories.
National Association of Testing Authorities, Australia. https://g1ar.files.wordpress.com/2010/11/nata-
assessment-of-uncertainties.pdf (Accessed May 24, 2019).
Finsterle, W. 2006. WMO international pyrheliometer comparison IPC-X, final report. IOM report No. 91,
TD 1320. Geneva, Switzerland: WMO.
Fröhlich, C. 1991. History of solar radiometry and the World Radiometric Reference. Metrologia 28:
111–115.
Gueymard, C. 1998. Turbidity determination from broadband irradiance measurements: A detailed multicoef-
ficient approach. Journal of Applied Meteorology 37: 414–435.
Habte, A., M. Sengupta, A. Andreas, S. Wilcox, and T. Stoffel. 2016. Intercomparison of 51 radiometers for
determining global horizontal irradiance and direct normal irradiance measurements. Solar Energy
133: 372–393.
International Organization for Standardization (ISO). 2018. Solar Energy--Specification and Classification
of Instruments for Measuring Hemispherical Solar and Direct Solar Radiation. ISO 9060. Geneva,
Switzerland: ISO. https://www.iso.org/standard/67464.html (Accessed May 24, 2019).
Joint Committee for Guides in Metrology (JCGM). 2008. Evaluation of measurement data—Guide to the
expression of uncertainty in measurement. GUM 1995 with minor revisions. Paris: Bureau International
des Poids et Mesures. https://www.bipm.org/utils/common/documents/jcgm/JCGM_100_2008_E.pdf
(Accessed May 24, 2019).
Kopp, G., and J. L. Lean. 2011. A new, lower value of total solar irradiance: Evidence and climate significance.
Geophysical Research Letters 38: L01706. doi:10.1029/2010GL045777.
LI-COR, Inc. 1982. LI-2020 Automatic Solar Tracker Instruction Manual. Publication No. 8203-27. Lincoln,
NE: LI-COR.
Michalsky, J. J., G. P. Anderson, J. Barnard, J. Delamere, C. Gueymard, S. Kato, P. Kiedron, A. McComiskey,
and P. Ricchiazzi. 2006. Shortwave radiative closure studies for clear skies during the atmospheric
radiation measurement 2003 aerosol intensive observation period. Journal of Geophysical Research
111: D14S90. doi:10.1029/2005JD006341.
Michalsky, J. J., J. A. Augustine, and P. W. Kiedron. 2009. Improved broadband solar irradiance from the
multi-filter rotating shadowband radiometer. Solar Energy 83: 2144–2156.
Michalsky, J., E. G. Dutton, D. Nelson, J. Wendell, S. Wilcox, A. Andreas, P. Gotseff et al. 2011. An extensive
comparison of commercial pyrheliometers under a wide range of routine observing conditions. Journal
of Atmospheric and Oceanic Technology 28: 752–766.
90 Solar and Infrared Radiation Measurements

Myers, D. R., S. Wilcox, W. Marion, R. George, and M. Anderberg. 2005. Broadband model performance for
an updated national solar radiation database in the United States of America. Proceedings of the ISES
Solar World Conference. Orlando, FL: ISES.
National Renewable Energy Laboratory (NREL). 2010. Broadband Outdoor Radiometer Calibration
(BORCAL 2010-02), August 5. Golden, CO: NREL.
Ohmura, A., H. Gilgen, H. Hegner, G. Müller, M. Wild, E. G. Dutton, B. Forgan, et al. 1998. Baseline sur-
face radiation network (BSRN/WCRP): New precision radiometry for climate research. Bulletin of the
American Meteorological Society 79: 2115–2136.
Perez, R., P. Ineichen, K. Moore, M. Kmiecik, C. Chain, R. George, and F. Vignola. 2002. A new operational
satellite-to-irradiance model. Solar Energy 73: 307–317.
Reda, I. 1996. Calibration of a Solar Absolute Cavity Radiometer with Traceability to the World Radiometric
Reference, NREL/TP-463-20619. Golden, CO: NREL.
Reda, I., D. Myers, and T. Stoffel. 2008. Uncertainty estimate for the outdoors calibration of solar pyranom-
eters: A metrologist perspective. Journal of Measurement Science 3: 58–66.
Reda, I., M. Doorhagi, A. Andreas, and A. Habte. 2017. NREL Pyrheliometer Comparisons, September
25–October 6 (NPC-2017). NREL/TP-1900-70436, November. Golden, CO: NREL. http://www.nrel.
gov/publications.
Sengupta, M., A. Habte, S. Kurtz, A. Dobos, S. Wilbert, E. Lorenz, T. Stoffel, et al. 2015. Best practices hand-
book for the collection and use of solar resource data for solar energy applications. NREL/TP-5D00-
63112. Golden, CO: NREL. http://www.nrel.gov/docs/fy15osti/63112.pdf.
Stoffel, T., D. Renne, D. Myers, S. Wilcox, M. Sengupta, R. George, and C. Turchi. 2010. Concentrating
solar power: Best practices handbook for the collection and use of solar resource data. NREL/TP-550-
47465. 146 pp. National Renewable Energy Laboratory, Golden, Colorado. https://www.nrel.gov/docs/
fy10osti/47465.pdf (Accessed June 5, 2019).
Stokes, G. M., and S. E. Schwartz. 1994. The  Atmospheric Radiation Measurement (ARM) program:
Programmatic background and design of the cloud and radiation test bed. Bulletin of The American
Meteorological Society 75: 1201–1221. doi:10.1175/1520-0477.
Taylor, B. N., and C. E. Kuyatt. 1994. Guidelines for evaluation and expressing the uncertainty of NIST mea-
surement results. NIST Technical Note 1297. Gaithersburg, MD: National Institute of Standards and
Technology. http://physics.nist.gov/cuu/Uncertainty/basic.html.
Vignola, F., and F. Lin. 2010. Evaluating calibrations of normal incident pyrheliometers. Proceedings of the
SPIE Conference. San Diego, CA: SPIE.
Vignola, F., and F. Lin. 2011. Characterizing the performance of an Eppley normal incident pyrheliometer.
Proceedings of the Solar 2011. Raleigh, NC: American Solar Energy Society.
Vignola, F., P. Harlan, R. Perez, and M. Kmiecik. 2005. Analysis of satellite derived beam and global solar
radiation data. Proceedings of the ISES Solar World Conference. Orlando, FL: ISES.
Vignola, F., P. Harlan, R. Perez, and M. Kmiecik. 2007. Analysis of satellite derived beam and global solar
radiation data. Solar Energy 81: 768–772.
Vuilleumier, L., M. Hauser, C. Félix, F. Vignola, P. Blanc, A. Kazantzidis, and B. Calpini. 2014. Accuracy
of ground surface broadband shortwave radiation monitoring. The  Journal of Geophysical Research
Atmospheres 119: 13,838–13,860. doi:10.1002/2014JD022335.
Wesely, M. L. 1982. Simplified techniques to study components of solar radiation under haze and clouds.
Journal of Applied Meteorology 21: 373–383.
World Meteorological Organization (WMO). 2017. WMO guide to meteorological instruments and methods
of observation. WMO-No. 8, Chapter 7. https://library.wmo.int/opac/doc_num.php?explnum_id=4147.
5 Broadband Global Irradiance

This instrument, as its name (from the Greek, fire, up, a measure) indicates, is intended to
measure the heat equivalent of radiation received from or going out toward the complete
hemisphere above the plane of the measuring surface.
C. G. Abbot and L. B. Aldrich
The Pyranometer: An Instrument for Measuring Sky Radiation, 1916

5.1 INTRODUCTION TO GLOBAL HORIZONTAL


IRRADIANCE MEASUREMENTS
Global horizontal irradiance (GHI) is the total solar flux available from the hemispheric sky dome
that is incident on a horizontal surface. By convention global irradiance implies the amount of
shortwave solar radiation in the spectral range of about 280–4000 nanometers (nm) received by a
hemispheric, 2π steradian, field of view on any surface. The solar spectrum is continuous and var-
ies significantly with wavelength. In terms of irradiance measurements, the solar spectrum can be
separated into the ultraviolet, visible, and near infrared as is discussed in more detail in Chapter 7.
This chapter covers instruments used to measure shortwave radiation in total. GHI is the sum of
the solar radiation coming directly from the solar disk, also called the direct normal irradiance
(DNI) or “beam irradiance,” on a surface normal to the rays from the Sun, projected (Equation 5.1)
onto a horizontal plane plus the solar radiation coming from all other directions of the sky dome,
or “sky irradiance” (Figure 5.1). The solar radiation coming from all parts of the sky dome other
than directly from the Sun is called diffuse horizontal irradiance (DHI). Chapter 6 discusses DHI
in detail. Planar surfaces of arbitrary orientation may receive solar radiation directly from the Sun
plus diffuse radiation from the sky and reflected radiation from the ground. Historically, this has
been termed total solar radiation on a tilted surface. Those involved with photovoltaic systems
refer to this total irradiance as plane of array (POA) irradiance. Studies of the thermal perfor-
mance of buildings rely on the amount of solar irradiance on various parts of a building envelope
and refer to this total irradiance as global irradiance on tilted surfaces. Traditionally, the study of
non-concentrating thermal collectors also has used the term global irradiance available to tilted flat
plate collectors. In this book, global horizontal irradiance (GHI) will refer to the total solar irradi-
ance on a horizontal surface. When discussing irradiance on a surface oriented at any angle other
than horizontal, the irradiance will be referred to as global tilted irradiance (GTI) (see Chapter 9).
Global irradiance is measured by a pyranometer. The root of the word pyrano is derived from the
Greek “pyr” meaning fire or heat and “ano” meaning sky. Therefore, a pyranometer is a meter for
measuring heat from the sky. Earlier versions of the pyranometer were referred to as a 180° pyrheli-
ometer, but the terminology was changed to pyranometer to help avoid the confusion with the term
pyrheliometer that is used to describe the instrument for measuring the DNI or beam irradiance
coming directly from the Sun (see Chapter 4).
As described in Chapter 2, GHI is equal to the DNI projected onto the horizontal surface plus the
DHI that is from all other areas of the sky.

GHI = DNI ⋅ cos ( sza ) + DHI (5.1)

where sza is the solar zenith angle of the Sun as measured from the vertical (zenith) at the time of
interest. Near solar noon, sza is smallest (determined by the location and date), and at sunrise or

91
92 Solar and Infrared Radiation Measurements

FIGURE  5.1 Drawing showing the sources of DHI as the DNI passes through the atmosphere. DHI can
come from scattering by the constituents of the atmosphere, clouds, and a portion of ground-reflected irradi-
ance that is further scattered by the clouds or atmospheric constituents.

sunset, sza is 90°. When using angles in the cosine function, most software programs interpret the
value in radians instead of degrees. For example, one of the most common errors in computing GHI
is to use degrees instead of radians when calculating the cosine of the solar zenith angle. Always
confirm that calculations use the appropriate angular units.
On a clear day, the GHI is predominately composed of the DNI projected onto the horizontal surface,
and the DHI contribution is often a small fraction of GHI in Equation 5.1. The relationship over a clear
day between GHI, DNI, and DHI is illustrated in Figure 5.2. The direct horizontal irradiance (DrHI),
is the projection of DNI on a horizontal surface (i.e., DNI times the cosine of the solar zenith angle) as
shown by the dashed line in Figure 5.2. This produces the bell-shaped curve just below the GHI values
for the clear-sky conditions seen in the figure. DNI, shown by the dotted gray line, can be greater than
GHI, especially in the early morning and late afternoon hours. Depending on the location and season,
the DNI can exceed the GHI near solar noon. The cosine of the solar zenith angle is smallest during the
morning and evening hours, and the DrHI should always be smaller than GHI, as seen in Figure 5.2.
As described in Chapter 3, GHI has been measured in many ways during the past 100 years.
Abbot and Aldrich (1916) discussed a design for pyranometers. A  variety of early pyranometers
were developed and used in the field from the Robitzsch actinograph and the more sophisticated
Eppley bulb-type pyranometer up to today’s modern designs (see Chapter 3). As evident from their
historical development, the design goals for these instruments included improved measurement per-
formance and durability. Understanding their development is useful because a considerable amount
of irradiance data was and continues to be gathered using these historical pyranometers. However,
pyranometers in wide use today are much more accurate than their predecessors, and they output
their data to automated data-logging equipment instead of analog strip-chart recorders that must be
digitized for analysis.
Broadband Global Irradiance 93

FIGURE 5.2 Plot of 1-minute data for GHI, DNI, and DHI irradiance on a clear day in Eugene, Oregon. GHI
is the solid line, the dotted line is DNI, and the dot-dash line is DHI. The dashed line is the DNI projected
onto a horizontal surface (DrHI). Under very clear skies DHI can be less than 100 Wm−2 and decreases in
the afternoon as the amount of atmosphere traversed by the scattered light increases. After 19:12, the direct
sunlight is blocked by an obstruction near the horizon.

Based on their detector design, three types of pyranometers will be discussed in more detail in
this chapter:

• Black-disk thermopile pyranometers


• Black-and-white thermopile pyranometers
• Photodiode-based pyranometers

The basis for all thermopile radiometers is the thermoelectric effect, the generation of voltage from
temperature differences in two dissimilar metals. The thermoelectric effect was discovered in 1821
by Thomas Seebeck, a German physicist. The  Seebeck Effect predicts 41  microvolts (µV) from
junctions of copper and constantan (a copper-nickel alloy) per Kelvin (as determined by T2–T1 in
Figure 2.17 from Chapter 2) at room temperature. A thermocouple consists of one such junction.
A thermopile is formed by connecting a number of thermocouples in series. A copper constantan
thermopile is shown in Figure 5.3. Constantan wire is wrapped around a solid core, plated with
copper, and covered by a blackened surface to form the detector. The  detector is mounted in a
metal body and positioned under a set of precision-ground optical glass domes for protection and
thermal insulation. The resulting “single-black” thermopile pyranometer has the hot junctions of the
thermopile detector attached to a blackened disk that absorbs incoming solar radiation. The voltage
produced by the detector is proportional to the heat flow between the irradiated black-disk and the
cold junction of the thermopile that is attached to the body of the pyranometer. Some pyranometers
have a pair of thermopiles in parallel, called “bucking” thermopiles, with one thermopile being
heated by the Sun and the other thermopile being shaded from the Sun and typically looking at
the body of the pyranometer. This combination is purported to compensate for the thermal offset
experienced by a single thermopile due to the exchange of infrared radiation between the “hot”
detector and the “cold” sky, but the exact details of this construction are proprietary. Alternatively,
black-and-white or star-type pyranometers are designed to mitigate these thermal offsets by using
thermopiles made from alternating black (hot junction) and white (cold junction) segments form-
ing the receiver surface. Although this design results in a relatively larger detector surface than the
94 Solar and Infrared Radiation Measurements

FIGURE  5.3 Picture of a copper-constantan thermopile manufactured by The  Eppley Laboratory, Inc.
Constantan wire is wrapped around a solid core and then dipped into a copper bath to create a series of ther-
mocouples called a thermopile. (Courtesy of Warren Gretz, NREL staff photographer.)

single-black method, limiting the instrument’s time response and challenging the ability to con-
struct a planar surface, the thermal offset is nearly eliminated by equally exposing the hot and cold
thermopile junctions to the same infrared (sky) environment (see Figures 5.15 and 5.16).
In  1905, Albert Einstein explained the photoelectric effect as the emission of electrons by a
surface, typically a metal, in response to the absorption of photons (i.e., individual quanta of light).
The development of photovoltaic (PV) devices, also called solar cells, is based on the photoelectric
effect. The photodiode pyranometer uses a solar cell to generate a current across a precision resister
to produce a voltage that is proportional to the incident radiation. Each type of pyranometer will be
described in more detail in subsections of this chapter.
The  International Organization for Standardization (ISO) and the World Meteorological
Organization (WMO) have undertaken the classification of pyranometers based on their measure-
ment performance characteristics (ISO, 2018; WMO 2008). ISO classifies pyranometers into three
classes: Class A, with the least uncertainty over a range of conditions; Class B with an overall larger
uncertainty under the same range of conditions; and Class C with the overall largest set of uncertain-
ties (Table 5.1). These classifications are based on overall uncertainty, and some characteristics of
pyranometers with a lower classification can have less uncertainty for a specific feature than those
with a higher classification. Roughly speaking the new ISO classification correspond to the old ISO
classifications of Secondary Standard (A), First Class (B), and Second Class (C). Although most
of the ISO and WMO standards are identical, the characterizations specified are not the same in all
cases. ISO values are specified with a ± range and ± “guard band” uncertainty, i.e., a response time
of <10 s (1 s) for a Class A instrument means the response time is less than 10 seconds and has a
guard band of one second. The guard band signifies that the response time has to be measured to
within plus or minus one second. When ISO and WMO specifications differ, the ISO specifications
will be listed first indicating the acceptance interval and guard band width (BIPM, 2012), followed
by a “/”, and then the WMO characteristics. The ISO and WMO specifications evolve over time and
requirements change and characteristics of instruments improve.
The main difference between the ISO and the WMO standards is in the specification of spec-
tral sensitivity. The WMO standard specifies about twice the spectral range that the old ISO stan-
dard required. Silicon photodiode-based pyranometers do not respond to solar irradiance beyond
Broadband Global Irradiance 95

1,100 nm, so while some photodiode pyranometers perform better than some class B or C pyranom-
eters, they technically do not qualify under these WMO or the 1990 ISO standards.
The  new ISO spectral specifications rely on the degree to which the pyranometer’s spectral
response matches an ideal pyranometer with a uniform or flat spectral response. A set of standard
clear sky spectrum is used to evaluate the pyranometer’s spectral dependence. The spectra are spec-
ified for an air mass of 1.5 and 5 and the change in the pyranometer’s average responsivity is calcu-
lated for each air mass. For example, the average spectral transmission of the pyranometer’s dome
is evaluated at an air mass of 1.5 and 5 for several different atmospheric conditions. The difference
in responsivity is then used to determine the pyranometer’s spectral classification. The maximum
differences for each ISO classification are given in Table 5.1. For the WMO classification, the spec-
tral classification depends on the spectral transmission of the dome or the spectral range to which

TABLE 5.1
Specifications for Pyranometer Classification
ISO Specifications/WMO
Specifications Class A/High-Quality Class B/Good-Quality Class C/Moderate-Quality
Response time (to 95% of final <10 s (1 s) / <15 s <20 s (1 s) / <30 s <30 s (1 s) / <60 s
value)
Zero offset response: ±7 Wm−2 (2 Wm−2)  ±15 Wm−2 (2 Wm−2) ±30 Wm−2 (3 Wm−2)
1. To a 200 Wm−2 net radiant
loss to sky (ventilated)
Zero offset response: ±2 Wm−2 (0.5 Wm−2) ±4 Wm−2 (0.5 Wm−2) ±8 Wm−2 (1 Wm−2)
2. To a 5°Kh−1 change in
ambient temperature
Total zero offset response: type ±10 Wm−2 (2 Wm−2) ±21 Wm−2 (2 Wm−2) ±41 Wm−2 (3 Wm−2)
(1), (2) and other sources/no
WMO equivalent
Non-stability (change in ±0.8% (0.25%) ±1.5% (0.25%) ±3.0% (0.5%)
sensitivity per year)
Non-linearity (deviation from ±0.5% (0.2%) ±1.0% (0.2%) ±3.0% (0.5%)
sensitivity at 500 Wm−2 over
100 to 1,000 Wm−2 range)
Directional response for beam ±10 Wm−2 (4 Wm−2) ±20 Wm−2 (5 Wm−2) ±30 Wm−2 (7 Wm−2)
radiation for incident angles up
to 90º (maximum uncertainty
in the direct horizontal
component assuming DNI of
1,000 Wm−2)
(a) Spectral effect on clear sky (a) ±0.5% (0.1%) (a) ±1% (0.5%) (a) ±5% (1%)
GHI: Maximum deviation
obtained using reference clear
sky spectra at air mass 1.5 and
5.0 as compared to a
pyranometer with a spectrally
flat response (ISO).
(b) Spectral selectivity (deviation (b) ±3% (b) ±5% (b) ±10%
of the product of spectral
absorptance and transmittance
from the mean within the 300
to 3,000 nm range (WMO))
(Continued)
96 Solar and Infrared Radiation Measurements

TABLE 5.1 (Continued)


Specifications for Pyranometer Classification
ISO Specifications/WMO
Specifications Class A/High-Quality Class B/Good-Quality Class C/Moderate-Quality
Temperature response: Percent ±1% (0.2%) / ±2% ±2% (0.2%) / ±4% ±4% (0.5%) / ±8%
deviation in response within
the −10 to 40°C range relative
to the signal at 20°C/error due
to 50°C ambient temperature
change
Tilt response (deviation from ±0.5% (0.2%) ±2% (0.5%) ±5% (0.5%)
horizontal responsivity due to
tilt from horizontal to vertical
at 1,000 Wm−2)
Additional signal processing ±2 Wm−2 (2 Wm−2) ±5 Wm−2 (2 Wm−2) ±10 Wm−2 (2 Wm−2)
errors, not in WMO
Only in WMO standard/ ±1 Wm−2 ±5 Wm−2 ±10 Wm−2
resolution (smallest detectable
change)
Achievable uncertainty (CIMO
2017)
Hourly totals 3% 8% 20%
Daily totals 2% 5% 10%
Suitable applications Working standard Network operations Low-cost network

the pyranometer is responsive. For  example, the optical domes for the thermopile pyranometers
are used to isolate the thermopile sensor from the environment. The material used in these domes
transmits a limited portion of the solar spectrum to the sensing thermopile. Ideally, the spectral
transmittance would be the same for all solar irradiance wavelengths (generally 300–3,000 nm).
The domes of several class A and class B thermopile pyranometers are made of materials that meet
the old ISO spectral standards but do not transmit wavelengths over the whole 300 to 3,000 nm
range and hence do not meet WMO specifications. By changing the standard to performance based
instead of specifying the range of wavelengths to which the pyranometer is sensitive, the new ISO
standard gives an operational standard that the pyranometer must meet.
The  rational for characterizing the performance of pyranometers in specific categories will
become more apparent from the more detailed descriptions found later in this chapter. Four main
performance characteristics will help determine the quality of pyranometers:

• Deviation from true angular response of the detector to receive radiation within a 2π ste-
radian field of view
• Change in responsivity as a function of ambient temperature
• Dependence of pyranometer responsivity on the wavelength of incident solar radiation
• Thermal offsets of the pyranometer as a function of the net infrared radiation

The characteristic response of the pyranometers to changes in these factors will be discussed for
each of the pyranometer types along with other important factors, such as changes in response as the
solar azimuth angle changes. These factors all affect the accuracy of pyranometer measurements,
and the magnitude of these effects will be used to determine the uncertainty in the irradiance values
Broadband Global Irradiance 97

obtained using the instrument. The specifications are one way to differentiate between the capa-
bilities of each pyranometer and to provide guidance in matching the application’s particular solar
radiation measurement requirements with a suitable pyranometer.

5.2 BLACK-DISK THERMOPILE PYRANOMETERS


Major design features of black-disk thermopile pyranometers are as follows:

• A single all-black thermopile measures the heat flow between the black-disk receiver and
the heat sink in the body of the pyranometer.
• Two quartz domes reduce convective and radiative thermal losses.
• Optical design minimizes solar azimuth angle dependence.
• GHI and DHI measurements are subject to discernable thermal offsets.

Class A and most class B instruments use the single black thermopile design. Single black-disk ther-
mopile pyranometers measure the heat flow between a black-disk receiver exposed to the incident
radiation and the heat sink in the body of the pyranometer. Radiant energy is absorbed by the black
receiver that is thermally attached to the “hot junction” of the thermopile, while the “cold junction”
is connected to the heat sink for the pyranometer. The temperature difference between the hot junc-
tion of the thermopile and the cold junction produces a voltage typically on the order of 10 millivolts
(mV) for a flux of 1,000 Wm−2 incident on the detector. The thermoelectric effect was observed by
Seebeck, Peltier, and Thomson (Bunch and Hellemans 2004; Velmre 2007). Danish physicist Hans
Christian Ørsted helped explain this thermoelectric effect (Velmre 2007).
The voltage created in the circuit is on the order of several microvolts (µV) per Kelvin (K) dif-
ference in temperature. In  some pyranometers, a copper–constantan thermopile is used because
copper–constantan has a relatively large Seebeck coefficient of 41  µV K−1  at room temperature.
Many thermocouples are connected in series to form the thermopile that produces a larger, more
easily measured output voltage.
For accurate solar irradiance measurements, the heat flow should be only between the disk and
the pyranometer’s heat sink that has considerable thermal mass compared to the disk. A thinner
sensor (less thermal mass) results in a faster response time. Heat losses by other means such as con-
vection, advection, conduction, and radiation reduce the accuracy of the measurements. Convection
is the movement of heat in a medium such as water or air. An example of convection is using a ceil-
ing fan to circulate warm air near the ceiling of a room to the lower part of the room. Advection is
the process that transfers heat from a surface to a gas or liquid. An example of advection is when
a fan is used to blow air across hot elements in a heater, resulting in the transfer of heat to the air.
Conduction is the movement of heat in a material. An example of conduction is when a spoon is put
in a hot cup of coffee and the handle of the spoon becomes hot as heat travels up the spoon. Heat
also is transferred through electromagnetic radiation. An example of radiative heating is microwave
heating of water or sunlight heating a solar collector to heat water (see Section 2.12 in Chapter 2 for
more information on thermodynamic fundamentals).
To reduce heat losses by convection and advection from wind blowing across the pyranometer
receiver, high-quality pyranometers use two glass domes. One glass dome prevents advective heat
losses from wind blowing across the pyranometer receiver being heated by the Sun. A second glass
dome acts much like a double-pane window to reduce the heat loss with a layer of air between the
domes acting as a buffer to further reduce convective heat losses.
Conductive heat losses are reduced by isolating the thermopile hot junction from other areas of
the pyranometer. The better the thermal isolation, the slower the heat transfer from the pyranometer’s
receiver to unintended heat sinks. The heat sink is a large thermal mass that maintains a temperature
near the ambient temperature. Most thermopile pyranometers use the body of the pyranometer for
the heat sink. Ideally, only the heat flow from the receiver disk to the heat sink should be allowed.
98 Solar and Infrared Radiation Measurements

Stopping radiative heat loss from a pyranometer is difficult because an accurate pyranometer
should pass all solar radiation through the glass domes, including the near-infrared (NIR) radiation
to be measured. Therefore, the best pyranometers have precision-ground glass domes that uni-
formly transmit irradiance from around 300 to 3,000 nm. However, the receiver disk that absorbs
solar radiation also detects and emits thermal infrared (IR) radiation (wavelengths longer than
3,000  nm). The  glass domes that allow the IR portion of the solar spectrum to impinge on the
receiver disk also allow thermal IR radiation to radiate to the dome and then to the sky. This action
causes the receiver disk to become cooler as it radiates to the cooler sky. This reaction is the basis
for the so-called thermal offset found in pyranometers designed with a single black disk (Dutton
et al. 2001).
As described previously, thermopile-type pyranometers produce a voltage proportional to the
energy flow from the receiver disk to the cold junction or body of the instrument. At  night, the
energy flow reverses as the receiver disk and protective domes cool due to exposure to the cold sky,
and a negative output voltage from the thermopile results. During the day, there is a similar heat
flow (thermal IR radiation) from the sensor disk to the cooler sky. This path to the sky reduces the
energy flow through the thermopile and, hence, reduces the voltage that would be produced other-
wise. This radiation to the sky is called the IR radiative loss or thermal offset, resulting in a negative
bias in the pyranometer readings (Dutton et al. 2001; Gulbrandsen 1978).
Because the effective sky temperature varies with the amount and vertical distribution of clouds,
atmospheric water vapor, and aerosol loading, the pyranometer thermal offsets vary with location,
time of day, and season, making it difficult to account for this thermal offset. It can range anywhere
from zero during foggy conditions to as much as −30 Wm−2 for high-altitude clear-sky conditions,
depending on the pyranometer construction. The  use of ventilators also will affect the thermal
offset contributions to uncertainties in pyranometer calibration and field measurement performance.
A new, improved pyranometer design has significantly reduced this thermal offset.
Examples of three black-disk thermopile pyranometers in common use are examined here.
The Kipp & Zonen model CM 22 pyranometer (Figure 5.4) is considered a class A pyranometer.

FIGURE 5.4 Kipp & Zonen CM 22 pyranometers in VENs during outdoor calibrations at the Solar Radiation
Research Laboratory (SRRL) in Colorado.
Broadband Global Irradiance 99

TABLE 5.2
Specifications of the Kipp & Zonen CM 22 and CM 21 Pyranometers
Characteristic CM 22 (Class A) CM 21 (Class A)
Spectral range 250–3,500 nm (50% transmission points) 285–2,800 nm (50% transmission points)
290–3,500 nm (95% transmission points) 335–2,200 nm (95% transmission points)
Typical sensitivity 10 µV/Wm−2 10 µV/Wm−2
Typical impedance 10–100 ohms (Ω) 10–100 Ω
Linearity ±0.2% < 1,000 Wm−2 ±0.2% < 1000 Wm−2
Temperature dependence ±0.5% (−20 to +50°C) ±1.0% (−20 to +50°C)
Non-stability ±0.5% sensitivity change per year ±0.5% sensitivity change per year
Cosine response ±1% deviation at 60° solar zenith angle ±1% deviation at 60° solar zenith angle
Directional response <5 Wm−2 up to 80° solar zenith angle <10 Wm−2 up to 80° solar zenith angle
(for beam irradiance)
Response time <5 seconds for 95% <5 seconds for 95%
Tilt response <0.2% <0.2%
Quartz dome High quality quartz dome 2 mm thick Schott K5 optical glass 2 mm thick, 30 mm
inner and 50 mm outer dome diameter

The factory specifications for the CM 22 are given in Table 5.2. The CM 22 is similar to the Kipp &
Zonen CM 21 except the CM 22 uses a high-quality quartz dome and the CM 21 uses a Schott K5
optical glass dome that transmits a narrower range of the solar spectrum in a less uniform man-
ner. The CM 21 is also classified as a class A pyranometer, and its specifications are also given in
Table 5.2. Older versions of the Kipp & Zonen CM 11 and CM 21 pyranometers have domes with
nonuniform transmission values at the short-wavelength end of the solar spectrum. The 50 percent
transmission point in the older pyranometers started around 310 nm instead of 285 nm for the cur-
rent versions of these pyranometers.
The Eppley Laboratory, Inc. standard precision pyranometer (SPP) (Figure 5.5a and b) is con-
sidered a class A  pyranometer, and its factory specifications are given in Table  5.3. Note that
only some of the factory specifications are described in terms that translate directly into ISO or

FIGURE 5.5 (a) Image of an Eppley SPP pyranometer. Although structurally the SPP looks much like the older
Eppley PSP pyranometer, several changes were made to reduce the thermal offset and improve cosine response.
(b) Basic schematic drawing of an Eppley SPP pyranometer. Sun shield represented by the dotted oval.
100 Solar and Infrared Radiation Measurements

TABLE 5.3
Specification of an Eppley SPP Pyranometer (Class A)
Characteristic Specification
Spectral range 295–2,800 nm
Spectral selectivity 300–1,500 nm <2%
Typical sensitivity 8 µV/Wm−2
Typical impedance 700 Ω
Linearity ±0.5% from 0 to 2,800 Wm−2
Temperature dependence ±.5% (−30 to +50°C)
Zero offset—nighttime thermal offset ±5 Wm−2
Zero offset at 700 Wm−2 with a temperature change from 30 to 25°C ±2 Wm−2
Stability ±0.5% sensitivity change per year
Directional response ±10 Wm−2
95% Response time 51 second (1/e signal)
Glass domes Clear WG295 glass

WMO specifications. In fact, specifications among manufacturers are not uniform. There are many


reasons for this diversity in specification, and much of this variance results from the way the manu-
facturers conduct their standardized tests. Many performance specifications are best determined
indoors under controlled conditions. Manufacturers also perform outdoor tests to determine the
responsivities under conditions experienced in the field. There  are many ways to perform tests
on pyranometers, and descriptions of factory specifications often relate directly to the testing
method that is used. Therefore, if manufacturers use different testing methods, it is not  always
easy to compare specifications because the details in the specifications are different. In addition,
the performance of various pyranometer models is affected differently by environmental condi-
tions such as air temperature and wind speed. The best measurement comparisons are done side
by side under conditions likely to be experienced by the pyranometer in the field. Unlike pyrheli-
ometers (described in Chapter 4), no absolute reference pyranometer exists (i.e., no ISO 9060 class
AA) that can be used for calibrations and accurate comparisons. At best, a reference GHI value is
obtained by projecting the DNI, measured by an absolute cavity radiometer, onto the horizontal
surface and adding the DHI obtained by using an offset-corrected class A pyranometer or a class
B black-and-white type pyranometer (because of their minimal thermal offset characteristics) with
the direct solar radiation blocked by a shade ball or disk. The  reference diffuse pyranometer is
calibrated by the Shade-Unshade Method, described in Section 5.5.1 and discussed in Chapter 13,
Section 13.2. The black-and-white pyranometer will be discussed in Section 5.3.
The  following characteristics of a black-disk thermopile pyranometer are discussed in more
detail:

• Thermal offsets
• Nonlinearity
• Spectral response
• Angle of incidence response
• Response degradation over time
• Temperature dependence
• Ice and snow on dome—ventilator section
• An optical anomaly
• Care and maintenance
Broadband Global Irradiance 101

5.2.1 tHeRmal offsets


The  thermal offset of a pyranometer is a subtle thermal loss mechanism that affects the perfor-
mance of single black thermopile detector pyranometers. Ground, ambient, and sky temperatures
are terms used when discussing the thermal offset and it is important to clearly define each term.
Ground or surface temperature is the temperature of the ground at the surface or at different depths
in the soil. Ambient temperature is the temperature of the air surrounding an instrument and is
typically measured at a height of 2 meters (m) above the surface. Sometimes ambient and ground
temperature are used interchangeably, but they differ. Sky temperature is the effective temperature
of the sky above a site and is measured using a pyrgeometer or infrared thermometer (IRT).
By design, the black disk is a good absorber of solar radiation as well as a good emitter in
the IR portion of the radiation spectrum. The  body of the pyranometer is close to the ambient
temperature of the air surrounding the instrument and is generally warmer than the effective sky
temperature. Therefore, there is IR radiative transfer from the pyranometer’s black disk to the sky.
The transfer actually consists of several steps with the outer dome of the pyranometer radiating to
the sky, the inner dome radiating to the outer dome, and the black disk radiating to the inner dome,
lowering its temperature with respect to the body and producing a negative signal (Dutton et al.
2001; Gulbrandsen 1978; Long et al. 2001). Convective and conductive heat losses also play a role.
Thermal offsets usually vary from minus a few Wm−2 to as much as −20 to −40 Wm−2, depend-
ing on the amount of net IR radiation (Sanchez et al. 2016). One key finding from the research of
Sanchez and others (2016) is that wind speed plays a key role in modulating the thermal offset.
During the day, thermal offsets are larger (more negative) than the nighttime values as the ground,
the air surrounding the pyranometer, the body of the pyranometer, and the sky are heated by solar
radiation (Figure 5.6). This difference causes the thermal offsets to increase during the day. The mag-
nitude of the pyranometer thermal offset is proportional to the net IR at the time of measurement.
The net IR is a function of the effective sky, ambient air, and ground temperatures. The temperature
of the cold heat sink, the body of some pyranometers, is closely related to the ambient air temperature
and the incident solar radiation. The amount of atmospheric precipitable water vapor (PWV) affects

FIGURE  5.6 Calculated thermal offset for an Eppley PSP pyranometer in Eugene, Oregon, during June
2010. One-minute data for all hours in June 2010 are plotted. Using nighttime readings and pyrgeometer data,
the methodology of Younkin and Long (2003) is employed to calculate the thermal offset of the PSP during
the day. During cloudy periods there is minimal thermal offset as the sky temperature is similar to the ambient
temperature. During clearer weather, the thermal offset reaches its maximum values.
102 Solar and Infrared Radiation Measurements

the net IR (i.e., higher PWV generally reduces net IR), and in turn the pyranometer thermal offset.
The  amount of PWV also is affected by barometric pressure and, therefore, the thermal offset is
related to the station elevation. That is, under cloudless sky conditions, lower thermal offsets can be
expected from pyranometers used in Eugene, Oregon, at 150 m (−2 to −10 Wm−2) than the offsets
found in measurements made in Golden, Colorado, at 1,829 m (−15 to −40 Wm−2). The magnitude of
the thermal offsets is significantly less during cloudy periods as clouds act as a thermal blanket and
produce warmer skies, reducing net IR, as opposed to clear periods with cooler effective sky tem-
peratures resulting in higher net IR at the surface. The body of the black-disk thermopile pyranometer
acts as the cold reference junction and is usually at or near ambient air temperature, especially when
installed in a non-heated ventilator. During clear weather, sky temperatures are much colder than
ambient surface air temperatures. Because different sites exhibit different ranges of sky temperatures
due to elevation and amounts of atmospheric water vapor and aerosols, the thermal offsets will be
more severe at some sites than at others. The fact that thermal offsets vary from site to site makes it
difficult to clearly determine the proper calibration constant of a pyranometer, especially ones that
have large thermal offsets. Starting around 2005, some pyranometers have been calibrated to deter-
mine both the broadband responsivity and the IR responsivity (Rnet) (Reda et al. 2005).
Thermal offsets in pyranometer measurements can be explored most easily with nighttime data.
In the absence of solar radiation negative nighttime values from a pyranometer result from ther-
mal offsets. The  thermal offsets differ depending on whether the pyranometer is ventilated and
whether a direct current (DC) or alternating current (AC) motor powers the ventilator (Michalsky
et  al. 2017). Most ventilators heat the body of the pyranometer, exacerbating the thermal offset.
For an Eppley PSP with an Eppley ventilator, the thermal offset at night was observed to be −2 to
−3 Wm−2 greater, more negative, for pyranometers in ventilators with fans powered by AC motors
compared with those powered by DC motors. AC motors use more power than DC motors and
slightly warm the air passing over the body of the pyranometer (Michalsky et al. 2017).
It  is easiest to correlate thermal offsets at night because there is no incoming solar radiation
present to heat the receiver, and the pyranometer output is expected to be zero. Key to modeling
thermal offsets is the use of pyrgeometers from which net IR can be determined from the pyrgeom-
eter thermopile signal. Correlations between net IR and thermal offsets, relative humidity, and sky
temperature have been developed using nighttime pyranometer values (Younkin and Long 2003).
These correlations then can be used with data gathered during the day to adjust the DHI or GHI
measurements based on pyrgeometer thermopile signals during the day. These correlations are spe-
cific to pyranometer and pyrgeometer pairs due to their individual thermal properties, and they can
be used to produce more accurate measurements of GHI and, in particular, DHI (see Figure 5.6).
Unfortunately, there are a limited number of pyrgeometers paired in the field with pyranometers.
Some work has been done to estimate the thermal offsets without the pyrgeometer data (Reda et al.
2005; Vignola et al. 2007, 2008, 2009), but to date a general meteorological model that can be asso-
ciated with a pyranometer has not been developed.
With thermal offsets changing from site to site, it is very difficult to determine the appropri-
ate pyranometer responsivity without correcting for the thermal offset. Systematic biases result if
calibration values are obtained from sites where large thermal offsets occur and are then used at
locations where the thermal offsets are much less.
Three options are available to address this calibration issue:

1. Use calibration values determined without consideration of thermal offsets.


2. Use calibration values determined with mean nighttime offsets subtracted during the day.
3. Use calibration values determined with thermal offsets corrected.

If a pyrgeometer and meteorological data necessary to calculate the pyranometer thermal offsets
is not available, then calibrate the instrument and use it at the same location to minimize thermal
offset issues (also see Chapter 13, Sections 13.2 and 13.3).
Broadband Global Irradiance 103

An alternative that accounts for some of the thermal offsets is to subtract the average night-
time values from the night before and the night after from the daytime values (Dutton et al. 2001).
This method has been shown to account for roughly half of the thermal offsets during the day (Vignola
et al. 2007). If this method is used, the calibration constant should be determined from a calibration
that subtracts the nighttime values from daytime measurements. In other words, the calibration value
used for a pyranometer should be determined in a manner similar to how the pyranometer is used.
The best option is when pyrgeometer and other meteorological data are available from which
pyranometer thermal offsets can be calculated. Then the calibration constant determined with the
thermal offsets subtracted should be used.
In  summary, pyranometer thermal offsets can be an important contribution to the estimated
measurement uncertainty for historical data and current measurements using older pyranometer
designs. Recognizing the importance of this pyranometer response characteristic, manufacturers
have improved their instrument designs to greatly reduce thermal offsets in their pyranometers by
using single black thermopile detectors; for example, see Reda and Andreas (2017).

5.2.2 nonlineaRity
Nonlinearity is the deviation of the radiometer’s responsivity (µV/Wm−2) as the solar irradiance level
increases or decreases. The  ISO specification for nonlinearity is less than ±0.5 percent for a class
A instrument (see Table 5.1). Nonlinearity is defined as the deviation from the responsivity measured at
500 Wm−2 over a range of 100 to 1,000 Wm−2. For example, if the measured responsivity at 500 Wm−2 is
8.00 µV/Wm−2, then the responsivity from 100 to 1,000 Wm−2 has to be between 7.96 and 8.04 µV/
Wm−2. Examples of factory specifications include the Eppley SPP with a linear response better than
±0.5 percent from 0 to 2800 Wm−2, Kipp & Zonen models CM 22 and CM 21 claim a linear response
better than ±0.25 percent for irradiances below 1,000 Wm−2 and Hukseflux model SR30 with a stated
non-linearity of less than ± 0.2% from 100 to 1000 Wm−2 (see Appendix G).
At  high levels of solar irradiance, black-disk thermopile pyranometers have a good linear
response. It is at lower levels of irradiance that the thermopile pyranometers may exhibit apparent
nonlinearity problems. For example, if a pyranometer exhibits a 20 Wm−2 thermal offset at a solar
irradiance of 1,000 Wm−2, then its readings are offset by 20/1000, or 2 percent. If the thermal offset
is 20 Wm−2 and the incident irradiance is 100 Wm−2, then the readings are off by 20 percent.
Consequently, nonlinearity tests are usually performed under indoor laboratory conditions with the
instrument and surroundings at a stable (room) temperature. Under such conditions, there are no thermal
offsets caused by the detector radiating to a colder target. A stable lamp with a chopper wheel is used
to test for response change with intensity. The intensity of light is proportional to the area blocked by
the chopper wheel. The National Institute of Standards and Technology (NIST) has developed a more
exacting way to test for linearity (see Walker et al. 1987). Under laboratory conditions, ISO classifica-
tion specifies a pyranometer as class A if it deviates from a linear response by less than 0.5 percent.

5.2.3 sPeCtRal ResPonse


The spectral response characteristics of thermopile-based pyranometers is largely determined by the
spectral transmission of the material used for the protective and insulating domes and the spectral
absorption of the detector coating. Ideally, the pyranometer output should be proportional to the sum
of the solar flux across all solar wavelengths. The Kipp & Zonen model CM 22 uses two high-quality
quartz domes to protect the detector. The spectral transmission range for these domes is greater than 50
percent from 290 to 3,500 nm. The transmission of light through the domes changes with wavelength
by about 2 percent from 300 to 2,800 nm. The transmission of light through the dome decreases below
300 nm and above 2,800 nm. The transmission of light through an Infrasil II window that is used on The
Eppley Laboratory, Inc. normal incidence pyrheliometer (NIP) [Model NIP] is shown in Figure 5.7.
The quartz dome used with a Kipp & Zonen CM 22 has similar properties to the Infrasil II window.
104 Solar and Infrared Radiation Measurements

FIGURE 5.7 Spectral transmission of light through an Infrasil window from 280 to 3200 nm. Inset shows
transmission from 280 to 400  nm. The  percent transmission varies from around 89 percent at 280  nm to
around 92 percent at 1,100 to 2,600 nm. The transmission drops to 80 percent around 2,950 nm and returns
to around 90 percent around 3,200 nm. Measurements were made by Fuding Lin at the University of Oregon
Support Network for Research and Innovation in Solar Energy.

The spectral distribution of solar energy changes as sunlight takes a longer path through the atmo-
sphere (see Chapter 7). The amount of atmosphere through which sunlight passes is called air mass
(see Section 7.5.1 for a more detailed discussion of air mass). As air mass increases from 1.5 to 6.0,
the percentage of incident solar radiation that is transmitted through each Infrasil II dome decreases
by ~0.1 to ~0.3 percent. The CM 21 uses Schott K5 optical domes and the SPP uses WG295 optical
glass. The transmittance of WG295 Schott glass used in the SPP is shown in Figure 5.8. The WG295
glass cuts off some ultraviolet and some IR radiation. Similar transmission properties are seen in the
Schott K5 domes. The Hukseflux model SR25 pyranometer has a sapphire outer dome providing
uniform transmittance from 285 to 3,000 nm. Spectral ranges can be quoted for a 2 percent change
over a range, at 95 percent transmission points or 50 percent transmission points. Therefore, it is
important to understand the specifications for the range quoted. The current ISO 9060 2018 stan-
dards for pyranometer performance classifications now bases the spectral response characteristics
on the comparison of the spectral responsivity of the pyranometer using reference clear sky spec-
trum at air masses of 1.5 and 5 for a variety of atmospheric conditions (ISO, 2018). The change in
responsivity is compared to the responsivity of an instrument with a flat spectral response. The dif-
ference between the calculated or measured responsivities at the different air masses is then used
to determine the spectral sensitivity of the pyranometer. This  methodology yields a quantitative
measure of the effect of changes in the spectral distribution of incident radiation.

5.2.4 angle of inCidenCe ResPonse


Ideally, the responsivity of a pyranometer to the solar radiation from any direction within its hemi-
spherical field of view, 2π steradians, depends exactly on the cosine of incidence angle. Lambert’s
cosine law states that the irradiance received at a surface should be proportional to the cosine of the
Broadband Global Irradiance 105

FIGURE 5.8 Spectral transmission of light through a Schott WG295 glass window from factory specifica-
tions. Inset shows spectral transmission from 260 to 400 nm.

incident angle. Therefore, a pyranometer with an ideal angular response, or Lambertian response,
would respond to incident radiation in accordance with Lambert’s cosine law. When the pyranom-
eter measurements are not Lambertian, the instrument’s responsivity is said to vary with incident
angle. It is difficult, if not impossible, to produce a pyranometer with a perfect angular response,
especially for a pyranometer with a relatively large receiver that is covered by glass domes. In most
literature and in this book, angular response is referred to as cosine response and a perfect or
Lambertian angular response is called a true angular or cosine response. When describing an instru-
ment’s angular response, the comparison is normalized against a true cosine response. By making
the comparison against an ideal Lambertian response, deviations of a few percent become easier
to detect. Optical leveling of the receiving disk is very important, especially at large solar zenith
angles where small differences in the angle of incidence can translate into large errors in angular
response. The spirit levels used to orient a pyranometer are not always consistent with the optical
plane defined by the detector surface. Imperfections in the glass domes and even refraction of light
as it passes through the domes can cause some deviations from true cosine (Lambertian) response.
For solar irradiance measurements, the sun is considered an idealized point source with the colli-
mated light coming from a given direction.
Results from a calibration of a Kipp  & Zonen CM 10 pyranometer are shown in Figure  5.9.
The  reference GHI is calculated from the DNI measured with an absolute cavity radiometer
(an Eppley automatic Hickey-Frieden [AHF] in this example) projected onto a horizontal surface
plus the diffuse irradiance obtained by a pyranometer with minimal thermal offset (a Schenk star
or Eppley 8-48 pyranometer) shaded by a shade ball. In this book, the reference GHI is obtained by
this method using either an absolute cavity radiometer or a thermopile pyrheliometer for the DNI.
The source of the DNI will be referenced for each example. The GHI readings from the CM 10 were
divided by the reference GHI measurements and normalized to 1.00 at a solar zenith angle of 45°.
During this calibration, there were a limited number of data points near 45 degrees in the after-
noon, hence the average at 45 degrees is heavily weighted by the morning data. Since the absolute
106 Solar and Infrared Radiation Measurements

FIGURE 5.9 Cosine response of Kipp & Zonen CM 10 pyranometer. The reference GHI is calculated using
the summation technique where the direct-horizontal component, obtained from the DNI by multiplying by
the solar zenith angle, is added to the DHI. The DNI is measured using an AHF cavity radiometer and the DHI
is measured with a Schenk Star pyranometer mounted on a tracker using a shade ball. The responsivities are
shown with and without taking the thermal offset into account (see Section 5.2.1). The results are for morning
and afternoon measurements.

accuracy of these reference measurements is less than 1 percent, it will be assumed for these discus-
sions that any deviation from the ratio of 1.00 represents deviation from a calibrated instrument with
a perfect angular response.
In  Figure  5.9, the ratio of the pyranometer output to reference values is plotted against solar
zenith angle. While some of the deviations from 1.00 may result from changes in temperature and
possibly tilt of the instrument, much of the deviation is associated with a nonperfect Lambertian or
cosine response. In Figure 5.9, the correction for thermal offset is up to 0.5 percent, a small fraction
of the deviation from an ideal angular response. Of course, the thermal offset in Eugene for this
analysis is only about −6 Wm−2 due to the amount of water vapor in the atmosphere.
Understanding the effect introduced by imperfect leveling of the pyranometer illustrates the
difficulty in obtaining a true Lambertian response. When the sun is at a 30° solar zenith angle, a
0.5 degree shift in leveling will change the results by about 0.1 percent, at 70° angle of incidence
leads to about a 0.6 percent effect in the measurement, and at 80° leads to a 1.2 percent effect.
Therefore, it is important that the pyranometer is leveled at the site and that the platform support-
ing the pyranometer is securely anchored. It also is prudent to occasionally check the level of the
pyranometer as determined by the integral spirit level.
Most of the deviation from true cosine response comes from the detector failing to absorb inci-
dent irradiance uniformly at all angles of incidence. The thickness of the glass domes, the varia-
tion in the thickness, the concentric alignment of the domes with respect to the receiver disk, the
uniformity of the receiver disk coating, the “flatness” of the receiver surface, and optical leveling of
the receiver disk within the instrument body all contribute to a non-Lambertian response. Dust or
moisture on the dome can significantly contribute to the apparent angular response problem.
Broadband Global Irradiance 107

Ideally, the receiver should be flat or shaped to give the sensor a better cosine response. One
problem that has been observed in a few pyranometers is that the circular receiver disk glued to
the thermopile will become deformed and will take on the rectangular shape of the thermopile
below. This deformation can be caused during manufacture, by vibrations, or by rough treatment
during shipping. If the receiver disk takes on the shape of the underlying thermopile, devia-
tion from a true cosine response becomes pronounced, and the pyranometer should be repaired
(see Figure 5.10). It must be noted that these examples represent older instruments. The manufac-
turers now offer greatly improved designs with higher performance characteristics.

5.2.5 ResPonse degRadation


The  black receiver of the pyranometer is protected from direct contact with the elements but is
exposed to global irradiance, and, like paint, the exposure to sunlight, especially ultraviolet (UV)
radiation, will change the “color” of the paint and the absorptivity of the receiver over time. In the
case of a thermopile pyranometer, the change in detector color is not easily perceived even after
many years, but the change in responsivity typically decreases between 0.5 and 1.0 percent per year.
The  greater the exposure to sunlight, especially UV light, the faster the change in responsivity.
Wilcox et al. (2001) showed that the formula used to measure paint aging also works for modeling
the change in pyranometer responsivity. Using analyses of clear-sky records and long-term calibra-
tion records, Riihimaki and Vignola (2008) confirmed the model of Wilcox et al. (2001) and showed
this effect continues at a linear rate over the decades that a pyranometer may be in the field.
Reference pyranometers usually have limited exposure to sunlight between calibrations.
This limited exposure greatly reduces the degradation rate of the reference pyranometer and again
confirms the paint-aging hypothesis.
Paint coatings on pyranometer and pyrheliometer detectors can differ as the coating for a pyr-
heliometer does not have to produce a good cosine response while the coating for the pyranometer
should have a good Lambertian response to light incident from all angles.

FIGURE 5.10 Figure of a distorted disk on an Eppley PSP pyranometer. The disk should be flat for a good
cosine response. This  disk has been deformed to take on the shape of the thermopile to which the disk is
attached.
108 Solar and Infrared Radiation Measurements

5.2.6 temPeRatuRe dePendenCe


All pyranometers exhibit some change in responsivity as ambient temperature changes. Many
pyranometer manufacturers provide a temperature response curve for the specific model or model
type. Thermopile pyranometers have a temperature dependence even if first-order temperature-
compensating circuits are used to minimize the primary temperature effects. However, correcting
the temperature response at one temperature will affect the temperature response in a different tem-
perature range. Most class B pyranometers have a responsivity that changes by less than 1 percent
over the −20 to 40°C range after temperature compensation circuitry is added. Temperature tests
are performed in the laboratory because the responsivity of pyranometers outdoors is dependent on
many meteorological variables that are not under the control of those making the test. Examples of
the pyranometer temperature dependence are shown in Figure 5.11 for a Kipp & Zonen CM 22 and
an Eppley PSP pyranometer.

5.2.7 iCe and snoW on dome—ventilatoRs


Ventilators are extremely useful for keeping the domes of pyranometers free of moisture, frost,
snow, and dust. If obtaining a complete and accurate measurement record is important, a ventilator
is a necessary addition for any pyranometer installation.
A ventilator is basically a short cylinder for mounting a pyranometer with a fan mounted near the
bottom to blow air under the pyranometer shield and across the dome. Three ventilators are shown in
Figure 5.12. The ventilator on the left is a standard Eppley model ventilator (VEN) with a clear acrylic
plenum surrounding the pyranometer. The ventilator in the middle is a Schenk ventilator, and the ven-
tilator on the right is a Physikalisch-Meteorologisches Observatorium Davos (PMOD) design installed
on an Eppley VEN base. The ventilator fans can be AC or DC powered, and some ventilators include
a resistive element to heat the air slightly just before it blows across the dome, such as the PMOD and
the Schenk VENs shown in Figure 5.12. This arrangement may be helpful in reducing the temperature
difference between the dome and body that is responsible for the IR losses (thermal offsets) discussed
in Section 5.2.1. AC fans generate and dissipate more heat from their motors than DC fans. To a small
degree, VENs can either enhance or reduce the thermal offsets experienced by the instrument.

FIGURE 5.11 Example of the temperature dependence of a Kipp & Zonen CM 22 and an Eppley PSP. These
measurements were done in the laboratory under controlled conditions. Each instrument has its own unique
temperature dependence. The temperature dependence of both instruments meets the factory specifications.
Broadband Global Irradiance 109

FIGURE 5.12 Three VENs mounted on a Sci-Tech tracker. Going from right to left, the VEN on the left is
a Swiss PMOD VEN holding a Kipp & Zonen CM 22 pyranometer. The VEN is mounted on an Eppley VEN
base so the height of the dome matches the height of the domes of the other instruments. PMOD also makes a
base for leveling. The middle VEN is a Schenk VEN for Schenk Star pyranometers. The VEN on the left is an
Eppley VEN holding an Eppley model Precision Infrared Radiometer (PIR) pyrgeometer.

VENs usually have a screen beneath the fan to prevent larger debris from being added to the airflow.
This screen should be cleaned regularly, as often as weekly, depending on local conditions because it
can become blocked and hinder the airflow. In some environments the ventilator works best without
the screen. Sometimes ventilators are mounted above a hole in the platform to allow the free flow of air
and reduce the accumulation of debris that can clog the screen at the bottom of the ventilator.

5.2.8 an oPtiCal anomaly


One anomaly observed with some pyranometers occurs when the Sun is at a specific location in a
cloudless sky. Over a period of about an hour, the irradiance readings first drop and then increase
to an anomalously high reading before falling back to normal behavior. This change is the result
of reflections from the domes that will create a bright spot or “caustic” near the receiver disk.
Normally the caustic doesn’t cause any problems but with some pyranometers during a specific
time of year at a given zenith and azimuthal angle, this caustic will move across the ring surround-
ing the detector and illuminate a small portion of the absorbing disk. As the Sun moves across the
sky, the spot moves off the detector and again crosses the metal ring surrounding the detector. When
the spot is on the ring, the reading from the pyranometer is low and when the spot is on the detec-
tor, the reading can increase by 50 percent or more. The caustic on an Eppley 8-48 pyranometer is
illustrated in Figures 5.13 and 5.14 for a PSP.
Some pyranometers don’t exhibit this effect, while others by the same manufacturer will exhibit
this effect at different but specific times of the year at certain times of the day. Dust or film on the
dome can cause or enhance this effect. Therefore, it is important to keep the domes as clean as
possible. Pyranometers mounted on two-axis trackers usually do not  exhibit this effect because
the pyranometers on trackers always present the same azimuthal geometry to the solar disk. When
analyzing pyranometer data to identify problems, it is important to watch for the occasional clear
days that exhibit this anomalous behavior.
110 Solar and Infrared Radiation Measurements

FIGURE 5.13 Photo of a caustic on an Eppley 8-48 pyranometer. Picture is taken in the lab and an image
of the light source can be seen in the reflection on the top. The caustic is the bright spot on the bottom of the
white wedge.

FIGURE 5.14 Photo of the caustic on an Eppley PSP pyranometer. Moisture on the dome focused the light
on the receptor disk.
Broadband Global Irradiance 111

5.3 BLACK-AND-WHITE PYRANOMETERS


Black-and-white type pyranometers consist of alternating black and white surfaces that are attached
to the hot and cold junctions, respectively, of a thermopile detector. When exposed to solar radia-
tion, the resulting temperature difference between the hot and cold junctions of the thermopile
produces a voltage.
Major features of black-and-white pyranometers are as follows:

• Use of thermopiles with alternating black and white surfaces


• Performance specifications consistent with class C pyranometers
• Responsivity has a measurable temperature dependence
• Responsivity varies with azimuth
• Responsivity affected by pyranometer tilt
• Limited thermal offset

5.3.1 CHaRaCteRistiCs of blaCk-and-WHite PyRanometeRs


The  specifications of the Eppley Laboratory, Inc. model 8-48 (more commonly referred to as a
black-and-white pyranometer) (Figure  5.15) and the Schenk Star pyranometer (Figure  5.16) are
given in Tables 5.4 and 5.5, respectively.
Black-and-white pyranometers are classified as class C pyranometers because the responsivity of
the rather large detector surface changes significantly as a function of azimuth angle and tilt. One of
the main reasons for this is that their alternating black and white surfaces make them sensitive to the
Sun’s azimuth position. The responsivity of the pyranometer changes if the Sun is directly aligned
with a white surface as opposed to being directly aligned with a black surface. This problem can
be addressed to some extent by making the surface areas smaller or using other geometric patterns.
In addition, if the pyranometer is mounted on a tracker, the pyranometer would always have the
same orientation to the Sun and thus be less susceptible to azimuth affects.
Like other pyranometers, black-and-white pyranometers have imperfect angular responses.
The  receiver tends to be much larger than those found in single black-disk designs, making it
more difficult to have a planar surface. The angular response of a Schenk Star type pyranometer is

FIGURE 5.15 Image of Eppley black-and-white type pyranometer.


112 Solar and Infrared Radiation Measurements

FIGURE 5.16 Image of Schenk Star type pyranometer.

TABLE 5.4
Eppley Model 8-48 Pyranometer Specifications
Characteristic Specification
Spectral sensitivity 0.285–2.8 µm
Typical sensitivity 10 µV/Wm−2
Typical impedance 350 Ω
Temperature dependence ±1.5% over ambient temperature range −20 to +40°C
Linearity ±1.0% from 0–1,400 Wm−2
Response time 5 seconds (1/e signal)
Typical cosine accuracy ±2% from normalization 0°–70° incident angle
±5% from normalization 70°–80° incident angle
Mechanical vibration tested up to 20 g’s without damage
Calibration method integrating hemisphere
Size 5.75 inch diameter, 2.75 inches high
Weight 2 pounds
Broadband Global Irradiance 113

TABLE 5.5
Schenk Star Pyranometer Specifications
Characteristic Specification
Measuring range 0–1,500 Wm −2

Spectral sensitivity 0.3–3 µm


Typical sensitivity 15 µV/Wm−2
or 4... 20 mA = 0... 1500 Wm−2
Typical impedance 35 Ω
Temperature dependence <1% of the value between −20 and +40°C
Linearity <0.5% in the range 0.5–1,330 Wm−2
Response time <25 second (95%), <45 second (99%)
Typical cosine accuracy <3% of the value, incident angle 0°–80°
Azimuth response <3% of the value
Stability <1% per year
Resolution <1 Wm−2
Operational ambient temperature range −40 to +60°C
Weight 1.0 kilograms
Cable 2-polar shielded, 3 meter length

FIGURE 5.17 Cosine response of a Schenk Star pyranometer.

illustrated in Figure 5.17. This specific instrument has an excellent cosine response, increasing by


only 2 percent from a solar zenith angle of 22.5° to 82° with peak responsivity around 72°.
As with other black-and-white receiver designs, the Schenk pyranometer has a single glass dome.
Having a single dome also leaves the pyranometer subject to more convective and advective thermal
losses.
A  major concern with black-and-white pyranometers is that their responsivity changes when
they are tilted. Gravity starts to affect the convective heat flow as the pyranometer is tilted. As the
pyranometer tilt increases, its responsivity decreases. Therefore, black-and-white pyranometers
114 Solar and Infrared Radiation Measurements

were not recommended for measuring irradiance on tilted surfaces. Design changes have been made
to Eppley 8-48 pyranometers, and these redesigned pyranometers are reported to have a three per-
cent tilt effect (http://www.eppleylab.com/instrument-list/black-white-pyranometer/).
The Schenk star is often difficult to level because Schenk’s bubble level is not easy to read. It is
often easier to put a level on the metal ring that holds the dome in place. The desiccant is in a con-
tainer under the pyranometer. This maneuver requires lifting the pyranometer from its platform or
VEN to check and replace the desiccant. When the desiccant is examined, the pyranometer has to
be releveled. The Eppley black-and-white pyranometer (model 8-48) has its desiccant on the front
side of the pyranometer and is therefore easier to view and change.
As with other types of pyranometers, responsivity of black-and-white type pyranometers also
changes with temperature (Figure 5.18). Temperature compensation is typically available between
−20 and +40°C.
Black-and-white pyranometers should be mounted in a VEN or with an air gap between the
instrument base and the mounting surface to maintain thermal isolation from the mounting platform.

5.3.2 laCk of tHeRmal offset


For accurate measurements, the heat flow should be between the thermopiles behind the black and
white surfaces. Other heat losses, such as from wind blowing across the pyranometer surfaces,
reduce the accuracy of the measurements. In addition, temperature-compensating components are
added to the detector circuitry to minimize the effect of nonlinear thermopile response to ambient
temperature. Conduction losses are minimized by isolating the thermopile from the other compo-
nents of the pyranometer.
From Kirchhoff’s law, as discussed in Chapter 2, a good absorber of radiation is also a good emit-
ter of radiation, and a good reflecting surface is also a poor emitting surface. At thermal equilibrium,
absorption is equal to emission. As a first guess, it may seem that the highly reflective white surface
would also be a good reflector (poor emitter) at all wavelengths, including infrared. This is not the
case because the emittance and reflectance characteristics are dependent on wavelength and the
highly reflective white paint used in the black-and-white type pyranometer becomes as good an emit-
ter at the mid-IR wavelengths as are the black surfaces. Thermal losses result from the pyranometer

FIGURE 5.18 Typical temperature response of an Eppley 8-48 (black-and-white) pyranometer.


Broadband Global Irradiance 115

dome radiating to the sky. As the dome cools, the receiver surface cools by radiating to the cooler
dome. At ambient temperatures, almost all of this radiation is in the mid-IR. Therefore, both surfaces
emit nearly the same amount of mid-IR radiation, resulting in a minimal differential thermal offset.
As discussed in Section 5.2.1, black-disk thermopile pyranometers can have offsets as large as
−20 to −40 Wm−2. This size can lead to intolerable errors if using these instruments to measure dif-
fuse irradiance that can be on the order of 100 Wm−2 or less on clear days. This radiative loss sug-
gests that black-and-white type pyranometers are superior to black-disk thermopile pyranometers
for diffuse measurements unless the latter are corrected precisely for thermal offsets. This near-zero
offset characteristic makes black and white pyranometers excellent for DHI measurements despite
their other shortcomings (Reda et al., 2002). This characteristic will be discussed more thoroughly
in Chapter 6 on high-quality diffuse measurements.

5.4 PHOTODIODE-BASED PYRANOMETERS


Photodiode-based pyranometers offer a lower-cost option for measuring solar radiation. Because
they use silicon photodiodes responsive only to wavelengths from 400 to 1,100 nm, these pyranom-
eters did not meet the old ISO-9060:1990 spectral response requirements for wavelength range from
350 to 1,500 nm or the WMO specifications for the 300–3,000 nm wavelength range. However, the
new ISO-9060:2018 specification calls for the calculation of a spectral mismatch factor (MMF)
under clear sky conditions and this MMF is used to determine the pyranometer classification.
The  MMF is determined by measuring or calculating the change in test instrument’s responsiv-
ity averaged over the 280–4,000  nm wavelength range using reference spectra at air masses 1.5
and 5. This change is compared to responsivity determined as if the instrument had no spectral
dependence. Several different reference spectra are used that simulate a variety of atmospheric
conditions. The MMF is obtained by dividing the responsivities determined at air masses 1.5 and
5 by the responsivity that would be obtained if the instrument had a totally flat spectral response.
The spectral bias is the MMF minus one (1). The largest absolute value of spectral bias is used to
determine the spectral bias classification.
Photodiodes are used to measure energy flux, and those packaged in a pyranometer are spe-
cifically designed to monitor incident solar radiation. Although some photodiodes can utilize a
transimpedance amplifier to generate the current signals or a reverse voltage, most photodiode
pyranometers do not use these techniques because they require a power source. Photodiodes mea-
sure the incident flux by monitoring the current through a load resistor in series with the photodiode
series resistor. The shunt resistor is on the order of 10–1,000 megohms (MΩ) and is much larger
than the series and load resistor combined. The voltage Vo across the load resistor Rl is then mea-
sured (see Figure 5.19). The voltage across Rl is then related to the current generated by the incident
radiation Is using Equation 5.2.

Vo =
[ I s + I l + I n ] ⋅ Rl ⋅ Rsh (5.2)
[ Rl + Rsh + Rs ]
where
Il is the leakage current
In is the noise current

Both Il and In are small. The ratio between Vo and Is is then obtained through calibration.
Conceptually this is equivalent to measuring the short-circuit current of a solar cell. As opposed
to photodiodes, the much larger area solar cells are used to convert the incident solar energy into
electrical power. Power is equal to current times voltage (P  =  I  ×  V) and photodiodes operate
near zero voltage and hence produce zero power while solar cells operate at a current and voltage
that produce the maximum power. Therefore, solar cells are operated differently than photodiodes.
116 Solar and Infrared Radiation Measurements

FIGURE 5.19 Simplified circuit for measuring current from an illuminated photodiode. Light is incident on
the photodiode and generates a current Is that is proportional to the incident radiation. The shunt resistor Rsh is
much larger than the series Rs and load resistor Rl.

Although solar cell output is related to incident solar energy, the photocurrent from a photodiode
pyranometer provides a more direct and accurate way to measure the incident irradiance. The mea-
surement accuracy of a photodiode pyranometer can be enhanced by using a fore-optic to improve
the angular response of the instrument.
Major features of photodiode pyranometers are as follows:

• Fast response
• Based on short-circuit current from solar cells (silicon photodiode)
• Output signal can be monitored either in a current or voltage mode
• Responsivity has a temperature dependence
• Responsivity depends on the spectral characteristics of the incident radiation
• Responsivity not affected by tilt of instrument
• Tilted measurements affected by spectral characteristics of sky and ground-reflected
irradiance

Understanding the characteristics of the silicon photodiode is helpful in identifying the size and
nature of the bias in measurements made using photodiode-based pyranometers and to grasp the
rationale for methods used to ameliorate the problems faced when using a photodiode pyranometer
(Stoffel and Myers 2010). The cell used in a photodiode is often enhanced to have a better response
to the blue portion of the solar spectrum than a typical solar cell.
Properly designed, photodiode output is linearly related to incident energy flux to better than 1
percent at any given wavelength. Because of this linear relationship between output and incident
energy flux and the ready availability of components, photodiodes have been used in a variety of
sensors that measure photocurrents from photometers to pyranometers.
At first glance it would seem relatively easy to obtain the incident irradiance from a pyranometer
using a photodiode. However, the external quantum efficiency of photodiodes that generate the cur-
rent is dependent on the wavelength of the incident light, and photodiodes using silicon-based solar
cells respond well to wavelengths between 450 and 900 nm but do not respond to wavelengths shorter
than ∼300 nm and longer than ∼1,100 nm (see Figure 5.20). Knowledge of the spectral response of
silicon and the spectral distribution of the incident radiation allows estimation of the spectral biases
of photodiode pyranometers. The spectral response is the amount of current that is produced divided
by the energy incident on the cell at a given wavelength. A plot of the relative spectral response of
the same reference solar cell is given in Figure 5.21. Although photons in the 450–900 nm range
produce electrons nearly 90 percent of the time, any energy above the energy needed to separate the
electron hole pair goes into heat. Therefore, the response to higher-energy photons (photons with
shorter wavelengths) does not produce a higher current. The closer the photon energy is to the energy
needed to cross the band gap, the higher the spectral response. Another way of looking at this is that a
photon at a wavelength of 450 nm will produce one electron and a photon at a wavelength at 900 nm
will produce one electron. Since the photon with a wavelength of 450 nm has twice as much energy
Broadband Global Irradiance 117

FIGURE  5.20 External quantum efficiency of a silicon solar reference cell as a function of wavelength.
Quantum efficiency is the percentage of electrons produced divided by the number of photons incident on the
solar cell. A 90 percent quantum efficiency means that 90 percent of the photons of the particular wavelength
incident on the material produce an electron. Measurements are made with voltage near zero—short-circuit
current.

FIGURE 5.21 A plot of the relative spectral response of a typical silicon reference solar cell. Data comes
from the same silicon cell used in Figure 5.20. The relative spectral response is the amount of current at a
given wavelength divided by the maximum current produced. Calibration involves determining the average
overall responsivity to the incident broadband solar radiation. Measurements are made with voltage near
zero—short-circuit current.
118 Solar and Infrared Radiation Measurements

as a photon with a wavelength of 900 nm, the responsivity per unit of energy of the 450 nm photon
will be half that of the 900 nm wavelength photon. In most cases these photodiode pyranometers use
an acrylic diffuser that limits the response to wavelengths greater than 400 nm. In addition, the dark
current (signal generated in the absence of solar radiation) and the photocurrent are affected by the
temperature of the photodiode. Therefore, the response of the photodiode pyranometer varies with
temperature. The dark current is minimized in photodiode pyranometers and most of the temperature
effect is caused by the change in the band gap characteristics related to increased kinetic energy
associated with higher temperatures. As atoms and molecules are heated, the faster they vibrate, their
kinetic energy increases along with their ability to jump into the conduction band.
Photodiode-based pyranometers are used for measuring global irradiance for clear and cloudy skies
and are calibrated under clear-sky conditions. The use of these pyranometers for measurements under
conditions where the total incident irradiance has a significantly different spectral distribution as com-
pared to clear sky irradiance is discouraged. The LI-COR model LI-200 user’s manual states that the
pyranometer should not be used under vegetation or artificial lights, in a greenhouse, or for reflected
solar radiation. Photodiode pyranometers should not be used for diffuse or direct normal solar mea-
surements without adjustments correcting the responsivity’s spectral dependence (see Chapter 10).
The photodiode-based pyranometer has two major weaknesses. The spectral distribution of inci-
dent irradiance varies over the day (i.e., with changing airmass) and under different weather condi-
tions. An example of this change is shown in Figure 5.22 where in the morning and afternoon, the
path length of the DNI though the atmosphere increases and the amount of scattering increases.
Under clear skies, the main source of scatter is Rayleigh scattering that preferentially scatters blue
(short wavelength) light. High air masses cause a significant reduction in the blue light in the DNI
and a shift in the spectral distribution to longer wavelengths. Since the peak responsivity of a photo-
diode is near 900 nm, a shift in the spectral distribution to longer wavelengths increases the average
responsivity of a photodiode. While the response to the number of photons at any given wavelength
is very linear, the photodiode response to each wavelength is different. Therefore, as the spectral
distribution of the incident radiation changes over the day, the magnitude of the current produced

FIGURE 5.22 Plot of the DNI spectral distribution at 9:00 a.m., noon, and 3:00 p.m. on January 2, 2018, at
Golden, Colorado. The sky was clear, and the morning and afternoon spectral irradiance are nearly identical.
For wavelengths greater than 900 nm, the spectral contributions for the three times are similar. There is a
considerable difference between the noontime DNI spectral values and the morning and afternoon values.
This difference results from the increased Rayleigh scattering that scatters more blue (short wavelength) light.
Broadband Global Irradiance 119

changes per given amount of incident energy flux. In most cases, it is only in the morning and eve-
ning hours when the change in spectral distribution is significant.
The second concern is that the photodiode only responds to a limited range of wavelengths. While
the photodiode’s range of response covers the wavelengths that provide approximately 75 percent
of the incident energy, the energy contained in wavelengths outside this region are not always pro-
portional to the energy in the range measured by the photodiode. The amount of perceptible water
vapor in the atmosphere affects the incident irradiance at wavelengths outside the range monitored
by the photodiode and this can affect the comparison of energy contained in the range measured
by the photodiode and the energy contained in wavelengths outside this range (Guechia et al. 2011).
Both these effects increase the uncertainty in the measurements, and for GHI measurements, these
uncertainties are in the range of class C pyranometers. A series of articles on the biases in rotat-
ing shadowband radiometers (Vignola 2006; Vignola et al. 2016, 2017, 2018; Wilber et al. 2015;
Augustyn et al. 2004;) provide further insight into the behavior of photodiode pyranometers.
When photodiode-based pyranometers are used on tilted or tracking surfaces, the spectral effects
are enhanced, especially at larger tilts and solar zenith angles. This effect is discussed in more detail
in Chapter 10.
The peak spectral response of a typical single crystalline silicon solar cell is around 900 nm.
The quantum efficiency of the solar cells is highest between 450 and 900 nm but falls off rapidly
below 450 nm and above 1,000 nm (King and Myers 1997) (see Figure 5.20).
At higher temperatures, the photodiode is more responsive (Hishikawa et al. 2018). The tem-
perature response is actually wavelength dependent with a small, negative temperature dependence
below 900 nm and a large positive temperature dependence above 900 nm resulting in a net positive
temperature response (see Figure 5.23).

FIGURE 5.23 Typical temperature dependence of the responsivity of a silicon photodiode versus wavelength.
The values determined in the plot are from quantum efficiency measurements of RCO reference cell measured
at 24 and 45°C. The RCO reference cell is a silicon cell whose output is measured at the short circuit value.
The spectral temperature response of the RCO reference cell is the same as expected from the silicon photodi-
ode used in the LI-200SA. The spectral response will vary depending on the processing; however, the general
features are driven by the effect of temperature on the ability to jump into the conduction band. Theoretically
the ratio should go to infinity as higher temperature will allow a few electrons to jump the band gap when
stimulated by longer wavelengths whereas at lower temperatures those longer wavelengths will not allow any
electrons to be promoted to the conduction band. The shape of the curve at wavelengths beyond 1,175 nm has
large uncertainties resulting from a limited number of electrons being promoted into the conduction band.
The ratio goes to infinity when zero electrons jump to the conduction band at the lower temperature while a few
reach the conduction band at higher temperatures. (Quantum efficiency measurements were done at NREL.)
120 Solar and Infrared Radiation Measurements

The temperature response of a photodiode is different than the temperature response of solar


cells when used in PV modules. Photodiodes operate at near short-circuit currents with very little
voltage drop in the circuit. Solar modules that are used to produce electricity operate at a maximum
power point that is nearer the open-circuit voltage than the zero voltage of a genuine short-circuit
voltage. The open-circuit voltage decreases significantly with temperature, and this causes the PV
module output to drop with increasing temperature. The temperature dependence of the maximum
power point output of a solar panel is not directly related to the slight increase in short-circuit cur-
rent as the temperature increases.
PV reference cells are used to compare actual PV system performance with that expected from
reference conditions (Pulfrey 1978), such as ASTM standard spectra for 1,000  Wm−2  and 25°C
(ASTM 1997). Used properly, PV reference cells should be made of the same material as the PV
modules being tested. Reference cells are made to test whether system performance is as expected
and not to measure the precise broadband irradiance incident on the PV system. Metallic current
collection grids on the surface of the cell and antireflective coatings on the protective glazing of
the PV cell also affect the cell’s output depending on the angle of incidence of the incoming irradi-
ance. PV reference cells do not make good pyranometers since they have a poor cosine response.
The cosine response of a reference cell is illustrated in Figure 5.24 that plots the ratio of a reference
cell divided by a reference GHI calculated by using an AHF cavity radiometer for the DNI and
DHI from a horizontally mounted Schenk pyranometer in a VEN shaded by a disk on an automatic
tracker. The dashed line in the Figure 5.24 is from a model for the angle of incident effect (Marion
2017). This model did not factor in the change of spectral distribution over the day.
The deviation from a Lambertian cosine response is dominated by the angle of incidence effects
associated with the light reflecting from the glazing covering the solar cell. As with photodiode
pyranometers, the shift to longer wavelengths in the spectral distribution during the early morning
or late afternoon increases responsivity of the reference cell as compared to the middle of the day.
This increase in reference cell responsivity brought about by the change in the spectral distribution
is the order of a few percent and runs counter to the larger loss of responsivity cause by the reflec-
tions (angle of incident effects).

FIGURE 5.24 Calibration of an IMT reference solar cell normalized to 1 at 45°. The reference global irradi-
ance was calculated by using a Hickey-Frieden (HF) Cavity radiometer for the reference DNI and a Schenk
star pyranometer mounted on an automatic tracker with a shade ball for the DHI irradiance. The dashed line is
a model estimate of the angle of incidence effect associated with the glazing on the reference cell.
Broadband Global Irradiance 121

5.4.1 CHaRaCteRiZing a PHotodiode PyRanometeR


The  factory specifications for the LI-COR model LI-200 (Figure  5.25) and the Kipp  & Zonen
SP Lite (Figure  5.26) pyranometers are given in Tables 5.6 and 5.7, respectively. Photodiode
pyranometers do not meet the ISO and WMO specifications as class C pyranometers because of
their limited spectral response. However, these pyranometers may perform as well as many class
C pyranometers in certain applications. In  fact, many of the specifications of these photodiode
pyranometers are significantly better than class C specifications and even meet or surpass some
of the class B specifications. For  example, the response time of a LI-COR LI-200  is extremely

FIGURE 5.25 LI-COR model LI-200 pyranometer.

FIGURE 5.26 Kipp & Zonen model SP Lite pyranometer.


122 Solar and Infrared Radiation Measurements

TABLE 5.6
LI-COR 200R Pyranometer Specifications
Characteristic Specification
Sensitivity 75 µA per 1,000 Wm −2

Stability <±2% change over a 1-year period


Response time Less than 1 µs
Temperature dependence <0.15%/°C
Cosine correction Corrected up to an 82° angle of incidence
Azimuth <±1% error over 360° at 45° elevation
Tilt No error induced from orientation
Linearity Maximum deviation of 1% up to 3,000 W/m−2
Operating temperature −40 to 65°C
Relative humidity 0 to 95% non-condensing
Calibration Calibrated against an Eppley PSP under natural daylight conditions,
typical uncertainty is ±3% within ±60° angle of incidence
Detector High-stability silicon photovoltaic detector (blue enhanced)
Sensor housing Weatherproof anodized aluminum case with acrylic diffuser and
stainless steel hardware

TABLE 5.7
Kipp & Zonen SP Lite 2 Pyranometer Specifications
Characteristic Specification
Spectral range 400 to 1,100 nm
Sensitivity 60–100 (option, 10 ± 0.5) µV/Wm−2
Sensitivity change per year <2%
Response time (95%) <500 ns
Directional error up to 80° <10 Wm-2
Temperature dependence 0.15%/°C
Operating temperature range −40 to +80°C
Non-linearity 0 to 1,000 Wm−2 <1%
Maximum solar irradiance 2,000 Wm−2
Field of view 180°

fast (approximately  10  microseconds (µs)), about five orders of magnitude faster than a thermo-
pile pyranometer. Therefore, photodiode pyranometers are often used when fast response times are
required (see RSR in Chapter 10).
Pyranometer calibration plots of an Eppley PSP, a LI-COR LI-200, and a Schenk Star are given
in Figure 5.27 and summarized in Table 5.8. The data in Figure 5.27 were taken on July 6, 2010, and
show the typical angular response characteristics of each instrument. The majority of the deviation
from a flat response is related to the angular response, but the effects of thermal offsets, temperature,
and change in spectral distribution also contribute to the shape of the ratios. The thermal offsets
can be different depending on location, but the cosine or angular response should be the same. It is
sometimes difficult to untangle the two affects. A calibration ratio is created by dividing the mea-
sured GHI of each instrument, using previous responsivities, by the reference GHI calculated from
the reference DNI and DHI. The reference DNI was measured using an AHF absolute cavity radiom-
eter, and the reference DHI was obtained from an Eppley PSP shaded by a shade ball and previously
Broadband Global Irradiance 123

FIGURE  5.27 Calibration ratio data for an Eppley PSP pyranometer, a Schenk Star pyranometer, and a
LI-COR model LI-200 pyranometer plotted against solar zenith angle. These are 1-minute data. The reference
data are the DNI from an Eppley AHF absolute cavity radiometer projected onto a horizontal surface plus the
diffuse irradiance from an Eppley PSP that has been corrected for thermal offsets.

TABLE 5.8
Relative Responsivity of Three Pyranometers Normalized to 1 at sza of 45°
Zenith Angle Std Dev Schenk Star Std Dev Eppley PSP Std Dev Li-COR Li-200 Std Dev
22.53 0.99 0.9926 0.0024 1.0154 0.0046 1.0078 0.0030
26.90 1.39 0.9954 0.0055 1.0146 0.0063 1.0082 0.0087
32.15 1.64 0.9969 0.0030 1.0116 0.0046 1.0082 0.0110
36.96 1.19 0.9958 0.0069 1.0042 0.0065 1.0072 0.0101
42.19 1.79 0.9956 0.0077 0.9985 0.0078 1.0055 0.0112
44.54 0.47 1.0000 0.0080 1.0000 0.0089 1.0000 0.0129
47.26 1.35 1.0002 0.0081 0.9957 0.0103 1.0012 0.0137
52.41 1.71 1.0003 0.0125 0.9880 0.0141 1.0042 0.0125
57.49 1.72 1.0012 0.0095 0.9816 0.0133 1.0037 0.0157
62.68 1.36 1.0031 0.0097 0.9783 0.0118 1.0095 0.0188
67.81 1.82 1.0095 0.0089 0.9728 0.0184 1.0055 0.0228
72.70 0.64 1.0105 0.0124 0.9536 0.0221 1.0136 0.0231
77.42 1.75 1.0088 0.0090 0.9277 0.0274 1.0390 0.0339
82.04 NA 0.9956 NA 0.8899 NA 1.1214 NA

calibrated using the Shade-Unshade method described in Chapter 13, Section 13.2. The thermal off-
set from the Eppley PSP was corrected by use of pyrgeometer and relative humidity measurements
(Younkin and Long 2003). Normally a Schenk Star or an Eppley model 8-48 pyranometer are used
for the DHI measurements to minimize the thermal offset issue, but in this case the Schenk Star type
pyranometer was being calibrated and no pyranometer with minimal thermal offset was available.
The responsivities have been normalized to give the calibration ratio of 1 at a 45° solar zenith angle.
124 Solar and Infrared Radiation Measurements

It is standard practice to specify the pyranometer responsivity at a 45° solar zenith angle (Myers
et al. 2002, 2004). Because the responsivity changes with solar zenith angle, only data taken within
±1 degree of 45° were used to determine the instrument’s responsivity at 45°. In this instance the
data at 44.54° were used to determine the responsivity of the instruments. Typically, the variance in
the PSP data at each solar zenith angle is less than the LI-200 and Schenk Star pyranometer because
the PSP has a large thermal mass and hence is less influenced by rapid changes in the environment.
The Schenk Star pyranometer responsivity varies more than the PSP because it has a smaller thermal
mass and only one glass dome. However, this instrument has a better cosine response than this PSP
at the station. The LI-COR LI-200 pyranometer exhibits more variance because the spectral compo-
sition of the irradiance changes, but the average irradiance measured is a good representation of the
reference values. The cosine response of the LI-200 pyranometer starts to significantly deviate from
true cosine response when the zenith angle is greater than 75°.
Figure 5.28 shows a calibration ratio plot of 15 LI-200 pyranometers. Some instruments repre-
senting the same model do perform slightly better than others, and while the shapes of the cosine
response curves are similar among the same model of instrument, they do vary. This is especially
true at large angles of incidence. The large deviation starting around 80° is related to the shape of
the diffuser and the artificial horizon on the LI-200 pyranometer.

5.4.2 Removing biases in PHotodiode PyRanometeR measuRements


The designers of photodiode pyranometers consider two factors in their construction (Kerr et al. 1967):

1. The  diffuser of the pyranometer must be designed to produce, as close as possible, a


Lambertian response to the incident solar radiation. With a Lambertian response, the inci-
dent energy from a point source is proportional to the intensity of the source times the
cosine of the incident angle.
2. The photodiode selected for the pyranometer needs to have a broad spectral response that
yields an average responsivity that doesn’t change significantly as the spectral distribution
of incident irradiance changes throughout the day.

FIGURE 5.28 Average calibration ratio plots of 15 LI-200 pyranometers that have been normalized so that
the ratio is 1 at a zenith angle of 45°. The oldest pyranometer show the best cosine fit at large angles.
Broadband Global Irradiance 125

FIGURE 5.29 Plot of the ratio of a LI-COR LI-200 pyranometer divided by GHI calculated from an Eppley
NIP for DNI and a Schenk Star pyranometer for the DHI against time. Values are 1-minute averaged data.
Clear days in July 2010 were selected for the plot.

The typical responsivity of the LI-COR LI-200 pyranometer is fairly flat across the day and shows
sharp spikes in early morning and late afternoon hours (see Figure  5.29). Only clear days were
selected for the plot. For LI-200 pyranometers, two peaks in Figure 5.29 are observed. These angles
correspond to large solar zenith angles around 82.5° in the morning and the afternoon. The plot
reminded Augustyn et al. (2002) of “cat ears” when he was developing a model to correct the cosine
response of the LI-200 pyranometer. The cause of the cat ears effect is the shape of the diffuser that
improves the uniformity of the responsivity over much of the day (Augustyn et al. 2004). As with
all pyranometers, modifications to correct for one limitation often lead to another limitation in a
different situation. Therefore, it is always useful to know how the instrument performs under a wide
variety of circumstances.
There is a distinct slope to the responsivity over the day with the ratio lower in the morning than
in the afternoon. Temperature dependence of the pyranometer is likely to be partially responsible
for some of this apparent slope. Multiplying the LI-200 pyranometer GHI values by a temperature
adjustment factor of

1 − 0.0035 ⋅ ( temperature − 25°C) (5.3)

results in a more symmetric responsivity ratio over the day (see Figure 5.30). The temperature used
in the equation is the ambient air temperature in Celsius. An ambient temperature of 25°C was cho-
sen as the reference temperature because it is the standard temperature used when specifying solar
cell performance. The temperature data in Figure 5.30 range from 10 to 35°C. For 3 days during the
month, shown as gray circles in Figure 5.30, the ratios were from a tight fit that has a slightly lower
ratio than the rest of the clear sky periods in the month. There was no indication as to why the ratio
for those 3 days were 2 to 3 percent below the ratios for the other clear sky periods. The probable
explanation is a different aerosol or water vapor content in the atmosphere. This issue shows the
difficulty in getting consistent photodiode calibration values.
These temperature adjustments are opposite in sign and are two to four times larger than exist
for silicon solar modules. Characteristically, the short-circuit current of a solar cell increases
with temperature while the power output of a solar panel decreases with increasing temperature.
126 Solar and Infrared Radiation Measurements

FIGURE 5.30 Plot of the ratio of a LI-COR LI-200 pyranometer divided by GHI calculated from an Eppley
NIP for DNI and a Schenk Star pyranometer for the DHI against time. The LI-200 GHI values have been mul-
tiplied by 1−0.0035* (Temperature−25°C). Values are 1-minute averaged data. Clear days in July 2010 were
selected for the plot. Grey circles are from 3 days during the near the beginning of the month.

Solar panels operate at voltages closer to the open circuit voltage to output the maximum power
and not close to the zero voltage of a short-circuit current. The operating voltage of a solar mod-
ule decreases with increase in temperature. Other factors may mimic this assumed temperature
dependence (Kerr et al. 1967). For example, the optical alignment of the detector could be partially
responsible for the asymmetry seen in Figure 5.29 (Michalsky et al. 1995). In addition, for a photo-
diode, the responsivity of wavelengths closer to the band-gap of the solar cell have a larger response
to temperature increase than shorter, more energetic wavelengths (see Figure 5.23).
Many other factors can contribute to the shape and distribution seen in Figure 5.29, and care
should be taken to untangle spectral change effects from cosine and temperature response effects.
It  is important to theoretically and experimentally identify the magnitude and direction of such
effects on measurement performance before claiming that results show the physical effects of a
specific variable. Before systematic biases can be removed with confidence, it has to be verified for
a wide variety of measurement situations and at a number of different geographic locations.
For solar cells, the temperature coefficient introduces a bias of a few percent in measurements
made in the 0 to 40°C range. The  temperature coefficient for short-circuit current from silicon
solar cells varies from −0.0004/°C to around 0.0015/°C when measured at different wavelengths
(see Figure 5.23). A more detailed examination shows that the temperature coefficient for a silicon
photodiode is positive in the wavelengths between 850 and 1,100 nm and zero or slightly negative
for wavelengths below 850 nm (Kerr et al. 1967). To minimize the uncertainty associated with tem-
perature dependence, some photodiode pyranometers are heated to maintain a constant (higher than
ambient) temperature, while others measure the pyranometer temperature and apply a correction
algorithm to account for the change in responsivity with temperature. Some of the least expensive
photodiode pyranometers ignore the instrument calibration change with temperature.
Pyranometers are calibrated during clear days and produce results such as shown in Figure 5.27.
The measured irradiance under cloudy conditions compares favorably to measurements made with
higher precision pyranometers. Clouds scatter all solar wavelengths nearly equally since the cloud
particles are large compared with the solar wavelengths (geometric scattering limit).
Black-disk thermopile pyranometers such as an Eppley PSP have a fairly uniform spectral
response over the wavelengths of interest. However, photodiode pyranometers, such as the LI-COR
Broadband Global Irradiance 127

FIGURE 5.31 Percent relative response of a LI-COR LI-200 pyranometer plotted against wavelength. Data
taken at NREL using a reference solar lamp. Peak responsivity of the LI-200 pyranometer is to wavelengths
between 900 and 1,000 nm.

LI-200 and Kipp & Zonen SP Lite, have a marked spectral response to the incident solar radiation.
As shown in Figure 5.31, the LI-200 has the maximum sensitivity to wavelengths near 900 nm and
does not respond to solar radiation with wavelengths over 1,100 nm or under 400 nm. The effects
of the acrylic diffuser on the LI-200 responsivity can be seen by comparing Figure  5.31 with
Figure 5.21, which shows the responsivity for a silicon solar cell without an acrylic diffuser.
King et al. (1997) provide a careful analysis of the spectral responsivity of the LI-200 pyranom-
eter. They examined the responsivity under clear skies when approximately 90 percent of the global
irradiance came from DNI. The  responsivity of the photodiode-based pyranometer to clear-sky
DHI is significantly different from the responsivity for the DNI (Michalsky et al. 1991; King et al.
1998; Vignola 1999). The ratio of GHI irradiance from a photodiode to reference values is near one
(see Figure 5.30) and the ratio of DHI from a photodiode to a reference DHI measurement is about
70 percent (see Figure 5.32). Under clear skies, the responsivity of the photodiode to the blue-sky
dome (the DHI component) is about 70 percent of the responsivity determined for GHI or DNI.
The photodiode pyranometers are not as responsive to the blue portion of the solar spectrum as
seen in Figures 5.21 and 5.31, and when the diffuse spectrum consists largely of blue light, such
as on a clear day, the photodiode will underreport the DHI. The DHI responsivity also changes as
the solar zenith angle increases and the diffuse spectrum shifts toward red. The responsivity of the
LI-200 to DHI under clear skies increases as the DHI spectral distribution reddens with increasing
air mass. At a solar zenith angle of 85° the difference between the GHI and DHI responsivity is half
the difference that is observed at 45° (Figure 5.32). Some of the spread in the responsivity ratios
in Figures 5.29 and 5.30 results from different atmospheric composition and hence differences in
the spectral characteristics of the incident radiation. The spread of the data in Figure 5.32 is related
to the difference in spectral distributions of the incident radiation. While models do exist that can
mimic the bias shown in Figure 5.32, the uncertainty around this average still exists.
On clear days, the spectral characteristics of both the DNI and DHI are different and their
spectral distribution changes over the day. The  DNI spectrum shifts toward the IR as the path
length through the atmosphere increases and more of the blue portion of the spectrum is scat-
tered from the DNI. The DHI on clear days contains a considerable amount of blue light caused
128 Solar and Infrared Radiation Measurements

FIGURE  5.32 Comparison of diffuse measurements from LI-200 pyranometer in a rotating shadowband
pyranometer and a Schenk Star pyranometer with a shade ball. Five clear summer days were chosen for the
plot. The LI-200 diffuse measurements are around 30 percent below the measurements of broadband thermo-
pile pyranometers. As the spectral distribution of the incident radiation shifts toward the near infrared por-
tion of the spectrum as the sunlight has a longer path length through the atmosphere, the responsivity of the
photodiode-based pyranometer increases.

by Rayleigh scattering. Near sunrise or sunset, both DNI and DHI are redder as the blue light has
been preferentially removed by Rayleigh scattering. This removal results in the reddish color of the
sun and sky in the morning and evening. This red shift in the morning and evening hours occurs
for DHI on cloudy as well as clear days because clouds are neutral attenuators.
On clear days it is difficult to separate angular response problems from spectral response changes.
This can be seen in Figure 5.29 where the cat ears effects dominate the change in responsivity at
larger solar zenith angles. On cloudy days, the scattering by clouds provides a more uniform distri-
bution of irradiance, and the angular response does not change over the day. Figure 5.33 plots the
ratio of DHI from a LI-COR LI-200 pyranometer divided by the DHI from the Schenk pyranom-
eter on cloudy days. This ratio isolates the effects of changes in spectral distribution on the LI-200
responsivity. The  angular responses of both instruments affect the ratio, but the effect remains
approximately constant over the day because the radiation is coming from all parts of the sky. Note
that the responsivities of the instruments for cloudy periods are about the same during most of the
day because the DHI spectrum that the LI-COR LI-200 pyranometer sees is much like the GHI
spectrum because the clouds act as neutral filters scattering all wavelengths equally. However, as
the solar zenith angle increases, the total spectrum shifts toward red and becomes similar to the
spectrum that the LI-200 sees on clear days near sunrise and sunset; therefore, the responsivity of
the LI-200 decreases to about 85 percent at 85°, the same value as in Figure 5.32. If the photodiode
pyranometer is to be used for DHI measurements, large systematic biases will occur unless adjust-
ments are made to remove the biases brought about by the changes in the spectral distribution and
the resulting changes in pyranometer responsivity.
The response of the photodiode pyranometer to clear-sky diffuse is about 30 percent lower than
a thermopile pyranometer. Under cloudy skies, the spectral distribution is similar to the GHI solar
spectrum because the clouds are neutral scatters. Therefore, the photodiode pyranometer measure-
ment of cloudy sky diffuse is comparable to thermopile pyranometer measurements.
Broadband Global Irradiance 129

FIGURE 5.33 Plot of the responsivity ratio of a LI-200 pyranometer as a function of zenith angle for periods
in July 2010 when the DNI less than or equal to 1 Wm−2. Data are 1-minute values.

The  shade–unshade and summation calibration techniques will be discussed in detail in


Section 5.5, but there is a systematic difference in responsivities determined for photodiode pyranom-
eters using the two techniques resulting from the different DNI and DHI spectra. For the shade-
unshade calibration, only the DNI spectra affect the calibration as the DHI contribution is canceled
by the technique. The calibration of the photodiode pyranometer by summation technique is margin-
ally affected by the lower diffuse responsivity since the diffuse component may make up only about
10 percent of the total irradiance on a clear day. A 30 percent error in a 10 percent component results
is only a 3 percent factor, which is less than the typical uncertainty of ±5 percent associated with the
calibration of a photodiode pyranometer. This calibration assumes very clear sky conditions when
DHI is about 10 percent of the GHI. The poorer response of the photodiode pyranometer to clear-
day DHI results in the responsivity determined by the summation method being about 3 percent
lower than the responsivity obtained by a shade–unshade calibration. For calibrations done on days
with higher DHI values, the summation method will yield even lower responsivities.
It is recommended that a photodiode pyranometer be calibrated by the summation technique as
this is similar to the way the pyranometers are used in the field for GHI measurements. For cali-
brations where it is not possible to use the summation technique, a side-by-side comparison can be
made near a solar zenith angle of 45° with a reference photodiode pyranometer of the same model
that has been calibrated using the summation technique. The estimated measurement uncertainties
associated with side-by-side pyranometer comparisons are larger than methods referenced directly
to reference pyrheliometer and DHI pyranometer measurements.
Pyranometers operate under all weather conditions, unlike the clear-sky conditions used to cali-
brate the pyranometers. The absolute uncertainty of a photodiode pyranometer is often quoted as
±5 percent. In the field, the measurement uncertainty of these devices depends on many condi-
tions, such as air temperature, time of day (solar zenith and azimuth angles), irradiance levels,
spectral composition, and soiling or moisture on the diffuser. The temperature response of the LI-200
pyranometer under all sky conditions is illustrated in Figure 5.34. The responsivity of the LI-200
pyranometer in this example appears to increase by 0.3 percent per degree Celsius; however, the tem-
perature dependence of the reference instruments has not been corrected. Other environmental factors also
130 Solar and Infrared Radiation Measurements

FIGURE  5.34 Ratio of LI-200  pyranometer readings divided by reference global values plotted against
temperature. The ratio increases at a rate of approximate 0.3 percent per degree Celsius.

may have influenced the distribution of the data in Figure 5.34. If the temperature dependence of the
pyranometer has been accurately determined, these systematic biases may be removed. A multivari-
ate correlation is needed to more accurately analyze the dependence of the responsivity on a vari-
able because some factors can mimic changes seen in other variables. For example, the temperature
dependence of a photodiode varies with the spectral distribution of the incident radiation. Also, other
issues such as the temperature dependence of the reference instruments need to be considered.
Figure  5.35 contains a plot of global irradiance measured by a LI-COR LI-200 pyranometer
divided by the reference global irradiance plotted against the clearness index for May 2010 in
Eugene, Oregon. The clearness index (kt) is computed as the ratio of measured GHI to the corre-
sponding extraterrestrial value. The points with the low response ratios around clearness index of
0.2 occur mostly at zenith angles above 70° and probably represent the same spectral phenomena
as shown in Figure 5.33.

5.4.3 RefeRenCe solaR Cells


Reference solar cells are sometimes used to check the relative performance of PV systems. Reference
solar cells are specially packaged and characterized according to standard practices (ASTM 2008,
2009, 2010; IEC 2007), have a high-quality calibration history, and are used to check performance
of PV modules made of similar materials when tested in solar simulators. In addition, reference
cells are commonly used to adjust and monitor the intensity of solar simulators. The reference solar
cells are designed to mimic the performance of PV modules and include some elements such as
glazing that are included in the PV module but generate biases when compared to pyranometers
used to measure incident radiation. Reference cells are not designed to measure incident broadband
solar radiation consistent with the ISO or WMO performance specifications and as such are
not pyranometers. Reference cells are useful for checking the performance of PV modules relative
to standard test conditions (STC) and evaluating the degradation of performance over time, as long
as periodic calibrations of the reference cell are well maintained.
Reference cells are normalized to standard operating conditions (an ambient temperature of
25°C, global irradiance of 1,000 Wm−2, air mass of 1.5, and no wind). The output of one reference
Broadband Global Irradiance 131

FIGURE 5.35 Plot of ratio of LI-200 pyranometer data to reference global irradiance plotted against zenith
angle. One-minute data from May 2010 were used in the plot. May was a particularly cloudy month. Note that
under cloudy conditions (clearness index <0.4) and zenith angles greater than 70° the LI-200 pyranometer
deviates by more than 10 percent from reference measurements. Reflection off clouds accounts for the high
number of data points with a clearness index above 0.8.

cell is not comparable to a reference cell made with a different cell technology, although they should
compare well to identical solar cells in identical packaging. Reference cells often have the capabil-
ity to determine the cell’s temperature and adjust their measurements to the reference temperature
by formulas that depend on the average change in reference cell output at various temperatures.
This adjustment is necessary when reference cells are used to evaluate PV modules because the ref-
erence cell measurements based on short circuit currents increase output as temperature increases
and PV modules decrease in output as temperature decreases because the modules are operated
for optimum power output. Because reference solar cells are behind a flat glazing, they suffer from
a non-Lambertian response similar to solar modules (Figure 5.24). Pyranometers are designed to
measure the total solar spectrum from 300 to 3,000 nm under all weather conditions. Therefore, a
reference solar cell is not a pyranometer and should not be treated as one.

5.5 CALIBRATION OF PYRANOMETERS


Proper calibrations depend on traceability to an internationally accepted measurement standard.
A critical aspect of pyranometer calibrations is its traceability to the World Radiometric Reference
(WRR). The  WRR is the internationally recognized measurement reference for solar irradiance
(Fröhlich 1991). A detector-based reference, the WRR is realized by measurements from the World
Standard Group (WSG) of electrically self-calibrating absolute cavity radiometers (ACRs) limited
to DNI above 700 Wm−2 under clear-sky conditions. Another important consideration is to calibrate
an instrument in the same environment it will be used for routine measurements. Although this con-
cept may not always lead to practical calibration methods, the estimated measurement uncertainties
are generally more applicable.
There  are two accurate outdoor methods used to calibrate pyranometers. The  shade–unshade
method uses a DNI measurement with alternate shading and unshading of the pyranometer under
calibration (see Chapter 13, Section 13.2). The summation method compares an unshaded pyranom-
eter output signal (in µV) with the reference GHI calculated from the reference DNI measurement
132 Solar and Infrared Radiation Measurements

times the cosine of the solar zenith angle plus the reference DHI (Equation 5.1). Standards for cali-
brations are found in ASTM (2015, 2017) and ISO (1993). The estimated measurement uncertainty
in the pyranometer calibrations is larger than the uncertainty in pyrheliometer calibrations because
there is no absolute pyranometer available to obtain a reference GHI measurement. In addition, the
responsivity of pyranometers is dependent on additional factors resulting from the 180 degree field
of view. The most noticeable problem with pyranometers is the lack of a true Lambertian angular
response (true cosine response) over the range of solar zenith angles experienced during the day.
Examples of this angular response issue for a thermopile pyranometers is shown in Figure 5.27.

5.5.1 sHade–unsHade CalibRation metHod


The  most accurate way to calibrate a pyranometer is to use the shade–unshade method (ASTM
1997, 2015; ISO 1993). This is the preferred method to establish reference pyranometers calibration
for use in side by side comparison calibrations or calibrations of pyranometers used for measuring
DHI (see Chapter  13, Section  13.2 for an overview on broadband shortwave radiometer calibra-
tions). When using the shade-unshade calibration, the responsivity (µV/Wm−2) of the pyranometer is
determined by subtracting the voltage when the horizontally-mounted pyranometer is shaded from
the voltage when the pyranometer is unshaded and dividing this difference by the reference DNI
projected onto the horizontal (DrHI).

(Unshaded Voltage − Shaded Voltage)


R = (5.4)
( DNI ⋅ cos ( sza ) )
The accuracy of the shade–unshade method is dependent on the accuracy to which the DrHI is
determined in the denominator. Either a high-quality thermopile pyrheliometer or, preferably, an
absolute cavity radiometer can be used for the DNI value. A cavity radiometer can measure the
DNI to an accuracy of better than 0.5 percent, whereas a thermopile pyrheliometer has an absolute
accuracy of between 0.7 and 1.7 percent (Michalsky et al. 2011). The field of view of a cavity radi-
ometer is five degrees, while that of some older thermopile pyrheliometers is 5.7 degrees. A five
degree field of view is now standard for modern pyrheliometers (WMO, 2008). A shade disk or
sphere has to be built that will shade the pyranometer such that the field of view blocked by the disk
is the same as the field of view of the instrument used to obtain the DNI value. For example, a disk
with a 4.97 centimeter (cm) radius would be equivalent to the 5.7 degree field of view for the disk
if it was 1 m away. If the disk was only 50 cm away, the radius needed would be 2.49 cm. For a 5.0
degree field of view the radius for a distance of 1 m would be 4.36 cm, and for a distance of 50 cm
the radius would be 2.18 cm. The size of the disk should be such that it shades the dome of the
pyranometer and, hence, prevents reflections from the dome impinging on the receiver. The disk
should be painted black, using an ultra-flat black paint such as Rust-Oleum Black to coat the disk
and the supports. For  information on standards for the shade–unshade calibrations, see ASTM
(1997) and McArthur (2004).
A  plot of two-second data from a shade-unshade calibration of a Kipp  & Zonen CMP 22 is
shown in Figure 5.36. Included in the figure are DfHI data from a Schenk pyranometer and DrHI
from a Kipp & Zonen CHP1 pyrheliometer. When performing a shade–unshade calibration, it is
necessary to wait until the pyranometer reaches equilibrium when the instrument is shaded and
when it is unshaded. For  photodiode pyranometers, the response time is ∼10  µs; therefore, the
shade–unshade intervals are very short. For a thermopile pyranometer, the response time may be
several seconds depending on the design and thermal mass of the pyranometer. Response time is
the amount of time it takes for the instrument to respond to a step change in irradiance. In this
chapter, response time is the time it takes to be within 1/e (~37%) of the new value. Ideally, the
reading should be the true value or the equilibrium reading. During a shade–unshade calibration
Broadband Global Irradiance 133

FIGURE  5.36 Shade-unshade calibration of a Kipp & Zonen CMP 22 pyranometer, data obtained every
2 seconds. The dashed line is direct horizontal irradiance from a CHP 1 pyranometer, and the dotted line is
the reference DHI from a Schenk pyranometer. DrHI from an ACR taken at 20 second intervals is not shown.
A  shade ball was placed in front of the pyranometer every 2 minutes and then removed 2 minutes later.
The CMP 22 pyranometer reached equilibrium after about 20 seconds.

clear-sky conditions are required. This requirement is necessary because the shade and unshaded
measurements are taken at different times. Under clear-sky conditions, the DHI values are fairly
constant and can be interpolated between the two shaded values to obtain the reference DHI values
while the instrument is unshaded. There is an order of magnitude change in irradiance from GHI
to DHI measurements. It takes between 20 and 60 times the 1/e response time for consistent read-
ings (Reda et al. 2003). For some instruments it takes three times longer to achieve an equilibrium
state when the shaded pyranometer is unshaded and exposed to full Sun than when the unshaded
pyranometer is shaded for the calibration. At full Sun (e.g., DNI = 1,000 Wm−2), a reading taken
after 60 times the 1/e response time is representative of the true GHI. During the shaded period, it
takes 20-time constants for the readings to reach a consistent DHI level. Some pyranometers, such
as a Schenk star pyranometer, have a very slow response time that can exceed five minutes and
it is recommended that the summation technique be used on these instruments. A more detailed
examination of the reference DHI in the shade-unshade calibration for a CMP 22 pyranometer is
shown in Figures 5.37 and 5.38. When the pyranometer is first shaded, its readings are below the
reference DHI values. Gradually after about 20 seconds, the ratio of the DHI irradiance from the
instrument under calibration approaches a constant. This effect appears in two-second data and
would essentially disappear in 20-second data because the instrument reaches equilibrium by
that time. A similar but opposite effect exists when the pyranometer is unshaded (Figure 5.38).
The GHI from the pyranometer under calibration is plotted against GHI calculated using the sum-
mation technique and scaled to illustrate the response time. Although the unshaded pyranometer
initially has increased output, it settles down to the reference GHI values after about 20 seconds.
The absolute cavity radiometer data are not shown in the plot because the data were taken every 20
seconds. The response time is related to the thermal mass, thermal conductivity, and other design
and construction elements of the pyranometer. Each style of pyranometer will have its own unique
response time characteristics.
134 Solar and Infrared Radiation Measurements

FIGURE  5.37 Detail of a shade-unshade calibration when the pyranometer is shaded. A  shadeball is used
to shade the pyranometer every 2 minutes. It is then removed after 2 minutes. The solid line is the 2 second
data from the CMP 22 pyranometer and the dotted line is the reference DHI data from a Schenk pyranometer.
The CMP 22 drops below the reference DHI value when it is first shaded and slowly climbs to be about 75 W/m2
above the reference DfHI after about 20 seconds. The DHI values track after thermal equilibrium is reached.

FIGURE 5.38 Detail of a shade-unshade calibration when the pyranometer is unshaded. In this example, the
pyranometer is unshaded every 2 minutes. The solid line is the 2-second data from the CMP 22 pyranometer
and the dotted line is the sum of the DrHI from a CHP1 pyrheliometer and DHI from a Schenk pyranometer.
The ACR used in the calibration was not plotted because its values were recorded every 20 seconds. When
first unshaded, the CMP 22 jumps above the reference GHI value and slowly decreases to the reference GHI
value after about 20 seconds. The GHI values track after thermal equilibrium is reached. The two scales are
matched to show the time needed to return to equilibrium. The scale on the left is the output of the CMP 22
pyranometer and the scale on the right is the GHI calculated from the sum of the DfHI and DrHI.
Broadband Global Irradiance 135

As in Figure 5.37, the best way to determine when the signal has stabilized is to plot the out-
put voltage from the pyranometer when it is shaded and unshaded. Identifying this stable period
requires sound judgment and experience to separate voltage creep (or overshoot) due to true irradi-
ance change from instrument stabilization. The data used in determining the responsivity are taken
when it is judged that equilibrium has been attained.
When performing a shade–unshade calibration, the timing of the measurements is important to
accurately compute the solar zenith angle representative of the two reference irradiance values used
for the calibration. The ambient air temperature also should be measured because thermopile- and
photodiode-based pyranometer readings have a known dependence on temperature. Shade–unshade
calibrations should be made only when the skies are clear of clouds and the DNI readings are stable
(e.g., changing by less than 3 Wm−2 during a 1-minute interval). Ideally, the measurements for estab-
lishing a reference pyranometer calibration should be repeated over multiple days throughout a year
to evaluate the pyranometer’s change in responsivity with different combinations of solar zenith and
azimuth angles. In lieu of this challenging requirement it is often better to fit a curve (responsivity as
a function of solar zenith angle or air mass) over a few days of calibration data that average random
variations, which makes it easier to detect and account for biases in the pyranometer measurements.
One significant advantage of the shade–unshade method is the elimination of the thermal offsets
(Section 5.2). Thermopile pyranometers emit IR radiation to the sky, resulting in reduced voltage
readings. The  thermal offsets during shaded and unshaded measurements are nearly identical. By
subtracting the shaded from the unshaded voltage, the offset caused by the thermal offsets is elimi-
nated. Therefore, the responsivity determined in Equation 5.4 is unaffected by the thermal offset.
Of course, when the pyranometer is used in the field, it is subject to the thermal offsets, and these
offsets need to be corrected or the irradiance is underestimated. If the summation method, discussed
next, is to be used, best practices recommend that the diffuse measurements be made with a reference
pyranometer that has minimal thermal offset and has been calibrated using the shade–unshade method.

5.5.2 summation metHod CalibRation


A less tedious and more efficient method for calibrating multiple pyranometers simultaneously is
to compare the voltage output from the pyranometers under test with a reference global irradiance
determined by a DNI measurement made with an ACR, ISO 9060 class AA, or a class A thermo-
pile pyrheliometer and a DHI measurement made with a shade–unshade-calibrated and offset-
corrected pyranometer mounted under a shade disk or sphere as described in the previous section.
The DNI value is projected onto a horizontal surface by multiplying this irradiance by the cosine
of the solar zenith angle to get DrHI, and the reference GHI value is obtained by adding the DHI
value to the DrHI value (Myers el al. 2002, 2004; Myers, 2005; Reda et  al. 2003; Wilcox et  al.
2002). This method can be automated for the simultaneous calibration of multiple pyranometers
with responsivity values obtained throughout the day (Dooraghi et al. 2019). Ideally, this should
be performed near the summer solstice to allow for the widest range of solar zenith angles, but
calibrations can be obtained as long as the solar zenith angle at solar noon is less than 45°. The key
to the accuracy of the summation calibration method is to determine the reference GHI value using
a DHI value from a pyranometer that has minimal or correctable thermal offset. By design, black-
and-white type thermopile pyranometers have minimal thermal offsets. There also are single black
thermopile pyranometers that exhibit minimal or correctable thermal offsets (See Appendix G).
For  consistent responsivity results, pyranometer calibrations are performed on clear days with
minimal variations in atmospheric transmission (i.e., aerosol optical depth). There  are many ways
to specify a stable atmosphere. Instruments with fast response times, like photodiode pyranometers,
can be used to check for rapid variations in the GHI due to variations in atmospheric water vapor,
aerosols, and thin (subvisual) clouds, or the variability of the DNI and DHI irradiance over sub
1-minute intervals can be used to select the best calibration periods (Reda et al. 2003; Wilcox et al.
2002; Dooraghi et al. 2019).
136 Solar and Infrared Radiation Measurements

For a comprehensive determination of responsivity, the calibrations should be made over a full
day or parts of 2 days and include a wide range of solar zenith and azimuth angles. To understand
the change in responsivity of a pyranometer over time, it is important to measure the environmental
parameters that influence the instrument responsivity. (See the broadband outdoor radiation cali-
bration (BORCAL) certificate in Appendix H for examples of these parameters.) Specifically, in
addition to the irradiance measurements, concurrent ambient temperature, relative humidity, wind
speed and direction, and barometric pressure should be recorded (Dooraghi et al. 2019). If an initial
responsivity is obtained when the ambient temperature is 10°C and an additional responsivity is
obtained when the ambient temperature is 25°C, some of the difference between the two measured
responsivities may be related to the change in temperature. If it is known how the responsivity var-
ies with temperature, the change in responsivity over time can be better estimated by adjusting the
responsivities to a standard temperature. Influence of wind speed, downwelling infrared, spectral
distribution of the incident radiation, and other factors influence the responsivity of the pyranom-
eter. It is not straightforward to separate the influence of these factors on the responsivity because
the effects are often small, and they may be inseparable. For example, wind speed affects the tem-
perature difference between the dome and body of the pyranometer, and relative humidity affects
the thermal offsets and the advective efficiency of wind on heat transfer. Therefore, calibrations
should be performed over a wide variety of meteorological conditions.
A  single responsivity can be assigned to a pyranometer for all measurements, or, for more
accuracy, a responsivity that is a function of the primary variables that affect the responsivity can
be used. A prime example of the latter is a pyranometer responsivity that varies with solar zenith
angle. Generally, a single responsivity of the instrument, near 45°, is used to specify the calibra-
tion for irradiance measurements. An example of a pyranometer responsivity changing over a day
is given in Figure 5.39, and the responsivity plotted against the solar zenith angle is illustrated in
Figure 5.40.

FIGURE  5.39 Plot of 1-minute responsivity data against time of day. An ideal pyranometer would have
a constant responsivity over the day. Often the deviation from a flat responsivity is attributed to imperfect
Lambertian (or cosine) response. Note: thermal offsets and other factors also play a role in the deviation from
a true cosine response.
Broadband Global Irradiance 137

FIGURE 5.40 Same data as in Figure 5.39 with the responsivity plotted against zenith angle. A true cosine
response would yield a straight line. Note that there is an asymmetry between the morning values and the after-
noon values. Leveling and other factors such as changes in the azimuth response of the pyranometer and tem-
perature effects on the responsivity, and net IR influences on thermal offsets can contribute to this asymmetry.

5.6 PYRANOMETER CALIBRATION UNCERTAINTIES


As discussed in Chapter 4, the Guide to the Expression of Uncertainty in Measurements (GUM;
JCGM 2008) presents the GUM protocols that specify a uniform manner for obtaining uncertainties
(see Appendix A for a more detailed discussion). With a set of measurements, the mean and stan-
dard deviation of the data set can be determined with respect to a reference value. This information
provides only a subset of the uncertainties. Many other factors contribute to the uncertainty such as
the calibration of reference instruments and accuracy of the measured radiometer output voltages.
The  GUM protocols help classify the uncertainties and prescribe a methodology for combining
the uncertainties to give an overall uncertainty of the measured values. Estimated measurement
uncertainty is a very important part of understanding and applying radiometer data. Uncertainties
as applied to the calibration and use of pyranometers are covered in this section (the GUM pro-
cedure was applied in Chapter 4 to pyrheliometer measurements). Uncertainty determinations for
pyranometers are more complex and will be approached from a slightly different perspective than
covered in Chapter 4. A brief review of the GUM procedure will be followed by examples of uncer-
tainties calculations taken from the National Renewable Energy Laboratory (NREL) calibration
procedures for broadband radiometer calibration (Myers et al. 2004; Reda et al. 2008; Reda 2011).
The GUM procedure can be summarized in six steps:

1. Determine the measurement equation.


2. Identify the sources of uncertainty
3. Calculate or estimate the standard uncertainty for each variable in the measurement
equation.
4. Calculate the sensitivity coefficients.
5. Calculate the combined standard uncertainty (uC ).
6. Calculate the expanded uncertainty (UE ) by multiplying the combined standard uncer-
tainty by the coverage factor (k), typically by applying student’s t-distribution factors for
95 or 99 percent.
138 Solar and Infrared Radiation Measurements

There are two types of uncertainties in a measurement. Type A uncertainties are those that can be
calculated from a series of observations (labeled u A). Type A analysis is statistical in nature and
determines the random errors associated with any measurement. Type B uncertainties are all the
other uncertainties that are not subject to statistical analysis of the observations (labeled uB). Type B
uncertainties are associated with a variety of effects that can affect the measurements and add uncer-
tainty. This type of uncertainty includes the meteorological and situational influences that affect
the measurement, systematic biases caused by the data logger used for the voltage measurement,
and uncertainties in the calibration equipment and methodology used to calibrate the pyranometer.
When discussing a measurement, it is common to use the term “error” to describe the difference
between the true value of the measurand and the measured value. In practice the true value of the
measurand is only approximated by some reference. There is the uncertainty in the measurement
and a difference between the reference value and the measurement. Type A uncertainties arise from
the nature of the measurement and the instruments used. They form a distribution. This distribution
is not an error in the common sense of the term but characteristic of the measurement. Some instru-
ments have systematic biases, such as a pyranometer’s deviation from a true Lambertian response.
These systematic biases are often referred to as systematic errors. In general, systematic errors and
systematic biases are referring to the same issue; however, the GUM methodology uses systematic
uncertainties or biases when referring to these issues.
GUM (JCGM 2008) treats the two uncertainty types separately and then determines a combined
uncertainty (uC ). Uncertainties with a small “u” refer to uncertainties that are within 1 standard
deviation (a 68% confidence level), while uncertainties labeled with a capital “U” refer to uncertain-
ties at the 95 percent level of confidence. For uncertainties that are from a Gaussian (or “normal”)
distribution with a large number of data points, 68 percent of the data points lie within 1 stan-
dard deviation of the mean, while 95 percent of the data points lie within 1.96 times the standard
deviation (see Table  5.9 for the relationship between multiples of the standard deviation and the
confidence levels for a Gaussian distribution). The ustd for Type B uncertainties is the estimate of
an equivalent standard deviation (of a specified distribution) of the source of the bias (Stoffel et al.
2010). The combined uncertainty (uC ) is computed from the Type A and Type B standard uncertain-
ties summed in quadrature. When terms are added in quadrature, the sum of the square of the terms
is calculated, and the square root of the total is the combined standard uncertainty of all terms. It is
important to use the same confidence level for all the uncertainty terms. If one term is assigned at 68
percent confidence level (1 standard deviation), then all terms should be referenced to the 68 percent
confidence level. When the square root of the sum of the squares is determined, then the confidence
level can be expanded to the desired level. In the GUM methodology, the expanded uncertainty, UE,
is determined from the combined uncertainty uC determined at the 68 percent level of confidence
and multiplied by a coverage factor k to obtain the desired level of confidence:

∑u + ∑u
1/ 2
UE = k ⋅  B
2
A
2  (5.5)
 

TABLE 5.9
Standard Deviation and Confidence Level for a
Gaussian Distribution
Number of Standard Deviations Confidence Level (%)
1σ 68.27
1.96σ 95.0
2σ 95.45
3σ 99.73
4σ 99.99
Broadband Global Irradiance 139

For a large enough statistical sample, a coverage factor of 1.96 is used to obtain a 95 percent level
of confidence or a coverage factor of 2.58 is used to obtain a 99 percent confidence level. This all-
inclusive uncertainty is the expanded uncertainty (UE ).
To be more precise, the combined standard uncertainty for a measurement is obtained from
the square root of the sum of the variances of the variables that affect the measurement, weighted
according to how the measurement results change with changes in these quantities (that is, by sensi-
tivity coefficients). These sensitivity coefficients are explained in the example uncertainty analysis
given in the next section.

5.6.1 unCeRtainty analysis aPPlied to PyRanometeR CalibRation


The procedure for determining the uncertainty in calibrating a pyranometer based upon the sum-
mation technique will be illustrated. This example is taken from Reda et al. (2008). For calibrating
pyranometers with a thermal offset, the equation for the responsivity is

V − offset
R= (5.6)
DNI ⋅ cos ( sza ) + DHI

where
R = responsivity in [µV/(Wm−2)]
V = thermopile output voltage in (µV)
offset = To determine the offset, the offset voltage is first determined from nighttime measure-
ments of the pyranometer offset as a function of the net radiation measured by an infrared
radiometer (pyrgeometer). After the pyranometer responsivity to net IR irradiance Rnet is
established, this information can be used during the daytime with net infrared measure-
ments Wnet to derive the offset, which will equal Rnet·Wnet.
Rnet = longwave net responsivity of the pyranometer in [µV/(Wm−2)]
Wnet = infrared net irradiance measured by collocated pyrgeometer in (Wm−2)
DNI = DNI measured by the reference pyrheliometer in (Wm−2)
sza = solar zenith angle
DHI = DHI measurement (Wm−2)

Thermopile pyranometers are subject to an IR radiative loss or thermal offset. As discussed in


Section  5.3.2, black-and-white pyranometers have a minimal thermal offset. Sometimes it is
possible to determine both the broadband solar and IR responsivities (Rnet) for the pyranometer.
The instrument-specific Rnet can be measured in a blackbody chamber (Reda et al. 2005). Generic
offsets have been determined using a blackbody chamber for Eppley and Kipp & Zonen pyranome-
ters by model, but there are still some variations within model types. The Rnet is typically 10 percent
or smaller than the broadband solar responsivity (R). Having Rnet  is useful only when pyrgeom-
eter data are available to determine net IR flux and compute the thermal offset for a specific or
generic pyranometer. If a pyrgeometer is not present, the nighttime thermal offsets can be averaged
from the night before and night after the daytime measurements to estimate an offset correction.
This method will, of course, carry a larger uncertainty than the measured offset estimate (Dutton
et al. 2001; Vignola et al. 2007).
The sensitivities of the coefficients in Equation 5.6 to changes in the variables are found by cal-
culating partial derivatives of the responsivity (R) with respect to each of the variables:

∂R 1
cV = = (5.7)
∂V  DNI ⋅ cos ( sza ) + DHI 
140 Solar and Infrared Radiation Measurements

∂R −Wnet
cRnet = = (5.8)
∂Rnet  DNI ⋅ cos ( sza ) + DHI 

∂R − Rnet
cWnet = = (5.9)
∂Wnet  DNI ⋅ cos ( sza ) + DHI 

cDNI =
∂R
=−
(V − Rnet ⋅ cos ( sza ) ) (5.10)
 DNI ⋅ cos ( sza ) + DHI 
2
∂DNI

∂R DNI ⋅ sin ( sza ) ⋅ (V − Rnet ⋅Wnet )


csza = = (5.11)
 DNI ⋅ cos ( sza ) + DHI 
2
∂sza

∂R − (V − Rnet ⋅Wnet ) (5.12)


cDNI = =−
 DNI ⋅ cos ( sza ) + DHI 
2
∂DHI

Table 5.10 contains the Type B standard uncertainties for the variables in Equation 5.6. All these
uncertainties are assumed to come from rectangular distributions implying that any value between
plus or minus the u-value in column three of Table 5.10 is equally likely. For this type of distribu-
tion, the standard uncertainty is calculated as u/√3.

uB2 = cV2 ∆V 2 + cR2net ∆Rnet


2
+ cW2 net ∆Wnet
2
+ cDNI
2
∆DNI 2 + cZ2 ∆Z 2 + cD2 HI ∆DHI 2 (5.13)

The pyranometer responsivity using the data in Table 5.10 and Equation 5.6 is 8.0736 µV/Wm−2.


The combined uncertainty of the calibration is determined by combining the standard uncertainties
with sensitivity coefficients applied in quadrature.
Using the data in Table 5.10, uB is calculated to be 0.022 µV/ Wm−2. The uncertainties of Type
A arise from statistical sampling of the measured voltage V and the measured irradiances DNI and
DHI used in the determination of R. Therefore, an equation of the form

u 2A = cV2 ∆V 2 + +cDNI
2
∆DNI 2 + cDHI
2
∆DHI 2 (5.14)

determines the combined uncertainties of the Type A variables. The total combined uncertainty
is computed from the Type A and Type B standard uncertainties by adding them in quadrature.
A coverage factor k is applied to the combined uncertainty to determine an expanded uncertainty,
usually, a 95 percent confidence interval or a 99 percent confidence interval

TABLE 5.10
Estimated Standard Uncertainties of Type B
Variable Value u Offset a = u + offset Distribution DF ustd
V 7930.3 µV 0.079 µV 1 µV 1.079 µV Rectangular 1000 0.62
Rnet 0.4 µV/(Wm−2) 0.04 µV/(Wm−2) – 0.4 µV/(Wm−2) Rectangular 1000 0.02
Wnet –150 Wm−2 7.5 Wm−2 – 7.5 Wm−2 Rectangular 1000 4.33
DNI 1,000 Wm−2 4 Wm−2 – 4 Wm−2 Rectangular 1000 2.31
szaa 20° 0.00002 – 0.00002 Rectangular 1000 0.00001
DHI 50 Wm−2 1.5 Wm−2 1 Wm−2 2.5 Wm−2 Rectangular 1000 1.44

a Since the uncertainty in sza is 0.003°, at 20° the uncertainty is cos(sza + 0.003) − cos(sza).
Broadband Global Irradiance 141

U E = k ⋅ u A2 +uB2 
(5.15)
The expanded uncertainty (UE) is defined as the interval that is expected to contain 95 percent (k ≅ 2)
or 99 percent (k ≅ 3) of the values from measurements. The +UE and –UE can be different values as some
errors do not have symmetrical distributions. Understanding the GUM procedure requires some statistics
background, especially for determining the coverage factor. For example, a student’s t-distribution should
be used to determine k when the number of samples is small (n < 30) and the distribution is nearly
normal. For a large number of samples, the student’s t-distribution yields a coverage factor of 1.96 for
95 percent uncertainty as it would be for a Gaussian distribution of an infinite number of samples.

5.6.2 an examPle of tHe GUM PRoCeduRe aPPlied to tHe


CalibRation unCeRtainties of a PyRanometeR
The  responsivity of pyranometers changes as a function of solar zenith angle. The  responsivity
behavior is similar among pyranometers of the same model. Because of this change with solar zenith
angle, it is necessary to be specific when stating a single calibration constant for a pyranometer.
The NREL has a calibration procedure based on the summation method called Broadband Outdoor
Radiometer CALibration (BORCAL) (Myers et al. 2002; Dooraghi et al. 2019). Pyranometers cali-
brated with the BORCAL method are assigned the responsivity determined when the solar zenith
angle is 45° with the expanded uncertainty based on the calibration results for solar zenith angles
between 30 and 60°. Responsivities are also reported at 2  degree solar zenith angle increments,
and a composite responsivity based on all the solar zenith angles in the calibration is reported (see
Appendix G for a sample BORCAL report).
Recent BORCAL reports stipulate that the responsivity given in the report are good for the
pyranometer calibrated on the given day. What does this mean? The responsivity of the pyranom-
eter depends on the mix of DNI and DHI. The responsivity to DNI changes with the sza, while the
DHI responsivity can be considered fixed and in general is very close to the pyranometer respon-
sivity to DNI at 45°. The GHI responsivity is then a weighted average of the DrHI and the DHI
responsivity and depends on the percentage of the GHI that comes from DrHI and DHI. Therefore,
as the percentage of DrHI and DHI that make up the GHI change, the responsivity changes. With
pyranometers that have a good cosine response, the change in GHI responsivity with sza is not as
dependent on the mixture of DrHI and DHI irradiance. Other effects such as thermal offsets, tem-
perature response, and spectral response can also play a role. Providing the responsivity at a given
angle of 45° from one ideal day at one location is a way of tracking the change in the responsivity
of the instrument, but use of the responsivity in the field operation has additional uncertainties that
must be acknowledged. A more detailed discussion on uncertainties is provided in Section 5.6.3.
In Section 5.6.1 the Type B uncertainties were discussed in detail. Examples of the Type A uncer-
tainties will be discussed here for two different methods of defining the pyranometer responsivity.
As shown in Figures  5.39 and 5.40, the responsivity of a pyranometer changes over the day with
the solar zenith angle. For the uA uncertainty for the responsivity at 45°, a simple standard deviation is
determined for the data used to obtain the responsivity. The combined uncertainty is obtained by add-
ing in quadrature the uB determined in Section 5.6.1 with the uA determined by the statistical analysis.
The expanded uncertainty (UE) is obtained by multiplying the combined uncertainty by the coverage factor.
The Type A uncertainties for the responsivity data binned by solar zenith angle are determined
in a similar manner to the Type A  uncertainty determined for the responsivity at 45°. These u A
uncertainties are the statistical uncertainties of the data used to obtain the responsivity for the given
2 degree solar zenith angle bin.
The discussion here is limited to calibrations performed under idealized conditions with mini-
mal diffuse irradiance. This discussion emphasizes the dependence on sza of the DrHI responsivity
and the effects of thermal offset as there are no clouds that minimize the thermal offset.
142 Solar and Infrared Radiation Measurements

Including the complexity of the solar zenith angle effect using the GUM methodology is chal-
lenging because it involves the combined uncertainty over the full range of solar zenith angles or
over a given range of solar zenith angles. The first task is to create a spline fit through the data points
corresponding to the morning (AM) and afternoon (PM) measurements:
3

Ri , AM ( Z ) = ∑a ( sza − sza )
j
j i (5.16)
j =0

Ri , PM ( Z ) = ∑a ( sza − sza )
j
j i (5.17)
j =0

This is a third-degree polynomial through four data points. A natural cubic spline interpolation is a
piecewise cubic fit that is twice differentiable, and the second differential is zero at the end points.
The  aj are determined by the spline fit to the nearest data points. With the spline responsivities
determined for each measurement, the root mean square error (RMSE) can be calculated for all
data. The RMSE is the difference between the measured responsivity and the responsivity deter-
mined for each bin by the piecewise fit.

∑ ∑ 
m n
( Ri ,meas − Ri , AM ) ( Ri ,meas − Ri ,PM )
2 2
 
r =
2 i =1
+ i =1
/2
m n
 
(5.18)

∑ ∑ 
m n
2 2
 r i , AM r i , PM

= i =1
+ i =1
/2
m n
 

where r2 represents the sum of the squares of the AM and PM residuals with respect to the spline fit
divided by the total number of data points. The number of responsivity data points in the AM is m
and the number of points in the PM is n.
Next, the task is to calculate the standard deviation of the residuals from the spline fit to deter-
mine the average residual. Since there are separate spline fits for the AM and PM, a standard devia-
tion for the morning and afternoon hours can be determined:


m
( rAM − ri , AM )
2

σ AM = i =1
(5.19)
m −1
where,
rAM in this equation represents the average residual for the spline fit in the morning
ri, AM is the individual residual of each data point

The afternoon standard deviation is the same format except the AM values are replaced by the PM
values. The standard uncertainty (uint) from the Type A analysis is then

uint = (r 2
AM + rPM
2
)
/ 2 + σ AM
2
+ σ PM
2
(/2 ) (5.20)

and the degrees of freedom are m + n − 2. For this example, the uncertainty for the interpolated function
(spline fit in this example) is ±0.21 percent, and the Type B uncertainty (uB) is ±0.61 percent. Combining
their uncertainty in quadrature gives the combined uncertainty (uC) of ±0.65 percent. The coverage
factor for this data set is 1.96, which yields an expanded uncertainty at a 95 percent confidence interval
of ±1.27 percent.
Broadband Global Irradiance 143

In  the NREL BORCAL reports, the assigned responsivity is specified at the 45° solar zenith
angle. This responsivity is accompanied by a plus-minus (±) range of uncertainty at the 95 percent
level of confidence for measurements made when the solar zenith angle is between 30 and 60°.
This confidence level is determined in the following manner. First Type A and Type B uncertainties
for the 45° responsivity are determined. The percent expanded uncertainty for the 45° responsivity
is then calculated. Next, the minimum and maximum responsivities in the range from 30 to 60° are
determined from the calibration measurements (see Figure 5.41). After these plus-minus differences
are obtained, the percent uncertainty in the responsivity is calculated from

%U ( differences) ± = ( Rmax R min − R45 ) ± R45 (5.21)

The expanded uncertainty is calculated by

U E ± = U B − 45± + U differences ± (5.22)

In this example, the expanded uncertainty (UE ) for the responsivity at 45° is ±0.64 percent. The dif-
ference between the responsivity at 45° and the maximum and minimum responsivities in the 30
to 60o range provides a +Udifferences and a −Udifferences. The differences shown in Figure 5.41 are used
in Equation 5.22 to calculate the percent differences +Udifferences  = 1.54% and −Udifferences  = −3.01%.
Therefore, the expanded uncertainty for the plus range would be 2.18 percent and −3.64 percent for
the negative range. The offset uncertainties are at the 95 percent confidence level and are the values
from simple addition—not added in quadrature. Uncertainties for other zenith angle ranges can be
calculated in a similar manner.
Calibration uncertainties have been quoted here in percentages as calibrations occur under clear-
sky conditions with a significant amount of GHI, and the DrHI is much greater than the DHI. When
performing calibrations at large sza, the GHI the DrHI contributions are much smaller and the
responsivity of the DHI and its uncertainty plays a much bigger role. In addition, uncertainties that
do not scale with intensity can play bigger roles.
As mentioned earlier in this chapter, calibrations should be performed with the instrument con-
figured as it would be in the field. If the instrument will be in a VEN in the field, the most appropri-
ate calibrations are done with the instrument in a similar type of VEN.

FIGURE 5.41 Responsivity plotted against zenith angle as in Figure 5.32. The uncertainty offsets U+(offset)
and U−(offset) are shown on the plot.
144 Solar and Infrared Radiation Measurements

The uncertainties discussed so far are for calibrations. Uncertainties associated with field condi-
tions that are different from the calibration conditions must be added to the uncertainties discussed
for the calibration (e.g., data logger measurement specifications, stability and leveling of mounting,
degree of solar access, electrical interference, etc.). The estimated measurement uncertainty of field
measurements is largely dependent upon how the instruments are installed and maintained in the
field and the mixture of DrHI and DHI that occur.

5.6.3 imPoRtanCe of undeRstanding limitations of PeRCent unCeRtainties


Pyranometer specifications come with a considerable amount of detailed information. It is impor-
tant to understand the meaning of these specifications and to understand that when the pyranometer
is used outside the limits set in the specifications, uncertainties can increase. In addition, some spec-
ifications are given in units of irradiance while others are given as a percentage of the reading. Just
because an instrument says it has an uncertainty of ±2.5 percent at a 95 percent level of confidence
does not mean that all measurements will have a 2.5 percent uncertainty. In fact, this uncertainty
is related to conditions under which the instrument was calibrated. When the irradiance levels are
very low, the uncertainties can be 100 percent and the instrument will still meet the specifications
of the manufacturer because some uncertainties are quoted as given values and not as a percentage
or the conditions are outside those conditions specified by the manufacturer.
A glaring example is the cosine response of the instrument. Calibrations and uncertainties are often
quoted at an incident angle of 45° or over a 30 to 60° range. However, there are many occasions when
the sza zenith angle is outside this range. For example, for a latitude of 40° or greater from the equator,
during the winter, the sun is always low in the sky and always has a sza greater than 60°. This condi-
tion results in measurements being made outside the specifications, and the measurements likely will
have larger uncertainties. Figures 5.42 and 5.43 show the deviation of the cosine responses for an older
Eppley PSP and a newer Kipp & Zonen CMP22 divided by reference GHI values obtained during

FIGURE 5.42 Plot of the ratio of the output of an Eppley PSP and a Kipp & Zonen CMP22 pyranometer
divided by reference GHI calculated from DrHI from a Kipp & Zonen CHP1 and the DfHI from a Schenk
pyranometer. Data taken on July 8, 2017, in Eugene, Oregon, on a clear day. Large scattering of the data during
the early morning or late evening results from objects on the horizon.
Broadband Global Irradiance 145

FIGURE 5.43 Plot of the ratio of the output of an Eppley PSP and a Kipp & Zonen CMP22 pyranometer
divided by reference GHI calculated from DrHI from a Kipp & Zonen CHP1 and the DfHI from a Schenk
pyranometer. Data taken on December 6, 2017, in Eugene, Oregon, on a clear day. Large scattering of the data
during the early morning or late evening results from objects on the horizon.

winter and summer by summing the DrHI from a Kipp & Zonen CHP1 and DfHI from a Schenk
pyranometer. When the instrument is operating with the sza less than 60° the measurements are within
specifications. When the sza gets much larger than 60° the differences increase.
Therefore, when talking about the uncertainty in measurements, it is important to understand
the caveats that go with the statement and the consequences of any larger uncertainties on results
obtained using the data.
Under cloudy skies, the differences between the measured DHI and the GHI can be very small
and likely relate to the linearity of the responsivity or a small thermal offset. This results from the
responsivity of the pyranometer under diffuse skies is close to the responsivity of the instrument
at 45°. In this case, the uncertainties determined under clear-sky conditions may be the extreme
uncertainties. Because of dependence of the responsivity pyranometer on the mixture of DrHI and
DHI and the thermal offset, GHI obtained by the summation method provides a better estimate
when the instruments are well maintained.

QUESTIONS
1. If the DNI is 1,000 Wm−2 and the DHI is 100 Wm−2 and the solar zenith angle is 45°, what
is the GHI? What is the direct irradiance on a horizontal surface?
2. What is meant by a Lambertian response?
3. Is a class A pyranometer more precise that a class B pyranometer? List three reasons one
type of pyranometer is better than the other.
4. Briefly explain how a black-disk thermopile pyranometer measures solar energy.
5. Why do many class A and class B pyranometers have two quartz domes?
6. What is meant by a thermal offset in a pyranometer? What is one way to estimate the ther-
mal offset of a pyranometer?
7. What is a ventilator, and why is it used?
146 Solar and Infrared Radiation Measurements

8. Define non-linearity. If a pyranometer has a responsivity of 8.05 µV/Wm−2 at a solar inten-


sity of 1,000 Wm−2, a responsivity of 8.00 µV/Wm−2 at 500 Wm−2, and 7.95 µV/Wm−2 at
100 Wm−2. What is the magnitude of non-linearity of the pyranometer?
9. Why should a pyranometer be calibrated on a yearly basis?
10. What is meant by advective heat loss?
11. Describe how a black-and-white pyranometer works.
12. In equilibrium, a black surface has an emissivity of 95 percent; what is its absorptivity?
13. Can a white surface be highly reflective in the visible portion of the spectrum and highly
absorptive in the infrared portion of the spectrum?
14. Explain why black-and-white pyranometers have smaller thermal offsets than some class
A pyranometers.
15. List a number of reasons photodiode-based pyranometers are better at measuring solar
radiation than a PV module.
16. Will a silicon-based photodiode pyranometer respond to sunlight with a wavelength of
1,400 nm? Explain why.
17. Why does the LI-COR LI-200 photodiode pyranometer exhibit its cat ears response?
18. If the ambient temperature is 25°C, how does the responsivity of the Eppley Model
8-48  pyranometer change? If the temperature is 0°C, how does the responsivity of the
pyranometer change?
19. On a clear day, a photodiode pyranometer responds differently to DHI than to DNI. Explain
why? During the middle of the day, is there a large effect on the GHI measurements?
20. Explain the shade–unshade method of calibration.
21. Explain the summation calibration technique.
22. Name several causes of Type A uncertainties.
23. Name several causes of Type B uncertainties.
24. If the Type A uncertainties are ±3 percent and the Type B uncertainties are ±4 percent,
what is the combined uncertainty? What is the expanded uncertainty at a 95 percent level
of confidence?

REFERENCES
Abbot, C. G., and L. B. Aldrich. 1916. The  pyranometer—An instrument for measuring sky radiation.
Smithsonian Miscellaneous Collection 66 (7).
ASTM. 1997. Standard Method for Calibration of Reference Pyranometers with Axis Vertical by the Shading
Method, E913-97. West Conshohocken, PA: ASTM.
ASTM. 2015. Standard Test Method for Calibration of a Pyranometer Using a Pyrheliometer, G167. West
Conshohocken, PA: ASTM.
ASTM. 2017. Standard Guide for Evaluating Uncertainty in Calibration and Field Measurements of
Broadband Irradiance with Pyranometers and Pyrheliometers, G213. West Conshohocken, PA: ASTM.
ASTM. 2008. Standard Test Methods for Electrical Performance of Non-concentrator Terrestrial Photovoltaic
Modules and Arrays Using Reference Cells, Standard E1036-08. West Conshohocken, PA: ASTM.
ASTM. 2010. Standard Test Method for Calibration of Non-Concentrator Photovoltaic Secondary Reference
Cells, Standard E1362-10. West Conshohocken, PA: ASTM.
ASTM. 2009. Standard Test Method for Electrical Performance of Photovoltaic Cells Using Reference Cells
Under Simulated Sunlight, Standard E948-09. West Conshohocken, PA: ASTM.
Augustyn, J., T. Geer, T. Stoffel, R. Kessler, E. Kern, R. Little, and F. Vignola. 2002. Improving the accuracy
of low cost measurement of direct normal solar irradiance. In  Conference Proceedings of the 2002
American Solar Energy Society, Solar 2002. Reno, NV: American Solar Energy Society.
Augustyn, J., T. Geer, T. Stoffel, R. Kessler, E. Kern, R. Little, F. Vignola, and B. Boyson. 2004. Update
of algorithm to correct direct normal irradiance measurements made with a rotating shadow band
pyranometer. In  Conference Proceedings of the 2004  American Solar Energy Society, Solar 2004.
Portland, OR: American Solar Energy Society.
Broadband Global Irradiance 147

BIPM. 2012. Evaluation of measurement data-The role of measurement uncertainty in conformity assessment.
Joint Committee for Guides in Metrology, JCGM Volume 106. https://www.bipm.org/utils/common/
documents/jcgm/JCGM_106_2012_E.pdf (Accessed May 23, 2019).
Bunch, B., and A. Hellemans. 2004. The History of Science and Technology: A Browser’s Guide to the Great
Discoveries, Inventions, and the People Who Made Them, from the Dawn of Time to Today. Boston,
MA: Houghton Mifflin.
Dooraghi, M., A. Andreas, M. Kutchenreiter, I. Reda, M. Stoddard, M. Sengupta, and A. Habte. 2019. Broadband
Outdoor Radiometer Calibration (BORCAL) Process for the Atmospheric Radiation Measurement
(ARM) Program, Second Edition. NREL/TP-5D00-71476. Golden, CO: National Renewable Energy
Laboratory, 54 pp. https://www.nrel.gov/docs/fy19osti/71476.pdf (Accessed May 25, 2019).
Dutton, E. G., J. J. Michalsky, T. Stoffel, B. W. Forgan, J. Hickey, D. W. Nelson, T. L. Alberta, and I. Reda.
2001. Measurement of broadband diffuse solar irradiance using current commercial instrumentation
with a correction for thermal offset errors. Journal of Atmospheric Oceanic Technology 18:297–314.
Fröhlich, C. 1991. History of solar radiometry and the world radiometric reference. Metrologia 28(3):111–115.
http://iopscience.iop.org/0026-1394/28/3/001 (Accessed May 25, 2019).
Guechia, A., M. Chegaarb, and A. Merabet. 2011. The effect of water vapor on the performance of solar cells.
Physics Procedia 21:108–114.
Gulbrandsen, A. 1978. On the use of pyranometers in the study of spectral solar radiation and atmospheric
aerosols. Journal of Applied Meteorology 17:899–905.
Hishikawa, Y., M. Yoshita, H. Ohshima, K. Yamagoe, H. Shimura, A. Sasaki, and T. Ueda. 2018. Temperature
dependence of the short circuit current and spectral responsivity of various kinds of crystalline silicon
photovoltaic devices. Japanese Journal of Applied Physics 57:08RG17.
IEC. See International Electrotechnical Commission.
International Electrotechnical Commission. 2007. IEC 60904-2: Photovoltaic Device—Part 2: Requirements
for Reference Solar Cells. Geneva, Switzerland: International Electrotechnical Commission.
International Organization for Standardization. 2018. Solar energy — Specification and classification of
instruments for measuring hemispherical solar and direct solar radiation. ISO 9060:2018. https://www.
iso.org/standard/67464.html (Accessed May 23, 2019).
International Organization for Standardization. 1993. Solar energy—Calibration of a pyranometer using a
pyrheliometer. ISO 9846. https://www.iso.org/standard/17724.html (Accessed May 23, 2019).
ISO. See International Organization for Standardization.
JCGM. See Joint Committee for Guides in Metrology.
Joint Committee for Guides in Metrology. 2008. Evaluation of Measurement Data—Guide to the Expression
of Uncertainty in Measurement. GUM 1995 with Minor Revisions. Bureau International des Poids et
Mesures. http://www.bipm.org/en/publications/guides/gum.html (Accessed May 26, 2019).
Kerr, J. P., G. W. Thurtell, and C. B. Tanner. 1967. An integrating pyranometer for climatological observer
stations and mesoscale networks. Journal of Applied Meteorology 6:688–694.
King, D. L., and D. R. Myers. 1997. Silicon-photodiode pyranometers: Operational characteristics, historical
experiences, and new calibration procedures. Paper Presented at the 26th IEEE Photovoltaic Specialists
Conference. September 29–October 3, Anaheim, CA.
King, D. L., J. A. Kratochvil, and W. E. Boyson. 1997. Measuring solar spectral and angle-of-incidence effects
on photovoltaic modules and solar irradiance sensors. Paper Presented at the 26th IEEE Photovoltaic
Specialists Conference. September 29–October 3, Anaheim, CA.
King, D. L., W. E. Boyson, B. R. Hansen, and W. I. Bower. 1998. Improved accuracy for low-cost solar irra-
diance sensors. Paper Presented at the 2nd World Conference and Exhibition on Photovoltaic Solar
Energy Conversion, July 6–10. Vienna, Austria.
Long, C. N., K. Younkin, and D. M. Powell. 2001. Analysis of the Dutton et al. IR loss correction technique applied
to ARM diffuse SW measurements. In Proceedings of the Eleventh Atmospheric Radiation Measurement
(ARM) Science Team Meeting. ARMCONF-2001. Washington, DC: United States Department of Energy.
Marion, B. 2017. Numerical method for angle-of-incidence correction factors for diffuse radiation incident
photovoltaic modules. Solar Energy 147:344–348.
McArthur, L. J. B. 2004. Baseline Surface Radiation Network (BSRN). Operations Manual. WMO/TD-No.
879. Geneva, Switzerland: World Climate Research Programme/World Meteorological Organization.
Michalsky, J., J. Liljegren, and L. Harrison. 1995. A comparison of sun photometer derivations of total col-
umn water vapor and ozone to standard measures of same at the Southern Great Plains Atmospheric
Radiation Measurement site. Journal of Geophysical Research 100 (D12): 25995–26003.
148 Solar and Infrared Radiation Measurements

Michalsky, J, M. Kutchenreiter, and C. N. Long. 2017. Significant improvements in pyranometer offsets


using ventilation high-flow DC. Journal of Atmospheric and Oceanic Technology 34 (6): 1323–1332.
doi:10.1175/JTECH-D-16-0224.1.
Michalsky, J., E. G. Dutton, D. Nelson, J. Wendell, S. Wilcox, A. Andreas, P. Gotseff, et al. 2011. An extensive
comparison of commercial pyrheliometers under a wide range of routine observing conditions. Journal
Atmospheric Oceanic Technology 28:752–766.
Michalsky, J. J., R. Perez, L. Harrison, and B. A. LeBaron. 1991. Spectral and temperature correction of sili-
con photovoltaic solar radiation detectors. Solar Energy 47:299–305.
Myers, D. 2005. Solar radiation modeling and measurements for renewable energy applications: Data and
model quality. Energy 30:1517–1531.
Myers, D. R., I. Reda, S. Wilcox, and A. Andreas. 2004. Optical radiation measurements for photovoltaic
applications: Instrument uncertainty and performance. National Renewable Energy Laboratory (NREL)
Conference Paper NREL/CP-560-36320. Denver, CO: NREL.
Myers, D. R., T. L. Stoffel, I. Reda, S. M. Wilcox, and A. Andreas. 2002. Recent progress in reducing the
uncertainty in and improving pyranometer calibrations. Transactions of the American Society of
Mechanical Engineers 124:44–49.
Pulfrey, D. L. 1978. Photovoltaic Power Generation. New York: Van Nostrand Reinhold.
Reda, I. 2011. Method to Calculate Uncertainties in Measuring Shortwave Solar Irradiance Using Thermopile
and Semiconductor Solar Radiometers. NREL/TP-3B10-52194. 20 pp. Golden, CO: National Renewable
Energy Laboratory. https://www.nrel.gov/docs/fy11osti/52194.pdf (Accessed May 25, 2019).
Reda, I., and A. Andreas. 2017. Calibration procedure of Hukseflux SR25 to establish the diffuse reference
for the outdoor broadband radiometer calibration. Technical Report NREL/TP-1900-68999. Golden,
CO: National Renewable Energy Laboratory. https://www.nrel.gov/docs/fy17osti/68999.pdf (Accessed
May 26, 2019).
Reda I., T. L. Stoffel, and D. R. Myers. 2002. Establishing the Clear Sky Diffuse Reference for BORCAL
Using EPLAB Model 8-48 Pyranometers at the National Renewable Energy Laboratory. In Proceedings
of the Twelfth ARM Science Team Meeting, Ed. by D. Carrothers. Richland, WA: U.S. Department
of Energy. https://www.arm.gov/publications/proceedings/conf12/extended_abs/reda-i.pdf (Accessed
May 24, 2019).
Reda, I., D. Myers, and T. Stoffel. 2008. Uncertainty estimate for the outdoor calibration of solar pyranom-
eters: A meteorologist perspective. National Conference of Standards International (NCSLI) Journal
of Measurement Science 3 (4): 58–66.
Reda, I., J. Hickey, C. Long, D. Myers, T. Stoffel, S. Wilcox, J. J. Michalsky, E. G. Dutton, and D. Nelson.
2005. Using a blackbody to calculate net longwave responsivity of shortwave solar pyranometers to cor-
rect for their thermal offset error during outdoor calibration using the component sum method. Journal
of Atmospheric and Oceanic Technology 22 (10): 1531–1540.
Reda, I., T. Stoffel, and D. Myers. 2003. A method to calibrate a solar pyranometer from measuring reference
diffuse irradiance. Solar Energy 74:103–112.
Riihimaki, L., and F. Vignola. 2008. Establishing a consistent calibration record for Eppley PSPs. Paper pre-
sented at the Proceedings of the 37th ASES Annual Conference, San Diego, CA.
Sanchez, G., M. L. Cancillo, and A. Serrano. 2016. An intercomparison of the thermal offset for different
pyranometers. Journal of Geophysical Research: Atmospheres 121 (13): 7901–7912.
Stoffel T. L., and D. R. Myers. 2010. Accuracy of Silicon Versus Thermopile Radiometers for Daily and
Monthly Integrated Total Hemispheric Solar Radiation. NREL/CP-5500-48697. Golden, CO: National
Renewable Energy Laboratory. doi:10.1117/12.859743. https://dx.doi.org/10.1117/12.859743 (Accessed
May 26, 2019).
Stoffel, T., D. Renné, D. Myers, S. Wilcox, M. Sengupta, R. George, and C. Turchi. 2010. Concentrating solar
power: Best practices handbook for the collection and use of solar resource data. Technical Report
NREL/TP-550-47465. Golden, CO: NREL.
Velmre, E. 2007. Thomas Johann Seebeck (1770–1831). Proceedings of the Estonian Academy of Sciences,
Engineering 13 (4): 276–282.
Vignola, F. 1999. Solar cell based pyranometers: Evaluation of the diffuse response. Proceedings of the
1999 Annual Conference American Solar Energy Society. Portland, ME: ASES.
Vignola, F. 2006. Removing systematic errors from rotating shadowband pyranometer data. In Proceedings
of the American Solar Energy Society, ed. R. Campbell-Howe. Boulder, CO: American Solar Energy
Society.
Broadband Global Irradiance 149

Vignola, F., Z. Derocher, J. Peterson, L. Vuilleumier, C. Felix, J. Grobner, and N. Kouremeti. 2016. Effects of
changing spectral radiation distribution on the performance of photodiode pyranometers. Solar Energy
129:224–230.
Vignola, F., C. Long, and I. Reda. 2007. Evaluation of methods to correct for IR loss in EppleyPSP diffuse
measurements. Paper Presented at the SPIE Conference. San Diego, CA.
Vignola, F., C. Long, and I. Reda. 2008. Modeling IR radiative loss from Eppley PSP pyranometers. Paper
Presented at the SPIE Conference. San Diego, CA.
Vignola, F., C. Long, and I. Reda. 2009. Testing a model of IR radiative losses. Paper Presented at the SPIE
Conference. San Diego, CA.
Vignola, F., J. Peterson, F. M Mavromatakis, S. Wilbert, A. Forstinger, M. Dooraghi, and M. Sengupta. 2018.
Removing biases from rotating shadowband radiometers. Proceeding of SolarPACES 2018. Casablanca,
Morocco.
Vignola, F., J. Peterson, S. Wilbert, P. Blanc, N. Geuder, and C. Kern, 2017. New methodology for adjust-
ing rotating shadowband irradiometer measurements. AIP Conference Proceedings. Vol. 1850, Issue 1.
doi:10.1063/1.498452.
Walker, J. H., R. D. Saunders, J. K. Jackson, and D. A. McSparron. 1987. NBS Measurement Services:
Spectral Irradiance Calibrations. Washington, DC: U.S. Department of Commerce, National Bureau
of Standards.
Wilber, S., N. Geuder, M. Schwandt, B. Kraas, W. Jessen, R. Meyer, and B. Nouri. 2015. Best Practices for
Solar Irradiance Measurements with Rotating Shadowband Irradiometers. Report of IEA-Task46-B1,
68 pp.
Wilcox, S., A. Andreas, I. Reda, and D. Myers. 2002. Improved methods for broadband outdoor radiom-
eter calibration (BORCAL). Paper Presented at the 12th Atmospheric Radiation Measurement (ARM)
Science Team Meeting Proceedings. St. Petersburg, FL.
Wilcox, S., D. Myers, N. Al-Abbadi, and M. Bin Mahfoodh. 2001. Using irradiance and temperature to deter-
mine the need for radiometer calibrations. Paper Presented at the Forum on Solar Energy, the Power
to Choose. Washington, DC.
World Meteorological Organization. 2008. WMO Guide to Meteorological Instruments and Methods of
Observation. WMO-No. 8, 7th ed. Geneva, Switzerland: WMO.
WMO. See World Meteorological Organization.
Younkin, K., and C. N. Long. 2003. Improved Correction of IR Loss in Diffuse Shortwave Measurement: AN
ARM Values-Added Product. RM TR-009. Richland, WA: Pacific Northwest National Laboratory, U.S.
Department of Energy.

USEFUL LINKS
WMO guide to meteorological instruments and methods of observation, https://library.wmo.int/opac/doc_
num.php?explnum_id=4147
BSRN operator’s manual, http://www.bsrn.awi.de/fileadmin/user_upload/Home/Publications/WCRP21_TD1274_
BSRN.pdf.
Seebeck values, http://www.omega.com/temperature/z/pdf/z021-032.pdf.
Standards, http://www.ictinternational.com.au/brochures/pyranometer-classification.pdf.
Standards, http://www.middletonsolar.com/documents/PyrClass.pdf.
Solar cell temperature dependence, http://pvcdrom.pveducation.org/CELLOPER/TEMP.HTM.
Transmission of wg295 glass, http://www.optical-filters.com/wg295.html.
CIMO Guide http://www.wmo.int/pages/prog/www/IMOP/CIMO-Guide.html
6 Diffuse Irradiance

A little later in the morning, if one hides the solar disk by holding the tip of a single finger at
arm’s length before his eyes, the deep blue sky color holds right up to the edge of the sun itself.
Charles Greeley Abbot (1872–1973)

6.1 INTRODUCTION
The  direct solar (beam) radiation passing through the Earth’s atmosphere is absorbed and/or
scattered by molecules, aerosols, water vapor and other gases, and clouds. The resulting diffuse (sky)
radiation reaching a horizontal surface is termed the diffuse horizontal irradiance (DHI) as depicted
in Figure 6.1. Alternatively, DHI can be considered as the skylight portion of the total hemispheric,
or global horizontal irradiance (GHI), remaining after removing the direct normal irradiance (DNI).
DHI is measured by a properly shaded pyranometer. It can also be calculated from measurements
of DNI and GHI using Equation 4.1, but with much larger uncertainty. The distribution of diffuse
radiance in the sky is not uniform, or isotropic, due to the complex interactions of solar radiation
with the atmosphere. From Earth, the Sun’s disk subtends about 0.52 degrees. When not obscured
by clouds, the intense forward scattering near the solar disk creates a circumsolar radiation field that
can enlarge the Sun’s disk appearance to several degrees, depending mostly on the aerosol burden,
as illustrated in Figure 6.2. The second contributor to the non-uniform distribution of solar radiation
in the sky are multiple reflections near the horizon, termed horizon brightening. For these, and other
factors, measuring diffuse solar radiation is challenging.
Accurately measuring DHI has implications for a number of potential applications, including:

• Characterizing the radiation balance at the Earth’s surface


• Determining the solar resources available for energy conversion
• Utilizing natural daylight in buildings
• Calibrating pyranometers
• Automating data quality assessments of solar radiation measurements.

In this chapter, some key interactions of solar radiation with the atmosphere are briefly discussed,
the available options for measuring DHI are described, aspects of solar calibrations of pyranometers
used for DHI measurements are discussed, and the importance of accurate DHI measurements is
stressed.

6.2 ATMOSPHERIC SCATTERING CONCEPTS


A very clear atmosphere with minimal aerosols gives the sky a deep blue color because of Rayleigh
scattering of solar radiation by air molecules (Scattering and absorption of solar radiation are
discussed in more detail in Chapter 7). This blue color is produced because air molecules scatter
radiation nearly inversely with the fourth power of the wavelength, that is, scattering α λ−4. Thus,
the shorter, blue wavelengths are scattered more efficiently by the very small gas molecules than are
other parts of the visible spectrum. As a result, the Sun appears yellow or red near the horizon since
Rayleigh scattering preferentially removes the bluest light from the transmitted DNI.

151
152 Solar and Infrared Radiation Measurements

FIGURE  6.1 This figure demonstrates the source of diffuse solar radiation that falls on a horizontal
surface and measured by a shaded pyranometer. The black ball on the tracker blocks direct beam radiation
(DNI) from the Sun plus some of the aureole, but allows sunlight scattered by air molecules, aerosols,
and clouds to reach the pyranometer. Ground-scattered solar radiation, beneath the pyranometer horizon,
does not reach the pyranometer detector unless it is further scattered by the atmosphere. (Courtesy of
Bobby Hart.)

FIGURE 6.2 An illustration of an all-sky radiance distribution on a clear, but hazy day with the sun
near the horizon. There is a bright aureole about the sun. There is a minimum brightness near an angle
of 90 degrees from the sun. A bright horizon is produced from light scattering multiple times. (Courtesy
of Bobby Hart.)
Diffuse Irradiance 153

Aerosol-scattered light depends on the aerosol particle size. For “typical” continental aerosols,
the scattering is approximately α λ–1.3. Heavily polluted skies typically have an even larger negative
exponent, and they scatter blue light more effectively than typical continental aerosols, but not as
strongly, as in Rayleigh scattering. This wavelength dependence for aerosols is most noticeable in
extremely polluted conditions where the solar disk may appear red even with the Sun high in the
sky, when the effect of Rayleigh scattering would ordinarily be insufficient to cause noticeable
reddening. Although the Sun’s disk appears red for these hazy conditions, skylight appears milky
white, not blue, especially near the horizon because photons are scattered multiple times during
these heavy pollution episodes. The range of sky color from dark blue to milky blue to milky white
is a rough gauge of the amount of aerosol in the atmosphere.
Cloud particles have large sizes relative to visible wavelengths, and their scattering wavelength
dependence is nearly neutral (i.e., a combination of all colors); therefore, they appear as the color of the
sun. The spatial distribution of skylight is complex and can change rapidly for partly cloudy conditions,
but it can be described mathematically for some conditions. Moon and Spencer (1942) developed
a simple and useful approximation of the spatial distribution of skylight for thick uniform clouds.
This sky is, basically, three times brighter in the zenith than at the horizon and varies as the cosine of
the zenith angle (the angle measured from a point in the sky to the local vertical, not the solar zenith
angle). For very clear skies, where Rayleigh scattering dominates, the skylight is approximately half as
bright 90 degrees from the Sun as it is near the Sun. This interpretation is not strictly correct because
of multiple scattering (photons scattered more than once) that occurs in the atmosphere. Cloudless,
but aerosol-laden skies have a more complex mathematical description for the spatial distribution
of skylight. Figure 6.2 is a low-resolution depiction of the skylight distribution for clear conditions
with moderate aerosol. A shading disk normally blocks the area of the sky surrounding the sun for
DHI measurements, but it may not block all of the enhanced bright aureole around the Sun, called
circumsolar radiation. Similarly, the figure illustrates a bright horizon caused by multiple scattering
at the borders of the image. The darkest area in the figure illustrates the minimum sky brightness
roughly 90 degrees from the Sun due to Rayleigh scattering. Forward-scattered radiation (from the
Sun direction to 90 degrees) is higher than backscattered radiation (90 to 180 degrees from the sun) for
cloudless, aerosol-laden skies. This enhancement in the forward direction depends mainly on the size
of the aerosols with larger aerosols producing more scattering in the forward hemisphere.

6.3 MEASURING DIFFUSE IRRADIANCE


Diffuse irradiance measurements have become more sophisticated with the development of
automated tracking devices that dependably keep the direct solar radiation from reaching the
detector of the pyranometer. The first part of this section describes the use of a fixed shadowband,
a simple device still in use, and the attempts to correct the measurements for excessive skylight
blocked by the band. Most of the correction schemes for this skylight blockage, unfortunately, use as
their “best estimate” the diffuse calculated by taking the difference between the measured GHI and
the horizontal component DNI. “Best estimate” DHI values obtained using this method are prone
to large systematic errors mostly due to the pyranometer’s angular response characteristics, making
comparisons of shadowband correction schemes unreliable and site dependent. The tracking shade
disk or ball is the preferred method for measuring DHI and will be discussed in Section  6.3.2.
A  discussion of a third method for measuring DHI based on a rotating shadowband completes
this section.

6.3.1 fixed sHadoWband measuRements of diffuse iRRadianCe


Before the development of automatic solar trackers capable of dependably following the Sun in two
dimensions (solar elevation and azimuth or hour angle and declination), most DHI measurements
were performed using a fixed shadowband or shade ring that blocked the Sun from sunrise to sunset
154 Solar and Infrared Radiation Measurements

but required manual adjustments for solar declination changes. By design, this band blocked a large
part of the sky in addition to the Sun and had to be adjusted for the Sun’s declination; more often
when the Sun was near the equinoxes when the solar declination changes rapidly (~0.4° per day)
and less often when the Sun was near the solstices when the solar declination changes slowly (<0.1°
per day). Operationally, the shadowband should be aligned to keep the pyranometer detector within
1.5 degree of the center of the shadow cast by direct sunlight.
To align the fixed shade, the plane of the arc formed by the shadowband or the shade ring should
be perpendicular to the polar axis. The  black detector surface of the pyranometer is adjusted to
intersect the polar axis. This latter adjustment allows the shade to block the pyranometer surface
throughout the day after the band is adjusted on the polar axis according to the declination of
the sun. Adjustment details are explained in the installation instructions from the manufacturers.
Kipp & Zonen B.V. continues to offer a fixed shade ring. Figure 6.3 illustrates the proper alignment
of The Eppley Laboratory, Inc. shadowband for a clear summer day in Eugene, Oregon.
Because the fixed shadowband or shade ring blocks more of the sky than where the Sun is
located, a correction is needed for the excluded diffuse irradiance that can be more than 20
percent of the measured diffuse irradiance. This  correction factor for the fixed shadowband or
shade ring has been a frequent topic in the solar radiation literature. Drummond (1956) developed
a geometrical correction for the blocked skylight, assuming an isotropic (uniform) distribution of
skylight. This method is applicable for any location on Earth. However, as previously discussed,
diffuse radiation is never uniformly distributed. Researchers since the study by Drummond (1956)
have developed improved corrections or evaluated available correction schemes for the actual
distribution of skylight for different conditions from overcast to clear skies. Corrections derived for
one climate are often not applicable to another because different sky conditions prevail from location
to location. Additionally, the accuracy of these corrections is limited by determining a “reference”
DHI from measurements of GHI and DNI with their associated uncertainties (see Section 6.2.2), as
discussed earlier.

FIGURE 6.3 The Eppley Laboratory, Inc., model SBS shadowband blocks direct sunlight on a clear summer
day in Eugene, Oregon. Note that the detector surface is well centered beneath the shadowband. Under
cloudless skies, the circumsolar irradiance is the brightest region of the sky and the shadowband blocks a
significant part of this brightest region.
Diffuse Irradiance 155

Kudish and Evseev (2008) evaluated four of these correction methods: Drummond (1956),
LeBaron et al. (1990), Battles et al. (1995), and Muneer and Zhang (2002). Gueymard and Myers
(2009) pointed out, however, that this evaluation contains systematic biases because the data used
for establishing the “reference diffuse values” used the subtraction method to calculate diffuse
values from GHI and direct horizontal values. Diffuse values obtained by the subtraction method
contain significant systematic biases due in large part to the angular response (or cosine) errors
of the unshaded pyranometer and do not have the accuracy to establish a reliable diffuse value.
Most, maybe all, of the correction methods developed have used the subtraction method for
diffuse values as the standard for comparison. It is clear that more research is needed to establish
a more reliable shadowband adjustment factor; however, fixed shadowband DHI values with these
corrections will always have larger uncertainties than the current best methods of measurement.
Our recommendation, as will become clear, is to measure diffuse with a tracking disk (ball) such
as the one shown in Figure 6.4.

6.3.2 CalCulated diffuse iRRadianCe veRsus sHading disk diffuse


Rearranging Equation  2.12 from Chapter  2, DHI can be calculated from measurements of GHI
and DNI:

DHI = GHI − DNI ⋅ cos ( sza ) (6.1)

where sza is the solar zenith angle. With an estimated 95 percent confidence interval, a very good
measurement of the DNI has an uncertainty (U95) of around ±2 percent, and similarly a very good
GHI measurement has a U95 of ±3 percent or higher because of additional uncertainties associated

FIGURE 6.4 An example of a shaded-disk measurement of DHI. The arm holds disks that shade each of
three pyranometers mounted on the automatic solar tracker.
156 Solar and Infrared Radiation Measurements

with thermal offsets and imperfect angular response (Stoffel et al. 2010). Therefore, the uncertainty
of calculating diffuse irradiance from GHI and DNI is significant. To illustrate the uncertainty in
calculating diffuse in this way, assume a value of 1,000 Wm−2 for DNI and GHI at a solar zenith
angle of 30°. The calculated diffuse then would be

DHI = 1, 000 − 1, 000 ⋅ cos ( 30° ) = 134  Wm −2  (6.2)

with an expanded uncertainty of

( 30 ) + ( 20 ⋅ 0.866 )
2 2
U95 = = 35  Wm −2  (6.3)

where the U95 estimates for GHI and DNI are summed in quadrature (e.g., 3% and 2% of the mea-
sured irradiances). Comparing the above results indicates a ±26 percent uncertainty in the calculated
diffuse using two well-made, first-class measurements.
A vast improvement in uncertainty is possible if the diffuse is, instead, measured with a tracking
disk, or ball, shading a pyranometer mounted on an automatic solar tracker. Applying the ±3 percent
uncertainty of the pyranometer measurement to this example, the U95 for the DHI measurement
would be about 4  Wm−2 (134  Wm−2  ×  0.03), for a reduction of nearly an order of magnitude in
uncertainty of the DHI value. Therefore, the preferred method to evaluate fixed shadowband
corrections (discussed in Section 6.2.1) is to use tracking disk diffuse measurements for the reference
diffuse irradiance. As more recent applications have shown, the marked improvements in measuring
DHI with a shading disk often outweigh the increased costs of the automatic solar tracker.
As discussed in Chapter 5, first-class pyranometers with thermopile-based detectors can have
significant thermal offsets. These offsets can be large when compared with diffuse irradiance
levels. Depending on the net infrared (IR) radiation, which is generally larger during very clear-
sky conditions, pyranometer thermal offsets can be 20 percent or more of the DHI value. In fact,
pyranometers that are uncorrected for thermal offsets have reported diffuse irradiances lower
than predicted for purely Rayleigh scattering atmospheres (Cess et al. 2000). Therefore, it can be
preferable to use second-class pyranometers based on black-and-white thermopile detectors with
very small thermal offsets for DHI measurements. High-cost, secondary-standard pyranometers
also have minimal thermal offsets and better cosine responses, so the best DHI measurements may
be made using secondary-standard pyranometers. A comparison of DHI measurements made using
most of the commercially available, ventilated pyranometers in the early 2000s can be found in
Michalsky et al. (2003).

6.3.3 Rotating sHadoWband diffuse measuRements


The  rotating shadowband radiometer (RSR), also known as a rotating shadowband pyranometer
(RSP) or rotating shadowband irradiometer (RSI),1 offers a third option for the measurement of
diffuse irradiance with at least five suppliers of these types of instruments (see Appendix G).
Irradiance, Inc. (http://www.irradiance.com) offers the RSR (RSR2TM), which uses a rapidly
moving band to alternately expose and shade the diffuser of a silicon photodiode pyranometer with
10 microsecond (µs) time response. Hundreds of measurements are made during the band sweep
with many near, but not quite shading, the diffuser. The measurements made in the solar aureole,
the bright disk surrounding the Sun, are used to derive an approximation for the excess skylight that

1 The  International Energy Agency (IEA) Solar Heating and Cooling Task 46, subtask B has adopted the term RSI to
denote an instrument measuring irradiance by use of a rotating shadowband (Wilbert et al. 2015).
Diffuse Irradiance 157

is blocked by the band during the shaded measurement. This is like a real-time fixed shadowband
correction, which was discussed earlier, but uses actual measurements for the correction. The most
significant issue with the RSR2TM is that the silicon photodiode sensor has a non-uniform spectral
response that significantly underestimates clear-sky diffuse if the instrument is calibrated for global
or direct irradiance (see Section 5.4.2). Software to correct this shortcoming is included with the
instrument control and data acquisition system. The effectiveness of this or other correction schemes
remains under investigation (Sengupta et al. 2017).
Yankee Environmental Systems (YES), Inc. (http://www.yesinc.com) makes a seven-channel
multifilter RSR (MFR-7) and a single detector radiometer (SDR) that makes a spectral and broadband
measurement of diffuse irradiance. Initially, they also used a silicon photodiode detector and had to
deal with a few of the same measurement issues previously outlined for the RSR. YES now offers
the MFR-7 with a fast-response thermopile in place of the unfiltered silicon cell (with six additional
channels of narrowband wavelength measurements). The performance of this thermopile device is
still under study since this thermopile-based radiometer is a relatively new development. Although
the detector is broadband, the diffuser/integrating cavity of this instrument introduces a wavelength
dependence; further investigation is needed.
Prede, Inc. (http://prede.com/english/PDF/PRB-100e.pdf) makes the rotating shadow blade
(RSB-100), which works with commercial pyranometers (thermopile or silicon cell versions) to
make diffuse and global irradiance measurements. There  is little information on the design and
measurement performance of this device.
The  RSP (RSP 4G) is manufactured by Reichert GmbH (http://reichertgmbh.de/).
This  instrument is sold to other companies that provide solar resource assessment services
(e.g., Suntrace GmbH at the website (http://suntrace.de/services/solar-resource-meteorological-
services.html) provides a clear picture of this instrument). It appears to operate similarly to the
Irradiance RSR2TM.
CSP Services is another German company that manufactures what they call an RSI. Besides
this instrument, called the Twin RSI, which has two sensors, they also sell the RSP 4G of Reichert
GmbH (http://www.cspservices.de/products-services/measurement-systems/mdi). Note that the
RSR2TM, the RSP 4G, and the Twin-RSI all use the LI-COR LI-200 silicon cell pyranometer as
their solar sensor. More details of RSR-style instruments are presented in Chapter 10.

6.4 CALIBRATION OF DIFFUSE PYRANOMETERS


Michalsky et al. (2007) provide a thorough discussion of the calibration and characterization of
pyranometers with the goal of defining a standard for DHI measurements. The  preferred, but
labor-intensive, method for calibrating any pyranometer is to alternately shade and unshade the
pyranometer while measuring the reference DNI with an absolute cavity radiometer (ACR).
The  pyranometer response, which is the inverse of the calibration factor (a multiplier), is then
calculated using

R=
( Vunshaded − Vshaded ) (6.4)
DNI ⋅ cos ( sza )

where
Vx are the voltage readings of the pyranometer unshaded and shaded
sza is the solar zenith angle at the time of measurement

These calibration measurements should be made when the Sun is at a solar zenith angle of
45° ± 3 degrees. The details of performing this calibration procedure are in Michalsky et al. (2007)
and are discussed in Chapter 13.
158 Solar and Infrared Radiation Measurements

FIGURE  6.5 The  geometry used to integrate the radiance distribution of skylight to calculate diffuse
horizontal irradiance.

Why calibrate at 45°? If skylight were isotropic (i.e., uniform in all directions), then the diffuse
irradiance falling on a horizontal surface could be calculated as

π
2 2π

DHI = Rsky ⋅
∫∫ cos (θ ) ⋅ sin (θ ) dθ dϕ
0 0
(6.5)

where Rsky is the constant radiance of the isotropic sky. The term sin(θ) dθdφ comes from spherical
geometry (see Figure 6.5), and the cos(θ) term comes from the Lambert cosine law, which states
that the irradiance flux on a surface is proportional to the cosine of the angle of incidence. Since
sin(θ)cos(θ)  =  sin(2θ)/2, it is clear that the largest contribution to the diffuse irradiance is from
the annulus of skylight near 45° and falls monotonically and symmetrically to zero at 0° and 90°.
This means that the effective angle of incidence for diffuse radiation is 45°, which does not change
throughout the day for isotropic skies. Skylight is not isotropic, but contributions from all parts of
the sky are such that the effective incident angle for diffuse is never far from the nominal 45° and
changes little throughout the day (Michalsky et al. 2007).
Calibrating pyranometers using the shade–unshade technique eliminates problems with thermal
offsets since offsets for the global and diffuse measurements are approximately equal and are
eliminated by subtraction of the two components, as can be deduced from Equation 6.4. However,
when making DHI measurements with a single black detector thermopile pyranometer, even if
calibrated with the shade-unshade method, a correction must be made to subtract the thermal offset.
Having calibrated a reference pyranometer for diffuse irradiance measurements through the shade-
unshade method just described, it is possible to calibrate similar pyranometers for diffuse irradiance
by simply comparing to this standard under totally overcast, low-cloud conditions. Although not
strictly isotropic, the sky is uniform enough to yield a reliable calibration. Further, the thermal offset
for these cloud-cover conditions will be near zero since the effective sky temperature will be close to
the instrument temperature resulting in low values of net IR and, consequently, low thermal offsets.

6.5 VALUE OF ACCURATE DIFFUSE MEASUREMENTS


Because there are systematic errors with all pyranometers, the most accurate GHI values can be
obtained from summed DNI and DHI measurements (McArthur 2005). As with the example for
calculating diffuse from GHI and DNI, the uncertainty of GHI also can be estimated when based
Diffuse Irradiance 159

on measurements of DNI and DHI. Under clear-sky conditions used in Equation 6.2 with a properly
shaded pyranometer used for the diffuse measurement with a ±3 percent uncertainty, the calculated
GHI would be

GHI = 134 + 1, 000 ⋅ cos ( 30° ) = 1, 000  Wm −2  (6.6)

with an uncertainty of DHI and DNI summed in quadrature:

( 4 ) + ( 20 ⋅ 0.866 )
2 2
U 95 = = 18  Wm −2  (6.7)

The uncertainty (U95) in the GHI then is around ±2 percent, which is better than the ±3 percent
U95 from well-maintained GHI measurements. The reason for the improved uncertainty is that
most of the irradiance comes from the DNI contribution. Under cloudy conditions, the DNI
is zero, and the uncertainty in the GHI is ±3 percent, the same as the uncertainty in the DHI
measurement.
Because of the relationship between DHI, DNI, and GHI (Equation  4.1), a three-parameter
comparison is the best way to validate the accuracy of solar irradiance measurements. If rain or
soiling affects the measurements, then the three-component check can be used to confirm the
reliability of the measurements.
Accurate DHI measurements are also essential for the calibration of pyranometers using the
summation technique. This method allows the simultaneous calibrations of multiple pyranometers
as described in Chapter  13. Diffuse irradiance contributes to the plane of array (POA) available
to flat-plate solar collectors and is used in all performance calculations. Therefore, an accurate
performance analysis requires DHI measurements with the smallest possible U95. Because DHI is
much smaller (typically less than 20%) than the DNI contribution on sunny days, the uncertainty in
the DHI contribution to POA irradiance is often neglected. However, as the importance of precision
measurements increases for the analysis of solar energy conversion system performance, accurate
DHI data will become indispensable.

QUESTIONS
1. On a clear day, what are two of the several indirect sources of diffuse light on a horizontal
surface (the Sun is not a correct answer)?
2. What is the most accurate way to measure DHI?
3. Why is it possible to measure DHI values with a shaded pyranometer based on a single-
black detector that are less than a theoretical pure Rayleigh sky (no clouds or aerosols)
would produce?
4. What are the two sources of measurement error when using a fixed shade band to
measure DHI?
5. What is the major advantage of the black-and-white pyranometer in measurements of DHI?
6. On a clear day with some aerosols present, which is likely to be brighter: (1) a point in the
sky 90 degrees from the Sun in the same vertical plane as the Sun and the zenith, or (2) a
point near the horizon?
7. What is the most significant problem with using a photodiode pyranometer for DHI
measurements?
8. Under what circumstances is skylight isotropically distributed during the daylight hours?
160 Solar and Infrared Radiation Measurements

REFERENCES
Battles, F. J., L. Alados-Arbodelas, and F. J. Olmo. 1995. On shadow band correction methods for diffuse
irradiance measurements. Solar Energy 54:105–114.
Cess, R. D., T. Qian, and M. Sun. 2000. Consistency tests applied to the measurement of total, direct,
and diffuse shortwave radiation at the surface. Journal of Geophysical Research 105:881–887.
doi:10.1029/2000JD900402.
Drummond, A. J. 1956. On the measurement of sky radiation. Archives for Meteorology, Geophysics, and
Bioclimatology–Series B: Climatology, Environmental Meteorology, Radiation Research 7:413–436.
Gueymard, C. A., and D. R. Myers. 2009. Evaluation of conventional and high-performance routine solar
radiation measurements for improved solar resource, climatological trends, and radiative forcing. Solar
Energy 83:171–185.
Kudish, A. I., and E. G. Evseev. 2008. The  assessment of four different correction models applied to the
diffuse radiation measured with a shadow ring using global and normal beam radiation measurements
for Beer Sheva, Israel. Solar Energy 82:144–156.
LeBaron, B. A., J. J. Michalsky, and R. Perez. 1990. A new simplified procedure for correcting shadow band
data for all sky conditions. Solar Energy 44:249–256.
McArthur, L. J. B. 2005. Baseline Surface Radiation Network (BSRN) Operations Manual Version 2.1. WRCP-
121, WMO/TD-No.1274. World Meteorological Organization. https://epic.awi.de/id/eprint/30644/
(Accessed May 24, 2019).
Michalsky, J. J., R. Dolce, E. G. Dutton, M. Haeffelin, G. Major, J. A. Schlemmer, D. W. Slater, et  al.
2003.  Results from the first ARM diffuse horizontal shortwave irradiance comparison. Journal of
Geophysical Research 112:D16111. doi:10.1029/2007JD008651.
Michalsky, J. J., C. Gueymard, P. Kiedron, L. J. B. McArthur, R. Philipona, and T. Stoffel. 2007. A proposed
working standard for the measurement of diffuse horizontal shortwave irradiance. Journal of
Geophysical Research 112:D16112. doi:10.1029/2007JD008651.
Moon, P. and D. Spencer. 1942. Illumination from a non-uniform sky. Illuminating Engineering 37:707–726.
Muneer, T., and X. Zhang. 2002. A new method for correcting shadow band diffuse irradiance data. Journal
of Solar Energy Engineering 124:34–43.
Sengupta, M. A. Habte, C. Gueymard, S. Wilbert, D. Renne, and T. Stoffel. 2017. Best Practices Handbook
for the Collection and Use of Solar Resource Data for Solar Energy Applications, 2nd ed. NREL/
TP-5D00-68886. Golden, CO: National Renewable Energy Laboratory. https://www.nrel.gov/docs/
fy18osti/68886.pdf (Accessed May 24, 2019).
Stoffel, T., D. Renné, D. Myers, S. Wilcox, M. Sengupta, R. George, and C. Turchi. 2010. Concentrating
Solar Power: Best Practices Handbook for the Collection and Use of Solar Resource Data. NREL/
TP-550-47465. Golden, CO: National Renewable Energy Laboratory. https://www.nrel.gov/docs/
fy10osti/47465.pdf (Accessed May 24, 2019).
Wilbert, S., N. Geuder, M. Schwandt, B. Kraas, W. Jessen, R. Meyer, and B. Nouri. 2015. A report of IEA SHC
Task 46, Solar Resource Assessment and Forecasting, August 18, 2015. http://task46.iea-shc.org/Data/
Sites/1/publications/INSRSI_IEA-Task46B1_BestPractices-RSI_150819.pdf.
7 Solar Spectral Measurements

I shall without further ceremony acquaint you that in the year 1666, I procured me a triangular
glass prism, to try therewith the celebrated phenomena of colors.
Sir Isaac Newton (1642–1727)

7.1 INTRODUCTION
Solar radiation is not uniformly distributed in wavelength. The spectral distribution of solar radia-
tion as it arrives at the top of the atmosphere (TOA) is approximately a blackbody spectrum with
an effective temperature of 5772 Kelvin (K) that is modified by selective absorption and emissions
in the Sun’s own atmosphere (NASA 2018). This radiation is further and significantly modified by
scattering and absorption processes as it passes through the Earth’s atmosphere before it reaches
the surface.
This  chapter begins with a discussion of the extraterrestrial solar spectrum and its modest
variability with the 11-year solar cycle. Its variability with Sun–Earth distance is covered in
Chapter 2. Clear-sky interactions of solar radiation with the atmosphere also are discussed in this
chapter including molecular, that is Rayleigh, scattering, aerosol scattering and absorption, and
absorption by atmospheric gases. Scattering and absorption by clouds is a very complex topic
and is not addressed in this book. The textbook by Liou (1992) provides an introduction to this
subject.
Some useful broadband spectral measurements including photometric, photosynthetically
active radiation (PAR), and ultraviolet (UV) photometry are covered in this chapter. A discussion
of narrowband photometry follows with a focus on its use in retrieving aerosol optical depth and
precipitable water vapor (the most important clear-sky constituents affecting the spectral distribu-
tion of solar radiation at the Earth’s surface). Spectroradiometry also is finding use in analysis of
the performance of photovoltaic (PV) systems. Instruments for making these measurements then
are described with some advice on the specifications to consider if purchasing these instruments.
The chapter concludes with brief sections on moderate spectral resolution spectrometry and moder-
ate resolution spectral irradiance models.

7.2 THE EXTRATERRESTRIAL SOLAR SPECTRUM


The  solar spectrum covers a wide wavelength range; however, for practical purposes virtually
all of the Sun’s energy that reaches the Earth’s surface lies between the wavelengths of 280 and
4,000  nanometers (nm). Figure  7.1 is a plot of the spectral irradiance in watts per square meter
per nanometer versus wavelength in nanometers (Gueymard 2018). The  integral of this spectral
distribution over all wavelengths, if measured just above the Earth’s atmosphere for one 11-year
solar cycle, yields 1,361 watts per square meter (Wm−2) at the mean distance between the Sun and
the Earth (149,597,887 kilometers (km) or 92,955,818 miles). This mean distance is defined as one
astronomical unit (1 AU). Fröhlich (2006) summarized how extraterrestrial irradiance has varied
over the last two 11-year solar cycles. The maximum to minimum solar cycle variation averages
about 0.1 percent of the total energy with short-term excursions about three times this amount.

161
162 Solar and Infrared Radiation Measurements

200 400 600 800 1,000 1,200 1,400 1,600 1,800 2,000 2,200 2,400 2,600 2,800 3,000 3,200 3,400 3,600 3,800 4,000

FIGURE  7.1 Extraterrestrial solar spectral irradiance distribution at mean Sun–Earth distance. (After
Gueymard, C.A., Solar Energy, 169, 434–440, 2018.)

The  total irradiance monitor (TIM) on the Solar Radiation and Climate Experiment (SORCE)
satellite (Rottman 2005) measured a lower mean value for the integrated spectral irradiance at 1 AU
of 1,361 W/m2 than the long-accepted value of 1366 W/m2 (Fröhlich 2006). Fehlmann et al. (2012)
resolved the issue and found that the earlier value based on satellite radiometers traceable to the
World Radiometric Reference (WRR) was too high by about 0.34 percent because of an underesti-
mate of the stray light in radiometers that were used to establish the WRR. Therefore, 1,361 Wm−2
is very close to the currently accepted solar irradiance at 1 AU.
Harder et al. (2009) used the solar irradiance monitor (SIM) on the SORCE satellite to inves-
tigate the variations in portions of the solar spectrum. Separately assessing the variability in the
spectral bands from 201–300, 300–400, 400–691, 691–972, 972–1,630, and 1,630–2,423  nm,
they found maximum changes of about 0.6 Wm−2 with the largest changes in the shortest wave-
length bands. The  measurements spanned only 4 years or about one-half of the maximum
change in total irradiance expected over a solar cycle; however, even twice this change would
amount to a small fraction of 1 Wm−2 by the time the radiation reaches the surface of the Earth.
Consequently, it is reasonable to assume that the Sun’s inherent variability is lower than can
reasonably be detected at the surface given the scattering and absorption that occurs within the
Earth’s atmosphere. These interactions with the Earth’s atmosphere are discussed in the next
section.
Solar Spectral Measurements 163

7.3 ATMOSPHERIC INTERACTIONS


When solar radiation passes through the atmosphere, complex interactions occur depending on the
wavelength of the radiation and the composition of the atmosphere at the time. More specifically,
solar radiation either passes unscathed to the surface, or it is scattered or absorbed by molecules,
aerosols, cloud water droplets, and ice crystals.

7.3.1 RayleigH sCatteRing


Molecular scattering is the elastic scattering of solar radiation by the gaseous molecules in the atmo-
sphere first explained by Lord Rayleigh (1871). Many historical papers have been written describing
the scattering of solar radiation by atmospheric gases and it continues to be the subject of ongoing
research (Eberhard 2010). Bodhaine et al. (1999) thoroughly examined the problem of calculating
the molecular scattering optical depth as a function of wavelength. Optical depth, in general, is a
measure of the wavelength dependent extinction (by scattering or absorption) that occurs as a beam
of radiation propagates through a medium. It can be defined using the Beer-Lambert law:

I (λ ) / I0 (λ ) = e ( )
−τ λ m
(7.1)

where I0 is the strength of the source (e.g., the spectral solar irradiance) before it enters the atmo-
sphere, I is the strength after it has passed through the atmosphere to the surface, τ is the optical
depth in a vertical path, m is the amount of atmosphere traversed relative to the vertical path, and λ
is the wavelength of interest. A good approximation to optical depth by Rayleigh scattering is given
by Hansen and Travis (1974):

P
τR =
P0
(
0.008569λ −4 1 + 0.0113λ −2 + 0.00013λ −4 ) (7.2)

where wavelength λ is in µm, P is the pressure at the measurement site within the Earth’s atmo-
sphere in kilopascals, and P0 is the standard pressure at sea level equal to 101.325  kilopascals
(100 kilopascals = 1000 millibars). Evaluation of Equation (7.2) leads to the conclusion that Rayleigh
scattering is most effective in the shortest wavelengths of the solar spectrum and decreases dramati-
cally at longer wavelengths (approximately as λ–4). There is less Rayleigh scattering at higher eleva-
tion because the atmospheric pressure is lower and hence there are fewer molecules to scatter the
incident radiation.

7.3.2 aeRosol sCatteRing and absoRPtion


Aerosols are always present in the atmosphere. They manifest themselves as the haze often noticed
when looking toward the horizon. Generally, aerosols are concentrated near the surface but may
appear in layers especially when transported into a region from long distances. When volcanic
eruptions occur that are strong enough to inject dust and gases into the stratosphere, such as Mount
Pinatubo in 1991, a stratospheric aerosol layer may persist for years from the sulfur gases that are
chemically transformed into aerosols. The dust particles from an eruption, however, are large and
therefore are removed from the atmosphere quickly because of gravitational settling. Figure  7.2
is the long-term record of transmission from Mauna Loa Observatory using the technique devel-
oped by Ellis and Pueschel (1971). The largest excursions are caused by the three named volcanic
eruptions (Agung, El Chichon, and Pinatubo). Lower transmission persists for years after the events.
164 Solar and Infrared Radiation Measurements

0.95 ●
●●●
● ● ● ●
●● ● ● ● ●
● ● ● ● ● ● ● ● ●● ●●
● ●● ● ●●● ● ●
●● ● ●●
● ● ● ●●
●● ● ● ●
Monthly−Averaged, Clear−Sky Transmission

● ● ● ● ● ● ●●● ●● ●
● ● ●●● ●● ● ● ● ● ●● ●● ● ● ● ● ● ●
● ● ● ●
●●● ●● ●● ● ●
● ● ● ●● ● ● ● ●● ●● ● ● ● ●● ● ● ● ● ●● ● ● ●● ● ●●● ●● ● ●
● ●● ● ● ● ● ● ● ●● ● ●●● ● ●● ● ● ● ● ● ● ● ●● ● ● ● ●●●● ●● ● ●● ● ● ●
● ●● ●● ●●●
● ●●
● ●● ● ● ● ●● ●
● ●
● ● ●
● ● ●● ● ● ● ● ●●●
● ● ● ● ●●●● ●● ● ● ●
● ● ● ●● ●


● ●● ● ● ● ● ● ●

● ●●
● ● ●● ● ● ● ●
● ● ● ●●
●●●

●●
●● ● ●●● ● ● ●● ●
●● ●●
● ●● ●


●● ●●


●●
● ●● ● ● ● ● ●●●●●
● ● ● ● ● ● ● ● ● ● ●● ●● ●● ● ● ● ● ● ●● ●● ● ●● ●●
● ● ● ●● ● ● ● ● ●●
● ●●
●●
● ● ● ●● ●● ● ● ● ●● ● ●● ● ●● ● ● ● ● ● ●● ● ●● ● ● ● ●●
● ● ● ●● ● ● ● ●● ● ●● ● ●● ●●
● ● ● ●●
●● ● ●
●● ● ● ● ●
●● ● ● ● ● ● ● ●●
●● ● ● ● ● ●● ● ● ● ● ● ●●
● ● ●
● ● ● ● ● ●
● ● ● ●●●● ● ● ●● ●●●

● ●● ● ●●●● ● ● ● ●
● ● ● ● ● ● ●●
● ●●● ●

● ● ● ● ● ● ● ● ● ●
●● ●● ●
●● ●● ● ●● ● ●
● ● ●●


● ● ● ● ● ● ● ● ● ● ● ●
● ●● ●

● ● ● ● ● ● ● ●
● ●
● ● ● ●● ● ● ●● ●●
● ● ● ● ●● ●● ●●● ● ● ● ● ●
● ● ● ● ● ● ●● ●
● ● ● ● ● ●●
● ● ● ● ● ● ●● ● ●
● ●● ●● ● ● ● ● ●
● ●
●●● ● ● ● ●

● ● ●● ●
● ● ● ● ● ● ● ● ●●
● ●●
● ● ●
● ●
●●
● ●
● ●
● ● ● ● ●

● ●
●●
●●

● ●
● ● ●
0.90

● ●

Agung El Chichon Pinatubo


● ●
●● ●


● ●

●●





● ●

● ●


0.85

● ●●●


● ●


0.80

1960 1970 1980 1990 2000 2010


Year

FIGURE 7.2 Transmission record from Mauna Loa Observatory. The large dips in transmission are associ-
ated with stratospheric aerosols from three volcanoes. Note that the transmission takes 3–5 years to recover
after the initial eruptions.

Aerosols can scatter and absorb solar radiation depending on their chemical makeup and on the
wavelength of the incident solar radiation. The probability of a photon scattering from an aerosol
is called the single scattering albedo, ϖ0. Often ϖ0 exceeds 0.9 at visible wavelengths for typical
continental aerosols. For sea salt aerosols, ω0 is very close to 1.0 at visible wavelengths. The term
co-albedo is used for the quantity (1 – ϖ0) and is the probability of the photon being absorbed upon
encountering an aerosol. Therefore, sea salt aerosols would have a co-albedo of almost zero in the
visible spectral region. Another quantity that describes the scattering of aerosols is the asymmetry
parameter, often denoted g. The probability of scattering in any given direction is given, approxi-
mately, by the Henyey–Greenstein function (Henyey and Greenstein 1941):

1 1− g2
p ( cos θ ) = (7.3)
( )
3
2 1 + g 2 − 2 g cos θ 2

where θ is 0 degrees in the direction of photon propagation and 180 degrees in the opposite direc-
tion. This probability function exhibits these properties:

∫ p ( cosθ ) d ( cosθ ) = 1
−1
(7.4)

∫ p ( cosθ ) cosθ d ( cosθ ) = g


−1
(7.5)
Solar Spectral Measurements 165

The  larger g is, the more the scattering into the forward hemisphere. A  g value of zero would
imply isotropic scatter or equal probability of scattering in all directions. Typical values of g are
between 0.5 and 0.7 for continental aerosols. Sea salt aerosols tend to be large and have large values
of asymmetry parameter, while anthropogenic pollution aerosols are small with lower asymmetry
parameters. The use of spectral models that use these two parameters ϖ0 and g as inputs is dis-
cussed in Section 7.6.2.

7.3.3 gas absoRPtion


The dominant gas absorber affecting the spectral distribution of the incident solar irradiance in the
solar spectrum is water vapor with several major absorption bands in the near-infrared (near-IR).
For  example, Figure  2.15 in Chapter  2 indicates the strong water vapor absorption band in the
global horizontal irradiance (GHI) spectrum centered near 940 nm. This measurement is of par-
ticular importance to many solar PV collectors since the peak response of crystalline silicon PV
is near the strong water vapor band around 940 nm. Ozone is another important absorber in the
atmosphere. However, its strongest absorption is below 380  nm in the Hartley–Huggins bands.
The Chappuis ozone band is broad and has modest absorption centered near 610 nm. The Wulf
ozone bands in the near-IR are even less absorbing. Other gases such as oxygen (O2) and carbon
dioxide (CO2) are ever-present but less significant in the sense that they do not  remove a large
fraction of the total solar radiation. Nitrogen dioxide (NO2) is a significant gas when air pollu-
tion concentrations are high. This gas produces a reddish-brown tinge to the skylight by prefer-
entially absorbing blue solar radiation. Motor vehicles produce most of the NO2 and are also the
main source of ozone production in cities in which this molecule exceeds the U.S. Environmental
Protection Agency (EPA) standards.
Optical depths associated with ozone absorption τozone may be calculated using

τ ozone = ηozone ⋅ α ozone (7.6)

where ηozone is the ozone abundance, and αozone is the ozone absorption coefficient. Typically, for
mid-latitudes, the ozone column (the amount of ozone present from the surface to TOA) is approx-
imately 300 Dobson units (DU) but varies typically between 200 and 400 DU. A Dobson unit
of ozone is equal to 2.687 × 1016 molecules per cubic centimeter. This is 1/1000 of Loschmidt’s
number, or the number of molecules in a cubic centimeter of a gas under standard tempera-
ture and pressure. A typical value of 300 DU of ozone at the standard temperature of 273.15 K
and standard pressure of 101.325  kilopascals would form a column of ozone 0.3  centimeters
(cm) high. Current ozone data from the National Aeronautics and Space Administration (NASA)
Ozone Monitoring Instrument (OMI) satellite are posted on the NASA website ftp://toms.gsfc.
nasa.gov/pub/omi/data/Level3e/ozone/ 2  days after they are acquired. A  convenient website
to get interpolated ozone for a particular site is hosted by National Oceanic and Atmospheric
Administration (NOAA): https://www.esrl.noaa.gov/gmd/grad/neubrew/SatO3DataTimeSeries.
jsp. Ozone absorption coefficients αozone in units of molecules−1 at 1 nm resolution between 407
and 1086 nm for the visible and near-IR (Chappuis and Wulf bands, respectively) based on Shettle
and Anderson (1995) are given in Table 7.1. As an example, using Equation (7.6) 300 DU will
yield a vertical optical depth at 602 nm of

τ ozone = 300 ⋅ 2.687 ⋅1016 ⋅ 5.190 ⋅10 −21 = 0.04184 (7.7)


166 Solar and Infrared Radiation Measurements

TABLE 7.1
Ozone Absorption Coefficients (molecules−1) at the Given Wavelengths (nm)
Absorption Absorption Absorption Absorption
Wavelength Coefficient Wavelength Coefficient Wavelength Coefficient Wavelength Coefficient
407 2.001e-24 577 4.706e-21 747 4.262e-22 917 1.862e-23
408 4.739e-24 578 4.665e-21 748 4.216e-22 918 1.795e-23
409 6.727e-24 579 4.623e-21 749 4.118e-22 919 1.737e-23
410 8.178e-24 580 4.576e-21 750 3.947e-22 920 1.682e-23
411 8.984e-24 581 4.532e-21 751 3.740e-22 921 1.657e-23
412 8.255e-24 582 4.490e-21 752 3.544e-22 922 1.650e-23
413 7.163e-24 583 4.452e-21 753 3.337e-22 923 1.633e-23
414 7.526e-24 584 4.423e-21 754 3.139e-22 924 1.626e-23
415 9.642e-24 585 4.401e-21 755 2.977e-22 925 1.637e-23
416 1.233e-23 586 4.385e-21 756 2.827e-22 926 1.646e-23
417 1.442e-23 587 4.382e-21 757 2.709e-22 927 1.642e-23
418 1.553e-23 588 4.382e-21 758 2.612e-22 928 1.628e-23
419 1.569e-23 589 4.395e-21 759 2.536e-22 929 1.645e-23
420 1.561e-23 590 4.414e-21 760 2.473e-22 930 1.651e-23
421 1.655e-23 591 4.456e-21 761 2.421e-22 931 1.669e-23
422 1.954e-23 592 4.507e-21 762 2.379e-22 932 1.662e-23
423 2.466e-23 593 4.575e-21 763 2.607e-22 933 1.622e-23
424 3.148e-23 594 4.659e-21 764 2.586e-22 934 1.631e-23
425 3.828e-23 595 4.742e-21 765 2.552e-22 935 1.703e-23
426 4.229e-23 596 4.838e-21 766 2.522e-22 936 1.834e-23
427 4.203e-23 597 4.931e-21 767 2.492e-22 937 2.025e-23
428 3.928e-23 598 5.013e-21 768 2.503e-22 938 2.284e-23
429 3.722e-23 599 5.084e-21 769 2.513e-22 939 2.564e-23
430 3.707e-23 600 5.139e-21 770 2.515e-22 940 2.900e-23
431 4.002e-23 601 5.173e-21 771 2.570e-22 941 3.277e-23
432 4.690e-23 602 5.190e-21 772 2.626e-22 942 3.623e-23
433 5.316e-23 603 5.184e-21 773 2.696e-22 943 3.913e-23
434 5.506e-23 604 5.156e-21 774 2.795e-22 944 4.049e-23
435 5.465e-23 605 5.115e-21 775 2.924e-22 945 4.019e-23
436 5.501e-23 606 5.058e-21 776 3.027e-22 946 3.848e-23
437 5.812e-23 607 4.983e-21 777 3.092e-22 947 3.493e-23
438 6.625e-23 608 4.906e-21 778 3.141e-22 948 3.135e-23
439 7.903e-23 609 4.821e-21 779 3.155e-22 949 2.740e-23
440 9.554e-23 610 4.730e-21 780 3.119e-22 950 2.435e-23
441 1.110e-22 611 4.641e-21 781 3.035e-22 951 2.129e-23
442 1.217e-22 612 4.554e-21 782 2.935e-22 952 1.938e-23
443 1.299e-22 613 4.475e-21 783 2.791e-22 953 1.683e-23
444 1.299e-22 614 4.395e-21 784 2.642e-22 954 1.512e-23
445 1.237e-22 615 4.321e-21 785 2.478e-22 955 1.512e-23
446 1.158e-22 616 4.251e-21 786 2.359e-22 956 1.436e-23
447 1.120e-22 617 4.181e-21 787 2.266e-22 957 1.368e-23
448 1.150e-22 618 4.122e-21 788 2.169e-22 958 1.271e-23
449 1.260e-22 619 4.062e-21 789 2.068e-22 959 1.161e-23
450 1.391e-22 620 4.011e-21 790 1.999e-22 960 1.066e-23
451 1.478e-22 621 3.963e-21 791 1.919e-22 961 9.900e-24
452 1.514e-22 622 3.911e-21 792 1.833e-22 962 9.293e-24
(Continued)
Solar Spectral Measurements 167

TABLE 7.1 (Continued)


Ozone Absorption Coefficients (molecules−1) at the Given Wavelengths (nm)
Absorption Absorption Absorption Absorption
Wavelength Coefficient Wavelength Coefficient Wavelength Coefficient Wavelength Coefficient
453 1.501e-22 623 3.866e-21 793 1.764e-22 963 8.866e-24
454 1.504e-22 624 3.821e-21 794 1.696e-22 964 8.599e-24
455 1.572e-22 625 3.774e-21 795 1.632e-22 965 8.295e-24
456 1.735e-22 626 3.721e-21 796 1.597e-22 966 7.920e-24
457 2.001e-22 627 3.673e-21 797 1.571e-22 967 7.137e-24
458 2.326e-22 628 3.620e-21 798 1.553e-22 968 6.496e-24
459 2.658e-22 629 3.563e-21 799 1.522e-22 969 6.058e-24
460 2.986e-22 630 3.510e-21 800 1.489e-22 970 5.446e-24
461 3.224e-22 631 3.457e-21 801 1.477e-22 971 5.003e-24
462 3.326e-22 632 3.403e-21 802 1.472e-22 972 4.651e-24
463 3.312e-22 633 3.347e-21 803 1.469e-22 973 4.341e-24
464 3.135e-22 634 3.290e-21 804 1.473e-22 974 4.207e-24
465 2.934e-22 635 3.231e-21 805 1.498e-22 975 3.977e-24
466 2.834e-22 636 3.172e-21 806 1.525e-22 976 3.942e-24
467 2.935e-22 637 3.111e-21 807 1.569e-22 977 3.827e-24
468 3.128e-22 638 3.048e-21 808 1.638e-22 978 3.777e-24
469 3.300e-22 639 2.989e-21 809 1.713e-22 979 3.701e-24
470 3.475e-22 640 2.926e-21 810 1.795e-22 980 3.676e-24
471 3.558e-22 641 2.867e-21 811 1.866e-22 981 3.841e-24
472 3.601e-22 642 2.814e-21 812 1.937e-22 982 4.032e-24
473 3.594e-22 643 2.761e-21 813 2.009e-22 983 4.367e-24
474 3.707e-22 644 2.710e-21 814 2.072e-22 984 4.989e-24
475 3.990e-22 645 2.662e-21 815 2.114e-22 985 5.915e-24
476 4.457e-22 646 2.616e-21 816 2.135e-22 986 7.191e-24
477 4.997e-22 647 2.569e-21 817 2.127e-22 987 8.732e-24
478 5.624e-22 648 2.527e-21 818 2.089e-22 988 1.041e-23
479 6.256e-22 649 2.487e-21 819 2.028e-22 989 1.246e-23
480 6.932e-22 650 2.445e-21 820 1.946e-22 990 1.802e-23
481 7.408e-22 651 2.405e-21 821 1.839e-22 991 3.049e-23
482 7.730e-22 652 2.366e-21 822 1.717e-22 992 2.713e-23
483 7.779e-22 653 2.324e-21 823 1.591e-22 993 2.195e-23
484 7.552e-22 654 2.283e-21 824 1.471e-22 994 2.195e-23
485 7.243e-22 655 2.242e-21 825 1.359e-22 995 2.102e-23
486 7.026e-22 656 2.201e-21 826 1.252e-22 996 1.724e-23
487 6.901e-22 657 2.161e-21 827 1.154e-22 997 1.345e-23
488 6.963e-22 658 2.119e-21 828 1.071e-22 998 1.134e-23
489 7.225e-22 659 2.083e-21 829 1.001e-22 999 9.501e-24
490 7.590e-22 660 2.046e-21 830 9.396e-23 1000 7.456e-24
491 7.815e-22 661 2.009e-21 831 8.855e-23 1001 7.291e-24
492 7.963e-22 662 1.973e-21 832 8.463e-23 1002 5.993e-24
493 8.033e-22 663 1.935e-21 833 8.143e-23 1003 5.603e-24
494 8.198e-22 664 1.899e-21 834 7.848e-23 1004 5.164e-24
495 8.383e-22 665 1.858e-21 835 7.679e-23 1005 4.140e-24
496 8.752e-22 666 1.817e-21 836 7.605e-23 1006 4.452e-24
497 9.275e-22 667 1.774e-21 837 7.514e-23 1007 3.733e-24
498 9.929e-22 668 1.732e-21 838 7.446e-23 1008 3.303e-24
(Continued)
168 Solar and Infrared Radiation Measurements

TABLE 7.1 (Continued)


Ozone Absorption Coefficients (molecules−1) at the Given Wavelengths (nm)
Absorption Absorption Absorption Absorption
Wavelength Coefficient Wavelength Coefficient Wavelength Coefficient Wavelength Coefficient
499 1.074e-21 669 1.692e-21 839 7.426e-23 1009 3.501e-24
500 1.170e-21 670 1.653e-21 840 7.458e-23 1010 2.965e-24
501 1.269e-21 671 1.614e-21 841 7.562e-23 1011 2.673e-24
502 1.367e-21 672 1.575e-21 842 7.712e-23 1012 2.804e-24
503 1.454e-21 673 1.537e-21 843 7.874e-23 1013 2.559e-24
504 1.517e-21 674 1.501e-21 844 8.120e-23 1014 2.322e-24
505 1.561e-21 675 1.470e-21 845 8.523e-23 1015 2.200e-24
506 1.570e-21 676 1.437e-21 846 9.008e-23 1016 2.250e-24
507 1.557e-21 677 1.404e-21 847 9.562e-23 1017 2.078e-24
508 1.533e-21 678 1.373e-21 848 1.024e-22 1018 1.892e-24
509 1.500e-21 679 1.347e-21 849 1.105e-22 1019 1.889e-24
510 1.480e-21 680 1.322e-21 850 1.196e-22 1020 1.808e-24
511 1.470e-21 681 1.300e-21 851 1.290e-22 1021 1.631e-24
512 1.491e-21 682 1.278e-21 852 1.368e-22 1022 1.532e-24
513 1.522e-21 683 1.260e-21 853 1.412e-22 1023 1.452e-24
514 1.559e-21 684 1.239e-21 854 1.415e-22 1024 1.463e-24
515 1.594e-21 685 1.217e-21 855 1.388e-22 1025 1.406e-24
516 1.628e-21 686 1.191e-21 856 1.347e-22 1026 1.329e-24
517 1.661e-21 687 1.166e-21 857 1.298e-22 1027 1.351e-24
518 1.691e-21 688 1.134e-21 858 1.243e-22 1028 1.402e-24
519 1.723e-21 689 1.101e-21 859 1.176e-22 1029 1.384e-24
520 1.761e-21 690 1.071e-21 860 1.091e-22 1030 1.420e-24
521 1.817e-21 691 1.040e-21 861 1.002e-22 1031 1.410e-24
522 1.877e-21 692 1.009e-21 862 9.202e-23 1032 1.463e-24
523 1.952e-21 693 9.845e-22 863 8.438e-23 1033 1.525e-24
524 2.041e-21 694 9.576e-22 864 7.720e-23 1034 1.564e-24
525 2.137e-21 695 9.294e-22 865 7.025e-23 1035 1.662e-24
526 2.233e-21 696 9.030e-22 866 6.419e-23 1036 1.764e-24
527 2.337e-21 697 8.758e-22 867 5.939e-23 1037 1.876e-24
528 2.434e-21 698 8.524e-22 868 5.533e-23 1038 1.944e-24
529 2.527e-21 699 8.303e-22 869 5.203e-23 1039 2.073e-24
530 2.609e-21 700 8.094e-22 870 4.958e-23 1040 2.282e-24
531 2.678e-21 701 7.904e-22 871 4.745e-23 1041 2.512e-24
532 2.731e-21 702 7.757e-22 872 4.475e-23 1042 2.745e-24
533 2.764e-21 703 7.596e-22 873 4.239e-23 1043 3.104e-24
534 2.782e-21 704 7.449e-22 874 4.075e-23 1044 3.730e-24
535 2.784e-21 705 7.322e-22 875 3.960e-23 1045 5.270e-24
536 2.778e-21 706 7.205e-22 876 3.899e-23 1046 7.263e-24
537 2.777e-21 707 7.100e-22 877 3.852e-23 1047 5.683e-24
538 2.791e-21 708 7.000e-22 878 3.847e-23 1048 5.129e-24
539 2.820e-21 709 6.904e-22 879 3.879e-23 1049 5.166e-24
540 2.860e-21 710 6.820e-22 880 3.928e-23 1050 4.980e-24
541 2.908e-21 711 6.753e-22 881 3.991e-23 1051 4.406e-24
542 2.968e-21 712 6.690e-22 882 4.102e-23 1052 3.500e-24
543 3.029e-21 713 6.640e-22 883 4.246e-23 1053 3.188e-24
544 3.083e-21 714 6.626e-22 884 4.401e-23 1054 2.659e-24
(Continued)
Solar Spectral Measurements 169

TABLE 7.1 (Continued)


Ozone Absorption Coefficients (molecules−1) at the Given Wavelengths (nm)
Absorption Absorption Absorption Absorption
Wavelength Coefficient Wavelength Coefficient Wavelength Coefficient Wavelength Coefficient
545 3.135e-21 715 6.566e-22 885 4.587e-23 1055 2.647e-24
546 3.171e-21 716 6.487e-22 886 4.769e-23 1056 2.225e-24
547 3.209e-21 717 6.349e-22 887 4.887e-23 1057 2.262e-24
548 3.244e-21 718 6.186e-22 888 4.974e-23 1058 1.937e-24
549 3.268e-21 719 5.985e-22 889 5.038e-23 1059 1.926e-24
550 3.297e-21 720 5.775e-22 890 5.035e-23 1060 1.834e-24
551 3.329e-21 721 5.569e-22 891 4.985e-23 1061 1.705e-24
552 3.365e-21 722 5.378e-22 892 4.981e-23 1062 1.849e-24
553 3.406e-21 723 5.230e-22 893 5.100e-23 1063 1.852e-24
554 3.451e-21 724 5.070e-22 894 5.313e-23 1064 1.987e-24
555 3.505e-21 725 4.946e-22 895 5.597e-23 1065 2.347e-24
556 3.568e-21 726 4.801e-22 896 5.900e-23 1066 2.290e-24
557 3.641e-21 727 4.699e-22 897 6.117e-23 1067 1.889e-24
558 3.722e-21 728 4.567e-22 898 6.197e-23 1068 1.746e-24
559 3.818e-21 729 4.430e-22 899 6.143e-23 1069 1.879e-24
560 3.910e-21 730 4.307e-22 900 5.958e-23 1070 1.671e-24
561 4.010e-21 731 4.213e-22 901 5.685e-23 1071 1.422e-24
562 4.106e-21 732 4.141e-22 902 5.385e-23 1072 1.280e-24
563 4.195e-21 733 4.073e-22 903 5.068e-23 1073 1.116e-24
564 4.275e-21 734 4.035e-22 904 4.759e-23 1074 1.073e-24
565 4.343e-21 735 4.005e-22 905 4.410e-23 1075 1.104e-24
566 4.408e-21 736 3.988e-22 906 3.999e-23 1076 9.390e-25
567 4.464e-21 737 3.970e-22 907 3.670e-23 1077 1.014e-24
568 4.519e-21 738 3.948e-22 908 3.415e-23 1078 8.850e-25
569 4.579e-21 739 3.956e-22 909 3.168e-23 1079 8.400e-25
570 4.634e-21 740 3.966e-22 910 2.922e-23 1080 7.620e-25
571 4.683e-21 741 4.006e-22 911 2.683e-23 1081 6.970e-25
572 4.730e-21 742 4.063e-22 912 2.488e-23 1082 7.090e-25
573 4.756e-21 743 4.108e-22 913 2.329e-23 1083 7.230e-25
574 4.766e-21 744 4.176e-22 914 2.143e-23 1084 6.850e-25
575 4.763e-21 745 4.221e-22 915 2.014e-23 1085 6.620e-25
576 4.739e-21 746 4.265e-22 916 1.945e-23 1086 7.860e-25

7.3.4 tRansmission of tHe atmosPHeRe


Figure 7.3 illustrates the transmission of direct normal (beam) irradiance (DNI) for the parameters
given in the legend. Superimposed on the plot is the spectral irradiance (blue line) for the given
trace species with the sun in the zenith position at 1 AU. Rayleigh scattering is a smooth function of
wavelength and the larger dips (and many of the smaller ones) that depart from this smooth function
result from mixed gas absorption by O2 and CO2. Water vapor (red) and ozone (green) absorption
bands are plotted separately. Even a modest aerosol optical depth (dashed line) of 0.15 at 500 nm
with a wavelength dependence given by

τ = βλ −α (7.8)
170 Solar and Infrared Radiation Measurements

FIGURE 7.3 Direct solar transmission of major components of the clear-sky atmosphere (left axis); superim-
posed on the transmission spectra is the spectral irradiance calculated for these conditions (right axis).

where β  =  0.06092, and α  =  1.3 gives a reduced transmission that is typically the second most
important contributor to solar radiation extinction after Rayleigh scattering. This condition is true
for clear skies because the extinction from aerosols is typically larger in the mid-visible wavelengths
where solar irradiance is highest. Some spectral absorption features in the spectral irradiance curve
are not  in the transmission spectra. These extra absorption features are from the extraterrestrial
solar spectrum and occur in the outer layers of the Sun’s atmosphere (see Figure 7.1).

7.4 BROAD FILTER RADIOMETRY


There are many reasons to measure isolated portions of the solar spectrum with broad filters (100 nm
wide or wider). For example, broad filter measurements of red, green, and blue light are used for
specifying color for both commercial and artistic reasons. Broad filter measurements of the spectral
distribution of sunlight can be used for calculating the efficiency of spectrally dependent PV devices
(Michalsky and Kleckner 1984). Although not the preferred method, broad filters have been used
with a pyrheliometer to approximately measure aerosol optical depths (Dutton and Christy 1992).
This section provides a discussion of some common types of broad spectral measurements with
solar applications.

7.4.1 PHotometRy
Photometry is the measurement of light that the human eye can sense. Although there have
been slight modifications to the definition, the Commission Internationale de l’Eclairage (CIE)
defined the basic luminosity function for daytime (photopic) vision for the average human eye
in 1924 (CIE, 1926). Figure 7.4 illustrates how eye response varies with wavelength with a peak
near 555 nm. Scotopic response is the wavelength response of the eye in darkened conditions.
The response distribution with wavelength is similar but more skewed, and it shifts to shorter
wavelengths with the peak response near 507 nm. Since the interest is mostly in daylight appli-
cations, the focus will be only on the photopic response that most photometers are designed
to measure.
Solar Spectral Measurements 171

FIGURE 7.4 The typical response of the human eye for sunlit conditions as specified by the CIE. (From
https://en.wikipedia.org/wiki/File:Srgbspectrum.png.)

To convert any spectral irradiance, such as sunlight, into illuminance L, a very good approxima-
tion (CIE, 1926) is

770


L = 683 y ( λ ) I ( λ ) d λ
380
(7.9)

where y is the photopic function plotted in Figure 7.4, I(λ) is the spectral irradiance in W/(m2-nm),
and L has units of lumens/m2. (Note: 1 watt at 555 nm equals 683 lumens based on the peak photopic
response of the human eye.)
Manufacturers of photometers try to match the photopic function as closely as possible using
combinations of filters and detector responses. The  success with which this is achieved varies;
an example is given in Figure 7.5 from the tutorial offered on UDT Instrument’s website (www.
gamma-sci.com/photometry.pdf). When searching for a photometer, make sure that the search is for
a photometer with a photopic response since the term photometer is used for many types of instru-
ments that do not attempt to match the spectral response of the eye in daylight. If the photometer
is to be used outdoors, make certain that it is designed to be waterproof and rugged. Although
manufacturers try to match the CIE response curve, they all fail to some extent, so look for a reason-
able photopic match. If the goal is to measure solar radiation for daylighting purposes, for example,
performing the integral in Equation (7.9) using a modeled clear-sky irradiance first with the phot-
opic function y and then using the manufacturer’s estimate of the photopic function, for example y ′,
a bias can be calculated. Most daylight photometers will measure total horizontal illuminance. To
do this well, a respectable cosine response for the instrument is required; that is, the signal should
172 Solar and Infrared Radiation Measurements

FIGURE 7.5 An example of one manufacturer’s (Macam) attempt to match the CIE photopic response using
filters. (From Austin, R., The Guide to Photometry, UDT Instruments, San Diego, CA, n.d. With permission.)

fall off approximately as the cosine of the angle of incidence. If the response is within 10 percent
of the normalized response (actual response as a function of incident angle divided by cosine of the
incident angle) out to about 80 degrees incidence, then it can be considered acceptable. The left side
of Figure 7.6 illustrates an acceptable cosine response for an actual instrument. Some manufacturers
may illustrate cosine response as shown on the right side of Figure 7.6, but without the normalized
cosine response plotted for comparison. These plots may look approximately correct but should be
plotted normalized as in the left half of Figure 7.6 for proper evaluation. Other factors to consider
are instrument stability (responsivity changes are less than 3% per year) and absolute uncertainty
(better than 5%). A final note on cost: It was possible to purchase a good photopic photometer for
outdoor use for a little over US$500 in 2018.

FIGURE 7.6 An example of the cosine response of a good receiver compared with an exact cosine response
plotted normalized and unnormalized.
Solar Spectral Measurements 173

7.4.2 PHotosyntHetiCally aCtive Radiation (PaR)


Plants use PAR to transform water and carbon dioxide into sugars and oxygen through photosyn-
thesis. PAR is measured with sensors that filter light to accept radiation between about 400 and
700 nm. Plants mostly use the light near either end of this wavelength range for photosynthesis and
reflect more of the green light, thus giving plants their color. Photosynthesis is a quantum process.
In other words, photons cause the reactions that occur in plants; therefore, it would seem logical to
define PAR in terms of photons per unit time per unit area. If the spectral distribution of sunlight as
a function of wavelength I(λ) is known, for example in watts per square meter-nanometer, and the
energy of a photon is given by hc/λ in joules, then the photosynthetic photon flux density (PARPPFD)
in units of photons/seconds per square meter (s-m2) can be calculated as

700
1
PAR PPFD =
hc ∫ λ I ( λ ) dλ
400
(7.10)

where hc is the product of Planck’s constant and the speed of light, respectively. To make the num-
bers reasonably convenient to work with, the photon number is usually expressed in units of micro-
moles (6.022 × 1017). PAR with full Sun (e.g., broadband DNI ~ 1,000 W/m2) on a cloudless summer
day would measure around 2,000 micromoles per second square meter (µmoles/(s-m2)) near solar
noon. PAR is sometimes defined by

700

PAR =
∫ I ( λ ) dλ.
400
(7.11)

PAR, in this case, is simply the integrated energy in watts per square meter between the wave-
lengths of 400 and 700  nm. Clearly, the two equations  are different as they account for the num-
ber density of photons and power density at the collector respectively. Figure  7.7 illustrates the
different spectral weighting. For  the energy definition of PAR, the entire energy spectrum in the
wavelength interval between 400 and 700 nm is given equal weighting illustrated by the thin black

FIGURE 7.7 The weighting for PAR measured in energy units of W/m2 (thin black line) and in photons/
(sec-m2) (thick dashed line).
174 Solar and Infrared Radiation Measurements

line in the figure. For the photosynthetic photon flux density (PPFD) definition in Equation (7.10),


weighting is proportional to the wavelength. A 400 nm photon has more energy than a 700 nm photon,
yet both produce a chemical reaction in plants; therefore, an energy spectrum should be weighted by the
dashed curve in Figure 7.7 to calculate its photosynthetic effect. Federer and Tanner (1966) and Biggs
et al. (1971) began to standardize PAR measurements with some early sensor designs that attempted
to weight the sensitivity proportionally to the wavelength. For example, EKO Instruments, Kipp &
Zonen, LI-COR Biosciences, and Skye Instruments Ltd are four instrument manufacturers with sensor
designs, models ML-020P, PQS1, LI-190R and SKP 215, respectively, that approximate this weighting
function using colored glass or interference filters in combination with a silicon photodiode detector
to achieve proper wavelength sensitivity. Other considerations in purchasing a PAR sensor, besides the
agreement between the theoretical and actual photosynthetic response (dashed line in Figure 7.7), are
all-weather ruggedness, good cosine response, and calibration stability. An annual stability of better
than 3 percent and an absolute response of better that 5 percent are reasonable specifications to expect.
Good PAR sensors could be found for about US$500 in 2018.

7.4.3 uva and uvb


Ultraviolet (UV) radiation has wavelengths below 400 nm and above 10 nm. The practical lower
wavelength limit for the UV for solar radiation at the Earth’s surface is around 280 nm since ozone
in the stratosphere absorbs virtually all of the radiation below this wavelength. High spectral resolu-
tion measurements of the solar UV spectrum are best made with a spectrometer with two dispersive
elements, such as a double monochromator, to minimize stray light from the longer and stronger
wavelength portion of the solar spectrum reaching the detector, especially to achieve the best mea-
surements at the shortest UV wavelengths. Because of the high cost of a double monochromator, it
is more common to measure UV in broad wavelength bands. UVA covers the wavelengths between
315 and 400 nm and UVB covers the wavelengths between 280 and 315 nm. UVC includes wave-
lengths shorter than 280 nm but will not be discussed here because no significant amounts of UVC
reach the Earth’s surface due to atmospheric absorption and scattering.
UVB is the most common UV measurement because of its deleterious effects on humans and
materials. Overexposure to UVB is associated with sunburn, eye disease (including cataracts), and
suppression of the immune system. However, underexposure to UVB leads to lower vitamin D
production, which has deleterious effects as well. Exposure to UVB at levels before sunburn may
produce a net positive effect (Holick 2005). The World Health Organization (WHO) has, however,
taken the conservative position on its website (http://www.who.int/uv/faq/uvhealtfac/en/index.html)
that the harmful effects of UVB exposure generally outweigh the positive ones.
UVA radiation penetrates deeper into the skin layers and causes premature aging. It also pro-
duces hydroxyl and oxygen radicals that cause DNA damage. In fact, some scientists believe that the
higher incidence of malignant melanomas in sunblock users may result from sunblock ingredients
blocking UVB but not UVA (Autier et al. 1995).
UVB sensors are usually tailored to match as closely as possible the Diffey erythemal response
function (McKinlay and Diffey 1987). UVA  sensors measure the 315–400  nm passband with a
uniform response. Most modern UVB sensors use a filter to define the passband of the radiation
for isolation followed by a phosphor that is sensitive to UVB. The phosphor produces visible radia-
tion that is more easily detected by solid-state detectors. For these instruments to perform well and
with good stability at the expected low signal levels, they must be temperature stabilized. Another
important consideration is their ability to approximate a Lambertian cosine response; that is, they
should mimic a receiver with a perfect cosine angular response as described in the section on pho-
tometry. As far as absolute accuracy, UVB sensors that are better than 10 percent are likely the best
that one can expect; relative calibration stability should be better than 5 percent per year. A list of
UVB instrument manufacturers (see Appendix G) generally includes EKO Instruments, Kipp &
Zonen, Solar Light, and Yankee Environmental Systems, Inc.
Solar Spectral Measurements 175

7.5 NARROW-BAND FILTER RADIOMETRY


Narrow-band filter radiometers are used for the measurement of atmospheric aerosol and
molecular constituents, such as water vapor and ozone. Direct normal solar irradiance instruments
with spectrally selective detectors are called sun photometers or, more generally, sun radiometers
since they are used outside the photopic response band as well as within it. A sun radiometer mea-
sures in several narrow wavelength bands with responsivities defined by full widths (wavelength)
at half-maximum (response) of (typically) 5–10 nm. The instrument should track the sun using a
narrow field of view to observe the disk of the sun plus a small annulus around the sun that allows
for tracking uncertainty and atmospheric refraction (e.g., 2–2.5 degrees full view angle).

7.5.1 aeRosol oPtiCal dePtH


For a clear view of the sun with no clouds, one can use the Beer-Lambert law to calculate the col-
umn aerosol optical depth using

−τ aerosol ( λ )⋅maerosol −τ ozone ( λ )⋅mozone −τ Rayleighh ( λ )⋅mRayleigh


I (λ ) = I0 (λ ) e (7.12)

where Equation (7.12) is Equation (7.1) expanded to explicitly show the optical depth components
that are normally measured in the visible and near-IR spectrum. For sun radiometers to measure
aerosols, filter wavelengths and bandwidths are carefully chosen to avoid water vapor and molecu-
lar oxygen absorption bands. If I(λ) is measured, the m’s calculated using the Sun’s elevation, τozone
obtained from satellite or ground-based measurements, τRayleigh calculated using Equation (7.2), and
the measured atmospheric pressure, then τaerosol can be calculated if Io(λ) is known. Determining
the extraterrestrial solar irradiance, Io(λ), is paramount when it comes to performing good sun
radiometry and is the most demanding task in making valid aerosol optical depth measurements.
For example, a 1 percent uncertainty in Io(λ) leads to an approximate uncertainty of 0.01 in aerosol
optical depth.
The air masses for ozone, aerosol, and Rayleigh scattering differ because each is vertically dis-
tributed differently in the Earth’s atmosphere. However, an adequate approximation for many pur-
poses is to assume that the air masses for ozone, aerosol, and Rayleigh scattering are the same, in
which case Equation (7.12) reverts to Equation (7.1). Taking natural logarithm of Equation (7.1) gives

ln ( I ( λ ) ) = ln ( I 0 ( λ ) ) − τ ( λ ) ⋅ m (7.13)

that has the form of a linear equation y = yo + b ⋅ x. If ln(I) versus air mass m is plotted, then for
a perfectly stable atmosphere the result should be a straight line whose intercept is the ln(Io) and
whose slope is the negative of the total optical depth (an example plot appears later in the chapter
in Figure 7.9). If ln(Io) is known, then everything else in Equation (7.13) can be calculated or mea-
sured to derive total optical depth at any other time. Subtracting optical depths caused by Rayleigh
scattering and ozone absorption yields the aerosol optical depth.
Although the assumption is that the air masses are the same for the different extinction com-
ponents to simplify the explanation of the Langley plot technique, the more correct procedure for
Langley analysis is to plot the left-hand side of the following equation versus aerosol air mass,
that is,

ln ( I ( λ ) ) + τ ozone ⋅ mozone + τ Rayleigh ⋅ mRayleigh = ln ( I 0 ( λ ) ) − τ aerosol ⋅ maerosol (7.14)

Using Equation (7.13) versus Equation (7.14) for the determination of the Langley intercept results


in much less than 1 percent bias at visible and longer wavelengths, but it becomes crucial to use
176 Solar and Infrared Radiation Measurements

Equation  (7.14)  for studies in the UV because Rayleigh and ozone are generally the dominant
contributions to the total optical depth.
The  air mass relative to the zenith direction typically used in Equation  (7.13)  is for a pure
Rayleigh atmosphere. Kasten and Young (1989) developed a revised approximation for a Rayleigh
atmosphere, namely,

1
mRayleigh = (7.15)
(
cos sza + 0.50572 ⋅ ( 96.07995 − sza )
−1.6364
)
where the solar zenith angle sza on the right-hand side of the equation is in degrees. Almost always
it is the case that more than 90 percent of the atmospheric ozone resides in the stratosphere between
20 and 40 km and varies with season. The ozone layer is lower in the winter than in the summer.
The altitude also depends on the latitude with higher layers at the equator than at the poles. If one
considers the geometry of an elevated layer, ozone air mass relative to the zenith direction is some-
what smaller than the Rayleigh air mass. Ozone air mass as calculated by the Dobson spectropho-
tometer community (Komhyr and Evans 2008) is given by

R+h
mozone = (7.16)
( R + h) − ( R + r ) ⋅ sin 2 ( sza)
2 2

where R is the mean radius of the earth 6371.299 km, r is the station height above mean sea level in
kilometers, and h is the mean height of the ozone layer in kilometers. Although aerosol layers can be
elevated, especially as a result of long-range transport in the free troposphere, they typically reside
in the boundary layer near the surface. Therefore the same air mass expression for aerosol that is
used for water vapor is used here (Kasten 1966; based on Schnaidt 1938), namely,

1
maerosol = (7.17)
cos ( sza ) + 0.0548 ⋅ ( 92.65 − sza )
−1.452

Two well-known locations for obtaining the clear, stable atmospheres needed for this straightfor-
ward determination of ln(Io) are Izaña Observatory on the island of Tenerife in the Canary Islands
and at Mauna Loa Observatory on the island of Hawaii in the Hawaiian Islands. Even at these
sites, a number of Langley plots are required using only morning data since upslope winds can
bring lower tropospheric air over the site. Examining multiple Langley plots allows identifica-
tion of measurement outliers and provides a robust estimate of the uncertainty of the calibration
constant ln(Io).
Alternatively, indoor lamp calibrations, in principle, can be used to calibrate sun radiometers.
A  very careful study of lamp calibrations versus Langley calibrations from a high mountain site
(Jungfraujoch in Switzerland) by Schmid and Wehrli (1995) demonstrated that the calibrations dif-
fer from as little as 0.2 percent to as much as 4.1 percent depending on the wavelength with most
differences greater than 1.4 percent. A follow-on study (Schmid et al. 1998) suggested that typical
differences between the two calibration techniques were at best 2–4 percent and became more vari-
able than this if different published extraterrestrial spectra (Io(λ)) were used for the extraterrestrial
irradiance. Lamp calibration, however, may be a viable means for checking the stability of an instru-
ment, but Langley plots are suggested for establishing the calibration of the radiometer. For example,
if a good Langley calibration is associated with a near-concurrent lamp calibration, then future lamp
calibrations could be used to track and correct for the degradation of the sun radiometer if needed for
a site where conditions for good Langley plots occur infrequently or are impossible.
Solar Spectral Measurements 177

Michalsky et al. (2001a) and Augustine et al. (2003) describe two calibration methods that do
not require high mountains or lamps. The fundamental idea is to use all possible Langley plots to
develop a robust and continuously updated calibration of the sun radiometer. These methods have
demonstrated uncertainties around 1 percent.
AERONET and its affiliates form the largest aerosol optical depth network in the world (Holben
et al. 1998). The AERONET network uses measurements on Mauna Loa to calibrate their stan-
dards using Langley plots over several good, stable days. These standard instruments then are
transported to the Goddard Space Flight Center in Maryland, where they are used to transfer
calibrations to the network instruments via side-by-side comparisons. Comparisons between sun
radiometers calibrated using this method and the method of Michalsky et al. (2001a) found biases
in daily-averaged aerosol optical depths of less than 1 percent over a 3-year period (Michalsky
et al. 2010).

7.5.2 WateR vaPoR


Using narrow-band filter radiometers to measure water vapor is more difficult than measuring aero-
sol optical depth. Thome et al. (1992) reviewed the sun radiometric techniques, both empirical and
theoretical, that had been used in the previous 80 years for water vapor column retrievals. The tech-
nique that was used most often in the last 20 years to retrieve water vapor is based on the modified
Langley method first introduced by Reagan et al. (1987). If measurements are made in the water
vapor band, then Equation (7.1) can be modified; thus

I = I 0e −τ ⋅mTH2O (7.18)

where τ and m are the optical depths associated with Rayleigh scattering and aerosol scattering and
absorption. TH2O is the transmission in the water band due to water vapor absorption. Bruegge et al.
(1992) modeled the transmission in the water band using

b
TH2O = e ( )
− k u⋅m
(7.19)

where k and b are constants. This equation accounts for the product of water vapor air mass m and
column water u not behaving linearly in extinction. In other words, while doubling aerosol doubles
aerosol optical depth, doubling water vapor increases the water vapor optical depth by about a factor
of 2. Although Reagan et al. (1987) used the approximation that b was equal to 0.5, it should be
calculated for the specific filter used for the measurements because the optical depth is measured
over a finite filter width of, typically, 10 nm and includes water vapor and water vapor continuum
contributions. A  radiative transfer model is used to calculate the transmission over the expected
range of u ⋅ m. Each calculated transmission value t(λ) for a given water vapor path is formed by
convolving with the filter function f(λ) according to

T=
∫ t ( λ ) f ( λ ) dλ
λ
(7.20)
∫ f ( λ ) dλ
λ

Taking the natural logarithm of Equation (7.19) twice and rearranging terms gives

ln ( ln (1 / TH2O ) ) = ln( k ) + b ⋅ ln(u ⋅ m) (7.21)


178 Solar and Infrared Radiation Measurements

A plot of the left-hand side versus ln(u ⋅ m) should be linear, and a linear least-squares fit to the
calculations yields the ln(k) and b coefficients. Taking the natural logarithm of Equation (7.18) after
substituting Equation (7.19) and rearranging terms yields

(
ln( I ) + τ scat ⋅ mscat = ln ( I 0 ) − k (u ⋅ m)b ) (7.22)

If the left-hand side of Equation  (7.22) is plotted versus mb on clear days that have stable water
vapor, the intercept of the linear plot is ln(Io), which can then be used to solve for water vapor for
any subsequent clear-sky measurement. Langley plots for aerosol channels are less of a challenge
than modified Langley plots for the water channel. Even seemingly clear days generally experience
large water vapor changes; therefore, many more modified Langley plots are required to obtain even
an approximate calibration compared to the aerosol channels.
AERONET (Holben et al. 1998) implemented its version of this algorithm to produce a column
water vapor product for its sites. Their latest methodology for water vapor retrieval is described in
Smirnov et al. (2004).
Michalsky et  al. (2001b) introduced an alternative technique for retrieving water vapor that
depends on calibration with a stable spectral lamp and a trusted extraterrestrial spectral irradi-
ance near the 940 nm water vapor band and the continuum near it. Besides the uncertainty in the
extraterrestrial spectrum and the lamp calibration, the overall uncertainty of this procedure also
depends on estimating the aerosol and Rayleigh scattering optical depths and on the accuracy of
the water vapor transmission calculations, as was the case for the previously described modified
Langley approach.
If column water vapor is required, for example, for the spectral irradiance model calculations
that are described in Section 7.6.2, and a sun radiometer with a water channel is not available, there
are other possible sources of column-integrated water vapor.
The  National Weather Service launches radiosondes on balloons twice each day as part
of a worldwide network of measurements that profile temperature, humidity, pressure, and
wind speed and direction. The launches are coordinated with launch times centered on 0 and
12 GMT. The website http://www.wmo.int/pages/prog/www/ois/volume-a/vola-home.htm con-
tains a list of stations launching radiosondes throughout the world. Although relative humidity
is measured on these radiosondes, dew point temperature is the reported value. The  actual
vapor pressure e for any measurement in the profile can be calculated from the dew point tem-
perature Td using

 17.502 ⋅ Td 
e = 6.1121 ⋅ exp   (7.23)
 240.97 + Td 

The particulars of radiosondes measurements as practiced in the United States are explained in the
Federal Meteorological Handbook No. 3 (1997, https://www.ofcm.gov/publications/fmh/FMH3/00-
entire-FMH3.pdf). Column water vapor column can then be calculated using

P2
1 0.622 ⋅ e
W=
g ∫
P1
p−e
dp (7.24)

In Equation (7.24) from Fleagle and Businger (1963) W is the integrated water column, g is the accel-
eration due to gravity, and p is the pressure with the integration between pressure level P1 and P2.
Alternatively, the water vapor column can be derived from the global positioning system (GPS)
measurements (Bevis et al. 1992; Businger et al. 1996; Gutman et al. 1994, 1995; Kuo et al. 1993;
Rocken et al. 1993; Yuan et al. 1993). These papers explain how the delay time of the signal received
Solar Spectral Measurements 179

from the average of multiple GPS satellite transmissions is used to derive water vapor column.
The  website https://www.unidata.ucar.edu/data/suominet/ can be used to find up-to-date column
water vapor for many sites.

7.5.3 sun RadiometeRs


Most sun photometers (or sun radiometers in this book) are made to point at the Sun with a cir-
cular and narrow field of view, ideally just large enough to encompass the solar disk and allow
for refraction and small tracking errors. Generally, this alignment is done with a field of view
smaller than 2.5 degrees full-angle. Some sun radiometers contain a moveable wheel that positions
narrow-band (5–10 nm full width half-maximum) filters to permit sequential sampling over a fixed
detector. Some sun radiometers contain channels with individual filter–detector combinations to
ensure simultaneous measurements in all channels. Most sun radiometers must be mounted on a
separate tracker, although some come with a tracker as part of the design. Some sun radiometers are
temperature controlled, which is recommended, but others rely on temperature measurements made
independently for correction of any temperature sensitivity of the sun radiometer detector or filter.
Multi-filter rotating shadowband radiometers (MFRSRs) (Harrison et al. 1994) are different in
that they measure the GHI spectrum and then shade the receiver from the direct solar and measure
the DHI spectrum. The DNI and DHI spectral output from a MFRSR over a mostly clear day is
shown in Figure 7.8. By subtraction of DHI from GHI and division by the cosine of the solar zenith
angle, direct normal spectral irradiance is calculated; the result must be further corrected for the

Direct Normal Spectral Irradiance Diffuse Horizontal Spectral Irradiance

415 nm
1.5

500 nm
0.15

615 nm
spectral irradiance in W/(m2−nm)

spectral irradiance in W/(m2−nm)

673 nm
500 nm
1.0

415 nm
0.10

870 nm

940 nm
0.5

615 nm
0.05

673 nm

870 nm
940 nm
0.0

0.00

6 8 10 12 14 16 18 6 8 10 12 14 16 18
17 October 2012 17 October 2012

FIGURE 7.8 Clear sky plot of MFRSR DNI (left) and DHI (right) data on October 17, 2017, for Boulder,
Colorado, USA. The DNI and DHI plots are shown to illustrate the difference in spectral distribution of DNI
and DHI irradiance over the day. The GHI spectral distribution (not shown) is a combination of the DNI and
DHI spectral irradiance and can be obtained by the summation technique.
180 Solar and Infrared Radiation Measurements

2.0
1.5 Langley Plots for 17 October 2012

Vo Values

500 nm

673 nm
1.0

615 nm



●●


●●


●●

●●


●●

●●


●●

●●


●●

●●


●●

●●

●●

●●

●●

●●

●●

●●

●●

●●

●●
●●

●●●

●●

●●

●●
●●

●●
●●

●●
●●

870 nm

●●

●●
●●

●●●

ln(V)

●●●
●●

●●
●●

●●
●●
●●
●●
●●
●●
●●
●●
●●●
●●●
●●
●●
●●
●●
●●
●●
●●
●●
●●
●●
●●
●●
●●
●●
●●
●●
●●
●●●
●●●
●●●
●●

● ●
●●

●●●●
●●●●
●●●●
●●●
●●●●●
●●
●●●●●
●●●●●
●●
●●●
●●●●
●●
●●●
●●●
●●●●●
●●●●
●●
●●●

●●●
●●
●●●
●●●
●●●
●●
0.5

●●●●
●●
●●
●●●
●● ●

●●●
●●●
●●
●●
●●
●●●● ●

●●
●●
●●

●●
●●

●●●

●●
●●
●●●


●●●
●●

●●
●●

●●●
●●



940 nm
●●●

415 nm



●●
●●


●●
● ●

● ●

● ●
● ●
● ●

● ●
● ●

0 1 2 3 4 5 6
Air Mass

FIGURE 7.9 Langley plots for several wavelengths. The linear extension of the Langley plots to zero air
mass are within a few percent of the best estimates for the extraterrestrial spectral response at each wavelength
except for the 940-nm wavelength that is in the spectral region significantly affected by water vapor.

bench-measured departure of the receiver from a perfect cosine response. Since this instrument
involves subtraction of two measurements and a cosine correction, it is inherently somewhat less
accurate than a pointing sun radiometer that makes a single measurement of the DNI spectrum.
The  calibration of the MFRSR can be determined using Langley plots of the natural log of the
narrowband DNI irradiance versus air mass on multiple clear days. If the aerosol content does
not change significantly over the day, a linear plot will be obtained and the intercept at zero air mass
will be equal to the extraterrestrial spectral response* at the specific wavelength outside the atmo-
sphere (see Figure  7.9). Accuracies of an estimated extraterrestrial spectral instrument response
of 1 percent are obtainable in these narrow bands. Measurements in the 940-nm wavelength band
(a wavelength region affected by water vapor) do not produce a linear Langley plot even when the
column water vapor is stable as discussed in Section 7.5.2. Note the highly variable water vapor
channel in Figure 7.9.
Sun radiometers use interference filters, which have improved over the years, but continue to
have trouble maintaining their transmission stability over time, often due to their hygroscopic
nature. Ideally, the specifications for these instruments would indicate less than 1 percent response
change per year. The  spectral band pass profiles of these instruments are typically very stable

* The term “response” rather than irradiance in used here because sun radiometers are not often calibrated for irradiance.
In Equation (7.1) I/I0 is a ratio; therefore, any calibration factor needed to convert measured instrument voltages to irradi-
ances will cancel. Optical depths can be measured using measured instrument responses only.
Solar Spectral Measurements 181

even if the transmission changes. Temperature stability is the ideal, but it is possible, although less
desirable, to use ancillary temperature measurements to correct instruments whose temperature
response has been characterized.
Integrated, out-of-band light getting through the interference filters should be less than 0.1
percent. This detail is extremely important for good aerosol optical depth measurements. This mea-
surement can be tested using cut-on filters that pass light longward of the tail of the interference pass
band. For example, for a 500 nm filter, a Schott glass filter OG 570 should not pass any radiation if
placed over a good 500 nm filter with a 10 nm pass band. Measurements of the direct Sun on a clear
day around solar noon should be made with a completely dark cover to determine zero offset, then
with full Sun, followed by full Sun with the OG 570 filter covering the radiometer over the 500 nm
filter; two or three measurements with this sequence should be averaged. After subtracting the dark
offsets from the measurements, the ratio of the measurements with the OG 570 filter to the uncov-
ered measurement average should be less than 0.001.

7.6 SPECTROMETRY
Spectrometers separate light into wavelengths typically using a diffraction grating or a prism.
The fundamental layout of a spectrometer is illustrated in Figure 7.10. There is an entrance slit on
which the source of light is incident; a collimating lens, whose focal length is the distance from
the slit, that causes the light exiting the slit to become parallel (collimated) before it reaches the
dispersing element; either a prism or a diffraction grating, that may be rotated or held stationary;
and after the dispersing element is another lens with a focal length that allows the light to come to a
focus either on an exit slit or on a detector array. If the last element is a slit, then a detector is placed
after the slit, and the dispersing element is rotated to scan different wavelengths of radiation. Of
course, in practice there are other complications that will not be discussed here; for example, optical
aberrations and order sorting for grating instruments.
Spectrometers can be calibrated applying standard procedures and using absolute spec-
tral irradiance sources available from the national standards laboratories of many countries,
such as the National Institute of Standards and Technology (NIST) in the United States, the
Physikalisch-Technische Bundesanstalt (PTB) in Germany, and the National Physical Laboratory
(NPL) in the United Kingdom. These sources are typically 1,000 W FEL lamps operated at pre-
scribed electrical currents at a fixed distance between source and spectrometer (e.g., Walker
et al. 1987).

FIGURE 7.10 Schematic of the key components of a spectrometer.


182 Solar and Infrared Radiation Measurements

7.6.1 sPeCtRometeRs
Manufacturers of spectrometers are numerous; spectrometers that operate unattended in an outdoor
environment are few. LI-COR Biosciences offered a weatherproofed scanning spectrometer for
many years. The LI-1800 scanned solar spectral irradiances between 300 and 1,100 nm with about
6 nm spectral resolution, although higher resolution using narrower slits was possible. Production
of this instrument stopped many years ago, and support for existing instruments ceased at the end
of 2012.
A prototype Precision Solar Spectroradiometer (PSR) built by the Physikalisch-Meteorologisches
Observatorium Davos (PMOD) was tested at Payerne in 2012 (Schmutz and Ebert 2013). The PSR
measures the incoming DNI spectral composition from 320 to 1,030 nm with a resolution full width
at half maximum (FWHM) between 1.6 nm in the ultraviolet and 4.0 nm in the IR portion of the
solar spectrum. A solar spectrum is obtained by averaging over ten 500-ms scans. This averaging
takes between 15 and 20 seconds and the spectral data are stored every minute. Under clear skies,
the standard deviation between the individual scans is on the order of 0.3 percent or better over
most of the range. Of course, under changing skies conditions, the variation between each scan can
be significant. However, the averaging provides a more representative value under rapidly changing
weather conditions. Before and after each solar spectrum measurement, the dark counts measured
with the shutter closed are used to determine the offset of the instrument for no incident radiation.
The instrument has a 2 degree field of view and a temperature stabilized optical bench. The follow-
ing data treatment is applied to the solar measurements:

1. The dark counts are subtracted from the solar measurements.


2. A temperature correction based on the temperature of the detector is applied to the spectra
resulting from step 1.
3. The solar spectra are compared to a high-resolution extraterrestrial solar reference spec-
trum and the resulting spectral wavelength shifts are applied to the wavelength of the solar
spectra.
4. The sensitivity of the instrument, determined prior and after a campaign in the labora-
tory of PMOD/WRC by measuring the irradiance of a reference source (1,000 W FEL
tungsten-halogen lamp, calibrated at PTB) is applied to the measurements to convert them
from counts per second to watts per square meter-nanometer. This lamp has an uncer-
tainty that varies from 1.1 percent at 300 nm to 0.9 percent at 1,000 nm (see Chapter 13,
Calibrations).
5. As a final step, the spectra are interpolated to put them on a uniform wavelength grid of
0.5 nm in wavelength.

The resulting uncertainties of the solar spectra obtained for the PSR following the procedure out-
lined in steps 1 through 5 is shown in Figure 7.11. The calibration uncertainty (U95%) of the PSR
measurements during the campaign in 2012 was 3 percent between 300 and 390  nm, 2 percent
between 390 and 800 nm, and 4 percent above 800 nm. The transfer standard 1,000 W lamp used
for calibrations received a calibration certificate for the lamp giving a 1 percent uncertainty in this
wavelength range (U95%).
The uncertainty for the PSR was determined by combining the following uncertainty compo-
nents and assuming they are uncorrelated (root mean square sum):

1. Calibration uncertainty
2. Repeatability (standard deviation of the measurements)
3. Uncertainty in the wavelength
4. Uncertainty in the applied temperature correction
Solar Spectral Measurements 183

FIGURE 7.11 Calculated uncertainty of the PMOD prototype PSR spectroradiometer. The uncertainty is


less than 2 percent for the range between 400 and 800 nm and less than 4 percent for 800–1,000 nm. At the
extreme ends of the range, the uncertainties increase. This change in uncertainties over the 300–1,100 range
is typical of spectroradiometers using silicon photo sensors. Development and characterization of the PSR
continues. (From Raptis, P. et al., Atmos. Meas. Tech., 11, 1143–1157, 2018.)

The  spectroradiometer was mounted on an automatic tracker with a four-quadrant photoelectric


sensor to improve the tracking. The PSR spectroradiometer is waterproof and it was deployed dur-
ing all weather conditions. Accurate aiming of the PMOD spectroradiometer is important because
of its narrow field of view.
EKO Instruments Co., Ltd. offers several all-weather spectroradiometers that cover the UV to
near-IR between 300 and 2,550  nm. As an example, the specifications for an EKO MS-711 are
given in Table 7.2. The EKO spectroradiometers can be configured to measure spectral GHI or, by
replacing the dome on the spectroradiometer with a collimating tube, measure DNI as illustrated

TABLE 7.2
EKO MS-711 Specifications
Wavelength range 300–1,100 nm
Optical resolution FWHM <7 nm
Wavelength accuracy ±0.2 nm
Directional response at 1,000 Wm−² <5%
Temperature response −10 to 50°C <2%
Temp. control 25°C
Operating temperature range −10 to 50°C
Exposure time 10–5,000 ms
Dome material Quartz
Field of view (FOV) 180°
184 Solar and Infrared Radiation Measurements

in Figures 7.12 and 7.13. In Figures 7.12 and 7.13 the three spectroradiometers cover different spec-
tral ranges. The photograph was taken during the 2016 NREL Pyrheliometer Comparison (NPC)
(Reda et al. 2016). As mentioned earlier, spectroradiometers are calibrated against standard lamps
with the sensor perpendicular to the light source. For spectroradiometers used for measuring DNI,
this is a traceable method when using a standard lamp with a certified spectral irradiance calibration.
For pyranometers used for outdoor spectral measurements, a correction for the cosine response of
the instrument should be made. The cosine response calibrations should be performed on an opti-
cal bench with a stable lamp. An analysis of GUM methodology for assessing the uncertainty of
GHI spectrorradiometer calibrations has been presented by Peterson et  al. (2017). This  method-
ology demonstrates how to combine the various sources of measurement uncertainty to obtain
an overall calibration uncertainty. A plot of the deviation from a Lambertian cosine response for
an EKO spectroradiometer is shown in Figure  7.14. The  shade-unshade methodology similar to
the method used to calibrate broadband pyranometers could provide a more accurate estimate of

FIGURE 7.12 Image of three EKO spectroradiometers used for GHI spectral measurements over different
wavelength ranges.
Solar Spectral Measurements 185

FIGURE  7.13 Three EKO spectroradiometers used for DNI spectral measurements, the MS-701
(300–400 nm), MS-711 (300–1,100 nm), and the MS-712 (300–1,700 nm). Picture taken at the 2016 NPC
at NREL. The fore optics are the same for the three instruments in this photo. Other DNI fore optics are
available from EKO.

FIGURE 7.14 Cosine response of an EKO MS-711 spectroradiometer. The instrument is mounted on an opti-


cal bench and calibrated in different orientations with respect to a reference lamp. The instrument is rotated
90 degrees between each set of calibrations yielding four different curves as the angle of incidence between
the lamp and instrument is changed.
186 Solar and Infrared Radiation Measurements

200 300 400 500 600 700 800 900 1,000 1,100 1,200 1,300 1,400 1,500 1,600 1,700

FIGURE  7.15 Morning GHI spectra obtained in Golden, Colorado, USA, using the EKO MS-711 and
MS-712 spectroradiometers.

any angle of incidence bias; however, this requires a reference spectral pyrheliometer for the DNI
spectral measurements. An example of spectral GHI measurements made using the EKO MS-711
and MS-712 spectrometers is shown in Figure 7.15.
EKO now produces a rotating shadowband spectroradiometer (RSR) (RSB-01S). This is similar
to the RSR discussed in Chapter  10 except a spectroradiometer is used for the measurements
instead of a photodiode pyranometer (see Figure  7.16). Since spectroradiometer measurements
take a considerable time, ~20 seconds in some cases, the instrument has to be in near continu-
ous use to determine the DNI spectrum. The shadowband stops just before shading the sensor
for the first measurement series, shades the instrument for the next measurement series, and then
just after shading the sensor for the third series of measurements. The  first and third spectral
measurements then are used to estimate the change in diffuse irradiance due to blockage by the
band and this is added back into the diffuse measurement when the sensor is shaded by the band.
Coincident DNI spectral measurements then can be used to obtain the cosine response of the
RSB-01S.
There are many companies offering diode array and charge-coupled device (CCD) spectrometers
but not  generally for continuous outdoor use. Custom-built outdoor instruments based on these
devices have been developed, such as Herman et al. (2009) for UV, and this instrument is now a
commercial product (http://sciglob.com/products.html).

7.6.2 sPeCtRal models


Mathematical models for calculating the spectral distribution of solar radiation can be useful for
estimating responses of broadband and spectrally selective devices, for example, PV modules. These
models do very well at predicting the clear-sky spectral irradiance if the key inputs are well specified.
Solar Spectral Measurements 187

FIGURE 7.16 An image of the RSB-01S that uses an EKO spectrometer and a rotating shadowband to enable
measurements of spectral GHI and DHI and calculation of the spectral DNI. (Courtesy of EKO Trading, Ltd.,
Kowloon, Hong Kong.)

Some models work for cloudy skies; however, specifying the cloud properties is challenging.
For clear skies, the critical inputs are the aerosol optical depth for at least two wavelengths, integrated
water vapor, total column ozone, and the spectral surface reflectivity at a number of wavelengths.
For concentrating systems, the surface albedo is irrelevant because the DNI is the only irradiance
component of interest. Time, latitude, and longitude are required to calculate the solar position. Two
models in the public domain that are widely used for solar calculations are the Simple Model of the
Atmospheric Radiative Transfer of Sunshine (SMARTS; Gueymard 1995, 2001; the latest release
is SMARTS Version 2.9.7, June 2017) and the Santa Barbara DIScreet Ordinate Radiative Transfer
(SBDART; Ricchiazzi et al. 1998).
The solar energy community is the primary user of SMARTS for clear-sky calculations. It is use-
ful for calculating DNI radiation and global tilted irradiance (GTI) falling on a surface of any ori-
entation. It is useful, therefore, for calculating the radiation that would be incident on a rooftop PV
panel or a solar hot water installation, for example. It also has useful built-in functions to calculate,
for example, PAR, photometric flux, and various UV action spectra. The software and instruction
manual are available from the National Renewable Energy Laboratory’s Renewable Resource Data
Center at http://www.nrel.gov/rredc/smarts/.
SBDART (Ricchiazzi et al. 1998) was written to calculate radiative transfer in cloudy as well
as clear atmospheres. This  code is widely used in the atmospheric radiation community. It  is
accessible online and as downloadable code. An introduction and explanation was available at
188 Solar and Infrared Radiation Measurements

http:// www. ncgia.ucsb.edu/projects/metadata/standard/uses/sbdart.htm, although at this time it is


no longer supported. Both SMARTS and SBDART take a little effort to understand the input struc-
ture, but once this is learned, many useful calculations can be performed.
For calculations that focus on the UV spectrum, a very useful, well-tested, and straightforward
code is available from the National Center for Atmospheric Research (https://www2.acom.ucar.
edu/modeling/tropospheric-ultraviolet-and-visible-tuv-radiation-model). Madronich (1993) and
Madronich and Flocke (1999) developed this model for biological and atmospheric chemistry appli-
cations, respectively.
For solar spectral irradiance measurements in the UV range, a double monochromator scanning
spectroradiometer is recommended because it is still the best solution to prevent stray-light con-
taminated measurements at very short UV wavelengths. Examples of instruments that could be used
are the Brewer spectroradiometers, or Bentham spectroradiometers, but the latter are not normally
considered all-weather instruments (J. Gröbner, personal communication).

7.6.3 disCRete sPeCtRal measuRements and modeling Combined


With the six spectral channels of the MFRSR, it becomes possible to retrieve aerosol optical depth
as a function of wavelength and to retrieve integrated precipitable water vapor, the most significant
molecular absorber of shortwave radiation. With a second down-viewing MFRSR of matching filters
(referred to as an MFR due to the absence of a rotating shadowband), spectral surface reflectivity
can be determined from simultaneous measurements by the two instruments. With these quantities
it is possible to use radiative transfer codes to calculate the spectral irradiance of direct normal
and diffuse horizontal radiation. This  broader capability of the MFRSR has not  been exploited;
however, spectrally sensitive technologies like PVs are becoming increasingly important, and it
will be necessary to measure the solar resource spectrally to evaluate and optimize PV efficiencies
(Brennan et al. 2014).
Along these lines, the company Spectrafy™ (spectrafy.com) sells a sun photometer (SolarSIM-D2)
and software that produces a spectral distribution of direct normal radiation (Tatsiankou et  al.
2016). The sun photometer measures DNI at selected wavelengths to estimate the aerosol optical
depth as a function of wavelength, water vapor column, and total column ozone amounts. These
are the three main atmospheric variables that affect the amount of DNI at the surface. The other
atmospheric constituents have fixed values that have much less effect on the DNI. These are input to
the SMARTS model (Gueymard 2006) to produce a DNI spectrum at the SMARTS spectral resolu-
tion (0.5 nm between 280 and 400 nm, 1 nm between 400 and 1,700 nm, and 5 nm between 1,705
and 4,000 nm). Spectrafy™ also sells a spectral pyranometer (SolarSIM-G) that produces a spectral
distribution of GHI from irradiance measurements at nine wavelengths and proprietary algorithms
to generate solar spectra over the full 280–4,000 nm wavelength range (https://spectrafy.com/wp-
content/uploads/2017/08/7th-Int.-Spectral-and-Broadband-Intercomparison-Spectrafy.pdf). A  new
rotating shadowband version of the SolarSIM-G called the SolarSIM-E is now available.
Outdoor comparisons of measurement performance of various spectroradiometers have been
made to characterize and improve this type of purpose-built radiometer (Galleano et al. 2013, 2014,
2016; Habte et al. 2014; Martinez-Lozano et al. 2003).

QUESTIONS
1. To the nearest watt per square meter, what is the currently accepted spectrally integrated
solar irradiance at the TOA of the Earth for an average separation of the Sun and Earth
(That is, solar constant)?
2. Explain why the clear winter sky is blue. Explain why the setting sun is red.
3. What is the term that is used to quantify aerosol absorption? What does it mean?
4. What is the most significant gas absorber in the solar spectrum?
Solar Spectral Measurements 189

5. What wavelength is the easiest for the eye to see? What color would you say this is? What
term is used for the eye’s response to daylight?
6. In general, the shorter the wavelength of UV light, the more dangerous it is; why do we
not worry about UVC when we go outside?
7. Given a number of measurements of the direct solar radiation on a clear day through a nar-
rowband filter centered at 500 nm, how would the estimate of the response of that instru-
ment at TOA be done?
8. What is the ozone optical depth at 555 nm if the OMI satellite measures 308 DU of ozone?
9. Calculate the Rayleigh air mass if the Sun is at a solar zenith angle of 81°.
10. What are the two types of sun radiometers? How do they differ?
11. What are two ways to separate colors in a spectrometer?
12. What is the biggest impediment to performing a Langley calibration for a sun radiometer
with a filter in a water vapor band?
13. Name two ways to get information on the water vapor column.

REFERENCES
Augustine, J. A., C. R. Cornwall, G. B. Hodges, C. N. Long, C. I. Medina, and J. J. DeLuisi. 2003. An automated
method of MFRSR calibration for aerosol optical depth analysis with application to an Asian dust out-
break over the United States. Journal of Applied Meteorology 42:266–278.
Austin, R. n.d. The Guide to Photometry. San Diego, CA: UDT Instruments.
Autier, P., J. F. Dore, E. Schifflers, J. P. Cesarini, A. Bollaerts, K. F. Koelmel, O. Gefeller, et al. 1995. Melanoma
and use of sunscreens: An EORTC case control study in Germany, Belgium and France. International
Journal of Cancer 61:749–755.
Bevis, M., S. Businger, T. A. Herring, C. Rocken, R. A. Anthes, and R. H. Ware. 1992. GPS meteorology:
Remote sensing of atmospheric water vapor using the global positioning system. Journal of Geophysical
Research 97:15787–15801.
Biggs, W., A. R. Edison, J. D. Eastin, K. W. Brown, J. W. Maranville, and M. D. Clegg. 1971. Photosynthesis
light sensor and meter. Ecology 52:125–131.
Bodhaine, B. A., N. B. Wood, E. G. Dutton, and J. R. Slusser. 1999. On Rayleigh optical depth calculations.
Journal of Atmospheric and Oceanic Technology 16:1854–1861.
Brennan, M. P., A. L. Abramase, R. W. Andrews, and J. M. Pearce. 2014. Effects of spectral albedo on
solar photovoltaic devices. Solar Energy Materials and Solar Cells 124:111–116. doi:10.1016/j.
solmat.2014.01.046.
Bruegge, C. J., J. E. Conel, R. O. Green, J. S. Margolis, R. G. Holm, and G. Toon. 1992. Water vapor column
abundance retrievals during FIFE. Journal of Geophysical Research 97:18759–18768.
Businger, S., S. R. Chiswell, M. Bevis, J. Duan, R. A. Anthes, C. Rocken, R. H. Ware, M. Exner, T. Van Hove,
and F. S. Solheim. 1996. The  promise of GPS in atmospheric monitoring. Bulletin of the American
Meteorological Society 77:5–18.
CIE. See Commission Internationale de l’Éclairage.
Commission Internationale de l’Éclairage. 1926. Commission Internationale de l’Éclairage Proceedings,
1924. Cambridge, UK: Cambridge University Press.
Dutton, E. G., and J. R. Christy. 1992. Solar radiative forcing at selected locations and evidence for global
lower tropospheric cooling following the eruptions of El Chichon and Pinatubo. Geophysical Research
Letters 19:2313–2316.
Eberhard, W. L. 2010. Correct equations and common approximations for calculating Rayleigh scatter in pure
gases and mixtures and evaluation of differences. Applied Optics 49:1116–1130.
Ellis, H. T., and R. F. Pueschel. 1971. Solar radiation: Absence of air pollution trends at Mauna Loa. Science
172:845–846.
Federer, C. A., and C. B. Tanner. 1966. Sensors for measuring light available for photosynthesis. Ecology
47:654–657.
Fehlmann, A., G. Kopp, W. Schmutz, R. Winkler, W. Finsterle, and N. Fox. 2012. Fourth World Radiometric
Reference to SI radiometric scale comparison and implications for on-orbit measurements of the total
solar irradiance. Metrologia 49:S34–S38.
Fleagle, R. G., and J. A. Businger. 1963. An Introduction to Atmospheric Physics. New York: Academic
Press.
190 Solar and Infrared Radiation Measurements

Fröhlich, C. 2006. Solar irradiance variability since 1978: Revision of the PMOD composite during solar cycle
21. Space Science Reviews 125:53–65.
Galleano, R., W. Zaaiman, D. Alonso-Alvarez, A. Minuto, N. Ferretti, R. Fucci, M. Pravettoni, et al. 2016.
Results of the Fifth International Spectroradiometer Comparison for improved solar spectral irradiance
measurements and related impact on reference solar cell calibration. IEEE Journal of Photovoltaics
6(6): 1587–1597.
Galleano, R., W. Zaaiman, C. Strati, S. Bartocci, M. Pravettoni, M. Marzoli, R. Fucci, et al. 2014. Second
international spectroradiometer intercomparison: Results and impact on PV device calibration. Progress
in Photovoltaics 23:929–938.
Galleano, R., W. Zaaiman, A. Virtuani, D. Pavanello, P. Morabito, A. Minuto, A. Spena, et al. 2013. Intercomparison
campaign of spectroradiometers of a correct estimation of solar spectral irradiance: results and potential
impact on photovoltaic devices calibration. Progress in Photovoltaics 22(11): 1128–1137.
Gueymard, C. 1995. SMARTS2, Simple model of the atmospheric radiative transfer of sunshine:
Algorithms and performance assessment. Rep. FSEC-PF-270-95. Cocoa Beach, FL: Florida Solar
Energy Center.
Gueymard, C. 2001. Parameterized transmittance model for direct beam and circumsolar spectral irradiance.
Solar Energy 71:325–346.
Gueymard, C. A. 2006. SMARTS2 code, version 2.9.5, user’s manual. Tech. Rep. Solar Consulting Services.
Gueymard, C. A. 2018. Revised composite extraterrestrial spectrum based on recent solar irradiance observa-
tions. Solar Energy 169:434–440.
Gutman, S. I., R. B. Chadwick, D. E. Wolfe, A. M. Simon, T. Van Hove, and C. Rocken. 1994. Toward an
operational water vapor remote sensing system using GPS. In  Forecast Systems Laboratory Forum,
13–19. Boulder, CO: Forecast Systems Laboratory.
Gutman, S. I., D. E. Wolfe, and A. M. Simon. 1995. Development of an operational water vapor remote sens-
ing system using GPS: A progress report. In Forecast Systems Laboratory Forum, 21–32. Boulder, CO:
Forecast Systems Laboratory.
Habte, A., A. Andreas, L. Ottoson, C. Gueymard, G. Fedor, S. Fowler, J. Peterson, et al. 2014. Indoor and out-
door spectroradiometer intercomparison for spectral irradiance measurement. NREL/TP-5D00-61476.
https://www.nrel.gov/docs/fy14osti/61476.pdf.
Hansen, J. E., and L. D. Travis. 1974. Light scattering in planetary atmospheres. Space Science Reviews
16:527–610.
Harder, J. W., J. M. Fontenla, P. Pilewskie, E. C. Richard, and T. N. Woods. 2009. Trends in solar spec-
tral irradiance variability in the visible and infrared. Geophysical Research Letter 36:L07801.
doi:10.1029/2008GL036797.
Harrison, L., J. Michalsky, and J. Berndt. 1994. Automated multifilter rotating shadow-band radiometer: An
instrument for optical depth and radiation measurements. Applied Optics 33:5118–5125.
Henyey, L., and J. Greenstein. 1941. Diffuse radiation in the galaxy. The Astrophysical Journal 93:70–83.
Herman, J., A. Cede, E. Spinei, G. Mount, M. Tzortziou, and N. Abuhassan. 2009. NO2 column amounts
from ground-based Pandora and MFDOAS spectrometers using the direct-sun DOAS technique:
Intercomparisons and application to OMI validation. Journal of Geophysical Research 114:D13307.
doi:10.1029/2009JD011848.
Holben, B. N., T. F. Eck, I. Slutsker, D. Tanre, J. P. Bluis, A. Setzer, E. Vermote, et al. 1998. AERONET—A fed-
erated instrument network and data archive for aerosol characterization. Remote Sensing of Environment
66:1–16. doi:10.1016/S0034-4257(98)00031-5.
Holick, M. F. 2005. The  vitamin D epidemic and its health consequences. Journal of Nutrition
135:2739S–2748S.
Kasten, F. 1966. A new table and approximate formula for relative optical air mass. Archives for Meteorology,
Geophysics, and Bioclimatology–Series B: Climatology, Environmental Meteorology, Radiation
Research 14:206–223.
Kasten, F., and A. T. Young. 1989. Revised optical air mass tables and approximation formula. Applied Optics
28:4735–4738.
Komhyr, W. D. 2008. Operations Handbook—Ozone Observations with a Dobson Spectrophotometer, rev.
R.D. Evans, AW No. 183, WMO/TD-No. 1469.
Kuo, Y., Y. Guo, and E. R. Westwater. 1993. Assimilation of precipitable water vapor measurements into a
mesoscale numerical model. Monthly Weather Review 121:1215–1238.
Solar Spectral Measurements 191

Liou, K. N. 1992. Radiation and Cloud Processes in the Atmosphere. New York: Oxford University Press.
Madronich, S. 1993. UV radiation in the natural and perturbed atmosphere. In Environmental Effects of UV
(Ultraviolet) Radiation, ed. M. Tevini, 17–69. Boca Raton, FL: Lewis Publishers.
Madronich, S., and S. Flocke. 1999. The role of solar radiation in atmospheric chemistry. In Handbook of
Environmental Chemistry, ed. P. Boule, 1–26. Heidelberg, Germany: Springer-Verlag.
Martinez-Lozano, J. A., M. P. Utrillas, R. Pedros, and F. Tena. 2003. Intercomparison of spectroradiom-
eters for global and direct solar irradiance in the visible range. Journal of Atmospheric and Oceanic
Technology 20:997–1010.
McKinlay, A. F., and B. L. Diffey. 1987. A  reference action spectrum for ultraviolet induced erythema in
human skin. In Human Exposure to Ultraviolet Radiation: Risks and Regulations, ed. W. F. Passchier
and B. F. M. Bosnajakovic, 83–87. Amsterdam, the Netherlands: Elsevier.
Michalsky, J., F. Denn, C. Flynn, G. B. Hodges, P. Kiedron, A. Kountz, J. Schlemmer, and S. E. Schwartz. 2010.
Climatology of aerosol optical depth in north central Oklahoma: 1992–2008. Journal of Geophysical
Research 115:1–16. doi:10.1029/2009JD012197.
Michalsky, J., J. Schlemmer, W. Berkheiser III, J. L. Berndt, and L. C. Harrison. 2001a. Multiyear mea-
surements of aerosol optical depth in the atmospheric radiation measurement and quantitative links
programs. Journal of Geophysical Research 106:12099–12107.
Michalsky, J. J., and E. W. Kleckner. 1984. Estimation of continuous solar spectral distributions from discrete
filter measurements. Solar Energy 33:57–64.
Michalsky, J. J., Q. Min, P. W. Kiedron, D. W. Slater, and J. C. Barnard. 2001b. A differential technique to
retrieve column water vapor using sun radiometry. Journal of Geophysical Research 106:17433–17442.
NASA. See National Aeronautics and Space Administration.
National Aeronautics and Space Administration. 2018. NASA Sun Fact Sheet (updated February 23, 2018).
https://nssdc.gsfc.nasa.gov/planetary/factsheet/sunfact.html (Accessed May 18, 2018).
Peterson, J., F. Vignola, A. Habte, and M. Sengupta. 2017. Developing a spectroradiometer data uncertainty
methodology. Solar Energy 149:60–76.
Raptis, P., S. Kazadzis, J. Gröbner, N. Kouremeti, L. Doppler, R. Becker, and C. Helmis. 2018. Water
vapour retrieval using the precision solar spectroradiometer. Atmospheric Measurement Techniques
11:1143–1157. doi:10.5194/amt-11-1143-2018.
Rayleigh, L. 1871. On the light from the sky, its polarization and color. Philosophical Magazine 41:107–120,
274–279.
Reagan, J. A., K. Thome, B. Herman, and R. Gall. 1987. Water vapor measurement in the 0.94  micron
absorption band: Calibration, measurement, and data applications. Proceedings of the International
Geosciences and Remote Sensing Symposium IEEE 87CH2434-9: 63–67.
Reda, I., M. Dooraghi, A. Andreas, and A. Habte. 2016. NREL pyrheliometer comparisons: September
26–October  7, 2016 (NPC-2016). NREL/TP-3B10-67311. https://www.nrel.gov/docs/fy17osti/67311.
pdf.
Ricchiazzi, P., S. Yang, C. Gautier, and D. Sowle. 1998. SBDART: A research and teaching software tool for
plane-parallel radiative transfer in the earth’s atmosphere. Bulletin of the American Meteorological
Society 79:2101–2114.
Rocken, C., R. H. Ware, T. Van Hove, F. Solheim, C. Alber, J. Johnson, and M. G. Bevis. 1993. Sensing atmo-
spheric water vapor with the global positioning system. Geophysical Research Letters 20:2631–2634.
Rottman, G. 2005. The SORCE mission. Solar Physics 203:7–25.
Schmid, B., and C. Wehrli. 1995. Comparison of sun photometer calibration by use of the Langley technique
and the standard lamp. Applied Optics 34:4500–4512.
Schmid, B., P. R. Spyak, S. F. Biggar, C. Wehrli, J. Sekler, T. Ingold, C. Maetzler, and N. Kaempfer. 1998.
Evaluation of the applicability of solar and lamp radiometric calibrations of a precision sun photometer
operating between 300 and 1025 nm. Applied Optics 37:3923–3941.
Schmutz, W., and S. Ebert, eds. 2013. Annual Report 2012 PMOD/WRC. Davos, Switzerland: PMOD/WRC.
Schnaidt, F. 1938. Berechnung der relativen Schnichtdicken des Wasserdampfes in der Atmosphäre.
Meteorologische Zeitschrift 55:296–299.
Shettle, E. P., and S. Anderson. 1995. New visible and near IR ozone absorption cross-sections for MODTRAN.
In Proceedings of the 17th Annual Review Conference on Atmospheric Transmission Models, PL-TR-
95-2060, ed. G. P. Anderson, R. H. Picard, and J. H. Chetwynd, 335–345. Hanscom AFB, MA: Phillips
Laboratory.
192 Solar and Infrared Radiation Measurements

Smirnov, A., B. N. Holben, A. Lyapustin, I. Slutsker, and T. F. Eck. 2004. AERONET processing algorithms
refinement. AERONET Workshop, El Arenosillo, Spain.
Tatsiankou, V., K. Hinzer, J. Haysom, H. Schriemer, K. Emery, and R. Beal. 2016. Design principles and field
performance of a solar spectral irradiance meter. Solar Energy 133:94–102.
Thome, K. J., B. M. Herman, and J. A. Reagan. 1992. Determination of precipitable water from solar transmis-
sion. Journal of Applied Meteorology 31:157–165.
Walker, J. H., R. D. Saunders, J. K. Jackson, and D. A. McSparron. 1987. NBS Measurement Services: Spectral
Irradiance Calibrations. National Bureau of Standards Special Publication 250-20.
Yuan, L. L., R. A. Anthes, R. H. Ware, C. Rocken, W. D. Bonner, M. G. Bevis, and S. Bussinger. 1993. Sensing
climate change using the global positioning system. Journal of Geophysical Research 98:14925–14937.
8 Albedo

When you’re thinking of putting on a new roof, make it white.


Steven Chu
Secretary of the U.S. Department of Energy, 2009

8.1 INTRODUCTION
A surface reflects and scatters a fraction of the incident solar irradiance that falls on it, including
the direct normal irradiance (DNI) and the scattered sunlight that arrives from all parts of the sky,
that is, the diffuse horizontal irradiance (DHI). However, a surface reflects the DNI and DHI com-
ponents very differently, as will be discussed in this chapter on albedo. In general, albedo is defined
as the total reflected irradiance divided by the total incident irradiance. Snowfields and deep oceans
represent extremes in albedo. For  newly fallen snow, 90 percent or more of the incident radia-
tion may be reflected at visible wavelengths, but as snow ages, melts, and refreezes the reflectivity
decreases. For the deep ocean the albedo is around 4 percent for calm seas at all solar wavelengths.
However, if the seas have breaking waves, the albedo increases. If one looks toward the Sun at the
angle of incidence, a glint is detectable from the ocean surface. This glint obeys Fresnel’s equations.
Albedos of surfaces can be very different spectrally with some surfaces showing higher reflectivity
in the visible than in the near-infrared (IR) while other surfaces can show the opposite behavior.
The albedos of vegetated areas have considerable spectral structure and change appreciably as the
plants go through stages of growth, flowering, and senescence. Measurements by Bowker et  al.
(1985) resulted in a reflectance catalogue for many common surface types, and this remains a pri-
mary reference for surface albedo for real surfaces.
The  following section discusses the most frequently made albedo measurements, including
broadband solar and photosynthetically active radiation (PAR), followed by a section that explains
the usefulness of important but infrequently made spectral albedo measurements. The fourth section
of this chapter is devoted to explaining the bidirectional reflectance function and why it is important
to understanding the nature of solar radiation. The chapter concludes with a section on how these
measurements should be made, including calibration tips.

8.2 BROADBAND ALBEDO


The most frequent measurement of albedo uses broadband solar radiation detectors, that is, a pair
of horizontally oriented pyranometers: one mounted facing the ground for measuring the reflected
(upwelling) irradiance and another mounted facing skyward to measure the downwelling irradi-
ance, as shown in Figure 8.1. PAR is probably the second most frequent measurement of albedo.
Broadband albedos are ever changing. They depend on whether the clouds hide the Sun, and if
they do not, then it depends on the angle of the Sun with respect to the zenith; early morning and
late afternoon direct solar radiation produces a higher albedo than solar noon direct radiation. If the
day is overcast, the albedo is nearly constant. If surfaces are wet, they appear darker than when they
are dry; this is very obvious for PAR, which covers the wavelengths 400–700 nm, where the eye can
clearly detect the effect.
Understanding surface albedo characteristics is important in renewable energy applications,
atmospheric physics, agricultural research, oceanography, and climate studies. Yang et al. (2008)
found that snow-free surface albedos used in the National Centers for Environmental Prediction

193
194 Solar and Infrared Radiation Measurements

FIGURE 8.1 Two examples of commercial albedometers for measuring broadband albedos: (left) Kipp &
Zonen’s CMA albedometer; (right) the Hukseflux SRA01.

(NCEP) Global Forecast Systems (GFS) model had too little dependence on solar zenith angle
and were too low at all solar zenith angles. Further, they found that the National Aeronautics and
Space Administration (NASA) satellite, MODIS, parameterizations of albedo gave overestimates
for high Sun angles and underestimates for low Sun angles. These comparisons were based on
long-term surface-based observations of broadband albedos at three U.S. Department of Energy
Atmospheric Radiation Measurement (ARM) Program sites (Ackerman and Stokes 2003; Stokes
and Schwartz 1994) and at six of the National Oceanic and Atmospheric Administration (NOAA)
Surface Radiation (SURFRAD) network sites (Augustine et al. 2000).
PAR albedos have been found to be robust predictors of relative water content in maize leaves
(Zygielbaum et al. 2009). It is reasonable to assume that this would apply to other plants and serve
as an early predictor of plant stress. As the plant becomes water stressed, the PAR albedo increases
in a very predictable way. It may be that this effect of increasing vegetation albedos has not been
incorporated in climate modeling. Although broadband albedos are useful, and the measurements
required are relatively straightforward, much more can be learned from albedo measurements with
even moderate spectral resolution.

8.3 SPECTRAL ALBEDO


Surface radiation is not reflected equally across the solar spectrum. For many surfaces, the wave-
length dependence is dramatic. Figure 8.2 is an example of measured albedos (ratio of reflected
radiation to incident radiation) as a function of local time in fractional day at four visible and two
near-IR wavelengths. The clear-day albedos are for June 8, 2006, at the ARM Southern Great Plains
(SGP) site between Lamont and Billings, Oklahoma (36.604°N, 97.485°W). The  albedo, in this
case, was calculated as an equally weighted average of two measurements over a pasture and over
a fallowed field; the two measurements of albedo were used because these were the two dominant
surface types for the area that represents the central ARM SGP site. Figure 8.2 illustrates at least
three properties of spectral albedo: the wavelength dependence, the solar zenith angle dependence,
and a slight asymmetry in solar zenith angle dependence.
The reflectance is lowest at the four shortest wavelengths (those in the PAR region of the spec-
trum). Reflectivity in the visible window of the solar spectrum (400–700 nm) typically peaks around
550 nm and decreases on either side of this peak. The reflectivity at 615 nm is slightly higher than at
673 nm. The reflectivity at 500 and 415 nm are very low. Relative to the visible, there is a threefold
increase in reflectivity at the 870 and 940 nm infrared wavelengths that is seen in all green vegetation.
Albedo 195

FIGURE 8.2 Illustration of how the albedo changes on a clear day over a vegetated surface as a function of
time and as a function of wavelength. The wavelengths in the legend are in nanometers. Since the data are for
June 8 in the Northern Hemisphere, the Sun is high in the sky at solar noon where the albedo is the lowest.
Since it is a clear day, the solar zenith angle dependence is associated with the direct Sun’s incidence angle
on the surface.

Vegetation has evolved to reflect this infrared energy starting at around 700 nm that cannot be used
for photosynthesis and would heat the plant causing deleterious effects if this longwave energy were
absorbed in any significant amount.
Another property of albedo that is immediately obvious in Figure 8.2 is the large and smooth
dependence on the solar zenith angle. For the SGP station on June 8, the solar zenith angle at solar
noon is 14°, and the end points in this plot are at a solar zenith angle of 70° observed in the morn-
ing and afternoon. Although it is not apparent from the plots, the dependence is about the same at
each wavelength; that is, the ratio of the highest to lowest albedo at each wavelength is about 1.5.
Moreover, if plotting continued to larger solar zenith angles, this ratio would climb even higher.
These data are not shown as the noise in the measurements gets significantly larger as the reflected
and downwelling irradiances decrease in magnitude near sunrise and sunset.
The final property of albedo discussed is the subtlest of the three effects. Solar zenith angles
at the end points in Figure 8.2 in the morning and evening are the same, but the reflectances are
slightly different. This difference is best detected at the longest wavelengths but is present at all
wavelengths. A  dashed vertical line denotes solar noon. It  is clear, at the longer wavelengths at
least, that the lowest albedo is displaced to a time after solar noon. The reason for this asymmetry is
unclear: It may be morning dew on vegetation that reflects more (Minnis et al. 1997), it may be the
orientation of the vegetation, or it could be a changing partitioning of the diffuse and direct sunlight
throughout the day, which will be discussed later.
Although useful spectral information is gleaned from Figure 8.2, more spectral coverage with
higher spectral resolution allows an examination of the adequacy of discrete spectral sampling of sur-
face albedo. The data to calculate albedo in Figure 8.3 were taken with a portable spectrometer, the
LI-COR LI-1800. This instrument, whose manufacture was discontinued in 2003, takes 30 seconds
to acquire continuous spectral data with 6 nanometers (nm) resolution between 330 and 1,100 nm.
The data in Figure 8.3 were acquired in three locations from a height of about 1.5 meters (m) above
ground level under a 10 m albedo tower on Table Mountain, 12 kilometers (km) north of Boulder,
Colorado, which is used for continuous measurement of spectral albedo at the same wavelengths as
in Figure 8.2. The three sample areas were chosen to represent the extremes in visual diversity of
vegetation under the tower. The measurement sequence was to move quickly to acquire downwelling
196 Solar and Infrared Radiation Measurements

FIGURE 8.3 Three measurements of spectral albedo made using a spectrometer under a 10 m albedo tower.
The measurements were chosen to examine the visual extremes under the tower. Clearly, there are only subtle
differences in spectral albedo between the different vegetation. The notable features are the peak at 550 nm,
the sharp rise near 700 nm, and the liquid absorption in plants near 975 nm.

measurements before and after an upwelling measurement. The  two downwelling measurements
were averaged and divided into the upwelling measurement to calculate albedo. Irradiance condi-
tions were stable with the Sun high in the sky near solar noon and only a few clouds on the western
horizon. The 10 m tower site was surrounded by some grasses that were green, although past their
peak, cheat grass (Bromus tectorum) that was well on its way to senescence, some taller succulent
bushes, and a few exposed rocks, but no bare soil. From Figure 8.3, it is clear that albedo is a smooth
function over most of these wavelengths. The noisy albedo values in the ultraviolet (UV) are caused
by very low upwelling irradiance and losses in the fiber optic cable that is used to transfer sunlight
to the LI-1800 spectrometer from the receiver. Some notable features are the peak near 550 nm, the
sharp rise beyond 700 nm, and the absorption band near 1,000 nm. The peak spectral albedo over the
PAR region occurs because the predominant absorption features are in the blue and red regions of the
spectrum, giving a green reflective peak as a result. The rise in spectral albedo beyond 700 nm was
discussed earlier, but the sharpness of the increase is clearly illustrated in Figure 8.3. Liquid water in
the plants causes the absorption band around 975 nm (Pu et al. 2003). The gray arrows indicate the
central wavelengths of the spectral measurements made in Figure 8.2.
Although Figures  8.2 and 8.3 represent albedos from different locations on different days,
they are typical of growing vegetation. Their spectral dependencies, in fact, closely match if
we consider the high Sun (low solar zenith angle) albedos in Figure 8.2. The only exception is
that the 940  nm albedo in Figure  8.3 slightly exceeds the 870  nm albedo, while in Figure  8.2
the opposite is true. This difference suggests that even with sparse spectral sampling, using the
six wavelengths of Figure  8.2, and knowledge that the surface was vegetative, the continuous
spectral dependence over the 330 to 1,100 nm region in Figure 8.3 could be approximated with
some confidence that the general spectral behavior was captured. Obviously, the details would
Albedo 197

FIGURE 8.4 How the spectral albedo of vegetation changes from the vigorous growth stage through senes-
cence. The peak at 550 nm and the sharp rise near 700 nm disappear.

not be uncovered without the higher spectral resolution data. McFarlane et al. (2011) developed
a technique for approximating the continuous albedo using discrete spectral measurements and
information on the surface type and state.
Figure 8.4 illustrates the change in albedo as vegetation transforms from vigorous growth to
complete senescence. The discrete measurements were taken at the SGP site in Oklahoma that was
described in the beginning of this section, and continuous albedo spectra were estimated using
the method by McFarlane et al. (2011) for four different times of the year as listed in the legend
of Figure 8.4. The first measurements on May 14 were over green pasture and a green wheat field.
The next measurements on June 8 showed the grasses and wheat still green, but the grasses were
beginning to dry and the wheat was ripening, respectively. By July 14 the wheat was harvested,
and the grasses showed almost no hint of green. Also, a portion of the surface measured under one
of the two towers had been tilled resulting in a surface of exposed soil. On October 11 the surfaces
were dead grasses and weeds. Figure 8.4 illustrates the changes in albedo with the clearest green
peak and highest near-IR reflectance on May 14. The green peak is less distinct and the reflectivity
in the near-IR lower by June 8. After harvest on July 14, there is only a slight red gradient after
700 nm. By October 11 the only spectral signature is the monotonic increase with wavelength.
In Figure 8.5 the albedos of four distinct soil samples are examined. Dirt is not dirt. Soil does
not look the same to our eye, and, indeed, it does not look the same spectrally. Bowker et al. (1985)
pointed out two of the spectral features of soils in the 300–1,100 nm spectral region that depend
on their organic content and on their iron content. Even when organic and iron content is low, there
are some general differences that depend on water content and the specific mixture of sand, clay,
and silt. In Figure 8.5 the line labeled “dry soil” has both the organic feature that shows higher
reflectivity starting at about 0.55 micrometers (µm). Iron absorption is the broad feature in the 0.9 to
1.0 µm area. The other soil albedos are monotonically increasing with wavelength with no distinct
features. The wet sample has lower overall reflectivity but is higher than black loam until around
0.7 µm. The samples labeled “soil” and “black loam” are similar, but they show different slopes with
wavelength. Note that these three soil samples resemble the wavelength dependence of the dead
vegetation albedo on October 11, 2006, in Figure 8.4.
198 Solar and Infrared Radiation Measurements

FIGURE 8.5 Albedos for four different soils from Bowker et al. (1985). Note the sharp rise near 500 nm
due to organics and the depression center near 950 nm due to iron in the dry soil albedo spectrum. The other
samples are monotonically increasing at different rates. (From Bowker, D.E. et al., Spectral reflectances of
natural targets for use in remote sensing studies, NASA Ref. Publ. 1139, 1985.)

The final class of examined surfaces is the snow-covered surface. Determining the reflectivity of
snow is challenging because albedo changes with the condition of the snow. Newly fallen fine snow
has an extremely high albedo as shown in Figure 8.6. Granular or corn snow gives the next highest
reflectivity, and melting snow shows high reflectivity but less than the other two conditions. Clearly,
snow has little spectral dependence in the UV and visible but begins to indicate absorption in the
near-IR wavelengths.

FIGURE 8.6 The albedo of snow depends on how recently it was deposited on the surface. The older the
snow, the less reflective it is. As it starts to melt, there is a noticeable drop in reflectivity. The visible spectral
dependence is almost uniform, but the solid ice absorption starts to be a factor beyond 800 nm.
Albedo 199

8.4 BIDIRECTIONAL REFLECTANCE DISTRIBUTION FUNCTION


Nicodemus (1965) defined the term bidirectional reflectance distribution function (BRDF), which
describes how radiation that is incident on a surface will scatter from that surface as a function of the
angle of incidence and as a function of the angle to which it is reflected; these angles are defined with
respect to the normal to the surface. In Figure 8.7 the incident solar radiation direction is defined by
two angles: (1) the angular distance from the zenith and (2) an azimuth angle referenced to a fixed
direction. The reflected radiation is defined by zenith and azimuth angles that are, in general, dif-
ferent from the incident angles. As the solar zenith angle increases from the normal, the reflection
increases giving a higher albedo for the direct sunlight that is reflected; the BRDF integrated over all
angles is the albedo for the direct solar radiation. This albedo is also known as the black-sky albedo.
This increase with increasing zenith angle is primarily responsible for the increases in morning and
evening albedos in Figure  8.2. The  other contribution to the albedo is from the skylight (diffuse
irradiance) that is reflected from the surface; this albedo does not change appreciably throughout the
day. The scattered sunlight albedo is called the white-sky albedo. The total albedo changes with the
solar zenith angle and, therefore, depends on the partitioning of the DNI and DHI.
Why does the white-sky albedo remain approximately constant throughout the day? This condi-
tion was discussed in Chapter 6 on diffuse measurements and is repeated verbatim here. If skylight
radiances were isotropic, that is, uniform in all directions, then the diffuse irradiance falling on a
horizontal surface could be calculated as

π
2 2π

DHI = R
∫ ∫ cos (θ ) sin (θ ) dφ dθ
0 0
(8.1)

where R is the constant radiance of the isotropic sky. The term sin(θ) dθdφ comes from spherical
geometry that defines the radiance from every direction over the sky dome, and the cos(θ) term

FIGURE  8.7 Bidirectional reflectance from a surface depends on the incident angle and azimuth of the
incoming light and on the reflectance angle and azimuth from which this reflected light is viewed.
200 Solar and Infrared Radiation Measurements

comes from the assumption that the surface receives radiation with an exact cosine response.
Since 2sin(θ)cos(θ)  =  sin(2θ), it is clear that the largest contribution to the diffuse irradiance
is from the annulus near 45° and falls monotonically and symmetrically to zero at 0° and 90°.
This means that the effective angle of incidence for diffuse radiation (DHI) is 45°, which does
not change throughout the day for isotropic skies. However, skylight is not strictly isotropic but
has contributions from all parts of the sky, especially horizon-brightening and circumsolar radia-
tion; however, the effective incident angle for diffuse is never far from the nominal 45° and
changes little throughout the day.
Measuring the white-sky and black-sky albedos separately under clear-sky conditions is difficult.
However, a reasonable estimate of the black-sky albedo can be obtained from the measurement of
albedo and irradiance on cloudy-sky and clear-sky days that occur within a few days of one another.
The measurements need to be separated by no more than a few days so that the surface and solar
geometry do not  change appreciably between the measurements. The  estimate of the black-sky
albedo is based on the reasonable assumption that the cloudy-sky albedo is about the same as the
clear-sky albedo. With measurements of direct sunlight (DNI) and diffuse skylight (DHI) on the
clear day and diffuse skylight (DHI) on the cloudy day, the white-sky albedo on the cloudy day can
be calculated and can be used to subtract the diffuse skylight albedo on the clear day to derive the
black-sky albedo on the clear day. In equation form

DHI ↑
albedo white_sky = ( on_cloudy_day ) (8.2)
DHI ↓

GHI ↑ −albedo white_sky ⋅ DHI ↓


albedo black_sky = ( on_clear_day ) (8.3)
DNI ↓ ⋅ cos ( sza )

where the up arrow indicates reflected irradiance and the down arrow indicates downward
irradiance.

8.5 ALBEDO MEASUREMENTS


8.5.1 bRoadband albedo
Broadband albedo measurements are best made with two broadband pyranometers of the same
make and model mounted with one looking upward and one looking downward (see Figure 8.1).
It  is important to use a pyranometer with a good cosine response for the downwelling measure-
ment. It is less critical for the upwelling (reflected) measurement. For both instruments the hori-
zontal alignment is critical, especially for the low Sun angle albedos where direct beam radiation
could be detected in the upwelling measurement if alignment is not strictly horizontal. A shade ring
around the upwelling pyranometer can be used to ensure that direct Sun is never able to irradiate
this pyranometer. Figure 8.8 illustrates the solar shading used in SURFRAD for all of their instal-
lations; this is at the Sioux Falls, South Dakota, site with the solar measurement on the right, which
is the south side of the tower. The instrument on the left, which is also shielded, is a pyrgeometer
for measuring the upwelling infrared irradiance. For the upwelling albedo measurement, the typical
sun shield used for downwelling measurements is not used, and the inner shield is painted with flat
black paint to reduce scattered solar light from the inside of the shield. The height chosen for the
albedo measurements for the SURFRAD sites is 9 m. The height of the albedo measurement should
be chosen so that the surface area seen at 45 degrees to the nadir direction represents the surrounding
surface area. This selection is important since the highest contribution to the upwelling radiation is
from this direction. However, that said, it is best if the area under the measurement is similar out to
about four times the measurement height.
Albedo 201

FIGURE 8.8 Photograph of the SURFRAD albedo tower near Sioux Falls, South Dakota. The shields ensure
direct sunlight will not reach the upwelling detector even for minor horizontal leveling errors.

An ideal albedo measurement would be made using one instrument rotated to periodically face
upward and downward to guarantee the same broadband spectral response. Of course, the down-
welling solar radiation would need to be stable during this measurement sequence and the time-
response of the pyranometer would need to be consistent with the data acquisition intervals.
If these instruments were to be used only for albedo measurements, the only calibration needed
would be to force their agreement when both are pointed upward since albedo is a relative mea-
surement. This ideal measurement ignores one important point: the thermal offset issue. For the
pyranometer used to measure upwelling irradiance, the thermal offset is almost absent due to the
relatively low net IR irradiance (i.e., the ground temperature is relatively close to the pyranometer
case temperature). When it looks at the clear sky, the pyranometer thermal offset is negative (i.e., the
sky temperature is generally colder than the pyranometer case temperature) and the measurement
must be corrected as discussed in Chapters 4 and 5.

8.5.2 sPeCtRal albedo


Spectral albedo is becoming more important for climate science and satellite remote-sensing research;
however, there are few options for all-weather measurements. As seen in Figures 8.3 through 8.6, the
spectral behavior of albedo varies greatly, but high spectral resolution is not necessary to capture
the key features of the wavelength dependence. Because the wavelength dependence is not highly
structured like the solar spectrum, it is possible to judiciously sample the albedo spectrum with dis-
crete filters and interpolate, for example, using the technique described in McFarlane et al. (2011).
202 Solar and Infrared Radiation Measurements

The  Yankee Environmental Systems, Inc. seven-channel rotating shadowband radiometer (RSR)
(MFR-7) is an all-weather instrument that can be used for the discrete filter measurement in the
visible and near-IR. For UV filter measurements, there are several choices including the NILU-UV
(NILU Products AS), the GUV-511 and GUV-541 (Biospherical Instruments, Inc.), and the UVMFR-7
(Yankee Environmental Systems, Inc.). The EKO Instruments Co., Inc. MS-700 is an all-weather
grating spectrometer that measures between 350 and 1,050 nm with continuous resolution of 3.3 nm.
EKO now  makes pyranometer spectrometers that can cover the spectral range between 300 and
2,550 nm, not with a single instrument, but with three instruments operated in tandem. In purchas-
ing these instruments, response stability (2% per year), temperature control (±2°C) or correction, and
cosine response (better than 10% out to incidence angles of 80°) are the most important qualities to
consider.
A very different and useful approach to estimating the spectral albedo was developed by Kassianov
et al. (2014). This method only uses spectral downwelling measurements to derive white-sky albedos.
The authors derived the cloud transmission in five spectral channels of the MFRSR (Chapter 10) at
415, 500, 615, 673, and 870 nm. From Langley plots, which are discussed in detail in the spectral chap-
ter of this book (Chapter 7), the extraterrestrial response of the MFRSR in each of these channels are
obtained. After the extraterrestrial response is established, transmission is calculated by dividing the
measured MFRSR GHI under overcast conditions on a particular day by the extraterrestrial response,
adjusted for solar distance on that day, multiplied by the cosine of the solar zenith angle, that is,

Tλ = Vλ / (Vλ 0 ⋅ cos( sza) ) (8.4)

where Tλ is the transmission at wavelength λ, Vλ and Vλ0 are the measured signal under the overcast
sky and the extraterrestrial response adjusted for solar distance. Using an assumed albedo at 415 nm
of 0.03, which is a good estimate for vegetated surfaces, and an assumed asymmetry parameter of 0.87
for 415 nm, the cloud optical depth can be estimated at 415 nm and then can be applied to find albedo
at other wavelengths. This estimate yields area-averaged albedos that compare well with both surface
and satellite measurements of albedo. This method was then used for the very difficult measurement
problem of finding an effective albedo for surfaces near a land-water interface (Kassianov et al. 2017).

QUESTIONS
1. How is the broadband albedo of a surface measured?
2. Are albedos wavelength independent; that is, are they the same at all wavelengths? Name a
surface whose albedo is almost wavelength independent in the visible between 400 and 700 nm.
3. Does the white-sky albedo change with the angle of the Sun? Does the black-sky albedo?
4. Which surface type shows the greatest spectral diversity: A field of alfalfa or a field of
wheat stubble?
5. What would cause the ocean surface albedo to become higher?
6. Can you think of a way to measure albedo with a single sensor?
7. As the polar ice cap shrinks, the summer Sun sees more ocean and less ice. What happens
to the albedo of the polar region in summer?

REFERENCES
Ackerman, T., and G. M. Stokes. 2003. The  atmospheric radiation measurement program. Physics Today
56:38–45.
Augustine, J. A., J. J. DeLuisi, and C. N. Long. 2000. SURFRAD—A national surface radiation budget net-
work for atmospheric research. Bulletin of the American Meteorological Society 81:2341–2357.
Bowker, D. E., R. E. Davis, D. L. Myrick, K. Stacy, and W. T. Jones. 1985, June. Spectral reflectances of natu-
ral targets for use in remote sensing studies. NASA Reference Publication RP-1139. NASA Scientific
and Technical Information Branch.
Albedo 203

Kassianov, E., J. Barnard, C. Flynn, L. Riihimaki, L. K. Berg, and D. A. Rutan. 2017. Areal-averaged spectral
surface albedo in an Atlantic coastal area: Estimation from ground-based transmission. Atmosphere
8:123–135. doi:10.3390/atmos8070123.
Kassianov, E., J. Barnard, C. Flynn, L. Riihimaki, J. Michalsky and G. Hodges. 2014. Areal-averaged spectral
surface albedo from ground-based transmission data alone: Toward an operational retrieval. Atmosphere
5:597–621. doi:10.3390/atmos5030597.
McFarlane, S., K. Gaustad, C. Long, E. Mlawer, and J. Delamere. 2011. Development of a high resolution
spectral surface albedo product for the ARM Southern Great Plains Central Facility. Atmospheric
Measurement Technology 4:1713–1733.
Minnis, P., S. Mayor, W. L. Smith, Jr., and D. F. Young. 1997. Asymmetry in the diurnal variation of surface
albedo. IEEE Transactions on Geoscience and Remote Sensing 35:879–891.
Nicodemus, F. 1965. Directional reflectance and emissivity of an opaque surface. Applied Optics 4:767–775.
doi:10.1364/AO.4.000767.
Pu, R., S. Ge, N. M. Kelly, and P. Gong. 2003. Spectral absorption features as indicators of water status in
coast live oak (Quercus agrifolia) leaves. International Journal of Remote Sensing 24:1799–1810.
Stokes, G. M., and S. E. Schwartz. 1994. The  atmospheric radiation measurement (ARM) program:
Programmatic background and design of the cloud and radiation test bed. Bulletin of the American
Meteorological Society 75:1201–1221.
Yang, F., K. Mitchell, Y.-T. Hou, Y. Dai, X. Zeng, Z. Wang, and X.-Z. Liang. 2008. Dependence of land
surface albedo on solar zenith angle: Observations and model parameterization. Journal of Applied
Meteorology and Climatology 47:2963–2982. doi:10.1175/2008JAMC1843.1.
Zygielbaum, A. I., A. A. Gitelson, T. J. Arkebauer, and D. C. Rundquist. 2009. Non-destructive detection of
water stress and estimation of relative water content in maize. Geophysical Research Letters 36:L12403,
4 pp. doi:10.1029/2009GL038906.
9 Measuring Solar Radiation
on a Tilted Surface
It  is current practice, for evaluating the energy received by a tilted surface, to decompose
the solar radiation into three components which are treated independently: Direct beam, sky
diffuse and ground-reflected.
Richard Perez (1987)

9.1 INTRODUCTION
When evaluating the performance of a solar energy conversion system, it is necessary to know the
irradiance incident on the collector. The incident solar irradiance either can be estimated from direct
normal irradiance (DNI) and diffuse horizontal irradiance (DHI) values, or incident irradiance
can be measured. For  concentrating solar power (CSP) systems and concentrating photovoltaics
(CPV), DNI is the only irradiance measurement or model estimate needed to represent the energy
input to the collector. Global tilted irradiance (GTI) is needed for non-concentrating (flat plate) solar
collectors (ASTM 2008, 2009, 2010). GTI is the incident irradiance on a surface with a given tilt and
azimuth orientation. Surfaces of interest can vary from building façades facing north to flat rooftops
to any collector tilt and orientation in between.
When specifying collector tilts and orientations, the notation used in this book will be GTI
(tilt, orientation), where tilt is given in degrees from the horizontal, and orientation is in degrees
measured eastward from true north (0°). For  example, total irradiance on a surface tilted up 30
degrees from a horizontal surface and facing true south (180°) is designated GTI (30°, 180°).
For surfaces on one- or two-axis trackers, the tilt and orientation change over the day, and the nota-
tion GTI (one-axis, 0°) or GTI (two-axis) will be used. GTI (one-axis, 0°) is for a tracker that rotates
on a horizontal north-south axis. For trackers that rotate on an east-west axis, 90 degrees is used to
indicate the orientation of the axis.
It  is impractical to assess the solar resource by measuring solar irradiance on a surface at
every possible tilt and orientation. Therefore, models have been developed that use measured or
estimated DNI and DHI components to calculate an approximation for GTI. A tilted surface will
receive irradiance reflected from the ground and will receive reflected irradiance from nearby
objects that are within its field of view. The ground-reflected irradiance is often difficult to calcu-
late because the reflectance is not spatially uniform and the absorption and reflectance properties
of the surface change the spectral composition of the reflected irradiance from that of the incident
solar spectral distribution. For  a specularly reflecting surface, the angle of reflectance is equal
to the angle of incidence. However, natural surfaces are generally granular in character, so an
approximation is used to estimate the ground-reflected irradiance. A generic formula that is often
used to calculate GTI is

GTI = DNI ⋅ cos ( szaT ) + DHI ⋅ (1 + cos (T ) ) / 2 + GHI ⋅ ρ ⋅ (1 − cos (T ) ) / 2 (9.1)

where szaT is the angle of incidence of the DNI with respect to the tilted surface, T is the tilt angle
of the surface from the horizontal, and ρ is the average albedo or reflectivity of the ground near the
tilted surface (Liu and Jordan 1963). The angle between incident DNI and the normal to the tilted
surface (szaT ) can be calculated as the dot product of the two unit vectors.

205
206 Solar and Infrared Radiation Measurements

s = sin( sza) ⋅ cos( azm) x + sin( sza) ⋅ sin( azm) y − cos( sza) z (9.2)

Equations for sza and azm are given in Chapter 2, Equations (2.7) and (2.10), respectively.

s = sin(θ ) ⋅ cos(γ ) x + sin(θ ) ⋅ sin( a) y + cos(θ ) z (9.3)

where θ is the tilt of the surface from the horizon and γ is the azimuth of the surface with south equal
to 180°. The cosine of szaT is the dot product of unit vector to the Sun (Equation 9.2) and the unit
vector from the tilted surface (Equation 9.3), or

cos(szaT ) = sin(sza) ⋅ cos(azm) ⋅ sin(θ ) ⋅ cos(γ ) (9.4)


+ sin(sza) ⋅ sin(azm) ⋅ sin(θ ) ⋅ sin(γ ) − cos(sza) ⋅ cos(θ )

Although DNI can be measured with considerable accuracy and szaT can be calculated precisely
(Equation 9.4), the other two terms in the equation are approximations (see Appendix B for further
discussions on modeling GTI). Therefore, to know the solar radiation incident on a tilted solar col-
lector with any precision, GTI must be measured with a pyranometer.
Pyranometers were originally designed for the measurement of global horizontal irradiance
(GHI) and installed on horizontal surfaces. It is only with the need to measure the performance
of solar energy systems that measurements on other than horizontal surfaces were considered and
pyranometers were increasingly used on tilted surfaces. When tilted, the measurement performance
of the pyranometer can change. The nature and magnitude of the change depend on the design of the
pyranometer. The change in behavior of three types of pyranometers will be discussed:

1. Single black detector thermopile pyranometers


2. Black-and-white thermopile pyranometers
3. Photodiode pyranometers

9.2 EFFECT OF TILT ON SINGLE BLACK DETECTOR PYRANOMETERS


In single black detector thermopile pyranometers the conduction of heat from the black detector
through the thermopile to the body of the pyranometer is measured to determine the radiant energy
(solar irradiance) incident on the instrument. Extraneous heat flow by conduction, convection, and
radiative transfer systematically reduces heat flow through the thermopile and, hence, affects the
measurements. Each pyranometer model has its own design features to minimize these losses within
the pyranometer. Thermal isolation and insulation help reduce the conductive losses, and the use of
double domes over the black detector help to reduce the convective thermal losses. Still, there are
some losses due to conduction, radiation, and convection in these pyranometers.
Tilting a pyranometer from the horizontal affects the convective losses and radiative losses
(thermal offsets) of thermopile pyranometers. For single black detector thermopile pyranometers,
the conductive losses remain about the same as when they are used horizontally since orientation
typically does not affect conduction. The convective losses change because the heat flow patterns
within the domes change. Thermal offsets decrease as the pyranometer is tilted at a steeper angle
because the pyranometer sees more of the warmer ground and less of the colder sky. Figure 9.1
shows the thermal offsets from Eppley precision spectral pyranometers (PSPs) at different tilts
(Vignola et  al. 2007). As the pyranometer tilt increases, the observed thermal offsets decrease.
The pyranometers facing the ground (i.e., inverted) show almost no thermal offset. Of course, the
amount of radiative loss depends on the difference between the target temperature (e.g., the sky
dome temperature) and ambient temperature (instrument temperature) as discussed in Chapter 5.
Many factors influence sky temperature—primarily cloud cover and precipitable water vapor.
Measuring Solar Radiation on a Tilted Surface 207

FIGURE 9.1 Trend lines of nighttime values for tilted PSPs for June 2007. Note that as the pyranometer tilt
increases, the nighttime values get smaller. This result means that the IR loss becomes less as the pyranometer
“sees” more of the ground.

Sky  temperature can vary over time scales of seconds and minutes with underlying seasonal trends.
Therefore, pyranometers with minimal thermal offset losses are better suited for horizontal and
tilted measurements.
Tests of the effect of tilt on pyranometer performance are typically performed indoors under
controlled conditions and, consequently, do not show thermal offset effects. For single black detec-
tor thermopile instruments, the change in responsivity with tilt is minimal (on the order of ±1% to
±2%). It is presumed that these changes are related to changes in convective losses, but the exact
nature of the change and the magnitude vary with pyranometer model. Some pyranometers show a
slight responsivity increase with tilt, while others may have a slight decrease (Flowers 1984).
While there are accurate measurements of tilt effects indoors, outdoor evaluations of tilt effects
are very difficult because of the variation in ground reflection on the instrument and the absence of
an absolute pyranometer to provide a reference measurement. In addition, the reduction of thermal
offsets as the pyranometer is tilted and the imperfect Lambertian response of the pyranometer fur-
ther complicate the analysis. Typical uncertainties quoted for the tilt effect vary from 0.5 percent
for International Organization for Standardization (ISO) class A instruments and up to 2 percent for
class B instruments (Hulstrom 1989). Although these values were determined indoors and provide
an estimate of the uncertainties, they may not be as accurate under actual field operating condi-
tions. Note that the uncertainty caused by tilting a pyranometer must be added in quadrature to the
uncertainty determined for the pyranometer when it is installed horizontally. Generally, tilted mea-
surements with a given pyranometer will always have a larger uncertainty than GHI measurements
with the same instrument. The exception is when the reduction in the magnitude of thermal offsets
exceeds the added uncertainty of the tilt effect.
Pyranometers used to measure tilted irradiance should be in a ventilator (VEN) or shielded from
direct sunlight. Most pyranometer sun shields are designed to shield the body of the pyranometer
from direct sunlight in a horizontal position. When the pyranometer is on a tilted surface, sunlight
can often strike the body of the pyranometer. Since the heat sink for the pyranometer is associated
with the body of the pyranometer, direct sunlight will reduce the temperature difference between
the black disk sensor and the heat sink, thus reducing the tilted irradiance measured.
208 Solar and Infrared Radiation Measurements

9.3 EFFECT OF TILT ON BLACK-AND-WHITE PYRANOMETERS


The effect of tilt on black-and-white pyranometers is greater than on single black detector thermo-
pile pyranometers. It is postulated that convective losses are responsible for most of the effects of tilt
on pyranometer responsivity (Hulstrom 1989). Therefore, it might be expected that black-and-white
pyranometers, with only a single dome, are more affected by tilt, and this is indeed the case.
In the lab, the pyranometer performance is measured using a light source that is perpendicular to
the pyranometer, and the whole platform is tilted at various angles; the black-and-white pyranom-
eter shows greater tilt effects than the single black detector pyranometer.
The deviation of a black-and-white pyranometer response at tilt compared with a horizontal
pyranometer varies from about 3 percent (Wardle and McKay 1984; Wardle et  al. 1996) to as
much as 8 percent (McArthur et al. 1995). It is uncertain how these losses translate to field con-
ditions where convective losses are often larger than experienced in the laboratory as a result of
wind blowing across the pyranometer dome and the naturally changing temperatures and relative
humidity of the air. The lack of an absolute pyranometer makes evaluation of tilt effects in the field
difficult to measure and characterize. Furthermore, it is difficult to isolate tilt effects from other
issues in the field. Even the best pyranometers have at least 2–3 percent absolute uncertainty and
a non-uniform Lambertian response. In addition, the responsivity of black-and-white pyranom-
eters has an azimuthal dependence as well as an imperfect Lambertian response. Isolating a 2 to
8 percent tilt effect from the other effects is fraught with uncertainty. The change in responsivity
as a function of tilt is shown in Figure 9.2 for a Schenk Star pyranometer. The uncertainty result-
ing from the tilt effect has to be added in quadrature to the typical black-and-white instrument
uncertainty of ±5 percent. Since this uncertainty can lead to estimated measurement uncertain-
ties of over 9 percent, it is not considered good practice to use black-and-white pyranometers on
tilted surfaces.

FIGURE 9.2 Relative effect of tilt on a Schenk pyranometer. Data taken indoors. (From Andersson, H.E.B.
et  al., Calibration of pyranometers, Report SP-PAPP 1981: 7, Statens Provningsanstalt  [National Testing
Institute], Borås, Sweden, 1981.)
Measuring Solar Radiation on a Tilted Surface 209

9.4 EFFECT OF TILT ON PHOTODIODE PYRANOMETERS


It  would not  be expected for photodiode pyranometers to be affected by tilt because they are
not thermal instruments. Photodiode pyranometers, in fact, often are used to test the effect of tilt on
thermopile pyranometers under laboratory conditions where they serve as the control or reference
sensor. As described in Chapter 5, photodiode pyranometers typically fall under ISO type C classi-
fication because of their sensitivity to the spectral distribution of the incident radiation. In the field,
the change in spectral composition of reflected light affects the measurement of tilted irradiance
by photodiode pyranometers. In general, ground-reflected irradiance, especially from vegetation, is
shifted more to the infrared (IR) portion of the spectrum. On a 90 degree tilted surface, the ground-
reflected irradiance is roughly equal to one-half the GHI times the reflectivity of the surface (see
Equation 9.1). Typical ground reflectivity is approximately 0.2 for non-snow-covered surfaces cov-
ered with vegetation. Hence, the contribution of the ground-reflected irradiance to a 90 degree tilted
pyranometer is about 10 percent of the GHI, and for a 45 degree tilted pyranometer the reflected
contribution is only about 3 percent of the GHI. Therefore, even a 30 percent change in responsiv-
ity resulting from a spectral shift in the irradiance would lead to an increase in uncertainty in the
GTI measurement of only about 1 percent for the 45 degree tilt case. Of course, the initial absolute
uncertainty in the photodiode reading is about 5 percent.
Tilting a pyranometer toward the Sun enables it to increase the radiant flux incident on the sensor
by reducing the angle of incidence (Vignola et al. 2017a, 2017b, 2018). By tilting, the amount of inci-
dent solar radiation on the sensor during the morning and evening can be significantly increased.
This result is especially true for pyranometers on one-axis to two-axis tracking surfaces. Since the
change in spectral distribution of GTI is significantly shifted toward the red in the morning and
evening hours, there will a significant increase in the responsivity of the photodiode pyranometer
during these time periods (see Chapter 5 for an explanation). The ratio of a LI-200SA pyranom-
eter divided by a CM11 (thermopile-based) pyranometer on a two-axis tracking surface is shown
in Figure 9.3. When the solar zenith angle reaches 75° to 80° the change in spectral distribution

FIGURE  9.3 Ratio of a LI-200SA  pyranometer measurements divided by the CM 11 pyranometer mea-
surements on a two-axis tracking surface under clear skies on July 1, 2017, in Golden, Colorado (shown as
the black line). A model for the effects of changing spectral distribution based on air mass (see Chapter 7) is
included for comparison. The data are normalized to 1 at a sza of 45°.
210 Solar and Infrared Radiation Measurements

increases the LI-200SA measurements by 10–15 percent. The data are taken on a clear day and the
temperature effect on the measurements over the day is less than 0.5 percent. A model to account
for the effect of the change in spectral distribution and based on changes in air mass is shown in the
dashed gray line. The close comparison between the model and the changed results indicate that
most of this change is related to change in spectral distribution and not related to effects of ground-
reflected irradiance in this case. So, while a photodiode pyranometer does not have its responsivity
change when the instrument is tilted, the change in spectral distribution of the source does signifi-
cantly affect the measurement (King et al. 1998; Vignola et al. 2016).
A clear-day comparison of thermopile and photodiode pyranometers is shown in Figure 9.4 for
a southern orientation with tilts of 0, 40, and 90 degrees and in Figure  9.5 for 90 degree tilted
pyranometers facing north, east, south, and west. Data are taken from the National Renewable Energy
Laboratory (NREL) Solar Radiation Research Laboratory (SRRL) Baseline Measurement System
using PSP and LI-200 pyranometers (https://midcdmz.nrel.gov/srrl_bms/). The difference between
the photodiode and thermopile pyranometers is divided by the thermopile reading to obtain the
percent difference. Average nighttime readings were subtracted from the daytime values to reduce
the effect of the thermal offset on the values. Differences vary between –5 and +5 percent for the
south-facing pyranometers and 0 and 15 percent for the 90 degree tilted pyranometers, depending on
the time of day and the tilt and orientation of the instruments. The pattern of difference changes for
every orientation. By mounting the pyranometer on a two-axis tracker, such as shown in Figure 9.3,
shifts caused by changing spectral response can be separated from deviations from a Lambertian
cosine response. After the effect of changes in spectral distribution are well characterized, the devia-
tions from a Lambertian cosine response can be studied. Again, the absence of an absolute pyranom-
eter makes it difficult to precisely determine the biases caused by tilting the pyranometer.

FIGURE 9.4 Clear-day comparison of thermopile Eppley PSP with a LI-COR LI-200 photodiode pyranom-
eter on south-facing surfaces tilted 0, 40, and 90 degrees. Data from the Solar Radiation Research Laboratory
at NREL on September 29, 2008. The thermopile instrument agrees rather well over the day with the photodi-
ode response increasing in the morning and afternoon. At tilts of 40 and 90 degrees, the behavior is different
with deviations due to cosine response, spectral reflectivity response, and tilt response. It is difficult to isolate
and quantify the various sources of systematic biases.
Measuring Solar Radiation on a Tilted Surface 211

FIGURE  9.5 Clear-day comparison of thermopile Eppley PSP pyranometer with a LI-COR LI-200
photodiode pyranometer on north, east, south, and west surfaces tilted 90 degrees. Data taken at the Solar
Radiation Research Laboratory at NREL on September 29, 2008. Differences range from 0 to 15 percent over
the day. The peak differences of the east- and west-facing pyranometers around solar noon is likely the cat
ears effect of the LI-200 pyranometer discussed in Chapter 5.

9.5 RECOMMENDATIONS FOR TILTED IRRADIANCE MEASUREMENTS


The accuracy of tilted irradiance measurements will always be less than the accuracy of horizontal
measurements because the changes in pyranometer performance caused by tilting must be added
to the uncertainty of the performance when the pyranometer is horizontal. Most broadband ther-
mopile pyranometers are subject to almost negligible spectral shifts of ground-reflected irradiance
since glass domes used on thermopile instruments block none or only a small portion of the near-
IR irradiance. For the very best measurements on tilted surfaces, single black detector thermopile
pyranometers that use double domes made of material that uniformly pass a broad range of wave-
lengths yield the most accurate broadband measurements of GTI. Single black detector thermopile
pyranometers with double domes that pass a slightly narrower band of wavelengths are the second
choice. Photodiode pyranometers and reference solar cells have no tilt effect and are commonly
used to monitor photovoltaic (PV) panels because PV panels have a similar response to changes in
spectral distribution (IEC 2007). However, the spectral response is dependent upon the character-
istics of the PV technology used and even slight differences in the responsivities can cause measur-
able differences in the outcome (Chapter 7). In addition, models could be used to remove the some
of the biases of photodiode pyranometers used on tilted surface to obtain reliable estimates of GTI
on the tilted surface (Vignola et al. 2017a, 2017b, 2018). Knowing the limitations of photodiode
pyranometers, relative performance can be monitored as long as measurements are compared at
similar sza values and similar atmospheric conditions.
The specifications and measurement characteristics of the pyranometer on the horizontal surface
helps determine the quality of the tilted measurements from that pyranometer. Pyranometers that
make good GHI measurements have a likelihood of making good GTI measurements. Of course,
212 Solar and Infrared Radiation Measurements

the shallower the tilt, the less likely there will be a significant tilt effect for any pyranometer. Black-
and-white pyranometers have significant tilt effects and are not  recommended for use on tilted
surfaces.

9.6 MODELING PHOTOVOLTAIC SYSTEM PERFORMANCE


WITH DATA FROM PHOTODIODE PYRANOMETERS
Photodiode pyranometers have spectral responses similar to silicon-based PV modules, and, conse-
quently, these pyranometers may correlate better with the output of a PV system than a broadband
thermopile pyranometer. To model the performance of a PV system, not only is it necessary to accu-
rately know the total incident solar irradiance on the module, but also the incident solar radiation
must be separated into its diffuse and direct components to properly calculate the transmission of
the solar radiation through the glazing. The separation of diffuse and beam components presents a
special problem for tilted measurements made with photodiode pyranometers because the spectral
composition of the direct and diffuse components is significantly different on clear days and the
responsivity of the photodiode is different for the direct and diffuse components. Properly matched
reference cells should have glazings that are similar to the PV module and hence the transmission
through the glazing should be similar. This issue is the reason reference cells do not make good
pyranometers, but if the interest is mainly in direct comparison with a PV module, then they might
be a useful tool for determining PV performance relative to a reference cell calibrated to represent
standard rating conditions (i.e., AM 1.5 solar spectrum, temperature equals 25°C, and incident solar
radiation is 1,000 watts per square meter (Wm−2)).
The glazing on the front of a solar PV module reflects an increasing percentage of the irradi-
ance as the angle of incidence increases (Fresnel reflection) (see Figure  9.6). The  glazing does
not  exhibit a Lambertian response, which would be uniform with angle of incidence. The  differ-
ence from a Lambertian response is especially divergent for angles of incidence larger than 50

FIGURE  9.6 Generic plot of the reflectivity on incident solar radiation on a sheet of glass. The  percent-
age of reflected light remains fairly constant until the angle of incidence gets to about 45°. To illustrate
the potential complexity of the reflection of light from glass, the reflection of polarized light is also shown.
Antireflective coatings are used with PV modules, and these coatings can reduce the reflected light by a factor
of two or more. (From Andersson, H.E.B. et al., Calibration of pyranometers, Report SP-PAPP 1981: 7, Statens
Provningsanstalt [National Testing Institute], Borås, Sweden, 1981. With permission.)
Measuring Solar Radiation on a Tilted Surface 213

degrees. DNI  values and the glazing’s refractive index (often modified by an antireflective coat-
ing) are required for accurate estimates of the transmission through the glazing to the solar cell.
PV system performance models are developed and tested with DNI measurements obtained from
pyrheliometers and GTI measurements obtained using high-quality pyranometers. The PV system
performance models incorporate the effects of changes in the spectral composition incident on the
solar cell. This issue is the reason that there is a better spectral match of the photodiode pyranometer
response; however, using photodiode pyranometer data causes the performance model to overcorrect
for the changes in spectral composition. This outcome is analogous to counting the effect of changes
in the spectral composition twice.
The need to separate DNI contribution from GTI measurement to properly account for the reflec-
tivity of the glazing further complicates the performance estimates if a photodiode pyranometer is
used for GTI measurements. DNI values can be obtained from measurements made with pyrheli-
ometers, estimated from rotating shadowband radiometer (RSR) measurements, or modeled from
measured GHI values. As discussed in Chapter 5, photodiode pyranometers systematically under-
estimate the DHI value on clear days by 20–30 percent, and if a pyrheliometer is used to obtain the
DNI value, the diffuse and ground-reflected irradiance reading of the photodiode pyranometer will
underestimate contributions of the diffuse component to the total energy incident on the solar cells.
If not taken into account, this systematic underestimation of the incident energy can overestimate
the efficiency of the solar system.
If DNI is obtained with an RSR, it is important to thoroughly understand how DHI and DNI
values are obtained and whether responsivities have been adjusted to remove the systematic
biases caused by the differences in the spectral composition of DNI and DHI irradiance. If DNI
is obtained from an RSR that has its DHI measurements adjusted to remove the underestimation
of the diffuse component, then DNI can be treated as if it came from a pyrheliometer, except the
uncertainty of the DNI value is increased. If the DHI values were not spectrally corrected, then
an algorithm must be applied to correct for the effects caused by the underreporting of DHI on
clear days.
Correlations between DNI and tilted measurements are crude at best (Vignola 2003).
Therefore, if DNI measurements are not available, GHI measurements are needed to estimate
the DNI value required for modeling PV module performance. Models correlating DNI with
GHI are usually derived from GHI measurements made with high-quality pyranometers using
single black thermopile detectors. Because each type of pyranometer has its characteristic sys-
tematic biases, using correlations derived from one type of pyranometer with measurements
made with a different model or type of pyranometer increases the uncertainty in the results and
may introduce systematic biases. Therefore, if a photodiode-based pyranometer is used for GHI
measurement, then a DNI–GHI correlation derived using GHI measurements with a photodiode
pyranometer should be used. The same applies to GHI measurements obtained from thermopile-
based pyranometers.
Spectral effects from ground-reflected irradiance are similar for a tilted photodiode pyranometer
and solar PV cells. So, although good correlations with solar module performance can be obtained,
the absolute uncertainty in module efficiency is larger than the uncertainty obtained when irradi-
ance measurements are made with thermopile-based pyrheliometers and pyranometers. The  PV
Performance Modeling Collaborative offers information about this and other topics related to
modeling PV performance (https://pvpmc.sandia.gov/).

QUESTIONS
1. What is the best pyranometer to use for tilted surface measurements?
2. What type of pyranometer should not be used for tilted measurements?
3. Why do the design features of a photodiode pyranometer make it compare better with PV
module output and at the same time reduce its reliability for tilted measurements?
214 Solar and Infrared Radiation Measurements

REFERENCES
Andersson, H. E. B., L. Liedquist, J. Lindblad, and L. A. Norsten. 1981. Calibration of pyranometers. Report
SP-PAPP 1981:7. Borås, Sweden: Statens Provningsanstalt (National Testing Institute).
ASTM. 2008. Standard Test Methods for Electrical Performance of Non-concentrator Terrestrial
Photovoltaic Modules and Arrays Using Reference Cells, Standard E1036-08. West Conshohocken,
PA: ASTM International.
ASTM. 2009. Standard Test Method for Electrical Performance of Photovoltaic Cells Using Reference Cells
Under Simulated Sunlight, Standard E948-09. West Conshohocken, PA: ASTM International.
ASTM. 2010. Standard Test Method for Calibration of Non-concentrator Photovoltaic Secondary Reference
Cells, Standard E1362-10. West Conshohocken, PA: ASTM International.
Flowers, E. C. 1984. Solar radiation measurements. In Proceedings of the Recent Advances in Pyranometry,
International Energy Agency IEA  Task IX Solar Radiation and Pyranometer Studies. Downsview,
Canada: Atmospheric Environment Service.
Hulstrom, R. L., ed. 1989. Solar Resources. Cambridge, MA: MIT Press.
IEC. See International Electrotechnical Commission.
International Electrotechnical Commission. 2007. IEC 60904-2:Photovoltaic Devices—Part 2: Requirements
for Reference Solar Cells. London, UK: British Standards Institution.
King, M. L., W. E. Boyson, B. R. Hansen, and W. I. Bower. 1998. Improved accuracy for low-cost solar irra-
diance sensors. Paper Presented at the 2nd World Conference and Exhibition of Photovoltaic Solar
Energy Conversion, July 6–10, Vienna, Austria.
Liu, B. Y. H., and R. C. Jordan. 1963. A rational procedure for predicting the long-term average performance
of flat-plate solar energy collectors. Solar Energy 7(2): 53–74.
McArthur, L. J. B., L. Dahlgren, K. Dehne, M. Hämäläinen, L. Liedquist, G. Maxwell, C. V. Wells, and D. I.
Wardle. 1995. Using Pyranometers in Tests of Solar Energy Converters. International Energy Agency,
Solar Heating and Cooling Programme, Task 9, IEASNCP-9F1. Downsview, Canada: Experimental
Studies Division, Atmospheric Environmental Service.
Vignola, F. 2003. Beam-tilted correlations. Paper Presented at the Solar 2003 American Solar Energy Society
Conference, Austin, TX.
Vignola, F., C. Chiu, J. Peterson, M. Dooraghi, and M. Sengupta. 2017a. Comparison and analysis of instru-
ments measuring plane-of-array irradiance for one-axis tracking PV systems. IEEE PVSC Conference,
Washington, DC.
Vignola, F., Z. Derocher, J. Peterson, L. Vuilleumier, C. Felix, J. Grobner, and N. Kouremeti. 2016. Effects of
changing spectral radiation distribution on the performance of photodiode pyranometers. Solar Energy
129:224–230.
Vignola, F., C. N. Long, and I. Reda. 2007. Evaluation of methods to correct for IR loss in Eppley PSP diffuse
measurements. Paper Presented at the SPIE Conference, San Diego, CA.
Vignola, F., J. Peterson, M. Dooraghi, M. Sengupta, and F. Mavromatakis. 2017b. Comparison of pyranom-
eters and reference cells on fixed and one-axis tracking. American Solar Energy Society Conference,
Denver, CO, October 9–12.
Vignola, F., J. Peterson, R. Kessler, M. Dooraghi, M. Sengupta, and F. Mavromatakis. 2018. Evaluation of
photodiode-based pyranometers and reference solar cells on a two-axis tracking system. IEE PVSC
Conference, Honolulu, HI.
Wardle, D. I., L. Dahlgren, K. Dehne, L. Liedquist, L. J. B. McArthur, Y. Miyake, O. Motschka, C. A. Velds,
and C. V. Wells. 1996. Improved Measurements of Solar Irradiance by Means of Detailed Pyranometer
Characterisation. International Energy Agency, Solar Heating and Cooling Programme, Task9,
IEA-SHCP-9C-2. Downsview, Canada: National Atmospheric Radiation Centre, Atmospheric
Environmental Service.
Wardle, D. L., and D. C. McKay, eds. 1984. Recent advances in pyranometers. Paper Presented at the
Symposium Proceedings of the Swedish Meteorological and Hydrological Institute, Norrkoping,
Sweden.
10 Shadowband Radiometers

The total assembly is called the “dial” radiometer both because the rotating shade resembles
the face of a dial with a revolving arm and because the device is used to measure “D-and-I-AL”
quantities (Diffuse, Incident direct-beam, and globAL irradiance).
Marvin L. Wesely (1944–2003)

10.1 INTRODUCTION
To fully characterize the solar resource, the three components global horizontal irradiance
(GHI), direct normal irradiance (DNI), and diffuse horizontal irradiance (DHI) that contribute
to the incident energy must be known. Measuring the three components is expensive especially
because a tracker is necessary to measure the DNI and DHI components with a pyrheliometer
and shaded pyranometer, respectively. Efforts to provide a more economical approach for pro-
ducing the three solar irradiance components have been ongoing since 1982 when the rotating
shadowband radiometer (RSR) was introduced (Wesely 1982). In this chapter, the RSR and the
work to improve the measurements from this type of measurement system will be discussed.
A  device using similar methodology called a Multi-Filter Rotating Shadowband Radiometer
(MFRSR) was developed to take three component spectral measurements (Harrison et al. 1994)
and will be discussed next. More recently the SPN1 Sunshine Pyranometer was introduced with
a novel fixed shadow mask and seven fast-response thermopile sensors (Badosa et  al. 2014).
The measurement concept and performance characterization of the SPN1 Sunshine Pyranometer
will also be discussed.

10.2 INTRODUCTION TO THE ROTATING SHADOWBAND RADIOMETER


A  simple RSR was introduced by Wesely (1982), as discussed in Chapter  3; it used a LI-COR
LI-200 silicon photodiode pyranometer because a rapid response time (10 microseconds (µs) for the
LI-200) was needed as the shading band passed over the detector to block direct sunlight. The band
rotated at constant speed and completely shaded the diffuser every 4 or 5 minutes depending on
the motor selected. The output from the pyranometer was a current; therefore, a low-temperature
coefficient resistor was used in the circuit to produce a voltage that was output to a chart recorder.
An upper envelope of GHI bound the output on a clear day, and a lower envelope represented DHI
(Figure 10.1). DNI could be calculated from the difference in the GHI and DHI after division by the
cosine of the solar zenith angle. The labor-intensive process of analyzing the chart manually limited
the analyses of these paper charts to produce numerical data.
Michalsky et  al. (1986) built a microprocessor-controlled version of the RSR that recorded
four measurements from the photodiode pyranometer for each rotation of the stepper motor-
driven band. The measurements were made with the band in the nadir position to measure unob-
structed global horizontal irradiance, blocking the Sun to measure DHI, and with the band offset
symmetrically slightly before and slightly after the Sun-blocking position to estimate the excess
diffuse sky radiation blocked by the band during the Sun-blocked measurement. The Sun-blocked
position of the band prevents the direct Sun from reaching the receiver but also blocks consider-
able diffuse skylight. An estimate of this band-correction is calculated using the two side-band
measurements.

215
216 Solar and Infrared Radiation Measurements

FIGURE 10.1 Chart recorder output from a Wesely-type constant-speed RSR. The bottom of the envelope
represents DHI, and the upper envelope GHI. By differencing these quantities the direct irradiance on the
horizontal can be calculated and is proportional to the length of the vertical lines. (Courtesy of the American
Meteorological Society, Boston, MA.)

Yankee Environmental Systems, Inc. (http://www.yesinc.com) currently manufactures a seven-


channel RSR (MFR-7) (Figure 10.2). Six of the channels are 10 nanometer–wide spectral channels
in the visible and near infrared (IR); the seventh channel is now  a broadband thermopile solar
sensor, although the initial version used a silicon photodiode. It uses the same band rotation and
sampling algorithm as the Michalsky et al. (1986) design.
Edward Kern developed the first version of the RSR that is controlled by a commercial data log-
ger (Figure 10.3). This instrument rotates a band at high speed with concurrent and rapid sampling
of the photodiode pyranometer. The minimum signal is used as the first-order diffuse signal, and
measurements just outside this minimum value are used to estimate a shadowband correction for
excess forward-scattered diffuse sunlight blockage. It is described at http://www.irradiance.com/.

FIGURE 10.2 The Yankee Environmental Systems, Inc. seven-channel MFRSR model MFR-7.


Shadowband Radiometers 217

FIGURE 10.3 The Irradiance, Inc. RSR version 2 (RSR2). (Courtesy of Ed Kern.)

Solar Millennium AG developed a version of the RSR that operates in a similar manner to
the Irradiance, Inc.’s instrument. That instrument is now  available from Reichert GmbH (www.
reichertgmbh.de) and is called the RSP G4.
CSP-Services GmbH manufactures the RSI, another variant of the Irradiance RSR that appears
more mechanically robust than the Irradiance, Inc. or Reichert GmbH instruments (see http://www.
cspservices.de/media/csps/CSPS-RSI-1510.pdf).
All of the latter three manufacturers use the LI-COR 200 silicon photodiode sensor for their mea-
surements but process the raw output data into corrected data using their own proprietary routines.
Prede Co. Ltd. of Japan (http://www.prede.com) makes the rotating shadow blade (PRB-100),
which is advertised to work with either thermopile or photodiode sensors of any manufacturer.
The authors have no experience with this device.
RSRs are popular because they provide estimates of all three solar components for a much lower
cost than do traditional thermopile radiometers that require an expensive tracker to perform the DNI
and DHI measurements and use more expensive thermopile pyranometers and a thermopile pyrhe-
liometer. To date, the RSR’s major drawback is that they require several corrections that limit their
accuracy compared with the accuracy of thermopile radiometers. When a few percent accuracy in
solar irradiance matters, as in multimillion-dollar solar power plants, then a properly maintained
Class A or better thermopile measurement system is the preferred choice. The next two sections
describe the two fundamentally different implementations of the RSR.

10.3 ROTATING SHADOWBAND RADIOMETER USING SILICON DETECTOR


Before the development of the MFRSR there was a rotating shadowband single silicon cell
pyranometer (LI-COR, Inc. LI-200) instrument that sampled according to the MFRSR algorithm
described in Section 10.1 (i.e., only four measurements were made rather than continuous sampling).
218 Solar and Infrared Radiation Measurements

Michalsky et al. (1987) describe a simple empirical technique to correct the silicon response of this
instrument to mimic the thermopile responses. In another paper, Michalsky et al. (1991) used correc-
tions based on a parameterization of sky conditions to improve the estimates of GHI, DNI, and DHI.
The  rest of this section describes instruments that use a sampling algorithm similar to that
first developed by Edward Kern that use a commercial photodiode pyranometer, a commercial
data-logger, and a band driven to pass over the pyranometer on command. The  basic correction
algorithms originally used for the Irradiance, Inc. RSR data are available in Augustyn et al. (2004).
The  need to automatically measure the direct and diffuse irradiance inexpensively led to the
development of the RSR2 by Irradiance, Inc., which was originally called the rotating shadowband
pyranometer (RSP). When the shadowband arm starts its motion to shadow the pyranometer, the
data logger goes into a rapid data-sampling mode and measures the minimum irradiance during
the sweep of the band (Figure 10.4). This minimum irradiance is the uncorrected estimate of dif-
fuse irradiance. Measurements just before and just after shading are used to estimate the diffuse
irradiance blocked by the band during the measurements when the solar disk is totally eclipsed.
This  action provides a positive correction that is added to the uncorrected DHI for a corrected
DHI estimate. The samples taken when the band is outside the horizontal field of view is the GHI.
The difference between GHI and DHI is the direct irradiance projected on the horizontal (DrHI).
DNI is calculated by dividing the DrHI by the cosine of the incident angle, which is the solar zenith
angle for a horizontal pyranometer. (The standard abbreviation for DHI should not be confused with
the direct horizontal irradiance DrHI.)
Specifically, DNI is calculated using

DNI = DrHI / cos( sza) (10.1)

where sza is the solar zenith angle.


In principle, the installation and operation of the RSR2 is simple and straightforward. The RSR2
can run unattended in the field for long periods, and the operation of the instrument does not require
a complex and, hence, expensive tracker to point the instrument at the Sun. The  instrument is
powered by a small photovoltaic (PV) module, is easy to align, and is easily integrated into an
automated monitoring station (see Figure 10.3).

FIGURE 10.4 The response of the LI-COR, Inc. LI-200 to shading by the RSR2 band.
Shadowband Radiometers 219

Although only one pyranometer is required to obtain GHI and DHI measurements to compute
the DNI, an auxiliary pyranometer is an option for automated data quality control, and newer
models integrate a second unshaded pyranometer into the design (hence, the RSR2 designation).
The National Renewable Energy Laboratory (NREL) developed a quality control check for an RSR
with a primary pyranometer for the GHI, DHI, and DNI measurements and a secondary pyranom-
eter for GHI measurements. If the performance of either pyranometer degrades because of soiling,
sensor degradation, or other problems, then the output ratio between the two pyranometers changes.
The auxiliary pyranometer also can be tilted at other angles, for example, due south by the station
latitude from the horizontal. A tilted pyranometer tends to be less affected by dirt, moisture, frost,
or snow, especially if it is tilted at 90 degrees. However, the quality control check becomes com-
plicated and less exact, not only because the pyranometers are no longer making the same mea-
surement, but because ground-reflected light has a significant wavelength dependence, as does the
photodiode pyranometer. Further, tilted pyranometers will not see the DNI component for hours
during the months when the Sun rises and sets behind the tilted pyranometer.
In practice, the accuracy of the RSR2 is only as good as the pyranometer used, and the uncer-
tainties in measured DHI and computed DNI are typically greater than the uncertainties associated
with GHI measurements. There are significant statistical uncertainties associated with infrequent
sampling and large systematic biases associated with using a photodiode pyranometer, especially
for DHI measurements. To record DHI when the band sweeps in front of the pyranometer, the
pyranometer needs a fast response time. This  requirement limits the choice of pyranometers.
Thermopile pyranometers are generally better instruments for broadband shortwave measurements
because they have a uniform response over most solar wavelengths; however, most have a response
that is too slow for typical shadowband operation. Consequently, photodiode pyranometers are pre-
ferred in RSRs.
Large random uncertainties are introduced by the lower sampling rates of GHI and DHI using
the RSR compared with the high frequency sampling possible with thermopile instruments. Variable
cloudiness will increase the scatter of the RSR values for DNI and DHI when compared with ther-
mopile pyrheliometers and pyranometers that sample continuously over the measurement period.
Newer models of the Irradiance, Inc. RSR2 are able to take several readings during a minute and
initiate more frequent band rotation and sampling during periods of rapidly changing GHI, whereas
the original model RSR took a single reading, typically every minute. The scatter in the measured
data is random, and when averaged over longer periods the two measurements approach the same
mean value. Figure 10.5 clearly indicates the reduction in scatter when 1-minute data are compared
with 15-minute data for the same data set. It is evident in Figure 10.5 that there is still room for
improvement in DNI corrections.
A systematic error in some of the early RSRs was the calculation of the cosine of the solar zenith
angle. Because of limited memory in the data acquisition and control system, an approximate for-
mula for the solar position was used. This approximation led to uncertainties in the cosine of the
solar zenith angle, and these errors are imbedded in the calculated values of DNI (Vignola 2006).
Systematic errors in DHI occur because the shadowband obscures more than just the Sun when
it passes in front of the pyranometer. Samples taken near the shading minimum are used to partially
correct for the excess sky blockage by the shadowband (Figure 10.4).
The photodiode pyranometer has a spectral dependence. If this pyranometer is calibrated using
a component sum GHI measurement with clear skies, for example, then a DHI measurement will be
nearly correct if the sky is cloudy because the cloudy sky spectrum is similar to the GHI clear-sky
spectrum. If there is a clear blue sky, then RSR DHI measurements can be 30 percent lower than a
reference DHI measurement. Said another way, the DHI responsivity of the photodiode pyranometer
can be almost 30 percent less for a clear blue sky than for an overcast sky. The systematic reduction
of DHI for a clear, blue sky results in an overestimate of the calculated DNI by about 5 percent during
clear-sky episodes. The deviation from true cosine response and the temperature dependence of the
photodiode pyranometer also cause systematic biases. King et al. (1998) characterized many of the
220 Solar and Infrared Radiation Measurements

FIGURE 10.5 While there is much scatter caused by the less frequent sampling of the photodiode pyranom-
eter, the mean of the photodiode DNI measurements approaches the thermopile results. Note the reduction in
scatter with time averaging.

systematic biases of the LI-COR Model LI-200 pyranometer that is used in the Irradiance, Inc. RSR2
instruments. Typical deviations from true cosine response were shown in Figure 5.28 in Chapter 5.
Early versions of the algorithms used to remove the systematic biases of the LI-200 in RSRs used the
angular response and temperature adjustment formulas based on the analysis of these instruments by
King et al. (1998). When plotted against time of day, the LI-200’s responsivity takes on a shape simi-
lar to a cat’s front profile with peaks occurring near a solar zenith angle of 81° (Augustyn et al. 2002).
This cat ears effect (see Figure 5.29 in Chapter 5) is similar for all LI-COR pyranometers and is
related to the shape of the diffuser and artificial horizon on the pyranometer. This diffuser produces
a good cosine response over most of the range of solar zenith angles but has difficulties producing a
precise cosine response at solar zenith angles greater than 75° (as do all pyranometers whether they
are photodiode or thermopile based). GHI adjustments in the original Irradiance, Inc. RSR adopted
the formulas suggested by King et al. (1998) and the cat ears corrections in Augustyn et al. (2004).
Shadowband Radiometers 221

Vignola (1999) published a study on the diffuse spectral responsivity of the LI-COR pyranom-
eter. This study showed the magnitude of the deviation and also provided a model of the diffuse
responsivity as a function of the clearness index. In 2005, NREL funded a study to see if the results
from an RSR with a LI-COR pyranometer could be improved by removing the systematic biases
(Vignola 2006). The RSR was similar to the previous RSR, but an adjustment algorithm was added
to the data-logging program in an attempt to remove the systematic biases that were produced.
First, GHI and DHI values were adjusted to eliminate the effects of temperature, and then the
cat  ears corrections helped improve the cosine response. Next, the diffuse spectral responsivity
was corrected using an algorithm based on the study by Vignola (1999) but using the corrected GHI
values. Finally, a more accurate algorithm was used to calculate the cosine of the solar zenith angle,
which was then used to calculate DNI from DrHI values.
Other improvements in design and operation were introduced at the same time. The sampling
frequency has been increased from 1 per minute to as many as 20 per minute. The higher sampling
rates better emulate pyrheliometer values that are obtained from 1- or 2-second samples and aver-
aged into 1-minute values. Temperature measurements of the pyranometer also were introduced to
better determine the temperature correction to the responsivity. Tests show that the adjustments to
the RSR data significantly improve the comparison to reference measurements. With the DHI adjust-
ment, DHI values obtained from a Schenk (black-and-white pyranometer) and DHI values from
RSRs agree to within a few percent over the full range of clearness index (Figure 10.6). Figure 10.6
also indicates the improvement in scatter with averaging time, as was shown for DNI in Figure 10.5.
Two additional measurement artifacts can appear in the RSR data. The first relates to the LI-COR
pyranometer. During some mornings, a bead of moisture collects on the diffuser of the pyranom-
eter. This bead directs sunlight striking at low angles onto the photodiode. This sunlight slightly
increases the GHI but does affect the DHI as it comes from all parts of the sky. When the DHI
is subtracted from the GHI, the slight increase in GHI increases the calculated DNI significantly
because early in the morning the cosine of the large solar zenith angle is small, and when dividing
this into the DrHI, the error is amplified. Occasionally the band does not rotate when it should, and
the DHI value exhibits a sudden spike while DNI has a sudden dip. This problem is most evident on
clear days. In the original RSPs this problem appears to increase as the instrument ages. A software
modification has helped to remove this problem, but not eliminate it entirely, in later versions of
the RSR.
The  adjustment methods outlined in this chapter should provide useful corrections for all
RSR-type instruments since they all use LI-COR pyranometers. These methods of corrections and
more recent efforts to improve on these are more comprehensively described in Wilbert et al. (2015)
and Vignola et al. (2016, 2017). Since the spectral distributions of DNI and DHI irradiance differ,
these newer correction methods are based on the calibration of DNI using a DNI standard pyrheli-
ometer. It has been suggested that DNI and DHI components be treated separately and that unique
adjustments be applied to DNI and DHI components. GHI is obtained by summing DNI projected
onto the horizontal surface and DHI.
A thorough investigation of the uncertainties of measuring GHI, DNI, and DHI with all three of
the RSR-type instruments was conducted in Payerne, Switzerland, for 15 months (Vuilleumier et al.
2017). Class AA or Class A thermopile instruments of the Payerne station that were used to compare
to the RSRs included two shaded-disk measurements of the DHI, two DNI measurements, and three
GHI measurements. Sampling for the thermopile instruments was 1 Hz with 1-minute averaging.
RSR sampling varied, but 1-minute averages of the samples were compared to the standards. All of
the RSR corrections suggested by the manufacturers were applied. Based on this study, the RSRs
measured GHI, DHI, and DNI adjusted values with expanded uncertainties of ±10, 13, and 5 to 6
percent, respectively. Additionally, there was some bias (up to about 3%) in one or two of the three
irradiance measurements for each type of RSR, but the irradiance component bias differed among
the RSR types. This  difference likely results from the calibration methodology of the different
manufacturers and the irradiance component used for determining the adjustment parameters.
222 Solar and Infrared Radiation Measurements

FIGURE 10.6 Comparison of corrected RSR DHI and Schenk black-and-white pyranometer DHI measure-
ment. Note the reduction in scatter with time averaging.

Of course, corrections removing systematic biases can be applied only to well-maintained instru-
ments with known calibration histories. If corrections are applied to instruments with unknown or
poor calibration records or to instruments that are soiled, corrections may actually make the values
deviate further from the reference values. Although Michalsky et al. (1988) and Wilbert et al. (2015)
find that silicon cell pyranometers are more resistant to soiling than other pyranometers, their rate
of soiling is entirely dependent on local conditions and time of year, thus good maintenance is a
necessity for the best results. To further address the soiling problem, continuing efforts are under-
way to develop automated cleaning for RSRs.
Shadowband Radiometers 223

10.4 MULTIFILTER ROTATING SHADOWBAND RADIOMETER


The  multifilter version of the RSR (MFRSR) was developed in the early 1990s (Harrison et  al.
1994). The instrument had a broadband silicon channel plus measurements with six filters that had
full widths at half maximum (FWHM) transmission of around 10 nanometers  (nm). Five of the
filters are between strong absorption bands and are used to calculate column aerosol optical depths
near 415, 500, 615, 673, and 870 nm. One is in the water vapor band centered near 940 nm that can
be used to calculate water vapor column. The MFRSR differs from the Irradiance, Inc. RSR and
similar instruments in the rotating shadowband makes three stops during its rotation: one stop just
short of shading the sensor, one stop that shades the sensor, and a final stop just after shading the
sensor. The extra stops do not quite block the direct Sun on the diffuser and this allows an estimate
of the diffuse irradiance that is shaded by the wide band when it stops to block DNI from reaching
the diffuser of the MFRSR.
The “raw” measurements from the MFRSR operating outdoors are the uncorrected GHI and
the blocked DHI, which is further corrected by using the average of the two side-band mea-
surements subtracted from the uncorrected global horizontal to approximately account for the
excess skylight that is blocked during the Sun-blocked measurement. This  modified diffuse
is subtracted from the uncorrected GHI to calculate the direct beam on a horizontal surface.
In equation form

modified_diffuse = global −
( sideband1 + sideband2 ) + blocked (10.2)
2

direct_horizontal = global_horizontal − modified_diffuse (10.3)

Even though the broadband channel of the MFRSR initially used a silicon photodiode detector, the
RSR adjustments for the LI-200 pyranometers discussed earlier could apply to the MFRSR silicon
sensor with parameters specifically determined for the MFRSR. The  angular response for each
MFRSR has always been characterized in the laboratory before they are deployed. These instru-
ment-specific responses are intended to remove most of the bias in the MFRSR angular response.
Therefore, angular adjustments applied to the LI-200 pyranometer should not  be applied to the
MFRSR broadband measurements. Specifically, before its deployment in the field, each MFRSR
head unit is mounted in a laboratory angular response bench so that the center of rotation of the
head is on the axis and perpendicular to a nearly collimated beam of a xenon arc lamp located about
5 meters from the MFRSR head (Michalsky et al. 1995). Measurements are made every 1 degree
from the horizon to the zenith in the north, south, east, and west directions. These measurements
are normalized to the zenith measurement producing an angularly interpolated cosine response for
each of the seven MFRSR channels.
Although the head temperature was controlled in the original silicon detector MFRSR, it was
only marginally held at the set temperature. Further, there was a seasonal temperature dependence
associated with the Spectralon® diffuser. Therefore, some temperature correction, different from that
of RSRs, had to be applied. Most importantly, the significant spectral dependence of a photodiode’s
response had to be corrected. Zhou et al. (1995) had some success in correcting this photodiode ver-
sion of the MFRSR to produce reasonable estimates of GHI, DHI, and DNI.
While the cosine correction is applied to the DrHI using interpolated measurements of the
cosine response for the position of the Sun in the sky, there is also a cosine correction that is applied
to the diffuse assuming an isotropic distribution of skylight. Although this is not a realistic distribu-
tion for most skies, calculations show that realistic spatial radiance distributions functions for clear
and cloudy skies give the same correction as isotropic skies to within 1 percent. The corrected GHI
224 Solar and Infrared Radiation Measurements

then is calculated by adding the cosine-corrected DrHI and cosine-corrected DHI. The  DNI is
obtained by dividing the cosine-corrected DrHI by the cosine of the solar zenith angle at the time
of measurement.
Michalsky (2001) further developed corrections for the spectral response characteristics associ-
ated with using the unfiltered silicon channel of the MFRSR for broadband solar measurements.
The results indicated almost zero bias for all components in the best methods and large root mean
square (RMS) differences associated with differences in sampling frequency (1 Hz versus 1/20 Hz)
under partly cloudy skies. Lastly, Michalsky et al. (2009) used the filtered measurements and the
unfiltered photodiode channel to find channel combinations that explained most of the variance in
a multivariate approach that consisted of regressing the photodiode’s measured solar components
against the same reference thermopile solar components. They found that it was unnecessary to use
the unfiltered photodiode channel to produce equivalent broadband estimates; therefore, it was pos-
sible to eliminate the unfiltered photodiode channel for another useful filter if needed. Surprisingly,
using only three carefully chosen filtered channels produced equivalent broadband estimates to
those produced by using all six filtered channels. This outcome can be plausibly explained because
two spectrally separated filters outside strong absorption bands of the shortwave spectrum capture
the broad spectral extinction caused by Rayleigh and aerosol extinction. Water vapor absorption
causes the largest molecular extinction in the shortwave spectrum; therefore, including 940  nm
filter measurements captures this source of variability. Since the filter measurements are over a very
narrow range of the spectrum (~10 nm), they do not require spectral corrections.
These calculations of GHI, DNI, and DHI depend on the accuracy of the narrow-band filter
measurements. The transmissions of the narrow-band filters degrade over time, and these affect the
measured irradiance values. These changes can be monitored by use of clear-day Langley plots.
According to Beer’s law,

I λ = I λ 0 ⋅ e − mτ λ (10.4)

where Iλ is the beam irradiance at a wavelength λ, Iλ0 is the extraterrestrial beam irradiance at
wavelength λ, m is air mass, and τλ is optical depth at wavelength λ. The air mass m can be cal-
culated accurately, and Iλ is measured with the MFRSR. By taking the logarithm of each side of
Equation (10.4), a linear equation in optical depth is obtained:

ln ( I λ ) = ln ( I λ 0 ) − mτ λ (10.5)

On clear days a Langley plot of ln(Iλ) versus m can be made. If τλ is constant, a linear plot results,
and extrapolating to zero air mass yields ln(Iλ0). Changes in ln(Iλ0) can be used to determine the
change in filter transmission. In practice, there is considerable scatter in ln(Iλ0) (see Section 7.5.1),
but a robust determination can be made with a sufficient number of Langley plots to follow the
changes in filter transmission and make adjustments as needed. Periodic lamp calibrations have
confirmed that this method tracks filter degradation well.
The  Yankee Environmental Systems MFRSR now  uses a thermopile detector for broadband
shortwave measurements and the filter chamber has been modified from its original design such
that the temperature in the chamber where the filters and detectors reside is held to a set tempera-
ture within a range of about 0.2°C. There remains a small temperature dependence associated with
changes in the Spectralon® diffuser transmission with ambient temperature.
The spectral correction is still a problem even though the thermopile responds to all solar wave-
lengths. The solar radiation that is transmitted to the detector is modified in a way that depends on
wavelength because of the scattering from the Spectralon® diffuser and within the integrating cav-
ity. Spectralon®’s wavelength-dependent, diffuse hemispheric transmission from 200 to 2600 nm
can be found at Spectralon®_UV-VIS-NIR_Diffusion. If GHI on a clear day is used to calibrate the
thermopile detector of the MFRSR, the diffuse measurement of clear skylight will be somewhat
Shadowband Radiometers 225

underestimated. If this underestimated diffuse is subtracted from the global to get direct irradiance
as in an RSR, the direct will be slightly overestimated. Attempts to derive spectral corrections and
to fix the remaining temperature dependence have not been made yet.

10.5 SPN1 SUNSHINE PYRANOMETER


A novel pyranometer that also estimates the GHI, DrHI, and DHI irradiance simultaneously is the
SPN1 Sunshine Pyranometer from Delta-T Devices Ltd. The  SPN1 has seven small thermopile-
based detectors under diffusers space strategically under a black mask that shades exactly half
of the sky (Figure 10.7). In essence, the SPN1 is making seven irradiance measurements. At least
one thermopile is shaded for any position of the Sun and at least one of these small thermopiles is
unshaded. All detectors receive equal amounts of diffuse solar irradiance from the rest of the sky.
Because the shade mask covers half of the sky, the DHI measurement is assumed to be twice the
reading from the lowest value of the thermopile readings. Similarly, GHI is determined by the high-
est reading thermopile plus the measurement of the lowest-reading thermopile. The  contribution
from the lowest reading thermopile compensates for the shading caused by the shield. The DrHI is
obtained by subtracting the lowest reading thermopile from the highest reading thermopile.
The  advantage of the SPN1 is that it can measure the DHI and GHI irradiance components
simultaneously and on a time frame similar to other thermopile-based pyranometers. While it is
important to ensure that the instrument is mounted horizontally, the specifications claim that the
instrument does not need to be oriented in any particular azimuthal direction. Not specifying the
orientation does have its drawbacks if the development or utilization of algorithms is desired to
remove systematic biases associated with this instrument. Orienting the signal cable to true north
would be a proper guide for this consistency.
The design of the SPN1 is based on the premise that the DHI is evenly distributed, or isotropic
across the sky. Even for totally cloudy skies DHI does not have a uniform distribution due to multiple
scattering with the zenith brighter than the horizon. Under clear or partially cloudy conditions the

FIGURE 10.7 Image of an SPN1 Sunshine Pyranometer showing the mask that shades some of the thermo-
piles from direct sunlight.
226 Solar and Infrared Radiation Measurements

distribution is definitely not uniform due to circumsolar (forward scattering near the Sun) and horizon
brightening. Models often divide the DHI into at least three separate categories or into three distinct
areas of the sky: the circumsolar component, the horizon brightening component, and the rest of
the sky. Under clear-sky conditions, the circumsolar component behaves much like the DNI compo-
nent during clear skies when it is projected onto a horizontal surface. Therefore, these three compo-
nents yield diverse contributions to the DHI and the assumption that diffuse irradiance is isotropically
distributed across the sky dome is not an accurate assumption. This skews the DHI and DrHI values.
It  is difficult to obtain a uniform and consistent calibration for all seven thermopiles. In  some
instances, a jump can be seen where the maximum measurement shifts from one thermopile to another.
While these jumps are small, they do introduce additional scatter and uncertainty into the data.
As with RSRs discussed in Chapter 10.1, three SPN1 pyranometers were compared against high
quality reference measurements during an experiment at the Baseline Surface Radiation Network
(BSRN) station in Payerne, Switzerland.1 A Kipp & Zonen CHP1 pyrheliometer that was checked
against an absolute cavity radiometer (ACR) during the experiment was the reference DNI instru-
ment. A Kipp & Zonen CMP22 was the reference GHI instrument, and a CMP22 under a shad-
ing ball provided the reference DHI data. Under all weather conditions, the expanded uncertainty
for the GHI measurements for the SPN1 was ±25 Wm−2 (±10%) when compared to the reference
measurements (Vuilleumier et al. 2017). This result is similar to the RSR GHI measurement uncer-
tainties that were conducted at the same time. The SPN1s DHI values exhibited a large range of
expanded uncertainty from −45 Wm−2 to 20 Wm−2 (−35% to 13%) with a median bias of −10 Wm−2.
For the SPN1 DNI values, the uncertainty range ran from −45 Wm−2 to 125 W/m−2 (−6% to 19%).
On a yearly basis, the GHI values exhibit little bias while the DHI values from the SPN1s have a bias
on the order of −8 percent and approximately 10 percent for DNI values.
Without some method for removing the DHI bias, the uncertainty in the DNI values will remain
large. Any attempt to develop a model to adjust for the DHI bias would require knowledge of the
orientation of the SPN1 because the DHI components have distinct distributions in the sky dome
that relate to the position of the Sun and the extent of cloudiness. Therefore, a uniform orientation
for SPN1s should be established. Work has been done to develop algorithms to adjust the SPN1 data
to remove the systematic effects (Badosa et al. 2014). These algorithms require that the SPN1 be
oriented in a specific manner so that more sophisticated models of diffuse irradiance can be used
in the adjustment algorithm.
Besides an algorithm designed to remove the systematic diffuse biases from the SPN1, other
biases such as the deviation from a Lambertian cosine response, thermal offset, and variance char-
acteristics among the seven thermopiles should be evaluated. With an adequate model to remove
systematic biases, the role of SPN1 sunshine pyranometers could play an expanded role in measuring
the solar components.

QUESTIONS
1. What was the major disadvantage of the Wesely (1982) simple RSR compared with modern
instruments?
2. Explain the difference in the measurement algorithms used by Yankee Environmental
Systems, Inc. and Irradiance, Inc. for their distinct versions of the RSR.
3. What is the major reason for the popularity of RSRs? What is their major shortcoming?
4. Why are traditional thermopile pyranometers not used in RSRs?
5. What are two of the four corrections that have to be performed for RSRs?
6. Given the low and high readings from the thermopiles of an SPN1 Sunshine Recorder, how
is GHI, DrHI, and DHI calculated?
7. How is DNI obtained from an SPN1 Sunshine Recorder?

1 See also, Habte et al. (2014) for similar comparisons conducted in Golden, Colorado.
Shadowband Radiometers 227

REFERENCES
Augustyn, J., T. Geer, T. Stoffel, R. Kessler, E. Kern, R. Little, and F. Vignola. 2002. Improving the accuracy
of low cost measurement of direct normal irradiance. In Proceedings of the American Solar Energy
Society, ed. R. Campbell-Howe and B. Wilkins-Crowder. Boulder, CO: American Solar Energy Society.
Augustyn, J., T. Geer, T. Stoffel, R. Kessler, E. Kern, R. Little, F. Vignola, and B. Boyson. 2004. Update
of algorithm to correct direct normal Irradiance measurements made with a rotating shadow band
pyranometer. In  Proceedings of the American Solar Energy Society, ed. R. Campbell-Howe and
B. Wilkins-Crowder. Boulder, CO: American Solar Energy Society.
Badosa, J., J. Wood, P. Blanc, C. N. Long, L. Vuilleumier, D. Demengel, and M. Haeffelin. 2014. Solar irra-
diances measured using SPN1 radiometers: Uncertainties and clues for development. Atmospheric
Measurement Techniques 7:4267–4283.
Habte, A., S. Wilcox, and T. Stoffel. 2014. Evaluation of radiometers deployed at the National Renewable
Energy Laboratory’s solar radiation research laboratory. NREL/TP-5D00-60896. https://www.nrel.gov/
docs/fy14osti/60896.pdf.
Harrison, L., J. Michalsky, and J. Berndt. 1994. Automated multifilter rotating shadow-band radiometer: An
instrument for optical depth and radiation measurements. Applied Optics 33:5118–5125.
King, M. L., W. E. Boyson, B. R. Hansen, and W. I. Bower. 1998. Improved accuracy for low-cost solar irra-
diance sensors. Paper Presented at the 2nd World Conference and Exhibition of Photovoltaic Solar
Energy Conversion, Vienna, Austria, July 6–10.
Michalsky, J. J. 2001. Accuracy of broadband shortwave irradiance measurements using the open silicon
channel of the MFRSR. Paper Presented at the 11th ARM Science Team Meeting, Atlanta, GA. http://
www.arm.gov/publications/proceedings/conf11.
Michalsky, J. J., J. A. Augustine, and P. W. Kiedron. 2009. Improved broadband solar irradiance from the
multi-filter rotating shadowband radiometer. Solar Energy 83:2144–2156.
Michalsky, J. J., J. L. Berndt, and G. J. Schuster. 1986. A  microprocessor-controlled rotating shadowband
radiometer. Solar Energy 36:465–470.
Michalsky, J. J., L. Harrison, and B. A. LeBaron. 1987. Empirical radiometric correction of a silicon photodi-
ode rotating shadowband pyranometer. Solar Energy 39:87–96.
Michalsky, J. J., L. C. Harrison, and W. E. Berkheiser III. 1995. Cosine response characteristics of some radio-
metric and photometric sensors. Solar Energy 54:397–402.
Michalsky, J. J., R. Perez, L. Harrison, and B. A. LeBaron. 1991. Spectral and temperature correction of sili-
con photovoltaic solar-radiation detectors. Solar Energy 47:299–305.
Michalsky, J. J., R. Perez, R. Stewart, B. A. LeBaron, and L. Harrison. 1988. Design and development of a
rotating shadowband radiometer solar radiation/daylight network. Solar Energy 41:577–581.
Vignola, F. 1999. Solar cell based pyranometers: Evaluation of diffuse responsivity. In Proceedings of the
American Solar Energy Society, ed. R. Campbell-Howe. Boulder, CO: American Solar Energy Society.
Vignola, F. 2006. Removing systematic errors from rotating shadowband pyranometer data. In Proceedings
of the American Solar Energy Society, ed. R. Campbell-Howe. Boulder, CO: American Solar Energy
Society.
Vignola, F., Z. Derocher, J. Peterson, L. Vuilleumier, C. Felix, J. Gröbner, and N. Kouremeti. 2016. Effects of
changing spectral radiation distribution on the performance of photodiode pyranometers. Solar Energy
129:224–230.
Vignola, F., J. Peterson, S. Wilbert, P. Blanc, N. Geuder, and C. Kern. 2017. New methodology for adjust-
ing rotating shadowband irradiometer measurements. AIP Conference Proceedings 1850(1).
doi:10.1063/1.498452.
Vuilleumier, L., C. Felix, F. Vignola, P. Blanc, J. Badosa, A. Kazantzidis, and B. Calpini. 2017. Performance
evaluation of radiation sensors for the solar energy sector. Meteorlogische Zeitschrift 26:485–505.
Wesely, M. L. 1982. Simplified techniques to study components of solar radiation under haze and clouds.
Journal of Applied Meteorology 21:373–383.
Wilbert, S., N. Geuder, M. Schwandt, B. Kraas, W. Jessen, R. Meyer, and B. Nouri. 2015. Best practices
for solar irradiance measurements with rotating shadowband irradiometers. Report of IEA-Task46-B1.
IEA_Task46B1_BestPractices.pdf.
Zhou, C., J. Michalsky, and L. Harrison. 1995. Comparison of irradiance measurements made with the multi-
filter rotating shadowband radiometer and first-class thermopile radiometers. Solar Energy 55:487–491.
11 Infrared Measurements

The World Infrared Standard Group has been operating continuously since 2003, showing an
excellent relative stability between individual instruments of ±1 W m−2.
Julian Gröbner (2009)

11.1 INTRODUCTION
A body with a temperature above zero Kelvin (K) emits radiation. For the Earth’s surface and atmo-
spheric temperatures, this radiation is in the infrared (IR) portion of the electromagnetic spectrum.
The  Stefan–Boltzmann law states that the power per unit area L emitted by a body of absolute
temperature T in K is

L = εσ T 4 . (11.1)

In this equation, σ is the Stefan–Boltzmann constant equal to 5.670 × 10−8 W m−2 K−4, and ε is the
emissivity of the body. If the emitting body is a gray body, then the emissivity is <1; for a black-
body, ε is equal to 1. For example, if a pyrgeometer, which measures IR radiation (see Section 11.2)
is used to look at the surface of a parking lot, the temperature of the tarmac can be calculated very
accurately, assuming an emissivity of 1. When looking at a clear sky, the temperature of the sky is
more difficult to measure because there is no longer a near-perfect blackbody or gray body with a
known emissivity. Figure 11.1 illustrates blackbody spectral irradiance in units of power per unit
area per wavelength at two temperatures, 257 and 299 K (representative of subarctic and tropical
skies). The two clear-sky spectral distribution curves in this figure do not follow any of the black-
body curves at all wavelengths but they do follow the 257 and 299 K blackbody distributions at
some of the wavelengths. So, what is going on in the sky? Parts of the IR spectrum are completely
absorbed by the molecules in the Earth’s atmosphere; water vapor and carbon dioxide are especially
significant absorbers. Good absorbers are also good radiators and the pyrgeometer sees the IR radi-
ated by these atmospheric constituents. There are regions in the IR spectrum where there is little
molecular absorption or there is only weak broadband extinction caused by aerosols. The infrared
spectrum will have “windows” that allow outgoing IR radiation to escape. These windows are also
regions where there is little IR radiation from the sky. The “299 K Sky” plot in Figure 11.1 is for a
modeled tropical clear sky. The “257 K Sky” plot in the same figure is for a modeled subarctic clear
sky. The tropical and subarctic clear skies both exhibit significantly reduced irradiance centered
near 10 micrometers (µm). This decreased irradiance, or window if thinking of IR radiation from
the Earth passing through the atmosphere, is caused by the reduction in the number of molecules
that emit, or absorb, radiation at these wavelengths, mainly, water vapor. Note that the subarctic
clear sky is very transparent around 10 µm and the 20 µm region is semitransparent, whereas the
20 µm region is blackbody-like for the tropical sky. If the sky is covered with thick clouds, the spec-
trum fills in (the windows are no longer open) and the irradiance from the sky takes on the distribu-
tion of a blackbody with a temperature that is the effective temperature of the bottom of the clouds
as viewed from the surface of the Earth.
Ground-based, spectrally resolved IR measurements are not  discussed in this book. The  best
measurements of spectral IR are made using Fourier transform spectrometers. These measurements
are radiance measurements as opposed to the irradiance measurements that are the focus of this
book. The University of Wisconsin’s Space Science and Engineering Center has led the way in devel-
oping ground-based spectral IR measurements (Knuteson et al. 2004a, 2004b). Feltz et al. (2003)

229
230 Solar and Infrared Radiation Measurements

30
299 K
299 K Sky
257 K
257 K Sky
25
irradiance (W/m2 per micrometer)

20
15
10
5
0

5 10 15 20 25 30

wavelength(micrometers)

FIGURE 11.1 Blackbody irradiances for true blackbodies at the temperatures in the legend plus clear-sky
irradiances for a tropical sky and a subarctic sky. Note the 10 µm window for skies and the semitransparent
window for the subarctic sky in the 20 µm region.

described continuous retrievals of temperature and moisture that are possible using these measure-
ments; however, a wealth of information on retrieving other atmospheric trace species such as car-
bon dioxide and ozone can be found in the literature.
In this chapter, the discussion begins with broadband instruments for measuring infrared radia-
tion, including how they function and three similar but slightly different equations that have been
used to produce measured infrared irradiance from the same inputs. Next, laboratory calibrations
using a blackbody calibrator are discussed along with two different methods for using the same
blackbody to obtain calibrations and the more recent development of an interim standard for IR
measurements. In  Section  11.4, recent developments in calibrating pyrgeometers are discussed.
These methods use laboratory blackbody and outdoor measurements against a standard group
of well-calibrated pyrgeometers for the derivation of the coefficients of the governing equations.
Section  11.5 outlines information about other pyrgeometer manufacturers not  presented in this
chapter. The  final section contains some operational considerations for using a pyrgeometer and
estimates uncertainties that one may obtain in routine operations.

11.2 PYRGEOMETERS
Downwelling and upwelling IR radiation measurements in the broadband are made with instru-
ments called pyrgeometers (see, e.g., Figure  3.17 in Chapter  3). Broadband instruments use
a thermopile detector covered by a protective silicon dome that has been coated with a solar
blind filter (Figure  11.2). The  shortwave cutoff of the transmission of the filter is generally in
the 3 to 5 µm range with a longwave cutoff at 30 to 50 µm. Sample transmission curves for the
Eppley Laboratory, Inc. model precision infrared radiometer (PIR), the Kipp  & Zonen model
CG 4, and Hukseflux model IR20 pyrgeometer domes are given in Figure 11.3. Their response
Infrared Measurements 231

FIGURE  11.2 Schematic of a pyrgeometer thermopile detector under a hemispherical interference filter
designed to block solar (short wave) radiation. Typically, the dome and pyrgeometer body temperatures
are measured (TD and TB). The  thermopile signal depends on the difference in temperatures between the
exposed surface TE and the pyrgeometer body temperature TB. Generally, the body temperature is warmer
than the exposed surface because the exposed surface cools with exposure to the dome cooled by exposure to
the generally cooler sky. The pyrgeometer equation is derived by equating the three incoming contributions
to the exposed thermopile surface (1–3) to the one outgoing term from this surface (4). (See Equation 11.2.)

Transmission of IR Domes
60
50
transmission (%)

40
30
20

cg4
pir
ir20
10
0

0 5 10 15 20 25 30

wavelength(micrometers)

FIGURE 11.3 Typical pyrgeometer transmission curves for three different manufacturers’ domes. The cut-on
wavelengths vary somewhat and depend on temperature to some extent.
232 Solar and Infrared Radiation Measurements

is not spectrally flat, and their short wavelength cutoffs are similar but not identical. What this
implies with regard to IR measurement under changing IR spectral conditions will be discussed
in Section 11.4.
In  this discussion of pyrgeometer measurements, it is assumed that the dome temperature
and the thermopile junction body temperature of the Eppley PIR are measured and that the less
accurate practice of using the internal battery-powered circuit to compensate for the radiation loss
from the exposed junction of the thermopile is not used. The equation for determining the incident
IR irradiance on a pyrgeometer is derived by adding the three most significant contributing terms
to the received flux on the exposed junction of the thermopile (see Figure 11.2) and subtracting
the term that represents the outgoing flux from the exposed junction of the thermopile; this sum-
mation of four terms is equal to the net flux at the exposed junction of the thermopile. The three
incoming terms are (1) the incoming IR that is transmitted through the dome and is absorbed at
the exposed junction of the thermopile, (2) the IR that is emitted from the dome and is absorbed
by the exposed junction of the thermopile if they are not at the same temperature, and (3) the IR
emitted by the exposed junction of the thermopile and reflected back to the exposed junction of
the thermopile after reflection from the dome. The outgoing flux (4) is the IR radiation emitted
from the exposed junction of the thermopile that passes through the dome. The exposed junction
of the thermopile is the emitting surface; although, there is no temperature measurement at the
exposed junction. However, there is a temperature measurement at the unexposed junction of
the thermopile, referred to here as the body (B) temperature. The  temperature of the exposed
junction is calculated using TE = T B + αV, where α is a product of the Seebeck coefficient for the
thermopile type, the number of junctions, and the thermopile efficiency factor. The voltage, V,
is produced by the temperature difference between the unexposed and exposed junctions of the
thermopile.
Derivations of the pyrgeometer equation to calculate the incident IR flux are usually expressed
in terms of the measurable dome and body temperatures; therefore, only the highest-order term that
includes the product of αV and TB3 is kept since αV is very small. For comparison, a large tempera-
ture difference between the exposed (hot) and hidden (cold) junctions of a thermopile would be on
the order of 0.5 K, while the dome and case temperatures typically are in the range 260 to 310 K for
mid-latitudes.
Albrecht et  al. (1974) derived the basic equation  for how an Eppley PIR is used to measure
incoming infrared radiation. Their Equation (11.1) as given in Albrecht and Cox (1977) using modi-
fied symbols is

( ) (
L = U emf c1 + c2TB3 + 0σ TB4 − kο TD4 − TB4 ) (11.2)

This equation has an analytic form similar to the Philipona et al. (1995) Equation (11.11)

U emf
L=
C
( ) (
1 + k1σ TB3 + k2σ TB4 − k3ο TD4 − TB4 ) (11.3)

In  these equations  L is the measured longwave irradiance, T D is the dome temperature, T B is
the unexposed thermopile junction (body) temperature, and Uemf is the measured voltage across
the thermopile. Albrecht and Cox (1977) considered c1 ≫ c2T B3 in Equation 11.2 and ignored the
T B3 dependence. Therefore, the most common form of the Albrecht and Cox (1977) formula is
written

U emf
L=
C
(
+ σ TB4 − kσ TD4 − TB4 ) (11.4)
Infrared Measurements 233

where εo in Equation 11.2 is assumed to be unity, and C is the responsivity factor, a constant. Reda
et al. (2002) derived a form of the pyrgeometer equation that used the estimated thermopile exposed
junction temperature TE as given above rather than TB

( )
L = k0 + k1U emf + k2σ TE4 − k3σ TD4 − TE4 . (11.5)

The addition of an offset k0 is unique among pyrgeometer equations and was added by Reda et al.
(2002) to account for the offset of the pyrgeometer.

11.3 CALIBRATION
Historically, pyrgeometer calibration methods have been based on temperature-controlled black-
body chambers designed to provide a uniform field of infrared radiation computed from the shape,
emissivity, and effective temperature of the blackbody source. In this section, we provide informa-
tion about blackbody calibration methods and introduce the recent efforts to provide an interna-
tional reference for pyrgeometer calibrations. Additional calibration information is available from
Chapter 13.
The Eppley Laboratory model PIR is calibrated at Eppley using a circulating water bath black-
body that can be held at a constant, uniform temperature (Eppley Laboratory Inc. 1995). The PIR is
raised into the Parsons black-painted recessed hemisphere of this blackbody. After about 1 minute
the instrument reaches equilibrium at 5°C and later at 15°C. From Equation 11.4, the pyrgeometer
thermopile responsivity (C) can be obtained:

U emf U emf
L = σ TBB
4
= + σ TB4 ⇒ C = (11.6)
C (
σ TBB
4
− TB4 )
where the thermopile and dome temperatures are assumed to be equal or negligibly different once
equilibrium is obtained, and TBB is the temperature of the laboratory blackbody. The values of C at 5
and 15°C are averaged for the final determination of the calibration. The value of k in Equation 11.4
is not provided by Eppley but is assumed to be 4 for KRS-5 domes found in early production units;
however, a typical value of k equal to 3.5 is suggested for the silicon domes now used for the Eppley
PIR (Thomas Kirk, personal communication).
Laboratory calibration blackbodies, not  unlike the Eppley blackbody, have been used in a
procedure that determines both the thermopile responsivity C and the dome correction factor k.
One technique that yields both C and k in Equation 11.4 is described in Albrecht and Cox (1977).
Their blackbody calibrator is made from a copper cylindrical block that has a black recessed cone.
The calibrator is chilled to a low temperature overnight, that is, lower than the environment where
it will be used, and then allowed to warm to room temperature throughout the several-hour cali-
bration process. Before placing the pyrgeometer upside down in the cone, the dome of each pyr-
geometer to be calibrated is first heated so that it is a few degrees above the pyrgeometer body
temperature. Measurements of the blackbody calibrator temperature are made in two places to
establish the blackbody’s temperature uniformity; typical agreement is within 0.2°C. Thermopile
body and dome temperatures and the thermopile voltage output are measured simultaneously with
the calibrator temperature. The dome will cool to have a lower temperature than the pyrgeometer
body passing through a point where the dome and body temperatures are the same. This procedure
is repeated several times during calibration as the blackbody warms to room temperature over a few
hours. The thermopile sensitivity C can be determined from Equation 11.6 at the body and dome
temperature-crossing points by averaging these, or it can be calculated from the slope of a plot of
234 Solar and Infrared Radiation Measurements

the thermopile voltage Uemf versus σ TBB


4
( )
− TB4 at the crossing points. With C determined, k can be
calculated from an average of multiple determinations of k using

 U 

(
k = σ TB4 − TBB
4
)
+ emf
C
(
 / σ TD − TB
4 4
) (11.7)

or from the slope of

U emf
(
σ TBB
4
− TB4 − ) C
(
versus σ TD4 − TB4 )
Either method permits the calculation of the uncertainty of C and k from the standard deviation of
the multiple determinations of k or the standard deviation in the slope of the fit.
At the National Oceanic and Atmospheric Administration (NOAA) this same blackbody is oper-
ated in much the same way to obtain the basic data; however, C and k are estimated by using all of
the simultaneous measurement points. Equation 11.4 is used in a multivariate linear regression to
solve for the two parameters. Dutton (1993) demonstrated that the thermopile coefficient obtained
in this way is within 2 percent of the Eppley thermopile coefficient provided by the manufacturer.
Further, in a round-robin comparison of five Eppley PIRs sent to 11 laboratories, Philipona et al.
(1998) identified six laboratory calibrations of the thermopile coefficient, including those of the
NOAA and the Eppley Laboratory, that agreed to within 2 percent of the median value; the other
five laboratory calibrations showing inconsistent behavior with mean differences all greater than
5 percent and many more than 10 percent different from the median.
Thus far the discussion has been about the Eppley PIR. There is a different pyrgeometer designed
by Kipp & Zonen, the CG4 (updated in 2014 to models CGR4 and SGR41), for which the calibration
strategy is different. A photo of the Kipp & Zonen CG4 is shown in Figure 11.4. The CG4 could
be calibrated using a blackbody, but the manufacturer asserts that its outside calibration using sky
irradiance is superior. The  CG4 to be calibrated is compared to a reference CG4 that has been

FIGURE 11.4 The Kipp & Zonen CG 4 is constructed to minimize the difference in temperature between


the dome and pyrgeometer body. This difference has the effect of removing the last term of Equations 11.3
through 11.5 and making the measurement problem simpler.

1 Updated design includes low window heating offset and optimal spectral cut-on wavelength allowing for accurate day-
time measurements without the need for dome shading. The SGR4 has an RS-485 Modbus® digital interface.
Infrared Measurements 235

calibrated in Davos, Switzerland, at the World Radiation Center (WRC) against the interim World
Infrared Standard Group (WISG), which is discussed in the following section. The manufacturer
contends that the thermal contact between the dome and the body of the CG4 is such that there is
never a significant difference in temperatures between dome and the body of the CG4. This means
that the thermopile sensitivity can be easily calculated using Equation 11.4 with the last term set to
zero and L set to the irradiance measured by the reference CG4.

11.4 IMPROVED CALIBRATIONS


Since 2004, at the recommendation of the Commission for Instruments and Methods of Observation
(CIMO), the WRC in Davos, Switzerland has led the research to improve broadband IR radiation
calibrations of pyrgeometers and establish an interim World Meteorological Organization (WMO)
pyrgeometer reference (Gröbner et  al. 2014). The  initial foray into this line of research was led
by Rolf Philipona who developed an infrared absolute sky-scanning radiometer (ASR) (Philipona
2001). His sky scanner was calibrated by pointing a 6 degree field of view radiometer into a
well-characterized blackbody, then scanning the sky at 32 points (four elevation and eight azimuth
positions) with a dwell time at each position of 30 seconds, and then pointing into the blackbody at
the end of the sky scan. The 32-point scan takes about 24 minutes to complete; therefore, a stable
sky is absolutely necessary. The instrument is unfiltered (i.e., designed as an absolute instrument
with no optical surface affecting incoming radiation) so it can only be used at night. The integral
of the interpolated radiance measurements over the sky provides an absolute irradiance calibra-
tion for pyrgeometers. Two International Pyrgeometer and Absolute Sky-scanning Radiometer
Comparisons (IPASRCs) were held. IPASRC-I (Philipona et al. 2001) was at mid-latitude in the early
autumn at the U.S. Department of Energy Atmospheric Radiation Measurement (ARM) Program’s
Radiometer Calibration Facility at the Southern Great Plans (SGP) site in northern Oklahoma.
Measurements were obtained over a range of 260 to 400 Wm−2 downwelling IR for clear to cloudy
skies and irradiances typical of late summer conditions. IPASRC-II was conducted at the ARM
North Slope of Alaska site near Barrow, Alaska, during late winter (Marty et al. 2003). The long-
wave irradiances varied between 120 and 240 watts per square meter (Wm−2) with frequent clear
skies. The two comparisons included 15 and 14 pyrgeometers, in IPASRC-I and -II with many, but
not all pyrgeometers, the same in both campaigns.
In both comparisons better agreement among the pyrgeometers for outdoor measurements was real-
ized using the calibrations based on the sky-scanning radiometer than using calibrations performed
on all of the instruments using a good blackbody calibration system. In turn, the uniform blackbody
calibration on the instruments provided better agreement among the pyrgeometers than using calibra-
tions that came with the instruments either from the manufacturer or their own laboratories. These
measurements also were compared to IR models of the atmosphere including MODTRAN (Berk
et al. 2000) and the line-by-line radiative transfer model (LBLRTM) (Clough et al. 1992). Both mod-
els were slightly low with respect to pyrgeometer and scanning radiometer measurements; however,
neither model included the effects of aerosols because no aerosol measurements in the infrared were
made that could be used as inputs to these models. Any aerosol effect would have likely improved the
agreement, but it is not clear by how much. In summary, blackbody-calibrated, sky-scanner-calibrated
measurements, and models of downwelling infrared irradiance agreed to within 1–2 Wm−2 at night.
The  improvement in the agreement when the pyrgeometer calibrations were obtained using
the sky scanner versus the blackbody can be plausibly explained when the spectral response of
the pyrgeometer and the spectral distribution of the downwelling IR are considered. Figure 11.1
demonstrates that spectral IR sky irradiance is not that of a blackbody. Figure 11.3 illustrates the
non-uniform spectral response of the filters that transmit infrared. The product of the filter trans-
mission and blackbody emission would be expected to be different from the product of the filter
transmission and the sky emission leading to the differences seen in the IPASRC results when
blackbody calibrations were used versus sky-scanner calibrations.
236 Solar and Infrared Radiation Measurements

The previous discussion of the absolute sky scanner and improvement in IR measurements made
with pyrgeometers using sky-scanner calibrations suggested a methodology to improve the stan-
dard for infrared measurements. The interim WISG for the pyrgeometer IR standard (https://www.
pmodwrc.ch/en/world-radiation-center-2/irs/) consists of four pyrgeometers: an Eppley PIR and a
Kipp & Zonen CG 4 that participated in both IPASRCs plus another of each type of these pyrgeom-
eters that were calibrated using these two IPASRC participant pyrgeometers.
Pyrgeometers are calibrated at the WRC in Davos by obtaining the coefficients in Equations 11.3
or 11.4 in the WRC blackbody except for the thermopile coefficient; the latter is obtained by com-
parison with the averaged WISG irradiance when nighttime sky conditions are stable. To transfer
the calibrations to other pyrgeometers, regressions may be performed using either Equation 11.3
or 11.4 using the prygeometers that have been calibrated at WRC. Alternatively, Reda et al. (2006)
demonstrated a pyrgeometer calibration method where the nighttime atmosphere is used as the IR
reference.
Julian Gröbner, who is currently in charge of the WISG at the WRC, has developed a pyrge-
ometer with a very different design (Gröbner 2012) called the Infrared Integrating Sphere (IRIS).
Independently, Ibrahim Reda, at the National Renewable Energy Laboratory (NREL), in collabo-
ration with colleagues at NREL, the National Institute of Standards and Technology (NIST), and
Labsphere, has developed a new Absolute Cavity Pyrgeometer (ACP) to measure IR irradiance
with traceability to the international system of units (Reda et al. 2012). In an initial comparison of
these two instruments and the WISG, Gröbner et al. (2014) found that they agreed with each other
to within ±1 Wm−2. However, they measured higher IR irradiance by 2 to 6 Wm−2 than the WISG.
The difference was found to be dependent on the water vapor column above the instruments. Based
on this study, they suggested that WISG should be corrected; however, Philipona (2015) argued
that it was premature to make this adjustment. A response to the Philipona note from Gröbner et al.
(2015) suggested that more formal comparisons should be performed before a new IR standard
could be established.

11.5 OTHER PYRGEOMETER MANUFACTURERS


There are other pyrgeometer manufacturers, but the previous discussion covers the two basic designs
for a commercially available pyrgeometer. Other commercial pyrgeometers known to the authors are
the Kipp & Zonen CG3, which differs from the CG4 and CGR42 in that the field of view of the CG3
is 150 degrees as opposed to 180 degrees (www.kippzonen.com). Hukseflux makes the IR20 and
IR20WS with a 180 degree field of view and the IR02 pyrgeometer with a 150 degree field of view
(www.hukseflux.com). EKO Instruments Co., Ltd. makes the MS-202 pyrgeometer, which appears
to function much like the Eppley PIR including circuitry to compensate for radiation loss from the
exposed junction of the thermopile (the σTB4 in Equation 11.4). Thermistors or, alternatively, plati-
num resistance thermometers, to measure the body and dome temperatures are available as options
(eko-eu.com). Apogee Instruments offers their model SL-510 pyrgeometer with a 150 degree field
of view for downwelling measurements and model SK-610 for upwelling irradiance measurements
with similar performance specifications (www.apogeeinstruments.com). Middleton Solar offers their
model PG01 Pyrgeometer with a 170 degree field of view (www.middletonsolar.com).

11.6 OPERATIONAL CONSIDERATIONS


Figure 11.5 is a photograph of three Eppley PIRs in ventilated housings. The three pyrgeometers are
mounted on an automatic solar tracker to keep the instruments shaded from the direct solar beam
at all times of the day. This is the preferred way to operate a pyrgeometer, ventilated and shaded
(McArthur 2004). The  ventilation, with optional heating, helps prevent dew and frost formation

2 Kipp & Zonen also manufactures SGR3 and SGR4 pyrgeometers providing an RS-485 Modbus® digital interface.
Infrared Measurements 237

FIGURE  11.5 Three Eppley PIR pyrgeometers operated in the optimum manner. The  instruments are
continuously shaded from direct sunlight using an automatic tracker and are placed in ventilators to minimize
dew formation and dust deposition on the instrument domes.

and keeps the dome cleaner than it would be without ventilation. Shading with a tracking ball or
disk minimizes the effects of possible shortwave leaks associated with the interference filter cut-on
wavelength and with pinholes in the interference filter, plus the heating of the dome by sunlight is
minimized. Operating without a shaded pyrgeometer will, as expected, result in higher uncertainty
for the daytime measurements.
Two (for CG4-type instruments) or three (for PIR-type instruments) voltages need to be mea-
sured from a pyrgeometer, and one (CG4) or two (PIR) voltages need to be supplied to excite the
thermistor circuits used for temperature measurements.
The thermopile voltage is produced by a temperature difference between the exposed and unex-
posed junctions of the thermopile. This cumbersome language (exposed and unexposed) is substi-
tuted for the usual terms hot and cold junctions because of the reversal of the roles these junctions
play in the pyrgeometer relative to their usual use in pyranometer measurements. The unexposed
junction is at the body temperature of the pyrgeometer, and the exposed junction is generally at a
lower temperature because it is cooled by exposure to the dome that radiates to the colder sky, as
explained earlier in this chapter. This  arrangement results in a negative voltage output from the
thermopile most of the time. In fact, a good test to verify that the thermopile is connected correctly
is to briefly cover the dome with your warm hand and watch the signal become positive.
The  thermistor circuit typically used (rarely are platinum resistance thermometers used for
these measurements) to measure the temperature of the body and dome may differ in the thermis-
tors used, but most use a configuration similar to that in Figure 11.6. The thermistor’s resistance
has a reproducible behavior with temperature. Slight imperfections in thermistor resistance ver-
sus temperature are included in the calibration coefficients (k’s) in Equations  11.3 through 11.5.
The equations and algebra implied by Figure 11.6 permit the temperature of the dome and body
to be determined. For  the CG4 there is only one thermistor for the body temperature, since the
manufacturer asserts that the dome and case are nearly the same temperature and does not normally
supply a dome thermistor. In support of this claim, Marty et al. (2003) suggested that the dome and
body temperature difference term is 20 times smaller for the CG4 than it is for an Eppley PIR. In an
238 Solar and Infrared Radiation Measurements

FIGURE 11.6 One method to derive the dome or pyrgeometer body temperatures using a thermistor circuit.

unpublished study, Ellsworth Dutton (personal communication) found that more than 95 percent of
these dome–body corrections for the PIR are smaller than 30 Wm−2; this suggests that the CG4 may
have a bias, but it is less than 1.5 Wm−2.
Given that

V = i ⋅ Rfixed + i ⋅ Rthermistor

where i is the current through the low temperature coefficient fixed resistor Rfixed and the tempera-
ture dependent variable resistor Rthermistor. It can be shown with a little algebra that:

 V − VR 
Rthermistor = Rfixed  .
 VR 

Therefore, the resistance of the thermistor placed in contact with the body of the pyrgeometer or the
dome can be calculated by measuring the voltage drop across the fixed resistor. Consequently, this
resistance can be related to the temperature using the Steinhart-Hart equation:

T = C1 + C2 ⋅ ln ( Rthermistor ) + C3 ⋅ ln ( Rthermistor ) ,
3

where C1, C2, and C3 are the Steinhart-Hart coefficients for the thermistor type used in the
pyrgeometer.
Philipona et al. (2001) and Reda et al. (2012) provide guidance on calculating the uncertainty
of pyrgeometer measurements. The uncertainty depends on the provider of the calibration, the for-
mula used to reduce the data to irradiances, and whether a calibration is from a blackbody or from
a comparison to the WISG IR standard. Furthermore, daytime uncertainties are generally twice
the nighttime uncertainties. Philipona et  al. (2001) show that the nighttime 95 percent level of
confidence uncertainties range from 0.8 to 4.6 Wm−2 and daytime uncertainties range from 1.6 to
8.2 Wm−2. The uncertainty of the WISG needs to be included in the total estimated uncertainty,
which is estimated to be approximately ±1.7  Wm−2 (Julian Gröbner, personal communication).
Adding these uncertainties as the square root of the sums of the square implies nighttime uncer-
tainties between 1.9 and 4.9 Wm−2 and daytime uncertainties between 2.3 and 8.4 Wm−2 depending
on the calibration procedure followed.
Infrared Measurements 239

QUESTIONS
1. What distinguishes a blackbody from a gray body?
2. At about what central wavelength is the atmospheric window that keeps a clear atmosphere
from resembling a perfect blackbody?
3. What is the shortest wavelength that typical pyrgeometer domes transmit?
4. What is the current working calibration standard for broadband infrared radiation?
5. Why is the thermopile signal from a pyrgeometer only rarely positive?
6. Why are two temperature measurements required for the Eppley PIR and only one for the
Kipp & Zonen CG 4?
7. A human body should emit like a blackbody at what temperature?
8. Why do the best daytime measurements of IR use a Sun-shaded instrument?

REFERENCES
Albrecht, B., and S. K. Cox. 1977. Procedures for improving pyrgeometer performance. Journal of Applied
Meteorology 16:188–197.
Albrecht, B., M. Poellot, and S. K. Cox. 1974. Pyrgeometer measurements from aircraft. Review of Scientific
Instruments 45:33–38.
Berk, A., G. P. Anderson, P. K. Acharya, J. H. Chetwynd, L. S. Bernstein, E. P. Shettle, M. W. Matthew,
and S. M. Adler-Golden. 2000. Modtran4 User’s Manual. Hanscom Air Force Base, MA: Air Force
Research Laboratory, Space Vehicles Directorate, Air Force Material Command.
Clough, S. A., M. J. Iacono, and J.-L. Moncet. 1992. Line-by-line calculations of atmospheric fluxes and cool-
ing rates: Application to water vapor. Journal of Geophysical Research 97:15761–15785.
Dutton, E. G. 1993. An extended comparison between LOWTRAN7-computed and observed broadband
thermal irradiances: Global extreme and intermediate surface conditions. Journal of Atmospheric and
Oceanic Technology 10:326–336.
Eppley Laboratory, Inc. 1995. Blackbody Calibration of the Precision Infrared Radiometer, Model PIR. Tech.
Procedure 05. Newport, RI: Eppley Laboratories.
Feltz, W. F., W. L. Smith, H. B. Howell, R. O. Knuteson, H. Woolf, and H. E. Revercomb. 2003. Near-continuous
profiling of temperature, moisture, and atmospheric stability using the atmospheric emitted radiance
interferometer (AERI). Journal of Applied Meteorology 42:584–597. doi:10.1175/1520-0450.
Gröbner, J. 2009. Final report of CIMO expert team on meteorological radiation and atmospheric composition
measurements, first session, February 6–10, 2006, Davos, Switzerland. Geneva, Switzerland: World
Meteorological Organization. CIMO/OPAG-SURFACE/ET-MR&ACM-2/Doc. 5.1, Item: 5.
Gröbner, J. 2012. A  transfer standard radiometer for atmospheric longwave irradiance measurements.
Meteorologia 49:S105–S111. doi:10.1088/0026-1394/49/2/S105.
Gröbner, J., I. Reda, S. Wacker, S. Nyeki, K. Behrens, and J. Gorman. 2014. A new absolute reference for
atmospheric longwave irradiance measurements with traceability to SI units. Journal of Geophysical
Research Atmospheres 119:7083–7090. doi:10.1002/2014JD021630.
Gröbner, J., I. Reda, S. Wacker, S. Nyeki, K. Behrens, and J. Gorman. 2015. Reply to comment by
R.  Philipona on “A  new absolute reference for atmospheric longwave irradiance measurements
with traceability to SI units.” Journal of Geophysical Research Atmospheres 120: 6885–6886.
doi:10.1002/2015JD023345.
Knuteson, R. O., H. E. Revercomb, F. A. Best, N. C. Ciganovich, R. G. Dedecker, T. P. Dirkx, S. C. Ellington,
et  al. 2004a. Atmospheric emitted radiance interferometer. Part I: Instrument design. Journal of
Atmospheric and Oceanic Technology 21:1763–1776.
Knuteson, R. O., H. E. Revercomb, F. A. Best, N. C. Ciganovich, R. G. Dedecker, T. P. Dirkx, S. C. Ellington,
et al. 2004b. Atmospheric emitted radiance interferometer. Part II: Instrument performance. Journal of
Atmospheric and Oceanic Technology 21:1777–1789.
Marty, C., R. Philipona, J. Delamere, E. G. Dutton, J. Michalsky, K. Stamnes, R. Storvold, T. Stoffel,
S. A. Clough, and E. J. Mlawer. 2003. Downward longwave irradiance uncertainty under arc-
tic atmospheres: Measurements and modeling. Journal of Geophysical Research 108:4358–43669.
doi:10.1029/2002JD002937.
240 Solar and Infrared Radiation Measurements

McArthur, L. J. B. 2004. Baseline Surface Radiation Network (BSRN). Operations Manual. WMO/TD-No.
879, World Climate Research Programme/World Meteorological Organization.
Philipona, R. 2001. Sky-scanning radiometer for absolute measurements of atmospheric long-wave radiation.
Applied Optics 40:2376–2383.
Philipona, R. 2015. Comment on “A new absolute reference for atmospheric longwave irradiance measure-
ments with traceability to SI units” by Gröbner et al. Journal of Geophysical Research Atmospheres
120:6882–6884. doi:10.1002/2014JD022990.
Philipona, R., E. G. Dutton, T. Stoffel, J. Michalsky, I. Reda, A. Stifter, P. Wendling, et al. 2001. Atmospheric
longwave irradiance uncertainty: Pyrgeometers compared to an absolute sky-scanning radiometer,
atmospheric emitted radiance interferometer, and radiative transfer model calculations. Journal of
Geophysical Research 106:28129–28141.
Philipona, R., C. Fröhlich, and C. Betz. 1995. Characterization of pyrgeometers and the accuracy of atmo-
spheric long-wave radiation measurements. Applied Optics 34:1598–1605.
Philipona, R., C. Fröhlich, K. Dehne, J. Deluisi, J. A. Augustine, E. G. Dutton, D. W. Nelson, et al. 1998.
The  baseline surface radiation network pyrgeometer round-robin calibration experiment. Journal of
Atmospheric and Oceanic Technology 15:687–696.
Reda, I., J. R. Hickey, J. Gröbner, A. Andreas, and T. Stoffel. 2006. Calibrating pyrgeometers outdoors inde-
pendent from the reference value of the atmospheric longwave irradiance. Journal of Atmospheric and
Solar-Terrestrial Physics 68:1416–1424.
Reda, I., J. R. Hickey, T. Stoffel, and D. Myers. 2002. Pyrgeometer calibration at the National Renewable
Energy Laboratory (NREL). Journal of Atmospheric and Solar-Terrestial Physics 64:1623–1629.
Reda, I., J. Zeng, J. Scheuch, L. Hanssen, B. Wilthan, D. Myers, and T. Stoffel. 2012. An absolute cavity
pyrgeometer to measure the absolute outdoor longwave irradiance with traceability to the ineternations
system of units, SI. Journal of Atmsopheric and Solar-Terrestrial Physics 77:132–143.
12 Net Radiation Measurements

Changes in the atmospheric abundance of greenhouse gases and aerosols, in solar radiation
and in land surface properties, alter the energy balance of the climate system. These changes
are expressed in terms of radiative forcing, which is used to compare how a range of human
and natural forces drive warming or cooling influences on global climate.
Intergovernmental Panel on Climate Change (2007)

12.1 INTRODUCTION
Net radiation (Rn) is the algebraic sum of the downwelling broadband solar, the same as global
horizontal irradiance (GHI), also called shortwave (S↓), the downwelling broadband infrared (IR),
also called longwave (L↓), the upwelling broadband shortwave (S↑), and the upwelling broadband
longwave radiation (L↑), that is,

Rn = S ↓ −S ↑ + L ↓ − L ↑ (12.1)

Net radiation is the energy that evaporates water or sublimates ice, produces sensible heat that raises
the temperature of the atmosphere, and heats surface soil or water. A small amount of net energy is
used to photosynthesize carbon dioxide and water into organic compounds, especially sugars and
oxygen in plants. Measuring the net energy with the highest accuracy is important because it is the
energy that drives weather and climate.
Figure  12.1 plots all of the variables in Equation  12.1 for two of the National Oceanic and
Atmospheric Administration (NOAA) Surface Radiation (SURFRAD) network sites on the same
day as a function of the local standard time at the site (http://www.srrb.noaa.gov/surfrad/index.
html). The downwelling components are in solid lines, and the upwelling components are dashed
lines. Obviously, it was clear at both sites with GHI reaching nearly the same maximum values
of about 800  watts per square meter (Wm−2). Since desert terrain at the Nevada station has a
higher albedo than the vegetated site in Mississippi, the reflected GHI (S↑) is higher on this
day at Desert Rock. Upwelling IR is elevated (daily maxima ranging from 400 to 600 Wm−2) at
both sites because the surface temperature and emissivity are greater than the atmosphere’s tem-
perature and emissivity. The difference in upwelling and downwelling IR at Goodwin Creek is
smaller than the difference at Desert Rock because of the higher atmospheric water vapor content
at Goodwin Creek (midday relative humidity was 10% and 40%, respectively). At both sites, the
net energy is negative at night because the ground is warmer than the atmosphere and there is no
solar radiation contribution to the energy budget. The daily integrated net energy (areas under
the dark solid curves) is larger at Goodwin Creek than at Desert Rock with higher values during
the day and night.
Four different types of commercial instruments are used to measure net radiation. The most
basic instrument measures net radiation by determining the thermopile voltage created by con-
necting the hot and cold junctions of a thermopile to the top and bottom absorbing surfaces of
the radiometer, respectively. The  thermopile is covered with a dome, often polyethylene, that

241
242 Solar and Infrared Radiation Measurements

FIGURE 12.1 Plots of all of the components of net radiation at the surface and the calculated net radiation
throughout the same clear day at two SURFRAD sites. GHI^ = (S↑) and IR^ = (L↑).

transmits most wavelengths of the shortwave and longwave spectrum. A second type of net radi-
ometer consists of two sensors that measure total (shortwave and longwave) downwelling and
total upwelling radiation separately. The  third type of net radiometer measures net shortwave
and net longwave separately. The fourth method to measure net radiation is to construct net radi-
ometers using four sensors that separately measure the four components of net radiation. These
instruments and their accuracies will be discussed in the next three sections.

12.2 SINGLE-SENSOR (ALL-WAVE) NET RADIOMETERS


The simplicity of concept and construction of a single-unit-construction net radiometer would
appear to be a good design. The hot junction of the thermopile is connected to the top of the
detector, which receives downwelling shortwave and longwave radiation, and the cold junction
is connected to the bottom of the detector, which receives upwelling shortwave and longwave
radiation. The difference in temperature of these matched receivers is proportional to the net
radiation. To reduce the effects of wind, dust, and precipitation on the detector, a material that
transmits both solar and IR radiation protects the surfaces, usually polyethylene. Since thermo-
pile responses are sensitive to temperature, it is best to include a measurement of the thermopile
temperature needed to make this correction. A bubble level is used to ensure the proper hori-
zontal orientation of the instrument, and a desiccant holder located within the housing is often
included to eliminate internal condensation. The cost of this sensor tends to be much lower than
the multi-sensor net radiometers. Examples of commercially available instruments of this type
of net radiometer are the Radiation and Energy Balance Systems (REBS) Q-7.1 and the Kipp &
Zonen net radiometer (NR) Lite2. Figure 12.2 is an example of a single-sensor (all-wave) net
radiometer.
The  polyethylene domes transmit most of the radiation in the short and long wavelengths.
The  bubble level allows for a proper (horizontal) orientation. Desiccated air flows through the
Net Radiation Measurements 243

FIGURE 12.2 This is an example of an all-wave net radiometer. It is a REBS model Q-7.1. Note the poly-
ethylene domes (the bottom dome is barely visible), the built-in level, and the air tube that keeps the domes
inflated.

FIGURE 12.3 This net radiometer is an alternate design for an all-wave net radiometer. It is a Kipp &
Zonen NR Lite2 that uses black Teflon-coated conical absorbers over the thermopile. (Courtesy of
Kipp & Zonen.)

support arm. Figure 12.3 is a novel design of the all-wave net radiometer using a rugged shield that
does not require constant airflow to keep the protective shield inflated. The vertical shaft is to dis-
courage birds from perching on this instrument.
There are design and operational issues when using these types of net radiometers. Some sensors
of this type use a flexible polyethylene dome to protect the detector from dust, wind, and precipita-
tion yet transmit most of the solar and IR wavelengths. These domes use a continuous desiccated
air or nitrogen stream to keep the domes inflated and free of condensation on the interior surfaces
(Figure 12.2). Some manufacturers now use thicker polyethylene domes so that continuous inflation
244 Solar and Infrared Radiation Measurements

is no longer required, but the air inside continues to need desiccation. Manufacturers suggest that
the domes be changed every 3–6 months, mainly because they can be easily scratched, and they
become brittle and crack under freezing conditions. The NR Lite2 (Figure 12.3), which has a black
Teflon conical absorber covering the sensors, is more rugged than the polyethylene domes, but this
instrument is sensitive to wind speed. All of these sensors show some dependence on wind speed
(Brotzge and Duchon 2000; Smith et al. 1997), and manufacturers provide corrections for wind-
speed dependency. Additionally, Brotzge and Duchon (2000) found that the effects of precipitation
are pronounced in the NR Lite. Imperfect Lambertian responses (responses that are proportional to
the cosine of the angle of incidence) are not thoroughly discussed in the literature, although Brotzge
and Duchon (2000) pointed out some problems with the NR Lite, the predecessor to NR Lite2.
Calibration is problematic because there is no accepted standard for this measurement. Calibrations
are discussed at the end of the chapter.

12.3 TWO-SENSOR NET RADIOMETERS


As mentioned in the introduction, two types of net radiometers employ two separate sensors;
however, their mode of operation is fundamentally different. The Schulze-Däke net radiometer
is composed of two separate pyrradiometers (2π steradian field of view instruments measuring
both long-wavelength and short-wavelength irradiance) that measure all downwelling radia-
tion with one and all upwelling radiation with the other. Polyethylene domes are used to shield
the sensors’ surfaces from the weather but are thick enough to be self-supporting. Halldin
and Lindroth (1992) found this instrument “showed superior performance, and its output was
almost entirely within the accuracy of the reference measurements” (p. 762). This was a com-
parison to the net radiation instruments of the day, many of which still exist in some form.
It appears that the Schulze-Däke net radiometer is no longer available. The Schenk 8111 pyr-
radiometer is similarly configured as the Schulze-Däke net radiometer and is available from
Ph. Schenk GmbH Wien (see Figure 12.4). Also, Radiation and Energy Balance Systems, Inc.
makes the THRDS7.1 similarly configured for net radiation measurements. A third company

FIGURE 12.4 This is the Schenk 8111. This instrument separately measures the downwelling and upwelling
longwave plus shortwave. It is similar to the Schulze-Däke net radiometer that is no longer available. It uses
thick polyethylene domes. (Courtesy of Schenk GmbH Wien & Co.)
Net Radiation Measurements 245

FIGURE 12.5 This instrument is the Kipp & Zonen CNR 2. It has two separate signals like the Schenk 8111,
but it measures net shortwave and net longwave. The domes are durable glass and silicon windows. (Courtesy
of Kipp & Zonen.)

making a net radiometer with this configuration is Middleton Solar. They began manufacturing
the NSR1 in 2013.
The Kipp & Zonen model CNR 2 measured net shortwave and net longwave using separate sen-
sors (Figure 12.5). When introduced, the CNR 2 was a relatively new design using glass domes over
the shortwave detector and silicon windows over the long-wave detector. The IR sensor field of view
is 150 degrees. This instrument was the only one of its type that was found. Its manufacture was
discontinued in early 2012.

12.4 FOUR-SENSOR NET RADIOMETERS


The Kipp & Zonen model CNR1 is the oldest of this genre and was replaced in 2009 by the CNR4
offering many design improvements. Introduced in 2007, Hukseflux’s model NR01 is another four-
component sensor available for net radiation measurements (see Figure 12.6). The NR01 IR mea-
surements have a field of view that is 150 degrees, as was the case for the older Kipp  & Zonen
CNR1. The  150 degree field of view instruments use a flat surface for better deposition of the
interference filter for the IR compared with the convex surfaces of the 180 degree instruments.
These four-component systems have the advantage that each sensor can be calibrated separately,
and cosine response of the downwelling shortwave sensor can be measured. The sensors provide
details of the incident radiation for each of the components of net radiation. These four-component
instruments do not, however, use the manufacturers’ top-of-the-line pyranometers or pyrgeometers
because of the high cost of such a system.
In a very careful study using the current and best pyranometer and pyrgeometer calibration tech-
niques, Michel et al. (2008) found that a heated and ventilated CNR1 could be field calibrated to
yield net radiation uncertainties smaller than 10 percent as claimed by the manufacturer. However,
using the manufacturer’s indoor calibrations led to significantly higher uncertainties. They found
that even with field calibrations that an unheated and unventilated CNR1 could not measure net
radiation with uncertainties under 10 percent.
246 Solar and Infrared Radiation Measurements

FIGURE 12.6 This instrument is an example of a four-component net radiometer, the NR1 by Hukseflux.


Each of the four components of net radiation is available separately. A  12  VDC, 1.5  W resistor-heater is
installed in each pyrgeometer case to mitigate dew and frost accumulation. (Courtesy of Hukseflux USA, Inc.,
Center Moriches, NY.)

12.5 ACCURACY OF NET RADIOMETERS


Blonquist et al. (2009) addressed the question of the accuracy for many of the currently manufac-
tured net radiometers. They found that, in general, the accuracy increases with the cost of the net
radiometer. The four-component systems were the best, followed by the two-component systems,
and then the all-wave net radiometers. However, one problem with this comparison is that there was
no true reference. As a proxy for the reference, this study used the average of each separate sensor
on the two Kipp & Zonen CNR1’s and the three Hukseflux NR01’s. The most significant problem
is the lack of a longwave, or broadband IR, standard. The lack of a broadband IR standard has been
a concern for a long time and has been mentioned early and often in the literature (Blonquist et al.
2009; Brotzge and Duchon 2000; Halldin and Lindroth 1992; Ohmura et al. 1998). An interim stan-
dard for the measurement of IR irradiance was developed at the World Radiation Center (WRC)
in Davos, Switzerland. Established in 2004, the interim World Infrared Standard Group (WISG)
(https://www.pmodwrc.ch/en/world-radiation-center-2/irs/) consists of four pyrgeometers that have
been calibrated using an absolute sky-scanning radiometer (ASR) and blackbody characterizations
(Gröbner et al. 2014; Marty et al. 2003; Philipona et al. 2001). This IR standard, along with the
World Radiometric Reference (WRR) for solar radiation (Finsterle 2006), presents an excellent
opportunity to develop a true net radiation standard.

12.6 A BETTER NET RADIATION STANDARD


With improvements in the measurement of the diffuse and direct components of solar radiation
(Michalsky et al. 2007, 2011) and the establishment of the interim WISG at the WRC in Davos,
Switzerland, for the measurement of IR (https://www.pmodwrc.ch/en/world-radiation-center-2/irs/),
there is the potential to develop an accurate working standard for net radiation.
Net Radiation Measurements 247

The  very best measurement of net radiation requires four first-class horizontal radiometers
(two pyranometers for shortwave and two pyrgeometers for IR) plus a first-class pyrheliometer
and a solar tracker with shading for the two downwelling, horizontally mounted radiometers.
These measurements are not  widely made, but several Baseline Surface Radiation Network
(BSRN) sites, which field first-class radiation sensors (Ohmura et  al. 1998), are making these
measurements. Due to the complexities associated with a 2π steradian field of view and other
pyranometer design considerations, the most difficult measurement of shortwave irradiance is the
downwelling component. The downwelling shortwave irradiance is much larger than the other
net radiation components and a small bias in the downwelling can overwhelm uncertainties in
the other components. The cosine response bias common to current pyranometers and the ther-
mal offset correction for properly calibrated pyranometers must also be taken into consideration
(Michalsky et al. 2007). Adding the independently measured direct normal component and the
DHI (measured with a zero-biased pyranometer) to compute the total horizontal shortwave irradi-
ance can minimize both the cosine response biases and the thermal offsets. Further, by shading
the downwelling IR from direct solar radiation, the solar correction to the IR irradiance measure-
ment is minimized.
Characterization of commercial net radiometers using any one of the BSRN sites that measure
all four radiation components would be a worthwhile study that should be undertaken in light of the
recent improvements in solar and IR radiometry.

12.7 NET RADIOMETER SOURCES


This list identifies known manufacturers who continue to produce net radiometers:

• EKO Instruments Co., Ltd. (http://www.eko.co.jp)


• Hukseflux Thermal Sensors B.V. (http://www.hukseflux.com/)
• Kipp & Zonen BV (http://www.kippzonen.com/)
• Middleton Solar (http://www.middletonsolar.com/)
• Ph. Schenk GmbH Wien (http://www.schenk.co.at/schenk/)
• Radiation and Energy Balance Systems, Inc. (no website; phone: 1(206) 624-7221)

QUESTIONS
1. What is considered the most accurate of the four methods for measuring net radiation at the
surface?
2. What are the four different types of net radiometers?
3. Why is polyethylene a popular dome material for net radiometers?
4. Why is the fractional accuracy of net radiation lower than the individual radiation
measurements?
5. Why is the measurement of net radiation important?

REFERENCES
Blonquist, Jr., J. M., B. D. Tanner, and B. Bugbee. 2009. Evaluation of measurement accuracy and comparison
of two new and three traditional net radiometers. Agricultural and Forest Meteorology 149:1709–1721.
Brotzge, J. A., and C. E. Duchon. 2000. A  field comparison among a domeless net radiometer, two four-
component net radiometers, and a domed net radiometer. Journal of Oceanic and Atmospheric
Technology 17:1569–1582.
Finsterle, W. 2006. WMO International Pyrheliometer Comparison IPC-X. September 26–October  14,
Davos, Switzerland, Final Report 91, WMO/TD No.  1320, PMOD/WRC Internal Report. Davos,
Switzerland: WMO.
248 Solar and Infrared Radiation Measurements

Gröbner, J., I. Reda, S. Wacker, S. Nyeki, K. Behrens, and J. Gorman. 2014. A new absolute reference for
atmospheric longwave irradiance measurements with traceability to SI units. Journal of Geophysical
Research: Atmosphere, 119. doi:10.1002/2014JD021630.
Halldin, S., and A. Lindroth. 1992. Errors in net radiometry: Comparison and evaluation of six radiometer
designs. Journal of Oceanic and Atmospheric Technology 9:762–783.
Marty, C., R. Philipona, J. Delamere, E. G. Dutton, J. Michalsky, K. Stamnes, R. Storvold, T. Stoffel,
S. A. Clough, and E. J. Mlawer. 2003. Downward longwave irradiance uncertainty under arc-
tic atmospheres: Measurements and modeling. Journal of Geophysical Research 108:4358–4369.
doi:10.1029/2002JD002937.
Michalsky, J., E. G. Dutton, D. Nelson, J. Wendell, S. Wilcox, A. Andreas, P. Gotseff, et al. 2011. An extensive
comparison of commercial pyrheliometers under a wide range of routine observing conditions. Journal
of Atmospheric and Oceanic Technology 28:752–766.
Michalsky, J. J., C. Gueymard, P. Kiedron, L. J. B. McArthur, R. Philipona, and T. Stoffel. 2007. A  pro-
posed working standard for the measurement of diffuse horizontal shortwave irradiance. Journal of
Geophysical Research 112:D16112. doi:10.1029/2007JD008651.
Michel, D., R. Philipona, C. Ruckstuhl, R. Vogt, and L. Vuilleumier. 2008. Performance and uncertainty
of CNR1 net radiometers during a one-year field comparison. Journal of Oceanic and Atmospheric
Technology 25:442–451.
Ohmura, A., H. Gilgen, H. Hegner, G. Müller, M. Wild, E. G. Dutton, B. Forgan, et al. 1998. Baseline Surface
Radiation Network (BSRN/WCRP): New precision radiometry for climate research. Bulletin of the
American Meteorological Society 79:2115–2136.
Philipona, R., E. G. Dutton, T. Stoffel, J. Michalsky, I. Reda, A. Stifter, P. Wendling, et al. 2001. Atmospheric
longwave irradiance uncertainty: Pyrgeometers compared to an absolute sky-scanning radiometer,
atmospheric emitted radiance interferometer, and radiative transfer model calculations. Journal of
Geophysical Research 106:28129–28141.
Smith, E. A., G. B. Hodges, M. Bacrania, H. J. Cooper, M. A. Owens, R. Chappell, and W. Kincannon. 1997.
BOREAS Net Radiometer Engineering Study. NASA  Contractor Report, NASA  Grant NAG5-2447.
Greenbelt, MD: NASA Goddard Space Flight Center.
13 Radiometer Calibrations

I have been struck again and again by how important measurement is to improving the human
condition.
Bill Gates

13.1 INTRODUCTION
Proper calibration of measurement and test equipment is essential for producing accurate and reli-
able data. Radiometry continues to be one of the more challenging types of measurement due to
instrument exposure to varying weather conditions during outdoor measurements and the damag-
ing effects of the very radiation the instrument is designed to measure. Additionally, the calibration
of solar and infrared (IR) radiometers depends on detector-based reference standards rather than
a measurement standard that relies on the first principles of physics. For  example, based on the
Josephson effect, the volt is known to within a few tenths of a part per million (ppm) (Taylor 1989),
while the World Radiometric Reference (WRR) for solar direct normal irradiance (DNI), based on
measurements by the World Standard Group (WSG), has an estimated accuracy of 3,000 ppm or
±0.3 percent (Fröhlich 1978).
This chapter introduces the key elements of radiometer calibration and provides context for the
more detailed discussions found in other chapters describing specific radiometer types, their esti-
mated measurement uncertainties, and common measurement applications.

13.1.1  WHat is CalibRation?


In  metrology,1 the science of measurement, calibration is a process of comparing measurement
values produced by a device under test with those from a calibrated instrument with measurement
performance that is traceable to an internationally recognized reference measurement standard.
When these calibrated instruments with traceability to international measurement standards are
used to calibrate many other instruments, they are referred to as reference instruments. The cur-
rently recognized international solar and IR radiation reference measurement standards are:

• World Radiometric Reference (WRR): Broadband shortwave (solar)


• Interim World Infrared Standard Group (WISG): Broadband longwave (infrared)
• Standard lamps: Spectral shortwave (solar)

In  1971, following decades of developing competing scales for solar radiation measurements,
the World Meteorological Organization (WMO) designated the Physikalisch-Meteorologisches
Observatorium Davos (PMOD) as the World Radiation Center (WRC). The PMOD/WRC devel-
oped the WRR and WISG and continues to improve these measurement references (Nyeki et al.
2017). Reference standard lamps have been developed and maintained by various National
Metrology Institutes (NMIs), such as the National Institute of Standards and Technology (NIST)
in the United States, the National Physical Laboratory (NPL) in the United Kingdom, and the

1 Metrology addresses the needs for (1) establishing measurement units, reproducing these units based on measurement
standards, and ensuring the uniformity of measurements, (2) developing methods of measurement, (3) analyzing the
accuracy of measurement methods, establishing uncertainty of measurement, researching causes of and mitigation meth-
ods for measurement biases and uncertainties.

249
250 Solar and Infrared Radiation Measurements

Physikalisch-Technische Bundesanstalt (PTB) in Germany. Radiometer calibrations and their sub-


sequent measurements should be traceable to internationally recognized references and measure-
ment units.
The concept of calibration traceability addresses the needs for a process to achieve consistent
measurements within stated uncertainties. Radiometers, like any piece of measurement and test
equipment, should provide traceable measurements through an unbroken chain of comparisons to
a known measurement standard. The  comparisons should include an accredited facility that has
demonstrated technical competence (ISO/IEC 2017) and provide field instruments with calibration
traceability as illustrated in Figure 13.1.
Radiometer calibrations are done indoors using artificial sources (lamps or blackbody chambers)
or outdoors using the Sun and sky as the source of radiation. These calibrations are performed
by radiometer manufacturers (see Appendix G) and may also be done by government-funded
national standards laboratories or by private metrology services (Habte et  al. 2016; Nyeki et  al.
2017). Existing internationally recognized reference measurement standards for broadband solar
and infrared radiation are detector-based and maintained by the WRC. Solar spectral irradiance
scales traceable to Système International d’Unite ́s (SI) units are realized with lamps at NIST in the
United States, NPL in the United Kingdom, and PTB in Germany.
Calibration is not  characterization. A  radiometer calibration provides the instrument-specific
responsivity (Rs) specifying the value for converting the radiometer output signal to irradiance (micro-
volts per watts per square meter (µV/Wm−2)) and its estimated measurement uncertainty. Calibration
results are based on comparisons with a reference instrument whose calibration is traceable to a rec-
ognized reference measurement standard, such as the WRR, under the specific conditions at the time
of calibration. Radiometer characterization, on the other hand, is based on calibrations obtained with

FIGURE 13.1 Radiometer calibration traceability is an unbroken chain of comparisons to reference mea-


surement standards.
Radiometer Calibrations 251

FIGURE 13.2 Calibration and characterization results for a pyranometer indicating responsivity at 45° solar
zenith angle Rs(45) and Expanded Uncertainty U95 of the calibration results plotted against solar zenith angle
in (a) and against time of day in (b).

the variation of parameters affecting the measurement performance (i.e., solar position, air tempera-
ture, irradiance levels, wind conditions, etc.; Andreas and Wilcox 2016). An example of an outdoor
calibration of a thermopile-based pyranometer is shown in Figure 13.2. Notice the calibration result
is given as the instrument’s responsivity at 45° solar zenith angle and the associated estimated uncer-
tainty is based on the results for solar zenith angles between 30 and 60 degrees.
Recognizing the needs for improved calibrations with results that can be accepted by other
organizations, the international measurements community has developed standards for accrediting
measurement and test equipment calibrations. ISO 9000 and ISO 17025 standards address the design
of Quality Management Systems that include documented calibration methods and procedures for
producing traceable calibration values with statements of estimated measurement uncertainty to a
specified confidence level (ISO 2015).
Terminology as applied to radiometry is varied and can be confusing. For improved clarity and
consistency in this book, the types of instruments used for solar and IR radiation measurements based
on their performance specifications, estimated measurement uncertainty, and position in the calibra-
tion traceability chain are presented in Table 13.1. Internationally recognized Reference Measurement
Standards for broadband radiometry are traceable to SI units and based on well-characterized instru-
ments such as the WSG used to realize the WRR. Spectral solar irradiance standards are currently
based on lamps available from commercial suppliers and the NMIs as listed in Table 13.1. Reference
Instruments are limited to those radiometers that have been compared directly with the radiometers
defining the measurement standard. Examples include absolute cavity radiometers (ACRs) that have
participated in an International Pyrheliometer Comparison (IPC) (Finsterle 2016). WMO regional
radiation centers, radiometer manufacturers, and research laboratories typically maintain from one
to three radiometers at this level of calibration traceability and measurement uncertainty to represent
the reference measurement standards for their organization. These radiometers are then used at the
discretion of the owner to develop and maintain Transfer Reference Instruments specifically for the
purpose of protecting the reference standards as part of their internal calibration chain. Typically, a
small group of these transfer radiometers then is used to calibrate and maintain Working Reference
Instruments that are used for more frequent calibrations of Process Control Instruments, Field
Instruments, and Spare Instruments. This last group of radiometers are generally production instru-
ments of the same type and model with the largest acceptable estimated measurement uncertainties
for a measurement project. The Process Control Instruments, however, are used only during calibra-
tions of the field and spare radiometers to monitor the consistency of a calibration process based on
the performance of the Working Reference Instruments forming the basis of the calibration.
252 Solar and Infrared Radiation Measurements

TABLE 13.1
Terminology for Radiometer Calibration Hierarchy
No. Radiometer Name Example
1 Reference Measurement Standard World Radiometric Reference (WRR) developed and maintained by the
World Radiation Center (WRC) in Davos, Switzerland
2 Reference Instrument Absolute cavity radiometer with direct traceability to the WRR by
intercomparison at PMODa
3 Transfer Reference Instrument Absolute cavity radiometer traceable to the WRR by outdoor comparison
with a reference instrumentb
4 Working Reference Instrument Absolute cavity radiometer used to perform regular calibrations of field
instruments
5 Calibration Process Control Radiometers used only during calibration events to monitor the consistency of
Instrument(s) results
6 Field Instruments Radiometers used for routine outdoor measurements
7 Spare Instruments Radiometers held in reserve for exchange of field instruments due for
calibration or damaged during service

Note: Calibration of these radiometers must be currently valid prior to deployment.


a PMOD Davos, Switzerland conducts IPCs every 5 years to transfer the WRR to participating instruments (Finsterle 2016).

b NREL hosts regular comparisons of reference and transfer standard pyrheliometers at the SRRL in Golden, Colorado,

USA (Reda et al. 2016).

13.1.2 WHy is CalibRation needed?


Radiometer calibration is an important part of the overall operations and maintenance of a mea-
surement station or network of stations to ensure data quality. Estimating measurement uncertainty
of field data begins with the contribution of radiometer calibration uncertainty to the overall mea-
surement uncertainty budget. Regular calibrations are also important for addressing radiometer
measurement performance changes over time due to exposure and material aging. Compliance
with accepted measurement protocols (i.e., WMO/ Commission for Instruments and Methods of
Observation (CIMO),2 BSRN,3 and Atmospheric Radiation Measurement (ARM)4) requires peri-
odic calibration using prescribed methods.
Radiometer calibration also may be needed for these conditions:

• Installing a new instrument


• After an instrument has been repaired or modified
• After a specified time of operation has elapsed
• Before and/or after a critical measurement period
• After an instrument has been exposed to a shock, excessive vibration, or physical damage,
which may have compromised the calibration integrity
• When radiometer readings appear questionable or are not consistent with other measure-
ments or general atmospheric conditions
• As specified by a requirement, e.g., customer specification, radiometer manufacturer rec-
ommendations, or an applicable standard.

2 World Meteorological Organization, Commission on Instruments and Methods of Observation—http://www.wmo.int/


pages/prog/www/IMOP/CIMO-Guide.html.
3 Baseline Surface Radiation Network—http://bsrn.awi.de/.
4 Atmospheric Radiation Measurement Program—https://www.arm.gov/.
Radiometer Calibrations 253

Calibration values and their associated uncertainties are essential when using measured data to
evaluate the performance of a solar energy system, especially when this information is being evalu-
ated by investors considering purchasing or providing loans to a project. These calibrations are
essential to the “bankability” of the dataset because without the calibration information, there is no
way to evaluate the relative production estimates produced by using the data.

13.1.3 HoW fRequently sHould a RadiometeR be CalibRated?


The frequency of radiometer calibration is one of several important factors affecting the accuracy
of solar and IR radiation measurements:

• Radiometer performance specifications


• Equipment installation
• System maintenance procedures
• Regular data quality assessments

In general, the best practice is to follow the manufacturer’s recommendations for the radiom-
eter calibration interval as part of a well-defined operations and maintenance procedure. WMO
recommends field radiometers be regularly calibrated at 1- to 2-year intervals, depending on the
type of use and measurement conditions (WMO 2017). When planning radiometer calibration
intervals, it is important to consider the data quality needs, the local environmental conditions
at the measurement site, and the duration of instrument exposure. Determining the optimum
calibration interval can be based on initially scheduling annual or more frequent radiometer
calibrations, depending on the degree and type of exposure of the radiometer. This calibration
history then can be used to identify any measurable changes or trends in radiometer perfor-
mance with use and their effects on the resulting measurement uncertainty estimates for the
field measurements.
After many decades of technological advances, the demand for more accurate measurements
of solar or IR radiation continues to challenge radiometer manufacturers to improve their radi-
ometer designs and provide affordable products. A  major challenge is the degradation of the
materials used in radiometers caused by exposure to extreme weather conditions and to the
damaging radiation they are designed to measure. The resulting “drift” in radiometer respon-
sivity (µV/Wm−2) over time affects the performance classification according to stability limits
set by the International Organization for Standards (ISO) and WMO (ISO 2018; WMO 2017).
The “non-stability” limits for broadband radiometers defined in ISO 9060 as the maximum per-
centage change in responsivity per year range from ±0.01 percent per year for Class AA pyrhe-
liometers to ±3 percent per year for Class C pyranometers (see Chapters 4 and 5 for radiometer
classifications).
Examples of pyrheliometer and pyranometer calibration histories shown in Figure 13.3 illustrate
the long-term trends of control instruments calibrated with the Broadband Outdoor Radiometer
Calibration (BORCAL) method developed by the National Renewable Energy Laboratory (NREL;
Myers et al. 2002). These control instruments are deployed only during calibration measurement
periods to monitor the repeatability of the calibration process. The quality assurance procedure
for certifying each BORCAL event requires the inspection of results for the control instruments to
confirm repeatability of the calibration values (i.e., responsivity values expressed in microvolts per
watt per square meter, µV/Wm–2) with previous calibrations. The average expanded measurement
uncertainties (+U95 and −U95) shown in Figure 13.3 are the averages of the 95 percent uncertainty
values (i.e.,  where the coverage factor equals 1.96 as described in Appendix A) determined for
each instrument during the BORCAL events conducted over the 22-year period. Because the mea-
surement performance of the BORCAL reference instruments was regularly characterized during
the period and the calibration process remained fundamentally unchanged, the rather consistent
254 Solar and Infrared Radiation Measurements

FIGURE 13.3 Calibration histories for two pyrheliometers (a) and three pyranometers (b) used as Control
Instruments during BORCAL events over a 22-year period.

trends of decreasing responsivities for the pyrheliometers could be considered mostly within the
estimated measurement uncertainties. However, the precision of the BORCAL results suggests a
measurable trend in the performance of all radiometers that could be attributed to the aging of
materials in spite of their limited exposure to solar radiation as control instruments. For compari-
son, the calibration histories for two field instruments deployed in the Tropical Western Pacific5
for continuous measurements, except during the calibration periods, over 15 years exhibit similar
changes over time as shown in Figure 13.4. The expanded measurement uncertainties with a 95
percent confidence interval (+U95 and −U95) for each calibration of the two operational radiometers
are shown in the figure.

5 These and other radiometers are used by the U.S. Department of Energy ARM Program at their Solar Infrared Radiation
Stations (SIRS), Ground Radiation (GNDRAD), and Sky Radiation (SKYRAD) locations: https://www.arm.gov/.
Radiometer Calibrations 255

FIGURE  13.4 Calibration results for an operational pyrheliometer (a) and pyranometer. (b) deployed for
more than 15 years in the Tropical Western Pacific.

13.2 BROADBAND SHORTWAVE RADIOMETER CALIBRATION


The WRR is a detector-based reference measurement standard composed of electrically self-calibrating
ACRs called the WSG. The WRR is the internationally recognized reference measurement standard
for broadband solar irradiance measurements (Fröhlich 1991). The WRR applies only to DNI measure-
ments above 700 W/m2. As shown in Figure 13.1, the WRC provides calibration traceability for short-
wave radiometer calibrations by hosting IPCs at their laboratory every 5 years in Davos, Switzerland
(Finsterle 2016). Efforts to demonstrate the traceability of WRR to the SI have indicated a relative
difference of about −0.3 percent with the SI scale being lower (Walter et al. 2017). At the time of this
writing, WMO/CIMO is reviewing the formal transition from the WRR to the SI scale based on the per-
formance of the recently developed Cryogenic Solar Absolute Radiometer (CSAR; Nyeki et al. 2017).
256 Solar and Infrared Radiation Measurements

With the advances in radiometer designs and measurement performance, a series of consensus
standards for radiometer calibrations have been revised to meet the increased demands for accurate
and traceable calibrations. Table 13.2 summarizes the standards as set forth by the American Society
of Testing and Materials (ASTM International) and the International Organization for Standards
(ISO) for calibrating pyrheliometers, pyranometers, ultraviolet radiometers, and spectroradiometers.
Additional instrument-specific information is available in Chapters 4, 5, and 7.
To provide cost-effective products, radiometer manufacturers generally rely on indoor calibra-
tion methods compliant with the appropriate standards listed in Table 13.2. Independent and accred-
ited calibration service providers also base their work on these standards.
As previously mentioned, the BORCAL method is one approach for calibrating pyrheliometers
and pyranometers outdoors with traceability to the WRR (Andreas et al. 2018; Andreas and Wilcox
2016; Reda et al. 2008; Wilcox et al. 2002). The method follows the calibration traceability concept
shown in Figure  13.1, incorporates procedures consistent with the applicable standards listed in
Table 13.2, and involves the following steps for calibrating up to 100 radiometers per event:

1. Establish and maintain Transfer Reference Instruments: Multiple electrically self-calibrating


absolute cavity radiometers with direct traceability to the WRR by regular participation in
the IPCs at the WRC (Finsterle 2016).

TABLE 13.2
Standard Methods for Radiometer Calibration
Standard Title Description
ISO 9846:1993 Solar energy—Calibration of field Outdoor with pyranometer in horizontal position and
pyranometers by using a pyrheliometer an indoor method using an integrating sphere and
lamp.
ISO 9847:1992 Solar energy—Calibration of field Outdoor with field pyranometer in horizontal, tilted, or
pyranometers by comparison to a normal incidence and an indoor method using an
reference pyranometer integrating sphere and lamp.
ASTM E816-15 Standard test method for calibration of Outdoor comparison with an electrically self-
pyrheliometers by comparison to reference calibrating absolute cavity pyrheliometer as the
pyrheliometers primary reference instrument.
ASTM E824-10 Standard test method for transfer of Harmonized with ISO 9847 for pyranometer
calibration from reference to field calibrations.
radiometers
ASTM G130-12 Standard test method for calibration of Primary reference instrument can be a scanning or
narrow- and broad-band ultraviolet linear-diode-array spectroradiometer.
radiometers using a spectroradiometer
ASTM G138-12 Standard test method for calibration of a Calibration of spectroradiometers using a standard of
spectroradiometer using a standard source spectral irradiance (lamp) traceable to a national
of irradiance metrological laboratory that has participated in
inter-comparisons of standards of spectral irradiance.
ASTM G167-15 Standard test method for calibration of a Harmonized with ISO 9847 for pyranometer
pyranometer using a pyrheliometer calibrations.
Type I calibrations use self-calibrating absolute cavity
radiometers. Type II use secondary reference
pyrheliometers as the Class A reference standard as
defined in ISO 9060:2018.
ASTM G207-11 Standard test method for indoor transfer of Calibrations use artificial light sources (lamps).
calibration from reference to field
pyranometers
Radiometer Calibrations 257

2. Establish and maintain Working Reference Instruments: Apply ASTM E816-15  to out-
door measurement comparisons of additional ACRs identified specifically for use during
field radiometer calibration measurement periods with the Transfer Standard Instruments
to establish individual WRR Transfer Factors for the Working Standard radiometers.
3. Develop a Reference Diffuse group: Apply the methods described in ASTM G167-15 for
this shade-unshade pyranometer calibration method to a group of commercially available
thermopile-based pyranometers (Reda et al. 2003; Reda and Andreas 2017). This method
provides for a low estimated calibration uncertainty but can require considerable effort in
the absence of an automated solar tracker or groups of trackers capable of performing this
calibration for multiple pyranometers.
4. Collect calibration measurements: Apply the semi-automated Radiometer Calibration
and Characterization (RCC) process to install, monitor, and collect clear-sky measure-
ments from the Working Reference Absolute Cavity Radiometer, Reference Diffuse
Pyranometers, and the radiometers under test (Andreas and Wilcox 2016).
5. Produce Radiometer Calibration Certificates: Review calibration measurements from
the reference radiometers and each radiometer under test for compliance with calibration
event selection criteria, generate the individual radiometer responsivities (µV/Wm−2) and
estimated expanded uncertainties (±U95%) as computed by RCC based on the summation
technique for pyranometers and on the comparison technique for pyrheliometers, dissemi-
nate the results in a BORCAL report (https://aim.nrel.gov/borcal.html), and update the
calibration archival database. See Appendix H for an example of a BORCAL file.

There are other calibration procedures in use by organizations responsible for providing metrology
services. The methods must comply with the individual Quality Management System developed by
each organization and be based on the recognized standards. Accredited calibration facilities must
also demonstrate periodic comparability of their results with other recognized organizations.

13.3 BROADBAND LONGWAVE RADIOMETER CALIBRATION


Longwave radiation is measured with a pyrgeometer designed to detect IR irradiance over the wave-
length range of approximately 4–50 µm. This section provides a brief introduction to the concepts
of pyrgeometer design, the use of blackbody calibrators, the development of an internationally
recognized reference measurement standard for downwelling infrared irradiance, and the recent
advances in instrument design that could lead to an improved reference measurement standard for
this important form of radiation.
Pyrgeometers are designed to measure radiation at wavelengths beyond our light-adapted phot-
opic response. Invisible to the human eye, IR radiation can be distinguished from the “light” we
see, nominally wavelengths between 400 and 700 nm, and the “heat” we perceive at wavelengths
greater than 3000 nm. Understanding the design and operation of pyrgeometers requires us to think
in terms of temperature and heat exchanges rather than intensity and colors of visible light.
Most molecules in the atmosphere are poor absorbers of visible and IR radiation and most radi-
ation passes through. Some molecules such as water vapor and carbon dioxide (CO2) are good
absorbers of IR radiation. Clouds and atmospheric conditions significantly affect short and long-
wave radiation that, in turn, significantly affects the temperature of Earth and the atmosphere.
For example, clouds generally absorb or reflect most direct sunlight that decreases the shortwave
irradiance at the surface. These same clouds, composed primarily of water vapor or ice, absorb or
reflect IR radiation from the Earth and re-emit most of it back to Earth. Without clouds, most of the
IR radiation from the Earth escapes into space. For a sense of scale, pyrgeometer measurements
from a warm and humid location can experience a range of downwelling IR irradiance sampled
over a few days of between 260 and 420 Wm−2 while a colder and drier location can produce short-
term pyrgeometer measurements from 120 to 240 Wm−2 (Marty et al. 2003; Philipona et al. 2001).
258 Solar and Infrared Radiation Measurements

As validated by surface measurements made at eight BSRN stations over a 10-year period (1984–
1993), the nominal globally averaged value of downwelling IR irradiance was estimated to be
between 342 and 344 Wm−2 (Pavlakis et al. 2004).6 More recent studies used 343 Wm−2  to model
the global energy balance (Wild et al. 2013).
Historically, the calibration of pyrgeometers for measuring longwave (IR) radiation has been
based on so-called “blackbody calibrators.” Using custom-designed systems, pyrgeometer calibra-
tions made by the instrument manufacturers, NMIs, and others have been based on accurately
measuring the temperature and characterizing the effective emissivity of the blackbody source and
applying the Stefan-Boltzmann law relating the radiated power and temperature:

E = εσ T 4 (13.1)

where:
E is the power emitted in Wm−2
ε is the emissivity of the source (0–1)
σ is the Stefan-Boltzmann constant equal to 5.6704 × 10−8 Wm−2K−4
T is the source temperature in Kelvin (K)

Similar to pyranometers used for shortwave or solar irradiance measurements, commercially avail-
able pyrgeometers have a black thermopile-type detector mounted under a protective window (dome
or disk). The window also acts as a solar blind filter, limiting the radiation available to the detector
to the 4–50 micrometers (µm) (4000–50000 nanometers (nm)) longwave spectral region. A pyrge-
ometer also requires the means for accurate measurements of internal temperature or temperatures
to determine the thermal environment within the instrument at the time of measurement. The pyr-
geometer’s thermopile detector signal (µV) is compared with the emitted IR flux while accounting
for the thermal energy exchange within the instrument. Figure 13.5 illustrates the basic pyrgeometer
design elements and the principal IR flux exchange:

Wnet = Win − Wout + ∆W (13.2)

FIGURE  13.5 Basic design elements and IR fluxes of a pyrgeometer. (Based on The Eppley Laboratory,
Model PIR construction elements and signal cable connections: Thermopile = A and C, Dome Thermistor = F
and G, and Case Thermistor = D and E).

6 Pavlakis et al. (2004) stated their model bias was −2.4 Wm–2 with a scatter (RMSE) of 12 Wm–2 when compared with
measurements from the eight BSRN stations. Summed in quadrature, this result suggests an expanded uncertainty of
±24 Wm–2 with a coverage factor of 2.0 (see Appendix A).
Radiometer Calibrations 259

where:
Win = Incoming IR flux, e.g., blackbody or sky [Wm−2]

ε skyσ Tsky
4

Wout = Outgoing IR flux, [Wm−2]

ε receiverσ Treceiver
4

∆W = Dome—Case flux interaction at the detector, [Wm−2]

ε caseσ Tcase
4
− ε domeσ Tdome
4

Wnet = C Vtp [Wm−2]
ε = Effective emissivity [Ratio]
σ = Stefan-Boltzmann Constant (see Equation 13.1)
T = Temperature [K]
C = Pyrgeometer thermopile calibration factor [Wm−2/µV]
Vtp = Pyrgeometer thermopile output voltage [µV].

Details of the IR flux exchange and the development of the equation used for IR irradiance measure-
ments by a pyrgeometer are available in Chapter 11. The extended Albrecht and Cox equation used
by PMOD/WRC is presented here for reference in this chapter (Philipona et al. 1995):

U
E=
C
( )
1 + k1σ TB3 + k2σ TB4 − k3σ TD4 − TB4( ) (13.3)
where:
E = Infrared radiation incident to the pyrgeometer [Wm−2]
U = Measured pyrgeometer thermopile voltage [µV]
C = Pyrgeometer thermopile responsivity [µV/Wm−2]
ki = Pyrgeometer constants from characterizations with reference blackbody
σ = Stefan-Boltzmann Constant (see Equation 13.1)
TB = Measured pyrgeometer body temperature [K]
TD = Measured pyrgeometer dome temperature [K].

By the mid-1990s, the growing interest to improve pyrgeometer calibrations and IR radiation measure-
ments resulted in the round-robin comparisons of 11 blackbody systems developed and operated by as
many international institutions (Philipona et al. 1998). Lacking a recognized pyrgeometer calibration
standard at the time, the blackbody designs used for the round-robin were of unique geometry and mate-
rials including ice cones of pure water, hemispherical or cylindrical brass or aluminum masses, and a
stainless-steel double-walled cylinder with a conical ceiling. Similarly, the data acquisition systems and
operating protocols were unique for each participant. The results of calibrating five Eppley Laboratory,
Inc. Model precision infrared radiometer (PIR) pyrgeometers at each of the 11 participating organiza-
tions provided the first benchmark for estimating pyrgeometer calibration uncertainties based on black-
bodies: The PIR round-robin calibration results indicated the PIR instrument design performance can be
considered stable. The various blackbody calibration systems, however, were shown to produce results
within 1 to 20 percent of the median thermopile responsivity for each of the 5 PIRs, depending on the
institution. The results confirmed the needs for a standard calibration reference and procedure.
Pursuing a new detector-based reference measurement standard and subsequent to the pyrgeom-
eter calibration round-robin, the WRC developed the ASR (Philipona 2001). The ASR used a pyro-
electric detector, an integral temperature-controlled reference blackbody source, and a rotating gold
mirror. The emittance of the blackbody source was well characterized and its temperature measure-
ments were traceable to an absolute temperature standard. Using the rotating gold mirror, longwave
260 Solar and Infrared Radiation Measurements

sky radiance measurements and self-calibrations based on the blackbody were alternated to deter-
mine the absolute value of the longwave irradiance produced by the ASR. The ASR participated in
two international intercomparisons representing unique ranges of longwave sky radiance. The First
International Pyrgeometer and Absolute Sky-scanning Radiometer Comparisons (IPASRC-I) was
conducted at the U.S. Department of Energy ARM’s Southern Great Plains (SGP) Radiometer
Calibration Facility near Lamont, Oklahoma (Philipona et al. 2001). The second IPASRC was con-
ducted at ARM’s North Slope of Alaska (NSA) site in Barrow (Marty et al. 2003). These inter-
comparison results showed that estimated uncertainties of longwave measurements and radiative
transfer models can approach ±2 Wm−2  and that improvements to longwave measurements would
require an internationally recognized reference measurement standard.
By 2004, PMOD/WRC had established the WISG of pyrgeometers to serve as an interim trans-
fer standard group with respect to its reference scale (WMO 2006). The  WISG consists of four
pyrgeometers—two modified Eppley Model PIRs and two Kipp  & Zonen Model CG4s)—that
were calibrated relative to the ASR and have not been changed since (Gröbner and Wacker 2015).
The estimated expanded measurement uncertainty for pyrgeometer calibrations based on the WISG
is ±2.6 Wm−2 with a 95 percent level of confidence (Nyeki et al. 2017). The continued need for a more
accurate reference measurement standard has prompted the independent development of two new
absolute measurement systems. The Infrared Integrating Sphere (IRIS) radiometer (Gröbner 2012)
and the Absolute Cavity Pyrgeometer (ACP) measurement system (Reda et al. 2012) are windowless
to eliminate the measurement performance complications due to the optical and thermal complexi-
ties presented by the presence of an enclosure window (see Figures 13.6 and 13.7). The windowless

FIGURE 13.6 Design elements of the IRIS radiometer. (1) Aperture for downwelling irradiance (the 8 mm
diameter opening has no window), (2) rotating shutters, 90 degrees out of phase, (3) windowless pyroelectric
detector, (4) reference blackbody cavity and thermistor, (5) shutter motors, (6) 60-mm diameter gold-plated
integrating sphere. (From Gröbner, J., Metrologia, 49, S105–S111, 2012.)
Radiometer Calibrations 261

FIGURE 13.7 ACP design concept. (1) Cylindrical, gold-plated compound parabolic concentrator (CPC),
(2) gold anodized external housing with signal cable, (3) two sets of three thermistors positioned at 120
degree intervals around the CPC, (4) thermal insulator, (5) pyrgeometer thermopile detector, (6) the Eppley
Laboratory, Inc. Model PIR with dome removed, and (7) temperature-controlled aluminum plate.

design of these advanced research measurement systems does, however, limit their use to fair
weather, nighttime conditions. The WISG instruments are designed for all-weather operations and
provide continuous measurements for pyrgeometer calibrations. The four WISG pyrgeometers typi-
cally agree to within ±1 Wm−2. Results of two measurement intercomparisons of the WISG, IRIS,
and ACP instruments under nighttime clear-sky conditions have demonstrated a need for revising
the existing interim WISG (Gröbner et al. 2014). The IRIS and ACP measurements under night-
time clear-sky conditions agree within ±1.5  W m−2 (Gröbner and Thomann 2018). Comparisons
of the WISG, IRIS, and ACP indicate the WISG is underestimating longwave irradiance during
these same conditions by 2 to 6 W m−2, depending on the amount of integrated atmospheric water
vapor (WV). During outdoor measurement conditions when WV exceeds 10 millimeters (mm), an
average thermopile sensitivity (parameter C in Equation 13.3) correction of +6.5 percent is needed
for the WISG pyrgeometers to better agree with the ACP and IRIS measurements (Gröbner et al.
2014; Gröbner and Thomann 2018). Until a more accurate measurement reference is developed for
broadband IR radiation measurements under all weather conditions, pyrgeometers calibrated to the
WISG will likely require a similar thermopile sensitivity increase determined from the adjusted
WISG and archived calibration data for U, TB, and TD (Nyeki et al. 2017).
The measured difference between the interim WISG and the IRIS and ACP instruments has prompted
a need for more investigation leading to a new reference measurement standard. At  the time of this
writing, additional IRIS and ACP instruments have been produced for further characterizations with
the goal of reducing the estimated measurement uncertainty of these absolute radiometers to ±2 W m−2.
A series of recent measurement comparisons at the WRC in Davos, the ARM SGP facility in Oklahoma,
and the NREL in Colorado indicate the WISG data are 4.4 to 6.6 W m−2 lower than the average irradiance
262 Solar and Infrared Radiation Measurements

measured by multiple ACP and IRIS systems, depending on the amount of total column WV (Reda et al.
2018a, 2018b). The WMO CIMO has initiated a task team on radiation references to address the needs
for an improved reference measurement standard for downwelling infrared irradiance.

13.4 SPECTRAL CALIBRATIONS


Calibrating radiometers that can resolve the intensity of solar radiation as a function of wavelength
(spectroradiometers) or measure irradiance within a specified wavelength range (photometers and
filter radiometers) introduces a level of complexity that is much greater than is needed for calibrat-
ing broadband pyrheliometers and pyranometers. In fact, it has been noted that “spectroradiometric
measurements are one of the least reliable of all physical measurements” (Kostkowski 1997).
A rather complex calibration chain is used by the NMIs to maintain traceability of spectral radiation
measurements to SI. International implementation of the calibration traceability outlined in Figure 13.1
for spectral solar radiation measurements requires various combinations of high-temperature black-
body sources, Planck’s radiation law, trap detectors, photomultiplier tubes, filter radiometers, spectro-
radiometers, cryogenic radiometers, and carefully controlled 1000 W, quartz-halogen, tungsten coiled
filament (FEL) lamps. Ultimately, spectral radiation measurements are traceable to SI by the unit of
electrical power used for electrical substitution radiometers and the unit of length used to determine
aperture areas and the distance measurements between sources and detectors.
The  unique design elements for spectrally resolving radiometers are presented in Chapter  7.
This chapter presents an overview of the fundamental calibration concepts for radiometers designed to
measure the wavelength distribution of solar radiation in a given wavelength interval or “band-pass.”

13.4.1 tHe measuRement equation


Measuring the amount of solar radiation at specific wavelengths requires a calibration that can
determine the spectral responsivity (Rsλ) of the radiometer:

S
Rsλ = (13.4)
Φ e ,λ

where S is the radiometer output signal (typically a voltage) corresponding to the input spectral
flux (Φe,λ)7 in W m−2 nm−1 at the specified wavelength (λ). The calibration process, therefore, must
include the ability to limit the radiometer detector to specific wavelengths and use a radiation source
of known radiant power at those wavelengths.
Filters selectively transmit radiation according to wavelength by using dyes or interference
coatings. Multiple filters are then required to measure radiation at more than a single wavelength.
To address this limitation, dispersion of solar radiation to spatially separate the different wavelengths
of solar radiation can be done by refraction using a prism or by a diffraction grating. As shown in
Figure 13.8, this approach is used in monochromators for spectral measurements at different wave-
lengths using a single detector or an array of detectors (not shown). Measuring instruments based
on this approach are called spectroradiometers.
Wavelength calibrations of spectroradiometers are based on known atomic emission lines as
determined by quantum mechanics for each element.8 Examples of gas-discharge lamps used for
wavelength calibrations include: mercury argon (253–1,700  nm), krypton (427–893  nm), neon
(540–754  nm), argon (696–1,704  nm), or xenon (916–1,984  nm).9 The  calibration is performed

7 SI units for spectral flux and other radiometric quantities are denoted with suffix “e” for energetic to avoid confusion with
photometric quantities.
8 See NIST Atomic Spectra Database Lines Data: https://physics.nist.gov/PhysRefData/ASD/lines_form.html
9 These wavelength calibration sources are available, e.g., from Ocean Optics, https://oceanoptics.com/product-category/
wavelength-calibration-sources/.
Radiometer Calibrations 263

FIGURE  13.8 Monochromator concept for measuring spectral irradiance from the Sun or an artificial
source. The diffraction grating is positioned to select the wavelength of radiation reaching the detector.

FIGURE 13.9 Emission spectra for mercury-argon source suitable for wavelength calibrations of a spectro-
radiometer. (From NIST Atomic Spectra Database: www.nist.gov/pml/atomic-spectra-database.)

by operating the spectroradiometer to measure a spectrum with the known emission wavelengths
(see  Figure 13.9). The comparison of a measured spectra and known values from the lamp(s) pro-
vides a wavelength calibration for the instrument.

13.4.2 standaRd lamPs


In  the absence of an “artificial sun” for use under controlled conditions, calibrating precision
instruments for measuring spectral solar irradiance continues to rely on standard lamps avail-
able from NMIs such as NIST (Yoon and Gibson 2011), NPL (Harrison et  al. 2000), and PTB
(Sperfeld et al. 2013). Originally designed as a projector lamp, the most common standard lamp
is the 1,000  W tungsten quartz-halogen modified FEL lamp serviceable from 250 to 2,400  nm
264 Solar and Infrared Radiation Measurements

FIGURE  13.10 NIST Standard Lamp type FEL rated at 1,000  W and calibrated over the wavelength
region from 250 to 2,400 nm. (From Yoon, H.W. and Gibson, C.E., NIST Measurement Services: Spectral
Irradiance Calibrations. NIST Special Publication 250-89, National Institute of Standards and Technology,
Gaithersburg, MD, 132 p, 2011.)

(see Figure 13.10). These lamps have modified mounting posts for precision alignment of the fila-
ment at 500 mm from the radiometer measurement plane and have a nominal useful operating life
of about 50 hours. After 50 hours the output of the lamps begins to change. Details of the lamp
calibrations with traceability to SI units using an electrical-substitution cryogenic radiometer can
be found in Yoon and Gibson (2011).
The estimated measurement uncertainty of a standard (reference) lamp depends on the wave-
length of the irradiance. An example of the estimated expanded uncertainties for a NIST-calibrated
FEL lamp is shown in Figure 13.11. The final estimated uncertainties of a radiometer calibration
based on such a lamp would be necessarily larger in order to account for the additional contributions
of the transfer process. Evaluations of the stability, estimated uncertainty, and other characteristics
of these standard lamps can be found in the work by Huang et al. (1998), Kiedron et al. (1999),
Harrison et al. (2000), and Fraser et al. (2007).

13.4.3 langley Plots


Narrow band, multiple wavelength radiometers are called sun photometers or sun radiometers.
These instruments are designed to estimate atmospheric extinction, also known as optical depth, by
measuring DNI in selected wavelengths at the Earth’s surface. As described in Chapter 10, Multi-
Filter Rotating Shadowband Radiometers (MFRSR) are also capable of estimating spectral DNI
values from measurements of global horizontal irradiance (GHI) and diffuse horizontal (DHI) irra-
diance at the same wavelength. Calibration of both types of these radiometers can be done accu-
rately by using the Sun as a source and applying the Bouguer-Lambert-Beer law for the extinction
of optical radiation:
Radiometer Calibrations 265

FIGURE  13.11 Estimated uncertainty at 10 calibration wavelengths for a NIST standard lamp type FEL
1000 W assuming a coverage factor of 2. (From Yoon, H.W and Gibson, C.E., 2011).

V =Vο e − mτ (13.5)

where:
V = Wavelength-dependent radiometer signal voltage
Vο = Instrument-specific extraterrestrial signal voltage
m = Air mass or optical path length
τ = Atmospheric optical depth

Values of air mass (m) depend on the apparent solar position in the sky and is approximately equal to
the secant of the solar zenith angle (e.g., m is 1.5 for a solar zenith angle of about 48.2°). The atmo-
spheric optical depth varies with the wavelength of the incident radiation (e.g., at 550 nm, an aerosol
optical depth (AOD) value of 0.01 represents an extremely clean and cloudless atmosphere and a
value of 0.4 or more corresponds to very hazy conditions. Average monthly values of estimated
AOD range from 0.1 to 0.15 for the USA and exceed 0.8 for parts of Africa and Asia based on satel-
lite observations.10
A simple linear equation results from taking the logarithm of Equation 13.5:

lnV = lnVο − mτ (13.6)

Named after Samuel Pierpont Langley, a Langley plot, is constructed from Equation  13.6 by
applying multiple DNI measurements in a narrow wavelength band at various times of the day

10 Global maps of monthly AOD distributions are available from National Aeronautics and Space Administration Earth
Observatory: https://earthobservatory.nasa.gov/global-maps/MODAL2_M_AER_OD
266 Solar and Infrared Radiation Measurements

(i.e., different solar zenith angles to achieve a range of air mass values) at a given location during
cloudless and stable atmospheric conditions. Stable conditions refer to minimal changes in τ result-
ing from changing atmospheric composition of aerosols, ozone, and other constituents during the
measurement period. For this reason, the Mauna Loa Observatory in Hawaii, the Izaña Observatory
in Tenerife, Spain, and other high-altitude locations are preferred for making measurements for sun
photometers used in ground-based measurement networks to estimate aerosol optical depth (e.g.,
AERONET (Holben et al. 1998), Global Atmospheric Watch- Precision Filter Radiometers (GAW-
PFR) (McArthur et  al. 2003), National Oceanic and Atmospheric Administration/Earth System
Research Laboratory (NOAA/ESRL) (Dutton et al. 1994), and SKYNET (Aoki et al. 2006)).
The  wavelength-dependent calibration factor (Vo) for each channel of the radiometer can be
determined by extrapolation of the calibration measurements as illustrated in Figure 13.12. As pre-
dicted by the Bouguer-Lambert-Beer law, the measurements during the cloudless day depicted in
the figure are quite linear when shown as a Langley plot. The atmospheric composition of aerosols
and other elements important for optical transmission of solar radiation on this particular day and
location were stable as indicated by the nearly overlapping morning and afternoon signals at the
higher air mass values.
As part of the WRC in Davos, Switzerland, the World Optical Depth Research and Calibration
Center (WORCC) calibrates sun photometers used for measuring aerosol optical depth (AOD)
and conducts research to improve the harmonization of data from AOD measurement networks
(Kazadzis 2018). Results from the fourth Filter Radiometer Comparisons (FRC-IV) in 2015
indicate the precision of AOD measurements from 30 instruments based on 6 different designs
is ±0.01. The  basis for international sun photometer measurement comparisons is a triad of
Precision Filter Radiometers (PFR) developed by the WRC and calibrated with the Langley
method at Mauna Loa, Hawaii, USA; Izana, Tenerife, Spain; and Davos, Switzerland (WMO/
GAW 2016).

FIGURE 13.12 Langley plot analysis of clear-sky measurements throughout the day during November 21,
2018, from a PREDE Model POM-2 sun photometer deployed at the Solar Radiation Research Laboratory
(SRRL) in Golden, Colorado, USA. (From https://midcdmz.nrel.gov/srrl_bms/.)
Radiometer Calibrations 267

QUESTIONS
1. Define metrology.
2. What is the estimated measurement uncertainty of the reference measurement standard for
solar irradiance?
3. When is instrument characterization a calibration?
4. Describe three instances when an instrument needs calibration.
5. True or False:
A radiometer calibration provides the instrument-specific responsivity (Rs) specifying the
value for converting the radiometer output signal to irradiance (µV/Wm−2) and its estimated
measurement uncertainty.
6. Describe the WRR and its application.
7. How can a copy of the publicly available BORCAL reports be obtained?
8. What is the Stefan-Boltzmann law and why is it important to radiometry?
9. What is the estimated expanded measurement uncertainty for pyrgeometer calibrations
based on the WISG?
10. Which of these instrument types for measuring IR irradiance not  suited for all-weather
operation: WISG, IRIS, or ACP?
11. Of all physical measurement domains, which is the least reliable?
12. What is the source of wavelength calibrations for spectroradiometers?
13. Name the three NMIs providing reference measurement standards for spectroradiometer
calibration.
14. What parameter can be determined from clear-sky measurements using a sun photometer/
radiometer?

REFERENCES
Andreas, A. M., and S. M. Wilcox. 2016. Radiometer Calibration and Characterization (RCC) User’s Manual:
Windows Version 4.0. National Renewable Energy Laboratory, NREL/TP-3B10-65844. https://www.
nrel.gov/docs/fy16osti/65844.pdf (Accessed June 11, 2018).
Andreas, A., M. Dooraghi, A. Habte, M. Kutchenreiter, I. Reda, and M. Sengupta. 2018. Solar and Infrared
Radiation Station (SIRS) Handbook. U.S. Department of Energy. DOE/SC-ARM/TR-025. http://www.
arm.gov/publications/tech_reports/handbooks/sirs_handbook.pdf (Accessed February 11, 2019).
Aoki, K., T. Hayasaka, T. Takamura, T. Nakajima, S.-W. Kim, and B. J. Sohn. 2006. Aerosol optical charac-
teristics measured by sky radiometers during the ABC/EAREX2005 observation, Korea-Japan-China
Joint Conference on Meteorology, KINTEX Conference Center, Goyang, Korea, October 1–13.
Dutton, E. G., P. Reddy, S. Ryan, and J. J. DeLuisi. 1994. Features and effects of aerosol optical depth observed
at Mauna Loa, Hawaii: 1982–1992. Journal of Geophysical Research 99: 829–8306.
Finsterle, W. 2016. WMO International Pyrheliometer Comparison, IPC-XII, September 28—October 16,
2015, Davos, Switzerland, Final Report. World Meteorological Organization, WMO IOM Report
No.  124. http://library.wmo.int/opac/index.php?lvl=notice_display&id=19643 (Accessed July  13,
2018).
Fraser, G. T., C. E. Gibson, H. W. Yoon, and A. C. Parr. 2007. “Once is enough” in Radiometric Calibrations.
Journal of Research of the National Institute of Standards and Technology 112 (1): 39–51.
Fröhlich, C. 1978. World Radiometric Reference, World Meteorological Organization, Commission for
Instruments and Methods of Observation, Final Report, WMO 490, 108–112.
Fröhlich, C. 1991. History of solar radiometry and the world radiometric reference. Metrologia 28: 111–115.
http://iopscience.iop.org/0026-1394/28/3/001 (Accessed February 12, 2019).
Gröbner, J. 2012. A  transfer standard radiometer for atmospheric longwave irradiance measurements.
Metrologia 49:S105–S111. doi:10.1088/0026-1394/49/2/S105.
Gröbner, J.. and C. Thomann. 2018. Report on the Second International Pyrgeometer Intercomparison,
September 27–October 15, 2015, PMOD/WRC. World Meteorological Organization (WMO), Geneva,
Switzerland: Instruments and Observing Methods (IOM) Report No.  129. https://library.wmo.int/
doc_num.php?explnum_id=4523 (Accessed February 12, 2019).
268 Solar and Infrared Radiation Measurements

Gröbner, J., and S. Wacker. 2015. Pyrgeometer calibration procedure at the PMOD/WRC-IRS. world meteo-
rological organization (WMO) Instrument and Observing Methods Report No. 120. http://library.wmo.
int/pmb_ged/iom_120_en.pdf (Accessed February 12, 2019).
Gröbner, J., I. Reda, S. Wacker, S. Nyeki, K. Behrens, and J. Gorman. 2014. A new absolute reference for
atmospheric longwave irradiance measurements with traceability to SI units. Journal of Geophysical
Research—Atmospheres 119 (12): 7083–7090. doi:10.1002/2014JD021630.
Habte, A., M. Sengupta, A. Andreas, I. Reda, and J. Robinson. 2016. The impact of indoor and outdoor radi-
ometer calibration on solar measurements. National Renewable Energy Laboratory, NREL/CP-5D00-
66659. doi:10.1109/PVSC.2016.7749561 (Accessed June 11, 2018).
Harrison, N. J., E. R. Woolliams, and N. P. Fox. 2000. Evaluation of spectral irradiance transfer stan-
dards. Metrologia 37: 453–456. https://www.researchgate.net/publication/230971557_Evaluation_of_
spectral_irradiance_transfer_standards (Accessed February 12, 2019).
Holben, B. N., T. F. Eck, I. Slutsker, D. Tanré, J. P. Buis, A. Setzer, E. Vermote, et al. 1998. AERONET—A fed-
erated instrument network and data archive for aerosol characterization. Remote Sensing of Environment
66: 1–16.
Huang, L. K., R. P. Cebula, and E. Hilsenrath. 1998. New procedure for interpolating NIST FEL lamp irradi-
ances. Metrologia 35: 81–386.
ISO/IEC. See International Organization for Standardization and International Electrotechnical Commission
International Organization for Standardization and International Electrotechnical Commission. 2017.
ISO/IEC 17025:2017, General Requirements for the Competence of Testing and Calibration Laboratories.
3rd ed. Geneva, Switzerland. https://www.iso.org/standard/66912.html (Accessed February 12, 2019).
International Organization for Standardization. 2015. ISO 9001:2015 Quality Management Systems—
Requirements. Genéva, Switzerland.
International Organization for Standardization. 2018. ISO 9060:2018 Specifications and classification
of instruments for measuring hemispherical solar and direct solar radiation. https://www.iso.org/
standard/67464.html (Accessed February 12, 2019).
Kazadzis, S., N. Kouremeti, S. Nyeki, J. Gröbner, and C. Wehrli. 2018. The World Optical Depth Research and
Calibration Center (WORCC) quality assurance and quality control of GAW-PFR AOD measurements.
Geoscientific Instrumentation Methods and Data Systems 7: 39–53. https://geosci-instrum-method-
data-sys.net/7/39/2018/ (Accessed February 12, 2019).
Kiedron, P. W., J. J. Michalsky, J. L. Berndt, and L. C. Harrison. 1999. Comparison of spectral irradiance
standards used to calibrate shortwave radiometers and spectroradiometers. Applied Optics 38 (12):
2432–2439.
Kostkowski, H.J. 1997. Reliable Spectroradiometry. La Plata, MA: Spectroradiometry Consulting.
Marty, C., R. Philipona, J. Delamere, E. G. Dutton, J. Michalsky, K. Stamnes, R. Storvold, T. Stoffel,
S. A. Clough, and E. J. Mlawer. 2003. Downward longwave irradiance uncertainty under arc-
tic atmospheres: Measurements and modeling. Journal of Geophysical Research 108 (D12): 4358.
doi:10.1029/2002JD002937.
McArthur, L. J. B., D. H. Halliwell, O. J. Niebergall, N. T. O’Neill, J. R. Slusser, and C. Wehrli. 2003. Field
comparison of network sun photometers. Journal of Geophysical Research 108 (D19): 4596.
Myers, D. R., T. L. Stoffel, I. Reda, S. M. Wilcox, and A. M. Andreas. 2002. Recent progress in reducing
the uncertainty in and improving pyranometer calibrations. Transactions of the ASME 124: 44–50.
doi:10.1115/1.1434262
Nyeki, S., S. Wacker, J. Gröbner, W. Finsterle, and M. Wild. 2017. Revising shortwave and longwave radiation
archives in view of possible revisions of the WSF and WISG reference scales: Methods and implica-
tions. Atmospheric Measurement Techniques 10 (8): 9076–9093.
Pavlakis, K. G., D. Hatzidimitriou, C. Matsoukas, E. Drakakis, N. Hatzianastassiou, and I. Verdavas. 2004.
Ten-year global distribution of downwelling longwave radiation. Atmospheric Chemistry and Physics 4:
127–142. doi:10.5194/acp-4-127-2004.
Philipona, R. 2001. Sky-scanning radiometer for absolute measurements of atmospheric longwave radiation.
Applied Optics 40: 2376–2383.
Philipona, R., C. Fröhlich, K. Dehne, J. DeLuisi, J. Augustine, E. Dutton, D. Nelson, et al. 1998. The baseline
surface radiation network pyrgeometer round-robin calibration experiment. Journal of Atmospheric and
Oceanic Technology 15: 687–696.
Philipona, R., E. G. Dutton, T. Stoffel, J. Michalsky, I. Reda, A. Stifter, P. Wendling, et al. 2001. Atmospheric
longwave irradiance uncertainty: Pyrgeometers compared to an absolute sky-scanning radiometer,
atmospheric emitted radiance interferometer, and radiative transfer model calculations. Journal of
Geophysical Research 106 (D22): 28,129-28,141. doi:10.1029/2000JD000196.
Radiometer Calibrations 269

Philipona, R., Fröhlich, C., and Betz, C. 1995. Characterization of pyrgeometers and the accuracy of atmo-
spheric long-wave radiation measurements. Applied Optics 34: 1598–1605.
Reda I., D. Myers, and T. Stoffel. 2008. Uncertainty estimate for the outdoor calibration of solar pyranom-
eters: A metrologist perspective. Journal of Measurement Science, Measure 3 (4): 58–66. doi:10.1080/
19315775.2008.11721448.
Reda, I., and A. Andreas. 2017. Calibration procedure of Hukseflux SR25 to establish the diffuse reference
for the outdoor broadband radiometer calibration. National Renewable Energy Laboratory, NREL/
TP-1900-68999. https://www.nrel.gov/docs/fy17osti/68999.pdf (Accessed February 12, 2019).
Reda, I., J. Gröbner, and D. Turner. 2018a. Recent updates about the absolute cavity pyrgeometer (ACP),
infraRed integrating sphere (IRIS), and atmospheric emitted radiance interferometer (AERI) compari-
sons. Golden, CO: National Renewable Energy Laboratory. NREL/PR-1900-71819. https://www.nrel.
gov/docs/fy18osti/71819.pdf (Accessed February 12, 2019).
Reda, I., J. Zeng, J. Scheuch, L. Hanssen, B. Wilthan, D. Myers, T. Stoffel. 2012. An absolute cavity pyr-
geometer to measure the absolute outdoor longwave irradiance with traceability to international
system of units, SI. Journal of Atmospheric and solar-Terrestrial Physics 77: 132–143. doi:10.1016/
j.jastp.2011.12.011.
Reda, I., M. Dooraghi, A. Andreas, and A. Habte. 2016. NREL pyrheliometer comparisons:
September 26–October  7, 2016 (NPC-2016). NREL/TP-3B10-67311. https://www.nrel.gov/docs/
fy17osti/67311.pdf (Accessed February 12, 2019).
Reda, I., M. Dooraghi, A. Andreas, J. Gröbner, and C. Thomann. 2018b. NREL comparisons between absolute
cavity pyrgeometers and pyrgeometers traceable to world infrared standard group and the InfraRed
Integrating Sphere. National Renewable Energy Laboratory Technical Report NREL/TP-1900-72633.
https://www.nrel.gov/docs/fy19osti/72633.pdf (Accessed February 12, 2019).
Reda, I., T. Stoffel, and D. Myers. 2003. A method to calibrate a solar pyranometer for measuring reference
diffuse irradiance. Solar Energy 74 (2): 103–112. doi:10.1016/S0038-092X(03)00124-5.
Sperfeld, P., S. Pape, and S. Nevas. 2013. The spectral irradiance traceability chain at PTB. AIP Conference
Proceedings 1531: 801–804. doi:10.1063/1.4804891.
Taylor, B. N. 1989. New internationally adopted reference standards of voltage and resistance. Journal of
Research of the National Institute of Standards and Technology 94 (2): 95–103.
Walter, B., R. Winkler, F. Graber, W. Finsterle, N. Fox, V. Li, and W. Schmutz. 2017. Direct solar irradiance
measurements with a cryogenic solar absolute radiometer. AIP Conference Proceedings 1810: 080007.
doi:10.1063/1.4975538.
Wilcox S. M., A. M. Andreas, I. Reda, and D. R. Myers. 2002. Improved methods for broadband outdoor
radiometer calibration (BORCAL). In Proceedings of the Twelfth ARM Science Team Meeting, ed. D.
Carrothers. Richland, WA: U.S. Department of Energy. https://www.arm.gov/news/events/post/50960
(Accessed February 12, 2019).
Wild, M., D. Folini, C. Schär, N. Loeb, E. G. Dutton, and G. König-Langlo. 2013. The global energy balance
from a surface perspective. Climate Dynamics 40 (11–12): 3107–3134. https://link.springer.com/article/
10.1007%2Fs00382-012-1569-8 (Accessed February 12, 2019).
WMO. See World Meteorological Organization.
World Meteorological Organization. 2006. CIMO-XIV, Activity Report, December 7–14, 2006, WMO
No.  1019 (Part II). https://www.wmo.int/pages/prog/www/CIMO/CIMO14-WMO1019/CIMO_14_
Activrep_E.pdf (Accessed November 11, 2018).
World Meteorological Organization. 2017. Guide to Meteorological Instruments and Methods of Observation,
No.  8, 2014  edition updated in 2017. Geneva, Switzerland: WMO. https://library.wmo.int/doc_num.
php?explnum_id=4147 (Accessed February 12, 2019).
WMO/GAW. See World Meteorological Organization-Global Atmospheric Watch.
World Meteorological Organization-Global Atmospheric Watch. 2016. Report No.  231, The  Fourth WMO
Filter Radiometer Comparison. https://library.wmo.int/index.php?lvl=notice_display&id=19806
(Accessed February 12, 2019).
Yoon, H. W., and C. E. Gibson. 2011. NIST Measurement Services: Spectral Irradiance Calibrations. NIST
Special Publication 250-89. Gaithersburg, MD: National Institute of Standards and Technology. https://
www.nist.gov/publications/sp250-spectral-irradiance-calibrations (Accessed November 11, 2018).
14 Ancillary Measurements

But unless we patiently and accurately carry on solar observing throughout this generation,
our successors will be as much at a loss as we are...
Charles Greeley Abbot (1872–1973)

14.1 INTRODUCTION
The World Meteorological Organization (WMO) is working to bring standards to the measurements
affecting weather, climate, and water. In 2014, WMO issued its eighth edition of the Commission
for Instruments and Methods of Observation (CIMO) guide to meteorological measurements
(http://www.wmo.int/pages/prog/www/IMOP/CIMO-Guide.html). The  CIMO guide is written to
encourage standardized meteorological measurements around the world.
This chapter covers the basic elements of the CIMO guide for the measurement of ambient tem-
perature, relative humidity, wind speed and direction, and pressure. In addition, ancillary instru-
ments such as global positioning system (GPS)-based instruments for water vapor measurement,
circumsolar measurement instruments, and sky imagers also will be discussed as they provide
detailed information that aides in solar resource assessment. The advantages and disadvantages of
typical practices using these ancillary instruments are discussed.

14.2 AMBIENT TEMPERATURE


Temperature is related to the average kinetic energy of molecules of a substance. The fundamental
temperature scale is the Kelvin scale based on the average kinetic energy of an “ideal” gas. An
ideal gas is composed of noninteracting point particles and is used because idealized gases can be
described and analyzed with great precision using statistical mechanics. The basic physical proper-
ties of gases are most easily understood using an ideal gas, and most gases behave in a manner very
similar to an ideal gas under most circumstances.
The triple point of water has a defined Kelvin temperature of 273.16 K (Preston-Thomas 1990).
The triple point of a substance is the state in which gas, solid, and liquid forms can coexist in ther-
modynamic equilibrium. A unit Kelvin is 1/273.16 of the thermodynamic temperature of the triple
point of water.
Historically, the centigrade scale is 1/100 the difference between the temperature of melting ice
and boiling water under standard atmospheric pressure (101,325 Pascal). Centigrade is now called the
Celsius scale after Celsius, who worked on developing the scale. The Celsius scale is now defined so
that 0°C = 237.15 K and a change of 1°C is equal to a change of 1 K. The International Temperature
Scale of 1990 also defines the triple point of water as 0.01°C (Preston-Thomas 1990). Note that when
referring to Kelvin, the “°” symbol is not used because it is an SI unit of measure like the meter and
gram. The degree symbol is used for the Fahrenheit and Celsius scales.
As with the precise definitions for the temperature scale, it is necessary to clearly define how
air temperature is measured to have consistency in measurements. Important considerations are
airflow over the sensor, height of the sensor above the ground, the local environment in which the
temperature is measured, and the method of shielding and ventilation of the sensor from direct sun-
light. These details are also important for most meteorological and solar and infrared (IR) radiation
measurements.

271
272 Solar and Infrared Radiation Measurements

14.2.1 tyPes of temPeRatuRe sensoRs


There are two common ways to electrically measure temperature: with a thermocouple or with a resis-
tance temperature detector (RTD), or a thermistor. A thermocouple is made of two dissimilar metals
that generate a voltage proportional to the temperature difference between the junctions of the two
metals and a reference or cold junction. This same device is used for thermopiles in pyranometers,
pyrheliometers, and pyrgeometers. Thermocouples are very versatile, operate over a wide range of
temperature with varying linearity, and have a fast response time (depending on the size and mass) but
also require a measuring device that provides a stable reference and linearization. Thermocouples are
difficult to recalibrate because their response is dependent on the environment in which they are used.
RTD detectors depend on the known change of resistance of the material as a function of tem-
perature. RTDs are stable over time, are more accurate over their given range than thermocouples,
and are easier to recalibrate. RTDs work over a limited temperature range compared with thermo-
couples and can be fragile in harsh industrial environments.
The most common types of thermometers used in automated weather stations are platinum resis-
tance thermometer (PRT) detectors. They are considered the most accurate of thermometers and
operate accurately over most ambient temperature ranges.
Specifications for a good temperature detector include an operating range of –30 to 45°C, with an
accuracy of better than 0.2°C with a deviation of less than 0.1°C over any 10°C range. Initial calibra-
tions should be done at accredited laboratories with periodic performance checks using reference PRTs.

14.2.2 ResPonse times


Air temperature can fluctuate up or down a K or two over a few seconds. However, the reported
value is the average ambient temperature, so it is not necessary to have thermometers that react on
timescales of seconds. The WMO guidelines recommend that a thermometer register 63.2 percent
(or more exactly 1-e –1) of a step change in air temperature in 20 seconds (WMO 2014).

14.2.3 measuRing temPeRatuRe


For  meteorological measurements, ambient air temperature (also called “dry bulb” temperature)
should be measured between 1.25 and 2  meters (m) above the ground. For  the best temperature
measurements, the ground should be level with little obstruction from surrounding objects such as
trees or structures. Temperature measurements on the tops of buildings are affected by the vertical
temperature gradient and the effect of the building itself.
Solar radiation directly from the Sun or the sky dome or reflected from the ground and sur-
rounding objects can heat the temperature sensor and give false readings of ambient air tempera-
ture. Therefore, it is necessary to shield the thermometer from such radiation. Radiation shields
also help protect the detector from rain and fog, which can artificially depress the tempera-
ture. Radiation shields can provide naturally aspirated or forced air measurement environments.
For example, the U.S. Weather Bureau (now the National Weather Service) adopted the Cotton
Region Shelter for air temperature measurements beginning in the nineteenth century (Köppen
1915) (see Figure 14.1). The current Automated Surface Observing System (ASOS) (Figure 14.2)
uses a ventilated instrument housing, or aspirator, to move air across the sensor. Regardless of
the mounting design, the thermometer should be thermally isolated from the radiation shield to
prevent conductive heating or cooling, especially if the thermometer is not ventilated.
Temperature measurements should have a steady airflow over the detector, and the airflow should
be between 2.5 and 10 meters per second (m/s). High-efficiency fan motors should be used because
the motor can slightly warm the air passing over the motor, especially if the flow rate is low enough.
An example of a meteorological tower is shown in Figure 14.3 with the temperature and relative
humidity sensors in a radiation shield and a tipping bucket rain gauge on the side.
Ancillary Measurements 273

FIGURE 14.1 Cotton Region Shelter use for temperature measurements. Temperature and relative humidity
sensors are inside this shelter.

FIGURE  14.2 Image of an ASOS weather station in Milford, Utah. (Courtesy of the Western Regional
Climate Center, Reno, NV.)
274 Solar and Infrared Radiation Measurements

FIGURE 14.3 Picture of a met tower with the temperature and relative humidity sensor in a radiation shield
mounted halfway up the Unistrut® beam. The lightning finial is on the top of the beam. The rain gauge is on
another Unistrut® beam on the right.

As with any measurement it is useful to fully describe the surroundings in which the measure-
ments are made and to record and date any changes that are made to the surroundings.
Information on temperature measurements made with glass or other types of thermometers,
minimum–maximum temperature measurements, and wet bulb measurements can be found in the
WMO Guide to Meteorological Instruments and Methods of Observation (WMO 2014). This guide
is recommended reading before purchasing meteorological equipment and setting up a station.

14.3 WIND SPEED AND WIND DIRECTION


Individual air molecules move at a variety of speeds and directions. We can describe the motion as
a vector in three-dimensional space with the speed assigned to the magnitude of the vector. As the
vector evolves over time, the generalized motion (or flow) of the air is obtained. Often with wind
measurements, we are dealing with surface flow and measurements are limited to wind speed and
Ancillary Measurements 275

wind direction in a plane 10 m above the ground level. A more complete characterization of wind
flow would require a measurement of the wind in the vertical direction as well as in the horizontal
plane. Various wind profilers have been designed based on remote-sensing techniques. Light detec-
tion and ranging (LIDAR) and sound detection and ranging (SODAR) technologies are two such
commercially available technologies (Lundquist et al. 2015).
In addition, there is considerable interest in the gustiness of the wind (i.e., rapid fluctuations).
Therefore, peak wind speed and standard deviations of the wind speed and direction are also useful.
Often for meteorological purposes wind speeds and directions are averaged over 10 minutes and
are used for forecasting. Wind speed measured over a 1-minute interval is often termed a long gust.

14.3.1 sensoR teRminology


The time response of an instrument is the time it takes to detect and indicate a change. More specifi-
cally, given an input step change, the response time to register 63.2 percent of the step change is com-
monly defined as the response time. For anemometers, it is more appropriate to use a representative
response length, or wind run, to characterize the instrument. For example, if the wind speed is 5 m/s
and the response time is 1 s, then the response length would be 5 m. For many anemometers, the
higher the wind speed, the quicker the response time. Therefore, the instrument response to changes in
wind speed is characterized as response length. Response length is approximately the passage of wind
(in meters) required for the output of a wind-speed sensor to indicate 63.2 percent of a step-function
change of the input speed. For most cup and propeller anemometers, the response time for acceleration
is faster than for deceleration. This difference leads to an overestimation of actual average wind speed.
An important topic of wind vanes is the critical damping factor. When there is a change in wind
direction, there is a tendency for the vane to overshoot and oscillate until it is pointing in the right
direction. The  critical damping factor is the damping value for a step change in wind direction
without overshooting the value. The damping ratio is the ratio of the actual damping factor to the
critical damping factor.

14.3.2 anemometeR
There  are two commonly used anemometers: the rotary cup and the propeller type. Pitot tubes
measuring static and dynamic air pressures can be used and produce satisfactory results but are
not commonly deployed. Advanced technologies such as sonic anemometers are new research tools
and are now being incorporated in some monitoring systems for routine measurements.
Both the cup and propeller type anemometers consist of two components: a rotor and a signal
generator. The rotors of the propeller and cup instruments turn at a speed proportional to the wind
speed. The responsivity of the instruments to changes in wind speed is characterized by a response
length that is related to the moment of inertia of the rotor and other geometrical considerations
(Busch and Kristensen 1976; Coppin 1982).

14.3.3 CuP anemometeRs


The first anemometer on record was invented by Leon Battista Alberti, an Italian architect, in 1450.
It consisted of a disk placed perpendicular to the airflow, and its rate of rotation depended on the wind
speed. The hemispherical cup anemometer was invented in 1846 by Irish astronomer John Thomas
Romney Robinson. The original cup anemometer had four cups, but work by Patterson (1926) showed
that three-cup anemometers responded faster to changes in wind speed. Patterson’s work also showed
that the larger the ratio of the cup radius to the arm length the better the linearity of the response.
The response to increases in wind speed is typically faster than the response to decreases in
wind speed, and this asymmetry in response results in overestimates of the wind speed (Kristensen
1993, 1998, 1999). In addition, the cup instruments are sensitive to the vertical motion of the air
276 Solar and Infrared Radiation Measurements

while the propeller instruments are insensitive to the vertical component. The  overestimate of
the average wind speed can be as much as 10 percent depending on the anemometer type and
characteristics.
The rate of rotation of the rotor is roughly the difference between the wind speed and the initial
wind speed necessary to rotate the rotor divided by the length of the column of air necessary to cause
the rotor to rotate 1 radian. Therefore, the wind speed (ws) is obtained by multiplying the rotation
frequency (rf) times a gain (g) plus the initial wind speed (inws) necessary to start the rotation (usu-
ally around 0.2 m/s). The gain is the length of the column of air necessary to cause the rotor to rotate
1 radian.

ws ≈ rf ⋅ g + inws (14.1)

The friction that prevents the anemometer from rotating at very low wind speeds introduces a non-
linear response. Friction also varies with temperature. In practice, the anemometer bearings have
a burn-in period before the instrument runs at a steady rate. This procedure is sometimes done at
the factory. After a year or two of use, the bearing friction starts to affect the anemometer and the
bearings should be replaced. Cup anemometers also have a slight dependence on temperature and
pressure. A thorough discussion of cup anemometers can be found in the IEA Wind Task 11 report
on wind speed measurements and cup anemometers (IEA 1999).
An example of a cup anemometer and a wind vane are shown in Figure 14.4.

FIGURE 14.4 A cup anemometer and a wind vane are mounted from a pole several feet from the railing of
a deck to provide ancillary measurements for outdoor radiometer calibrations.
Ancillary Measurements 277

14.3.4 PRoPelleR anemometeRs


H. W. Dines first published information on the propeller anemometer in 1887. In the early 1940s, the
Bendix Friez Company developed the “Aerovane” anemometer under contract with the U.S. Navy.
Modern helicoid (propeller) anemometers are based on the Aerovane design (propeller anemom-
eters are shown in Figure 14.5). The helicoid propeller is mounted on a rod with a vane at one end to
pivot the propeller into the wind. Above the wind speed that overcomes the friction of the bearing,
the rate of rotation of the propeller is proportional to the wind speed.
The propeller responds mainly to the airflow parallel to the axis of rotation. By mounting three
propeller anemometers in an orthogonal array, the three-dimensional airflow can be measured.
These instruments have been used to study turbulence. To achieve accurate results, corrections must
be applied to account for the cosine response of the instrument with respect to wind direction and
the interaction of the wind with the support arms.
Disadvantages of propeller anemometers are that they need to point into the wind and there
is a lag time, or they may overshoot the direction. The propeller is not always pointed directly
into the direction of wind flow; this is particularly a problem with low wind speed or if there
are eddies in the wind flow. Airflow that is perpendicular to the axis of rotation can also cause
problems.
For fixed-propeller anemometers, response decreases as the angle from the direction of airflow
increases. For example, when the angle reaches 85 degrees from the direction of airflow, the amount
of air needed to push the propeller through 1 radian increases by a factor of 3.
In theory, the instrument should not require calibration except at low wind speeds. In practice,
calibrations should be performed to obtain accurate measurements at low wind speeds and to check
for excessive drag as the bearings begin to wear. Overspeed problems should be less than with cup
anemometers because the propeller anemometers are pointed into the wind.

FIGURE 14.5 Wind anemometers—sonic and vane. Pacific Northwest National Laboratory has developed
a special filtering technique to enhance data gathered by an inexpensive cup anemometer. With the technique,
cup anemometer measurements are surprisingly close to those made with a costly and sophisticated sonic
anemometer (center of photo). (Courtesy of Warren Gretz, NREL staff photographer.)
278 Solar and Infrared Radiation Measurements

14.3.5 soniC anemometeRs


Sonic waves travel at different speeds depending on the direction of the path with respect to the
flow of the wind. Measuring the time for the pulse to travel between the source and a receiver can
then be used to calculate the wind speed. A sonic anemometer is shown in Figure 14.5. The typical
frequency of a sonic anemometer is about 100 kilohertz (kHz).

Wx = (∆ t ⋅ kRT )/ 2d (14.2)

where the wind speed in the “x” direction (W x) is equal to the time it takes to travel the distance
(∆t) times the velocity of sound squared (or equivalently kRT, where k is the ratio of specific heats,
R is the gas constant, and T is the air temperature) divided by twice the path length (d). Some sonic
anemometers can measure with a resolution as low as 0.005 m/s.
Sonic anemometers are ideally suited for measuring the structure of turbulence because they
have high resolution and precision. However, the presence of precipitation can affect the operation
of the instrument. In addition, the geometry of the sensing heads introduces distortions that can
affect wind speed measurements.

14.3.6 installing anemometeRs


When mounting anemometers on a meteorological tower, it is important to minimize distortions
caused by the tower and cabling. It is best to locate the anemometer at the top of the tower. Any
lightning rod should be far enough away from the anemometer to minimize its effect. The loca-
tion of the meteorological tower also should avoid local obstructions. The standard meteorological
measurement exposure for wind speed and direction is 10 m above the ground over level, open ter-
rain (WMO 2014). Because wakes from trees and buildings can easily extend downwind to 12 or
15 times the obstacles height, a properly located measurement system should be at least 10 obstruc-
tion heights away for proper “wind fetch.”
For  boom-mounted anemometers, the anemometer should be mounted above the cabling.
For cylindrical wind towers, the boom supporting the anemometer should be at least 6 diameters
from the support structure. If the wind comes from a prevalent direction, the minimum distortion
occurs when the boom is mounted 45 degrees to the prevailing direction of the wind.
For a lattice mast, the boom should be mounted perpendicular to the prevalent wind direction. To
reduce distortion of the measurements, the anemometer should be mounted 12 to 15 times the boom
diameter above the boom (Hunter 2003). To reduce distortions, cylindrical booms should be used.
In addition, angular structures should be avoided along with untidy cabling. A good distance also should
be kept from tower guys and lightning finials. An example of a wind tower is shown in Figure 14.6.
The tower is a narrow pole with wind measurements at two heights. The tower is well guyed for stabil-
ity, and a lightning finial is at the top of the pole and away from the wind sensors. It is important to
firmly guy the wind tower to the ground. The energy in a gust of wind is proportional to the cube of the
wind speed. Even wind gusts of 25 m/s (roughly 60 miles per hour) carry a lot of energy.
Note that some anemometers can freeze in position during cold periods, especially if there is
little airflow. This happens especially if there is freezing rain. Some anemometers have heaters to
prevent freezing. An additional source of information on wind monitoring is Schreck et al. (2008).

14.3.7 Wind vanes


Wind vanes have been around since at least 450 BCE and probably decorated homes and public
buildings much earlier. Besides providing decorative art to the building, wind vanes let people know
which way the wind blows. This important information is added to anemometer measurements, and
wind direction measurements often are taken alongside wind-speed measurements.
Ancillary Measurements 279

FIGURE 14.6 A wind tower with guys. Wind sensors are located at two levels with guy wires attached below
the wind sensors. The tower is a pole with the wires running to the wind sensors securely attached to the pole.
A lightning finial is mounted on top of the tower.

Wind direction is measured in a clockwise direction from true north (geographic north, not local
magnetic north). It is very important to align the wind vane due north when it is installed. Any
alignment error will be reflected in the measurement. When wind speeds are less than 1 knot or
0.514 ms−1, calm is reported, and direction is often given as 0. The abbreviation used for knot is kt
(kn is used in educational and naval applications and kt is used in technology, aviation, and military
and in the WMO/CMIO guide (2014)). Typically, it takes a wind speed of about 1 kt flowing at a
45 degree angle to move the wind vane. This will vary by model. Accuracies are typically reported
for steady winds above a certain speed, such as 5 ms−1. With fluctuating wind speeds and directions,
the wind vane can overshoot an angle and oscillate about the true direction. These oscillations are
usually damped exponentially.
There are many different ways to sense the direction the vane is pointing, ranging from poten-
tiometers to rotary switches. Wind vanes with potentiometers will be discussed here. Wind vanes
that use a potentiometer require a voltage to function. About one-half the excitation voltage should
be recorded when the vane is pointed at 180 degrees.
280 Solar and Infrared Radiation Measurements

The  shaft of the wind vane should be vertical. Deviations by more than a degree or two
will impart systematic errors into the readings. Resistors, hence potentiometers, are dependent
on temperature, and this temperature dependence can affect readings in extreme temperatures.
Like the anemometer, a heater is sometimes an additional component that can be purchased for
the wind vane. The heater should be powered by a separate unit from the one that powers the
potentiometer.
It is useful to calibrate the wind vane periodically to test for problems. Measurements can often
be started at 180 degrees and then subsequent measurements at 15 degree increments. Calibration
results that are outside of the manufacturer’s specifications require replacement. The manufacturer
should be contacted to make sure the component is replaced correctly.
Wind speed and wind direction information are often plotted in a figure  called a wind rose
(see Figure 14.7). The wind rose is a polar plot with the wind direction going from 0 to 360 degrees.

FIGURE 14.7 Example of a wind rose plot from the NREL Measurement and Instrumentation Center web-
site. Data are for all of 1989 for an anemometer 10 m above the ground. Wind rose plots can represent wind
data from days to years, depending on the time period selected. The wind directions are divided into eight
groups, one group every 45 degrees. Each circle represents 3.7 percent of the data. The different segments
represent different wind speed regimes. For example, the plot shows that approximately 3 percent of the time,
the wind blowing from the southwest is between 4 and 6 m/s.
Ancillary Measurements 281

The distance from the center of the plot is the frequency that the wind is from a given direction.
Often the length of the frequency is divided into wind-speed ranges.

14.4 RELATIVE HUMIDITY


The water vapor content of air can vary considerably over the day and affects many aspects of our
life from comfort level to the performance of electronic devices. There are several ways to measure
the water vapor content of our atmosphere, including most commonly the dew-point temperature
and relative humidity measurements. Relative humidity, U, is the percent ratio of the observed vapor
pressure at a given location to the saturated vapor pressure of water at the temperature and pressure
at the time of observation.
A  chilled-mirror hygrometer is a field and reference instrument used to measure dew-point
temperature. The  hygrometer works by cooling a surface until dew or ice forms on the sur-
face. The chilled-mirror hygrometer has a solid, non-hygroscopic surface, which does not absorb
or release water vapor. The  dew-point temperature (Td ), or frost-point temperature (Tf ), if it is
cold enough, is the temperature at which dew, or ice, forms on the surface. Most chilled-mirror
hygrometers use a polished metal mirror that is heated electronically by a Peltier-effect heater.
An optical detector determines when moisture starts to condense on the mirror, and the tempera-
ture of the surface is then recorded. A pressure measurement is needed to calculate the relative
humidity.
Still in use today, the sling psychrometer is used to measure relative humidity. The  sling psy-
chrometer measures both the ambient temperature and dew-point temperature, and the relative
humidity can be determined using psychrometric charts. Many hygrometers are used to measure
relative humidity such as the hair hygrometer, the chilled-mirror dew-point hygrometer, lithium chlo-
ride heated condensation hygrometer (dew cell), hygrometers using absorption of electromagnetic
radiation, and electrical resistive and capacitive hygrometers. Only the electrical resistive and capaci-
tive hygrometers will be discussed in this book because they lend themselves to automated measure-
ments. For information on the other methods of measuring relative humidity see the CIMO guide
(WMO 2014).
Electrical resistance hygrometers have a plastic surface with a conductive surface on top of a
nonconductive substrate. Water molecules adhere to the surface, changing the conductivity. A water-
attracting (hygroscopic) electrolyte is used, and the conductivity is a function of the amount of water
adsorbed. Alternating current is used to avoid polarizing the electrolyte.
Hygrometers based on electrical capacitance use solid hygroscopic material and measure the
dielectric properties of the material. Water binds to the polymer, and the large dipole moment
of water alters the dielectric property of the material. A capacitor is created by sandwiching the
polymer foil between two electrodes. The  capacitance is usually small, on the order of a few
hundred picofarads. Therefore, it is necessary to have the electrical interface near the probe to
minimize the electrical and capacitive effects. The effect of temperature on the probe must also
be considered.
Electrical resistive and capacitive hygrometers usually are packaged with PRT temperature sen-
sors. Enclosures that block the Sun also prevent rain from wetting the probes. These probes may be
damaged if they come in contact with liquid water.
The range of measurement for these probes can vary from 1 to 100 percent. These hygrometers
have a temperature dependence of around 0.05%/°C. The  response time is measured is seconds
(roughly 10–20 seconds) and depends on the temperature. Typically, the accuracy of these probes in
the field is 2 to 3 percent with the larger uncertainty above 90 percent. Stability of these hygrometers
is about 1 percent per year. Of course, manufacturer specifications and calibration certificate should
be consulted for the stated measurement performance of the instrument.
The response of these instruments to water vapor is nonlinear, and formulas are often supplied by
the manufacturer that render the measured values into percent relative humidity. Relative humidity
282 Solar and Infrared Radiation Measurements

probes should be calibrated annually, either at the factory or by use of salt solutions that have known
water vapor pressures. A  good hair hygrometer can be used to spot check the calibration of these
hygrometers.

14.5 ATMOSPHERIC WATER VAPOR


The  total amount of water in the sky above a surface observation can be estimated from rela-
tive humidity measurements (Garrison and Alder 1990). However, estimates from ground-based
relative humidity measurements make broad assumptions of the water vapor content higher in
the atmosphere. At selected locations, weather balloons are launched periodically to measure the
water vapor content of the upper atmosphere. Water vapor content is important for many reasons.
For example, it is used for weather forecasting and for modeling solar radiation. These atmospheric
water vapor measurements, relative humidity measurements at different altitudes, are part of the
radiosonde instrument package that is carried by balloon to measure various atmospheric param-
eters and use battery-powered telemetry to transmit the measurements to ground stations. A discus-
sion on the accuracy of water vapor measurements using various radiosonde packages can be found
in Miloshevich et al. (2006).
With the large number of GPS satellites now circling the Earth, it has become possible to deter-
mine atmospheric water vapor by comparing the GPS signals from several satellites at the time.
GPS satellites emit signals that enable devices that receive these signals to precisely calculate their
position. There are at least 24 Navstar GPS satellites orbiting the Earth with more planned. Other
countries also have similar networks such as Russia with its GLONASS system and Europe has the
Galileo GPS satellites. Improvements continued to be made, and more uses for these systems are
being exploited.
Around 2000, researchers starting looking into accurately monitoring atmospheric water vapor
using GPS satellites (Keihm 2001). Typically, 4 to 10 of these GPS satellites are accessible at any
one time. These satellites use atomic clocks to regulate the frequency of signals sent from the satel-
lite. A minimum of signals from four GPS satellites is needed to compute the three-dimensional
location of the station and the time offset between the satellite and the receiver clock. Typically, two
different frequencies are generated by the satellites. Ground-based antennas pick up these signals
and receivers using reference satellite clocks match the signal from the satellite with identical sig-
nals they produce. The signals generated in the satellite contain patterns and by matching the pat-
tern in the satellite signals with the signals produced on the ground, the phase difference between
when the signal was generated and when it was received is obtained.
The distance to the satellite from the ground antenna is roughly equal to the time it takes the
signal to reach the ground observer times the speed of light. Because the light travels through the
atmosphere, adjustments are necessary to account for the effects of the atmosphere on the speed
of light. The ionosphere consists mainly of ions and ranges from 50 to 500 miles above the earth.
Models exist that can account for the slowing of the signal as it passes through the atmosphere, but
there is still some uncertainty in the time it takes for light passing through the ionosphere.
Next the signals pass through the troposphere. Important factors that affect the time it takes
for the signal to pass through the troposphere are temperature, pressure, and water vapor content.
Models exist to estimate the effects of the atmospheric constituents and these models are more
accurate than the models used to estimate the delay in the signal through the ionosphere. The angle
between the observer and the satellite also must also be considered because the greater the angle,
the more atmosphere is traversed before the signal is received.
There  are two ways that these uncertainties can be reduced. The  first is to use two different
wavelengths. In general, the longer the wavelength of the signal the more it is slowed by the atmo-
sphere. Therefore, the use of dual frequencies can be used to better determine the characteristics
of the atmosphere through which the signal passes, and this reduces the uncertainty of the distance
estimate.
Ancillary Measurements 283

The uncertainties also can be reduced by using signals from more than one GPS satellite. It takes
signals from four satellites to get a good estimate of the distance and location relative to the satellite
and water vapor content. The more signals analyzed, the more representative is the estimate of the
water vapor content of the atmosphere.

14.5.1  using gPs satellites to measuRe PReCiPitable WateR vaPoR


A GPS antenna attached to a receiver can be used to measure the precipitable water vapor. Because
water vapor slows the GPS signal speed, measurements of this delay can be used to calculate the
precipitable water vapor in the atmosphere. If two signals are available, a differential or dual signal
methodology can be used. The phase of this signal changes as the distance between the ground station
and satellite changes as the satellite passes across the sky. The position of the satellite with respect to
the ground station can be determined. The amount of atmosphere between the satellite and the ground
station is known by the relative positions and the distance between the satellite and ground station.
Using atmospheric models with information on the atmospheric temperature and pressure, the amount
of water vapor along the path of the signal can be calculated. The percent uncertainty in the distance
between the satellite and the ground station is given in Table 14.1. The multipath uncertainty in the
table refers to reflected signals that also reach the antenna so that there are multiple paths by which
the signal can reach the antenna. The improved accuracy using differential signals is a significant
improvement over estimates with just a single signal as shown in the comparison in Table 14.1.

14.5.2  installing a global Positioning system antenna


The  antenna receiving GPS signals should be mounted with as clear a field of view as possible.
The goal is to receive signals from as many satellites as possible at any given time. Obstructions
above the horizon will block signals from satellites in that portion of the sky. The  antenna also
should be mounted away from surfaces that reflect the signal. For  example, the antenna should
not be mounted on a metal roof.
Although some preliminary analysis can be done by the receiver, the information is usually fed
to a central location where more thorough analysis can be performed. For example, the University
Corporation for Atmospheric Research (UCAR) office in Boulder, Colorado, takes signals from
many locations in the United States, analyses the data, and makes half-hourly water vapor data
available. The types of data available along with the precipitable water vapor (PWV) are given
in Table 14.2. The data in Table 14.2 were obtained using a Vaisala PTU300 pressure/humidity/
temperature transmitter, with the ASCII output string sent, via RS-232, to the Trimble Net9R
every 60 seconds. A pressure measurement is required for the PWV calculations. The locations
with PWV measurements generated at UCAR are available at https://www.suominet.ucar.edu/.
Useful information on PWV measurements is at https://www.suominet.ucar.edu/overview.html

TABLE 14.1
Estimated GPS Distance Uncertainties by Source (%)
Standard Differential GPS
Satellite clocks 1.5 0
Orbit errors 2.5 0
Ionosphere 5.0 0.4
Troposphere 0.5 0.2
Receiver noise 0.3 0.3
Multipath 0.6 0.6
284 Solar and Infrared Radiation Measurements

TABLE 14.2
Data Available Using a Vaisala
Water Vapor Transmitter
Measurement Unit
Precipitable water mm
Precipitable water (uncertainty) mm
Air temperature °C
Relative humidity %
Surface pressure mbar

14.6 PRESSURE
Atmospheric pressure is the force per unit area on a surface exerted by the weight of the atmo-
sphere above the surface. The common unit of pressure is the Pascal (Pa) or Newton per square
meter (Nm−2). Typically, hectopascals are the units used for pressure where 1 hPa = 100 Pa and
is equivalent to 1 millibar (mbar), the old scale of measurement. The value of the standard atmo-
spheric pressure is 1013.250 hPa and is equivalent to 760 millimeters (mm) of mercury (Hg).
The invention of the Hg barometer has been credited to Evangelista Torticelli in 1643, although
there were discussions of similar barometers around the same time or earlier. The Hg barometer
compares the pressure of a column of Hg against air pressure. The height of the column of Hg
is proportional to the atmospheric pressure. Temperature, gravity, and the meniscus of Hg in
the glass tube are among the most prevalent sources of error that can affect the reading of a Hg
barometer.
Several types of electronic barometers are used in automated weather stations. These barometers
use materials that respond to pressure and use transducers to translate these changes into electrical
signals. Some of these sensors are triply redundant to improve accuracy and have temperature sen-
sors mounted internally to provide automatic temperature compensation.

14.6.1 aneRoid disPlaCement tRansduCeRs


Aneroid displacement transducers use materials that expand or contract in response to changes in
pressure. Typically, a metal alloy of beryllium (Be) and copper (Cu) are used for the instrument and
expansion or contraction is measured by levers that are connected to a device to display the pressure.
Alternatively, designs can be created to generate electronic output.

14.6.2 PieZoResistive baRometeRs


Piezoresistive pressure sensors work by running a current through a crystal that changes resistance
resulting from deformations in the structure. The current then passes through temperature compen-
sation circuitry and is amplified. Piezoelectricity is the charge that occurs when pressure or strain
is applied to certain solid materials. Piezoresistivity and piezocapacitance are the changes in resis-
tivity and capacitance that occur when strain is applied. Piezoresistive barometers use crystalline
quartz as the sensor material because it has well-defined and stable piezoelectric properties. Strain
changes interatomic spacing, and this affects the bandgap, making it easier or harder for electrons
to be raised into the conduction band. The strain results in a change in resistivity of the semiconduc-
tor that is proportional to the strain being applied. A frequency control circuit is connected to the
Ancillary Measurements 285

piezoelectric oscillator, and alternating current (AC) resistance is proportional to the pressure being
applied. Again, crystalline quartz is usually chosen because frequency characteristics are precisely
reproducible and the dependence on temperature is relatively small.
The gauge factor, the change in resistance per change in length, is typically about 2 for metals,
10–35 for polycrystalline silicon (Si), and 50–150 for single crystalline Si (Schubert et  al. 1987;
Smith 1954; Tufte and Stelzer 1963).
Pressure at an inlet port moves a flexible bellows upward and applies an upward axial force
that creates a compressive force on the quartz crystal. The quartz crystal diaphragm is very rigid,
and the pressure creates minimal deflections. Strain gauges are attached around the circumference
of the diaphragm or diffused into the Si. A Si chip with diffused strain gauges can act as the dia-
phragm itself.
A Si wafer can be fabricated to serve as the pressure sensor (Singh et al. 2002). A thick Si rim
supports a monocrystalline Si diaphragm. Etching away the bulk of the Si creates the diaphragm.
Two to four piezoresistors are then diffused into the diaphragm. The Si is doped with n-type atoms
that provide an extra conduction electron to the Si. Diffusion of p-type atoms, atoms that seek an
electron to complete a pair bond, is applied to small squares to create piezoresistors. The piezoresis-
tors are arranged in the Wheatstone bridge configuration to achieve higher voltage sensitivity and
low temperature sensitivity.
A generic Wheatstone bridge (Figure 14.8) has three known resistances and one unknown. When
the bridge is balanced, there is no current flow from point B to point D in Figure 14.8. When the
bridge is unbalanced by the change in resistance of the strain gages, a current is created and this cur-
rent is proportional to the change in resistance and, hence, pressure. Some piezoresistive barometers
have thermostats to compensate for temperature effects.
Modern piezoresistive barometers measure two resonance frequencies of the piezoelectric
element. By calculating a linear function of these frequencies, and with an appropriate set of
variables obtained after calibration, a pressure is calculated by a microprocessor that is inde-
pendent of the temperature of the sensor. Accuracies of piezoresistance barometers can reach
0.01 percent.

FIGURE 14.8 Generic drawing of a Wheatstone bridge used to determine an unknown resistance. A gauge,
GV, measures the voltage between points D and B. Resistor R X is unknown resistance.
286 Solar and Infrared Radiation Measurements

14.6.3 PieZoCaPaCitanCe baRometeRs


In  a piezocapacitance barometer, a metal or Si diaphragm acts as one terminal of the capacitor.
The other terminal is a metal substrate on a ceramic or glass substrate. Pressure deflects the dia-
phragm, which in turn changes the gap spacing and the capacitance. In a differential piezocapaci-
tance barometer, the Si diaphragm is between the two electrodes and the pressure deforms the
diaphragm and increases the capacitance on one side and decreases the capacitance on the other.
Piezocapacitance barometers can have an accuracy of 0.1 percent.

14.7 SKY-IMAGING SYSTEMS


Sky-imaging systems are a relatively new development in meteorological instrumentation.
Hemispherical (whole sky) imagery from these ground-based automated systems provide details
of the local horizon, sky conditions, and other information helpful for understanding solar and IR
radiation measurements; they can be used to conduct site surveys of solar access and to develop
short-term solar forecasts supporting the operation of utility-scale solar energy conversion power
plants, as well as other research applications in meteorology and climatology.
Solar access or horizon surveys characterize site-specific shading and possible reflections from
fixed objects above the local horizon affecting irradiance measurements and solar collector perfor-
mance during the year. Portable sky-imaging systems can provide the necessary information for
properly siting resource measurement equipment and solar energy conversion systems.
Observing sky conditions, especially clouds, has been a long-standing meteorological practice.
Historically, cloud cover has been reported by trained weather observers who estimated the amount of
cloud cover, the types of clouds, and cloud layers once an hour. Using national and international stan-
dards for reporting, meteorological archives now contain decades of these hourly surface observations
made by national meteorological networks and inferred from satellite-based measurements (Rossow
1996; Smith et al. 2011). Cloud amounts are now recorded in oktas, where 1 okta = 1/8th of the sky
dome. Because clouds strongly effect the amount and distribution of solar and IR irradiance reaching
the surface, historical cloud observations were used in the early development of models for estimating
solar irradiance in the absence of radiometric measurements (Badescu 2008; Myers 2013).
The recent increased deployment of photovoltaic and solar thermal power systems has created
the need for nearly real-time cloud information from automated sky imagery to support the devel-
opment of operational solar forecasting for electric utilities (Kleissl 2013). The  increased levels
of information available from high-resolution sky imagery also have benefited climate research
(Kassianov et al. 2017; Long et al. 2001).
This section provides an overview of sky imager design concepts and presents some examples of
data applications. Sky imagers designed for site surveys will be briefly discussed first followed by
a more comprehensive description of sky imagers used to assess the solar resource. Detailed treat-
ments can be found in the references cited.

14.7.1 site suRveys


Performing a site survey to document the local horizon for fixed shading obstacles is among the
first steps for selecting a site for a solar measurement station or solar power conversion system. We
introduce a sample of commercially available instruments here for the convenience of the reader.
Portable instruments with a fisheye (hemispheric) field of view were among the first sky imag-
ers. Originally developed in the 1970s to provide rapid field assessments, the Solar Pathfinder™
continues to offer solar installers a portable tool and software for evaluating the solar access at a
specific location (Figure 14.9). To meet the demands of a growing solar industry in the last decade,
new approaches to field measurements of the sky and local horizon have been commercialized.
Solmetric produces their SunEye 210 fisheye digital camera system for estimating the amount
of shading from surrounding obstructions (Figure  14.10). Step Robotics developed Step Solar,
Ancillary Measurements 287

FIGURE 14.9 The Solar Pathfinder™ is one of the original portable tools for evaluating solar access in the
field. (From https://www.solarpathfinder.com.)

FIGURE 14.10 The SunEye by Solmetric is a portable shade measurement tool for determining solar access
potentials. (From http://www.solmetric.com/buy210.html.)
288 Solar and Infrared Radiation Measurements

a fisheye lens attachment for a smartphone and application software to produce shade analyses
(Figure 14.11). The Horicatcher from Meteotest consists of a hemispherical mirror, digital camera,
and a mounting device with a sun shade disk that are attached to a regular photographic tripod
(Figure 14.12). Each device has documentation and analysis software associated with the instru-
ment. For example, the associated analysis software for the Horicatcher is part of the Meteonorm
solar resource data and calculation tools (http://www.meteonorm.com/en/product/productpage/
mn-produkt-horicatcher).

FIGURE 14.11 The Step Solar by Step Robotics fisheye lens attaches to a smartphone to capture sky imag-
ery and produce shade analyses for solar applications. (From http://steprobotics.com/overview/.)

FIGURE 14.12 The Horicatcher from Meteotest has a camera and sun shade mounted above a hemispheri-
cal mirror for capturing images of the sky and local horizon. (From http://www.meteonorm.com/en/product/
productpage/mn-produkt-horicatcher.)
Ancillary Measurements 289

14.7.2 sky Conditions


Clouds are a major influence on the amount and distribution of solar radiation reaching the Earth’s sur-
face. Cloud types, heights, amounts, and relative locations in the sky dome are important pieces of mete-
orological information for interpreting and characterizing solar and infrared radiation measurements.
For example, one of the more difficult problems with measuring and modeling diffuse solar irradi-
ance is accounting for the anisotropic distribution of sky radiance depending on the changing solar
position and sky condition. Any given measurement of diffuse horizontal irradiance (DHI) can result
from a wide variety of these conditions. Under cloudless skies, DHI can be reliably partitioned into a
circumsolar component resulting from forward scattering by the atmosphere near the solar disk, hori-
zon brightening due to multiple scattering, and radiance emanating from the rest of the sky (Perez
et al. 1987). When clouds are present, this model breaks down because clouds in varying amounts
and types can be in different areas of the sky. Therefore, simply knowing the observed cloud cover
amount or amounts requires a statistical approach to modeling DHI (Maxwell 1998).
All-weather sky imagers have been developed to provide regular measurements of sky condi-
tions needed for improved solar resource modeling, including short-term forecasting, and ancillary
information for characterizing solar and IR measurements.
Early sky imager designs utilized a shading device that blocked the direct sunlight to reduce the
dynamic range of light incident on the camera as light from around the solar disc can be orders of mag-
nitude brighter than that from other areas of similar dimensions in the sky. Figure 14.13 illustrates the
installation of one of these systems and a graphical display of the image and resulting cloud analyses.
More recently, sky-imaging systems rely on improved digital camera detectors, wide-angle lenses
oriented skyward, and enclosures with temperature control and forced ventilation. Image analysis
software also has been improved by determining the blue/red and blue/green ratios to detect a wider
range of cloud optical depths, especially for very thin clouds (Kleissl 2013). The ASI-16 All-Sky
Imager from EKO Instruments B.V. is shown in Figure 14.14 with typical cloud analyses from a sta-
tion in Golden, Colorado, USA. Before evaluating the image for cloud amounts and positions, the
software first maps the image into a projection correcting the distortions created by the fish-eye lens.

FIGURE  14.13 A total sky imager, model TSI-880 from Yankee Environmental Systems installed at the
NREL Solar Radiation Research Laboratory (SRRL) and sample data product available from the Measurement
and Instrumentation Data Center. (From https://midcdmz.nrel.gov/srrl_bms/.) (Courtesy of Tom Stoffel.)
290 Solar and Infrared Radiation Measurements

FIGURE 14.14 The EKO Instruments B.V. Model ASI-16 All-Sky Imager has been recently installed at the
NREL SRRL for monitoring cloud conditions. Imagery and cloud analyses are available at 10-minute inter-
vals during daylight periods.

The cloud analyses are then based on evaluations using the blue/red and blue/green ratio of the image
pixels. A unique cloud detection and opacity classification algorithm determines the percentage of
“thick” and “thin” clouds when present.
Currently, sky imagers are being tested for forecasting the solar irradiance incident on a field of
solar collectors. A typical sky camera consists of a digital camera fitted with a fish-eye lens attached
to a data logger, a power supply, and a communication device. The device is enclosed in a weather-
proof case with a glass dome, effectively separating the instruments from the elements (Figure 14.15).
The interior of the sky camera built by the University of California at San Diego (UCSD) is shown in

FIGURE 14.15 Sky imager being tested at a University of California at San Diego photovoltaic facility.
Ancillary Measurements 291

FIGURE 14.16 Internal view of a sky imager. The sky imager is enclosed in an all-weather box. (Courtesy
of Jan Kleissl and Ben Kurtz of the University of California at San Diego, CA.)

Figure 14.16 (Urquhart et al. 2014, 2016). By obtaining a series of sky images and using automated
analysis tools, it is possible to estimate the movement of clouds and predict the amount of shading
that will occur in the next 15 minutes (Figure 14.17, Nguyen et al. 2016, Kurtz and Kleissl 2017).
The images must be automatically processed, and results made available to plant operators (Kurtz
et al. 2017). With possible multilayer clouds and the amorphous nature of clouds, prediction of cloud
movement, cloud genesis, and changing cloud structure will require more field experience (Nguyen
and Kleissl 2014, Yang et al. 2014, Wang et al. 2016). Forecasting using sky imagers is a relative new
field and modeling is expected to improve with continued testing and operational experience.

FIGURE 14.17 Images of the sky dome taken with a sky imager. The figures illustrate the variety of cloud
patterns that can exist and the need for computer programs that can recognize different cloud patterns to map
cloud motion.
292 Solar and Infrared Radiation Measurements

Another use of sky imagers is to estimate the aerosol content of the atmosphere (Kazantzidis
et al. 2017). Good estimates of the aerosol content of the atmosphere are scarce. Initial tests using
the digital data from sky imagers show that a fair estimate of the aerosol optical depth can be
obtained (Mejia et al. 2015, Kazantzidis et al. 2017). Use of sky imagers to measure circumsolar
irradiance is discussed in Section 14.8.3.
Sky imagers are providing new data from which the solar irradiance can be studied and charac-
terized. Because there is an increasing demand for more and better solar resource information, it
will be interesting to see what roles sky imagers will play in the future.

14.8 CIRCUMSOLAR INSTRUMENTS


After the underperformance of a concentrating solar power (CSP) facility in North Africa was
traced to overestimates of the DNI as measured by quality pyrheliometers, interest in measurements
of circumsolar irradiance has increased. Circumsolar radiation is the portion of sunlight that comes
from the area around the solar disk but not from the solar disk itself and is treated similar to DNI
irradiance in tilted surface and illuminance models because it behaves similar to DNI irradiance.
Under clear skies, the intensity of circumsolar irradiance falls off gradually as the angle from the
solar disk increases. A portion of circumsolar irradiance get counted as DNI irradiance because the
pyrheliometers have a field of view of 5 degrees while the solar disk itself subtends an angle of about
0.52°. CSP systems can only concentrate DNI coming from the solar disk or the area immediately
around the solar disk, depending on the collector’s effective field of view. Therefore, to accurately
evaluate and/or estimate the performance of a CSP facility, the portion of the circumsolar irradiance
available to the collector/receiver that can be concentrated must be measured.

14.8.1 sun and auReole measuRement


There has been a long history of interest in the circumsolar irradiance. From 1976 through 1981,
Lawrence Berkeley Laboratory took 180,000 sun shape profiles from 11 cities in the United States
using a circumsolar telescope (Bendt and Rabl 1980). In the 1990s, Deutsches Zentrum für Luft-
und Raumfahrt (DLR) obtained 2,300 sun shape profiles from three European cities. The profiles
then were used for the development of a sun shape model (Buie 2004). Equipment currently used
to obtain these data are like the Sun and Aureole Measurement (SAM) instrument (DeVore et al.
2010). These specialized instruments are expensive and require high maintenance.
The SAM system is mounted on an automatic tracker and uses two cameras, one to image the
solar disk and the other to measure the aureole, light coming from around the Sun. There is also a
lens that focuses light on a projection screen with a blackened cavity in the center. The instrument
can be equipped with a sun photometer to provide spectral as well as broadband sun shape data.
The system also comes with software to help provide autonomous measurements.

14.8.2 attemPts to automatiCally and Continuously


monitoR tHe CiRCumsolaR iRRadianCe
There have been several attempts to develop less expensive, easier to maintain instruments that can
continually monitor circumsolar irradiance (Al-Ansary et al. 2017; Kalapatapu et al. 2012; Wilbert
2018). The Sunshape Profiling Irradiometer (SPI) discussed in Kalapatapu et al. (2012), intended
for continuous circumsolar measurements, has been undergoing development and testing for sev-
eral years. The SPI being tested at the National Renewable Energy Laboratory (NREL) is shown
in Figure 14.18. The instrument tilted at latitude measures the irradiance using a shadowband that
sweeps across the sky and measures the solar intensity during the sweep. The instrument is simi-
lar to a Rotating Shadowband Radiometer (RSR) in concept except the hardware and sensor have
been adopted for the measurement of the circumsolar irradiance. A  narrow vertical entrance slit
Ancillary Measurements 293

FIGURE 14.18 Photograph of the SPI being tested at NREL. The white disk with the vertical slit covers
the pyranometer and is parallel to the edge of the shadowband as it passes in front of the sensor. (Courtesy of
Peter Armstrong.)

has been placed in front of the pyranometer to limit the irradiance incident on the pyranometer and
the rotation rate of the shadowband is precisely controlled. By tilting the pyranometer at latitude,
the pyranometer is aimed more directly toward the Sun to help optimize the DNI and circumsolar
contribution. By monitoring the solar intensity during the sweep and measuring the intensity drop
as the shadow band nears the Sun, the intensity of the circumsolar irradiance can be measured and
the magnitude of the circumsolar contribution to the DNI can be estimated. A sky map from an SPI
undergoing beta testing is shown in Figure 14.19. A circumsolar model, such as Buie (2004), then is
used to fit the instrument’s data using a range of circumsolar values. The circumsolar value yielding
the best fit then is chosen as the circumsolar value. In many areas, this circumsolar irradiance con-
tributes 5 percent or less of the measured DNI value under clear skies. Under dust laden skies, the
percentage can be 50 percent or more (Kalpatapu et al. 2012). Therefore, to accurately evaluate the
294 Solar and Infrared Radiation Measurements

FIGURE 14.19 Sky map data from SPI taken on a clear December day at noon under an arctic air mass. As
the shadowband rotates slowly between the sensor and the Sun different portions of the sky are shaded. When
the shadowband completely blocks the Sun a minimum occurs representing the diffuse sky irradiance and
ground reflected irradiance (about 0.075 in this example). Using a circumsolar model, such as Buie (2004), a
series of estimates are derived for various levels of circumsolar irradiance. The circumsolar value that yields
the best fit to the measured data is then chosen as the circumsolar irradiance.

production of a CSP system, especially in dust laden areas, it is important to account for the portion
of circumsolar irradiance contained in the DNI measurements.

14.8.3 use of tHe sky imageR to measuRe CiRCumsolaR iRRadianCe


Like many instruments, sky imagers can have a variety of uses. How the sky imager is used and
how the images are processed determines the type of the information gathered. Often instruments
that have multiple uses sacrifice one aspect of their capability to enhance another. Because there is
limited information available about circumsolar radiation and most radiometers only provide the
magnitude of the irradiance with temporal resolution, attempts have been made to coax circumsolar
information out of sky images (Al-Ansary et al. 2017).
Tests have shown that it is possible to pinpoint the location of the solar disc under clear skies
and obtain an estimate of the circumsolar irradiance using variations of brightness and intensity in
the image. Further work is needed to determine the uncertainties in these circumsolar estimates.
For example, cameras are not designed to measure the intensity of solar irradiance, and cameras
have a pronounced spectral response effect. More importantly scattered light in the optics makes
precise measurements of radiance problematic. Therefore, the magnitude of the DHI and circum-
solar irradiance measured by this approach have large uncertainties. However, given a lack of vali-
dated instruments to accurately measure circumsolar values, the sky imagers may yet provide useful
estimates of the circumsolar irradiance important for predicting the CSP electrical production. If
these attempts prove to provide useful circumsolar values, then sky imagers specifically designed to
measure circumsolar irradiance likely will be developed.

14.8.4 use of tHe Rotating sHadoWband RadiometeR


Like the sky imager, the RSR is increasingly used in conjunction with solar energy projects. During
the rotation of the shadowband over the detector, parts of the circumsolar radiation are blocked.
Ancillary Measurements 295

By examining the intensity of light measured in a burst (high-frequency) mode during the shadow-
band rotation, it is possible to estimate the circumsolar irradiance (Wilbert 2018) (see Chapter 10
for a discussion of the RSR). The burst mode sampling of the RSR occurs when the instrument’s
shadowband starts its rotation to shade the pyranometer. Multiple measurements of solar intensity
are made during this rotation that lasts about 1 second. Although the RSR can be used to get an
estimate of the circumsolar irradiance, it is not designed for circumsolar measurements. The instru-
ment discussed earlier in Section 14.8.2 contains enhancements that would facilitate the analysis
and accuracy of circumsolar irradiance. This is an example of where an instrument has a variety of
uses but enhancing one use degrades the other use.
Interest in measuring the circumsolar irradiance for CSP has been the impetus for develop-
ing instruments capable of continuously monitoring the circumsolar irradiance. Before these new
instruments provide circumsolar data, they will need to be tested against high quality circumsolar
measurements to ascertain the uncertainty in their measurements.

14.9 RECOMMENDED MINIMUM ACCURACIES


FOR OPERATIONAL INSTRUMENTS
When planning for meteorological measurements, it is important to get a feeling for the resolution,
data recording intervals, uncertainties, time constants, and achievable measurement uncertainties
of meteorological variables. The Guide to Meteorological Instruments and Methods of Observation
(WMO 2014) is a good reference for this type of information. Table 14.3 shows some of the informa-
tion for the measurements discussed in this chapter.

TABLE 14.3
Operational Measurement Characteristics
Achievable
Sensor Output Measurement
Variable Range Time Constant Averaging Time Resolution Uncertainty
Air temperature –80 to 60°C 20 s 1 min 0.1 K 0.2 K
Dew-point –80 to 35°C 20 s 1 min 0.1 K 0.25 K
temperature
Relative humidity 0 to 100% 40 s 1 min 1% 3%
Pressure 500–1,080 hPa 20 s 1 min 0.1 hPa 0.15 hPa
Wind speed 0–75 ms–1 2–5 m 2 min 0.5 ms–1 0.5 ms−1 for ≤5 ms–1
10% for > 5 ms–1
Wind direction 0–360° 1 s 2 min 1° 5°

QUESTIONS
1. What international organization sets the standards for meteorological measurements?
2. If the temperature is 300 K, what is the temperature in Celsius? Fahrenheit?
3. Name two types of temperature sensors.
4. What is an aspirator, and how does it help provide a more accurate temperature
measurement?
5. What is dew-point temperature, and how is it measured?
6. Name three types of anemometers.
7. What is an anemometer’s response length?
8. A zero reading for wind direction means the wind is blowing from what direction?
9. What are some of the problems that affect wind measurements?
296 Solar and Infrared Radiation Measurements

10. What is a wind rose?


11. What does a sling psychrometer measure, and how does it work?
12. Relative humidity sensors do not  respond linearly to changes in atmospheric moisture.
How is this nonlinear behavior corrected?
13. Name two types of pressure sensors and describe how each one works.
14. What is the difference between resolution and measurement uncertainty? Which is larger?
15. What are the international units for measuring atmospheric pressure?
16. What is a Horizon Survey and why is it useful for site selection?
17. If a cloud is moving with the wind and the wind speed is 5 miles an hour, how far does the
cloud shadow move in 15 minutes?
18. If the cloud is due south 15 degrees above the horizon and the Sun is also due south 45
degrees above the horizon, what other information is needed to know if the cloud will
block the Sun? How would this information be obtained?
19. How could satellite-based observations of cloud cover be used for solar resource forecasting?
20. How does circumsolar irradiance effect the power production of a concentrating solar
facility?

REFERENCES
Al-Ansary, H., T. Shafiq, A. Rizvi, and A. El-Leathy. 2017. Measurement of circumsolar ratio in high dust
loading regions using a photographic method. AIP Conference Proceedings 1850: 140003.
Badescu, V. ed. 2008. Modeling Solar Radiation at the Earth’s Surface. Berlin, Germany: Springer-Verlag.
Bendt, P., and A. Rabl. 1980. Effect of Circumsolar Radiation on Performance of Focusing Collectors,
SERI/TR-34-093. Golden, CO: Solar Energy Research Institute. https://www.nrel.gov/docs/legosti/
old/093.pdf (Accessed May 27, 2019).
Buie, D. C. 2004. Optical considerations in solar concentrating systems, Ph. D. thesis, Sydney, Australia:
University of Sydney.
Busch, N. E., and L. Kristensen. 1976. Cup anemometer overspeeding. Journal of Applied Meteorology 15:
1328–1332.
Coppin, P. A. 1982. An examination of cup anemometer overspeeding. Meteorologische Rundschau 35: 1–11.
DeVore, J. G., A. T. Stair Jr., A. J. LePage, and D. Villanucci. 2012. Using scattering calculations to compare
MODIS retrievals of thin cirrus optical properties with SAM solar disk and aureole radiance measure-
ments. Journal of Geophysical Research 117: D1. doi:10.1029/2011JD015858.
Garrison, J. D., and G. P. Adler, 1990. Estimation of precipitable water over the United States for application to
the division of solar radiation into its direct and diffuse components. Solar Energy 44: 225–241.
Hunter, R. S. ed. 2003. Expert Group Study on Recommended Practices for Wind Turbine Testing and
Evaluation—11. Wind Speed Measurement and Use of Cup Anemometry, 2d ed. International Energy
Agency Wind Productions.
International Energy Agency (IEA). 1999. Recommended Practice 11: Wind Speed Measurement and Use of
Cup Anemometry (Task 11). Olympia, Washington, DC: IEA Technology Collaboration Programme.
https://community.ieawind.org/viewdocument/recommended-practice-11-wind-speed (Accessed May
27, 2019).
Kalapatapu, R., S. Wilbert, M. Chiesa, and P. R. Armstrong. 2012. Measurement of sunshape with a sunshape
profiling irradiometer. SolarPACES 2012 Conference, Granada, Spain.
Kassianov, E., E. A. Riley, J. M. Kleiss, C. N. Long, L. Riihimaki, D. Flynn, C. Flynn, and L. K. Berg. 2017.
Macrophysical properties of continental cumulus clouds from active and passive remote sensing.
Proceedings of the SPIE 10424, Remote Sensing of Clouds and the Atmosphere XXII 104240A.
doi:10.1117/12.2278029.
Kazantzidis, A., P. Tzoumanikas, E. Nikitidou, V. Salamalikis, S. Wilbert, and C. Prahl. 2017. Application of
simple all-sky imagers for the estimation of aerosol optical depth. AIP Conference Proceedings 1850:
40012. doi:10.1063/1.4984520.
Keihm, S. J., Y. Bar-Sever, and J. Liljegren. 2001. Water Vapor Radiometer–Global Positioning System
Comparison Measurements and Calibration of the 20  to 32  Gigahertz Tropospheric Water Vapor
Absorption Model, TMO Progress Report 42-144, February 2001.
Kleissl, J. ed. 2013. Solar Energy Forecasting and Resource Assessment. Oxford, UK: Elsevier.
Ancillary Measurements 297

Köppen, V. 1915. A uniform thermometer exposure at meteorological stations for determining air temperature
and atmospheric humidity. Monthly Weather Review 42: 389–395.
Kristensen, L. 1993. The cup anemometer and other exciting instruments. Technical Report R-615(EN). Risø
National Laboratory, Roskilde, Denmark.
Kristensen, L. 1998. Cup anemometer behavior in turbulent environments. Journal of Atmospheric and
Oceanic Technology. 15: 5–17.
Kristensen, L. 1999. The perennial cup anemometer. Wind Energy 2: 59–75.
Kristensen, L. 2005. Fragments of the cup anemometer history. http://www.windsensor.com/application/
files/1314/2694/4639/Fragments_of_The_Cup_Anemometer_History_20050214.pdf (Accessed May
27, 2019).
Kurtz, B., and J. Kleissl. 2017. Measuring diffuse, direct and global irradiance from a sky imagery. Solar
Energy 141: 311–322.
Kurtz, B., F. Mejia, and J. Kleissl. 2017. A  virtual sky imager testbed for solar energy forecasting. Solar
Energy 158: 753–759.
Long, C. N., D. W. Slater, and T. Tooman. 2001. Total Sky Imager Model 880 Status and Testing Results.
U.S. Department of Energy, Atmospheric Radiation Measurement Program, DOE/SC-ARM/TR-006.
https://www.arm.gov/publications/tech_reports/arm-tr-006.pdf (Accessed May 27, 2019).
Lundquist, J. K., M. J. Churchfield, S. Lee, and A. Clifton. 2015. Quantifying error of lidar and SODOR
Doppler beam swinging measurements of wind turbine wakes using computational fluid dynamics.
Atmospheric Measurements Techniques 8: 907–920. doi:10.5194/amt-8-907-2015.
Maxwell, E. L. 1998. METSTAT: The solar radiation model used in the production of the national solar radia-
tion data base (NSRDB). Solar Energy 62: 263–279. doi:10.1016/S0038-092X(98)00003-6.
Mejia, F. A., B. Kurtz, K. Murray, L. Hinkelman, M. Sengupta, Y. Xie, and J. Kleissl. 2015. Coupling sky
images with three-dimensional radiative transfer models: A new method to estimate cloud optical depth.
Atmospheric Measurement Techniques Discussions 8: 11285–11321. doi:10.5194/amtd-8-11285-2015.
Miloshevich, L. M., H. Vömel, D. N. Whiteman, B. M. Lesht, F. J. Schmidlin, and F. Russo. 2006. Absolute
accuracy of water vapor measurements from six operational radiosonde types launched during AWEX‐G
and implications for AIRS validation. Journal of Geophysical Research 111: D09S10.
Myers, D. R. 2013. Solar Radiation, Practical Modeling for Renewable Energy Applications. Hoboken, NJ:
CRC Press.
Nguyen, A., M. Velay, J. Schoene, V. Zheglov, B. Kurtz, K. Murray, B. Torre, and J. Kleissl. 2016. High PV
penetration impacts on five local distribution networks using high resolution solar resource assessment
with sky imager and quasi-steady state distribution system simulations. Solar Energy 132: 221–235.
Nguyen, D. A., and J. Kleissl. 2014. Stereographic methods for cloud base height determination using two sky
imagers. Solar Energy 107: 495–509. doi:10.1016/j.solener.2014.05.005.
Patterson, J. 1926. The cup anemometer. Transactions of the Royal Society of Canada, Ser. III 20: 1–54.
Perez, R., R. Seals, P. Ineichen, R. Stewart, and D. Menicucci. 1987. A new simplified version of the Perez
diffuse irradiance model for tilted surfaces. Solar Energy 39: 221–232.
Preston-Thomas, H. 1990. The international temperature scale of 1990 (ITS-90). Metrologia 27: 3–10.
Rossow, W. B. 1996. International Satellite Cloud Climatology Project (ISCCP): Documentation of New
Cloud Datasets, WMO/TD-No. 737. Geneva, Switzerland: World Meteorological Organization.
Schreck, S., J. Lundquist, and W. Shaw. 2008. U.S. Department of Energy workshop report: Research needs
for wind resource characterization, June. NREL/TP-500-43521.
Schubert, D., W. Jenschke, T. Uhlig, and F. M. Schmidt. 1987. Piezoresistive properties of polycrystalline and
crystalline silicon films. Sensors and Actuators 11 (2): 145–155.
Singh, R., L. L. Ngo, H. S. Seng, and F. N. C. Mok. 2002. A silicon piezoresistive pressure sensor. Proceedings
of the First IEEE International Workshop on Electronic Design, Test and Applications (DELTA 2002),
Christchurch, New Zealand.
Smith, A., N. Lott, and R. Vose. 2011. The integrated surface database: Recent developments and partnerships.
Bulletin of the American Meteorological Society 92: 704–708. doi:10.1175/2011BAMS3015.1.
Smith, C. S. 1954. Piezoresistance effect in germanium and silicon. Physics Review 94: 42–49.
Tufte, O. N. and E. L. Stelzer. 1963. Piezoresistive properties of silicon diffused layers. Journal of Applied
Physics 34: 313–318.
Urquhart, B., B. Kurtz, and J. Kleissl. 2016. Sky camera geometric calibration using solar observations.
Atmospheric Measurement Techniques 9 (9): 4279–4294.
Urquhart, B., B. Kurtz, E. Dahlin, M. Ghonima, J. E. Shields, and J. Kleissl. 2014. Development of a sky imag-
ing system for short-term solar power forecasting. Atmospheric Measurements Techniques Discussion
7: 4859–4907. doi:10.5194/amtd-7-4859-2014.
298 Solar and Infrared Radiation Measurements

Wang, G., B. Kurtz, and J. Kleissl. 2016. Cloud base height from sky imager and cloud speed sensor. Solar
Energy 131: 208–221.
Wilbert, S., M. Röger, J. Csambor, M. Breitbach, F. Klinger, B. Nouri, N. Hanrieder et al. 2018. Sunshape
measurements with conventional rotating shadowband irradiometers. AIP Conference Proceedings
2033:190016. doi:10.1063/1.5067201.
WMO. See World Meteorological Organization.
World Meteorological Organization. 2014. WMO guide to meteorological instruments and methods of observa-
tion, WMO-No. 8 (2017 ed.). Chapter 8. http://www.wmo.int/pages/prog/www/IMOP/CIMO-Guide.html
(Accessed May 27, 2019).
Yang, H., B. Kurtz, A. Nguyen, B. Urquhart, C. W. Chow, M. Ghonima, and J. Kleissl. 2014. Solar irradiance
forecasting using a ground-based sky imager developed at UC San Diego. Solar Energy 103: 502–524.
15 Solar Monitoring Station
Best Practices
The question always arises as to the true reliability of any measurement performed with an
instrument under different conditions than those of the calibration.
John R. Hickey (1970)

15.1 INTRODUCTION
Considerable planning and forethought should go into the design and layout of a solar monitoring
station. Initial considerations should identify project goals and data accuracy requirements for the
project. After the basic parameters for the solar monitoring station are established, it becomes easier
to select between various design options. The optimum site specifications and location, instrument
selection, data acquisition system functions, and the operation and maintenance requirements for
the measurement equipment all directly affect the quality of the data that will be produced. Budget
constraints and continuity of funding for station operation and maintenance are also important.
Specifically, high-performance (expensive) instrumentation should not be considered if the project
resources are insufficient to support the maintenance and calibration schedule recommended by the
sensor manufacturer, and there are insufficient resources to purchase auxiliary supporting equip-
ment necessary to obtain the accuracy specified. A well-chosen, designed, and constructed solar
monitoring station will enhance the quality of collected data and help minimize the time and effort
necessary to service the station and validate the data. Three factors usually dictate the location of
a solar monitoring station: (1) a clear field of view (adequate solar access), (2) proximity to a site
operator who will maintain the station, and (3) most importantly, the requirements of the station
sponsor who will rely on the data.
While the main focus of this book is to characterize the instruments used in measuring solar and
infrared (IR) irradiance, the resulting data are only as good as the station installation, operation,
and maintenance. To expand on the limited number of publications available on this topic (Hickey
1970; McArthur 2005; Stoffel et al. 2010; Sengupta et al. 2017), this chapter discusses many of the
details necessary to locate, construct, and maintain a quality solar and IR radiation monitoring sta-
tion. Specifically, this chapter discusses factors to consider in choosing a site and determining the
requirements for the data-logging system. Maintenance of the station is described along with best
practices for cleaning the instruments and following standard procedures to record station mainte-
nance. The importance of viewing the data on a regular basis is discussed along with other quality
control measures. Next, suggestions for field calibrations of the radiometers are presented. Finally,
examples of solar monitoring stations are shown.

15.2 CHOOSING A SITE


Often the general location of the solar station is selected to fill certain requirements, such as provid-
ing data for a solar electric facility or characterizing the climate for a specific ecosystem. However,
just as important, and maybe more so, is the ease of access and maintenance. If the station is
located close to the person responsible for servicing the instrumentation and the measurement sys-
tem, the maintenance can be performed on a regular schedule. Access to electricity and to remote

299
300 Solar and Infrared Radiation Measurements

communication is important, although photovoltaic (PV) panels, cellular modems, and satellite data
links can provide sufficient power and communications to even the remotest stations.
Before selecting a location for the monitoring station, it is important to identify the person or group
who will maintain the station. Working with the station maintenance personnel, the potential locations
with the best field of view as well as optimum maintenance access and security should be identified.
It should be noted that finding a location that is totally unobstructed as the Sun moves across the sky
throughout the year is difficult. Trees, buildings, and power poles are among the many structures that
present problems. To identify the significance of the problem and to compare potential locations, it is
useful to obtain a sun path diagram for the site under consideration and take the diagram out to the loca-
tion under consideration. Sun path diagrams are discussed in Chapter 2 and can be created using the Sun
Chart Program from the University of Oregon (http://solardata.uoregon.edu/SunChartProgram.php). Sun
path charts for every 10 degrees of latitude are provided in Appendix D. Use a compass to locate mag-
netic north and make the magnetic declination corrections so that the sun path diagram can be aligned
facing the equator (south in the Northern Hemisphere and north in the Southern Hemisphere). A hori-
zontal plot of the obstructions can be made with the compass and clinometer to mark the location and
angular height of the obstructions. There are tools available that can take photographs of a location and
superimpose sun path charts on the panoramic image of the site. These tools are designed for siting solar
systems, and, while they can be used to quickly identify obstructions, they do not necessarily produce
accurate enough obstruction height information near the horizon (Figure 15.1).
To simplify descriptions for the rest of this chapter, it will be assumed that the station is located
in the Northern Hemisphere. For stations in the Southern Hemisphere, switch the terms north and
south. During the summer, observers will notice that the Sun will rise and set somewhat behind
them if they are facing the equator (south in the Northern Hemisphere and north in the Southern
Hemisphere). The farther away from the equator the site is located, the farther behind the observers’
back the Sun will rise or set during the summer months.
A site with no obstructions higher than 5 degrees above the horizon is ideal, but sites with obstruc-
tions up to 10 percent are acceptable, especially if the obstructions do not block the Sun’s path dur-
ing any part of the year. Sometimes obstructions cannot be avoided, such as nearby hills or trees.
If at all possible, keep obstructions out of the Sun’s path and as far back (north) as possible from
the east–west axis. If the obstructions are behind the site, they will not block the Sun’s path during
the winter when the Sun is lowest in the sky. If there is an obstruction that is not too tall (e.g.,  less

FIGURE 15.1 Image of Ashland, Oregon, station using the Solmetric Suneye. The Suneye first takes a fish-
eye view of the station, shown on the left-hand side of the image, and then places a grid of the path followed
by the Sun during different times of year, shown on the right-hand side of the image. Local standard time is
given in the image. Objects on the horizon that block the Sun’s path are then analyzed, and the percent of the
shaded area is then calculated.
Solar Monitoring Station Best Practices 301

than 15 degrees above the horizon) or wide, then it may be possible to locate the station so that the
obstruction is directly south of the station. Study the sun path chart, and make sure that the Sun
passes over the obstruction during the winter months. If the structure is too tall, it may be necessary
to move the station farther east or west to avoid shading as much as possible.
Trees can be a particular problem for solar-monitoring sites. Trees grow, and while they may
not block the Sun’s path at first this may not be the case several years later. Therefore, it is prudent
to know future plans for development near the site. Sometimes the station can be mounted on a
raised platform or a rooftop to avoid blockage by some local obstruction. If the station is located
on the roof of a building, adequate and safe access to the equipment should be incorporated into
the design.
Some solar stations are located near meteorological towers that are used to monitor the wind
resource. It is important to locate the station to the south of such towers and to make sure the tower
is not  highly reflective. In  fact, any object that is likely to reflect light toward the pyranometers
should be removed or painted with a flat black, gray, or nonreflective coating.
Some stations are located in areas that are subject to snow and snow drifts. The solar-monitoring
stations should be located high enough above ground to avoid direct reflections from the snow and
should be constructed so that minimal snow accumulates on the platforms around the pyranometers
and other instruments. Nonconductive gratings used for walkways make useful platforms. If the
temperature gets cold enough, enclosures containing the data logger or other equipment may need
electrical heaters controlled by thermostats.
The equipment in the boxes should be mounted so that moisture from condensation cannot
drip on electric connections. If appropriate, desiccant packs can be used to help keep sealed
boxes dry. Sometimes “Weep Holes” at the base of the enclosure are used to let the condensate
drain freely. Electrical connections to the boxes should be laid in conduit or contain a drip loop
so that water will not run down the cables and into the electronics box. Signal cables and power
cables should be separated as much as possible to avoid electrical interference with the instru-
ment signals.

15.3 GROUNDING AND SHIELDING


A solid electrical ground is important for lightning protection and signal noise protection. For most
areas, a copper-coated steel ground rod 1.5–3  meters (m) in length is sufficient for grounding.
The ground rod is typically between 10 and 20 millimeters (mm) in diameter. To drive the ground
rod into dirt, a heavy metal pipe with an end cap can be used. For areas of frequent lightning strikes,
a more robust grounding strategy might be needed. For more detailed information on grounding and
shielding see Morrison (1998).
All masts and metal poles should be grounded to allow an easy path for lightning to follow
to ground. A  lightning finial is often recommended for areas subject to more frequent lightning
strikes. Surge protectors such as metal oxide varistors or avalanche diodes can be used for lightning
protection. These surge protectors degrade over time, and manufacturers’ recommendations should
be followed when determining when these surge protectors need to be replaced.
Sensitive equipment usually has a ground lug that can be used for grounding. A 10- or 12-gauge
wire (AWG units) with 2.6–2.0  mm diameters, respectively, should be used for connecting the
equipment to the central grounding point.
Twisted pair shielded cable should be used for low-voltage signals and should be connected as a
differential measurement at the data logger. By isolating the signal low from ground, some of the noise
can be avoided. The shield should be connected at one end to the ground and should be left unattached
at the other end to prevent ground loops. Signal cables should be limited to 50 m in length and also
be physically isolated from power cables, and if they have to cross they should cross at right angles to
minimize any induced signals. More instruments are now available with digital output (i.e., RS-485)
allowing for longer cable lengths and easier interfacing.
302 Solar and Infrared Radiation Measurements

15.4 DATA LOGGER AND COMMUNICATIONS


Many data-logging systems are available; however, it is important to have data loggers that can
withstand field environmental conditions for many years maintaining their electrical specifications
over the range of temperatures experienced at the station. Providing a temperature-controlled enclo-
sure should help to reduce the subtle effects caused by temperature drift, but this increases the elec-
trical power requirements for the system. A good outdoor data logger should maintain specifications
for the temperature ranges expected at the measurement site (See Appendix G).
The clock on the data logger should have minimal drift. This requirement should be considered
for any new station, especially those that sample with high frequency such as every second. Some
data loggers come with software that will automatically synchronize the data-logger clock with the
computer clock. If that is the case, then it is important to ensure that the computer clock is correct
and does not drift. Coordinated Universal Time (UTC) and local time in the United States can be
found at http://www.time.gov, and either is recommended for data logging. Daylight saving time
is not recommended. It is important to monitor the data logger time and change it only if it has
drifted more than 1 or 2 seconds since the previous check. Some data loggers are equipped with
Global Positioning Satellite (GPS) sensors and these devices can be used to maintain an accurate
time record. Of course, it is still important to check the time occasionally in case a problem arises.
Timing errors effect solar position calculations.
The  data logger should be selected to have enough analog inputs, voltage excitation sources,
and control ports to handle the number of sensors being installed at the site. Plan to have room for
additional instruments. Over time, requirements change, and it is helpful if data loggers do not have
to be changed to accommodate additional measurements. Some data loggers come with auxiliary
multiplexers to handle additional channels of data.
The amount of storage in the data logger is an important consideration. The more frequent the
sampling and the greater the number of instruments, the greater the amount of storage required to
ensure continuous data reporting. The data storage should be sufficient to account for the longest
time between data transfers to a central computer.
Data loggers should be able to handle a variety of inputs. Some anemometers generate pulses that
need to be counted, temperature sensors require a reference voltage, and many pyranometers need
data loggers that can measure with microvolt precision. Make sure the data logger can handle the
signals from the sensors being chosen and from sensors that might be needed in the future.
Accuracy and linearity of the data logger is important and should be better than the sensors
being installed. Data loggers with a 0.5 percent accuracy or better are needed to avoid adding to the
measurement uncertainty of the data.
It is important to be able to communicate with and download data from the data logger on a regular
basis. It is very helpful if the data logger can be queried remotely and the data logging routines can be
changed from a central location. Remote communication enables troubleshooting from the central office
where technical expertise is available. Communications can be through a phone modem, internet, cel-
lular phone, satellite phone, or use of special communication networks. Over the long-term, communi-
cation costs can add up, and it is important to be able to change to new methods as they are developed.
For  ease of access, internet communication is best, but simple phone modems are very reliable if a
phone line is available. With cell service going to 5G networks, specifications for cell modems that work
with the network change, and phone modems may need to be upgraded to be compatible with the latest
technology. It is important to calibrate the data logger periodically, just as the radiometers should be
calibrated. This calibration helps ensure the quality of the data and also helps spot potential problems.
Power failures in the field are to be expected. The  data logger should have a battery backup
system and should be able to automatically recover data and programs if a power outage drains the
battery. It is very helpful if the data logger draws only small amounts of power. This requirement
minimizes the size of the backup power needed and provides for the opportunity to use PV panels
for the power source. The size of batteries and PV panels should be sufficient to last for long periods
Solar Monitoring Station Best Practices 303

during a winter dominated by clouds. Charge controllers that adjust for cold weather condition
should be used in conjunction with PV panel-charged batteries.
Software packages that come with the data logger are useful, but the software should be able to
output data from the logger in a format that is readable to a wide variety of computer programs or
comes with a translator that can export the data to an ASCII file.
With solar irradiance measurements, the ideal would be to have the sensor read continuously
over the time interval. This procedure would yield an integrated average of the sensor being read.
The practical alternative is to have the sensor read at very short intervals of a second or two and
these readings averaged over the time interval of interest. If interest is in the variance of the irradi-
ance over the time interval, it would be necessary to have a data logger that could record the stan-
dard deviation of the measurement over the time the measurement is averaged. Similarly, recording
the maximum and minimum signal during the averaging period can provide useful information.
Other factors to consider are the complexity of programming the data logger, readable and infor-
mative instruction manuals, and a quick response to requests for assistance with problems. The data
logger is a key component, and reliability and backup are needed to make the data set as complete
as possible. For larger networks it is useful to have spare components so that failure in the field can
be addressed quickly to minimize the loss of data.

15.5 MEASUREMENT INTERVAL


Choice of time interval for the measurements is often specified by the requirements of the user
of the data. For example, many electric utilities use 15-minute readings for their power monitor-
ing. However, irradiance data has many uses, and where one time interval is preferred by one user
another time interval may be preferred by another. With shorter interval data, it is possible always
to sum or average into longer time intervals, but a 15-minute data set cannot be subdivided into a
5-minute dataset without introducing additional uncertainties.
Lately, there has been interest in 1-minute or shorter time interval data. With short time intervals
it is important to look at the sensor’s response time, and unless examining the response time of a
sensor the time intervals chosen should be long enough so that the sensor can provide an accurate
measurement integrated over the time period. The Baseline Surface Radiation Network (BSRN)
recommendation (McArthur 2005) is to sample each second and store a 1-minute average. Storing
data in 10-minute intervals should be avoided because the data will not divide evenly into the utility
standard of 15 minutes.
Note that most satellite-derived data come from images that are analyzed each half-hour,
and the timing of the images does not necessarily fall on the hour or half-hour mark. The older
Geostationary Operational Environmental Satellite (GOES) East geosynchronous orbit weather sat-
ellite produces images of the Northern Hemisphere 15 and 45 minutes after the hour. The GOES
West geosynchronous orbit weather satellite produces images of the Northern Hemisphere on the
hour and 30 minutes after the hour. To validate models that use satellite images to estimate solar
data, the time sampling interval for ground-based measurements should be sufficiently short that it
captures the data at the point in time when the satellite image is taken (Wilcox and Marion 2008).
Ground-based solar data integrated over 5-minute intervals around the time the satellite images
were taken would provide a better comparison to the solar data derived from the satellite images
than ground-based data taken at 15-minute or hourly intervals.
Some data loggers and computers will automatically change from daylight saving time to local
standard time (LST) and vice versa. However, it is best to record data in either LST or UTC.
Keeping the time standard constant means that the data files will be continuous and will not have
discontinuities associated with daylight savings time. The  matter of consistent data record time
stamping is important for accurate solar position calculations needed for data quality assessment.
UTC relates global time to an atomic clock standard. GPS, computer time, and other equipment
times are related to UTC time. The advantage of UTC is that it is the same throughout the world; the
304 Solar and Infrared Radiation Measurements

disadvantage is that the 24-hour UTC day divides the local solar day over 2-day boundaries at many
locations. LST is best for small networks and local data, especially if the data are to be accessed by
the general public. UTC time will be found with larger databases or with satellite-derived data sets
(Andreas et al. 2018).
The granularity of the data (time interval) affects the precision to which changes in the solar
irradiance can be perceived. Short time intervals can help in spotting problems with the data such
as reductions in the output when the instrument is cleaned. Data applications dictate the maximum
time interval needed for recording. Sample plots of 1-minute, 5-minute, 15-minute, and hourly data
are shown in Figure 15.2a–d. Note that as the time interval increases, the information on the vari-
ability in the data becomes more and more obscured.

FIGURE  15.2 Plots of GHI, DNI, and DHI on August  14, 2011, at time intervals from (a) 1 minute,
(b) 5 minutes. (Continued)
Solar Monitoring Station Best Practices 305

FIGURE  15.2 (Continued) Plots of GHI, DNI, and DHI on August  14, 2011, at time intervals from
(c) 15 minutes, and (d) 1 hour. Information on the variability of the irradiance data becomes more and more
obscured as the averaging time interval of the data increases.

15.6 CLEANING AND MAINTENANCE


High-quality pyranometers require minimal maintenance other than cleaning the dome on a regular
basis and making sure the desiccant does not become saturated with moisture. It should be months
to a year between desiccant changes for a properly assembled pyranometer. If the desiccant needs
to be changed more often, that is usually an indication that a seal has developed a leak (typically at
the desiccant chamber/body interface). Gel desiccants that change color when they absorb moisture
are recommended because they do not leave dust that is associated with some desiccants (calcium
chloride should be avoided). The dust from a desiccant can travel to the surface of the sensor or to
the inside surface of the dome, thus affecting the responsivity of the pyranometer. Occasionally, the
306 Solar and Infrared Radiation Measurements

rubber O-rings sealing the desiccant holder will need to be greased to keep them from drying and
cracking. Grease that is used for O-rings on vacuum pumps works well as it does not outgas when
heated and leave a residue inside the pyranometer.
Moisture, frost, or snow also blocks or scatters incident solar radiation. While dust may build
up slowly over time and regular cleaning can minimize the effects of dust, frost or snow can affect
the instrument immediately. A properly designed ventilator can include electrical heating and will
keep most frost and light snow off a pyranometer while minimizing the accumulation of dust. Of
course, heavy snow can accumulate on the shield or around the ventilator, so the platform should
be constructed so that snow cannot accumulate around the instruments, for example, by mounting
instruments on a grate. More information on ventilators is given in Chapter 5.
Pyrheliometers are more susceptible to the effects of dust or moisture. A little dust or a drop of
moisture on the instrument window can scatter the incident irradiance so that it does not reach the
detector. For the most accurate results, the window of the pyrheliometer should be cleaned on a
daily basis, or no less often than twice each week. This schedule depends, of course, on the environ-
mental setting of the measurements and time of year. It is important to note on a log sheet when the
instruments were cleaned and what, if any, obstructions were added or removed.
To clean the instrument windows or domes, compressed air or distilled water are the preferred
cleaning agents. Ethyl alcohol (ethanol) may be used in icing conditions, although if the temperature
is above freezing during part of the day it is better to wait and clean with distilled water. Overuse of
alcohol can cause the seals to dry and crack. Cracked seals can allow moisture to enter the instru-
ment and degrade the measurements. If alcohol is used, 95 percent alcohol should be used because
the other 5 percent is water. With 99 percent alcohol, other chemicals are present that enable the
purity to reach 99 percent. These chemicals may leave a residue. A lint-free cotton cloth or paper
products designed for cleaning optical surfaces should be used to clean the radiometers. Any lint or
substance left behind affects the performance of the instrument. If a cotton cloth is used, it should
be laundered frequently to prevent grime from accumulating. It is also possible to use a can of pres-
surized air to dry the dome to avoid the microscopic scratching from rubbing the dome with a dry
cloth. Be sure that the spray can does not have an oil-based propellant or other chemicals that can
coat the surface of the window or dome.
Occasionally, chemicals from rainwater or treated water form a film on the instrument that cannot
be cleaned by water or alcohol. A little vinegar should remove these water stains. A little cesium oxide
that is used for polishing mirrors can be used to remove stubborn stains as a last resort. If the window
or dome becomes scratched or pitted, a substitute instrument should replace the damaged instrument
and the window or dome should be replaced or the instrument should be taken out of service.
On some older models of pyrheliometers, the seal around the window can crack and moisture can
enter the instrument. If moisture gets into the pyrheliometer, the data begin to show strange behav-
ior, especially if there is enough moisture to condense on the inside of the window. An example of
the effect of moisture in the pyrheliometer is shown in Figure 15.3. When cleaning the instrument,
if moisture can be seen condensed on the window then the instrument should be removed for repair.
Many pyranometers contain a desiccant to keep moisture from building up inside the instrument.
The desiccant should be checked each time the instrument is cleaned. Usually a quick glance will
show if the desiccant has changed color, indicating the need for replacement.
If the instrument is mounted inside a ventilator, the ventilator should be checked to see if the fan
is running and the ventilator has a free airflow. The filter screen below the fan should be cleaned at
least annually or more frequently according to the local conditions. Dust, mold, and other debris can
accumulate on the filter and inhibit the airflow. Removing this screen should be considered if it does
not adversely affect the operation of the pyranometer.
Although most snow is easy to dust off the domes of pyranometers, ice is a different story. Do not
mechanically remove the ice from the pyranometer dome or diffuser. Sometimes alcohol can be used to
melt the ice from a glass dome, but alcohol should not be used on white (acrylic) diffuser disks. Other
times, a warm hand on the dome or disk will melt the ice. A sponge soaked in warm water can also
Solar Monitoring Station Best Practices 307

FIGURE 15.3 Image of a chart record of DNI on a clear day. The spikes are hour marks and the scale on left
is in millimeters. 60 millimeters corresponds roughly to 700 W/m2. This is typical behavior when moisture
condenses on the interior of a pyrheliometer window. Often when this happens, moisture can be seen on the
inside of the window, especially after it has just been cleaned.

help. Hot water should be avoided because too large a temperature differential might break seals, crack
the glass, or cause other damage to this important optical element of the pyranometer or pyrheliometer.
At least one site visit per year is recommended for preventive maintenance. During an annual site
visit, the wiring should be checked to ensure that cables are not deteriorating. Batteries at the sta-
tion should also be checked and replaced or recharged if necessary. If cabinets or platforms require
painting, this can be done during the visit.
It is important to have a maintenance manual for the measurement station (see Appendix I for trouble-
shooting hints). This will help people who tend the station know what needs to be done, what to look out
for in the way of problems, and how to contact the appropriate person should a problem occur (Andreas
et al. 2018). These manuals can also help in training and can be used to refresh the person’s memory if
a procedure has not been performed in a while. The manual should contain simple information like how
to clean a pyranometer to more complex tasks like setting the clock on an automatic tracker.
Putting together the manual while planning the station is most helpful. Potential issues that might
have been overlooked in the initial station design can arise when describing how to operate and
maintain the station. This manual helps to document why certain decisions were made and is very
useful when other stations are established. Make an electronic version of the manual so that it can
be changed and modified as experience is gained with operating the station.

15.7 RECORD KEEPING


Keeping records of maintenance, site visits, instrument serial numbers, instrument calibration, and
all activities associated with the station is essential. If there are questions about the data, the records
can be examined to see if there were any activities that might explain the questioned data. Carefully
maintained records enable the ability to go back and see when instruments were changed and what
calibration numbers were used during particular time periods.
Log sheets should be available for those maintaining the site. They are useful reminders of the
daily tasks and help record what maintenance was performed. An example log sheet is shown in
Figure  15.4. Table  15.1 lists some information that might be included on the log sheet. National
Renewable Energy Laboratory (NREL) is working on an application for a table that would allow
the technician to record work with automatic time stamping of each action and photograph the
noteworthy observations at the time.
Room on the log sheet should be reserved to record the beginning and end times when the main-
tenance was performed. In addition, it is helpful to have a list of the solar declination for each day of
308 Solar and Infrared Radiation Measurements

FIGURE 15.4 Sample page from the log sheets used at the high-quality stations in the University of Oregon
Solar Radiation Monitoring Network. It is important to note when the person was maintaining the instruments
and who did the maintenance. It is also useful to leave room for comments. It is easier to remember all the
tasks if the log sheet has a space for each task that should be performed.
Solar Monitoring Station Best Practices 309

TABLE 15.1
Types of Information on a Daily Log Sheet
Preferred
Task Performance Schedule Status
Clean the instrument Daily Status before and after cleaning
(note unusual build-up)
Check if desiccant is OK Weekly Color OK
Check if ventilator is working Daily Air flow, no air flow
Pyrheliometer alignment Daily Show alignment status
Weather Daily Show type of weather
Shade ball shadow covering Daily Yes or no or partially; adjust
pyranometer (DHI) shadowing
Time tasks begin Daily When at the site
Person performing the Daily
maintenance
Additional notes

the year. This list is necessary if a fixed shadowband has to be adjusted or a manual tracker is used
at the site. It is also useful information when checking the data. It is important to leave room on the
log sheets for notes. Unusual events such as birds roosting or flying insects swarming near instru-
ments should be recorded. A digital form of the log sheet should be available and the information on
the paper log sheet should be transferred to the electronic database. The NREL Measurement and
Instrumentation Data Center (MIDC) online maintenance log can be view at https://midcdmz.nrel.
gov/apps/maint.pl?BMS. The resulting metadata records of maintenance findings for each instru-
ment are available as links from the data pages. For example, the maintenance records for the mea-
surement of global horizontal irradiance (GHI) using a CM22 pyranometer are accessible from
https://midcdmz.nrel.gov/apps/fmpxml.pl?type=maint;find=Global%20CM22;site=BMS.

15.8 IMPORTANCE OF REVIEWING DATA


It is helpful to review the data daily. To make viewing the data easier, a routine should be written to
plot the data as a time series. Data intervals of 15 minutes or less are best for visual inspection of
the data. The trained eye can be more precise than computer programs and can usually spot a wider
variety of problems. Viewing plots of the data routinely will help to spot problems quickly and
minimize the amount of bad data that needs to be flagged and possibly eliminated from the database.
For example, if there are problems with the alignment of the pyrheliometer, a quick scan of the data
will show the problem. Dirt can build on the dome of the pyranometer, and while it may not be eas-
ily visible to the eye, it will show up as a depression on the plot of the calculated diffuse irradiance
(see Figure 15.5 and caption for explanation). Misalignments of shadow balls or disks can be seen
in plots. Any anomaly that occurs in plots is easily spotted. Obstructions blocking the Sun also show
up on clear days. If a dip shows up in the same spot over several clear days, it is likely to be caused
by an obstruction. Raindrops on the window of the pyrheliometer will scatter the direct normal irra-
diance (DNI), and if there is intermittent Sun mixed with rain sometimes beam values can be seen
slowly return to normal values while the Sun evaporates the rain from the window (see Figure 15.6).
Some pyranometers’ domes or lenses can have beads of moisture that reflect additional irra-
diance onto the sensor. This  effect is usually small and mostly happens in the morning hours.
However, if the pyranometer is part of a rotating shadowband radiometer (referred to here as an
RSP - Rotating Shadowband Pyranometer), the small enhancement in the global measurement will
lead to a fairly significant increase in the calculated DNI (see Figure 15.7). For light snowfalls, snow
310 Solar and Infrared Radiation Measurements

FIGURE 15.5 Plot of GHI, DNI, and calculated DHI on July 1, 2011, in Burns, Oregon. There is dirt on the
pyranometer that causes a slight decrease in the GHI values. When subtracting the DNI times the cosine of
the sza, the calculated DHI shows unrealistic structure. Because the DNI is relatively smooth, the variations
in the DHI indicates there are problems with the GHI. The pattern of the bumps for several days confirms the
problem with the GHI data. Alcohol was required to clean the thin film that caused this problem.

FIGURE 15.6 Plot of DNI and DHI on May 27, 2011. Raindrops on the front of the pyrheliometer window
scatters incoming irradiance. As the rain evaporates, the DNI values return to normal. This  phenomenon
occasionally occurs when rain is followed by a bright sunny period. Dashed gray line is DNI from an RSP,
and the solid gray line is the DNI from an Eppley NIP. The solid black line is the DHI from an disk-shaded
Schenk pyranometer, the black dot-dash line is the DHI calculated from the GHI from an Eppley PSP minus
the DNI from the NIP times the cosine of the solar zenith angle (sza), and the dotted black line is the DHI
calculated from the GHI from a PSP minus the DNI from the RSP times the cosine of the sza. (Courtesy of
Eppley Laboratory, Newport, RI.)
Solar Monitoring Station Best Practices 311

FIGURE  15.7 Plot of 1-minute RSP data along with reference data. There  is dew on the diffuser of the
LI-COR LI-200 pyranometer in the morning between 6:00 and 7:00. The GHI increases slightly while there is
a large bump on the DNI, which is calculated by subtracting DHI from GHI and dividing by the cosine of the
solar zenith angle. The RSP in this plot was calibrated against the DNI values and no correction algorithms
have been applied to the DHI data.

will sometimes melt from the sun-facing side first, and the rest of the snow will act as a reflector and
concentrate the sunlight onto the sensor disk. The enhanced GHI can show up clearly when plotted.
A series of other plots used for daily data quality inspections of solar and meteorological mea-
surements is available from https://midcdmz.nrel.gov/apps/maintday.pl?BMS.
In summary, there are many ways of checking the data using plots that can be useful, and while
it does take a little time, a better, more complete dataset will result.

15.9 QUALITY CONTROL OF DATA


Quality control of data is very important. Quality control can spot problems as well as increase confi-
dence in the data. Since irradiance data can vary rapidly and significantly, it is useful to have more than
one instrument measuring the solar irradiance (Long and Shi 2008). The best quality control method is
to have three measurements: the DNI, GHI, and diffuse horizontal irradiance (DHI). The GHI is equal
to the DNI times the cosine of the solar zenith angle plus the DHI. If the values agree to within a few
percent (e.g., ±3%), all three measurements are likely to be correct. If raindrops on the window of the
pyrheliometer scatter some of the incident irradiance, the calculated global value may differ markedly
from the measured GHI. Therefore, these checks can identify problems in the data.
It is important to have a procedure to ensure the quality of the data and to identify any data that
are bad. For example, when an instrument is cleaned, a spike (either an increase or decrease) is
often observed in the reported irradiance. This outcome is especially true when 1 minute or shorter
time-interval data are obtained. The data can be flagged or even corrected as long as the corrected
data are labeled. The procedure for flagging the data or correcting the data should be documented.
With data intervals longer than 5 minutes, the ease of observing problems decreases along with the
accuracy to which corrections can be made.
One method of identifying problem data is to plot the data and compare them with notes on the log
sheets. If an instrument’s data start to deviate from the data of similar instruments at the station, then it
312 Solar and Infrared Radiation Measurements

is worthwhile to examine log sheets to see if there was any activity at the station that may have resulted
in this discrepancy. Plots of the data showing dips in the data that seem to be caused by cleaning can be
easily verified by checking the log sheet records. If a problem with the data is properly identified, then
the problem data should be flagged or corrected. At other times it might be easier to examine the logs
and see if the activity at the station might have influenced any of the data being taken at the station.
Normalizing the data by dividing by the extraterrestrial irradiance is one useful method in qual-
ity control because it allows a large range of data values to be grouped and any outliers to be identi-
fied that might need further examination. If these clearness indices are above 1 for any length of
time, there is likely to be a problem. However, especially with high time resolution data, clouds
passing near the Sun can reflect the incident irradiance onto a pyranometer and increase the incom-
ing irradiance to values greater than the extraterrestrial irradiance. This is why it is useful to have
all three coincident measurements of DNI, DHI, and GHI.
If only two of the three measurements are made (DNI and GHI or GHI and DHI) a clearness
plot of the data is useful. DNI divided by the extraterrestrial beam irradiance is kb; GHI divided by
the equivalent extraterrestrial global irradiance is kt. It is possible to plot either kb against kt or DHI/
GHI against kt to evaluate the data. Figures 15.8 and 15.9 illustrate these plots. The distribution of
1-minute data points in these plots is typical and the data have been validated. The majority of data
fall into bands, and any data that significantly deviate from these distributions or show an unusual
grouping should be examined in more detail. With 5-minute or hourly data, the distribution signifi-
cantly narrows. A learning curve is necessary to more efficiently identify suspect data efficiently
as the distribution of possible data points is fairly broad. Certain values, such as kb greater than kt,
are not generally possible. For partially cloudy periods when it is clear in the Sun’s direction, GHI
can be greater than the extraterrestrial radiation. These high values result from reflections off of

FIGURE 15.8 Plot of the diffuse fraction plotted against clearness index kt for 1-minute data. Most of the
data fall along the lower left-hand side of the plot. Those data points represent clear periods. A spread of data
points to the left represents data when clouds are present. When there is no DNI, the diffuse fraction is 1, and
GHI = DHI. When there is some Sun present, the clouds can act as reflectors, and a wide variety of GHI values
can occur for any diffuse fraction. Note the number of clearness index values near 1 or greater than 1. These
GHI values are enhanced by sunlight reflecting off clouds and onto the sensor. As the averaging time interval
increases, the number of data points with kt values near or greater than 1 significantly decreases. When plot-
ting hourly average values, there are very few data points with kt greater than 0.8.
Solar Monitoring Station Best Practices 313

FIGURE  15.9 Plot of kb versus kt for 1-minute data. As with the diffuse fraction relationship, shown in
Figure A.1, the majority of points during clear periods fall on the left-hand side of the plot. Values with the
largest clearness index are not values with the largest beam clearness index values. This results from sunlight
reflecting off clouds and into the global pyranometer. This reflected light is not seen by the pyrheliometer that
measures the DNI values. Note that there are very few data points with k b values above a maximum of 0.8.
The maximum kb values are somewhat dependent on altitude because higher altitudes have a lower air mass
through which sunlight has to traverse.

clouds adding to the GHI. This condition leads to clearness values greater than 1 and represents
real data occurrences. These periods when the GHI are greater than the extraterrestrial radiation
are short-lived and are not seen in 10-minute, 15-minute, or hourly data files (see kb vs. kt plots in
Appendix A). Offsets or fluctuations when GHI values are small, say less than 10 watts per square
meter (Wm−2), can lead to spurious data points. If the spread of the data is too great or if there are a
lot of points outside the main grouping, then there may be problems with the data. Common prob-
lems are dirt on the pyranometer, misalignment of the pyrheliometer, misalignment of the shadow-
band, systematic biases with the instrument, or poor calibrations for the instrument.
For  DNI and GHI measurements, NREL developed the SERI QC model (NREL 1993) that
defines the region where most data points should fall. The  more the data points fall outside the
region, the larger the SERI QC flags. The SERI QC flags are found in the National Solar Radiation
Data Base (Sengupta et al. 2018). With analysis of a lot of data, programs such as SERI QC can
isolate possible problem data without visual inspection of every data point. SERI QC does not spot
all problems, but it does spot the more egregious ones.

15.10 FIELD CALIBRATIONS


In general, pyranometers used to measure GHI decrease their responsivity by between 0.5 and 1
percent per year, whereas most pyrheliometer changes are smaller. Therefore, it is important to
keep track of the instrument calibration to minimize a systematic drift in the data values over time.
This calibration is especially true if one wants to use the data is to be used to track changes in the
resource with high confidence (bankability).
The following are two of the methods to track instrument calibration over time. The first method is
to replace the instrument with an auxiliary instrument and send it to a qualified facility for calibration
314 Solar and Infrared Radiation Measurements

against a standard traceable to the international standard. Although changing out instruments for
calibration can be done, the replacement instruments will not have precisely the same characteristics
as the original instrument and the absolute calibration has an uncertainty between 2 and 3 percent.
This difference makes obtaining a long-term record more difficult because exchanged instruments
increase uncertainty and variation in the data. This added uncertainty can be addressed by quickly cal-
ibrating the instrument and reinstalling it at the site and removing the temporary substitute instrument.
An alternate method of calibration is to take well-calibrated reference instruments to the field
site and calibrate the instruments in the field. For  field calibrations, it is important to use the
same type of instrument for calibration because each model of instrument deviates from a perfect
response in its own characteristic manner. Therefore, it would not be a good practice to calibrate
a photodiode pyranometer from one manufacturer with a photodiode pyranometer from another
manufacturer or with a thermopile pyranometer. The  best side-by-side calibrations are done by
comparing the same model instruments. The  reference instrument should be calibrated yearly
against standards traceable to the international standards as discussed in the calibration sections
of Chapters 4 and 5 and in the overview presented in Chapter 13.
For field calibration it is important use the chosen angle-of-incidence as the reference point, for
example a solar zenith angle (sza) of 45°. Use the calibration data gathered over a range of sza’s near
the chosen reference sza (e.g., 43–47°) to determine the instrument’s calibration value. The range
should be large enough so that a number of measurements will be available for analysis. Calibration
data gathered outside this range can be used to track the instrument’s deviation from a true cosine
response and other biases, but the values in the 43–47 degree range can be used to track the overall
calibration from year to year and determine any long-term trends.
The station should be constructed so that it is easy to place the comparison instrument next to
the instrument being calibrated. The instruments should be on the same level. If the field instrument
is in a ventilator, then the reference instrument should be housed in a ventilator. If the calibration is
being done on an automatic tracker, be sure that the instrument is secured to the platform. Cables
from these temporary instruments should have enough play so that movement of the tracker will
not stress the cable and be placed, or safety flagged, so that they will not present tripping hazards
while working around the instruments.
It is also helpful to run the calibration overnight, especially for instruments that are subject to
thermal offsets. An estimate of the thermal offset can be made looking at the nighttime values.
In addition, any offsets or problems with the data logger or cabling are more conspicuous during the
nighttime when there is no solar radiation input to the irradiance sensor.
Often other site maintenance tasks are performed when the instruments are being calibrated.
Make sure that these tasks do not interfere with the calibrations, or if they do, note the time, and the
questionable data can be removed.
Calibration numbers usually have a ±2 to 3 percent uncertainty and calibration values typi-
cally change less than 1 percent per year, and this is within the uncertainty of the calibra-
tion measurement. Less variation is introduced into the data set if the calibration number is
not changed each time a new one is obtained provided the new calibration number is within the
uncertainty of the last calibration. Plot the change in calibration over time. After a consistent
trend is observed, then that calibration trend can be used to adjust the data to the more accurate
value. See Chapter 13 and Appendix H (BORCAL Report) for examples of calibration trends for
a few instruments.
Calibration values often are affected by the ambient temperature. Note the temperature and irra-
diance level during the calibration. This information may be useful at a future date if the responsiv-
ity of the reference instrument is correlated with temperature.
Always keep good records during field calibrations. When the data are examined at a later date,
it will be easier to address questionable data values. It is good practice to keep these calibration
records in a designated file where they can be stored each year. It is also useful to develop software
for the standard evaluation of calibration field data.
Solar Monitoring Station Best Practices 315

15.11 PHYSICAL LAYOUT OF A SOLAR-MONITORING STATION


When setting up a solar-monitoring station, it is important to keep a few principles in mind:

1. The instruments at a solar-monitoring station should have minimal shading, especially by


other instruments at the station.
2. The instruments should be easy to access for cleaning and maintenance.
3. Wiring should be shielded from the environment and animals. Exposed wiring should be
kept to a minimum.
4. Access to the station should be limited to those who need to maintain or calibrate the
instruments.
5. Growth around the station should be manageable.
6. Station layout should allow for side-by-side calibrations when possible.
An example of a Surface Radiation (SURFRAD) Network station is shown in Figure 15.10. The var-
ious monitoring equipment are well separated to minimize shading. The meteorological tower is to
the north of the solar-monitoring equipment. Most wiring is in conduit, and power cables are buried
and protected. The equipment is on sturdy concrete pads, and the equipment is tied down securely
to the pads. The vegetation is low growing and has been cleared around the instruments. The veg-
etation should not be cleared beneath the tower out to a radius of about four times the instrument
height used for upwelling irradiance.
Two examples of the Solar Radiation Research Laboratory (SRRL) station in Golden, Colorado,
are shown in Figures 15.11 and 15.12. A wide variety of instruments are used and tested at SRRL.
The tower allows testing to be done from tilted irradiance to wind speed, temperature, and relative
humidity. Ground-emitted longwave and reflected shortwave radiation are also measured along with
net radiation. Some of these instruments are not easy to reach and hence are not as likely to receive as
much cleaning as instruments that are within easy reach. This tower is located near the SRRL main

FIGURE 15.10 Photograph of the SURFRAD station near Sioux Falls, South Dakota.
316 Solar and Infrared Radiation Measurements

FIGURE 15.11 Photograph of the SRRL station in Golden, Colorado, looking north. Sample of instruments
in continuous use (foreground) and under test at SRRL. Note platform on which the instruments are mounted
is a grate that helps to prevent snow from building up around instruments.

FIGURE  15.12 Photograph of the SRRL Baseline Measurement System’s radiometer tower in Golden,
Colorado. Located away from the elevated instrumentation platform shown in Figure  15.11, this auxiliary
tower provides for a wide variety of radiometers at various orientations for continuous measurements. (From
https://midcdmz.nrel.gov/srrl_bms/.)
Solar Monitoring Station Best Practices 317

FIGURE 15.13 Photograph of the University of Oregon SRML monitoring station in Eugene, Oregon. Note
that most tilted measurements use a shield to block some of the ground-reflected irradiance. The  level of
the shield is approximately even with the middle of solar sensor. This  way some horizontal irradiance is
blocked while a similar amount of ground reflect irradiance is incident on the sensor. This relationship is an
approximation.

station, and there are people there to provide maintenance on a regular basis. The reference solar
monitoring station for the University of Oregon Solar Radiation Monitoring Laboratory (SRML)
network in Eugene, Oregon, is shown in Figure 15.13.

QUESTIONS
1. What are a few issues to consider when selecting a solar monitoring station?
2. How will sun path charts help in this process?
3. Why are site logs useful?
4. Why do some networks prefer field calibrations as opposed to changing out instruments for
calibrations?
5. What is the advantage of having all three primary solar measurements, GHI, DNI, and DHI?
6. What is the primary cleaning fluid for solar instruments?
7. Explain why the calculated DHI is larger than the measured DHI in Figure 15.6.

REFERENCES
Andreas, A., M. Dooraghi, A. Habte, M. Kutchenreiter, I. Reda, and M. Sengupta. 2018. Solar Infrared
Radiation Station (SIRS), Sky Radiation (SKYRAD), Ground Radiation (GNDRAD), and Broadband
Radiometer Station (BRS) Instrument Handbook. DOE/SC-ARM-TR-025. U.S. Department of Energy.
https://www.arm.gov/publications/tech_reports/handbooks/sirs_handbook.pdf (Accessed May 27,
2019).
Hickey, J. R. 1970. Laboratory methods of experimental radiometry including data analysis. In  Advances
in Geophysics, ed H. E. Landsberg and J. Van Mieghem, Vol.  14, Precision Radiometry, ed. A. J.
Drummond. New York: Academic Press.
318 Solar and Infrared Radiation Measurements

Long, C. N., and Y. Shi. 2008. An automated quality assessment and control algorithm for surface radiation
measurements. Open Atmospheric Science Journal 2: 23–37.
McArthur, L. J. B. 2005. Baseline Surface Radiation Network (BSRN). Operations Manual, Version 2.1.
WMO/TD-No. 1274, World Climate Research Programme, World Meteorological Organization,
Geneva, Switzerland: WCRP/WMO. https://bsrn.awi.de/other/manuals/ (Accessed May 27, 2019).
Morrison, R. 1998. Grounding and Shielding Techniques, 4th ed. New York: John Wiley & Sons.
NREL. 1993. User’s Manual for SERI_QC Software—Assessing the Quality of Solar Radiation Data. NREL/
TP-463-5608. Golden, CO: National Renewable Energy Laboratory. https://www.nrel.gov/docs/legosti/
old/5608.pdf (Accessed May 27, 2019).
Sengupta, M., A. Habte, C. Gueymard, S. Wilbert, D. Renné, and T. Stoffel. 2017. Best Practices Handbook
for the Collection and Use of Solar Resource Data for Solar Energy Applications. 2nd ed. NREL/
TP-5D00-68886. Golden, CO: National Renewable Energy Laboratory. https://www.nrel.gov/docs/
fy15osti/63112.pdf (Accessed May 27, 2019).
Sengupta, M., Y. Xie, A. Lopez, A. Habte, G. Maclaurin, J. Shelby. 2018. The National Solar Radiation Data
Base (NSRDB). Renewable and Sustainable Energy Reviews 89: 51–60. doi:10.1016/j.rser.2018.03.003.
Stoffel, T., D. Renné, D. Myers, S. Wilcox, M. Sengupta, R. George, and C. Turchi. 2010. Concentrating solar
power: Best practices handbook for the collection and use of solar resource data (CSP). NREL Report
No.  TP-550-47465. Golden, CO: National Renewable Energy Laboratory. http://www.nrel.gov/docs/
fy10osti/47465.pdf (Accessed May 27, 2019).
Wilcox, S., and W. Marion. 2008. User’s manual for TMY3 data sets. Technical Report NREL/TP-581-43156,
revised May 2008. Golden, CO: National Renewable Energy Laboratory. https://www.nrel.gov/docs/
fy08osti/43156.pdf (Accessed May 27, 2019).
16 Solar Radiation Estimates
Derived from Satellite Images
and Auxiliary Measurements

16.1 INTRODUCTION
Quantifying the temporal and spatial variability of solar radiation for developing renewable
energy applications requires data representative of the long-term conditions at a specific location.
A continuing sparsity of ground-based solar monitoring stations has resulted in typically short-term
measurements representative of a limited number of locations around the world (see Appendix L).
Meanwhile, satellite observations of large areas of the Earth can provide decades of remote
sensing data with increasingly higher spatial and temporal resolution (Stackhouse et al. 2016).
However, extrapolating the measured solar resource data to nearby locations is less accurate
than  satellite-derived hourly values once the distance from the ground station exceeds 25 km
(Zelenka et al. 1999).
Among the recommended best practices for deploying solar energy conversion systems is
the combined use of ground-based solar and meteorological measurements and available sat-
ellite-based data. Short-term, site-specific measurements can be used to identify and adjust
any systematic differences in the satellite data which can then be used to better characterize
the long-term variability of the solar resource at the location of interest (BEST PRACTICES
HNDBK).
More specifically, if long-term, high-quality data are not available, the best option for a
bankable resource assessment study is to provide a year or two of high-quality ground-based
measurements with a 10–30 yearlong satellite-derived solar radiation database. The ground-based
data are used to validate the satellite-derived database and to identify any systematic biases in
the satellite-derived database. The adjusted satellite-derived data can then be used to characterize
the long-term variability of the solar resource and can help to identify any problems with the
ground-based measurements previously missed in the initial vetting process.
Historical geostationary satellites are most suitable for modeling solar irradiance as they moni-
tor the state of the atmosphere including cloud cover with a spatial resolution near 1 km in the vis-
ible range and with a 30-minute time resolution. Newer geostationary satellites provide even higher
spatial and temporal resolution (NASA 2016). However, the view of the Earth from these satellites
limits their effectiveness at latitudes greater that 60° N or 60° S (see Figure 16.1). Polar satellites
are closer to the Earth’s surface and provide a variety of measurements that can be transformed
into surface solar irradiance values, but they pass over a particular area only once during the day,
which limits their temporal coverage at the lower latitudes. For a good compilation of satellite-
derived data from polar-orbiting satellites for areas outside the range covered by geostationary
satellites, see Karlsson et  al. (2017). These databases have good spatial resolution but limited
temporal resolution.
The geostationary satellites measure the light reflected from clouds and the Earth’s surface to
estimate the solar radiation at the Earth’s surface. With knowledge of the solar irradiance from the
Sun at the top of the atmosphere (TOA) of the Earth and the amount of solar radiation reflected,

319
320 Solar and Infrared Radiation Measurements

FIGURE  16.1 Geostationary weather satellites cover most of the globe from latitude 60°S to 60°N. This
figure  shows the coverage of five geostationary satellites. (Courtesy of Angie Skovira of Luminate, LLC,
Denver, Colorado.)

augmented by measurements that enable estimates of the aerosols and gases in the atmosphere,
reasonable estimates of the irradiance at ground level can be made. The satellites used to derive
these measurements for the irradiance models will be discussed, along with a brief description of
the Perez et al. (2002) satellite irradiance model, as well as other similar models.

16.2 GEOSTATIONARY SATELLITES


Geostationary satellites are located approximately 35,880 km (22,300 miles) above the equator in a
geosynchronous orbit. The United States has two satellites, Geostationary Operational Environmental
Satellite (GOES)-West (135° W) and GOES-East (75° W), that cover North and South America.
The European Union operates two satellites, Meteosat-9 (0°) and Meteosat-7 (57.5° E), that cover
Europe, Africa, and the Middle East. The Japanese operate MTSAT (140° E), which covers Asia and
Australia (Figure 16.1). The curvature of the earth limits the useable images to between −66° and
+66° latitude, so there is some redundancy in the satellite images and this allows some coverage even
if one satellite has problems. Russia (GOMS), China (FY-2 series), and India (InSat and KALPANA)
also have geosynchronous satellites that provide meteorological data and images.
Geosynchronous satellites have to be replaced periodically because they become outdated or
run out of fuel for stabilization. For example, GOES-West and GOES-East are the seventeenth and
sixteenth satellites in the series, respectively. Historic meteorological satellite data from the United
States, the European Union, and Japan are more complete and readily available than satellite data
from China, India, and Russia.
Solar Radiation Estimates Derived from Satellite Images and Auxiliary Measurements 321

Prior to 2019, the GOES satellites made radiance observations at 5 wavelength bands, or spectral
regions. The  visible channel (640  nanometers (nm)) had a 0.01 degree resolution (approximately
1 km); the infrared channels (3,900, 6,500, 10,700, and 12,000  nm) had 0.04 degree resolution
(approximately 4 km). The GOES-16 satellite, launched in November 2016, has replaced GOES-13
(the GOES-East satellite). GOES-16 is the first of the next series of GOES satellites and has a new
instrument called the Advanced Baseline Imager (ABI) with 5-minute coverage at 0.01 degrees
resolution for 16 channels with 6 channels in the visible and near-infrared (IR) range (Sengupta
et al. 2017). In 2019, GOES-17 became the GOES-West satellite.
The latest European METEOSAT satellites and the Japanese Himawari satellite also are following the
trend to increase the number of spectral regions measured and to obtain more frequent measurements.

16.3 DERIVING IRRADIANCE FROM SATELLITES


Many models exist to derive irradiance values from satellite data. An overview of satellite models
will be presented in this section. For a more complete summary of satellite models see Chapter 4 of
the NREL Best Practices Handbook (Stoffel et al. 2010; Sengupta et al. 2017). The empirical satel-
lite models derive a cloud index (CI) from the satellite visible-channel measurements of reflected
surface radiance and use this index to modulate a clear-sky global irradiance model of the solar
resource. Physical models use cloud properties, water vapor, aerosol optical depth, and stratospheric
ozone measurements as inputs to radiative transfer models. These calculations are based on the
physical interactions of irradiance with atmospheric constituents on a quantum scale. These interac-
tions are well understood but tracing these interactions through the atmosphere requires consider-
able computer capacity. Physical models can either produce broadband or spectral results. Empirical
models rely on correlations to obtain an estimate of incident irradiance, thus bypassing some of the
more cumbersome and time-consuming atmospheric modeling procedures.

16.3.1 PHysiCal models


Physical models can take considerable computer time and require knowledge of the distribution of
the gases, aerosols, and particles that make up the atmosphere in addition to understanding how
each constituent affects the incoming irradiance. Physical models have been useful in understanding
radiative transfer and were instrumental in uncovering the IR radiative losses (thermal offset) that
skewed the diffuse horizontal irradiance (DHI) measurements made using high-quality thermopile
pyranometers (Cess et al. 1993).
A good example of a physical model is one developed by Pinker and Ewing (1985) that divides
the solar spectrum into 12 intervals and applies a radiative transfer model to a three-layer atmo-
sphere. A primary input for this model is cloud optical depth. Pinker and Laszlo (1992) enhanced the
model, and cloud information from the International Satellite Cloud Climatology Project (ISCCP)
(Schiffer and Rossow 1983) was used to develop irradiance data for the Surface Radiation Budget
database that was created for a 2.5 degree × 2.5 degree grid (Whitlock et al. 1995). The clouds in
the ISCCP climatology are separated into low, middle, and high clouds with three different opti-
cal thicknesses. Low and middle clouds are also categorized into water and ice clouds, whereas
high clouds are always ice clouds. This  division creates 15 different cloud types for the ISCCP.
The ISCCP climatology is used for cloud input for many models (Stoffel et al. 2010). The ISCCP
values often are used by programs modeling irradiance from satellite data.
A 2.5 degree by 2.5 degree grid is quite large and encompasses large area. Researchers have used
a technique called downsizing to reduce the information to a 1.0 degree by 1.0 degree grid and are
working to obtain hourly values instead of values every three hours (Stackhouse 2016). Still, for
most solar energy applications, spatial resolution down to 0.1 degrees by 0.1 degrees is necessary
with temporal intervals of an hour or less.
322 Solar and Infrared Radiation Measurements

16.3.2 emPiRiCal models


Empirical models take less computer time to run, are easier to apply, and do not require as detailed
input information as needed by the physical models (Perez et al. 2013; Xie et al. 2016). These empirical
models are based on regression relationships between satellite observations and ground-based instru-
ment measurements. The CI and regression relationships with other meteorological data are used to
estimate the solar irradiance. Accurate clear-sky direct normal irradiance (DNI) modeling is important
for all models because these clear-sky DNI values are modulated by a cloud index (Vignola et al. 2007).
Furthermore, good atmospheric turbidity values are necessary to improve clear-sky DNI estimates.

16.3.3 global iRRadianCe


Early work by Cano et al. (1986) is based on the observation that shortwave (solar) atmospheric trans-
missivity is linearly related to the Earth’s planetary albedo (Schmetz 1989). This relation is a good first-
order approximation. Without any clouds and with a fixed albedo, the irradiance incident on the Earth’s
surface is proportional to the intensity of the reflected irradiance as measured using the counts in the sat-
ellite image pixel. The initial task in modeling surface irradiance is to assign the image pixel to a precise
ground-based location. As the satellite wobbles in space, this limits the precision to which a pixel can
be assigned. The original Ineichen and Perez (1999), Ineichen et al. (2000), and Perez et al. (2002) pro-
cedure for extracting irradiance data is the basis for the following general discussion, and much of this
material is extracted from Vignola and Perez (2004). A more detailed description can be found in Perez
et al. (2002). As with many models, improvements are made over time, and a third-generation Perez
model was validated as of 2012 with the latest improvements announced in 2015 (Perez et al. 2015).
There are two distinct steps for this process:

1. Pixel-to-CI conversion
2. CI to global irradiance conversion

16.3.4 Pixel-to-Cloud index ConveRsion


Individual pixel brightness values in the satellite image are stored as counts. The pixel counts are first
normalized by the cosine of the solar zenith angle to account for the solar geometry. This normalized
pixel is then gauged against the satellite pixel’s dynamic range at the chosen location to estimate a
CI. The lowest pixel count during a series of baseline images is assumed to be the ground brightness
under clear-sky conditions.
An additional normalization is necessary to adjust for a secondary atmospheric air mass effect
and for the “hot spot” (Zelenka et al. 1999). The hot spot is cause by the Sun–satellite angle and
incorporates both atmospheric backscatter brightness intensification and the fact that the ground
surface becomes brighter as the Sun–satellite angle diminishes due to the reduction of ground shad-
ows seen by the satellite (e.g., Pinty and Verstraete 1991).
The dynamic range represents the range of values a normalized pixel can assume at a given location
from its lowest (darkest pixel, that is, clearest conditions) to its highest value (brightest pixel, that is,
cloudiest conditions). A record of the normalized pixel counts at a given location is kept to track the
changes in the dynamic range. While the upper bound of the range remains relatively constant except
for the drift in the satellite’s calibration, the lower bound evolves over time as the local ground albedo
changes (chiefly from changes in snow, moisture, and vegetation). Obtaining an accurate lower bound
that represents the true ground albedo is difficult, especially during periods with long-term cloudi-
ness. This condition can be especially important during the winter months when snow covers the
ground and significantly reduces the dynamic range. Techniques using the IR satellite image are
being used to better distinguish between snow on the ground and clouds (Perez et al. 2010).
In the Perez procedure, a sliding time frame is used to determine this lowest bound. The lower
bound is determined as the average of the 10 lowest pixels in the sliding time frame. A secondary
Solar Radiation Estimates Derived from Satellite Images and Auxiliary Measurements 323

normalization is applied to the lower bound of the dynamic range. Subtracting the lower bound from
the normalized pixel count and dividing this number by the normalized dynamic range defines the
CI. The CI then can vary from 0 (clearest) to 1 (cloudiest).

16.3.5 Cloud index to global HoRiZontal iRRadianCe ConveRsion


In the Perez procedure global horizontal irradiance (GHI) is determined by

GHI = ( 0.02 + (1 − CI ) ) ⋅ GHI c (16.1)

where GHIc is the clear-sky global irradiance per Kasten (1984). GHIc is adjustable for broadband
turbidity as quantified by the Linke turbidity coefficient (Kasten 1980) and ground elevation.

GHI c = ( 0.0000509 ⋅alt + 0.868 ) ⋅ I 0

− ( 0.0000392 ⋅ alt + 0.00387 ) ⋅ am 


  (16.2)
⋅ cos sza ⋅ exp  
⋅ ( exp( −alt / 8000 ) + exp ( −alt / 1250 ) ⋅ (TL − 1) ) 

(
⋅ exp 0.01 ⋅ am1.8 )
where:
Io is the extraterrestrial normal incident irradiance
sza is the solar zenith angle
am is the elevation-corrected air mass
TL is the Linke turbidity coefficient
alt is the ground elevation in meters.

16.3.6 diReCt iRRadianCe


Direct normal irradiance (DNI) is obtained, as GHI, by modifying the clear-sky direct irradiance
(DNIc). An example of the clear-sky model (Ineichen and Perez 2002) is

0.83 ⋅ I 0 ⋅ exp ( −0.09 ⋅ am ⋅ [TL − 1]) 


 
 
DNI c = min     (16.3)
0.196
⋅  0.8 +  , ( GHI c − DHI c ) / cos ( sza ) 
  exp [ −alt / 8000 ]  

where DHIc is the minimum clear-sky diffuse irradiance given by

DHI c = GHI c ⋅ {0.1 ⋅ [1 − 2 ⋅ exp ( −TL )]} ⋅ {1 / [ 0.1 + 0.882 / exp( −alt / 8000) ]} (16.4)

Unlike GHI, the direct modulating factor is not derived from CI but from global, using a GHI-to-
DNI model. One quasi-physical model, DIRINT (Perez et al. 1992), is based on the DISC model
developed by Maxwell (1987) that uses a clear-sky model that is a function of the clearness index kt.
In the DIRINT model, a modified clearness index is used along with atmospheric water vapor and
the change in clearness index.
DNI is obtained from DNI = DNIc × DIRINT(GHI)/DIRINT(GHIc), where DIRINT(GHI) is the
DIRINT model using the input global irradiance GHI. Other global–beam irradiance models also
have been used (Ineichen 2008).
324 Solar and Infrared Radiation Measurements

16.3.7 diffuse iRRadianCe


The DHI is obtained by subtracting the DNI times the cosine of the solar zenith angle from the GHI.
When the DNI value is very small, DHI is approximately equal to GHI and the uncertainty in the DHI
value is the same as the percent uncertainty in the GHI value. When DNI is large, then the DHI value
is small and the percent uncertainty in the DHI value is large compared with the percent uncertainty
in the GHI value. When it is neither clear nor totally overcast, the uncertainty in the DHI value can be
very large. Therefore, it is important to precisely state the conditions when specifying the uncertainty
in the DHI value.

16.3.8 tilted iRRadianCe


Like users of other solar data, the main interest is the solar radiation incident on the collector. Satellite-
based models have now been developed for looking at irradiance on tilted surfaces (Xie and Sengupta
2018). In addition, models are being developed and tested that produce spectral irradiance values from
satellite data by combining physical models with empirical models to produce spectral databases that
can be used to evaluate photovoltaic (PV) module performance (Xie et al., 2018a; 2018b).

16.4 STATUS OF SATELLITE IRRADIANCE MODELS


Satellite irradiance models provide a reasonable estimate of the location’s solar resource over vari-
ous time scales. The biases in annual satellite-derived GHI values range between −8 and +8 percent
(Šúri and Cebecauer 2014). Satellite-derived GHI hourly values have an uncertainty ranging from
±7 to ±30 percent when compared to ground-based GHI measurements. Of course, the differ-
ent perspectives of ground-based and satellite views account for a good portion of this random
error along with cloud cover, and to a lesser extent, snow cover and aerosols dynamics (Sengupta
et al., 2017). Improvements to satellite irradiance models and expanded satellite data with improved
spatial and temporal resolution should continue to reduce uncertainties in satellite-derived solar
resources values. There  is some room to modify models to better simulate statistical irradiance
distributions while reducing systematic biases. In  addition, considerable advancement has been
made by combining physical models with empirical correlations.
As with any modeling effort, it is important to be mindful of the accuracy of the data from which
the model is derived and validated. The estimated uncertainty of sub-hourly ground-based GHI is
±3 percent at best (see Chapter 5), so a claim that the data have a bias of 4 percent when compared
to ground-based measurements is problematic. When evaluating the uncertainty in satellite-derived
data, any effect of systematic biases in the ground-based measurement should be considered. All
measurements have a random uncertainty and the possibility of systematic bias. When an irradiance
measurement has an uncertainty of ±3 percent, some of this uncertainty is random and some of the
uncertainty is from a systematic deviation (bias) from reference measurements. When uncertainties
and biases are traced back to radiometric references, then a solid basis can be created for the uncer-
tainty and bias calculations (see Appendix A).
The  uncertainty in the National Aeronautics and Space Administration (NASA)-modeled
1-degree gridded data compared with high-quality Baseline Surface Radiation Network (BSRN)
data is given in Table 16.1 (Stoffel et al. 2010). Although the mean bias error (MBE) appears small
overall, it can vary several percent depending on the site examined. For example, the DNI MBE
varies from –15.7 percent above 60° latitude (north or south) to 2.4 percent below 60°. It is difficult
to compare ground-based BSRN-site-measured data with satellite-derived data on a 1-degree grid
because of the differences in the areal coverage and differences in timing between the “snapshot”
satellite observations and the averaged surface measurements.
According to Perez et  al. (1987), satellite-based GHI estimates are accurate from 10 to 12
percent. According to Zelenka et  al. (1999), the target-specific comparison with ground-based
Solar Radiation Estimates Derived from Satellite Images and Auxiliary Measurements 325

TABLE 16.1
Uncertainty in NASA-Modeled Satellite
Data for Monthly Averaged Values
Measurement MBE (%) RMS (%)
GHI −0.0 10.3
DHI 7.5 29.3
DNI −4.1 22.7

Note: The MBE and RMS errors are less for irradi-


ance values obtained between ±60° latitude
and larger for locations closer to the poles.

observations will have a root mean square error (RMSE) of at least 20 percent; the time-specific
area-wide accuracy is 10 to 12 percent on an hourly basis.
The RMSE decreases with averaging time as expected. An hourly RMSE for one site will decrease
from 20 to 25 percent for hourly estimates to 10 to 12 percent for daily estimates and down to 5 to
10 percent or less for monthly mean daily estimates. Mean bias errors generally range from −5 to
+5 percent, with most studies reporting absolute MBEs in the 2 to 3 percent range. Considering that
the best GHI measurements have an estimated expanded uncertainty of ±3 percent, it is necessary
to be sure of what is being compared and cognizant of the uncertainties in the comparison measure-
ments. The uncertainty in the GHI estimates increase if there is snow on the ground, the albedo in
the image pixel varies considerably, or the terrain creates significant shadowed areas.
DNI and DHI have larger fractional RMSE and MBE than GHI estimates with the DNI, being
up to twice the RMSE compared with the GHI estimate uncertainties and the DHI having a slightly
larger RMSE than DNI.
With satellite images not centered on the hour, but at other times, the correspondence between
ground-based data and data derived from satellite is not always easy to make. For example, when an
image is taken at 9:15, it has to be shifted to 9:00 to be compatible with other meteorological data.
The satellite-derived data in the National Solar Radiation Data Base is a weighted average of irradiance
values that correspond to the hourly averaged meteorological data (Habte et al. 2017). This smoothing
reduces some of the variability in the data values but gives an overall better representation when the
irradiance data are used with other meteorological values for system performance calculations.
Satellite-derived values near sunrise and sunset have higher uncertainties. This  difference is
caused by two factors: (1) the large incident angles and (2) the fact that sometimes the satellite images
are taken when the Sun is below the horizon but there is some irradiance during the hour. As an
illustration of the type of problem that can occur, if sunrise is 6:30 and the satellite image is taken at
6:15, there will be no irradiance recorded for the time period when there really is a finite GHI value
between 6:30 and 7:00. Using an averaging method with the 7:15 value will yield a value at 7:00. Of
course, the irradiance values are relatively small and large uncertainties do not significantly affect
the usefulness of the data, but it is important to understand the limitations of the data values obtained.

16.5 COMMENTS ON MODELING AND MEASUREMENT


As with any modeling effort, it is important to know the estimated uncertainty of the measurements
used to develop and validate a model. Specifically, the systematic errors, which are type B uncertain-
ties, are different for an Eppley PSP pyranometer, a LI-COR LI-200 pyranometer, a Kipp & Zonen
CM 21 pyranometer or other manufacturer and model. A model developed using one type of instru-
ment will work best with that type of instrument if the systematic biases are not removed. Therefore,
care should be taken when developing a model to use data where known systematic biases of the
326 Solar and Infrared Radiation Measurements

instruments have been removed. If these biases are not removed, then it is likely that the model will
agree best with the measurements made with the instruments used to develop the model, but these
are not necessarily the best measurements.
If an instrument is fully characterized, it is possible to remove most of the systematic errors
in the measurements, assuming adequate ancillary data are available. Alternatively, it is possible
to use very accurate, and usually more expensive, instruments like absolute cavity radiometers
(ACRs) to make the measurements. It should be possible to make significant refinement to models
based on very accurate measurements. However, to develop models with data taken with lesser-
quality instruments, the systematic biases associated with these less accurate instruments have to be
accounted for or else they will skew the modeled result. Alternatively, when using measured data
without correcting for systematic biases, the modeled results will be skewed. Therefore, it is impor-
tant to know the instrument’s measurement performance characteristics. Important information such
as temperature dependence, calibration uncertainty and frequency, and known systematic biases do
affect the usefulness of the data and the accuracy of any results.
Deriving irradiance data from satellite images is dependent on the characteristics of the local
albedo and also may be affected by local physical and environmental conditions. Although models
using satellite images are validated for a variety of sites, there is always the possibility that condi-
tions at a given site are such that the satellite-derived data may be skewed. For example, the satel-
lite pixel encompasses two types of surfaces, for example, a salt bed and a lake. Any wobble in the
pointing of the satellite will pick up more albedo from the salt bed or the lake. This wobble makes
it difficult to determine the surface albedo and hence decreases the accuracy of the modeled data.
Therefore, it is always useful to have some ground-based measurements to validate the satellite-
derived data for the specific location under study.

QUESTIONS
1. What type of satellites are used to estimate solar radiation?
2. Briefly describe the process of obtaining global solar radiation data from satellite images.
3. Compare and contrast the strengths and weaknesses of solar radiation values obtained
using satellite images and data from ground-based stations.

REFERENCES
Broesamle, H., H. Mannstein, C. Schillings, and F. Trieb. 2001. Assessment of solar electricity potentials in
North Africa based on satellite data and a geographic information system. Solar Energy 70: 1–12.
Cano, D., J. M. Monget, M. Aubuisson, H. Guillard, N. Regas, and L. Wald. 1986. A method for the determina-
tion of global solar radiation from meteorological satellite data. Solar Energy 37: 31–39.
Cess, R. D., S. Nemesure, E. G. Dutton, J. J. DeLuisi, G. L. Potter, and J. Morcrette. 1993. The impact of
clouds on shortwave radiation budget of the surface-atmosphere system: Interfacing measurements and
models. Journal of Climate 6: 308–316.
Habte, A., M. Sengupta, and A. Lopez. 2017. Evaluation of the National Solar Radiation Database (NSRDB
Version 2): 1998–2015. NREL/TP-5D00-67722. National Renewable Energy Laboratory, Golden, CO.
https://www.nrel.gov/docs/fy17osti/67722.pdf (Accessed May 29, 2019).
Ineichen, P. 2008. Comparison and validation of three global-to-beam irradiance models against ground mea-
surements Solar Energy 82: 501–512.
Ineichen, P., and R. Perez. 1999. Derivation of cloud index from geostationary satellites and application to the
production of solar irradiance and daylight illuminance data. Theoretical and Applied Climatology 64:
119–130.
Ineichen, P., and R. Perez. 2002. A new airmass independent formulation for the Linke turbidity coefficient.
Solar Energy 73 (3): 151–157.
Ineichen, P., R. Perez, M. Kmiecik, and D. Renne. 2000. Modeling direct irradiance from GOES visible chan-
nel using generalized cloud indices. In Proceedings of the 80th AMS Annual Meeting, Long Beach, CA.
Solar Radiation Estimates Derived from Satellite Images and Auxiliary Measurements 327

Karlsson, K.-G., K. Anttila, J. Trentmann, M. Stengel, J. F. Meirink, A. Devasthale, T. Hanschmann, et al. 2017.
CLARA-A2: CM SAF cLoud, Albedo and Surface RAdiation Dataset from AVHRR data—Edition 2.
Satellite Application Facility on Climate Monitoring. doi:10.5676/EUM_SAF_CM/CLARA_AVHRR/V002.
Kasten, F. 1980. A simple parameterization of two pyrheliometric formulae for determining the Linke turbid-
ity factor. Meteorologische Rundschau 33: 124–127.
Kasten, F. 1984. Parametriesirung der Globalstahlung durch Bedeckungsgrad und Trubungsfaktor. Annalen
der Meteorologie (Neue Folge) 20: 49–50.
National Aeronautics and Space Administration. 2016. GOES-R Series, Concept of Operations (CONOPS).
410-R-CONOPS-0008, Version 2.9. Washington, DC. https://www.goes-r.gov/syseng/docs/CONOPS.
pdf. (Accessed May 28, 2019).
NASA. See National Aeronautics and Space Administration. Maxwell, E. 1987. A quasi-physical model for
converting hourly global horizontal to direct normal insolation. SERI/TR-215-3087. 68 pages. National
Renewable Energy Laboratory, Golden, CO. https://www.nrel.gov/docs/legosti/old/3087.pdf (Accessed
May 29, 2019).
Perez, R., T. Cebecauer, and M. Suri. 2013. Semi-empirical satellite models. In J. Kleissl (Ed.), Solar Energy
Forecasting and Resource Assessment (pp. 21–48). Waltham, MA: Academic Press.
Perez, R., P. Ineichen, E. Maxwell, R. Seals, and A. Zelenka. 1992. Dynamic global-to-direct irradiance con-
version models. ASHRAE Transactions-Research Series 354–369.
Perez, R., P. Ineichen, K. Moore, M. Kmiecik, C. Chain, R. George, and F. Vignola. 2002. A new operational
satellite-to-irradiance model. Solar Energy 73 (5): 307–317.
Perez, R., S. Kivalov, A. Zelenka, J. Schlemmer, and K. Hemker Jr. 2010. Improving the performance of
satellite-to-irradiance models using the satellite’s infrared sensors. Proceedings of the ASES Annual
Conference, Phoenix, AZ. http://asrc.albany.edu/people/faculty/perez/2010/ir.pdf (Accessed May 29,
2019).
Perez, R., J. Schlemmer, K. Hemker, Jr., S. Kivalov, A. Kankiewicz, and C. Gueymard. 2015. Proc. 42nd IEEE
PV Specialists Conference, New Orleans, LA. http://asrc.albany.edu/people/faculty/perez/2015/V4.pdf
(Accessed May 29, 2019).
Perez, R., R. Seals, P. Ineichen, R. Stewart, and D. Menicucci. 1987. A new simplified version of the Perez dif-
fuse irradiance model for tilted surfaces. Description performance validation. Solar Energy 39: 221–232.
Pinker, R., and I. Laszlo. 1992. Modeling surface solar irradiance for satellite applications on a global scale.
Journal of Applied Meteorology 31: 194–211.
Pinker, R., and J. Ewing. 1985. Modeling surface solar radiation: Model formulation and validation. Journal
of Climate and Applied Meteorology 24: 389–401.
Pinty, B., and M. M. Verstraete. 1991. Extracting information on surface properties from bidirectional reflec-
tance measurements. Journal of Geophysical Research 96: 2865–2874.
Schiffer, R. A., and W. B. Rossow, 1983. The International Satellite Cloud Climatology Project (ISCCP):
The first project of the World Climate Research Programme. Bulletin of the American Meteorological
Society 64: 779–784.
Schmetz, J. 1989. Towards a surface radiation climatology: Retrieval of downward irradiances from satellites.
Atmospheric Research 23: 287–321.
Sengupta, M., A. Habte, C. Gueymard, S. Wilbert, D. Renné, and T. Stoffel. 2017. Best Practices Handbook
for the Collection and Use of Solar Resource Data for Solar Energy Applications, 2nd ed. NREL/
TP-5D00-68886. https://www.nrel.gov/docs/fy18osti/68886.pdf (Accessed May 29, 2019).
Stackhouse, P. W., Jr., P. Minnis, R. Perez, J. C. Mikovitz, J. Schlemmer, B. Scarino, T. Zhang, S. J. Cox, M.
Sengupta, and K. Knapp, 2016. An Assessment of New Satellite Data Products for the Development of a
Long-term Global Solar Resource at 10–100 km. Proceedings of the ASES National Solar Conference,
San Francisco, CA, July 10–13, 2016. American Solar Energy Society, Boulder, CO.
Stoffel, T., D. Renné, D. Myers, S. Wilcox, M. Sengupta, R. George, and C. Turchi. 2010. Concentrating Solar
Power best practices handbook for the collection and use of solar resource data. NREL/TP-550-47465.
National Renewable Energy Laboratory, Golden, CO. https://www.nrel.gov/docs/fy10osti/47465.pdf
(Accessed May 29, 2019).
Šúri, M., and T. Cebecauer. 2014. Satellite-based solar resource data: Model validation statistics versus user’s
uncertainty. Proceedings of the 43th ASES National Solar Conference, July 6–10, San Francisco.
American Solar Energy Society, Boulder, CO.
Vignola, F., and R. Perez. 2004. Solar resource GIS data base for the Pacific Northwest using satellite
data—final report. Project ID DE-FC26-00NT41011. University of Oregon, Eugene, OR. http://solardat.
uoregon.edu/download/Papers/SolarResourceGISDataBaseforthePacificNorthwestusingSatelliteData--
FinalReport.pdf (Accessed May 30, 2019).
328 Solar and Infrared Radiation Measurements

Vignola, F., P. Harlan, R. Perez, and M. Kmiecik. 2007. Analysis of satellite derived beam and global solar
radiation data. Solar Energy 81: 768–772.
Whitlock, C. H., T. P. Charlock, W. F. Staylor, R. T. Pinker, I. Laszlo, A. Ohmura, H. Gilgen, et al. 1995. First
global WCRP shortwave surface radiation budget dataset. Bulletin of the American Meteorological
Society 76: 905–922.
Xie, Y. and M. Sengupta, Submitted. 2018. A  fast all-sky adiation model for solar applications with nar-
rowband irradiances on tilted surfaces (FARMS-NIT): Part I. The clear-sky model. Solar Energy 174:
691–702.
Xie, Y., M. Sengupta, and M. Dooraghi. 2018b. Assessment of uncertainty in the numerical simulation of solar
irradiance over inclined PV panels: New algorithms using measurements and modeling tools. Solar
Energy. doi:10.1016/j.solener.2018.1002.1073.
Xie,Y., M. Sengupta, and C. Deline. 2018a. Recent advancements in the numerical simulation of surface irra-
diance for solar energy applications. Proceedings of the IEEE 44th Photovoltaic Specialist Conference
(PVSC), June 25–30, 2017, Washington, DC. pp. 116–119. https://www.nrel.gov/docs/fy17osti/67804.pdf
(Accessed May 29, 19).
Xie, Y., M. Sengupta, and J. Dudhia. 2016. Fast all-sky radiation model for solar applications (FARMS):
Algorithm and performance evaluation. Solar Energy 135: 435–445. doi:10.1016/j.solener.2016.06.003.
Zelenka, A., R. Perez, R. Seals, and D. Renné. 1999. Effective accuracy of satellite-derived irradiance.
Theoretical and Applied Climatology 62: 199–207.
Appendix A: Measurement
Uncertainty Principles

A.1 INTRODUCTION
All physical measurements and model estimates based on physical measurements include some
degree of uncertainty. The confidence in the measurements or model predictions is associated with
the uncertainty in the measurement or model. The greater the uncertainty, the more likely that value
of what is being measured will be differ from the best estimate of the measurand1 and it is necessary
to take into consideration the effect of this uncertainty on the conclusions drawn from use of this
value. Data without a stated estimate of the associated uncertainty are incomplete because the val-
ues cannot be readily evaluated with an associated level of confidence. For example, the estimated
uncertainty and stated confidence interval are essential information needed to assess the potential
risks associated with solar system performance predictions. That is, solar and infrared radiation
measurements along with estimated uncertainty bounds are required to provide a bankable dataset
from which funders and developers can assess the reliability of the performance estimates.
Measurement uncertainty is a rather complex subject that has gained more attention due to
the growing financial risks of applying measurements to larger projects incorporating a growing
list of advanced technologies. Central to the science of metrology, measurement uncertainty is an
evolving subject that needs to be done in a well defined and consistent manner by scientists, engi-
neers, analysts, policy makers, and other stakeholders who rely on measurements.
The purpose of this appendix is to provide an overview of the basic principles and terminology
used in measurement uncertainty as they apply to solar and infrared radiation measurements.
Detailed descriptions for applying these principles to radiometry can be found in the instrument-
specific chapters of this book. More detailed information on the advanced aspects of estimating
measurement uncertainty can be found in the references listed at the end of this appendix.

A.2 BACKGROUND
The quantity to be measured is called the measurand. No measurement of a measurand is exact.
In fact, a measurand cannot be completely described without an infinite amount of information.2
The measured value depends upon the circumstances under which it is being measured, the equip-
ment being used, and the methodology used to make the measurement (i.e., the number of samples
taken, the operating procedure, the skill of the operator, etc.). All these aspects introduce some
degree of uncertainty. Therefore, when making a measurement, the value and uncertainty obtained
for the measurand will involve probability and yield similar results only if the measurement is
repeated under identical circumstances. This is shown in Figure A.1 where the value and uncertainty
obtained by one measurement will be within the bounds of an estimated measurement uncertainty.
As described in more detail here, the measurement process involves many potential sources of uncer-
tainty generally attributed to the act of measurement, the instrumentation used for the measurement,
the equipment used to record the measurement, the installation and operation of the measurement
system, and the environment in which the measurement takes place. Uncertainty estimates provide

1 See section A.6 for the terminology used in this book for measurement uncertainty.
2 Guide to the Expression of Uncertainty in Measurement (GUM), Annex D, section D.1.1 (JCGM 2008).

329
330 Appendix A

FIGURE A.1 The value of a measurand is considered to be the best estimate and is within the expanded mea-
surement uncertainty bounds for a random distribution of measurements with an infinite number of degrees
of freedom.

a guide to the reliability or confidence in the measured value. The confidence level in the data is
especially important if the intent is to use the data in developing or comparing models or to persuade
investors to finance a project. Of equal importance to quoting an uncertainty is having a uniform
way to obtain the estimated uncertainty so comparisons can be made among datasets or analyses.
A typical example where more specificity in the estimated uncertainty is needed occurs in the
reports made on the changes in performance of photovoltaic systems over time. A wide variety of
inconsistent results have been reported and many of these reports use different measuring instru-
ments and/or different ways to analyze the uncertainty of the measurements. A standard method to
analyze the uncertainty in these reports in needed.
Over the years, methodologies to report uncertainties in measurements have become more
standardized and have gained wider acceptance in the solar radiation monitoring community.
According to the third edition of the International Vocabulary of Metrology (JCGM 2012), there
are two approaches that have been central to the development of measurement uncertainty analysis.
They stem from the need to define truth from a measurement perspective. The traditional Error
Approach relied on a true quantity value (i.e., what we seek to learn about a quantity from the
measurement process) that was considered unique and, in practice, unknowable. In the currently
accepted Uncertainty Approach, it is recognized that there is not a single true quantity value, but
rather a best estimate of the measurand with an uncertainty associated with this value. Given the
appropriate environment condition, the best available instruments, and the best practices, a set of
values for the measurand is obtained. One value is not more accurate than the other and the distribu-
tion of the values should be consistent with the calculated uncertainty.3
Currently, the accepted methodology for estimating measurement uncertainty is given by
the Guide to the Expression of Uncertainty in Measurements (GUM; JCGM 2008). First pub-
lished in 1995, the GUM is the result of cooperation among the Bureau International des Poids
et Mesures (BIPM), the International Electrotechnical Commission (IEC), the International

3 Fundamental constants are considered exceptions to this approach because these quantities are considered to have a
single true quantity value.
Appendix A 331

Federation of Clinical Chemistry (IFCC), the International Organization for Standardization


(ISO), the International Union of Pure and Applied Chemistry (IUPAC), the International Union
of Pure and Applied Physics (IUPAP), and the International Organization of Legal Metrology
(OIML).
The GUM uses the Uncertainty Approach that includes this note 4 about the true quantity value:

When the definitional uncertainty associated with the measurand is considered to be negligible com-
pared to the other components of the measurement uncertainty, the measurand may be considered to
have an “essentially unique” true quantity value. This is the approach taken by the GUM and associated
documents, where the word “true” is considered to be redundant.

A.3 ESTIMATING MEASUREMENT UNCERTAINTY


This section provides an overview of the approach used for estimating measurement uncertainty
in the instrument-specific chapters of this book. The reader is encouraged to review this appendix
to increase their familiarity with the terminology and methods for estimating uncertainty as it
applies to radiometry. A spreadsheet for computing the uncertainties for various types of radiom-
eters is also available from the National Renewable Energy Laboratory: https://midcdmz.nrel.gov/
radiometer_uncert.xlsx.
GUM (JCGM 2008) is an internationally recognized method for estimating measurement uncer-
tainty and is used in this book. There are six elements to the process for determining the expanded
uncertainty for a specified coverage interval:

1. Define the measurement equation (model) for the measurement system


2. Identify the sources of uncertainty
3. Compute the standard uncertainty (u) for each variable in the measurement equation
4. Compute the sensitivity coefficients (c) for each variable in the measurement equation
5. Determine the combined standard uncertainty (uc)
6. Compute the expanded uncertainty (U) for the appropriate coverage factor (k).

Application of the above elements for estimating measurement uncertainty of pyrheliometers,


pyranometers and pyrgeometers can be found in the appropriate chapters of this book.
Detailed information used by the NREL spreadsheet for calculating uncertainties of pyrheli-
ometers, pyranometers, and pyrgeometers can be found in the appropriate chapters of this book.
Following the above steps, a simplified application of the GUM approach is presented here for esti-
mating the expanded measurement uncertainty of a thermopile-based pyrheliometer and pyranom-
eter and a photodiode-based pyranometer used for operational monitoring of broadband direct
normal (DNI) and global horizontal irradiances (GHI) respectively. See section A.6 for descriptions
of the terminology used in this example.

Step 1. The simplest conversion of a radiometer output voltage to irradiance can be written,

E = V /Rs (A.1)

where:
E = solar irradiance [Wm−2]
V = radiometer output voltage [µV]
Rs = radiometer calibration responsivity [µV/Wm−2]

4 See note 3 in subsection 2.11, page 36 (JCGM, 2012).


332 Appendix A

TABLE A.1
Pyrheliometer Estimated Measurement Uncertainty (1-Minute Averages of 2-Second
Samples: Direct Normal Irradiance (DNI))
Note Source Estimated Uncertainty (EU ±%) Standard Uncertainty (u ±%)
1 Calibration to WRR (for sza = 45°) 1.00 0.577
2 Nonlinearity 0.20 0.115
3 Temperature dependence 0.50 0.289
4 Spectral response 1.00 0.577
5 DC voltage measurement @ 0.24 0.139
1000 µV
6 Nonstability 0.10 0.058
Combined uncertainty—uc (±%) 0.886
Expanded uncertainty—U95 (±%) 1.738
Notes:
1. Based on 11 years of calibration results for seven Model CH1/CHP1 pyrheliometers at NREL (Myers et  al. 2002)
(WRR = World Radiometric Reference; sza = solar zenith angle).
2. From manufacturer’s specifications for Model CHP1.
3. From manufacturer’s specifications for Model CHP1.
4. See Chapter 4 for a description of the spectral response contribution.
5. Manufacturer specifications for Model CR3000 data logger DC Voltage Accuracy: +/−0.09% of Reading + Offset (−40°C
to +85°C) where the Offset = 1.5 × Basic Resolution + 1.0 µV and the Basic Resolution = 0.33 µV @ +/−20 mV range.
Therefore, for a Reading of 1,000 µV (corresponding to 100 Wm−2 as measured by a radiometer with a responsivity of
10 µV/Wm−2), the DC Voltage Accuracy can be computed: Offset = (1.5 × 0.33 + 1) µV = 1.495 µV = 0.15% of Reading
DC Voltage Accuracy is then 0.09% + 0.15% = 0.24%.
6. Stability of Model CH1 pyrheliometers over a 9-year period at NREL has been demonstrated to be within the variability
of the calibration process.

Step 2. Lists of potential sources of uncertainty are shown for each radiometer type
in Tables  A.1  – A.3. The  sources of uncertainty and their estimated contribution, or
Estimated Uncertainty (EU) shown in the tables, are based on information available from
ISO 9060:2018 (see Chapters 4 and 5), the radiometer manufacturer’s performance specifi-
cations, and engineering judgments5. Given the numerous possible measurement consider-
ations, the lists are intended to serve as an illustration of the process rather than a rigorous
analysis.
Step 3. The EU values shown in the tables are considered as Type B, their distribution is
rectangular, and the degrees of freedom (DF) is assumed to be 1000. The resulting stan-
dard uncertainty (u) values for each source can be computed as:

u = EU /√ 3 (A.2)

Step 4. The sensitivity coefficients (c) for each variable in the measurement equation are
determined from the partial derivative with respect to each variable (Taylor and Kuyatt
1994). For this example, these coefficients are equal to 1.
Step 5. The Combined Uncertainty (uc) is determined by summing the standard uncertain-
ties in quadrature:

5 Engineering judgements include calibration reports such as those available from NREL: https://aim.nrel.gov/borcal.html
(Accessed June 4, 2019).
Appendix A 333

TABLE A.2
Thermopile-Based Pyranometer Estimated Measurement Uncertainty (1-Minute Averages
of 2-Second Samples: Global Horizontal Irradiance (GHI))
Note Source Estimated Uncertainty (EU ±%) Standard Uncertainty (u ±%)
1 Calibration to WRR (for sza = 45°) 2.00 1.155
2 angular response 1.00 0.577
3 thermal offset 0.45 0.260
4 Nonlinearity 0.20 0.115
5 Temperature dependence 0.50 0.289
6 Spectral response 1.00 0.577
7 DC voltage measurement @ 1000 µV 0.24 0.139
8 Nonstability 0.24 0.115
Combined uncertainty—uc (±%) 1.482
Expanded uncertainty—U95 (±%) 2.905
Notes:
1. Based on 7 years of calibration results for four Model CMP22 pyrheliometers at NREL (Myers et al. 2002) (WRR = World
Radiometric Reference; sza = solar zenith angle).
2. From results of outdoor calibration of CMP22 S/N 080018 at NREL.
3. From nighttime outdoor measurements from CMP22 S/N 080018 at NREL.
4. Based on manufacturer specifications for CMP22.
5. Based on manufacturer specifications for CMP22.
6. See Chapter 5 for a description of the spectral response contribution.
7. Manufacturer specifications for Model CR3000 data logger DC Voltage Accuracy: +/−0.09% of Reading + Offset (−40°C
to +85°C) where the Offset = 1.5 × Basic Resolution + 1.0 µV and the Basic Resolution = 0.33 µV @ +/−20 mV range.
Therefore, for a Reading of 1,000 µV (corresponding to 100 Wm−2 as measured by a radiometer with a responsivity of
10 µV/Wm−2), the DC Voltage Accuracy can be computed: Offset = (1.5 × 0.33 + 1) µV = 1.495 µV = 0.15% of Reading
DC Voltage Accuracy is then 0.09% + 0.15% = 0.24%.
8. Stability of Models CMP21/22 over a 9-year period at NREL has been demonstrated to be within the variability of the cali-
bration process.

uc = ∑(c * u) 2
(A.3)

Step 6. The expanded uncertainty (U) with a 95 percent confidence interval (U95) is found
by applying the coverage factor (k = 1.96) associated with DF:

U 95 = 1.96 × uc (A.4)

The  values shown in Tables A.1  – A.3 are indicative of radiometers that are properly installed
and maintained, including regular calibration. These abbreviated examples serve only as an intro-
duction to the more detailed analyses available in the chapters addressing each specific type of
measurement.
Uncertainties often come with qualifying conditions, especially when quoted as percentages.
For example, uncertainties can be specified for an angle of incidence of 45º or a range of 30º to 60º.
However, there are many occasions when the incident angle is greater than 60º where the percent
uncertainties can be greater. To cover more extreme conditions uncertainties are often quoted in
terms of W/m2 as well as percent.
334 Appendix A

TABLE A.3
Photodiode-Based Pyranometer Estimated Measurement Uncertainty (1-Minute Averages
of 2-Second Samples: Global Horizontal Irradiance (GHI))
Note Source Estimated Uncertainty (EU ±%) Standard Uncertainty (u ±%)
1 Calibration to WRR (for sza = 45) 3.00 1.732
2 Angular response 2.60 1.501
3 Thermal offset 0.0 0.0
4 Nonlinearity 1.00 0.577
5 Temperature dependence 0.30 0.173
6 Spectral response 3.00 1.732
7 DC voltage measurement @ 0.78 0.450
1000 µV
8 Nonstability 0.80 0.462
Combined uncertainty—uc (±%) 3.005
Expanded uncertainty—U95 (±%) 5.891
Notes:
1. Based on 15 years (40 events) of calibration results for LI-200SA  (S/N PY1755) at NREL (Myers et  al. 2002)
(WRR = World Radiometric Reference; sza = solar zenith angle).
2. From results of outdoor calibration of LI-200SA (S/N PY1755) at NREL.
3. From nighttime outdoor measurements from CMP22 S/N 080018 at NREL.
4. See Chapter 5 for a description of the stability of silicon-based detectors.
5. See Chapter 5 for a description of the temperature dependence of silicon-based detectors.
6. See Chapter 5 for a description of the spectral response contribution of silicon-based detectors.
7. Manufacturer specifications for Model CR3000 data logger DC Voltage Accuracy: +/−0.09% of Reading + Offset (−40°C
to +85°C) where the Offset = 1.5 × Basic Resolution + 1.0 µV and the Basic Resolution = 0.33 µV @ +/−20 mV range.
Therefore, for a Reading of 1,000 µV (corresponding to 100 Wm−2 as measured by a radiometer with a responsivity of
10 µV/Wm−2), the DC Voltage Accuracy can be computed: Offset = (1.5 × 0.33 + 1) µV = 1.495 µV = 0.15% of Reading
DC Voltage Accuracy is then 0.09% + 0.15% = 0.24%.
8. Based on 15 years of outdoor calibration results for LI-200SA S/N PY-17555 at NREL.

Uncertainties are also sensitive to the time period. For example, if a daily irradiance uncertainty
is quoted, the uncertainty can be weighted by the intensity as well as time of day. If the data have
an estimated uncertainty of ±2% around noon and ±5% in the early morning or late afternoon, the
daily uncertainty is closer to ±2% because the irradiance is greatest during the middle of the day
than in the early morning or late evening hours.

A.4 REDUCING MEASUREMENT UNCERTAINTY


Measurement uncertainty is important to a wide variety of stakeholders in renewable energy, cli-
mate science, and many other important areas that rely on knowing how the Sun’s radiant energy
interacts with the Earth’s atmosphere and surface.
For example, when evaluating the feasibility of a solar electric facility, the planners, engineers,
investors, and financers must account for data uncertainties. Although capital investment, operating
and maintenance costs and power purchase agreements play significant roles, it is the production
capacity of the facility that is critical to the development process along with the ability of the facility
to match the estimated load requirements.
Financiers look at the power production potential of such a facility to meet the loan costs.
They typically reduce the estimated power production by a portion of the estimated uncertainty of
the modeled plant performance. The amount of electricity produced and sold during the poor solar
Appendix A 335

resource years is often used to gauge the security of the loan and can affect the size of the loan and
the interest charged. Therefore, proposals using datasets with small and well-defined uncertainties
are often in a better position to be approved for financing with an interest rate commensurate with
a more precise knowledge of the power production potential.
Observing the following selected best practices for measuring solar and infrared radiation as part
of a comprehensive data quality management plan can limit measurement uncertainty:

• Select instrumentation, supporting equipment, and data acquisition system(s) with


performance specifications consistent with project goals.
• Follow manufacturer’s instructions for proper installation and operation of equipment.
• Develop and follow a maintenance plan consistent with data quality requirements and the
measurement environment.
• Use qualified staff and provide specialized training and proficiency checks.
• Implement regular re-calibration of instrumentation and data system(s).
• Calibrations should be conducted in a manner consistent with the way the equipment is
used and include an estimate of the measurement uncertainty.
• Regularly upload and view the data to quickly spot any problems.
• Maintain accurate records of measurements, calibrations, and maintenance activities.

In general, proper maintenance of the instruments, periodic calibrations, and sound record keeping
all enhance the quality of measurements and lower measurement uncertainty. Well-maintained
instruments and data that are regularly scanned produce more reliable values by quickly attending
to equipment failures and removing the dust, dew, snow, or frost that can cover the radiometers.
Periodic radiometer calibrations help the data user keep track of changes in the responsivity and,
if longer term records are kept, can provide a clear estimate of the instrument degradation rates.
Careful record keeping helps in the analysis of the data and can help identify operational problems
that can occur in the measurements. In addition, keeping records of routine maintenance helps to
insure operational maintenance does occur by providing a source for accountability. When the
maintenance and calibration records are submitted along with the measured data, the user’s confi-
dence in the data set increases.

A.5 DETERMINING A MEASURABLE DIFFERENCE


Including estimated measurement uncertainty is important when comparing different results to deter-
mine a measurable difference. For example, in radiometry a common need is to know when to change
an instrument’s responsivity value to reflect the results of a recent recalibration. Part of the decision for
changing the responsivity is to determine a measurable difference between the previous value and the
newer value from the recalibration. As shown in Figure A.2, a measurable difference exists between
the two measurements (M1 and M2) when that difference is greater than the sum of the expanded uncer-
tainties (U1 and U2) for the two measurements. In this example, the uncertainty of the difference (Δ)
between the two measurements (i.e., calibrations) is computed by taking the square root of the sum of
the squares of the two uncertainties. That is, the two uncertainties are summed in quadrature to deter-
mine the uncertainty of the difference. Figure A.2 illustrates the results of a radiometer recalibration
with four hypothetical outcomes.6 Using the scaled dimensions in the figure, a measurable difference
exists between the previous calibration (M1) and the fourth case (D) from the present calibration (M2)
because the difference is larger than the estimated uncertainty of the difference (UDiff).

6 The expanded uncertainties in this example are at a 95 percent level of confidence, which is the currently accepted default
level of confidence for expressing radiometer calibration uncertainties. A more conservative 99 percent level of confi-
dence can be approximated by simply adding the 95 percent uncertainties, assuming the data are normally distributed,
rather than summing in quadrature (Dieck 1992).
336 Appendix A

FIGURE  A.2 Detecting a measurable difference between two measurements, M1 and M2, with estimated
uncertainties, U1 and U2 based on four hypothetical measurement cases.

It is possible to refine the estimated change in calibration over time by modeling the calibration val-
ues, often using a linear model, and determining the uncertainty in the fit to the changing calibrations.
This way gradual changes in calibration values over time can be made to reflect new calibration results,
eliminating abrupt changes when there is a detectable difference from an older calibration value.
An alternative method for determining a measurable difference in two normally distributed
populations is called the Z-test (Sprinthall 2011). Applied to radiometry, this test can be used to
compute the number of standard deviations separating the two calibration results (Z):

M1 − M 2
Z= (A.5)
uc ( M1 )2 + uc ( M 2 )2

where
M1, M2 = Measured values
uc = Estimated standard uncertainty of the measurement result7

If the ratio (Z) determined by Equation A.5 is less than or equal to 2 (2 standard deviations), then
there is reasonably no detectable difference between the two measurement results with a 95 percent
confidence level and the two calibration values can be considered consistent. Ratios greater than or
equal to 3 can be considered to demonstrate a detectable difference with a probability exceeding 99
percent.
It  is important to note that although adjacent radiometer calibrations may be considered the
same within their stated uncertainties, it is important to track calibration results with respect to the
original “factory” value or the first recalibration result when determining a detectable change has
been observed from the latest calibration. Factory calibrations typically are done under controlled

7 Radiometer calibration results generally include the Expanded Uncertainty (U) based on a stated coverage factor (k).
The estimated Standard Uncertainty (uc) is computed by the ratio U/k as described in the next section.
Appendix A 337

conditions that differ from field calibrations. The location, calibration references, and environmen-
tal conditions are different. However, both results should be consistent with the estimated measure-
ment uncertainties associated with the instrument.

A.6 TERMINOLOGY
The following sample of relevant terminology is presented here as it applies to this appendix and
other instrument-specific chapters in this book. The sample listing is based on the terms defined by
the Joint Committee for Guides in Metrology (JCGM; JCGM 2012) as used in the GUM and
information available from the National Physical Laboratory (NPL; Bell 2001) and the National
Renewable Energy Laboratory (NREL; Sengupta et al. 2017). The reader is encouraged to refer to
the above references for complete and authoritative definitions.

Accuracy
Amount of agreement between the result of a measurement and the best estimate of the
measurand that has been corrected to remove systematic biases. Because some uncertainty
is always associate with the measurand the best that can be done is have a best estimate of
the measurand. The phrase “true value of the measurand” is not used in this book because
the word “true” is considered redundant in the GUM approach. (Note that the qualitative
term, accuracy. should not be confused with precision.)
Adjustment
An adjustment to a measurement is made to remove systematic bias that is caused by the
nature or environment in which the instrument is used for measurement. For example, an
instrument may respond differently depending on the temperature. By measuring the tem-
perature dependence of the instrument and the temperature under which the measurement
is made, an adjustment can be made to the measured data to remove the know bias caused
by the temperature dependence of the instrument. An adjustment has its own uncertainty
that must be included in the uncertainty analysis. An adjustment is sometimes called a
correction in that it is brings the measurement closer to the best available estimate of a
measurand.
Best Estimate of the Measurand
Given the state of the technology and the environmental conditions, it is an estimate of
the measurand with the least uncertainty and known biases removed. This concept has
replaced true value or true value of the measurand.
Bias (of a measuring instrument)
An estimate of a systematic change in the measurement instrument’s response dependent
on the environment or circumstance under which the measurement is made. For  exam-
ple, thermopile-based pyranometers and pyrheliometers are generally characterized for
ambient temperature response by the manufacturer. Irradiance measurements can then be
adjusted for this known bias using coincident measurements of air temperature.
Calibration
The  process of determining the responsivity and associated uncertainty of a measuring
instrument by comparison with a recognized reference of known traceability to reference
measurement standards.
Combined Standard Uncertainty (uc)
The  estimated combined uncertainty (uc) computed from Type A  (u A) and Type B (uB)
standard uncertainties associated with the input quantities in the measurement equa-
tion summed in quadrature:

uc = ∑u + ∑u
A
2
B
2
(A.6)
338 Appendix A

Confidence Interval (Level)


When multiplying the combined uncertainty (uc) by a coverage factor (k) to determine the
expanded uncertainty (U), the interval ±U has a probability (p) of containing the mea-
surand value. The specific probability (p) is called the level of confidence or the coverage
probability and approximates the probability coverage. When applied to radiometry, k is
typically assigned a value near 2 resulting in a 95 percent level of confidence.8
Correlation
A mathematical relationship or interdependence among quantities. When calculating the
combined standard measurement uncertainty, the correlation or interdependence between
quantities must be considered.
Coverage Factor (k)
A number, greater than one, multiplied by the combined standard uncertainty (u) to obtain
the expanded uncertainty (U).
Degrees of Freedom
The number of independent values that can vary in the uncertainty analysis within the mea-
surement constraints. For a random distribution, the degrees of freedom (DF) determine the
Student’s t-distribution value (t) that defines the interval that encompasses a specified fraction
(i.e., level of confidence) of the distribution. Example: For an average value based on 31 mea-
surements, DF is 30 which determines t as 2.04 for a 95 percent confidence interval. Therefore,
the Type A uncertainty for the measurement set would be proportional to the estimated stan-
dard deviation (see Equation A.7) multiplied by 2.04 for a 95 percent confidence interval. With
an infinite DF, the limiting value of t is 1.960 for a 95 percent confidence interval.
Error
An error, or more precisely called the uncertainty,9 is the difference between a measured
value and the best estimate of the measurand. This uncertainty can be due to random and/or
systematic effects. Error is associated with the difference between the true value and the
measurement value. Since true value is an idealized concept based on the best estimate of
the measurand, the concept of error cannot be known exactly. The uncertainty of a mea-
surement is related to the difference between the measured value and the best estimate of
the measurand but is dependent on the proper calculation of the uncertainty. An error is a
simplified version of uncertainty without the complete statistical analysis needed to deter-
mine uncertainty.
Estimated (Experimental) Standard Deviation
An unbiased indication of the differences between an individual measurement and the
average of a population of measurements. The standard deviation (s) for a sample sequence
of n measurements (x) with an average value (x ) is expressed as


n
( xi − x )
2

s= i =1
(A.7)
n −1
Note: When estimating measurement uncertainty, the population is considered to be infi-
nite and the number of recorded measurements is therefore a sample of that population.
The estimated standard deviation in Equation A.7 provides an estimate of the population
standard deviation based on the sample. The GUM refers to the experimental variance (s2)
and the experimental standard deviation of the mean (s/ n ) for estimating measurement
uncertainty.

8 The GUM does not name the uncertainty interval ±U because the term confidence interval has a precise statistical defini-
tion not typically met by the expanded uncertainty formulation.
9 The change in the GUM treatment of measurement uncertainty from an Error Approach or the traditional True Value
Approach to an Uncertainty Approach forced the reconsideration of some terminology concepts.
Appendix A 339

Expanded Uncertainty (U)


A quantity defining an interval about the result of a measurement that may be expected to
encompass a large portion (coverage) of the values that could reasonably be attributed to
the measurand. The combined standard uncertainty (uc) multiplied by a coverage factor (k)
yields the expanded uncertainty (U).
Gaussian Distribution
A  distribution of random values that is symmetric about the mean and becomes more
pointed with decreasing values of the standard deviation. Also called a normal distribution
or a Gauss-Laplace distribution.
Mean
The arithmetic average of a set of values computed as their sum divided by the number of
values in the set.
Measurand
The quantity to be measured including a description of the circumstances under which the
quantity is to be measured. The measurand often specifies certain physical states and con-
ditions (i.e., Direct Normal Irradiance (DNI)) at the Earth’s surface for solar zenith angles
greater than 80 degrees as measured by a pyrheliometer with a 5.0 degree field of view.
It is not possible to completely describe the measurand. The incompleteness of the descrip-
tion assures that the measurement of the measurand has an associated uncertainty.
Measurement Equation (Model)
In radiometry, the measurand (e.g., irradiance) is not measured directly, but is determined from
other quantities (e.g., voltage) through a functional relationship. A measurement equation is
the expression defining the transformation of repeated observations into a measurement result.
Metrology
The science of measurement and its application.
Normal Distribution
(See Gaussian Distribution)
Operator Error
A mistake made by the person(s) conducting the measurement.
Precision
A term used to describe the fineness of discrimination, or the number of digits given. The
term measurement precision should not be used for measurement accuracy or uncertainty.
Random Effects
The variation of measurements resulting from unpredictable or stochastic variations that
influence the measurement. Random effects were previously called random error.
Range
The span or difference between the highest and lowest values in a set.
Reading
The observed and recorded value at the time of measurement.
Rectangular Distribution
The distribution of values with an equal likelihood of being anywhere within a range.
Reference Scale (Measurement Standard)
A quantity-value scale defined by formal agreement (e.g., World Radiometric Reference
(WRR)).
Repeatability
Closeness of agreement among measurements of the same property (measurand) under the
same conditions.
Reproducibility
The degree of closeness of agreement among measurements of the same measurand con-
ducted under different conditions of measurement (e.g., at a different time, place, or by a
different operator).
340 Appendix A

Resolution
The smallest change in a quantity being measured that can be meaningfully distinguished
(e.g., a change in the last place of a digital display often referred to as precision).
Responsivity
The  relation between measurement results and indications determined by a calibration
process using traceable measurement standards. Responsivities must also include state-
ments of uncertainty (e.g., a thermopile-based pyranometer responsivity of 10.01  µV/
Wm−2 ± 1.5% (U95) for solar zenith angles between 30° and 60°).
Result (of a measurement)
A set of quantity values attributed to a measurand including any other available relevant
information. (Note: A measurement presupposes a description of the quantity commen-
surate with the intended use of a measurement result, a measurement procedure, and a
calibrated measuring system operating in accordance with a specified procedure, including
measurement conditions [JCGM 2012].)
Sensitivity (of a measuring system)
The amount of change in an indication of a measuring system (i.e., instrument response)
divided by the corresponding change in the stimulus (i.e., the quantity being measured).
Sensitivity Coefficient
The partial derivatives of the measurement equation with respect to an input. These coef-
ficients are used to identify the relative contributions of each input to the combined uncer-
tainty estimate.
Standard Deviation (of a population)
An indication of the differences between an individual measurement and the average of a
population. The standard deviation (σ ) for a known population (N) of measurements (x)
with an average value (μ) is expressed as:


N
( xi − µ )
2

σ= i =1
(A.8)
N
Note: When determining the variance in measurements, it is appropriate to assume an infi-
nite population. Because all measurements can be considered as a sample of the popula-
tion, the estimated standard deviation in Equation A.7 is used for calculating the estimated
measurement uncertainty.
Standard Uncertainty (u)
The uncertainty of the result of a measurement expressed as a standard deviation (s) in
Equation 4.3. The value depends on the source of uncertainty and its assumed distribution
as shown in Table A.4.
Student’s statistic
A statistical parameter (t) used in estimating the population mean (µ) from a distribution of
sample means (x ) when the population standard deviation (σ) is unknown for a normally
distributed population. Assuming the effective degrees of freedom (DF) is infinite, the
t-statistic (t) is 1.96 for a 95 percent confidence interval. The number of independent val-
ues that can vary in an analysis within the constraints are indicated by DF. When applied to
radiometry, the DF is generally the number of measurements recorded minus 1. For exam-
ple, the t parameter at 95 percent confidence is 2.043 (DF = 30), 2.0003 (DF = 60), and
1.9867 (DF = 90). The coverage factor used in GUM is determined by the t-statistic and
the degrees of freedom.
Systematic Bias
A  component of measurement uncertainty that in replicate measurements remains con-
stant or varies in a predicable manner and is separable from random measurement effects.
Appendix A 341

TABLE A.4
Standard Uncertainty Estimation for Common Distributions
Distribution Uncertainty Source Standard Uncertainty
Gaussian: Type A s U
u= OR u =
Standard deviation (s) (Statistical) n −1 k
Number of readings (n)
OR
Expanded uncertainty (U)
Coverage factor (k)
Rectangular Type B U
u=
(Non-statistical) 3
Triangular Type B U
u=
(Non-statistical) 6
Gaussian: Type B U U
u= OR u =
95% Confidence Interval Infinite Degree of Freedom: (Non-statistical) k t
k ≌ 2
OR
Specific Degree of Freedom:
Student’s t (t)

Source: Reda, I., Method to Calculate Uncertainties in Measuring Shortwave Solar Irradiance Using
Thermopile and Semiconductor Solar Radiometers. NREL/TP-3B10-52194. National Renewable
Energy Laboratory, Golden, CO, 20 p, 2019.

Systematic measurement bias, and its causes, can be known or unknown and impose either
a positive or negative bias or offset to the measurement (e.g., radiometer responsivity from
a calibration process).
Traceability
Metrological traceability is the property of a measurement result that can be related to a
reference through a documented unbroken chain of calibrations. Each calibration in the
series contributes to the measurement uncertainty of the instrument. Currently, the WRR
and the interim World Infrared Standard Group (WISG) are the recognized references for
calibrating solar and infrared radiometers, respectively.
True Value
The  observed measurand quantity obtained by a perfect measurement. A  true value is
unique and, in practice, unknowable. The GUM specifies “value of the measurand” and
considers “true” to be redundant. The best estimate of the measurand, with its associated
uncertainty, is the proper parameter to use for the true value.
Type A Evaluation of Uncertainty
A method of evaluation of uncertainty by the statistical analysis of a series of observations.
Type B Evaluation of Uncertainty
A method of evaluation of uncertainty by means other than the statistical analysis of series
of observations, such as calibration data, manufacturer specifications, and experimental or
engineering estimates.
Uncertainty Budget
A statement of the uncertainty calculations, including the measurement equation, compo-
nents of measurement uncertainty and their combination, the degrees of freedom, and any
coverage factor.
342 Appendix A

Uncertainty of Measurement
A quantified accuracy about the result of a measurement characterizing the dispersion of the
quantity values attributed to a measurand based on the information available (e.g., measurement
standards, operating procedures, systematic effects, and statistical analyses).
Uniform Distribution
(See Rectangular Distribution)
Value of the Measurand
The result of an ideal measurement that can never be known exactly due to random varia-
tions of the observations, systematic differences, and incomplete knowledge of relevant
physical phenomena. For a full description, the “value of the measurand” must be associ-
ated with an uncertainty. The GUM considers “true value of a measurand” and “value of a
measurand” as equivalent.

REFERENCES
Bell, S. 2001. A Beginner’s Guide to Uncertainty of Measurement, Measurement Good Practice Guide No. 11
(Issue 2). Middlesex, UK: National Physical Laboratory, Teddington. https://www.npl.co.uk/resources/
gpgs/guide-to-measurement-uncertainty (Accessed December 31, 2018).
Dieck, R. H. 1992. Measurement Uncertainty: Methods and Applications. Research Triangle Park, NC:
Instrument Society of America.
JCGM. See Joint Committee for Guides in Metrology.
Joint Committee for Guides in Metrology. 2008. Guide to the Expression of Uncertainty in Measurement.
Geneva, Switzerland: International Organization for Standardization. https://www.bipm.org/utils/
common/documents/jcgm/JCGM_100_2008_E.pdf (Accessed December 31, 2018).
Joint Committee for Guides in Metrology. 2012. International vocabulary of metrology—Basic and general
concepts and associated terms (VIM), 3rd ed. https://www.bipm.org/en/publications/guides/vim.html
(Accessed December 31, 2018.)
Myers, D. R., T. L. Stoffel, I. Reda, S. M. Wilcox, and A. M. Andreas. 2002. Recent progress in reducing
the uncertainty in and improving pyranometer calibrations. Transactions of the ASME 124: 44–50.
doi:10.1115/1.1434262.
Reda, I. 2011. Method to Calculate Uncertainties in Measuring Shortwave Solar Irradiance Using Thermopile
and Semiconductor Solar Radiometers. NREL/TP-3B10-52194. Golden, CO: National Renewable
Energy Laboratory. https://www.nrel.gov/docs/fy11osti/52194.pdf (Accessed June 1, 2019).
Sengupta, M., A. Habte, C. Gueymard, S. Wilbert, and D. Renné (eds.). 2017. Best Practices Handbook for
the Collection and Use of Solar Resource Data for Solar Energy Applications: Second Edition. NREL/
TP-5D00-68886. Golden, CO: National Renewable Energy Laboratory. https://www.nrel.gov/docs/
fy18osti/68886.pdf (Accessed June 1, 2019).
Sprinthall, R. C. 2011. Basic Statistical Analysis, 9th ed. Boston, MA: Pearson Publishing Co.
Taylor, B.N. and Kuyatt, C.K. 1994. Guidelines for Evaluating and Expressing the Uncertainty of NIST
Measurement Results. NIST Technical Note 1297. National Institute of Standards and Technology.
https://emtoolbox.nist.gov/Publications/NISTTechnicalNote1297s.pdf (Accessed June 5, 2019).
Appendix B: Modeling
Solar Radiation

B.1 INTRODUCTION
Before incident solar radiation is measured, it’s important to understand the characteristics and
magnitude of the incoming irradiance and how the radiation is changed as it passes through the
atmosphere. Solar radiation basics are discussed in Chapter 2, and spectral characteristics are cov-
ered in Chapter 7. From the existing data and with knowledge of the characteristics of solar irra-
diance, a wide variety of models have been created to estimate irradiance from meteorological
parameters, atmospheric constituents, and a limited set of irradiance measurements. Some models
exist to estimate solar radiation on clear days, and other models have been created to estimate dif-
fuse horizontal irradiance (DHI) or direct normal irradiance (DNI) from global horizontal irradi-
ance (GHI) data under all weather conditions. This appendix provides an overview of a variety of
solar radiation models used to estimate irradiance values when there are limited or no solar radia-
tion measurements at a site.
The first models covered in Appendix B are the clear-sky models that estimate the DNI irradi-
ance when clouds are not  present. Clear-sky models also are referred to as clear-day models in
some literature. The clear-sky modeling section is followed by the definition and initial discussion
of the clearness index parameters, Kt, Kdf, and Kb. The clearness index parameters are used with
many of the other models. Next, models that estimate the diffuse fraction are described followed by
a discussion of the beam–global (DNI–GHI) correlations. The diffuse fraction is the ratio of DHI
divided by GHI. Estimates of the diffuse irradiance then are obtained by multiplying the diffuse
fraction by GHI. Estimating the irradiance on tilted surfaces then will be demonstrated. A sum-
mary of the International Energy Agency (IEA) report on model comparisons then will be given
along with references to a variety of models.

B.2 CLEAR-SKY MODELS


As sunlight passes through the Earth’s atmosphere, the radiation is scattered, absorbed, and trans-
mitted in varying amounts by the atmosphere before reaching the surface. At the surface, the amount
of sunlight coming directly from the Sun is called DNI. If the exact constituents of the atmosphere
between the Sun and the observer are known, the DNI can be calculated to a high degree of accuracy
knowing the spectral composition of the extraterrestrial irradiance, the Earth–Sun distance at the
time of interest, and the length of the path through the atmosphere. The physics behind this calcula-
tion is described by the Beer–Lambert Law.
In general

DNI = DNI oexp( −m ⋅ (τ a + τ r + τ g + τ NO 2 + τ w + τ O 3 ) (B.1)

where DNIo is the extraterrestrial beam irradiance, m is the air mass, τa is the optical depth due
to aerosol absorption and scattering, τr is the optical depth due to Rayleigh scattering from atmo-
spheric molecules, τg is the optical depth due to absorption by gases other than those specified such
as carbon dioxide (CO2) and oxygen (O2), τNO2 is the optical depth due to absorption by nitrogen
dioxide, τw is the optical depth due to water vapor absorption, and τO3 is the optical depth due to
absorption by the ozone.

343
344 Appendix B

Exact calculations incorporate the wavelength dependence of the absorption and scattering
and the non-uniform distribution of the atmospheric constituents. The accuracy of the calcula-
tions is limited by the lack of information on the exact amount and location of the atmospheric
constituents at the time of interest.
Clear-sky models were developed to be simple to use and to give an approximate estimate of
DNI on clear days. Since the total solar irradiance (TSI) from the sun is constant to within a few
tenths of a percent (the currently accepted value of 1,361 watts per square meter (Wm−2) is used
in this book as described in Chapter 2) and the Earth–Sun distance is known with a consider-
able degree of accuracy, the incident irradiance outside the Earth’s atmosphere can be accurately
calculated. The extraterrestrial DNI is

DNI o = 1361Wm −2 ⋅ ( R /Ro )


2
(B.2)

where
R is the distance from the Earth to the Sun for the day of interest
Ro is the mean Earth–Sun distance

The mean Earth–Sun distance is called 1 astronomical unit (AU). Because Earth’s orbit is elliptical,
the distance between the Earth and the Sun varies by about ±1.6 percent over the year. Since the
Sun’s radiance can be treated as a point source, the irradiance is inversely proportional to the square
of the separation distance. The incident solar radiation at the top of the atmosphere (TOA) varies by
approximately 6.6 percent from the closest approach (perihelion) around January 3 to the farthest
distance (aphelion) around July 4.
During clear-sky periods, there are no clouds to interact with the DNI. However, other con-
stituents of the atmosphere absorb and scatter the irradiance as it passes. Since the clear-sky
model is an attempt to simplify a complex process, there is no exact prescription for obtaining
the simplified model. In fact, over a dozen clear-sky models attempt to estimate the incoming
irradiance as the DNI passes through the atmosphere (Atwater and Ball 1978, 1979; Badescu
2013; Bemporad 1904; Bird and Hulstrom 1979, 1980; Davies and Hay 1979; Hoyt 1978; Kasten
1964; Kondratyev 1969; Lacis and Hansen 1974; Liu and Jordan 1960; Mahaptra 1973; McDonald
1960; Moon 1940; Pérez-Burgos et al. 2017; Vignola and McDaniels 1988a; Watt 1978). These
models use the measured or average values of atmospheric elements that could interact with
the sunlight to model the scattering or absorbance of the irradiance. Specifically, the amount of
scattering depends on the amount and types of the aerosols in the atmosphere and the amount of
water vapor, ozone, and other gases. Since it is difficult to know the exact mixture of these ele-
ments in the atmosphere and since the models were developed at different sites with instruments
of varying quality used to validate the models, it is not  surprising that there are a variety of
model estimates given the same initial conditions. Bird and Hulstrom (1981) and Badescu et al.
(2013) provide a good survey of the clear-sky models.

B.3 CLEARNESS INDICES KT, KDF, AND KB


Normalization of the incident radiation by dividing by the equivalent extraterrestrial component
is very useful for irradiance modeling and data analysis. This method helps to compare data from
different times of year or different times of day. Normalization is especially useful when devel-
oping generalized correlations that hold true over the year or over the day. The averaging can be
done over a minute, an hour, a day, or a month, depending on the available data and information
required.
Appendix B 345

The  standard formula for extraterrestrial DNIo is given in the previous section for clear-sky
modeling. Since DNIo is used over various time periods, it is usual to obtain the value by averaging
over the time period.

ha2
24 24
DNI o =
2π ∫ DNI d(ha) = 2π DNI ⋅ ( ha − ha )
ha1
o o 2 1 (B.3)

where ha is the hour angle defined in Section  2.5. The  factor 24/(2π) changes from hour angles
to radians. The hour angles also should be expressed in radians. Since DNIo changes by less than
0.2 percent per day, it is generally assumed to be constant over the day.
The global extraterrestrial value, GHIo, is equal to DNIo projected onto a horizontal surface by
multiplying by the cosine of the solar zenith angle (computed from site location, day of year, and
time of day as described in Appendix E).

ha2
24
GHI o =
2π ∫ DNI (sin(lat )sin(dec) + cos(lat ) cos(dec) cos(hha)) ⋅ d(ha)
ha1
o (B.4)

or

GHI o =
24

(
⋅ DNI o ⋅ sin(lat ) ⋅ sin( dec) ⋅ ( ha2 − ha1 ) + cos(lat ) ⋅ cos( dec) ⋅ ( sin ( ha2 ) − sin ( ha1 ) ) )
(B.5)

The hour angles in Equation B.5 are in radians and not degrees. For daily GHI o , the sunrise and
sunset hours as determined in Equation 2.11 can be used. For monthly averaged GHI o the daily
GHI o should be averaged over the month. For many months, the middle of the month produces an
adequate monthly averaged GHI o . However, for both June and December, the average GHI o comes
later or earlier, respectively, because the solar declination reaches a maximum or minimum dur-
ing the month and the declination at the middle of the month is not typical of the average monthly
declination (Klein 1977; Klein and Theilacker 1981). The extraterrestrial DHIo is essentially zero as
there are few atoms, molecules, or particles in space to scatter the DNIoirradiance. To differentiate
between short interval, daily, and monthly average clearness values, the following convention is
used. For the global clearness index, kt is used for hourly or shorter time intervals, Kt is used for the
daily clearness index, and Kt is used for the monthly average index. The DHI and DNI clearness
indices have a similar structure with t replaced by df or b, respectively.

B.4 IRRADIANCE ON A TILTED SURFACE


To calculate global irradiance on a tilted surface (GTI), irradiance models sum the contributions from
the DNI, the DHI, and the ground-reflected radiation. Instantaneously, the DNI contribution is equal
to DNI times the cosine of the incident angle to the surface. The formula for the cosine of the incident
angle to the surface is given by Equation 2.13 in Chapter 2. When using longer time interval data, the
product must be integrated over time. Because the average beam irradiance (DNI) on the surface var-
ies systematically over the day, the average of the cosine should be weighted by the same systematic
change in the average beam irradiance over the time-period of interest. This is not always done, and for
hourly values the weighting is significant only in the morning and evening hours when the atmospheric
scattering and absorption significantly attenuate the incoming DNI (Vignola and McDaniels 1988b).
346 Appendix B

The projection of the DHI onto the tilted surface is complex as the diffuse radiance varies across
the sky (is anisotropic) and depends on the relative position of clouds and the Sun. In general, dif-
fuse radiation can be broken into four components: the forward-scattered radiation around the Sun
(aureole or circumsolar radiation), the diffuse radiance brightening near the horizon, the minimum
diffuse radiance 90 degrees from the Sun (i.e., anti-solar point), and the diffuse radiance from
the rest of the sky. Mathematically, it can be shown that if the diffuse radiance around the Sun is
evenly distributed, then it can be projected onto a surface much like DNI. Modeling other diffuse
components can be complex. One of the more rudimentary models treats the diffuse radiance as
distributed evenly across the sky (i.e., isotropic), and DHIT on a tilted surface is simply

DHIT = DHI ⋅ (1 + cos (θT ) ) / 2 (B.6)

where θT is the tilt angle of the surface from the horizontal. This  isotropic diffuse model is
the  simplest method (Kondratyev 1977; Kondratyev and Manolova 1960; Liu and Jordan 1963).
More comprehensive models that take into account the various diffuse components such as bright-
ening around the Sun or horizon brightening perform better than the simple isotropic diffuse mod-
els (ASHRAE 1971; Brink 1982; Bugler 1977; Cohen and Zerpa 1982; Gueymard 1983; Hay 1979;
Hay and Davies 1980; Hay 1984; Hay and McKay 1985; Hay et al. 1986; Hay 1993; Ineichen 1983;
Klucher 1979; Kusuda, 1976; Lawrence Berkeley Laboratories 1982; Ma and Iqbal 1983; Oegema
1971; Page 1978; Perez et al. 1983; Puri et al. 1980; Rogers et al. 1979; Skartveit and Olseth1986;
Temps and Coulson, 1977; Threkheld 1962).
A tilted surface also receives solar radiation reflected from the ground. Ground-reflected radia-
tion is very difficult to accurately model because the ground is not a uniform reflector and can scatter
light in a specular fashion. The radiative characteristics of this scattering are wavelength dependent.
In addition, many objects shade the ground from direct sunlight such as vegetation and structures,
further complicating any modeling. Therefore, most models that calculate irradiance on tilted sur-
faces use a simple isotropic ground-reflecting model to estimate the light reflected from the ground.

ground _ reflected = GHI ⋅ ρ ⋅ (1 − cos (θT ) ) / 2 (B.7)

where
θT is the tilt angle of the surface from the horizontal
ρ is the albedo of the ground, usually a value of 0.2 for vegetated surfaces, but the albedo can
vary from 0.1 for a dark surface to 0.9 for a surface covered with snow (see Chapter 8 for
more information about albedo)

Errors associated with modeling ground-reflected light are usually not  significant. For  example,
given the most extreme case when the surface is perpendicular to the ground (tilted 90°) and the
albedo is 0.2, the ground-reflected irradiance is just 10 percent of the GHI. Therefore, a 20 percent
error in the model will produce only 2 percent error in the calculated tilted irradiance, well within
the measurement uncertainty of most tilted surface irradiance measurements. Of course, if the sur-
face faces a snowfield with an albedo between 0.7 and 0.9 or faces a highly reflective surface, there
can be a large contribution from the ground-reflected light.
The IEA Task 9 compared many of the models that estimate the irradiance on tilted surfaces
(Hay and McKay 1988). The study divided the models into three types. The first type used hourly or
short time interval data to estimate diffuse irradiance on tilted surfaces. Twenty-one models were
compared (ASHRAE 1971; Brink 1982; Bugler 1977; Cohen and Zerpa 1982; Gueymard 1983; Hay
1979; Hay and Davies 1980; Hay and McKay 1985; Hay et al. 1986; Ineichen 1983; Klucher 1979;
Kusuda, 1976; Lawrence Berkeley Laboratories 1982; Oegema 1971; Page 1978; Perez et al. 1983;
Puri et al. 1980; Rogers et al. 1979; Skartveit and Olseth 1986; Temps and Coulson 1977), and the
Appendix B 347

models by Perez et al. (1983), Gueymard (1983), and Hay (1979) were rated the best when compared
with data from 27 data sets from around the world. The range of models varied from those using the
isotropic distributions to the sky distributions of Moon and Spencer (1942) to detailed diffuse sky
models such as that of Perez et al. (1983).
Seven models were tested that used daily GHI values to determine the daily GTI data values.
These models were used to estimate the DNI and then calculate the direct component on the tilted
surface (Bremer 1983; Desbucam et  al. 1986; Jones 1980; Klein and Theilacker 1981; Liu and
Jordan 1962; Page 1961; Revfeim 1976, 1979, 1982a, 1982b). To perform these calculations, some
model of the diurnal variation has to be included. Identical diffuse values were used for all compari-
sons (Collares-Pereira and Rabl 1979a, 1979b, 1979c). The Page (1961) model gave the best results
for the direct component on the horizontal surface. The models did have problems when estimating
irradiance data for surface orientations away from due south. Any asymmetry in the daily irradiance,
more irradiance in the morning or afternoon, would lead to such results (Vignola and McDaniels
1988b, 1989).
Estimation of daily diffuse irradiance on tilted surfaces from daily GHI values also was studied.
Either an isotropic diffuse model (Kondratyev and Manolova 1960; Liu and Jordan 1963) or a model
to estimate the diurnal variation over the day (Collares-Pereira and Rabl 1979a, 1979b, 1979c) can
be used, and then one of the hourly diffuse models can be used to estimate the diffuse on a tilted
surface. This  was the approach used by Gueymard (1985, 1986a, 1986b) and McFarland (1983).
All these approaches work better than the isotropic models.

B.5 DIFFUSE FRACTION


All models that calculate irradiance on a tilted surface require the use of DNI and DHI values rather
than starting with simply the GHI value. Since GHI data are more often available, models were
developed to estimate the DHI and DNI values from GHI values. The relationship between DHI and
GHI is double-valued and not linear. Low values of DHI occur when GHI values are low and the sky
is very cloudy or overcast and on very clear days when GHI is high.
Diffuse modelers address this problem by dividing DHI by GHI, creating the diffuse fraction.
When the diffuse fraction is plotted against kt, a functional relationship results. Modelers use this
relationship to estimate diffuse irradiance from GHI measurements. A sample plot of the diffuse
fraction plotted against kt is shown in Figure B.1 for hourly data. The densest area of data points
represents clear periods with little or no clouds. Clouds give the broad scatter to the possible GHI
values because there are numerous ways for clouds to reflect sunlight onto the pyranometer and
add to the measured GHI and DHI values. Gaps in the clouds that let some or most of the DNI
through lead to the widest scatter in the data points in Figure B.1. A general relationship between
the diffuse fraction and clearness index can be used for short interval, daily, or monthly average
correlations. The hourly, daily, and monthly relationships will differ. As the time period increases
over which the diffuse fraction is analyzed, the scatter in the data is greatly reduced. For hourly
data (Figure B.1), the spread of the data is reduced by about one-half from the scatter of 1-minute
data (see Figure 14.8). For daily or monthly averaged values, the scatter of the data is even further
reduced. In addition to the reduction of data scatter, the average relationship between the diffuse
fraction and clearness index also changes. Therefore, the relationship for daily data is not the same
as the relationship for monthly average data. The reduction in the scatter of the data resulting from
the average enables the seasonal relationship for daily and monthly average relationships to become
apparent with higher diffuse fractions in the spring than fall as shown in Figure B.2 (Vignola and
McDaniels 1984a; 1984b). This seasonal dependence is probably related to the varying aerosol opti-
cal depth determined by the dust loading in the atmosphere.
Scattering from dust and most aerosols is Mie-type scattering with a majority of the scattered
light going in the forward direction. With higher levels of aerosols and dust, the DHI component
increases, whereas the GHI decreases only slightly because the scattering is mostly in the forward
348 Appendix B

FIGURE B.1 Plot of the diffuse fraction plotted against clearness index kt for hourly data. Most of the data
fall along the lower right-hand side of the plot. Those data points represent clear periods. There is a spread of
data points to the left that represents data when clouds are present. When there is no DNI, the diffuse fraction
is 1 and GHI = DHI. When there is some Sun present, the clouds can act as reflectors and a wide range of GHI
values can occur for any diffuse fraction.

FIGURE B.2 Plot of monthly average diffuse fraction (DHI/GHI or in this figure DF/H) versus monthly aver-
age clearness index K t . Data are plotted in two groups: February through July and August through January.
There is a clear seasonal dependence of the diffuse fraction in relation to the monthly average clearness index.

direction. Therefore, it is expected that the relationship between the diffuse fraction and the clear-
ness index will vary at locations with different seasonal aerosol loading or during periods affected
by volcanic eruptions.
When the diffuse fraction is determined from the functional relationship, the DHI is estimated
by multiplying the fraction by GHI. The DNI can then be calculated by subtracting the DHI from
the GHI and dividing the result by the average value of the cosine of the solar zenith angle. For lon-
ger time periods, such as an hour, a more accurate way to determine the average cosine of the solar
zenith angle is to weight the cosine by a clear-sky DNI model (see Section B.7 for details).
Appendix B 349

For the daily data the cosine of the incident angle is integrated over the day, and for monthly aver-
aged data the cosine is averaged for the day in the month that is closest to the average extraterrestrial
irradiance for the month. For daily and monthly averaged cosine of the solar zenith angle, a constant
DNI value is assumed because the existing models were developed without a weighting factor.
Most of the diffuse fraction, clearness index studies have used DHI and GHI values that come
from instruments that have large uncertainties (±5% or more). Models derived from DHI values
obtained using pyranometers with fixed shadowbands (see Figure 6.3 in Chapter 6) have reduced
accuracy because other models are needed to estimate the shadowband correction factor to account
for blocking the sky as well as the solar disk. When the DHI values are obtained by subtracting
DNI times the cosine of the solar zenith angle from the GHI data, the systematic errors in the GHI
and DNI add significant uncertainties to the DHI values. More accurate DHI values are obtained
by shading pyranometers mounted on automatic trackers with a disk or sphere that blocks the DNI
component from the pyranometer. However, it remains to be seen how well new studies with better
data will be able to improve the results, especially considering the natural variability in the data and
the variation of the relationship over the year and from site to site. The studies by Gueymard (2009)
and Evseev and Kudish (2009) are examples of model testing and illustrate some of the problems
associated with verification of any model. Some of the models tested that were not referenced earlier
are Willmott (1982), Iqbal (1983), Perez et al. (1990), Reindl et al. (1990), Gueymard (1987), Muneer
(2004), and ASHRAE (2005).

B.6 BEAM–GLOBAL CORRELATIONS


A  more direct way to obtain the DNI value is by using a beam–global (DNI–GHI) correlation.
Using the results of a beam–global correlation, the DNI value can be calculated directly without
having to calculate the DHI first and then divide the difference between the GHI and DHI values by
the cosine of the solar zenith incident angle. It is important to remember that, at large zenith angles,
small errors in either the diffuse (DHI) or global (GHI) values can lead to large errors in the beam
(DNI) values because the cosine of large zenith angles is a small number and this small number
divides into the difference between GHI and DHI.
As with the diffuse fraction–global relationships, the beam–global relationships have a sea-
sonal dependence that reflects the seasonal change in aerosols and other atmospheric constituents.
The contrast is greatest between spring and fall where the solar angles are the same, but the aerosol
content of the atmosphere generally differs.
There are fewer beam–global models than the diffuse fraction–global models because histori-
cally the number of DHI measurements greatly outnumbers DNI measurements (e.g., see Randall
and Biddle 1981). Before the development of microprocessor-controlled automatic solar trackers
and rotating shadowbands, diffuse measurements with fixed shadowbands were easier to make and
less expensive than shading devices on clock-drive solar trackers. DNI measurements required man-
ual trackers that were difficult to keep in alignment and were more expensive than shadowbands.
Interest in deriving DNI directly from GHI started to increase when DNI values became of primary
interest, when more DNI data became available, and when it became clear that DNI measure-
ments were more accurate than DHI measurements. With a number of stations now using automatic
trackers that enable both high-quality DNI and DHI measurements, the relationship between GHI,
DNI, and DHI is once again being studied with hopes of improving on older models (Vignola and
McDaniels 1988b; Vignola 2003).
One of the main problems facing beam–global models is that the rise of the normalized DNI (kb)
as a function of kt is fairly steep, resulting in a given value of kt having a wide range of possible kb
values. One source of this range of kb values is the forward scattering nature of aerosols. Different
types and concentrations of atmospheric aerosols can lead to the same kt (i.e., the GHI values are
unchanged) while yielding different kb values (i.e., the redistribution of DNI and DHI). A typical plot
of kb versus kt is given in Figure B.3. The data obtained during clear skies fall into a concentrated
350 Appendix B

FIGURE B.3 Plot of kb versus kt for hourly data. The clear sky values fall along a line that increases linearly
with kt with a cluster of the highest clear sky values occurring during the middle of the day when the change in
sza is slowest. The data points to the left of the clear sky line is the result of sunlight reflecting off clouds and
into the global pyranometer. This reflected light adds to the GHI but is not seen by the pyrheliometer that mea-
sures the DNI values. There are very few data points with kb values above a maximum of 0.8. The maximum
kb values are somewhat dependent on station elevation because higher altitudes reduce the air mass through
which sunlight traverses.

group on the left side of the plot. When clouds are present, reflected light can greatly enhance the
GHI values. As opposed to 1-minute data (Figure 14.9 in Chapter 14), hourly data produces very few
kt values that are greater than 0.8 and does not result in any data with values of kt greater than 1.0 or
greater than the extraterrestrial irradiance.
The more accurate models for the relationship between kb and kt use the variability of GHI over
the day, the time of day, and other parameters that help separate partially cloudy from clear periods.
For hourly and other short time interval correlations, other meteorological variables can also affect
the correlation. These extra parameters may relate to changing responsivities of the instruments as
affected by the meteorological parameters and might not reflect true changes in the functional rela-
tionship. Haziness does affect the correlation, but it is often difficult to get good information about
the aerosols and incorporate that information into the function. Seasonal variance of the aerosols
also affects the function, as do volcanic aerosols that can be present in the air column through which
the DNI passes (Vignola and McDaniels 1984b).
Of course, once kb is obtained, the diffuse irradiance can be calculated by subtracting the DNI
projected on a horizontal surface from the global irradiance. DHI values obtained from disk-shaded
pyranometers should be the reference for validation. Unfortunately, the systematic biases in the GHI
and DNI can affect the comparison with the measured DHI. One question that sometimes occurs
is that if the GHI measurements were made with an instrument with systematic biases, should the
model used be one that was derived with similar instruments because the offsets from the system-
atic biases will be incorporated into the model? If the pyranometer’s systematic biases are consistent
and large enough, then the models would need to be pyranometer dependent. The systematic spec-
tral biases of photodiode-based pyranometers is an example where the biases in the data have to be
considered or incorporated into the model.
Appendix B 351

B.7 DETERMINING THE AVERAGE COSINE OF THE SOLAR ZENITH ANGLE


For short time periods, the average of the cosine of the solar zenith angle can be calculated to an
acceptable level of accuracy at the midpoint of the time in the interval. For 15-minute and espe-
cially hourly intervals, choosing the midpoint of the interval may result in systematic biases in
the estimated average cosine value. This outcome is especially true when using the cosine of the
incident angle to project DNI onto a horizontal (or tilted) surface during morning or evening hours.
For many models, the main reason to calculate the average cosine of the incident angle is to deter-
mine the direct component on the surface. For example, restating Equation 5.1 as an average over a
time interval results in

ha2 ha2 ha2 ha2

GHI =
∫ DNI ⋅ cos(sza)d(ha) / ∫ d(ha) + ∫ DHI ⋅ d(ha) / ∫ d(ha)
ha1 ha1 ha1 ha1
(B.8)

where
ha1 is the start time hour angle
ha2 is the end time hour angle

The second term on the right-hand side of Equation B.8 is just the average value of DHI for the
time period. The  first term of the right-hand side of Equation  B.8 is not  equal to the average
DNI times the average value of the cosine of the solar zenith angle, cos(sza), because DNI var-
ies in a systematic way over the interval (see how the DNI varies on a clear day in Figure 2.12
in Chapter  2). One method of weighting the cosine of the solar zenith angle is to assume that
DNI is equal to a constant multiplied by a clear-sky function of DNI (CS(DNI)). For a clear day,
the constant would equal 1. For cloudy periods, the envelope of the clear day still constrains the
DNI values, and on average the DNI can be approximated by a constant multiplied by CS(DNI)
(Vignola and McDaniels 1988b).

ha2


DNI ≈ a ⋅ CS ( DNI)dha
ha1
(B.9)

Substituting Equation B.9 into Equation B.8 and replacing a by the ratio DNI divided by the integral
of the clear-sky function over the period, a modified formula for GHI as a function of DNI and DHI
is obtained:

GHI = DNI ⋅ w ⋅ cos( sza) + DHI (B.10)

where the weighted average of the cosine of the solar zenith angle is
ha2

w ⋅ cos ( sza ) =
∫ ha1
CS ( DNI) ⋅ cos( sza)dha
ha2 (B.11)
∫ ha1
CS ( DNI)dha

The DNI contribution to the GHI can then be calculated by multiplying the average value of the
hour by the weighted average of the cosine of the solar zenith angle.
For accurate hourly calculations, this weighting of the cosine of the solar zenith angle becomes
important especially during morning and evening hours (Vignola and McDaniels 1988b).
When calculating DNI from the difference between GHI and DHI, using just the average cosine
of the zenith angle can lead to systematic biases. Weighting the cosine removes the systematic bias
352 Appendix B

but there will still be some variability because clouds affect the DNI during the period. This vari-
ability is mostly random, and weighting with an appropriate clear-sky model results in a more accu-
rate estimate of the average of the cosine of the solar zenith angle.

B.8 SPECTRAL IRRADIANCE MODELING


The spectral distribution of incident radiation is of growing importance, especially for modeling the
performance of photovoltaic (PV) systems because quantum efficiency of a solar cell varies with
wavelength. To model the output of a PV module, the spectral distribution of the incident radia-
tion and the spectral responsivity of the PV module need to be known along with temperature, tilt,
orientation, and inverter characteristics. The focus of this section will be on how to use spectral
irradiance data to model the performance of a pyranometer, PV module, or other devices that are
sensitive to the spectral distribution of incident radiation. Before such an effort can be undertaken,
it is necessary to have spectral data from various times of day and different times of year.

b.8.1 obtaining sPeCtRal data


There is a sparsity of measured spectral irradiance data and very few locations have spectral mea-
surements of the GHI, DNI, and DHI. Chapter 7 has a considerable discussion on the nature of spec-
tral irradiance and the instruments that monitor spectral data. Because of the sparsity of measured
spectral data, considerable effort has gone into modeling the GHI, DNI, and DHI spectral irradi-
ance. The interaction of sunlight (photons) with molecules in the atmosphere has been well docu-
mented on the molecular and quantum mechanical level. The modeling of light passing through
the atmosphere is known as radiative transfer modeling and is complex and requires knowledge
of the molecules, aerosols, and clouds and their distribution in the atmosphere. One such model
is the BRITE code by Blättner et al. (1974). While the physics is well understood, the computa-
tional effort to obtain realistic results is considerable. Computer models such as LOWTRAN (LOW
resolution atmospheric TRANsmission) and MODTRAN® (MODerate resolution atmospheric
TRANsmission) (Berk 2014) have been developed to emulate the more complex computations, but
even these programs require considerable computer power. There are several versions of each model
as improvements or enhancements have been added.
Simpler spectral computer models, SPCTRAL (Bird 1979, 1980, 1981), and Simple Model of the
Atmospheric Radiative Transfer of Sunshine (SMARTS; Gueymard 2004) have been developed to
emulate the MODTRAN® and LOWTRAN results. These models still require some knowledge and
effort to use but they are available for download (Myers 2017).
These models require specific information about the atmosphere to yield proper results and are
designed to work under cloudless conditions. The  models will generate spectral irradiance for a
specified time and place at specific wavelength intervals. To generate spectral data over the day,
these models have to be run multiple times and the input atmospheric parameters should be changed
over the day if the data are available. Precise data on the atmosphere is scarce, but improvements in
technology have enabled satellites to provide better estimates of these variables.
These models work under clear-sky conditions. Attempts are being made to incorporate satellite
images with these spectral models to generate spectral data during all weather conditions (Xie
and Sengupta 2017, 2018a, 2018b). Much like the satellite-derived solar radiation databases,
satellite-derived spectral solar radiation databases are to be expected. To validate the satellite spec-
tral database, ground-based spectral measurements will be essential.

b.8.2 use of sPeCtRal data


To use spectral data to analyze the performance of photodiode-based pyranometers, PV mod-
ules, or other spectrally sensitive instruments, it is important to know the spectral responsivity
Appendix B 353

R S of the devices. To use this R S with the spectral data, the spectral values must be for the same
wavelength intervals. For example, if the spectral data are at 1-nanometer (nm) intervals, the
R S has to be obtained for the same intervals. Splines or linear fits can be used to create 1-nm
intervals from the R S obtained at limited number of wavelengths. The  R S of an instrument
typically does not change rapidly. It is also necessary to be sure that the spectral values have a
uniform separation. If the spectral data are at 1-nm intervals from 350 to 1,050 nm and at 3-nm
intervals from 1,050 to 4,000 nm, then a spline or linear fit has to be used to get spectral val-
ues within each uniform interval. Often measured spectral data are over a limited wavelength
range. It is necessary to use a model to get spectral values over a range that covers all respon-
sive wavelengths of the device. Alternatively, the total irradiance measured by a pyranometer
can be substituted for the sum of spectral irradiance over all wavelengths (the denominator in
Equation B.12).
Then the average responsivity for the device at a specific time under a specific spectral distribu-
tion can be obtained by summing the product of the irradiance times RS and dividing that sum by
the sum of the spectral irradiance over the same interval:


λ =280 nm
I (λ ) ⋅ RS (λ )
R= λ = 4000 nm
(B.12)

λ =280 nm
I (λ )
λ = 4000 nm
where
λ is the wavelength
I(λ) is the spectral irradiance
RS(λ) is the spectral responsivity of the device

The value used for RS is often relative and not absolute. Therefore, it is best to calibrate R under a
given set of conditions and use that value to normalize RS . For example, the average responsivity of
a photodiode pyranometer is typically determined at a solar zenith angle of 45° under clear skies.
With this information and modeled spectral radiation, the magnitude of RS can be determined and
used in other analyses (Vignola 2016).

B.9 FORECASTING SOLAR IRRADIANCE


b.9.1 intRoduCtion
Appropriately integrating electricity generated by a solar power facility into the regional power grid
requires knowledge of the solar resource and the ability to forecast the variation in the amount of
electricity generated. Near-term forecasting is especially importation for utility operators charged
with maintaining the stability of the electrical grid and plant operators to manage the generation
efficiently. Various forecast methods are now in use and continue to be developed to address grid
operator needs with forecast horizons that range from intra-hour, intra-day, day-ahead, seasonal, and
beyond. Based on the forecast time-scale, these methods use on-site measurements, cloud motion
detection from ground-based sky imagery or satellite observations, numerical weather prediction
results, or climatological data.
Numerous references on the rational for solar forecasting are contained in Kleissl (2013) and
Sengupta et al (2017). The main focus in this section is to discuss the solar irradiance data that are
used by the various forecasting models. The solar resource varies over the day and over the seasons.
There is a saying in Oregon, if you don’t like the weather, wait 10 minutes. While this maxim is an
exaggeration, it does illustrate the variability of the solar resource and the need to understand this
variability. Much like predicting the weather, forecasting the future solar irradiance is a useful skill
but has limitations.
354 Appendix B

As described in Chapter 2, the GHI consists of a DNI component and a DHI component.
Each component has its unique characteristics. When considering global irradiance incident on a
tilted or tracking surface (GTI) the DNI and the DHI have to be projected onto the tilted or tracking
surface. In addition a ground reflected component also contributes to the GTI.
Many forecasting models concentrate on estimating the GHI component, but the DNI component
is the main contributor to a solar energy system, and the DHI component also should be forecast to
estimate the corresponding GTI. If only GHI is forecast, then the GHI has to be separated into the
DNI and DHI components and this adds uncertainty to the predicted values. A general overview of
the forecasting methods can be found in Inman et al. (2013), Kleissl (2013), and Antonanzas et al.
(2016).

b.9.2 foReCasting tHe dni ComPonent


Conceptually, forecasting DNI is simple. When there is no cloud blocking the Sun, the DNI can be
assumed to be the clear-sky value. When clouds block the Sun, the DNI is zero. DNI values should
rarely get above the clear-sky value because there is no physical mechanism that increases the DNI
component as there are unique atmospheric optics like reflections from clouds that can increase the
DHI component. Dividing the DNI by the clear-sky value normalizes the DNI values between 0
and 1. Therefore, the distribution of normalized DNI values peaks around 1 and 0. Since clear-sky
DNI models are dependent on atmospheric constituents that vary with time of day and over the year,
there will be times when the measured DNI will be slightly greater than the modeled clear-sky DNI
value. These values are valid and should be included in any analysis. The distribution of normal-
ized DNI values will change as the time interval for the DNI values increases from 1 minute to
1 hour. The hourly values will be averaged over 60 1-minute values that can be composed of a range
of normalize DNI values depending on the sky conditions. Therefore, forecasting for 1-minute,
15-minute, or hourly values likely will have different characteristics.
To predict DNI or other irradiance values, it is necessary to know and characterize the cloud
cover that will move over the area of interest. With knowledge of the forecasted cloud cover either
from the National Weather Service (NWS) or sky imagers and a history of the site-specific solar
resource climatology, the forecast can be generated. Some methods use the clear-sky values in the
models while others just use time of day to account for the systematic variation of irradiance under
clear-sky conditions. The clear-sky estimates of the DNI are affected by the amount and type of
aerosols in the atmosphere along with the amount of dust loading. Estimates of the amounts of aero-
sols and dust should be incorporated into irradiance estimates.

b.9.3 foReCasting tHe dHi ComPonent


Although there is a considerable body of work discussing the forecasting of the GHI, very little effort
has gone into forecasting the DHI component. It is possible to calculate the DHI component if the
GHI and DNI components are known, but it is worthwhile to discuss the characteristics of the DHI
component that should be considered when forecasting. Under clear skies, DHI increase from zero
in the first hour or two to a value that is relatively constant over the day until the hour or two before
sunset when it decreases to zero. The DHI component increases as aerosol and dust particles in the
atmosphere increase. Therefore, the aerosol and dust loading of the atmosphere are important factors
when estimating the clear-sky magnitude of the DHI component.
Under cloudy conditions, the magnitude of the DHI varies considerably and this variation depends
on the nature of the cloud cover. Under intermittent cloud cover, the DHI can vary dramatically, just
like the DNI. In addition, reflections from clouds can significantly increase the DHI component.
On short time intervals such as 1-minute, the DHI component can increase the GHI value above the
extraterrestrial value and often increases it above the clear-sky GHI value. Thin clouds can produce
DNI values close to 0 while yielding DHI values about one-half of the clear-sky GHI value. Under
Appendix B 355

heavy cloud cover, the DHI component can again be small. So, while the cloud cover can almost act
as a switch for the DNI irradiance, the effect on the DHI value is more varied. Forecasts of the DHI
component are significantly influenced by the nature of the cloud cover.

b.9.4 ConsideRations foR foReCasting global HoRiZontal iRRadianCe


As introduced in Chapter 2, GHI is equal to the DNI projected onto the surface plus the DHI.
Therefore, the considerations for forecasting the GHI are a combination of the considerations for
forecasting the DNI and DHI components. While the clear-sky estimates for GHI can be used,
GHI values above the clear-sky value do occur and any forecasting model should take this into
consideration. The  variation of the GHI is less than the DNI or DHI because some of the light
scattered from the beam radiation ends up as DHI. This outcome dampens the swings of the GHI
values. The exception to this is when the sky is partially cloudy, and the cloud passes near the Sun.
The cloud can reflect the radiation onto the sensor and greatly enhance the GHI.

b.9.5 ConsideRations foR foReCasting tHe global tilted iRRadianCe


As with GHI, the considerations for the GTI forecasts are similar to those for the DNI and DHI
forecasts. In addition, it is necessary to consider the ground-reflected irradiance. Ground-reflected
irradiance is dependent on the ground cover and the factors that affect the reflectivity of the ground
as well as the GHI irradiance on the ground. Snow or rain can significantly affect the reflectivity
of the ground and therefore these weather conditions should be included in the forecasting model
for GTI values. Changing vegetation over the year also can affect the ground-reflected irradiance.
The ground-reflected irradiance is more important for flat receivers with large tilts or flat receivers
on tracking systems because they intercept more ground-reflected irradiance.
Although irradiance is a key driver for solar power output, other environmental factors also
impact the power production. Ambient temperature, humidity, wind speed, and wind direction
can impact power production. Wind gusts, sand storms, hail, and other weather events also can
affect production and proper forecasting of these events should be and often are included in any
forecasting package.

b.9.6 foReCasting metHods


There are a variety of methods used for forecasting solar irradiance. They range from statistical
analysis of surface meteorological observations to the use of sky imagers or satellite images that
are used to calculate cloud motion and model cloud behavior. Forecast models can also be based
on numerical weather prediction (NWP).
The choice of forecast models depends on the time period of the forecast and the type of input
data available. Three types of forecasts will be briefly discussed: intra-hour forecast with high spatial
and temporal resolution; forecasts 4 to 6 hours ahead; and day-ahead forecasts. The performance of
a forecast model can be gauged by comparing the modeled results against a simple persistence fore-
cast model. The persistence forecast model essentially says that the weather now is like the weather
will be in the next interval. The better the model performs compared with the persistence model, the
more reliable the model is. See Jolliffee and Stephenson (2011) for an overview of forecast evaluation
methods.

B.9.6.1 Intra-hour Forecasts


Intra-hour forecasts require on-site irradiance and/or cloud cover data from ground-based irradiance
measurements or sky imagers. Satellite images typically do not  have enough temporal resolution
for intra-hour forecasts, but longer term forecasts always can be used to augment short-term fore-
casts. The irradiance data and cloud images are input into models using statistical methods, including
356 Appendix B

machine-learning techniques. For these models it is important to identify the cloud movements and
characteristics as they are the driving factor affecting the incident irradiance. The behavior of clouds
is quite chaotic with different types of clouds affecting the irradiance differently. Machine learning
and artificial intelligence are important tools that are used in intra-hour forecasts (Arbizu-Barrena
et al. 2017; Boland and Soubdhan 2015; Chow et al. 2011; Graditi et al. 2016; Huang et al. 2013;
Hammer et al. 1999; Lonij et al. 2013; Lee et al. 2017; Marquez and Coimbra 2013; Perez et al. 2010;
Quesada-Ruiz et al. 2014; Schmidt et al. 2016; Voyant et al. 2014). Conventional auto-regressive meth-
ods can provide useful forecasts a few hours ahead under stable conditions, but forecasts are particu-
larly important under changing conditions. The local sky conditions obtained from sky imagers help
provide forecasts on the order of a few minutes and have spatial resolutions of 10–100 meters (m).
These forecasts are usually used to predict the cloud conditions approximately 15 minutes ahead.

B.9.6.2 Four to Six Hour Ahead Forecasts


Forecasts four to six hours ahead often are referred to as “nowcasts.” Typically, nowcasts extrapo-
late cloud cover using vector models based on satellite images. The current generation of geosyn-
chronous weather satellites have images with a spatial resolution between 1 and 5 kilometers (km)
and images are taken every 10–30 minutes. Other satellite and ground-based measurements aug-
ment the images that are used to analyze cloud type and movement.

B.9.6.3 Day Ahead Forecasts


Forecasts that are six hours to several day ahead often use NWP data generated by the NWS. These
NWP data are used in local weather forecasts to predict the evolution and characteristics of clouds
and involve complex models that are based on atmospheric physics and data that are gathered from
many sources. Worldwide NWP models have a spatial resolution of 0.1–0.5 degrees and a temporal
resolution of 1–3 hours. Regional models exist that downscale the NWP data to local areas and often
are used in day ahead forecasts (Jimenez et al. 2016a, 2016b; Lara-Fanego et al. 2012; Mathiesen
and Kleissl 2011; Mathiesen et al. 2013; Ohtake et al. 2013; Perez et al. 2013; Pelland et al. 2013).
Combinations of these techniques also exist (Dambreville et al. 2014; Marquez and Coimbra
2011; Marquez et  al. 2013; Mazorra Aguiar et  al. 2016; Perez et  al. 2014; Wolff et  al. 2016).
For  more information on forecast modeling see (Antonanzas et  al. 2016; Inman et  al. 2013;
Sengupta et al. 2017).

B.10 WHICH MODELS TO USE?


There are a variety of models that exist for any one task. The best model to choose is not always
clear even in review articles that compare models with a single set of data from a variety of loca-
tions. It  is important to keep in mind the uncertainty and characteristics of the data input into
the forecasting model. Models derived from the best instruments available usually have the best
performance but using data with systematic biases will carry those systematic biases throughout
the modeling process. The user is left with three choices, (1) use the existing data and include the
uncertainties associated with these data into the uncertainty in the results; (2) attempt to remove
the systematic biases from these data (This method should only be attempted if the user has a clear
understanding of the systematic biases and documents the way the biases are removed); (3) is to
purchase better equipment for measuring the solar resources.
Of course, the quality of the starting data is very important. Proper maintenance and calibration of
the instruments are a major concern. Therefore, it is important to use forecast models developed with
the best solar resource data, otherwise systematic biases will skew the results. When using a model,
read the literature comparing the various models and understand the weaknesses and strengths of the
model being used including the quality of the data collection and instrumentation used to develop the
model. A comprehensive solar forecast addresses the expected irradiance(s), the variability (ramping)
within the forecast horizon, and the estimated forecast uncertainty.
Appendix B 357

QUESTIONS
1. What is meant by clearness index?
2. Why is the diffuse fraction used in correlations with GHI instead of DHI?
3. What is a clear-sky model?
4. Approximately what is the maximum value of hourly kt and kb values?
5. Should the model used to estimate a variable be derived with similar instruments used to
make the GHI measurement? Discuss.
6. When the average value of the cosine of the solar zenith angle is determined, would it be
appropriate to weigh the average with a clear-sky model?

REFERENCES
American Society of Heating, Refrigerating, and Air-Conditioning Engineers. 1971. Procedure for
Determining Heating and Cooling Loads for Computerized Energy Calculations—Algorithms for
Building Heat Transfer Subroutines. New York: ASHRAE.
American Society of Heating, Refrigerating, and Air-Conditioning Engineers. 2005. Handbook of
Fundamentals, SI edition. Atlanta, GA: ASHRAE
Antonanzas, J., N. Osorio, R. Escobar, R. Urraca, F.-J. Martinez-de-Pison, and F. Antonanzas-Torres. 2016.
Review of photovoltaic power forecasting. Solar Energy 136:78–111. doi:10.1016/j.solener.2016.06.069.
Arbizu-Barrena, C., J. A. Ruiz-Arias, F. J. Rodríguez-Benítez, D. Pozo-Vázquez, and J. Tovar-Pescador. 2017.
Short-term solar radiation forecasting by advecting and diffusing MSG cloud index. Solar Energy
155:1092–1103.
ASHRAE. See American Society of Heating, Refrigerating, and Air-Conditioning Engineers.
Atwater, M. A., and J. T. Ball. 1978. A numerical solar radiation model based on standard meteorological
observations. Solar Energy 21:163–170.
Atwater, M. A., and J. T. Ball. 1979. Erratam. Solar Energy 23:275.
Badescu, V., C. A. Gueymard, S. Cheval, C. Oprea, M. Baciu, A. Dumitrescu, F. Iacobescu, I. Milos, and
C. Rada. 2013. Accuracy analysis for fifty-four clear-sky solar radiation models using routine hourly
global irradiance measurements in Romania. Renewable Energy 55:85–103.
Bemporad, A. 1904. Zur Theorie der Extinktion des Lichtes in der Erdatmosphare. Mitteilungen der
Grossherzoglichen Sternmwarte Zu Heidelberg. No. 4, 1–78. Karlsruhe, Germany: G. Braun.
Berk, A., P. Conforti, R. Kennett, T. Perkins, F. Hawes, and J. van den Bosch. 2014. MODTRAN6: A major
upgrade of the MODTRAN radiative transfer code. Proceedings of the SPIE 9088, Algorithms and
Technologies for Multispectral, Hyperspectral, and Ultraspectral Imagery XX, 90880H (June  13,
2014). doi:10.1117/12.2050433.
Bird, R. E., and R. L. Hulstrom. 1979. Application of Monte Carlo Techniques to Insolation Characterization
and Prediction. SERI/RR36-306. Golden, CO: Solar Energy Research Institute.
Bird, R. E., and R. L. Hulstrom. 1980. Direct Insolation Models. SERI/TR-335-344. Golden, CO: Solar
Energy Research Institute.
Bird, R. E., and R. L. Hulstrom. 1981. Simplified clear-sky model for direct and diffuse insolation on horizon-
tal surfaces, Technical Report SERI/TR-642-761. Golden, CO: Solar Energy Research Institute.
Blättner, W. G., H. G. Horak, D. G. Collins, and M. B. Wells, 1974. Monte Carlo studies of the sky radiation
at twilight. Applied Optics 13:534–537.
Boland, J., and T. Soubdhan. 2015. Spatial-temporal forecasting of solar radiation. Renewable Energy
75:607–616. doi:10.1016/j.renene.2014.10.035.
Bremer, P. 1983. Modeles de Rayonnement Siples Testes Sous Ie Soleil Suisse., October 17–18, EPFL-GRES EDS.
Lausanne, Switzerland: Quatrieme Symposium Recherche et Developpementen Energie Solaire en Suisse.
Brink, G. J. van den. 1982. Climatology of solar irradiance on inclined surfaces, Pt II: Validation of calcula-
tion models. Report 803.229  IV-2 (ESF-006-80  NL B). Delft, the Netherlands: Technisch Physische
Dienst TNO-TH.
Bugler, J. W. 1977. The determination of hourly insolation on an inclined plane using a diffuse irradiance
model based on hourly measured global horizontal insolation. Solar Energy 19:477–491.
Chow, C.W., et al. 2011. Intra-hour forecasting with a total sky imager at the UC San Diego solar energy test-
bed. Solar Energy 85 (11): 2881–2893. doi:10.1016/j.solener.2011.08.025.
Cohen, B., and N. Zerpa. 1982. Spatial transformations of insolation measurements: Preliminary results. Solar
Energy 28:75–76.
358 Appendix B

Collares-Pereira, M., and A. Rabl. 1979a. Derivation of a method for predicting long term average energy
delivery of solar collectors. Solar Energy 23:223–233.
Collares-Pereira, M., and A. Rabl. 1979b. Simple procedures for predicting long term average performance of
nonconcentrating and of concentrating solar collectors. Solar Energy 23:235–253.
Collares-Pereira, M., and A. Rabl. 1979c. The average distribution of solar radiation: Correlations between
diffuse and hemispherical and between daily and hourly insolation values. Solar Energy 22:155–164.
Dambreville, R., P. Blanc, J. Chanussot, and D. Boldo. 2014. Very short-term forecasting of the global
horizontal irradiance using a spatio-temporal autoregressive model. Renewable Energy 72:291–300.
doi:10.1016/j.renene.2014.07.012.
Desbucam, U. V., B. G. Petrovic, and D. Desnica. 1986. Calculation of monthly average daily insolation
on tilted, variously oriented surfaces using anyalytically weighted Rb factors. Solar Energy 37:81–90.
doi:10.1016/0038-092X(86)90066-6.
Davies, J. A., and J. E. Hay. 1979. Calculation of the solar radiation incident on a horizontal surface.
In Proceedings of the First Canadian Solar Radiation Data Workshop, April 17–19. Ottawa, Canada:
Canadian Atmospheric Environment Service.
Evseev, E. G., and A. I. Kudish. 2009. The assessment of different models to predict the global solar radiation
on a surface tilted to the south. Solar Energy 83:377–388.
Graditi, G., S. Ferlito, and G. Adinolfi. 2016. Comparison of photovoltaic plant power production prediction meth-
ods using a large measured dataset. Renewable Energy 90:513–519. doi:10.1016/j.renene.2016.01.027.
Gueymard, C. 1983. Utilisation des donnees meteorologiques horaires pour Ie calcul du rayonnement solaire
sur des batiments solaires passifs. Ph. D. thesis, University of Montreal.
Gueymard, C. 1985. Radiation on tilted planes: A physical model adaptable to any computational time-step.
Presented at INTERSOL 85, Congress of the International Solar Energy Society, June, Montreal, Canada.
Gueymard, C. 1986a. Mean daily averages of beam radiation received by tilted surfaces as affected by the
atmosphere. Solar Energy 37:261–267.
Gueymard, C. 1986b. Monthly averages of the daily effective optical air mass and solar related angles for
horizontal and inclined surfaces. Journal of Solar Energy Engineering 108:320–324.
Gueymard, C. A. 1987. An anisotropic solar irradiance for tilted surfaces and its comparison with selected
engineering algorithms. Solar Energy 38:367–386.
Gueymard, C. A. 2009. Direct and indirect uncertainties in the prediction of tilted irradiance for solar engi-
neering applications. Solar Energy 83:432–444.
Gueymard, C. A. 2004. The sun’s total and spectral irradiance for solar energy applications and solar radiation
models. Solar Energy 76:423–453.
Hammer, A., D. Heinemann, E. Lorenz, and B. Lückehe. 1999. Short-term forecasting of solar radia-
tion: A  statistical approach using satellite data. Solar Energy 67 (1–3):139–150. doi:10.1016/
S0038-092X(00)00038-4.
Hay, J. E. 1979. A  study of shortwave radiation on non-horizontal surfaces. Final report, Contract Serial
Number OS878 00053. Downsview, ON: Atmospheric Environment Service.
Hay, J. E. 1984. Calculation of solar irradiances for inclined surfaces: Verification of models which use hourly
and daily data. Interim Report, Contract 02SE.KM147-3-3029 (DSS I AES), June 22. Downsview, ON:
Canadian Atmospheric Environment Service.
Hay, J. E. 1993. Calculating solar radiation for inclined surfaces: Practical approaches. Renewable Energy
3 (4–5): 373–380.
Hay, J. E., and D. C. McKay. 1985. Estimating solar irradiance on inclined surfaces: A review and assessment
of methodologies. International Journal of Solar Energy 3:203–240.
Hay, J. E., and D. C. McKay. 1988. Final Report IEA Task IX-Calculation of solar irradiances for inclined
surfaces: Verification of models which use hourly and daily data. Downsview, ON: International Energy
Agency Solar Heating and Cooling Programme.
Hay, J. E., and J. A. Davies. 1980. Calculation of the solar radiation incident on an inclined surface.
In  Proceedings of the First Canadian Solar Radiation Data Workshop, ed. E. Hay and T. K. Won.
Downsview, ON: Atmospheric Environment Service.
Hay, J. E., R. Perez, and D. C. McKay. 1986. Addendum and errata to the paper, estimating solar irradiance on
inclined surfaces: A review and assessment of methodologies. International Journal of Solar Energy
4:320–324.
Hoyt, D. V. 1978. A model for the calculation of solar global insolation. Solar Energy 21:27–35.
Huang, J., M. Korolkiewicz, M. Agrawal, and J. Boland. 2013. Forecasting solar radiation on an hourly time
scale using a coupled autoregressive and dynamical system (CARDS) model. Solar Energy 87:136–149.
doi:10.1016/j.solener.2012.10.012.
Appendix B 359

Ineichen, P. 1983. Quatre annees de mesures d’ensoleillement a Geneve 1978–1982. Thesis, Geneva University.
Inman, R. H., H. T. C. Pedro, and C. F. M. Coimbra. 2013. Solar forecasting methods for renewable energy
integration. Progress in Energy and Combustion Science 39:535–576. doi:10.1016/j.pecs.2013.06.002.
Iqbal, M. 1983. An Introduction to Solar Radiation. Toronto, Canada: Academic Press.
Jimenez, P. A., J. P. Hacker, J. Dudhia, S. E. Haupt, J. A. Ruiz-Arias, C. A. Gueymard, G. Thompson,
T. Eidhammer, and A. Deng. 2016a. WRF-solar: An augmented NWP model for solar power predic-
tion. Model description and clear-sky assessment. Bulletin of the American Meteorological Society
97:1249–1264. doi:10.1175/BAMS-D-14-00279.1.
Jimenez, P. A., S. Alessandrini, S. E. Haupt, A. Deng, B. Kosovic, J. A. Lee, and L. D. Monache. 2016b.
The  role of unresolved clouds on short-range global horizontal irradiance predictability. Monthly
Weather Review 144 (9): 3099–3107. doi:10.1175/MWR-D-16-0104.1.
Jones, R. E. 1980. Effects of overhang shading of windows having arbitrary azimuth. Solar Energy 24:305–312.
Jolliffe I.T., and D. B. Stephenson, Eds. 2011. Forecast Verification: A Practitioner’s Guide in Atmospheric
Science, 2nd edition. New York: John Wiley & Sons.
Kasten, F. 1964. A new table and approximation formula for the relative optical air mass. Technical Report 136.
Hanover, Germany: U.S. Army Material Command, Cold Region Research and Engineering Laboratory.
Klein, S. A. 1977. Calculation of monthly average insolation on tilted surfaces. Solar Energy 19:325–329.
Klein, S. A., and J. C. Theilacker. 1981. An algorithm for calculating monthly-averaged radiation on inclined
surfaces. Journal of Solar Energy Engineering 103:29–33.
Kleissl, J., Ed. 2013. Solar Energy Forecasting and Resource Assessment, 503pp. Oxford, UK: Elsevier.
Klucher, T. M. 1979. Evaluation of models to predict insolation on tilted surfaces. Solar Energy 23:111–114.
Kondratyev, K. Y., and M. P. Manolova. 1960. The radiation balance of slopes. Solar Energy 4:14–19.
Kondratyev, K. Y. 1969. Radiation in the Atmosphere. New York: Academic Press.
Kondratyev, K. Y. 1977. Radiation regime on inclined surfaces. Technical Note 152, WMO-No. 467. Geneva,
Switzerland: World Meteorological Organization.
Kusuda, T. 1976. Procedure employed by the ASHRAE task group for the determination of heating and cool-
ing loads for building energy analysis. ASHRAE Transactions 82:305–314.
Lacis, A. L., and J. E. Hansen. 1974. A  parameterization for absorption of solar radiation in the earth’s
atmosphere. Journal of Atmospheric Science 31:118–133.
Lara-Fanego, V., J. A. Ruiz-Arias, D. Pozo-Vázquez, F. J. Santos-Alamillos, and J. Tovar-Pescador. 2012.
Evaluation of the WRF model solar irradiance forecasts in Andalusia (Southern Spain). Solar Energy
86 (8): 2,200–2,217. doi:10.1016/j.solener.2011.02.014.
Lawrence Berkeley Laboratories. 1982. DOE-2 engineers manual, version 2.1 A, LBL11353 (LA-8520-M),
(DE83004575), Technical Information Center. Washington, DC: United States Department of Energy.
Lee, J. A., S. E. Haupt, P. A. Jimenez, M. A. Rogers, S. D. Miller, and T. C. McCandless. 2017. Solar irradi-
ance nowcasting case studies near Sacramento. Journal of Applied Meteorology and Climatology 56 (1):
85–108. doi:10.1175/JAMC-D-16-0183.1.
Liu, B. Y. H., and R. C. Jordan. 1960. The interrelationship and characteristic distribution of direct, diffuse
and total solar radiation. Solar Energy 4:1–9.
Liu, B. Y. H., and R. C. Jordan. 1962. Daily insolation on surfaces tilted toward the equator. Transactions,
American Society of Heating, Refrigerating and Air Conditioning Engineers 67:526–541.
Liu, B. Y. H., and R. C. Jordan. 1963. The long-term average performance of flat-plate solar energy collectors.
Solar Energy 7:53–74.
Lonij, V. P. A., A. E. Brooks, A. D. Cronin, M. Leuthold, and K. Koch. 2013. Intra-hour forecasts of solar
power production using measurements from a network of irradiance sensors. Solar Energy 97:58–66.
doi:10.1016/j.solener.2013.08.002.
Ma, C. C. Y., and M. Iqbal. 1983. Statistical comparison of models for estimating solar radiation on inclined
surfaces. Solar Energy 31:313–317.
Mahaptra, A. K. 1973. An evaluation of a spectroradiometer for the visible-ultraviolet and near-ultraviolet.
Ph. D. thesis, University of Missouri. (University Microfilms 74-9964).
Marquez, R., and C. F. M. Coimbra. 2011. Forecasting of global and direct solar irradiance using stochastic
learning methods, ground experiments and the NWS database. Solar Energy 85 (5):746–756.
doi:10.1016/j.solener.2011.01.007.
Marquez, R., and C. F. M. Coimbra. 2013. Intra-Hour DNI forecasting based on cloud tracking image analysis.
Solar Energy 91:327–336. doi:10.1016/j.solener.2012.09.018.
Marquez, R., H. T. C. Pedro, and C. F. M. Coimbra. 2013. Hybrid solar forecasting method uses satel-
lite imaging and ground telemetry as inputs to ANNs. Solar Energy 92:176–188. doi:10.1016/j.
solener.2013.02.023.
360 Appendix B

Mathiesen, P., and J. Kleissl. 2011. Evaluation of numerical weather prediction for intra-day solar forecasting
in the continental United States. Solar Energy 85(5):967–977.
Mathiesen, P., J. Kleissl, and C. Collier. 2013. Case studies of solar forecasting with the weather research
and forecasting model at GL-Garrad Hassan. In Solar Energy Forecasting and Resource Assessment,
ed. J. Kleissl, 357–382. Boston, MA: Academic Press. doi:10.1016/B978-0-12-397177-7.00014-0.
Mazorra Aguiar, L., B. Pereira, P. Lauret, F. Díaz, and M. David. 2016. Combining solar irradiance measure-
ments, satellite-derived data and a numerical weather prediction model to improve intra-day solar fore-
casting. Renewable Energy 97:599–610. doi:10.1016/j.renene.2016.06.018.
McDonald, J. E. 1960. Direct absorption of solar radiation by atmospheric water vapor. Journal of Meteorology
17:319–328.
McFarland, R. D. 1983. Solar radiation approximations. In  Appendix E: Passive Solar Design Handbook,
Vol. 3: Passive Solar Design Analysis and Supplement, ed. R. W. Jones, pp. 293–302. Boulder, CO:
American Solar Energy Society.
Moon, P. 1940. Proposed standard solar-radiation curves for engineering use. Journal of the Franklin Institute
230:583–617.
Moon, P., and D. E. Spencer. 1942. Illumination from a non-uniform sky. Transactions of the Illumination
Engineering Society 37:707–726.
Muneer, T. ed. 2004. Solar Radiation and Daylight Models, 2d ed. New York: Elsevier.
Myers, D. 2017. Solar Radiation: Practical Modeling for Renewable Energy Applications. Hoboken, NJ: CRC
Press.
Oegema, S. W. T. M. 1971. De digitale berekening van de zonnestraling. Report 332-111, Technisch Physische
Dienst TNO-TH, Delft, the Netherlands.
Ohtake, H., K. Shimose, J. Gari da Silva Fonseca Jr., T. Takashima, T. Oozeki, and Y. Yamada. 2013. Accuracy
of the solar irradiance forecasts of the Japan meteorological agency mesoscale model for the Kanto
Region, Japan. Solar Energy 98 (Part B): 138–152. doi:10.1016/j.solener.2012.10.007.
Page, J. K. 1961. The estimation of monthly mean values of total daily solar short-wave radiation on vertical
and inclined surfaces from sunshine records for latitudes 40N-40S. Proceedings of the United Nations
Conference on New Sources of Energy, Paper 35/5/98, 378–389.
Page, J. K. 1978. Methods for the Estimation of Solar Energy on Vertical and Inclined Surfaces. Sheffield,
UK: BS 46, Department of Building Science, University of Sheffield.
Pelland, S., G. Galanis, and G. Kallos. 2013. Solar and photovoltaic forecasting through post-processing of
the global environmental multiscale numerical weather prediction model. Progress in Photovoltaics:
Research and Applications 21:284–296. doi:10.1002/pip.1180.
Perez, R. R., J. T. Scott, and R. Stewart. 1983. An anisotropic model for diffuse radiation incident on slopes of
different orientations and possible applications to CPC’s. Progress in Solar Energy 25:85–90.
Perez, R., A. Kankiewicz, J. Schlemmer, K. Hemker, and S. Kivalov. 2014. A new operational solar resource
forecast model service for PV fleet simulation. Paper presented at the IEEE 40th Photovoltaic Specialist
Conference (PVSC), June 8–13, Denver, CO. doi:10.1109/PVSC.2014.6925204
Perez, R., E. Lorenz, S. Pelland, M. Beauharnois, G. Van Knowe, K. Hemker Jr., D. Heinemann et al. 2013.
Comparison of numerical weather prediction solar irradiance forecasts in the U.S., Canada, and Europe.
Solar Energy 94:305–326. doi:10.1016/j.solener.2013.05.005.
Perez, R., P. Ineichen, R. Seals, J. J. Michalsky, and R. Stewart. 1990. Modeling daylight availability and irra-
diance components from direct and global irradiance. Solar Energy 44:271–289.
Perez, R., S. Kivalov, J. Schlemmer, K. Hemker Jr., D. Renné, and T. E. Hoff. 2010. Validation of short
and medium term operational solar radiation forecasts in the U.S. Solar Energy 84 (12):2161–2172.
doi:10.1016/j.solener.2010.08.014.
Pérez-Burgos, A., M. Díez-Mediavilla, C. Alonso-Tristán, and M. C. Rodríguez-Amigo. 2017. Analysis of
solar direct irradiance models under clear-skies: Evaluation of the improvements for locally adapted
models. Journal of Renewable and Sustainable Energy 9:023703. doi:10.1063/1.4981798.
Puri, V. M., R. Jiminez, and M. Menzer. 1980. Total and non-isotropic diffuse insolation on tilted surfaces.
Solar Energy 25:85–90.
Quesada-Ruiz, S., Y. Chu, J. Tovar-Pescador, H. T. C. Pedro, and C. F. M. Coimbra. 2014. Cloud-
tracking methodology for intra-hour DNI forecasting. Solar Energy 102:267–275. doi:10.1016/j.
solener.2014.01.030.
Randall, C. M., and J. M. Biddle. 1981. September. Hourly estimates of direct insolation: Computer code
ADIPA user’s guide. The Aerospace Corporation, AIR-81 (7878)-1.
Reindl, D. T., W. A. Beckman, and J. A. Duffie. 1990. Evaluation of hourly tilted surface radiation models.
Solar Energy 45:9–17.
Appendix B 361

Revfeim, K. J. A. 1976. Solar radiation at a site of known orientation on the Earth’s surface. Journal of Applied
Meteorology 15:651–656.
Revfeim, K. J. A. 1979. Maximization of global radiation on sloping surfaces. New Zealand Journal of Science
22:293–297.
Revfeim, K. J. A. 1982a. Estimating global solar radiation on sloping surfaces. New Zealand Journal of
Agricultural Research 25:281–283.
Revfeim, K. J. A. 1982b. Simplified relationships for estimating solar radiation incident on any flat surface.
Solar Energy 2:509–517.
Rogers, G. G., J. K. Page, and C. G. Souster. 1979. Mathematical models for estimating the irradiance fall-
ing on inclined surfaces for clear, overcast and average conditions. In Proceedings of the Meteorology
for Solar Energy Applications, Conference (C18) at the Royal Institution, 48–62. Freiburg, Germany:
International Solar Energy Society.
Schmidt, T., J. Kalisch, E. Lorenz, and D. Heinemann. 2016. Evaluating the spatio-temporal performance
of sky-imager-based solar irradiance analysis and forecasts. Atmospheric Chemistry and Physics
16:3,399–3,412. doi:10.5194/acp-16-3399-2016.
Sengupta, M., A. Habte, C. Gueymard, S. Wilbert, D. Renné, and T. Stoffel. 2017. Best practices handbook
for the collection and use of solar resource data for solar energy applications, 2nd ed. NREL/TP-5D00-
68886, December 2017.
Skartveit, A., and J. A. Olseth. 1986. Modeling slope irradiances at high latitudes. Solar Energy 36 (4):
333–344.
Temps, R. C., and K. L. Coulson. 1977. Solar radiation incident upon slopes of different orientations. Solar
Energy 19:179–184.
Threkheld, J. L. 1962. Solar radiation of surfaces on clear days. ASHRAE Journal 69:43–54.
Vignola, F. 2003. Beam-tilted correlations. In Proceedings of the Solar 2003 American Solar Energy Society
Conference, Austin, TX.
Vignola, F., and D. K. McDaniels. 1984a. Correlations between diffuse and global insolation for the Pacific
Northwest. Solar Energy 32:161–168.
Vignola, F., and D. K. McDaniels. 1984b. Diffuse-global correlation: Seasonal variations. Solar Energy
33:397–402.
Vignola, F. and D. K. McDaniels. 1988a. Beam-global correlations in the Pacific Northwest. Solar Energy 36:
409–418.
Vignola, F., and D. K. McDanielZs. 1988b. Direct beam radiation: Projection on tilted surfaces. Solar Energy
40:237–247.
Vignola, F., and D. K. McDaniels. 1989. Direct radiation: Ratio between horizontal and tilted surfaces. Solar
Energy 43:183–190.
Vignola, F., Z. Derocher, J. Peterson, L. Vuilleumier, C. Felix, J. Gröbner, and N. Kouremeti. 2016. Effects of
changing spectral radiation distribution on the performance of photodiode pyranometers. Solar Energy
129:224–230.
Voyant, C., C. Darras, M. Muselli, C. Paoli, M.-L. Nivet, and P. Poggi. 2014. Bayesian rules and sto-
chastic models for high accuracy prediction of solar radiation. Applied Energy 114:218–226.
doi:10.1016/j.apenergy.2013.09.051.
Watt, D. 1978. On the Nature and Distribution of Solar Radiation. HCP/T2552-0l. Washington, DC: U.S.
Department of Energy.
Willmott, C. J. 1982. On the climatic optimization of the tilt and azimuth of flat-plate solar collectors.
Solar Energy 28:205–216.
Wolff, B., J. Kühnert, E. Lorenz, O. Kramer, and D. Heinemann. 2016. Comparing support vector regres-
sion for PV power forecasting to a physical modeling approach using measurement, numerical weather
prediction and cloud motion data. Solar Energy 135:197–208. doi:10.1016/j.solener.2016.05.051.
Xie, Y., and M. Sengupta. 2017. Toward improved modeling of spectral solar irradiance for solar energy appli-
cations, European PV Solar Energy Conference and Exhibition, Amsterdam, the Netherlands, September
25–29, 2017. NREL/CP-5D00-70159 https://www.nrel.gov/docs/fy18osti/70159.pdf (Accessed May 31,
2019).
Xie, Y., and M. Sengupta. 2018a. Assessment of Error Sources in the Modeling of POA  Irradiance under
Clear-sky Conditions, 2018 European PV Solar Energy Conference and Exhibition, Brussels, Belgium,
September 24–28, 2018.
Xie, Y., and M. Sengupta. 2018b. Assessment of Error Sources in the Modeling of POA Irradiance under
Clear-sky Conditions.  Presented at the 2018 World Conference on Photovoltaic Energy Conversion
(WCPEC-7), Waikoloa, Hawaii, June 10–15, 2018.
Appendix C: Sunshine Duration

The measurement of sunshine minutes or percent sunshine has been used to estimate irradiance over
the last few decades. This measurement is still made in many parts of the world. A thorough discus-
sion of ways to estimate sunshine duration, their differences, and their uncertainties is explained in
the Commission for Instruments and Methods of Observation (CIMO) Guide (WMO 2014).
Sunshine minutes have been used to estimate daily-integrated global horizontal irradiance using
this equation from Ångström (1924):

GHI / GHI 0 = a + b ⋅ S / S0 (C.1)


In this equation, GHI0 is the daily integrated global irradiance at the top of the atmosphere (TOA), S0 is
the total possible minutes of sunshine during the day, S is the sum of measured minutes of sunshine, and
a and b are constants that are empirically derived for each site. These estimates were used when the use
of pyranometers was not widespread, but they are of little use today where even low-cost pyranometers
make better irradiance measurements than the estimates that are possible using Equation C.1.
Aside: Sunshine duration estimates seem most useful to cities and resorts that want to promote
tourism. In this regard, consider what is meant by a sunny day. We have all heard of Sunny City, USA,
(not a real place, although there is a Sun City, AZ, USA), which has 300 days of sunshine a year. Is
this true? It depends on how we define a sunny day. As we shall see in the following discussion, a
sunshine minute is defined by the WMO as a minute where the average DNI exceeds 120 watts per
square meter (Wm−2). If we define a sunny day as one where we get at least 60 sunshine minutes, then
we could call this a place that receives 300 sunny days a year. The U.S. National Weather Service
(NWS) uses cloud cover observations to define a clear (sunny) day. If the average observation indi-
cates less than 30 percent cloud cover then the NWS classifies it as a clear day. This used to be done
by humans making observations once an hour. Cirrus clouds are considered clouds for this purpose,
but often they transmit DNI greater than 120 Wm−2. Ceilometers now make the cloud observations at
NWS sites, and whether cirrus clouds are detected depends on the maximum height that this instru-
ment reaches. Summarizing, sunny days are in the eye of the promoter. End Aside.
Qualitatively, the Sun is shining brightly whenever the shadow of an object can be discerned.
The predominant measurement of sunshine worldwide has been the Campbell–Stokes sunshine recorder
described in Section 3.3.2. It consists of a glass ball that focuses direct normal irradiance (DNI) on spe-
cially treated paper that is supposed to burn whenever the DNI exceeds 120 Wm−2. This is the WMO
(2014) definition of bright sunshine; however, the Campbell–Stokes burns over a range of conditions.
For  example, in a moist, cold environment, the DNI will need to be higher before the burn occurs.
Figure C.1 is the burn pattern from a Campbell–Stokes sunshine recorder. The handwritten numbers are
estimates of the fraction of the hour (measured in tenths) during which there was bright sunshine.
The very best measurement of sunshine duration is made using a well-calibrated pyrheliometer.
Since pyrheliometers have been built with either a 5 degree or 5.7 degree field-of-view, there is some
discussion concerning the difference, but discrepancies from these minor differences likely will be
insignificant.
The  next best estimate is made using two pyranometers of the same model and calibrated at
the same time. If one is shaded, and a shade-corrected value for diffuse derived as discussed in
Chapter 6, then sunshine minutes are tallied whenever

( GHI − DHI ) ≥ 120 Wm −2 (C.2)


cos ( sza )

363
364 Appendix C

FIGURE C.1 Northern Hemisphere summertime Campbell–Stokes measurement of sunshine duration.


Fractional hours with bright sunshine are estimated by an analyst and hand recorded.

Like this idea, but inferior to it, is the Foster–Foskett sunshine recorder (Foster and Foskett
1953) that was first used at airports by the NWS in 1953. The NWS appears to have decommis-
sioned these instruments in 2009 (see, e.g., http://www.startribune.com/local/62760757.html).
This  instrument used two matched selenium photocells with one shaded and one unshaded.
The shade ring was adjusted four times a year; therefore, it was much broader than the typical
fixed shade band described in Chapter 6. It records a minute of sunshine when the signal differ-
ence exceeds a threshold setting that is manually adjustable on the instrument. Michalsky (1992)
compared this instrument’s measurement of sunshine minutes to a nearby measurement using a
calibrated pyrheliometer. Although there was a reasonable correlation between the two instru-
ments, the two measurements of sunshine differed significantly compared with two collocated
pyrheliometers.
The burn method, employed in the Campbell–Stokes sunshine recorders, is inferior to the meth-
ods that use a pyrheliometer or two pyranometers, but, as stated earlier, it is still widely used outside
the United States.
New instruments have been developed that are probably superior to all methods except the pyr-
heliometer measurement. Two of these newer model sunshine recorders are the MS-093 from EKO
Instruments Co., Ltd. and the CSD 3 from Kipp and Zonen.

REFERENCES
Ångström, A. 1924. Solar and terrestrial radiation. Quarterly Journal of the Royal Meteorological Society
50:121–126.
Foster, N. B., and L. W. Foskett. 1953. A  photoelectric sunshine recorder. Bulletin of the American
Meteorological Society 34:212–215.
Michalsky, J. J. 1992. Comparison of a National Weather Service Foster sunshine recorder and the World
Meteorological Organization for sunshine duration. Solar Energy 48:133–141.
World Meteorological Organization. 2014. Measurement of sunshine duration. In  Guide to Meteorological
Instruments and Observing Practices, WMO-No. 8. http://www.wmo.int/pages/prog/www/IMOP/
CIMO-Guide.html.
WMO. See World Meteorological Organization.
Appendix D: Sun Path Charts

Before installing a solar monitoring station or evaluating the performance of a solar energy system,
it is useful to know the position of the Sun in the sky. Solar algorithms exist (Appendix E) to
calculate the solar position. Sometimes it is useful to get a visual image of the position of the
sun, especially when evaluating shading caused by nearby objects. Usage of sun path charts are
discussed in Chapters 2 and 15. In this appendix, both Cartesian and polar plots are given for every
10° of latitude. Sun path charts for a given latitude can also be created at http://solardat.uoregon.edu/
SunChartProgram.php. Sundial plots, showing the shadow from the tip of a pole as it moves across
the ground, can be created at http://solardat.uoregon.edu/SunDialProgram.html. All these plots are
created from the same solar position information, but the perspective is different.

D.1 CARTESIAN SUN PATH CHARTS


Sun path charts drawn using Cartesian coordinates are illustrated first. The solar elevation (90°–sza)
is the vertical axis, and the solar azimuth is the horizontal axis. True north has an azimuth angle
of 0°. Seven sun paths are shown on the chart: December 21; January 21/November 21; February
20/October 21; March 20/September 22; April 20/August 22; May 21/July 21; and June 21. The paths
on December 21 and June 21 represent the extremes of the sun path and illustrate the paths during
the winter and summer solstices. The days of the months of the paths in between were chosen to be
about 30 days apart so that the spring and fall paths would be identical.
Sun paths charts were created at latitudes 10 degrees apart and from the North Pole to the
South Pole. The time shown on the charts is solar time. This usage means that the Sun is highest in
the sky at solar noon. Plots can be made in local standard time, which is useful if the charts are used
to locate the azimuth angle of obstructions (Figures D.1 through D.19).

D.2 POLAR SUN PATH CHARTS


The sun path charts in Cartesian and polar coordinates show exactly the same solar zenith and solar
azimuth angles used in D.1. The difference is that the polar sun path charts project the sun path onto
polar coordinates. The highest point in the sky has a solar altitude of 90 degrees. Concentric circles
show the decreasing solar altitude at 10 degree increments with zero altitude being the horizon.
The spokes (or radii) of the wheel are the solar azimuth angle and vary from 0° to 360° starting from
0° at true north. The solar azimuth values are given on the outside of the largest circle.
The polar sun paths should be viewed as if the observer was standing in the center of the plot
and looking at the sky dome. The sun path and the hour lines form a grid that changes shape with
latitude. At the poles, the sun paths become circular (Figures D.20 through D.38).

365
366 Appendix D

FIGURE D.1 Cartesian sun path chart facing the equator at latitude 0°N.

FIGURE D.2 Cartesian sun path chart facing the equator at latitude 10°N.
Appendix D 367

FIGURE D.3 Cartesian sun path chart facing the equator at latitude 20°N.

FIGURE D.4 Cartesian sun path chart facing the equator at latitude 30°N.
368 Appendix D

FIGURE D.5 Cartesian sun path chart facing the equator at latitude 40°N.

FIGURE D.6 Cartesian sun path chart facing the equator at latitude 50°N.
Appendix D 369

FIGURE D.7 Cartesian sun path chart facing the equator at latitude 60°N.

FIGURE D.8 Cartesian sun path chart facing the equator at latitude 70°N.
370 Appendix D

FIGURE D.9 Cartesian sun path chart facing the equator at latitude 80°N.

FIGURE D.10 Cartesian sun path chart facing the equator at latitude 90°N.
Appendix D 371

FIGURE D.11 Cartesian sun path chart facing the equator at latitude 10°S.

FIGURE D.12 Cartesian sun path chart facing the equator at latitude 20°S.
372 Appendix D

FIGURE D.13 Cartesian sun path chart facing the equator at latitude 30°S.

FIGURE D.14 Cartesian sun path chart facing the equator at latitude 40°S.
Appendix D 373

FIGURE D.15 Cartesian sun path chart facing the equator at latitude 50°S.

FIGURE D.16 Cartesian sun path chart facing the equator at latitude 60°S.
374 Appendix D

FIGURE D.17 Cartesian sun path chart facing the equator at latitude 70°S.

FIGURE D.18 Cartesian sun path chart facing the equator at latitude 80°S.
Appendix D 375

FIGURE D.19 Cartesian sun path chart facing the equator at latitude 90°S.

FIGURE D.20 Polar sun path chart at latitude 0°N.


376 Appendix D

FIGURE D.21 Polar sun path chart at latitude 10°N.

FIGURE D.22 Polar sun path chart at latitude 20°N.


Appendix D 377

FIGURE D.23 Polar sun path chart at latitude 30°N.

FIGURE D.24 Polar sun path chart at latitude 40°N.


378 Appendix D

FIGURE D.25 Polar sun path chart at latitude 50°N.

FIGURE D.26 Polar sun path chart at latitude 60°N.


Appendix D 379

FIGURE D.27 Polar sun path chart at latitude 70°N.

FIGURE D.28 Polar sun path chart at latitude 80°N.


380 Appendix D

FIGURE D.29 Polar sun path chart at latitude 90°N.

FIGURE D.30 Polar sun path chart at latitude 10°S.


Appendix D 381

FIGURE D.31 Polar sun path chart at latitude 20°S.

FIGURE D.32 Polar sun path chart at latitude 30°S.


382 Appendix D

FIGURE D.33 Polar sun path chart at latitude 40°S.

FIGURE D.34 Polar sun path chart at latitude 50°S.


Appendix D 383

FIGURE D.35 Polar sun path chart at latitude 60°S.

FIGURE D.36 Polar sun path chart at latitude 70°S.


384 Appendix D

FIGURE D.37 Polar sun path chart at latitude 80°S.

FIGURE D.38 Polar sun path chart at latitude 90°S.


Appendix E: Solar Position
Algorithms
Solar position calculations are used for a wide variety of tasks from setting up solar monitoring
stations to positioning photovoltaic (PV) panels to maximize electricity output. Solar position is
used in the analysis of irradiance data and is essential for modeling solar irradiance for arbitrarily
tilted surfaces. Many algorithms are available that estimate the solar position to varying degrees of
accuracy; some of these solar algorithms and their accuracy are discussed in this appendix.
Every year the Astronomical Almanac (2018) publishes next year’s Sun position with accuracies
approaching 0.1 seconds of arc. To achieve these accuracies, the Nautical Almanac Office of the
United States Naval Observatory and Her Majesty’s Nautical Almanac Office within the United
Kingdom Hydrographic Office use an extremely complex algorithm that includes small effects on
the Earth’s orbit, such as lunar and planetary perturbations, plus optical effects such as stellar aber-
ration, parallax, and refraction.
Approximations to the Sun’s position relative to a location on the Earth’s surface have become
increasingly accurate. Blanco-Muriel et al. (2001) give a thorough history of the development of
the more accurate solar position algorithms up to the time of their paper. This  history is briefly
discussed here.
Spencer (1971) provided truncated Fourier series expressions for the equation of time and the
solar declination that can be used to approximate the solar position with maximum biases of between
10 and 20 minutes of arc. These are accurate enough for positioning fixed flat-plate thermal panels
or fixed flat-plate PV panels but are not useful for tracking the Sun for concentrating solar power
installations. As an example, Vant-Hull and Hildebrandt (1976) estimated that a tracking accuracy
of 3.5 minutes of arc would be needed for a concentration ratio of 1,000. Therefore, Spencer’s (1971)
approximations are inadequate for this application.
In the late 1970s, at about the same time and apparently unaware of each other’s efforts, Pitman
and Vant-Hull (1978) and Walraven (1978, 1979) developed simplifications of the Nautical Almanac’s
algorithms to produce much-improved approximations for the solar position. These were accurate
to approximately 1 arc minute. Walraven’s (1978) paper was noteworthy since it provided a Fortran
subroutine to calculate solar position. Walraven (1979) corrected two lines of code in an erratum
within a year of the first publication. Archer (1980) and Wilkinson (1981) followed with a comment
and a note that suggested a correction for refraction to the algorithm. To our knowledge there is no
hyperlink to the corrected code with refraction.
Progress continued with the Fortran program developed by Michalsky (1988a) using the low-
precision formulae published in the Astronomical Almanac (1984). These formulae were expected
to be good to about 0.01 degrees between the years 1950 and 2050. The  formulae in the 2018
version of the Astronomical Almanac (2018) are identical to those in the 1984 publication of this
almanac. An erratum was published that fixed some typesetting issues (Michalsky, 1988b) in the
original, and then another fixed typesetting issues in the erratum (Michalsky, 1989). However,
the most significant issue was corrected in a letter from Spencer (1989) that changed the code to
give the correct assignment to the azimuth in the Southern Hemisphere. The  Fortran program
for sunpos is given at the end of this appendix to illustrate the parameters needed and the ter-
minology used in calculating the position of the Sun. The ftp site ftp://aftp.cmdl.noaa.gov/user/
michalsky/asked_for_stuff/asunpos.f also contains the asunpos Fortran code with all corrections.

385
386 Appendix E

This program has been stable for years. This same algorithm is used in the National Renewable
Energy Laboratory (NREL) solpos program that calculates solar position and much more (avail-
able at http://rredc.nrel.gov/solar/codesandalgorithms/solpos/).
A  little more than a decade had passed since Blanco-Muriel et  al. (2001) published their
Plataforma Solar de Almeria (PSA) algorithm (available as C++ code at http://www.psa.es/sdg/
sunpos.htm). The accuracy was improved over Michalsky’s (1988a) code by restricting the years
approximated to the range 1999–2015, modifying the “low-precision” formulae in the Astronomical
Almanac (1984), and including a correction for parallax.
Reda and Andreas (2004) published solar position code based on Meeus’s (1998) algorithm.
This code is the most accurate to date with a stated accuracy of ±0.0003 degrees for the inclusive
years 2000 BC–AD 6000. Their C code name SPA can be found at http://www.nrel.gov/midc/spa/.
There is a link to an online calculator from that page. Cornwall provides a Meeus-based online
calculator for the solar position at http://www.esrl.noaa.gov/gmd/grad/solcalc/ that includes links to
spreadsheet versions of the solar position calculator for a day or year via the “Calculation Details”
link on that page.
An algorithm that provided higher accuracy than the Blanco-Muriel et al. (2001) algorithm and
lower accuracy than Reda and Andreas (2004) but with a shorter run time and less complexity was
provided in Grena (2008). The maximum error for this algorithm is stated to be 0.0027 degrees with
a run time comparable to somewhat less accurate algorithms (Blanco-Muriel et al. 2001; Michalsky
1988a). This accuracy is for the period 2003 to 2022. The code is included in Grena (2008) as an
appendix.
Table E.1 contains a summary of the available codes with maximum errors and years for which
the codes are valid. The maximum error is from the information in Grena (2008).
More complex codes run more slowly but deliver greater accuracy, sometimes with a limited
time window over which the accuracy is guaranteed. However, run time with today’s processors is
hardly ever an issue.
One point that does not seem to be emphasized in any of these papers is the issue of accuracy
given the generally unknown true refraction. Refraction is usually calculated for standard condi-
tions but depends on the true temperature and moisture structure of the atmosphere in the direc-
tion of the Sun and on atmospheric pressure. Van der Werf (2003) demonstrates arc minutes of
differences for a modified standard atmosphere with different temperature and humidity profiles.
These changes are exacerbated at the lowest sun angles, especially from the horizon to 5 degrees
above the horizon.
Since the publication of the first edition of this book several papers have been published
expanding on the solar position calculation. Two of these are suggested for those interested in
investigating this topic further. Grena (2012) introduced five new sun position algorithms and
Armstrong and Izygon (2014) produced software that not  only allows the solar position to be
calculated but compares several sun position algorithms with regards to their accuracy.

TABLE E.1
Summary of Solar Position Accuracies and Years of Applicability
Source Maximum Error Years Included
Spencer (1971) ~0.25 degrees —
Michalsky (1988a) 0.011 degrees 1950–2050
Blanco-Muriel et al. (2001) 0.008 degrees 1999–2015
Reda and Andreas (2004) 0.0003 degrees 2000 BC–AD 6000
Grena (2008) 0.0027 degrees 2003–2022
Appendix E 387

E.1 ASUNPOS.F
This Fortran program calculates the solar position given the year, day, time, latitude, and longitude.
It outputs solar azimuth, elevation, hour angle, declination, and air mass. The value output by this
program includes the effect of refraction calculated using standard conditions.

c   this program calculates the solar position given the year, day,
c   time, latitude, and longitude
c
c   it outputs solar azimuth, elevation, hour angle, declination, and
c   air mass
c
c
      program sunpos
      real lat,long, mn
      write(6,'(3x,"This program calculates sun position if you
     # specify time and location!")')
      write(6,'(/)')
      write(6,'(3x,"Type the latitude like this sxx.xx where s is a
     # sign!")')
      read(5,'(f8.4)')lat
      write(6,'(f8.4)')lat
      write(6,'(3x,"Type the longitude like this sxxx.xx!")')
      read(5,'(f9.4)')long
      write(6,'(f9.4)')long
      write(6,'(3x,"Type the year like this xxxx.!")')
      read(5,'(f5.0)')year
      write(6,'(f5.0)')year
      write(6,'(3x,"Indicate time zone, e.g., est=+5,mst=+7,etc.(The
     # sign is important)!")')
      read(5,'(i3)')itz
      write(6,'(i3)')itz
      tz=float(itz)
60    write(6,'(3x,"Type the day of year(Feb 1=32) like this xxx.!")')
      read(5,'(f4.0)')day
      write(6,'(f4.0)')day
50    write(6,'(3x,"Type the local standard time(not daylight savings
     # time) like this xx.xx.xx., e.g., 12.15.47.!")')
      read(5,'(3f3.0)')hour, mn,sec
      write(6,'(3f3.0)')hour, mn,sec
      tm=((tz+hour)*3600+ mn*60+sec)/3600.
      call sunae(year,day,tm,lat,long,az,el,ha,dec,soldst)
      write(6,'(//)')
      write(6,'("The solar azimuth and elevation are")')
      write(6,100)az,el
100   format(3x,f9.4,3x,f8.4,/)
      write(6,'("The solar hour angle and declination are")')
      write(6,100)ha,dec
      am=armass(el)
      write(6,'("The air mass is")')
      write(6,200)am
200   format(3x,f6.2)
      write(6,'("The solar distance is")')
      write(6,250)soldst
250   format(3x,f7.4)
      write(6,'(//)')
388 Appendix E

      write(6,'("If you want another time same day, type 9!")')
      read(5,'(i1)')i
      if(i.eq.9)go to 50
      write(6,'("If you want another day, type 99!")')
      read(5,'(i2)')i
      if(i.eq.99)go to 60
      end
c---------------------------------------------------------------------+
c                                                                     |
c   subroutine sunae(year,day,hour,lat,long,az,el,ha,dec,soldst)      |
c                                                                     |
c   this subroutine calculates the local azimuth and elevation of the |
c     sun at a specific location and time using an approximation to   |
c     equations used to generate tables in The Astronomical Almanac.  |
c     refraction correction is added so sun position is apparent one. |
c                                                                     |
c   The Astronomical Almanac, U.S. Gov't Printing Office, Washington, |
c     D.C. (1985).                                                    |
c                                                                     |
c   input parameters                                                  |
c     year=year, e.g., 1986                                           |
c     day=day of year, e.g., feb 1=32                                 |
c     hour=hours plus fraction in UT, e.g., 8:30 am eastern daylight  |
c       time is equal to 8.5 + 5(5 hours west of Greenwich) -1(for    |
c       daylight savings time correction)                             |
c     lat=latitude in degrees (north is positive)                     |
c     long=longitude in degrees (east is positive)                    |
c                                                                     |
c   output parameters                                                 |
c     a=sun azimuth angle (measured east from north, 0 to 360 degs)   |
c     e=sun elevation angle (degs)                                    |
c     plus others, but note the units indicated before return         |
c                                                                     |
c---------------------------------------------------------------------+
c
      subroutine sunae(year,day,hour,lat,long,az,el,ha,dec,soldst)
c   work with real variables and define some constants, including
c   one to change between degs and radians
      implicit real (a-z)
      data twopi,pi,rad/6.2831853,3.1415927,.017453293/
c
c   get the current julian date (actually add 2,400,000 for jd)
      delta=year-1949.
      leap=aint(delta/4.)
      jd=32916.5+delta*365.+leap+day+hour/24.
c   1st no. is mid. 0 jan 1949 minus 2.4e6; leap=leap days since 1949
c   the last yr of century is not leap yr unless divisible by 400
      if(amod(year,100.).eq.0.0.and.amod(year,400.).ne.0.0)jd=jd-1.
c
c   calculate ecliptic coordinates
      time=jd-51545.0
c   51545.0 + 2.4e6 = noon 1 jan 2000
c
c   force mean longitude between 0 and 360 degs
      mnlong=280.460+.9856474*time
Appendix E 389

      mnlong=mod(mnlong,360.)
      if(mnlong.lt.0.)mnlong=mnlong+360.
c
c   mean anomaly in radians between 0 and 2*pi
      mnanom=357.528+.9856003*time
      mnanom=mod(mnanom,360.)
      if(mnanom.lt.0.)mnanom=mnanom+360.
      mnanom=mnanom*rad
c
c   compute the ecliptic longitude and obliquity of ecliptic in radians
      eclong=mnlong+1.915*sin(mnanom)+.020*sin(2.*mnanom)
      eclong=mod(eclong,360.)
      if(eclong.lt.0.)eclong=eclong+360.
      oblqec=23.439-.0000004*time
      eclong=eclong*rad
      oblqec=oblqec*rad
c
c   calculate right ascension and declination
      num=cos(oblqec)*sin(eclong)
      den=cos(eclong)
      ra=atan(num/den)
c   force ra between 0 and 2*pi
      if(den.lt.0)then
          ra=ra+pi
      elseif(num.lt.0)then
          ra=ra+twopi
      endif
c
c   dec in radians
      dec=asin(sin(oblqec)*sin(eclong))
c
c   calculate Greenwich mean sidereal time in hours
      gmst=6.697375+.0657098242*time+hour 
c   hour not changed to sidereal time since 'time' includes
c   the fractional day 
      gmst=mod(gmst,24.)
      if(gmst.lt.0.)gmst=gmst+24.
c
c   calculate local mean sidereal time in radians 
      lmst=gmst+long/15.
      lmst=mod(lmst,24.)
      if(lmst.lt.0.)lmst=lmst+24.
      lmst=lmst*15.*rad
c
c   calculate hour angle in radians between -pi and pi
      ha=lmst-ra
      if(ha.lt.-pi)ha=ha+twopi
      if(ha.gt.pi)ha=ha-twopi
c
c   change latitude to radians
      lat=lat*rad
c
c   calculate azimuth and elevation
      el=asin(sin(dec)*sin(lat)+cos(dec)*cos(lat)*cos(ha))
      az=asin(-cos(dec)*sin(ha)/cos(el))
390 Appendix E

c
c   this puts azimuth between 0 and 2*pi radians
      if(sin(dec)-sin(el)*sin(lat).ge.0.)then
      if(sin(az).lt.0.)az=az+twopi
      else
      az=pi-az
      endif
c
c   calculate refraction correction for US stan. atmosphere
c   need to have el in degs before calculating correction
      el=el/rad
c
      if(el.ge.19.225) then 
         refrac=.00452*3.51823/tan(el*rad)
      else if (el.gt.-.766.and.el.lt.19.225) then
         refrac=3.51823*(.1594+.0196*el+.00002*el**2)/
     1   (1.+.505*el+.0845*el**2)
      else if (el.le.-.766) then
         refrac=0.0
      end if
c
c   note that 3.51823=1013.25 mb/288 C
      el=el+refrac
c   elevation in degs
c
c   calculate distance to sun in A.U. & diameter in degs
      soldst=1.00014-.01671*cos(mnanom)-.00014*cos(2.*mnanom)
      soldia=.5332/soldst
c
c   convert az and lat to degs before returning
      az=az/rad
      lat=lat/rad
      ha=ha/rad
      dec=dec/rad
c   mnlong in degs, gmst in hours, jd in days if 2.4e6 added;
c   mnanom,eclong,oblqec,ra,and lmst in radians
      return
      end
c
c
c   this function calculates air mass using kasten's
c   approximation to bemporad's tables
c
c
      function armass(el)
      z=(90.-el)*3.141592654/180.
      armass=1./(cos(z)+.50572*(6.07995+el)**-1.6364)
      return
      end
Appendix E 391

REFERENCES
Archer, C. B. 1980. Comments on “Calculating the position of the sun.” Solar Energy 25:91.
Armstrong, P., and M. Izygon. 2014. An innovative software for analysis of sun position algorithms. Energy
Procedia 49:2444–2453.
Blanco-Muriel, M., D. C. Alarcón-Padilla, T. Lópex-Moratalla, and M. Lara-Coira. 2001. Computing the solar
vector. Solar Energy 70:431–441.
Grena, R. 2008. An algorithm for the computation of the solar position. Solar Energy 82:462–470.
Grena, R. 2012. Five new algorithms for the computation of sun position from 2010 to 2110. Solar Energy
86:1323–1337. doi:10.1016/j.solener.2012.01.024.
Meeus, J. 1998. Astronomical Algorithms, 2nd ed. Richmond, VA: Willmann-Bell, Inc.
Michalsky, J. J. 1988a. The Astronomical Almanac’s algorithm for approximate solar position (1950–2050).
Solar Energy 40:227–235.
Michalsky, J. J. 1988b. Errata. Solar Energy 41:113.
Michalsky, J. J. 1989. Errata. Solar Energy 43:323.
Nautical Almanac Office of the United State Naval Observatory and Nautical Almanac Office of Great Britain.
1984. The Astronomical Almanac for the Year 1985. Washington, DC: U.S. Government Printing Office.
Nautical Almanac Office of the United State Naval Observatory and Nautical Almanac Office of Great Britain.
2018. The Astronomical Almanac for the Year 2019. Washington, DC: U.S. Government Printing Office.
Pitman, C. L., and L. L. Vant-Hull. 1978. Errors in locating the sun and their effect on solar intensity predictions.
Proceedings of the American Section of the International Solar Energy Society, Denver, CO.
Reda, I., and A. Andreas. 2004. Solar position algorithm for solar radiation applications. Solar Energy
76:577–589.
Spencer, J. W. 1971. Fourier series representation of the position of the sun. Search 2:172.
Spencer, J. W. 1989. Comments on “The  astronomical almanac’s algorithm for approximate solar position
(1950–2050).” Solar Energy 42:353.
Van der Werf, S. 2003. Ray tracing and refraction in the modified US1976 atmosphere. Applied Optics
42:354–366.
Vant-Hull, L. L., and A. F. Hildebrandt. 1976. Solar thermal power system based on optical transmission.
Solar Energy 18:31–39.
Walraven, R. 1978. Calculating the position of the sun. Solar Energy 20:393–397.
Walraven. R. 1979. Erratum. Solar Energy 22:195.
Wilkinson, B. J. 1981. An improved FORTRAN program for the rapid calculation of the solar position. Solar
Energy 27:67–68.
Appendix F: Useful
Conversion Factors

Key terms and units of measure important to radiometry are presented in Table  F.1. Converting
units of measure for other selected quantities can be accomplished with the information presented
in Table F.2. The multipliers needed to convert from the unit in the first column to the unit in the
second column are given in the third column of Table F.2.

TABLE F.1
Radiometric Terminology and Units of Measurement
Quantity SI Unit Abbreviation Description
Radiant Energy Joule J Energy (ability to do
work)
Radiant Flux Watt W Radiant energy per unit
time (radiant power)
Radiant Intensity Watt per steradian Wsr−1 Power per unit solid
angle
Radiant Power Density Watt per square meter Wm−2 Radiant power per unit
area
Radiant Energy Watt-hour per square Whm−2 Radiant power in 1 hour
Density meter (energy) per unit area
Radiant Emittance Watt per square meter Wm−2 Power emitted from a
surface
Radiance Watt per steradian per Wsr−1m−2 Power per unit solid
square meter angle per unit of
projected source area
Spectral Irradiance Watt per square meter Wm−2nm−1 Power incident on a
per nanometer surface per unit
wavelength

TABLE F.2
Useful Conversion Factors
To Convert From To Multiply By
Watt Joule (J)/sec 1.0000
Watt Btu/hour 3.4094
Watt Horsepower 1.3410 × 10−3
Joules/sec Watt 1.0000
Btu/hour Watt 0.2933
Horsepower Watt 745.71
Kilowatt-hour (kWh) Btu 3409.4
Kilowatt-hours/meter2 (kWhm−2) kilojoule/meter2 (kJm−2) 3600.0
Kilowatt-hours/meter2 (kWhm−2) Btu/foot2 316.74
(Continued)

393
394 Appendix F

TABLE F.2 (Continued)


Useful Conversion Factors
To Convert From To Multiply By
−2
Kilowatt-hours/meter (kWhm )
2 Langley (Ly) 86.04
Kilowatt-hours/meter2 (kWhm−2) Calories/centimeter2 86.04
Btu kilowatt-hour (kWh) 2.9331 × 10−4
Kilojoules/meter2 kilowatt-hour/meter2 (kWhm−2) 2.7777 × 10−4
Btu/foot2 kilowatt-hour/meter2 (kWhm−2) 3.1571 × 10−3
Langley (Ly) kilojoule/meter2 (kJm−2) 41.84
Langley (Ly) kilowatt-hour/meter2 (kWhm−2) 0.011622
Calories/centimeter2 kilowatt-hour/meter2 (kWhm−2) 0.011622
Joules (J) Btu 0.94706 × 10−3
Kilojoules/meter2 (kJm−2) kilowatt-hour meter−2 (KWhm−2) 0.27778 × 10−3
Kilojoules/meter2 (kJm−2) Langley (Ly) 0.023901
Kilojoules/meter2 (kJm−2) Btu foot−2 0.087982
Btu foot−2 Langley (Ly) 0.27144
Btu foot−2 kilojoules/meter2 (kJm−2) 11.366
Btu/hour-foot2 Watt meter−2 3.1572
Foot-candles Lux 10.76
Lux Foot-candles 0.09294
Watts/meter2-C Btu/hour-foot2-F 0.17611
Meters Inches 39.37
Meters Feet 3.2808
Meters Yards 1.094
Kilometer Mile 0.6213712
Meter2 Foot2 10.764
Kilometer2 Acre 247.10
Liter (L) Foot3 0.035316
Meter3 Gallon (U.S.) 264.172
Gallon Foot3 0.13368
Liter Gallon 0.26418
Kilogram Pound (Lb) 2.2046
Kilogram-meter−2 Pounds/square inch (Lb-in−2) 0.001422
Celsius Fahrenheit (°F) C°*1.8 + 32
Celsius Kelvin C° + 273.15
Fahrenheit (°F) Celsius (°F − 32)/1.8
Kelvin Celsius K − 273.15
Meters/second Miles/hr (mph) 2.2369
Meters/second Kilometer/hr 3.6
Meters/second Knots 1.9438
Millibars Pascals 100.0
Millibars Atmospheres 0.0009869
Millibars Pounds per square inch (Lb-in−2) 0.014504
Degrees Radians 0.017453293
Radians Degrees 57.2957795
Appendix F 395

F.1 THE MANY MEANINGS OF WATTS PER SQUARE METER


Solar irradiance, the amount of solar radiation received by a surface (i.e., radiant energy density)
for the minute, hour, day, month, or year, is often reported in units of average watts per square
meter (Wm−2) or averaged radiant power per unit area for a given time period. The  energy per
unit area per unit time (Wm−2) multiplied by the time interval becomes the radiant energy density
because energy is power multiplied by time. For example, one Colorado town received an average of
5.6 kilowatt-hours per square meter (kWhm−2) of solar radiation per day. This value is also reported
as an average of 233 Wm−2 over the day.

W ⋅ day W ⋅ 24 ⋅ hr W ⋅ hr kW ⋅ hr
233 2
= 233 2
= 5595 2
= 5.6 (F.1)
m m m m2

Another example of quantifying energy and power is to assume the measured solar irradiance (power
density) was 1,000 Wm−2 over a 5-minute period. The equivalent energy density is 83 Whm−2.

1 hr
W ⋅ 5 ⋅ min
60 min W ⋅ hr
1000 = 83 (F.2)
m2 m2

The general public is interested in kilowatt-hours (kWh) that can be produced by a photovoltaic


(PV) system. Consider that the average U.S. household consumes between 900 and 1,000 kWh of
electricity each month. In the previous example, a town in Colorado received an annual average of
5.6  kWhm−2 of solar energy per day on a surface that is tilted to the south by an angle equal to the
site latitude (the angle made with the horizontal is 40° for a site at 40°). In an average 30-day month
each square meter of a tilted surface will receive about 168 kWh of energy; therefore, 6 m2 would
be needed to intercept a little over 1,000 kWh of energy. However, if the PV panels receiving this
energy are about 16 percent efficient at converting sunlight into electricity, then 36 m2 of PV panels
are required to meet the average monthly electrical load of the house. This is roughly a square area
that is 6 m (~20 feet) on a side.

QUESTIONS
1. Given that an average 5.6  kWhm−2 is incident on a surface each day and solar panels
are 16  percent efficient, what is the area of PV panels needed to supply an average of
1,200 kWh per month to a house? What is the answer in meters? Feet?
2. What is the difference between energy and power?
Appendix G: Sources for Equipment

A list of manufacturers, distributors, and system integrators of equipment used for measuring solar
and infrared (IR) radiation and supporting meteorological elements is provided here. This list is
not complete, and inclusion in the list does not constitute an endorsement, expressed or implied, by
the authors or publisher. A thorough review of the literature should be conducted before purchasing
any scientific equipment.
Developing a measurement capability for a project of any scale requires thoughtful consideration
of these basic design criteria:

• What measurements are needed for the intended application(s)?


• What are the acceptable limits for estimated data uncertainty?
• Are the instrument performance specifications provided by the manufacturer consistent
with the acceptable estimated data uncertainties?
• What installation, operation and maintenance (including calibration) practices are needed
to produce the desired measurements?
• How will the measured data be acquired, reviewed for quality, analyzed, distributed, and
archived?

Note that the links below were checked in June 2019.

Apogee Instruments, Inc.


https://www.apogeeinstruments.com
Spectroradiometers, Radiometers
Campbell Scientific, Inc.
https://www.campbellsci.com/
Radiometers, Photometers, Data Loggers, Weather Stations, Data Management Software
Cimel Electronique
https://www.cimel.fr/?lang=en
Sun photometer, Weather Stations
CSP Services GmbH
http://www.cspservices.de/
Rotating Shadowband Irradiometers, Weather Stations, Data Systems, Sky Imagers, Data
management software, PV and CSP soiling rate measurement
Davis Instruments, Corp.
https://www.davisnet.com/
Weather Stations
Delta-T Devices Ltd
https://www.delta-t.co.uk/
U.S. Distributor:
Dynamax, Inc.
www.dynamax.com
Radiometers, Weather Stations & Data Loggers

397
398 Appendix G

EKO Instruments Co., LTD.


http://www.eko.co.jp/
European Office:
https://eko-eu.com/
USA Office:
https://eko-usa.com
Radiometers, Photometers, Spectroradiometers, Solar Trackers, Data Loggers
The Eppley Laboratory, Inc.
http://www.eppleylab.com/
Radiometers, Photometers, Solar Trackers, Data Loggers, Calibration Services
Geónica S.A.
http://www.geonica.com/en/
Solar Trackers, Radiometers, Weather Stations, Data Loggers
Gigahertz-OPTIK, INC.
https://www.gigahertz-optik.de/en-us/
Radiometers, Spectroradiometers, Photometers, Calibration Standards
GroundWork
http://grndwork.com
System integrator for solar and meteorological measurements, Monitoring station O&M, and
Data Management
Hukseflux Thermal Sensors B.V.
http://www.hukseflux.com/
Radiometers, Photometers and Calibration Services
Irradiance, Inc.
http://www.irradiance.com/
Rotating Shadowband Radiometer (RSR2)
Kipp & Zonen, B.V.
http://www.kippzonen.com
Radiometers, Photometers, Solar Trackers, PV Soiling Detection and Data Loggers
LI-COR, Inc.
https://www.licor.com/env/products/light/
Radiometers, Photometers, Data Loggers, Weather Stations
Malvern Panalytical
https://www.malvernpanalytical.com/en/products/product-range/asd-range Spectroradiometers
Middleton Solar
http://www.middletonsolar.com/
Radiometers, Photometers, Spectroradiometers, Solar Trackers
Ocean Optics, Inc.
https://www.oceanoptics.com
Spectroradiometers, Light Sources
Onset Computer Corporation
http://www.onsetcomp.com
HOBO® Logger-based Weather Stations
Appendix G 399

OWEL GmbH
http://www.owel-swiss.ch
Solar Tracker (Brusag INTRA)
Pace Scientific Inc.
http://www.pace-sci.com
Data Loggers, Radiometers, Meteorological Instruments
Philipp Schenk GmbH Wein
http://www.schenk.co.at
Radiometers
PMOD/WRC
https://www.pmodwrc.ch/
Absolute Cavity Radiometer, Spectroradiometer, Sky Imager
Prede Co., Ltd.
http://www.prede.com
Radiometers, Photometers, Solar Trackers, Shadow Ring
RainWise, Inc.
http://rainwise.com
Pyranometers, Rain Gauges, Wind Sensors, Data Loggers, Weather Stations
Schreder CMS
http://www.schreder-cms.com/en
Radiometers, Spectroradiometers, Sky Imagers, Calibration Services
SieltecCanarias S.L.
http://www.sieltec.es/
Radiometers, Sky Imagers
Skye Instruments Ltd
https://www.skyeinstruments.com
Radiometers, Photometers, Automatic Weather Station
Solar Light Company, Inc.
https://solarlight.com/
Radiometers, Photometers
Spectrafy
https://spectrafy.com/
Photometers and Spectral Modeling
Suntrace GmbH
http://suntrace.de/
System integrator for solar resource measurements
Trios Optical Sensors
http://www.trios.de
Spectroradiometers
Yankee Environmental Systems, Inc.
http://www.yesinc.com/
Radiometers, Spectroradiometers, Data Loggers, Sky Imagers
Appendix H: BORCAL Report
The first part of Appendix H provides a sample broadband outdoor radiation calibration (BORCAL)
report for shortwave (solar) instruments calibrated at the National Renewable Energy Laboratory
(NREL) Solar Radiation Research Laboratory (SRRL). The BORCAL process for shortwave instru-
ments is based on the summation method for calibration of pyranometers and direct comparison
of pyrheliometers with electrically self-calibrating absolute cavity radiometers traceable to the
World Radiometric Reference (see Chapter 4). The second part of this appendix provides a sample
BORCAL report for infrared (longwave) radiometers whose calibrations are traceable to the World
Infrared Standard Group (WISG; Chapter 11). BORCAL reports for many instruments can be found
at https://aim.nrel.gov/borcal.html. SRRL calibrations are accredited by the International Standards
Organization (ISO) 17025.

401
402 Appendix H

Broadband Outdoor Radiometer Calibration Shortwave

BORCAL –SW
2018 – 03

Generated by

Radiometer Calibration and Characterization

Customer
NREL – SRRL – BMS

Organization: NREL
Address: BMS, SRRL, Golden, CO 80401 USA
Phone: 303-384-6326

Calibration Facility
Solar Radiation Research Laboratory

Latitude: 39.742°N
Longitude: 105.180°W
Elevation: 1828.8
meters AMSL Time
Zone: 7.0

Calibration date
05/15/2018 to 05/16/2018

Report Date
May 23, 2018

NOTICE
This report was prepared as an account of work sponsored by an agency of the United States gov-
ernment. Neither the United States government nor any agency thereof, nor any of their employees,
makes any warranty, express or implied, or assumes any legal liability or responsibility for the
accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed,
or represents that its use would not infringe privately owned rights. Reference herein to any specific
commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does
not  necessarily constitute or imply its endorsement, recommendation, or favoring by the United
States government or any agency thereof. The views and opinions of authors expressed herein do
not necessarily state or reflect those of the United States government or any agency thereof.
Appendix H 403

Broadband Outdoor Radiometer Calibration Report

Introduction
This  report compiles the calibration results from a Broadband Outdoor Radiometer Calibration
(BORCAL). The work was accomplished at the Radiometer Calibration Facility shown on the front
of this report. The calibration results reported here are traceable to the International System (SI)
Units of Measurement.
This report includes these sections:

• Results Summary: A  table of all instruments included in this report summarizing their
calibration results and uncertainty.
• Control Instruments: A group of instruments included in each BORCAL event that pro-
vides a measure of process consistency.
• Instrument Details: The calibration certificates for each instrument.
• Environmental and Sky Conditions: Meteorological conditions and reference irradiance
during the calibration event.

Results Summary

TABLE 1
Results Summary
R@451 CF@451 U2 Rnet
Instrument (μV/W/m2) (W/m2/mV) (μV/W/m2) Page
140043 Kipp & Zonen CMP22 9.0413 110.60 1.0/–1.95 0.087 A1‐26
140108 Kipp & Zonen CHP1 8.0489 124.24 0.86/–1.9 0 A1‐29
21096 Eppley 8–48 11.292 88.561 2.7/–1.4 0 A1‐41

Note: Environmental Conditions for BORCAL starts on Pages A1‐104.


1 CF = 1000/R

2 See certificate for valid zenith angle range

3 Instrument’s Effective Net IR Response


404 Appendix H

Control Instrument History

FIGURE 1 Eppley NIP control instrument history.

FIGURE 2 Eppley PSP control instrument history.

Appendix 1
Instrument Details
Calibration Certificates: 3 pages for each radiometer (4 including Environmental Conditions)
Environmental Conditions for BORCAL: Last Page of a Calibration Certificate. Note:
This appears only once, at the end of Appendix 1.
Appendix H 405

Test Instrument: Pyranometer Manufacturer: Kipp & Zonen


Model: CMP22 Serial Number: 140043
Calibration Date: 5/16/2018 Due Date: 5/16/2019
Customer: NREL-SRRL-BMS Environmental Conditions: see page 4
Test Dates: 5/15-16

This  certifies that the above product was calibrated in compliance with ISO/IEC 17025:2005.
Measurement uncertainties at the time of calibration are consistent with the Guide to the Expression
of Uncertainty in Measurement (GUM) using Reda et al. (2008). All nominal values are traceable
to the International System (SI) Units of Measurement.
No statement of compliance with specifications is made or implied on this certificate; however,
the estimated uncertainties are the uncertainties of the calibration process. Users must add other
uncertainties that are relevant to their measuring system, environmental and sky conditions, out-
door set-up, and site location.
The Type-B Standard Uncertainty of using the responsivity at each even zenith angle is reported,
and the Expanded Uncertainty of the calibration is reported using two methods:

1. The Expanded Uncertainty of using the responsivity at zenith angle = 45°, within the zenith
angle range from 30 to 60 degrees.
2. The  Expanded Uncertainty of using Spline Interpolating Functions for the responsivity
versus zenith angle.

This certificate applies only to the item identified above and shall not be reproduced other that in full,
without specific written approval from the calibration facility. Certificate without signature is not valid.

TABLE 1
Traceability
Calibration
Measurement Type Instrument Calibration Date Due Date
Beam Irradiance† Eppley Absolute Cavity Radiometer Model HF, S/N 29219 09/25/2017 09/25/2018
Diffuse Irradiance† Hukseflux Pyranometer Model SR25, S/N 2541 04/11/2018 04/11/2019
Diffuse Irradiance† Hukseflux Pyranometer Model SR25, S/N 2542 04/18/2018 04/18/2019
Data Acquisition NREL Data Acquisition System Model RAP-DAQ, S/N 04/12/2017 04/12/2019
2005-998
Data Acquisition NREL Data Acquisition System Model RAP-DAQ, S/N 04/12/2017 04/12/2019
2005-999
Infrared Irradiance‡ Kipp & Zonen Pyrgeometer Model CG4, S/N FT002 04/16/2018 04/16/2022
†Through the World Radiometric Reference (WRR)
‡Through the World Infrared Standard Group (WISG)
406 Appendix H

Number of pages of certificate: 4

Calibration Procedure: BORCAL-P00-Calibration and QA Procedure; available upon request.

Setup: Radiometers are calibrated outdoors, using the sun as the source. Pyranometers and
pyrgeometers are installed for horizontal measurements, with their signal connectors ori-
ented north, if their design permits.
The shading disk for the reference diffuse subtends a solid angle of 5°. Pyrheliometers are
installed on solar trackers.

Calibrated by: RCC


______________________________ _________
Ibrahim Reda, Technical Manager Date

For questions or comments, please contact the technical manager at:


ibrahim.reda@nrel.gov; 303-384-6385; 15013 Denver West Parkway, Golden, CO 80401, USA
Appendix H 407

Calibration Results

140043 Kipp & Zonen CMP22


The  responsivity (R, µV/W/m²) of the test instrument during calibration is calculated using this
Measurement Equation:

R = (V − Rnet * Wnet) / I [1]

where
V = radiometer output voltage (microvolts)
I = reference irradiance (W/m²), beam (B) or global (G)
Rnet = radiometer net infrared responsivity (µV/W/m²), see Table 4
Wnet = effective net infrared measured by pyrgeometer (W/m²)

= Win − Wout

= Win − σ * Tc^4

where
Win = incoming infrared (W/m²)
σ = 5.6704e-8 W·m−2·K−4
Tc = case temperature of pyrgeometer (K)

G = B * COS(Z) + D,

where
Z = zenith angle (degrees),
D = reference diffuse irradiance (W/m²).

FIGURE 1 Responsivity vs Zenith Angle. FIGURE 2 Responsivity vs Local Standard Time.


408

TABLE 2
Instrument Responsivity (R) and Calibration Type-B Standard Uncertainty, u(B)
Zenith AM PM Zenith AM PM
Angle R u(B) Azimuth R u(B) Azimuth Angle R u(B) Azimuth R u(B) Azimuth
(deg.) (µV/W/m²) ± (%) Angle (µV/W/m²) ± (%) Angle (deg.) (µV/W/m²) ± (%) Angle (µV/W/m²) ± (%) Angle
0 N/A N/A N/A N/A N/A N/A 46 9.0310 0.29 102.04 9.0462 0.29 258.06
2 N/A N/A N/A N/A N/A N/A 48 9.0186 0.29 99.94 9.0525 0.29 260.12
4 N/A N/A N/A N/A N/A N/A 50 9.0173 0.29 98.11 9.0363 0.30 262.08
6 N/A N/A N/A N/A N/A N/A 52 9.0325 0.30 96.15 9.0782 0.30 263.96
8 N/A N/A N/A N/A N/A N/A 54 9.0247 0.30 94.43 9.0271 0.30 265.77
10 N/A N/A N/A N/A N/A N/A 56 9.0279 0.30 92.66 9.0377 0.31 267.53
12 N/A N/A N/A N/A N/A N/A 58 9.0292 0.31 90.84 9.0128 0.32 269.17
14 N/A N/A N/A N/A N/A N/A 60 9.0194 0.31 89.23 N/A N/A N/A
16 N/A N/A N/A N/A N/A N/A 62 9.0453 0.31 87.53 9.0352 0.33 272.25
18 N/A N/A N/A N/A N/A N/A 64 9.0374 0.32 85.94 9.0334 0.34 273.80
20 N/A N/A N/A N/A N/A N/A 66 9.0436 0.33 84.32 9.0499 0.35 275.43
22 9.0334 0.28 155.95 9.0323 0.28 203.93 68 9.0045 0.34 82.88 9.0410 0.37 277.04
24 9.0376 0.28 144.67 9.0358 0.28 215.53 70 8.9971 0.35 81.23 9.0442 N/A 278.55
26 9.0315 0.28 136.66 9.0400 0.28 223.33 72 9.0005 0.37 79.74 9.0752 N/A 280.15
28 9.0556 0.28 130.67 9.0409 0.28 229.48 74 8.9860 0.40 78.12 9.0648 N/A 281.76
30 9.0461 0.28 125.77 9.0424 0.28 234.21 76 8.9878 0.43 76.54 9.1151 N/A 283.35
32 9.0466 0.28 121.62 9.0343 0.29 238.59 78 8.9414 N/A 74.90 9.0730 N/A 284.94
34 9.0492 0.28 117.99 9.0211 0.29 242.26 80 8.9913 N/A 73.30 9.0127 N/A 286.58
36 9.0607 0.29 114.58 9.0253 0.29 245.28 82 9.0496 N/A 71.62 N/A N/A N/A
38 9.0402 0.29 111.69 9.0207 0.29 248.41 84 9.2084 N/A 69.95 N/A N/A N/A
40 9.0455 0.29 109.04 9.0325 0.29 251.03 86 N/A N/A N/A N/A N/A N/A
42 9.0490 0.29 106.62 9.0700 0.29 253.49 88 N/A N/A N/A N/A N/A N/A
44 9.0527 0.29 104.20 9.0426 0.29 255.90 90 N/A N/A N/A N/A N/A N/A

Abbreviation: N/A, Not Available.


Appendix H
Appendix H 409

FIGURE 3 Type-B standard uncertainty vs zenith FIGURE 4 Residuals from spline interpolation.
angle.

TABLE 3 TABLE 4
Uncertainty Using Spline Interpolation Calibration Label Values
Type-B standard uncertainty, u(B) (%) ±0.43 R @ 45° (µV/W/m²) Rnet (µV/W/m²)†
Type-A interpolating function, u(int) (%) ±0.21
9.0413 0.087000
Combined standard uncertainty, u(c) (%) ±0.48
Effective degrees of freedom, DF(c) 21690
†Rnet determination date: Estimated
Coverage factor, k 1.96
Expanded uncertainty, U95 (%) ±0.93
AM valid zenith angle range 22°–76°
PM valid zenith angle range 22°–68° TABLE 5
Uncertainty Using R @ 45
Type-B expanded uncertainty, U(B) (%) ±0.63
Offset uncertainty, U(off) (%) +0.41 / −0.32
Expanded uncertainty, U (%) +1.0 / −0.95
Effective degrees of freedom, DF +Inf
Coverage factor, k 1.96
Valid zenith angle range 30.0°–60.0°

FIGURE 5 History of instrument at zenith angle = 45°.


410 Appendix H

BIBLIOGRAPHY
Reda, I.; Hickey, J.; Long, C.; Myers, D.; Stoffel, T.; Wilcox, S.; Michalsky, J. J.; Dutton, E. G.; Nelson, D.
(2005). “Using a Blackbody to Calculate Net Longwave Responsivity of Shortwave Solar Pyranometers
to Correct for Their Thermal Offset Error During Outdoor Calibration Using the Component Sum
Method.” Journal of Atmospheric and Oceanic Technology, 2005; pp.  1531–1540; NREL Report
No. JA-560-36646. doi:10.1175/JTECH1782.1.
Reda, I.; Myers, D.; Stoffel, T. (2008). “Uncertainty Estimate for the Outdoor Calibration of Solar Pyranometers:
A Metrologist Perspective.” Measure. (NCSLI Journal of Measurement Science). Vol. 3(4), December
2008; pp. 58–66; NREL Report No. JA-581-4137.
Reda, I.; Andreas, A. (2004). “Solar Position Algorithm for Solar Radiation Applications.” Solar Energy.
Vol. 76(5), 2004; pp. 577–589; NREL Report No. JA-560-35518. doi:10.1016/j.solener.2003.12.003.
Stoffel, T.; Reda, I. (2009). “NREL Pyrheliometer Comparisons: 22 September–3 October 2008 (NPC-2008).”
54 pp.; NREL Report No. TP-550-45016.
Reda, I.; Stoffel, T.; Myers, D. (2003). ”Method to Calibrate a Solar Pyranometer for Measuring Reference
Diffuse Irradiance.” Solar Energy. Vol.  74, 2003; pp.  103–112; NREL Report No.  JA-560-35025.
doi:10.1016/S0038-092X(03)00124-5.
Reda, I. (1996). “Calibration of a Solar Absolute Cavity Radiometer with Traceability to the World Radiometric
Reference.” 79 pp.; NREL Report No. TP-463-20619.
Reda, I.; Gröbner, J.; Stoffel, T.; Myers, D.; Forgan, B. (2008). “Improvements in the Blackbody Calibration
of Pyrgeometers.” ARM 2008 Science Team Meeting (Poster).
Appendix H 411

Test Instrument: Pyrheliometer Manufacturer: Kipp & Zonen


Model: CHP1 Serial Number: 140108
Calibration Date: 5/16/2018 Due Date: 5/16/2019
Customer: NREL-SRRL-BMS Environmental Conditions: see page 4
Test Dates: 5/15-16

This  certifies that the above product was calibrated in compliance with ISO/IEC 17025:2005.
Measurement uncertainties at the time of calibration are consistent with the Guide to the Expression
of Uncertainty in Measurement (GUM) using Reda et al., 2008. All nominal values are traceable to
the International System (SI) Units of Measurement.
No statement of compliance with specifications is made or implied on this certificate. However,
the estimated uncertainties are the uncertainties of the calibration process; users must add other
uncertainties that are relevant to their measuring system, environmental and sky conditions, out-
door set-up, and site location.
The Type-B Standard Uncertainty of using the responsivity at each even zenith angle is reported,
and the Expanded Uncertainty of the calibration is reported using two methods:

1. The Expanded Uncertainty of using the responsivity at zenith angle = 45°, within the zenith
angle range from 30.0 to 60.0 degrees.
2. The  Expanded Uncertainty of using Spline Interpolating Functions for the responsivity
versus zenith angle.

This certificate applies only to the item identified above and shall not be reproduced other that in full,
without specific written approval from the calibration facility. Certificate without signature is not valid.

TABLE 1
Traceability
Measurement Type Instrument Calibration Date Calibration Due Date
Beam Irradiance† Eppley Absolute Cavity 09/25/2017 09/25/2018
Radiometer Model HF, S/N 29219
Diffuse Irradiance† Hukseflux Pyranometer Model 04/11/2018 04/11/2019
SR25, S/N 2541
Diffuse Irradiance† Hukseflux Pyranometer Model 04/18/2018 04/18/2019
SR25, S/N 2542
Data Acquisition NREL Data Acquisition System 04/12/2017 04/12/2019
Model RAP-DAQ, S/N 2005-998
Data Acquisition NREL Data Acquisition System 04/12/2017 04/12/2019
Model RAP-DAQ, S/N 2005-999
†Through the World Radiometric Reference (WRR)
412 Appendix H

Number of pages of certificate: 4

Calibration Procedure: BORCAL-P00-Calibration and QA Procedure; available upon request.

Setup: Radiometers are calibrated outdoors, using the sun as the source. Pyranometers and
pyrgeometers are installed for horizontal measurements, with their signal connectors ori-
ented north, if their design permits.
The shading disk for the reference diffuse subtends a solid angle of 5°. Pyrheliometers are
installed on solar trackers.

Calibrated by: RCC


______________________________ _________
Ibrahim Reda, Technical Manager Date

For questions or comments, please contact the technical manager at:


ibrahim.reda@nrel.gov; 303-384-6385; 15013 Denver West Parkway, Golden, CO 80401, USA
Appendix H 413

Calibration Results

140108 Kipp & Zonen CHP1


The  responsivity (R, µV/W/m²) of the test instrument during calibration is calculated using this
Measurement Equation:

R = (V − Rnet * Wnet) / I [1]

where
V = radiometer output voltage (microvolts)
I = reference irradiance (W/m²), beam (B) or global (G)
Rnet = radiometer net infrared responsivity (µV/W/m²), see Table 4
Wnet = effective net infrared measured by pyrgeometer (W/m²)

= Win − Wout

= Win − σ * Tc^4

where
Win = incoming infrared (W/m²)
σ = 5.6704e-8 W·m−2·K−4
Tc = case temperature of pyrgeometer (K)

G = B * COS(Z) + D,

where
Z = zenith angle (degrees),
D = reference diffuse irradiance (W/m²).

FIGURE 1 Responsivity vs Zenith Angle. FIGURE 2 Responsivity vs Local Standard Time.


414

TABLE 2
Instrument Responsivity (R) and Calibration Type-B Standard Uncertainty, u(B)
Zenith AM PM Zenith AM PM
Angle R u(B) Azimuth R u(B) Azimuth Angle R u(B) Azimuth R u(B) Azimuth
(deg.) (µV/W/m²) ± (%) Angle (µV/W/m²) ± (%) Angle (deg.) (µV/W/m²) ± (%) Angle (µV/W/m²) ± (%) Angle
0 N/A N/A N/A N/A N/A N/A 46 8.0534 0.29 102.08 8.0533 0.29 258.10
2 N/A N/A N/A N/A N/A N/A 48 8.0554 0.29 100.02 8.0529 0.29 260.14
4 N/A N/A N/A N/A N/A N/A 50 8.0650 0.29 98.13 8.0558 0.29 262.04
6 N/A N/A N/A N/A N/A N/A 52 8.0694 0.29 96.28 8.0546 0.29 263.92
8 N/A N/A N/A N/A N/A N/A 54 8.0605 0.29 94.53 8.0441 0.30 265.78
10 N/A N/A N/A N/A N/A N/A 56 8.0644 0.29 92.59 8.0523 0.30 267.57
12 N/A N/A N/A N/A N/A N/A 58 8.0563 0.29 90.81 N/A N/A N/A
14 N/A N/A N/A N/A N/A N/A 60 8.0715 0.30 89.21 7.9409 N/A 270.93
16 N/A N/A N/A N/A N/A N/A 62 8.0651 0.30 87.62 8.0664 0.30 272.18
18 N/A N/A N/A N/A N/A N/A 64 8.0715 0.30 85.97 8.0697 0.31 273.83
20 N/A N/A N/A N/A N/A N/A 66 8.0839 0.30 84.40 8.0677 0.31 275.46
22 8.0462 0.29 156.04 8.0388 0.29 203.74 68 8.0752 0.30 82.71 8.0659 0.31 276.93
24 8.0419 0.29 144.71 8.0418 0.29 215.49 70 8.0749 0.30 81.25 8.0697 N/A 278.58
26 8.0432 0.29 136.72 8.0456 0.29 223.29 72 8.0710 0.31 79.65 8.0661 N/A 280.18
28 8.0417 0.29 130.76 8.0407 0.29 229.27 74 8.0814 0.31 78.00 8.0710 N/A 281.74
30 8.0399 0.29 125.76 8.0437 0.29 234.28 76 8.0731 0.31 76.43 8.0757 N/A 283.33
32 8.0451 0.29 121.56 8.0435 0.29 238.56 78 8.0712 N/A 74.89 8.0762 N/A 284.97
34 8.0440 0.29 117.96 8.0452 0.29 242.25 80 8.0780 N/A 73.30 8.0722 N/A 286.57
36 8.0478 0.29 114.64 8.0436 0.29 245.37 82 8.0610 N/A 71.60 N/A N/A N/A
38 8.0552 0.29 111.73 8.0530 0.29 248.39 84 8.1081 N/A 69.94 N/A N/A N/A
40 8.0519 0.29 108.98 8.0499 0.29 251.16 86 N/A N/A N/A N/A N/A N/A
42 8.0496 0.29 106.46 8.0472 0.29 253.53 88 N/A N/A N/A N/A N/A N/A
44 8.0503 0.29 104.22 8.0521 0.29 255.89 90 N/A N/A N/A N/A N/A N/A
Abbreviation: N/A, Not Available.
Appendix H
Appendix H 415

FIGURE  3 Type-B Standard Uncertainty vs FIGURE 4 Residuals from spline interpolation.


Zenith Angle.

TABLE 3 TABLE 4
Uncertainty Using Spline Interpolation Calibration Label Values
Type-B standard uncertainty, u(B) (%) ±0.31 R @ 45° (µV/W/m²) Rnet (µV/W/m²)†
Type-A interpolating function, u(int) (%) ±0.15
8.0489 0
Combined standard uncertainty, u(c) (%) ±0.35
Effective degrees of freedom, DF(c) 22232 Rnet determination date: N/A

Coverage factor, k 1.96


Expanded uncertainty, U95 (%) ±0.68
AM valid zenith angle range 22°–76°
PM valid zenith angle range 22°–68° TABLE 5
Uncertainty Using R @ 45°
Type-B expanded uncertainty, U(B) (%) ±0.58
Offset uncertainty, U(off) (%) +0.28 / −1.3
Expanded uncertainty, U (%) +0.86 / −1.9
Effective degrees of freedom, DF +Inf
Coverage factor, k 1.96
Valid zenith angle range 30.0°–60.0°

FIGURE 5 History of instrument at zenith angle = 45°.


416 Appendix H

REFERENCES
[1] Reda, I.; Hickey, J.; Long, C.; Myers, D.; Stoffel, T.; Wilcox, S.; Michalsky, J. J.; Dutton, E. G.; Nelson, D.
(2005). “Using a Blackbody to Calculate Net Longwave Responsivity of Shortwave Solar Pyranometers
to Correct for Their Thermal Offset Error During Outdoor Calibration Using the Component Sum
Method.” Journal of Atmospheric and Oceanic Technology. 2005; pp.  1531–1540; NREL Report
No. JA-560-36646. doi:10.1175/JTECH1782.1.
[2] Reda, I.; Myers, D.; Stoffel, T. (2008). “Uncertainty Estimate for the Outdoor Calibration of Solar
Pyranometers: A  Metrologist Perspective.” Measure. (NCSLI Journal of Measurement Science).
Vol. 3(4), December 2008; pp. 58–66; NREL Report No. JA-581-4137.
[3] Reda, I.; Andreas, A. (2004). “Solar Position Algorithm for Solar Radiation Applications.” Solar Energy.
Vol. 76(5), 2004; pp. 577–589; NREL Report No. JA-560-35518. doi:10.1016/j.solener.2003.12.003.
[4] Stoffel, T.; Reda, I. (2009). “NREL Pyrheliometer Comparisons: 22 September–3 October 2008 (NPC-
2008).” 54 pp.; NREL Report No. TP-550-45016.
[5] Reda, I.; Stoffel, T.; Myers, D. (2003). “Method to Calibrate a Solar Pyranometer for Measuring
Reference Diffuse Irradiance.” Solar Energy. Vol. 74, 2003; pp. 103–112; NREL Report No. JA-560-
35025. doi:10.1016/S0038-092X(03)00124-5.
[6] Reda, I. (1996). “Calibration of a Solar Absolute Cavity Radiometer with Traceability to the World
Radiometric Reference.” 79 pp.; NREL Report No. TP-463-20619.
[7] Reda, I.; Gröbner, J.; Stoffel, T.; Myers, D.; Forgan, B. (2008). “Improvements in the Blackbody
Calibration of Pyrgeometers.” ARM 2008 Science Team Meeting (Poster).
Appendix H 417

Test Instrument: Black and White Pyranometer (Ventilated) Manufacturer: Eppley


Model: 8-48 Serial Number: 21096
Calibration Date: 5/16/2018 Due Date: 5/16/2019
Customer: NREL-SRRL-BMS Environmental Conditions: see page 4
Test Dates: 5/15-16

This  certifies that the above product was calibrated in compliance with ISO/IEC 17025:2005.
Measurement uncertainties at the time of calibration are consistent with the Guide to the Expression
of Uncertainty in Measurement (GUM) using Reda et al., 2008. All nominal values are traceable to
the International System (SI) Units of Measurement.
No statement of compliance with specifications is made or implied on this certificate; however,
the estimated uncertainties are the uncertainties of the calibration process. Users must add other
uncertainties that are relevant to their measuring system, environmental and sky conditions, out-
door set-up, and site location.
The Type-B Standard Uncertainty of using the responsivity at each even zenith angle is reported,
and the Expanded Uncertainty of the calibration is reported using two methods:

1. The Expanded Uncertainty of using the responsivity at zenith angle = 45°, within the zenith
angle range from 30 to 60 degrees.
2. The  Expanded Uncertainty of using Spline Interpolating Functions for the responsivity
versus zenith angle.

This certificate applies only to the item identified above and shall not be reproduced other that in full,
without specific written approval from the calibration facility. Certificate without signature is not valid.

TABLE 1
Traceability
Measurement Type Instrument Calibration Date Calibration Due Date
Beam Irradiance † Eppley Absolute Cavity 09/25/2017 09/25/2018
Radiometer Model HF, S/N 29219
Diffuse Irradiance† Hukseflux Pyranometer Model 04/11/2018 04/11/2019
SR25, S/N 2541
Diffuse Irradiance† Hukseflux Pyranometer Model 04/18/2018 04/18/2019
SR25, S/N 2542
Data Acquisition NREL Data Acquisition System 04/12/2017 04/12/2019
Model RAP-DAQ, S/N 2005-998
Data Acquisition NREL Data Acquisition System 04/12/2017 04/12/2019
Model RAP-DAQ, S/N 2005-999
†Through the World Radiometric Reference (WRR)
418 Appendix H

Number of pages of certificate: 4

Calibration Procedure: BORCAL-P00-Calibration and QA Procedure; available upon request.

Setup: Radiometers are calibrated outdoors, using the sun as the source. Pyranometers and
pyrgeometers are installed for horizontal measurements, with their signal connectors ori-
ented north, if their design permits.
The shading disk for the reference diffuse subtends a solid angle of 5°. Pyrheliometers are
installed on solar trackers.

Calibrated by: RCC


______________________________ ___________
Ibrahim Reda, Technical Manager Date

For questions or comments, please contact the technical manager at:


ibrahim.reda@nrel.gov; 303-384-6385; 15013 Denver West Parkway, Golden, CO 80401, USA
Appendix H 419

Calibration Results

21096 Eppley 8-48


The  responsivity (R, µV/W/m²) of the test instrument during calibration is calculated using this
Measurement Equation:

R = (V − Rnet * Wnet) / I [1]

where,
V = radiometer output voltage (microvolts),
I = reference irradiance (W/m²), beam (B) or global (G)
Rnet = radiometer net infrared responsivity (µV/W/m²), see Table 4,
Wnet = effective net infrared measured by pyrgeometer (W/m²),

= Win − Wout

= Win − σ * Tc^4

where,
Win = incoming infrared (W/m²)
σ = 5.6704e-8 W·m−2·K−4
Tc = case temperature of pyrgeometer (K)

G = B * COS(Z) + D,

where,
Z = zenith angle (degrees),
D = reference diffuse irradiance (W/m²).

FIGURE 1 Responsivity vs Zenith Angle. FIGURE 2 Responsivity vs Local Standard Time.


420

TABLE 2
Instrument Responsivity (R) and Calibration Type-B Standard Uncertainty, u(B)
AM PM AM PM
Zenith Zenith
R u(B) Azimuth R u(B) Azimuth R u(B) Azimuth R u(B)
Angle Angle Azimuth
(deg.) (µV/W/m²) ± (%) Angle (µV/W/m²) ± (%) Angle (deg.) (µV/W/m²) ± (%) Angle (µV/W/m²) ± (%) Angle
0 N/A N/A N/A N/A N/A N/A 46 11.301 0.29 102.13 11.282 0.29 258.09
2 N/A N/A N/A N/A N/A N/A 48 11.327 0.29 100.01 11.342 0.29 260.15
4 N/A N/A N/A N/A N/A N/A 50 11.397 0.29 98.06 11.316 0.30 262.09
6 N/A N/A N/A N/A N/A N/A 52 11.440 0.29 96.18 11.339 0.30 263.91
8 N/A N/A N/A N/A N/A N/A 54 11.460 0.30 94.42 11.332 0.30 265.83
10 N/A N/A N/A N/A N/A N/A 56 11.470 0.30 92.63 11.312 0.30 267.54
12 N/A N/A N/A N/A N/A N/A 58 11.502 0.30 90.90 11.353 0.31 269.00
14 N/A N/A N/A N/A N/A N/A 60 11.523 0.31 89.30 11.399 N/A 270.92
16 N/A N/A N/A N/A N/A N/A 62 11.596 0.31 87.58 11.370 0.33 272.20
18 N/A N/A N/A N/A N/A N/A 64 11.607 0.32 86.01 11.377 0.34 273.82
20 N/A N/A N/A N/A N/A N/A 66 11.644 0.33 84.39 11.411 0.35 275.45
22 11.179 0.28 155.67 11.207 0.28 203.52 68 11.628 0.34 82.79 11.388 0.36 276.97
24 11.167 0.28 144.43 11.220 0.28 215.46 70 11.652 0.35 81.16 11.448 N/A 278.57
26 11.186 0.28 136.71 11.229 0.28 223.40 72 11.634 0.37 79.64 11.436 N/A 280.17
28 11.190 0.28 130.62 11.276 0.28 229.35 74 11.645 0.39 78.10 11.401 N/A 281.78
30 11.203 0.28 125.74 11.246 0.28 234.33 76 11.628 0.42 76.51 11.468 N/A 283.37
32 11.216 0.28 121.64 11.303 0.28 238.04 78 11.551 N/A 74.88 11.446 N/A 284.96
34 11.224 0.28 117.99 11.268 0.28 242.27 80 11.515 N/A 73.29 11.429 N/A 286.56
36 11.245 0.29 114.58 11.245 0.29 245.44 82 11.399 N/A 71.59 N/A N/A N/A
38 11.237 0.29 111.60 11.256 0.29 248.37 84 11.535 N/A 69.98 N/A N/A N/A
40 11.257 0.29 109.01 11.276 0.29 251.11 86 N/A N/A N/A N/A N/A N/A
42 11.281 0.29 106.48 11.298 0.29 253.59 88 N/A N/A N/A N/A N/A N/A
44 11.284 0.29 104.24 11.296 0.29 255.93 90 N/A N/A N/A N/A N/A N/A

Abbreviation: N/A, Not Available.


Appendix H
Appendix H 421

FIGURE  3 Type-B Standard Uncertainty vs FIGURE 4 Residuals from spline interpolation.


Zenith Angle.

TABLE 3 TABLE 4
Uncertainty Using Spline Interpolation Calibration Label Values
Type-B Standard Uncertainty, u(B) (%) ±0.42 R @ 45° (µV/W/m²) Rnet (µV/W/m²)†
Type-A Interpolating Function, u(int) (%) ±0.32
10.866 0
Combined Standard Uncertainty, u(c) (%) ±0.53
Effective degrees of freedom, DF(c) 6178 †Rnet determination date: N/A
Coverage factor, k 1.96
Expanded Uncertainty, U95 (%) ±1.0
AM Valid zenith angle range 22°–76°
PM Valid zenith angle range 22°–68° TABLE 5
Uncertainty Using R @ 45°
Type-B Expanded Uncertainty, U(B) (%) ±0.62
Offset Uncertainty, U(off) (%) +0.57 / −0.32
Expanded Uncertainty, U (%) +1.2 / −0.95
Effective degrees of freedom, DF +Inf
Coverage factor, k 1.96
Valid zenith angle range 30.0°–60.0°

FIGURE 5 History of instrument at Zenith Angle = 45°.


422 Appendix H

REFERENCES
[1] Reda, I.; Hickey, J.; Long, C.; Myers, D.; Stoffel, T.; Wilcox, S.; Michalsky, J. J.; Dutton, E. G.; Nelson, D.
(2005). “Using a Blackbody to Calculate Net Longwave Responsivity of Shortwave Solar Pyranometers
to Correct for Their Thermal Offset Error During Outdoor Calibration Using the Component Sum
Method.” Journal of Atmospheric and Oceanic Technology. 2005; pp.  1531–1540; NREL Report
No. JA-560-36646. doi:10.1175/JTECH1782.1.
[2] Reda, I.; Myers, D.; Stoffel, T. (2008). “Uncertainty Estimate for the Outdoor Calibration of Solar
Pyranometers: A  Metrologist Perspective.” Measure. (NCSLI Journal of Measurement Science).
Vol. 3(4), December 2008; pp. 58–66; NREL Report No. JA-581-4137.
[3] Reda, I.; Andreas, A. (2004). “Solar Position Algorithm for Solar Radiation Applications.” Solar Energy.
Vol. 76(5), 2004; pp. 577–589; NREL Report No. JA-560-35518. doi:10.1016/j.solener.2003.12.003.
[4] Stoffel, T.; Reda, I. (2009). “NREL Pyrheliometer Comparisons: 22 September–3 October 2008 (NPC-
2008).” 54 pp.; NREL Report No. TP-550-45016.
[5] Reda, I.; Stoffel, T.; Myers, D. (2003). “Method to Calibrate a Solar Pyranometer for Measuring
Reference Diffuse Irradiance.” Solar Energy. Vol. 74, 2003; pp. 103–112; NREL Report No. JA-560-
35025. doi:10.1016/S0038-092X(03)00124-5.
[6] Reda, I. (1996). “Calibration of a Solar Absolute Cavity Radiometer with Traceability to the World
Radiometric Reference.” 79 pp.; NREL Report No. TP-463-20619.
[7] Reda, I.; Gröbner, J.; Stoffel, T.; Myers, D.; Forgan, B. (2008). “Improvements in the Blackbody
Calibration of Pyrgeometers.” ARM 2008 Science Team Meeting (Poster).
Appendix H 423

Environmental and Sky Conditions


for BORCAL-SW 2018-03
Calibration Facility: Solar Radiation Research Laboratory

Latitude: 39.742°N Longitude: 105.180°W Elevation: 1828.8 meters AMSL Time Zone: −7.0

Reference Irradiance:

FIGURE 6 Reference irradiance. FIGURE 7 Diffuse ratios.

Meteorological Observations:

FIGURE 8 Temperature. FIGURE 9 Humidity.

FIGURE 10 Pressure.
FIGURE 11 Effective net infrared.
424 Appendix H

FIGURE 12 Estimated broadband aerosol optical depth.

TABLE 6
Meteorological Observations
Observations Mean Min Max
Temperature (°C) 22.28 13.28 26.62
Humidity (%) 22.49 10.29 40.50
Pressure (mBar) 816.4 814.6 817.7
Est. Aerosol Optical Depth (BB) 0.077 0.032 0.881

For other information about the calibration facility visit: http://www.nrel.gov/solar_radiation/


Appendix H 425

Broadband Outdoor Radiometer Calibration
Longwave

BORCAL -LW
2018 - 06

Calibration Facility
Southern Great Plains
Latitude: 36.605°N
Longitude: 97.488°W
Elevation: 317.0 meters
AMSL Time Zone: -6.0

Calibration date
10/17/2018 to 11/06/2018

Report Date
November 6, 2018

NOTICE
This report was prepared as an account of work sponsored by an agency of the United States gov-
ernment. Neither the United States government nor any agency thereof, nor any of their employees,
makes any warranty, express or implied, or assumes any legal liability or responsibility for the
accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed,
or represents that its use would not infringe privately owned rights. Reference herein to any specific
commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does
not  necessarily constitute or imply its endorsement, recommendation, or favoring by the United
States government or any agency thereof. The views and opinions of authors expressed herein do
not necessarily state or reflect those of the United States government or any agency thereof.
426 Appendix H

Broadband Outdoor Radiometer Calibration Report

Introduction
This  report compiles the calibration results from a Broadband Outdoor Radiometer Calibration
(BORCAL). The work was accomplished at the Radiometer Calibration Facility shown on the front
of this report. The calibration results reported here are traceable to the World Infrared Standard
Group (WISG).
This report includes these sections:

• Results Summary: A  table of all instruments included in this report summarizing their
calibration results and uncertainty.
• Control Instruments: Aa group of instruments included in each BORCAL event that pro-
vides a measure of process consistency.
• Instrument Details: The calibration certificates for each instrument.
• Environmental and Sky Conditions: Meteorological conditions and reference irradiance
during the calibration event.

Results Summary

TABLE 1
Results Summary
K0 K1 U95
Instrument Customer (W/m2) (W/m2/μV) K2 K3 Kr* (W/m2) Page
30681F3 Eppley SGP −9.6 0.25570 1.0299 −3.96 7.044e-4 ±2.8 A1-11

Note: Environmental Conditions for BORCAL starts on Pages A1-20.


*Kr used to derive K0, K1, K2, K3

Appendix 1
Instrument Details
Calibration Certificates: 3 pages for each radiometer (4 including Environmental Conditions)
Environmental Conditions for BORCAL: Last Page of a Calibration Certificate. Note:
This appears only once, at the end of Appendix 1.
Appendix H 427

Control Instrument History

FIGURE 1 Eppley PIR control instrument History (K0 coefficient).

FIGURE 2 Eppley PIR control instrument History (K1 coefficient).

FIGURE 3 Eppley PIR control instrument History (K2 coefficient).

FIGURE 4 Eppley PIR control instrument History (K3 coefficient).


428 Appendix H

Test Instrument: Downwelling Pyrgeometer (Ventilated) Manufacturer: Eppley


Model: PIR Serial Number: 30681F3
Calibration Date: 11/6/2018 Due Date: 11/6/2019
Customer: SGP Environmental Conditions: see page 4
Test Dates: 10/17-24, 10/26-31, 11/1-6

This  certifies that the above product was calibrated in compliance with procedure listed below.
Measurement uncertainties at the time of calibration are consistent with the Guide to the Expression
of Uncertainty in Measurement (GUM) using Reda et al., 2008. All nominal values are traceable to
the World Infrared Standard Group (WISG).
No statement of compliance with specifications is made or implied on this certificate; however,
the estimated uncertainties are the uncertainties of the calibration process. Users must add other
uncertainties that are relevant to their measuring system, environmental and sky conditions, out-
door set-up, and site location.
This certificate applies only to the item identified above and shall not be reproduced other that
in full, without specific written approval from the calibration facility. Certificate without signature
is not valid.

TABLE 1
Traceability
Measurement Type Instrument Calibration Date Calibration Due Date
Data Acquisition NREL Data Acquisition System 01/09/2018 01/09/2019
Model RAP-DAQ, S/N 2009-1206
Data Acquisition NREL Data Acquisition System 01/09/2018 01/09/2019
Model RAP-DAQ, S/N 2009-1207
Data Acquisition NREL Data Acquisition System 01/09/2018 01/09/2019
Model RAP-DAQ, S/N 2009-1208
Data Acquisition NREL Data Acquisition System 01/09/2018 01/09/2019
Model RAP-DAQ, S/N 2014-1302
Infrared Irradiance‡ Eppley Downwelling Pyrgeometer 05/08/2017 05/08/2019
Model PIR, S/N 30835F3
Infrared Irradiance‡ Eppley Downwelling Pyrgeometer 06/27/2017 06/27/2019
Model PIR, S/N 31637F3
‡Through the World Infrared Standard Group (WISG)
Appendix H 429

Number of pages of certificate: 4

Calibration Procedure: SGP BORCAL-LW Calibration Procedure

Setup: Radiometers are calibrated outdoors, using the atmosphere as the source. Pyranometers
and pyrgeometers are installed for horizontal measurements, with their signal connectors
oriented north, if their design permits.

Calibrated by: Mike Dooraghi and Craig Webb

_________________________ __________
Michael Dooraghi, Technical Manager Date

For questions or comments, please contact the technical manager at:


Mike.Dooraghi@nrel.gov; 303-384-6329; 15013 Denver West Parkway, Golden, CO 80401, USA
430 Appendix H

Calibration Results

30681F3 Eppley PIR


The incoming irradiance (Win, W/m²) of the test instrument during calibration is calculated using
this Measurement Equation:

Win = K0 + K1*V + K2*Wr + K3*(Wd − Wr) [1]

where,
K0, K1, K2, K3 = calibration coefficeints,
V = thermopile output voltage (µV),

Wd = σ * Td^4 = dome irradiance (W/m²),

where,
Td = dome temperature (K),

Wr = σ * Tr^4 = receiver irradiance (W/m²),

where,
σ = 5.6704e-8 W·m−2·K−4,
Tr = Tc + Kr * V = receiver temperature (K),
Tc = case temperature (K),
Kr = efficiency coefficient (K/µV).

FIGURE  1 Residuals for calc. vs ref. irradiance FIGURE  2 Residuals for calc. vs ref. irradiance
using K0<>0 coefficients. using K0 = 0 coefficients.
Appendix H 431

TABLE 2 TABLE 3
Calibration Coefficients for K0<>0 Calibration Coefficients for K0 = 0
K0 −9.6 K0 0.0
K1 0.25570 K1 0.25593
K2 1.0299 K2 1.0042
K3 −3.96 K3 −4.33
Kr used to derive coefficients 7.044e-4 Kr used to derive coefficients 7.044e-4

TABLE 4 TABLE 5
Uncertainty Using K0<>0 Coefficients Uncertainty Using K0 = 0 Coefficients
Type-B Standard Uncertainty, u(B) (W/m²) ±1.4 Type-B Standard Uncertainty, u(B) (W/m²) ±1.4
Type-A Standard Uncertainty, u(A) (W/m²) ±0.25 Type-A Standard Uncertainty, u(A) (W/m²) ±0.57
Combined Standard Uncertainty, u(c) (W/m²) ±1.5 Combined Standard Uncertainty, u(c) (W/m²) ±1.5
Effective degrees of freedom, DF(c) +Inf Effective degrees of freedom, DF(c) +Inf
Coverage factor, k 1.96 Coverage factor, k 1.96
Expanded Uncertainty, U95 (W/m²) ±2.8 Expanded Uncertainty, U95 (W/m²) ±3.0

FIGURE 3 History of instrument (K0 coefficient).

FIGURE 4 History of instrument (K1 coefficient).


432 Appendix H

FIGURE 5 History of instrument (K2 coefficient).

FIGURE 6 History of instrument (K3 coefficient).

REFERENCE
[1] Reda, I.; Stoffel, T. (2010). “Pyrgeometer Calibration for DOE-Atmospheric System Research Program
using NREL Method (Presentation).” 9 pp.; NREL Report No.  PR-3B0-47756; http://www.nrel.gov/
docs/fy10osti/47756.pdf.
Appendix H 433

Environmental and Sky Conditions


for BORCAL-LW 2018-06

Calibration Facility: Southern Great Plains

Latitude: 36.605°N Longitude: 97.488°W Elevation: 317.0 meters AMSL Time Zone: −6.0

Reference Irradiance (hourly averages):

FIGURE 6 Reference irradiance.

Meteorological Observations (hourly averages):

FIGURE 7 Temperature. FIGURE 8 Humidity.


434 Appendix H

FIGURE 9 Pressure. FIGURE  10 Estimated precipitable water vapor


(PWV).

TABLE 6
Meteorological Observations
Observations Mean Min Max
Temperature (°C) 11.48 4.57 23.30
Humidity (%) 73.71 36.17 100.03
Pressure (mBar) N/A N/A N/A
Est. Precipitable Water Vapor (mm) 17.5 9.4 30.0

For other information about the calibration facility visit: http://www.arm.gov/docs/sites/sgp/sgp.html


Appendix I: Failure Modes

The following is a list of possible problems that have been compiled from past experiences, mainly
from the Atmospheric Radiation Measurement (ARM) research facilities1 and the experiences from
developing and operating a solar measurement station at our respective institutions.2
Equipment references indicate the original instrument systems as first installed. See Appendix G
for an updated list of equipment suppliers offering the latest instrumentation options. Some key
points before addressing possible equipment failures:

• Proper maintenance is the final step in quality control of measurements. After measure-
ments are recorded, only data quality assessment is possible.
• A situation with no data is often better than one that produces bad data. Unless bad data
are identified and flagged, they can destroy confidence in and value of the dataset.

The first section covers observed problems, and possible causes are noted. Next the corrective main-
tenance actions will be listed. Because all ARM and University of Oregon solar monitoring network
sites use Campbell data loggers for solar and infrared measurements, comments on data loggers will
be specific for Campbell data loggers, but the gist applies to all data loggers.

I.1 OBSERVED PROBLEMS


i.1.1 no signal
Possible Sources of Problem

1. An open circuit can be the result of a damaged signal cable, loose signal cable conductor
connections at the data logger input terminal, improperly seated connector at the radiom-
eter body, or broken thermopile winding (the latter typically due to lightning strike).
2. The data logger system program may also have been changed or lost.
3. The data logger internal battery voltage may be below minimal limits.

i.1.2 unstable signal


Possible Sources of Problem

1. Loss of proper electrical grounding and shielding of the low-level direct current (DC)
signal cable may cause erratic voltage readings.
2. The radiometers have a time-constant on the order of 1 second to changes in irradiance
(voltage) due to clouds. Electrical noise will generally have a higher frequency than this
natural variation.
3. Data logger system has failed.
a. The data logger system program may have been changed or lost
b. The data logger internal battery voltage may be below minimal limits

1 Atmospheric Radiation Measurement User Facility: https://www.arm.gov/capabilities/instruments/sirs.


2 University of Oregon Solar Radiation Monitoring Laboratory: http://solardat.uoregon.edu/NOAA/ESRL/GMD
Surface Radiation Budget (SURFRAD) Network: https://www.esrl.noaa.gov/gmd/grad/surfrad/
NREL Measurement and Instrumentation Data Center: https://midcdmz.nrel.gov/.

435
436 Appendix I

4. Other possible causes include


a. Moisture inside the radiometer case or signal cable connector
b. An improperly seated connector at the radiometer body
c. Loose signal cable connections at the data logger input terminal
d. Cold solder joint inside the signal cable connector
e. RF noise being rectified in the cable.

i.1.3 signal gReateR tHan PHysiCal limits


Possible Sources of Problem

1. The  global horizontal irradiance (GHI) pyranometer can, with unique reflections from
properly positioned towering cumulus clouds, exceed the solar constant for periods of less
than tens of minutes.
2. Moisture inside the pyranometer dome or the pyrheliometer window can focus radiation on
the radiometer detector.
3. Half-melted frost or snow on the dome can reflect sunlight onto the sensor.
4. More often, high signals are the result of an open circuit causing the data logger to over
range (6999), a ground loop introducing a spurious voltage, or an incorrect radiometer
calibration factor or logger program.

i.1.4 signal less tHan PHysiCal limits


1. Signal has been shorted by damaged cable.
2. Corrosion at signal cable connections has caused increased resistance.
3. Wrong calibration factor or application associated with instrument.
4. Moisture inside the pyranometer dome or behind the pyrheliometer window.
5. At night, thermal offsets can result in negative values.
6. Lightning can result in noise spikes, either negative or positive.

i.1.5 gHi, dni, and dHi not inteRnally Consistent


This inconsistency is most easily determined from Equation 2.1 on Chapter 2. In addition to bad
radiometer calibration factors (Cf), possible causes are

1. Direct normal irradiance (DNI) low


a. Incorrect tracker alignment
b. Dirty water, ice on normal incidence pyrheliometer (NIP) window
c. Nearby obstructions (trees, poles, structures) that shade or sometimes reflect sunlight
onto the detector
2. DNI high
a. Suspect electrical problem first (see the section on signal greater than physical limit)
b. Not many physical reasons for NIP to have prolonged increased output—can have
temporary increase when sun is next to the edge of a cloud
c. Contaminated NIP window, especially due to frost or dew, typically reduces reading,
but may cause increased signal with specific solar geometry
3. GHI low
a. Dirty/iced dome
b. Possibly bad electrical connections
c. Nearby obstructions, trees, poles, or structures shade the pyranometer
d. Alignment problem—pyranometer tilted toward the pole
Appendix I 437

4. GHI high
a. If persistent, suspect electrical problem first
b. Could be leveling problem, tilted toward equator
c. Water droplets (dew) on dome can focus sun’s rays
d. Ice or frost on dome opposite the sun’s location can reflect sun’s rays onto detector
e. Nearby reflective surfaces, such as buildings and poles
f. Artificial lights (at night)
5. Diffuse horizontal irradiance (DHI) low
a. Dirty/iced dome
b. Possibly bad electrical connections
c. Correct pyranometer for thermal-offset problem—black and white pyranometers will
not exhibit this behavior
d. Pyranometer not level, tilted to north
6. DHI high
a. Suspect misaligned solar tracker (dome must be in full shade of ball)
b. Improper pyranometer thermal offset correction applied to DHI measurement
(e.g., if shaded pyranometer and pyrgeometer are not coplanar).
c. Pyranometer not level, tilted to south
d. Possible ground loop or other electrical problems (see the section on no signal or unsta-
ble signal)
e. Nearby reflective surfaces, such as buildings and poles
f. Artificial lights (at night)
7. Upwelling GHI low
a. Dirty/iced dome
b. Possibly bad electrical connections
c. Field of view of local terrain biased by dark surfaces—water pooling?
d. Unlevel mounting platform
8. Upwelling GHI high
a. Fresh snow cover can reflect over 90 percent of the GHI
b. Frost on dome can reflect more radiation onto pyranometer detector
c. Possible ground loop or other electrical problems
d. Field of view of local terrain biased by bright surfaces—snow pooling?
e. Unlevel mounting platform

i.1.6 sHoRtWave elements (dni, gHi, and dHi)


Internally Consistent but WRONG (K-space)

1. Tracker failure (if pyrheliometer and shaded pyranometers are mounted on the same
tracker, then DNI = 0 and DHI = GHI, where GHI is indicative of clear sky irradiance
levels)
2. Tracker alignment problem?
3. Uniform soiling of optics

i.1.7 no data ColleCted by site data system


1. Suspect communications failure
2. Modem failure
3. Loss of electrical power to data logging system
438 Appendix I

i.1.8 no data ColleCted by CaRd stoRage module


1. Card not initialized at last site visit (card was full)
2. Card battery failure
3. Failed connection between logger and card storage module
4. Loss of electrical power to data logger system for more than 10 days

i.1.9 asymmetRiC diuRnal PRofiles (symmetRy WitH solaR noon, not CloCk noon
1. Incorrect data logger time (clock drift? improper switch to daylight savings time?)
2. Incorrect time zone
3. Incorrect longitude
4. Assignment of hours (0–23 vs 1–24)
5. Instrument alignment (pyranometer not level?)

i.1.10 bad data time sequenCe


1. Data buffer overrun due to corrupt memory?

I.2 CORRECTIVE MAINTENANCE ACTION OPTIONS


Based on experience and an understanding of the irradiance measurement processes, the following
actions are suggested for the previous failure modes. The following information should complement
the existing instructions for routine maintenance.

I.3 CORRECTIVE MAINTENANCE ACTIONS BY-THE-NUMBER


i.3.1 Clean oPtiCs

Note: Excessive dew, frost, snow, or dust on the precision spectral pyranometer (PSP) or precision
infrared radiometer (PIR) can indicate ventilator failure.
1. Contamination of optical surface
a. Wash dome or window with distilled water, wipe dry. A lint free cotton cloth should
be used for cleaning. Compressed air without contaminants can be used to clean dust
from the optics.
b. Wash dome or window with alcohol, wipe dry
c. Warm iced dome or window with palm of hand if necessary—do not scrape ice from
dome or window
d. Moisture inside the dome or window?
e. Check PSP desiccant canister seal is tight
f. Check PSP desiccant canister window is not cracked
g. Remove PSP sunshade and check dome collar screws are tight
h. Check NIP window screws are tight
i. Change desiccant if granule colors are changed showing moisture saturation.

i.3.2 CHeCk/adjust oRientation


1. Align Solar Tracker (north/south and base is level)
2. Adjust shading balls to shade PSP and PIR domes
3. Level the support arms when in the base position
Appendix I 439

4. PSP/PIR is level using circular spirit level


5. Signal cables not caught on tracker or mounting fixture

i.3.3 examine instantaneous data


1. For  Campbell Scientific Data Loggers that use Loggernet program in monitor mode to
view all data channels
a. Compare with expected readings3
b. Open channel value = “6999”
c. Note lightning can burn out delicate thermopile windings

i.3.4 ConfiRm ventilatoR is funCtioning


1. Confirm electrical power available
2. Sun shade must be level with black detector or base of PIR or pyranometer dome
a. The sun shade can shadow or reflect radiation onto detector if not properly positioned
at or slightly below the base of the outer dome

i.3.5 CHeCk signal Cable and ConneCtions


1. Confirm cable is not cut, stretched, or kinked
2. Check electrical connection at data logger wire panel and at radiometer connector
a. Hot and cold cycling loosens some wire connections

i.3.6 CHeCk data aCquisition system gRound ConneCtions


1. Check electrical connection at data logger, enclosure, and ground rod for corrosion or loose
fittings

i.3.7 ConfiRm seRial numbeRs WitH instRument loCation


1. Visually inspect radiometer serial numbers at each mounting location and compare with
data logger program locations (Monitor mode) or current inventory listing

i.3.8 ConfiRm CalibRation faCtoRs WitH instRument seRial numbeR


1. Compare calibration sticker information with data logger program locations (Monitor
mode) or current inventory listing

i.3.9 data loggeR PoWeR


1. The red LED indicator on power supply should be illuminated
2. Check logger battery voltage is at least 9.2 Vdc using Monitor mode
3. Check the “on–off” switch is in the “on” position
a. Note confusing label on CR10X-1M wiring panel—switch is pointing away from “on”
label when in operating position
4. If a PV panel charges the battery, use a charge controller to increases the charging rate
during cold weather.
3 Visit https://midcdmz.nrel.gov/apps/maintday.pl?BMS for examples of diagnostic plots comparing daily summaries of
1-minute measurements.
440 Appendix I

i.3.10 data loggeR CloCk


1. Confirm CR10X-1M logger clock is within 3 seconds of Greenwich mean time (GMT)
a. If a computer is checking the data logger clock, make sure the computer’s clock is
within 1 or 2 seconds of GMT.
b. Be sure that both the computer watching the clock and the data logger clock are both
set for standard time or daylight time.
Appendix J: How to Build
a Pyranometer with a Solar
Cell or Photodiode
Most thermopile-based pyranometers are expensive and require quality engineering and machining
in order to build a reliable instrument. Pyranometers based on the use of silicon photodiodes are
less expensive and it is relatively easy to build a photodiode-based pyranometer from items available
from any distributor that sells electronic components. Pyranometers using photodiodes are con-
structed with a silicon solar cell in a circuit designed to measure the photocurrent. Although home-
made pyranometers do not  have the accuracy and reliability of manufactured photodiode-based
pyranometers, they can be constructed from readily available parts and produce useful irradiance
values. In addition, building a pyranometer from off-the-shelf parts provides a learning experience
for those interested in solar energy and electronics.
There are two ways to build a pyranometer based on measuring the short circuit current of a solar
cell. The first involves an ordinary silicon solar cell about 8 by 8 centimeters (cm) (3 by 3 inches
(in)), a low value resistor (0.5 ohm (Ω) or less), and a voltmeter with low resistance leads to measure
the voltage drop across the resistor. The second way is to use a silicon photodiode, add a resistor
between the two leads and measure the voltage drop across the resistor. The methodology to create
a pyranometer from a solar cell will be discussed first followed by a description on how to use a
photodiode to create a pyranometer.

J.1 BUILDING A SOLAR CELL TO BUILD A PYRANOMETER


Understanding the fundamental characteristics of a solar cell and its behavior in a circuit helps when
designing and building a solar cell-based pyranometer. In a circuit, a solar cell behaves like a bat-
tery in that the relationship between voltage (V) and current (I) is not linear. Generally, the voltage
of a 1.5 volt AA battery is always around 1.5 volts independent of the load being powered by the
battery. As the battery ages, the voltage will decrease a little, but the current output by the battery
decreases markedly. A  “dead” battery may still measure over 1  volt. A  solar cell uses the Sun’s
energy to produce power instead of chemical energy in a battery. The voltage from a solar cell is
around 0.5 V and the current generated by the solar cell is dependent on the radiation incident upon
the cell. As with a chemical battery, the current output is fairly steady until the load is sufficient
to hinder the flow. A plot of the current (I) versus voltage (V), called the IV curve, for a 64 square
centimeters (cm2) (9 square inches (in2) solar cell is shown in Figure J.1. The maximum voltage is
called the open circuit voltage and occurs when the there is no current. The short circuit current is
the maximum current and is obtained when the voltage is 0.
Also drawn in Figure J.1 are plots of current verses voltage for given resistors. These straight
lines obey Ohm’s law that voltage equal current multiplied by resistance. When a resistance or load
is inserted into the circuit, the corresponding current and voltage change. For small loads, 0.1 Ω
in this example, the current is fairly constant and is dependent on the incident radiation and the
voltage in the circuit changes. For large loads, 2 Ω in Figure J.1, the current drops rapidly and the
voltage remains fairly constant. Ultimately, with very high loads (resistance), no current flows and
the open circuit voltage is obtained. For loads in between, 0.5 Ω in this example, both the current
and voltage change. The current times the voltage is the power generated by the solar panel. When

441
442 Appendix J

FIGURE J.1 Plot of the current (I) versus voltage (V) of a solar cell for three different intensities of solar
radiation. These are called IV curves. At low voltages the current is fairly constant and as the voltage increases
towards the maximum voltage the current decreases sharply to 0. The “knee” of the curve is where the solar
cell produces the most power. Power is equal to current times voltage. When a resistive load is in the circuit,
the current and voltage across the resistor form a straight line as the solar intensity changes, Ohm’s law.
The three straight lines plot current versus voltage for three different resistances.

using the solar cell as a pyranometer, interest is not in the power produced, but on the relationship
of the current to the incident solar radiation. The part of the IV curve that is most proportional to
the incident irradiance occurs with very small loads and at the short circuit current. When using just
one cell, the voltage and current are reduced dramatically. Therefore to measure the cell output near
the short circuit voltage, the resistance must be very small and it is difficult to accurately measure
the voltage across such a small resistance. An example of such a homemade solar cell pyranometer
can be found at http://www.instesre.org/construction/pyranometer/pyranometer.htm. This  type of
pyranometer works best with a larger solar cell say 8 by 8 cm (about 3 by 3 in). With a resistor about
0.06 Ω, a voltage drop across the resistor of 100 millivolts (mV) would occur under a full Sun.

J.2 BUILDING A PHOTODIODE-BASED PYRANOMETER


A  photodiode-base pyranometer is essentially a package using a solar cell to measure the
photocurrent. Using a silicon photodiode instead of a solar cell to build a pyranometer enables one to
have a smaller pyranometer while still producing a voltage that can be read by a typical multimeter.
As  discussed in Chapter  5, the photocurrent of a photodiode can be obtained by measuring the
voltage across a resistor in series with the cell’s series resistance as long as the combined resistance
of the resistor and series resistance is significantly smaller than the shunt resistance (see Figure 5.19
in Chapter  5). The  solar cell in the photodiode is much smaller than a typical solar cell (1 by
1 millimeter (mm) compared to 10 by 10 cm, respectively) and the shunt resistance is larger as shunt
resistance can increase as the area of the solar cell decreases, depending of the source of the shunt
resistance. Therefore, adding a resistor on the order of 100–200 Ω to the cell’s series resistance, is
still small compared to the shunt resistance. The voltage across this load resistor is proportional to
the photocurrent generated by the incident radiation (Equation 5.2 in Chapter 5). See Section 5.4 in
Chapter 5 for a more detailed discussion.
Appendix J 443

The series and shunt resistance of a solar cell result from the structure of the solar cell. The effects
of temperature on these resistances is small. To minimize the effect of temperature on the measured
values, the resistor placed between the two leads from the photodiode should have minimal temper-
ature dependence (the resistance should be nearly constant over the typical range of temperatures.
Final wiring will have one side connected to the positive terminal of the volt meter and the other
to the ground. Alternately one could measure the current directly with an ampmeter. A 100–500 Ω
resistor across the photodiode produces a voltage that is proportional to the incident solar radiation.
A resistor greater than ~500 Ω becomes more comparable to the shunt resistance and the relation-
ship between incident radiation and current becomes non-linear.
A ¼ watt metal film resistor has a very small temperature dependence and is the type of resistor
that should be used (Brooks and Mims 2001). A website that offers detailed instructions on how to
build these photodiode-based pyranometers is http://www.instesre.org/construction/pyranometer/
pyranometer.htm.
Constructing a pyranometer that can be used in the field during all weather conditions requires
an enclosure to protect the circuitry from the weather and a window so that sunlight can strike
the photodiode. A Teflon® disk can be used for the window as it is easy to cut and passes irradi-
ance in the 400–1,100 nanometer (nm) range that is used by a photodiode. Other components of
the enclosure are a small cylinder to house the photodiode, a box to contain the wiring and mount
to a surface, and a bulls eye level to help ensure that the photodiode is level. When building the
pyranometer, it is important that the base of the cylinder in which the photodiode is mounted and the
face of the photodiode are parallel. The Teflon® disk goes over the opening of the cylinder and must
be well sealed so that water does not enter. Superglue is used to form a water-tight seal between the
Teflon® disk and the cylindrical plastic pipe. The superglue should be in a small tube in liquid form
that is ease to apply accurately. Rubber grommets are needed so that wires coming from the box do
not let water enter. Wires connecting the photodiode pyranometer to the meter are needed (18 gauge
should be more than sufficient). Other equipment used to build the pyranometer are a soldering gun,
solder, and wire cutters. See Tables J.1 and J.2 for lists of components and tools.
The pyranometer still needs to be calibrated to change its millivolt readings into watts per square
meter proportional to the incident irradiance. The  typical way to calibrate a pyranometer is to
compare the output against a pyranometer that has its calibration traceable to the international
standard (See Chapter 13). This method can be an expensive proposition. There is an alternative
method that can supply a reasonable answer. On a cloudless day, set the pyranometer outdoors

TABLE J.1
Components Needed to Build a Solar Cell Pyranometer
Component or Tool Number
Silicon photodiode 1
Resistor—¼ watt 1% metal film 100–500 Ω 1
LED holder T 1-3/4 to hold photodiode 1
Teflon® disk 1 mm (~0.04ʺ) thick and 1 cm (3/8ʺ) in diameter approximately 2 cm (~3/4ʺ) long 1
Small “bulls eye” bubble level 1
Plastic case with holes for LED holder and wires to meter 1
PVC pipe—inside diameter 0.6 cm (~1/4ʺ) about 1
Cable—twisted shielded pair approximately 18 guage—2 to 3 m (~5–10 ft) long 1
Rubber Grommet that fits the cable and can seal the entry hole in the case 1
Cable ties 2
Solder with resin core—0.8 mm (~1/32ʺ) in diameter to connect components 1
Super Glue—small tube of liquid glue 1
444 Appendix J

TABLE J.2
Tools to Assemble the Kit
Tool
Soldering iron (15–40 watt)
Wire cutters
Pliers
Drill with drill bits that match the component sizes
Hole punch if case does not have the appropriate holes

and measure the voltage and record the time of day. Then search internet sites that can be used
to calculate the solar zenith angle such as https://www.esrl.noaa.gov/gmd/grad/antuv/SolarCalc.
jsp?mu=on&sza=on&el=on&az=on. Enter the time of day, the latitude, longitude, and altitude of
the site where the measurements were made. A fair estimate of the solar radiation incident upon
the pyranometer is 1,000 watts per square meter (W/m2) times the cosine of the solar zenith angle.
The calibration constant of the pyranometer is then equal to the estimate irradiance divided by the
voltage measurement of the reading. The irradiance at any other time can then be calculated by
multiplying the voltage output of the pyranometer by the calibration constant.

REFERENCE
Brooks, D. R., and F. M. Mims III. 2001. Development of an inexpensive handheld LED-based Sun photom-
eter for the GLOBE program. Journal of Geophysical Research 106:4733–4740.

OTHER USEFUL WEBPAGES


http://chuck-wright.com/projects/pv-measure.html
http://www.instesre.org/
http://www.instesre.org/Aerosols/order_form.htm
https://www.builditsolar.com/References/Measurements/KitPyr/KitPyr.htm
https://www.globe.gov/do-globe/for-students/student-research-reports/projectdetail/globe/influence-of-select-
atmospheric-data-on-winter-ground-level-aerosol-measurements-using-handheld-photometers
https://www.globe.gov/documents/10157/4178f8af-72d6-4db3-8bde-ea63dfdb5487
https://www.globe.gov/web/air-quality-campaign/overview/instruments
https://www.hukseflux.com/product/pyranometer-app?referrer=/product_group/pyranometer
https://www.schooot.nl/weerpyreng.htm
Appendix K: Content Required
for a Comprehensive Datafile
The goal of developing and operating a measurement station or network of stations for monitoring
solar and/or infrared (IR) radiation and other meteorological parameters is to create a database that
can be used to address the needs of a target audience and as many potential users in a simple yet
comprehensive manner. The information in this appendix provides an overview of the key elements
necessary for creating a successful measurements database.
Irradiance and meteorological data gathered by the University of Oregon Solar Radiation
Monitoring Laboratory (SRML) since 1975 have been available on the internet for many years.
By having the data available online, users do not have to contact SRML directly for data unless
they have specific questions. Originally, the data format was developed to be compatible with the
Research Cooperator Format originally developed by the National Climatic Data Center1 (NCDC)
and designed for the Cooperative Network for Renewable Resource Measurements (CONFRRM)2
network for data sharing (SERI 1988). At the time of its development, data were generally archived
on reels of magnetic tape, disk space was limited, and files had to be compact. Therefore, the
files consisted of ASCII tab-separated integers. Associated with the data files are documents that
explained the file structure and what is contained in the files. Today, such a file structure is archaic
and difficult to use. In addition, solar resource and meteorological data are now used by develop-
ers and financial institutions to evaluate operation and fiscal performance of solar facilities. More
detailed synopsis, or information about the data, is required to provide confidence in the analyses
that result from using the database. It  could be said that this information is necessary to decide
whether the dataset is “bankable.”
Using lessons learned over the years, the SRML developed a new more comprehensive file for-
mat that contains significantly more information about the station and the various measurements
(Peterson and Vignola 2017). This information is contained in the data file and it is not necessary
to seek other files to find this information. The SRML file format is used to illustrate information
that is valuable to include in a comprehensive data file and the file contains examples of addi-
tional features that might prove useful. The new file format contains detailed information on the
specific instruments used to make the measurements including their model and serial number as
well as calibration values used to translate radiometer voltage measurements into irradiance values
and the uncertainty in the calibration values. Additional information such as basic facts about the
monitoring station are included along with the solar zenith and azimuthal angles that are provided
to assist users of the data.

K.1 UO SRML DATA FILE STRUCTURE—OVERVIEW


The ASCII data files are in comma separated variable (CSV) format and are broken into monthly blocks
for a manageable file size. The files contain a synopsis that contains computed daily total or average val-
ues followed by the highest time resolved data. An overview of the file structure is shown in Table K.1.

1 In 2015, the National Climatic Data Center (NCDC) became part of the National Centers for Environmental Information
(NCEI). https://www.ncdc.noaa.gov/about.
2 The  Cooperative Network for Renewable Resource Measurements (CONFRRM) was developed by the National
Renewable Energy Laboratory (NREL) and others to conduce long-term solar radiation and wind measurements in the
United States. https://www.nrel.gov/grid/solar-resource/confrrm.html.

445
446 Appendix K

TABLE K.1
Overview of File Structure
Section 1: Header Station ID Information Data Column Information Comments
Daily Information
Section 2: Synopsis of Data Date and Calculated Final data Initial Data Auxiliary
Auxiliary Information Information
Short Interval Information
Section 3: Data Day and Time, Calculated Final Data Initial Data Auxiliary
Auxiliary Information Information

The header section contains general information about the station and a description of what is contained
in each column along with information about the various instruments used to collect the data.
The daily total and highest time resolved data regions are subdivided into four parts using
the following metric. The  left columns contain data such as date, time, solar zenith angle,
azimuthal angle, sunrise time, sunset time, extraterrestrial radiation, etc. The center columns
contain processed and calculated irradiance information. The right columns contain the original
measured irradiance quantities and other metrological information such as temperature and air
pressure. Auxiliary data then are provided. The process used to create the initial data columns
are described for each section. Room for comments is included to allow an explanation of non-
typical situations, such as a change of instruments that affect the dataset. The extreme right side
of the datafile can be used for additional comments related to the row of data or information
to the left. Comments in the header can be more generic, but each comment only fills one cell
(commas not allowed).

k.1.1 tHe HeadeR
The header should include information about the monitoring station, the equipment used to obtain the
data, and the contents of the data file. Table K.2 shows the station ID portion of the header from a file
for the Burns station in the UO SRML network and is used to illustrate the basic station information
and file information. The station name, location (altitude, longitude, and latitude), time zone (relative
to UTC), are essential. Other critical information contained in this section of the header are the time
interval of the stored data along with the month and year of the data in the file. To facilitate the use
of the file by spreadsheets additional commas should be avoided. For example, the station location
is connected by underlines, for example, “Burns_Oregon_USA” instead of “Burns, Oregon, USA.”

TABLE K.2
Sample Structure of Station ID Information
Labels Station Values
Station ID Number: 94170
Station Name: BUO
Station Location: Burns_Oregon_USA
Latitude: 43.5192
Longitude (+ East): −119.0216
Altitude (m): 1260
Time Zone (+ East): −8
Time Interval (Minutes): 1
Year//Month 2016//12
Appendix K 447

The year and month are separated by a “//” instead of a single “/” to prevent spreadsheet programs
from reformatting the dates and times into their own predetermined format. This technique is also
applied to times where “::” are used to separate the hours from the minutes.
The rest of the header is filled with information on the data in the file and any procedures used to
obtain or modify the data. In this example, shown in Table K.3, information pertinent to the global
horizontal irradiance data (GHI) is listed. This includes the instrument type and serial number and
information on the responsivity and the percent uncertainty of the responsivity at the 95 percent
confidence level. Each data column is accompanied by a quality flag column that is used to indicate
any uncertainty or issues with the data. Since the column information is aligned with the data in the
body of the file, flag columns are also included in the header description.
The recommended information includes the instrument model and serial number, responsivity
used in changing the measured voltage into irradiance or other measured quantity, the uncertainty
in the estimated data, sampling method, and data units (see Table K.3). Information on determining
estimated uncertainties is included in the chapter-specific measurement discussions. An overview of
measurement uncertainty concepts is presented in Appendix A. The uncertainty estimate is included
in the header to provide the user with an idea of the uncertainty in the data that are included in the
file. The sample method box indicates if the data are averaged over the time interval or sampled
like the diffuse horizontal irradiance (DHI) measurement in a rotating shadowband radiometer
(RSR) instrument (see Chapter 10). In addition, the sampling method should include information
on how the time stamp in the file relates to the data. For example, the time stamp can represent the
beginning, middle, or end of the time in which the data was gathered. The column notes are used to
indicate if any procedure has been performed on the measured data, e.g., gap-filling by interpola-
tion or estimated from other measurements. Since the file is intended to be used in spreadsheets and
computer programs, the standard ASCII character set should be used. For example, W/m2 is used in
the header instead of W/m2 (see Table K.3).
The type of measurement is given in the top row of the header. The initials GHI corresponds to
Global Horizontal Irradiance, DNI corresponds to Direct Normal Irradiance, and DHI correspond
to Diffuse Horizontal Irradiance. Direct Horizontal Irradiance is labeled as DrHI. Three different
types of GHI data are represented in these data columns. On the right had side of the table is infor-
mation for the initial (or raw) GHI data that comes from the data logger. In the UO SRML data-
set, nighttime values or offsets (withNO) are recorded. When the data are processed into the final
reported data, the average night time offsets are subtracted from all irradiance data during the day.
Chapter 5 in this book discusses the thermal radiation exchange with the sky that gives rise to these
night time offsets. In general, “raw” data are on the right side of the file while processed data, with
some of the biases removed, are on the left most side of the data columns as shown in Table K.3.
In addition to the measured data, three types of irradiance values are calculated if the correspond-
ing irradiance data are available.

DrHI_Calc = DNI*cos(sza)

GHI_Calc = DNI*cos(sza) + DHI

DHI_Calc = GHI – DNI*cos(sza)

where sza is the cosine of the solar zenith angle over the time of measurement.
If DNI and DHI data are available then a GHI value is calculate and labeled as GHI_Calc.
If DNI and GHI data are available, then the DHI value is calculated and labeled as DHI_Calc.
Because GHI instruments have a measurable deviation from a true cosine response (Chapter  5),
calculated GHI_Calc values are useful because they exhibit little of the systematic bias associ-
ated with pyranometer cosine response. Similarly, DHI values calculated from GHI and DNI data
have large uncertainties (see Chapter 6). In the column notes, the columns are listed as processed,
448

TABLE K.3
Information for the Data Column
Type of Measurement GHI GHI_Flag GHI_Calc GHI_Calc Flag GHI_withNO GHI_withNO Flag
Element: 1,000 — 1,001 — 1,000_withNO —
Instrument Serial Number: PSP (23973F3) — Computed from DHI and DNI — PSP (23973F3) —
Instrument Shorthand Name: P23 — NA — P23 —
Responsivity: 8.6844 µV/(Wm−2) NA — 8.6844 µV/(Wm−2)
Estimated Uncertainty (U95%): 3.587 — 2.6 — 3.587 —
Sample Method: Avg ending at time — — — Avg ending at time —
Data Units: Wm−2 — Wm−2 — Wm−2 —
Column Notes: Processed Column — Calculated Column — Measured Column —
Appendix K
Appendix K 449

calculated, or measured. The calculated data values do not need to be included in a comprehensive


data format file, but they are useful for the analysis of the quality of the data set (i.e., comparing
measured and calculated irradiance components) and for providing input to other models.
When the three solar irradiance values are measured independently, a quality control check can
be made. With an ideal dataset

GHI − DNI ⋅ cos(sza) − DHI = 0 (K.1)

Of course, when measured data are input into Equation K.1, the result is rarely zero as the conse-
quence of radiometer uncertainties. However, the smaller the result in Equation K.1, the better the
measurements are likely to be. Conversely, any large value that results from Equation K.1 is likely
to indicate a problem with at least one of the measurements, indicating further evaluation is needed
for each component.
Measured data are useful as others may want to run other processes to remove known biases and
it is better to start with the raw data than to modify data that have already been adjusted using some
other process.
A comment column on the right side of the header section is included so that other information
pertinent to the data in the file can be added. For example, information regarding an instrument
change during the month should be included in the comment section of the file. The header also has
an additional row at the bottom of the header to allow for possible expansion in the future if more
information is required.
The  new UO SRML comprehensive data file format has information that relates to previous
data archive formats and are not generally applicable to generic file formats. These are the element
numbers that refer to an old Research Cooperator Format (RCF) that was used with the original files
and the shorthand name for instruments that are kept for legacy reasons. The number of header rows
depends on the amount of information that is necessary to describe the content of the file. For ease
of use, the important factor is that the number of header rows is consistent for the whole dataset.

k.1.2 synoPsis of data


A synopsis of the data contained in the file follows the header rows. The synopsis is useful for an
overview of the monthly data block and enables one to quickly find data of interest if one does
not want to look at all the higher resolution data. The synopsis in the UO SRML file format include
daily total or average data, depending on what is being measured. In addition to the average tem-
perature, columns for minimum and maximum temperatures are included. Other information, such
as average nighttime values are also included in the synopsis. These values are used when a simple
process is used to remove some of the thermal offset bias in the data.
The first seven columns of the daily total section include the day of month, day of year, sunrise
and sunset times, solar noon times, daily total extraterrestrial energy on a horizontal surface (ETR),
and daily total extraterrestrial energy on a normal surface (ETRn). The sunrise and sunset times
included in the file are good to approximately ±30  seconds, account for atmospheric refraction,
and do not take into account obstructions on the local horizon. Table K.4 illustrates the first few
lines of the synopsis block. The time information has “::” between the hour and minute to prevent
the time information from being reformatted when imported into or exported from a spread sheet.
To provide a uniform block for the daily total information, there are always 31 rows for the daily
total information. By having the header and synopsis sections always contain the same number of
rows, the short interval data always starts at the same location in the file.
The next columns in the synopsis section contain data pairs. The first data column contains the
data value (either daily total or daily average followed by a column with an estimate of the uncer-
tainty of the daily value. The uncertainty is expressed in percent of the measurement at the 95 percent
level of confidence. Several meteorological columns in the synopsis section, such as temperature,
450 Appendix K

TABLE K.4
Time Information for the Daily Total Section
Daily Total Daily Total
Day of Time Sunrise Time Sunset Time Solar Noon Energy ETR Energy ETRn
Day of Month Year (HH::MM:SS) (HH::MM:SS) (HH::MM:SS) (kW h/m2) (kW h/m2)
1 153 04::46:46 20::10:40 12::28:30 11.473 20.405
2 154 04::46:16 20::11:29 12::28:40 11.493 20.422
3 155 04::45:47 20::12:17 12::28:50 11.513 20.438
4 156 04::45:21 20::13:03 12::29:01 11.531 20.476
5.… 31 … … … … … …

contain minimum and maximum values in addition to the average temperature value. The synopsis
serves as a quick quality check on the data and provides a broad snapshot of the weather for the day
and the month. Many users may be initially interested only in the summary data; however, the more
detailed short interval data are available to address most questions that can arise.
In addition to the data and uncertainty, the nighttime offset is recorded for each day for each irra-
diance instrument. The nighttime offset is computed using the following method. Missing or bad
data are first eliminated and only quality-controlled data are used to compute the nighttime value.
Astronomical night is defined as when the Sun has a solar zenith angle greater than 108 degrees.
Only data points that have an SZA > 108° are used in the calculation of the average nighttime offset.
If there are no good data points for a particular date, the average nighttime value from the entire
month is used for the daily value. If there still are no good nighttime values, a reasonable nighttime
offset from past history of the instrument at the location is used. Along with the average nighttime
offset, the standard deviation (1 sigma) is calculated for each night. The average nighttime offset
and standard deviation of the nighttime values are both given in watts per square meter. These
average nighttime offsets are recorded in the synopsis section under the raw data columns and the
standard deviation values are under the flag column of the corresponding instrument. This infor-
mation serves as a variable in the processing procedure where the nighttime values are subtracted
from the raw data values. The standard deviation values are incorporated into the uncertainty values
estimated for the daily values. If there are less than 31 days in the month, rows with “-”s are inserted
to fill in the missing data and keep the location of the short time interval data in the same location
in the file. At the end of the synopsis section there is a row with “-”s to visibly separate the sections.

k.1.3 sHoRt time inteRval data


Following the synopsis section is the main content of the file with the short time interval data.
The  information in the first row at the beginning of the short time interval data columns is the
same as the first row in the header and describes the data content of the information in each column
below. The units for each data column are given in the 8th row of the header in the same column.
The first seven columns consist of time associated with the measurement, solar zenith angle, azi-
muthal angle, ETR, and ETRn as shown in Table K.5. Time information is presented in three for-
mats for ease of use. The first column is the year plus the fraction of the year, the second column is
the day of the year plus the fraction of the day, and the third time column is a standard date format
with double “--”s inserted between the day and the hourly of the measurement. The double “--”s
are used to prevent spreadsheets from automatically reformatting the date and time into the spread-
sheets’ predefined format. In the example in Table K.5, the data are separated by one-minute. Room
for seconds are included if future data files have shorter time intervals.
The data in the next columns are the measured data accompanied by a flag column. The flag column
contains information on any modifications that have been made to the data (see Table K.6). Quality con-
trol flags should be used with any dataset as they help the user assess the quality of  the data. There are
Appendix K 451

TABLE K.5
First Seven Columns of the Short Interval Data
year.fraction day of year. YYYY-MM-DD--
of year fraction of day HH:MM:SS SZA AZM ETR (Wm−2) ETRn (Wm−2)
2015.413701 152.0006944 2015-06-01--00:01:00 114.42591 1.69212 0 0
2015.413702 152.0013889 2015-06-01--00:02:00 114.42005 1.94657 0 0
2015.413704 152.0020833 2015-06-01--00:03:00 114.41339 2.20098 0 0
2015.413706 152.0027778 2015-06-01--00:04:00 114.40593 2.45535 0 0
2015.413708 152.0034722 2015-06-01--00:05:00 114.39766 2.70968 0 0
2015.41371 152.0041667 2015-06-01--00:06:00 114.38859 2.96395 0 0
2015.413712 152.0048611 2015-06-01--00:07:00 114.37872 3.21817 0 0
2015.413714 152.0055556 2015-06-01--00:08:00 114.36804 3.47233 0 0
2015.413716 152.0062500 2015-06-01--00:09:00 114.35656 3.72643 0 0
2015.413718 152.0069444 2015-06-01--00:10:00 114.34427 3.98045 0 0

TABLE K.6
Definition of Quality Control Flags in the Datafile
Other Metrological Data Irradiance Data Irradiance Data Irradiance Data
Quality of Measurement (Temp, Pressure, etc.) (Measured) (Processed) (Calculated)
Good data 11 11 12 72
Corrected data from other 21 21 22 22
instrument
Corrected data from this 31 31 32 32
instrument
Questionable data 81 81 82 82
Bad or missing data 99 99 99 99

a variety of possibilities. For example, there are the SERIQC flags (Maxwell et al. 1993) that are used in
some files available from the National Renewable Energy Laboratory (NREL) National Solar Radiation
Data Base (NSRDB). The flags used in the UO SRML files relate to modifications and/or adjustments
to the data, see Table K.6. The flags used in the UO SRML data files are listed in Table K.6. If the data
are not changed or modified, then a flag of “11” is next to the data. Data that has be processed to remove
a potential bias has a flag “12.” Calculated data are flagged “72” unless the data used in the calculation
has a flag other than “11” or “12.” If the data from another instrument is used to replace questionable
data, the data are flagged “21.” If data are inserted in place of problem data, then a flag of “31” is used.
An example of this is if the instrument is cleaned and other data shows that the sky was clear, then the
questionable data were replaced with an extrapolation between good data points. If the data are suspect,
then the data are flagged 81.
When the raw data files are processed to remove some of the thermal offset bias, then a “2”
replaces the “1” of the second digit of the file. As discussed in Chapter 5, the thermal offsets at night
are approximately half of the thermal offsets during the day under clear skies. This process, used
with the UO SRML data, is a quick way to remove some of the thermal offset bias without needing
concurrent pyrgeometer data that provides a more accurate estimate of the thermal offsets. If night-
time offsets are to be subtracted from the daytime data values, the instrument should be calibrated
by subtracting the nighttime offsets from the values obtained during the day.
452 Appendix K

The raw data files without subtracting the average nighttime values are also included if the user
would like to work with the original data. Almost all irradiance measurements have some system-
atic bias. Models have been developed over time to adjust the data to remove the biases, and these
models are evolving over time.
This comprehensive data format is designed for archive files and contains information that is
not normally used by modeling programs. To facilitate the use of the data by modeling programs a
simplified data structure is also available that doesn’t contain the synopsis and flag data. The num-
ber of header lines is reduced in the simplified file structure. In addition, only one type of each
measurement is included in the file.
Other information is useful in assessing the quality of the data. This information includes main-
tenance and cleaning records, and a calibration history of the instrument. Calibration information
should include calibration data from all solar zenith and azimuthal angles that are in the calibration
record. This information is too large to put into each data file. However, this information should be
available on the website where one can gain access to the data.
For  ease of access and manipulation of information computer access to a data base is ideal.
An example of a good data structure can be found at https://midcdmz.nrel.gov/srrl_bms/. A useful
database allows easy access to the period of interest and enables downloading of data from only the
instruments of interest. With a computer database plots of the data also can be obtained instead of
the digital values. The drawback with a computer database is that they require a skilled person to
create the database and a knowledgeable person to maintain and/or modify the database.

K.2 BENEFITS OF OTHER STANDARDS DATA FORMATS


There  are standards that exist that help facilitate the use of array-oriented scientific data. These
standards help facilitate the portability of datafiles to different computer systems and are referred
to as “self-describing” formats. The advantage with having a recognized standard format is that
software packages, both proprietary and open source often support these formats. Since a variety
of models exist that estimate solar system performance, providing data in a standard manner that
facilitates importation of irradiance datasets into these computer models would simplify efforts.
Developed by the Unidata Program at the University Corporation for Atmospheric Research
(UCAR), the NetCDF (Network Common Data Format)3 is a popular, open standard, self-describing,
machine-independent data archival method with supporting software libraries for the development
of array-oriented scientific data repositories (Treinish and Gough 1987). Freely distributed collec-
tions of data access libraries are available for C, Fortran, C++, Java, Python, R, and other program-
ming languages. The Atmospheric Radiation Measurement (ARM) user facility, supported for more
than 20 years by the U.S. Department of Energy, is an important example of how the NetCDF has
benefited a large number of data users. By design, ARM has a unique collection of solar and infra-
red radiation data for multiple locations world-wide available in NetCDF (Turner and Ellingson
2017). Considerable effort has been made to provide on-line tutorials, data quick-look features, and
other tools for simplifying the user access to vast amounts and types of scientific data (https://www.
arm.gov/data/work-with-arm-data).
Besides modeling the performance of solar systems, irradiance databases are used to develop
and validate radiative transfer, satellite-based remote sensing, and other models. When data archi-
val standards are established and followed, then it becomes easier to use the data that have been
collected by diverse groups. There is always a question that if a model is developed and tested at one
location, how well does it work at another location with a different climatology. By having a well-
defined dataset with clearly determined uncertainties, one can gain a level of confidence testing
models in different locations. Different instruments, different data loggers, and different mainte-
nance routines can all affect the quality of the data. However, if that information is available, then it

3 https://www.unidata.ucar.edu/software/netcdf/docs/index.html.
Appendix K 453

is possible to have a level of confidence that the data has uncertainties of less than 2, 5, or 10 percent
depending on how the data were gathered and with what equipment. If the uncertainty for the data-
set is not given, then it is necessary to estimate or guess at the uncertainty in the dataset. The quality
of any comparison cannot be ascertained with a dataset that does not provide the information neces-
sary to assess the uncertainty of that data set. At a minimum, basic information about the instru-
ments and data loggers used should be available along with a maintenance and calibration schedule.

QUESTIONS
1. What is meant by a “Self-Documenting” data format and why is this a useful attribute?
2. Why are the station coordinates (i.e., latitude, longitude, elevation, and time zone) important
for archiving solar irradiance measurements?
3. Typically, GHI measurements made from pyranometers exhibit a systematic bias as the
solar zenith angle changes. Why does this occur?
4. Why do GHI data calculated by projecting the DNI irradiance onto a horizontal surface
and adding the DHI irradiance exhibit minimal bias as a function of solar zenith angle?
5. What programming languages are supported by NetCDF libraries?

REFERENCES
Peterson, J., and F. Vignola. 2017. Structure of a comprehensive solar radiation dataset. Conference Proceedings,
ASES National Solar Conference 2017, Denver, CO, October 9–12.
Maxwell, E., S. Wilcox, and M. Rymes. 1993. User’s Manual for SERI QC Software, Assessing the Quality
of Solar Radiation Data. NREL/TP-463-5608, DE93018210. Golden, CO: National Renewable Energy
Laboratory. https://www.nrel.gov/docs/legosti/old/5608.pdf (Accessed June 4, 2019).
SERI. See Solar Energy Research Institute.
Solar Energy Research Institute. 1988. SERI Standard Broadband Format: A Solar and Meteorological Data
Archival Format. SERI/SP-320-3305. Golden, CO: Solar Energy Research Institute. https://www.nrel.
gov/grid/solar-resource/assets/data/3305.pdf.
Treinish, L. A., and M. L. Gough. 1987. A software package for the data independent management of multi-
dimensional data. EOS Transactions American Geophysical Union 68:633–635.
Turner, D. D., and R. G. Ellingson, eds. 2017. The Atmospheric Radiation Measurement Program, The First
20 Years. AMS Meteorological Monograph Vol. 57. Boston, MA: American Meteorological Society.
www.ametsoc.org/bookstore.
Appendix L: Solar
Radiation Databases
Irradiance databases are required to assess, design, and operate solar electric facilities. It is only
since the late 1990s that reliable satellite-derived databases have been developed that enable an
estimate of incident solar radiation almost any place in the world on a 0.1 degree grid or smaller.
There are a limited number of long-term high quality ground-based solar radiation monitoring
sites that exist although the list grows a little each year. These ground-based measurement station
cannot duplicate the coverage available by satellite-derived databases. Current practice is to use
satellite-derived databases to survey potential site for solar facilities and then install a ground
station to measure irradiance that represents the climatology of the area of interest. The ground-
based data are then used to validate the irradiance values obtained from satellite images. Any sys-
tematic bias found by comparing the satellite-derived database and ground-based measurements
can be used to correct long-term satellite data. The short-term ground-based measurements can
also be used in engineering studies. If the final location of the new facility is not exactly where the
ground-based monitoring station is sited, the adjustments made to the satellite-derived database
at the ground-based monitoring site can be used on the satellite-derived database at the final loca-
tion of the facility to obtain a better estimate of the facility’s production. More detailed informa-
tion on available solar radiation data bases can be found at Arguez and Vose (2011), Gueymard
and Myers (2008, 2009), Hall et al. (1978), Hulstrom (1989), Kleissl (2013), Marion and Urban
(1995), Sengupta et al. (2014, 2017), and Wilcox (2007).
It is useful to discuss characteristics of the high-quality ground-based solar radiation dataset that
exist and other long-term databases that are derived mostly from satellite images. There are many
irradiance databases and the quality of these databases depends on the reliability of the information
used in the database and the length of record of the database.

L.1 GROUND-BASED SOLAR RADIATION DATABASES


There are many ground-based irradiance databases that can be used for analyzing the performance
of a solar electric system. These databases range from high quality measured data from the Baseline
Surface Radiation Network (BSRN) (http://bsrn.awi.de/) to agricultural databases that make irradi-
ance measurements, but do not require the highest quality for their purposes. The latter will not be
discussed here.
High quality measured solar radiation data requires more than just the use of high quality instruments.
It requires recordkeeping, regular maintenance, and periodic calibrations. These requirements limit
the number of high-quality stations with long-term irradiance measurements. Research facilities and
universities are typical hosts of such stations, along with selected national weather stations. One of the
most extensive high-quality networks is the BSRN (König-Langlo et al. 2013). It consists of approxi-
mately 60 stations worldwide in a variety of climates from 80°N to 90°S. These stations provide sam-
pling of the solar resource around the world, but do not cover all the areas of interest.
The basic requirements for the BSRN station serve as a guide to what is necessary for a very
high-quality solar monitoring station (McArthur 2005). The requirements are as follows:
• Long-term involvement of an expert in surface radiation at each site (station scientist).
• Station staffed during workdays, preferably all days. Ability to create station-to-archive file.
• Data has to be quality checked and submitted in regular intervals to the World Radiation
Monitoring Center (WRMC) within 1 year (https://bsrn.awi.de/).

455
456 Appendix L

• Data/Instruments required: Measurement of direct (DNI), diffuse (DfHI), global—short


wavelength (GHI) and downward long wavelength (LWD) radiation, air temperature, rela-
tive humidity, and air pressure.
• For DNI, DfHI, GHI and LWD, a sampling frequency of ~1 Hz is required; data submis-
sion of one minute-averages required.
• Instrument accuracies of around ±5 W/m² or 3 percent (whichever is greater) for GHI and DfHI;
±2 W/m² or 3 percent for DIR Normal and ±5 W/m2 or 2 percent for LWD is highly desirable.
• Frequent cleaning/maintenance/check of instruments (e.g., daily cleaning of domes during
workdays).
• Ventilation of all radiation instruments, especially at sites with large temperature shifts is
highly recommended!
• Calibration of instruments traceable to the World Radiation Center (WRC) (www.
pmodwrc.ch) standards is required.
• Site must allow reasonable access for frequent inspections and instrument service, reliable
electrical power.
• The  horizon has to be described and submitted to the WRMC (no obstructions above
5 degrees elevation).

These requirements are very stringent and most solar monitoring stations do not  meet  all these
requirements. These requirements do illustrate information that should be available if the data for
analysis of a location’s solar potential is going to be used. For example, regular analysis of the data,
good recordkeeping, calibration of instruments traceable to international standards, and frequent
maintenance are all essential to have a high degree of confidence in the irradiance. Regular analysis
is important because problems can arise with sensors and/or equipment that can fail and checking
the data can spot problems quickly and minimize potential gaps in the data.
SURFRAD/SOLRAD stations, University of Oregon SRML network, and ARM are a few net-
works with available solar data in the United States. Some countries from Switzerland to Saudi
Arabia and Mexico also have high quality irradiance networks or are setting up such facilities.
Certain research facilities and national laboratories also may have high-quality long-term solar
irradiance data sets (NREL, CIEMAT, SNL, DLR, and others).1 In addition, many stations world-
wide send their data to the World Meteorological Organization (WMO) site (World Radiation Data
Center—http://wrdc.mgo.rssi.ru/), but it is necessary to evaluate how these data were obtained
before judging the quality and appropriateness of these stations for evaluating solar facilities.
The usefulness of a dataset is determined by several factors:

• Location. Proximity to the site of interest. For satellite derived data sets, resolution can be
a small as a few hundred meters to as large as 1 degree or 2.5 degree grids.
• Period of record. Length of record and if the record covers the time period of interest.
• Types of measurements. GHI, DNI, and DHI measurements make up a complete set of
solar radiation components. However, auxiliary measurements such as ambient temperate
and wind speed can be necessary if the data are used to predict the performance of a solar
energy system. Measurement on tilted surfaces may also be included in the dataset.
• Type of instruments used to measure the data. Different instruments provide different lev-
els of accuracy and the systematic biases of each instrument is different from other similar
instruments. Knowledge of the instrument performance specifications will help determine
the reliability of the analysis made using the data gathered by the instrument.

1 National Renewable Energy Laboratory: https://midcdmz.nrel.gov/


Centro de Investigaciones Energéticas, Medioambientales y Tecnológicas: http://www.ciemat.es/
Sandia National Laboratories: https://energy.sandia.gov/category/solar-renewable/
Deutsche Zentrum für Raumfahrt: https://www.dlr.de/sf/en/desktopdefault.aspx/tabid-10224/17488_read-44933/.
Appendix L 457

• Available records. Maintenance records and calibration records help determine the accu-
racy of the data in the dataset. Sometimes this information is available with the dataset,
but this is not often the case.
• The uncertainty of the data. The uncertainty of the dataset should be specified. This infor-
mation is rarely available, but is becoming more of a necessity as the reliability of the
results generated using the data need to be described.
• Temporal resolution. Data intervals can vary from 1-minute averaged data to daily average
values. More recent 1-minute datasets are available and there are even some datasets with
1-second samples. Short interval data can always be summed to hourly daily values.
• Availability. Some datasets are free to download while others are available for a fee.
• Other considerations. A variety of time stamps are used ranging from local time to UTC.
Units also vary. This  variance is particularly true for short time interval dataset versus
hourly datasets. Well-structured data files will contain this information. It is also important
to specify if the time stamp is for the beginning or end of the sampling period or, perhaps,
the middle of the interval.
• How the data are generated. Are the data modeled from ground observations, estimated
using satellite images, or measured with ground-based instruments? The tables listing a
variety of datasets are separated into measured, modeled, and satellite-derived datasets
and some of the datasets have a mixture of all three. Eventually several of the datasets may
utilize all three sources of information, like the early typical meteorological year (TMY)
datasets, to provide more complete coverage.

A list of ground-based solar radiation datasets is given in Table L.1. The table lists the datasets that
are still active with the datasets with the longest record listed first. For sites that are no longer active,
the sites with the earliest year are listed first.

L.2 TYPICAL METEOROLOGICAL YEAR AND OTHER MODELED DATABASES


Typical Meteorological Year (TMY) databases consists of hourly irradiance and meteorological
values determined to be “typical” by month from long-term records and assembled to form a year
of data. The TMY databases were originally designed for modeling energy use for buildings and
have been used extensively to estimate the performance of a wide variety of solar systems. Before
discussing the strengths and limitations of the TMY databases, it is important to briefly describe
how they are constructed.
TMY datasets provide a serially complete year of hourly irradiance and meteorological data for
input into solar performance models (Habte et al. 2014). TMY datasets have no gaps and this allows
models to ignore complexities associated with missing or faulty data. TMY datasets are constructed
from historical records, typically 15 years or longer. Monthly average statistics such as average GHI
irradiance and temperature are determined and an algorithm using weighted statistics for up to 10
parameters identifies Typical Meteorological Months (TMM) that are the best representative of the
long-term dataset. This is done for all 12 months and the most typical months are concatenated to
create the TMY data file (Wilcox and Marion 2008).
There are a number of ways the typical month can be determined. This weighting of the aver-
age parameters has changed over the years and the importance of each parameter can be changed
depending on what is decided to be the most important indicator of a typical month. The more recent
TMY datasets give the DNI average about a 50 percent weight with the rest of the weighting divided
between average temperature, wind speed, and other parameters. Extreme years, such as those that
were affected by volcanic eruptions are not used in the analysis as these years are not deemed typical.
The advantages of the TMY datasets is that they provide a complete year of data representative of
a multi-year period of record without any gaps. The file structure is simple and can be input into a
number of system analysis programs.
458

TABLE L.1
Selection of Databases with Measured Solar Radiation Data
Period of Temporal Resolution Spatial Coverage Spatial Resolution Data Element and
Database Record Sources Website
WMO WRDC 1964–present Majority daily, some Global More than 1,000 Primarily daily total http://wrdc-mgo.nrel.gov
hourly measurement stations GHI, but some DHI and http://wrdc.mgo.
and DNI rssi.ru
University of Oregon 1977–present Hourly to 1 minute varies Idaho, Montana, 24 Active 11 historical GHI, DNI, DHI, and http://solardat.uoregon.
SRML by station and year Oregon, Utah, measurement stations other measurements edu/SolarData.html
Washington, and
Wyoming
NREL’s SRRL Data Center 1981–present 5 minutes to 1999, one Golden, Colorado One measurement station GHI, DNI, DHI plus http://www.nrel.gov/
(MIDC) minute since 1999 many other midc/srrl_bms/
measurements
BSRN 1992–present 1-minute Global 52 measurement stations GHI, DNI, DHI, http://www.bsrn.awi.de/
in operation; 7 additional downwelling en/home
stations (now closed) infrared irradiance,
with historical data upwelling infrared
irradiance, and more
NOAA Network—Surface 1993–present 3 minute before 2009, one Continental United Seven stations GHI, DNI, DHI, ftp://ftp.srrb.noaa.gov/
Radiation Network minute since 2009 States downwelling pub/data/surfrad www.
(SURFRAD) infrared irradiance, srrb.noaa.gov and www.
upwelling infrared bsrn.awi.de/
irradiance, and
upwelling (reflected)
shortwave irradiance
Atmospheric Radiation 1997–present 20-second instantaneous Southern Great Plains, 23 stations (southern GHI, DNI, DHI, DIR, http://www.arm.gov and
Measurement Program samples and 1-minute North Slope of Great Plains), 2 stations UIR, and other http://www.bsrn.awi.de/
(ARM) averages of 2-second Alaska, and tropical (North Slope of Alaska),
scans western Pacific and 2 stations (tropical
western Pacific)
(Continued)
Appendix L
TABLE L.1 (Continued)
Appendix L

Selection of Databases with Measured Solar Radiation Data


Period of Temporal Resolution Spatial Coverage Spatial Resolution Data Element and
Database Record Sources Website
Western Energy Supply 1976–1980 15-minute Southwest USA 52 measurement stations GHI, DNI (at three http://rredc.nrel.gov/solar/
and Transmission stations), DHI pubs/wa/wa_index.html
Associates Solar (shadowband)
Monitoring Network
NOAA Network 1977–1980 Hourly United States and 39 measurement stations GHI, DNI, DHI (7 http://www.ncdc.noaa.
territories stations) gov/
Solar Energy and 1979–1983 1-minute Fairbanks, Alaska; GHI, DNI, and DHI; http://rredc.nrel.gov/solar/
Meteorological Research Atlanta, Georgia; GTI, infrared old_data/semrts/
Training Sites Albany, New York; San irradiances,
Antonio, Texas ultraviolet and other
spectral irradiance
(varies)
Historically Black Colleges 1985–1996 5-minute Southeastern United Six measurement stations GHI, DNI (at three http://rredc.nrel.gov/solar/
and Universities Solar States: stations), DHI old_data/hbcu/
Measurement Network (shadowband)
International Daylight 1991–1994 Variable Global 43 measurement stations GHI, DNI, DHI, http://idmp.entpe.fr/
Measurement Program zenith luminance,
(IDMP) illuminance
(including vertical
surfaces), and other
measurements
NOAA Network— 1995–2018 15-minute to 1-minute Continental United Seven stations GHI, DNI, DHI, and ftp://aftp.cmdl.noaa.gov/
SOLRAD States Global UVB data/radiation/solrad/
Solar Data Warehouse Various Hourly and daily Continental United More than 3000 GHI http://solardat.uoregon.
lengths States measurement stations edu/SolarData.html
459
460 Appendix L

A major disadvantage of the TMY methodology is that the process can produce files that have
yearly average irradiances values that are either above or below any year contained in the dataset
from which it was constructed (Vignola et al. 2012). In addition, using just one typical year, varia-
tions in the output of the solar facility are missed and extreme situations are not part of the analysis
(Ryberg et al. 2015). These extreme situations are of importance if assessing the risks and reserves
necessary when production is curtailed by a year with low solar resource availability. In addition,
any long-term trends that exist in the dataset are missed. TMY datasets usually are limited to
hourly data because long-term datasets with shorter time intervals are scarce and only recently have
satellite-derived irradiance models produced shorter time interval data.
TMY datasets, as with any dataset, are limited by the uncertainty of the data used to create the
dataset. Any problems in the long-term input dataset will show up in the TMY dataset and systematic
problems can skew the results. While TMY datasets are good for initial surveys and point to prom-
ising areas one might consider for a solar facility site, more thorough analysis is usually required for
larger facilities. A selection of datasets with modeled data are given in Table L.2.

L.3 SATELLITE-DERIVED SOLAR RADIATION DATABASES


Many companies and government funded agencies provide satellite-derived solar radiation data-
bases. Most of these databases are derived from satellite weather images that are used to deter-
mine cloudiness and auxiliary meteorological datasets to characterize the atmospheric conditions
for the area of interest. Most of these auxiliary meteorological datasets, such as aerosol optical
depth are also derived from satellite data. Typically, clear-sky irradiance is calculated using the
auxiliary meteorological datasets and satellite weather images are used to estimate cloud cover.
The cloud cover is used to adjust the clear-sky values to estimate the irradiance values. The percent-
age cloud cover is determined by comparing the irradiance reflected by clouds and that reflected by
the ground. A satellite image consists of a number of pixels that represent an area and the range of
possible brightness values is finite. The first step is to align the image with ground coordinates and
relate the stability of the satellite and its pointing accuracy (Perez et al., 2002; Vignola and Perez,
2004). The amount of sunlight reflected from the surface depends on the surface albedo and this can
change over the day and the season. Because the range of brightness is limited, the more closely the
surface reflects like clouds, the smaller the range available to determine the amount of clouds and
the increase in the uncertainty of the predicted irradiance. It is particularly difficult to determine the
irradiance when the ground is covered by snow because the albedo of a snow covered field is similar
to the albedo of the cloud and only a small range of values are available to estimate the percent cloud
cover and hence the irradiance. The  parameters of irradiance models using satellite images are
adjusted or correlated to produce models that minimize the difference between the model estimates
and the ground-based measurements.
The perspective of ground-based measurements and satellites is complimentary to each other.
Ground-based measurements look directly at the incident irradiance from one location and provide
irradiance values averaged over a minute to an hour. Satellite images scan a much larger area at
an instance in time. Therefore it is not surprising that there is a large random scatter between the
ground-based measurements and satellite-derived irradiance values. Much of the bias is removed
using a correlation process although it is not certain that the satellite model will work as well in one
area as in another. Since the cause of much of the randomness is to be expected, it is more important
to reduce the bias and reproduce the statistical characteristics of the irradiance values.
TABLE L.2
Appendix L

Databases That Contain Mostly Modeled Data


Temporal Spatial Data Element and
Database Period of Record Resolution Coverage Spatial Resolution Sources Website
SOLMET/ERSATZ December 1951–December Hourly United States 26 measurement stations GHI, DNI (estimated http://www.ncdc.noaa.gov/
1976 (hour and territories and 222 modeled stations from GHI),
ending in
LST)
SOLDAY January 1952–December 1976 Daily Continental 26 measurement stations measured GHI from http://www.ncdc.noaa.gov/
United States mechanical integrators
and strip charts
TMY2 One year representative of the Hourly United States 239 stations representing Hourly GHI, DNI, DHI https://rredc.nrel.gov/solar/
1961–1990 NSRDB data and territories the 1961–1990 NSRDB other information old_data/nsrdb/1961-1990/tmy2/
period
NSRDB 1961–1990 1961–1990 Hourly United States 239 stations (56 stations Hourly GHI, DNI, DHI https://rredc.nrel.gov/solar/
and territories have some radiation other information old_data/nsrdb/1961-1990/
measurements)
NSRDB 1991–2005 1991–2005 Hourly United States 1454 locations and 10-km GHI, DNI, and DHI ftp://ftp.ncdc.noaa.gov/pub/data/
by 10-km grid nsrdb-solar/, ftp://ftp.ncdc.noaa.
(1998–2005) gov/pub/data/nsrdb-solar/ and
ftp://ftp.ncdc.noaa.gov/pub/data/
nsrdb-solar/
TMY3 1991–2005 Hourly United States 1020 locations GHI and illuminance, https://RReDC.nrel.gov/solar/
and territories DNI and illuminance, old_data/nsrdb/1991-2005/tmy3/
DHI and illuminance,
zenith luminance.
NSRDB 1991–2010 1991–2010 Hourly United States 1454 locations and 10-km GHI, DNI, and DHI. http://www.nrel.gov/docs/
by 10-km grid Measured or observed fy12osti/54824.pdf and https://
(1998–2010) data: RReDC.nrel.gov/solar/old_data/
nsrdb/1991-2010/
461
462 Appendix L

The National Solar Radiation Data Base (NSRDB) has undergone significant changes since it
was first introduced and is typical of how satellite-derived solar databases have evolved. The first
version of the NSRDB from 1961 to 1990 used some measured irradiance data but mainly relied
on ground-based observations of cloud cover to estimate hourly GHI, DNI, and DfHI irradiance
values for 237 national weather stations in the USA, Guam, and Puerto Rico. The  next version
contains data from 1,454 stations in the USA and its territories. This time all the irradiance data
were either modeled from meteorological observations or from satellite images. The most current
version NSRDB (v2.0.0) was developed using the Physical Solar Model (PSM), and offers users the
latest available data (1998–2016). This dataset is composed of 30-minute solar and meteorological
data for approximately 2  million 0.038-degree latitude by 0.038-degree longitude surface pixels
(~4 square kilometers). The area covered is bordered by longitudes 25°W on the east and 175°W on
the west, and by latitudes −20°S and 60°N. Development of the NSRDB and a description of the
current version is available from Sengupta et al. (2018).
This  dataset is not  designed to replace ground-based measurements for engineering due-dili-
gence or real-time operations but to provide a long-term view of the solar resource trends. In gen-
eral, ground-based measurements can be used to validate or adjust the satellite-derived data for
any systematic problems. Since a solar facility may not be located exactly where the ground-based
measurements are taken, any adjustments to the satellite-derived irradiance data that are necessary
to remove any systematic effects can be transferred to the satellite-derived dataset at the exact loca-
tion where the solar facility is being considered.
Among the sources listed in Table L.3 is another prominent satellite-derived solar radiation
database covering Europe and Africa from latitudes 66°N to 66°S and from longitudes 66°W to
66°E. The spatial resolution varies from 3 to 5 kilometers (km) depending on the distance from the
satellite nadir. The Helioclim data are from the MeteoSat satellites and run from 2004 to present.
The time resolution is 15 minutes and the data are updated in real time.
A good understanding of the uncertainties of the ground-based irradiance measurements and the
satellite-derived data is necessary so that systematic effects in the ground-based measurements are
not incorporated into the adjusted satellite-derived dataset (Table L.3).
TABLE L.3
Selection of Satellite-Derived Solar Radiation Databases
Appendix L

Temporal Data Element and


Database Period of Record Resolution Spatial Coverage Spatial Resolution Sources Website
NASA Modern-Era 1980–present Hourly Global 0.5° × 0.625° GHI http://eosweb.larc.nasa.gov/sse/
Retrospective analysis for
Research and Applications,
Version 2 (MERRA-2)
PVGIS 1981–present Hourly Europe, Africa, 1-km aggregated to GHI, DNI, DHI, http://re.jrc.ec.europa.eu/
Asia (partial) 5 arc-minutes (~8 km) and GTI pvg_tools/en/tools.html
SOLEMI 1991–present 30-minute Europe, Africa, 2.5 km GHI, DNI http://www.dlr.de/tt/en/
South America, desktopdefault.aspx/
Western Asia, tabid-2885/4422_read-6581/
Western Australia
SolarGIS 1994, 1999, 15- and 30-minute Land area, ~3 km (at the equator) DNI, GHI, DHI, https://solargis.info
2007–present (depends worldwide, down scaled to GTI
on region) between latitudes ~80 m using
60°N and 50°S SRTM-3 DEM
IrSoLaV 1994–present (Africa, 15-minute (except Europe, Africa, ≈3 km GHI, DNI, DHI, http://www.irsolav.com/en-gb/
Europe) 2000–present Americas), Americas, GTI home
(Americas, western 30-minute, western Asia
Asia) hourly, daily,
monthly
EnMetSol 1995–present Meteosat 30-minute for Continental 2.5 km for MFG; GHI, DHI, and https://uol.de/
first and second MFG; 15-minute Europe, Canary 1 km for MSG DNI energiemeteorologie/
generation for MSG Islands, Turkey, forschung/solarenergie/
and Israel
Green Power Labs: 1995–present 30-minute Americas, Asia, 1–4 km GHI, DNI, DHI, https://greenpowerlabs.com/
SolarSatDataTM (Americas) Australia, Europe GTI
2000–present (Europe)
2000–present (Europe)
(Continued)
463
464

TABLE L.3 (Continued)


Selection of Satellite-Derived Solar Radiation Databases
Temporal Data Element and
Database Period of Record Resolution Spatial Coverage Spatial Resolution Sources Website
3TIER (VAISALA) Solar January 1997–Present ~30-minute Global 2 arc-min (~3 km) GHI, DNI, and https://www.3tier.com/
Time Series instantaneous DHI products/
and 1-hour catalog/?tab=solar-catalog
averages
Clean Power 1998–present Hourly Continental United 10 km GHI, DNI https://www.solaranywhere.
Research—SolarAnywhere States and Hawaii com/
OSI-LAF 2001–present Hourly Africa, Americas, 0.05° × 0.05° GHI, Longwave http://www.osi-saf.
Europe, western infrared irradiance org/?q=content/
Asia radiative-fluxes-products
CAMS McClear Service 2004–present 1 minute, Global Various input data clear sky global, http://solar.atmosphere.
15 minutes, sources with direct, direct copernicus.eu/cams-mcclear
1 hour, 1 day, different spatial normal, diffuse
1 month, resolutions are irradiances
interpolated to the
location of interest
CAMS Radiation Service 2004–present 1 minute, Europe/Africa/ Various input data all-sky GHI, DNI, http://solar.atmosphere.
15 minutes, Middle East/ sources with DIR, DHI, and copernicus.eu/
1 hour, 1 day, Atlantic Ocean different spatial corresponding cams-radiation-service
1 month, resolutions are clear-sky
interpolated to the irradiances
location of interest
HelioClim v2-v5 2004–present 15-minute Europe and Africa 5 km Hourly and daily http://www.soda-pro.com/. See
GHI also: www.soda-is.com/eng/
index.html
CM-SAF Operational 2010–present Daily Europe, Africa 15 km GHI and the https://wui.cmsaf.eu
Products horizontal
projection of DNI
(Continued)
Appendix L
TABLE L.3 (Continued)
Selection of Satellite-Derived Solar Radiation Databases
Appendix L

Temporal Data Element and


Database Period of Record Resolution Spatial Coverage Spatial Resolution Sources Website
CM SAF cLoud, Albedo and 1982–2015 Monthly averages Global 0.25° × 0.25° Cloud properties, http://www.cmsaf.eu/
surface RAdiation data set surface albedo, and
from AVHRR data—Edition surface radiation
2 (CLARA-A2) parameters
NSRDB 1998–2016 1998–2016 half-hourly longitude: −25°E 4 km GHI, DNI, DHI, https://nsrdb.nrel.gov
to −175°W, clear-sky DHI and
latitude: −20°S to https://doi.org/10.1016/
60°N j.rser.2018.03.003
TMY 98-15 1998–2015 Hourly longitude: ~25°E to 4 km GHI, DNI, DHI https://nsrdb.nrel.gov
~175°W, latitude:
−20°S to 60°N
CERES SYN1deg 2000–2016 3-hourly Global 1° × 1° GHI, DHI, and http://solar.atmosphere.
DNI copernicus.eu/
cams-radiation-service
National Center for 1948–2009 6-hour (W/m2) Global 2.5° (nominal) GHI http://rda.ucar.edu/datasets/
Environmental Protection/ ds090.0/
National Center for
Atmospheric Research
Global Reanalysis Products
(NCEP)
DAYMET 1980–1997 Daily Continental United 1 km GHI http://daymet.ornl.gov/
States
ESRA 1981–1990 Monthly and Europe 10 km GHI, DNI, and http://www.mines-paristech.fr/
annual average DHI Ecole/Culture-scientifique/
daily totals Presses-des-mines/#54. See
(kWh/m2) also http://www.soda-is.com/
eng/index.html
(Continued)
465
466

TABLE L.3 (Continued)


Selection of Satellite-Derived Solar Radiation Databases
Temporal Data Element and
Database Period of Record Resolution Spatial Coverage Spatial Resolution Sources Website
NASA Surface Meteorology July 1983–June 2005 Monthly and Global 1 × 1° GHI, DNI, and http://eosweb.larc.nasa.gov/sse/
and Solar Energy annual average DHI
daily totals
(kWh/m2)
NASA Prediction Of July 1983–June 2005 Daily and Global 0.5° × 0.5° Clearness Indices, https://power.larc.nasa.gov/
Worldwide Energy Monthly Cloud Cover, GHI
Resources (POWER) and GTI
DLR ISIS July 1983–December 3-hour Global 280 km by 280 km DNI and GHI from http://www.pa.op.dlr.de/ISIS/
2004 a radiative transfer
model
Solar and Wind Energy Moderate resolution: Monthly and South America, Moderate resolution: GHI, DNI (DHI), https://openei.org/wiki/
Resource Assessment 1985–1991; high annual average Central America, 40 km; high and GTI Solar_and_Wind_Energy_
(SWERA) resolution: 1998–2002 daily totals Africa, South and resolution: 10 km Resource_
(kWh/m2) East Asia, Middle DLR are available Assessment_(SWERA) http://
East from http://www.dlr. www.inpe.br/ingles/index.php
de/tt/desktopdefault.
aspx/
tabid-2885/4422_
read-6548/
Management and Exploitation 1991–2005: Europe and Hourly Europe, Western 2.5 km GHI, DNI, DHI www.mesor.org/
of Solar Resource Africa; 1999–2006: Asia, Africa, parts from ground
Knowledge (MESoR) Asia of Australia, measurements and
South America modeling results
(Continued)
Appendix L
Appendix L

TABLE L.3 (Continued)


Selection of Satellite-Derived Solar Radiation Databases
Temporal Data Element and
Database Period of Record Resolution Spatial Coverage Spatial Resolution Sources Website
METEONORM 1991–2010 1-minute and Global Data from 8351 Measured: monthly https://www.meteonorm.com/
hourly modeled meteorological means of GHI,
data stations are Modeled:
interpolated to 1-minute and
establish weather data hourly typical
at any specified point. years radiation
Satellite: 2 km for parameters (GHI,
Europe, 8 km for rest DNI, DHI, GTI)
of the world
Satel-Light 1996–2000 30-minute Europe ~5 km DNI, GHI, DHI, http://www.satellight.com/
GTI, horizontal indexgS.htm
illuminance, tilted
illuminance, and
sky luminance
distribution.
467
468 Appendix L

REFERENCES
Arguez, A., and R. S. Vose. 2011. The definition of the standard WMO climate normal: The key to deriving
alternative climate normals. Bulletin of the American Meteorological Society 92:699–704.
Gueymard, C. A., and D. R. Myers. 2008. Solar radiation measurement: Progress in radiometry for improved
modeling. In Modeling Solar Radiation at the Earth Surface, ed, V. Badescu, 1–27. Berlin, Germany:
Springer.
Gueymard, C. A., and D. R. Myers. 2009. Solar resource for space and terrestrial applications. In Solar Cells
and their Applications, ed. L. Fraas and L. Partain, 427–461. Hoboken, NJ: John Wiley & Sons.
Habte, A., A. Lopez, M. Sengupta, and S. Wilcox. 2014. Temporal and Spatial Comparison of Gridded TMY,
TDY, and TGY Data Sets. NREL/TP-5D00-60886. 30 pp. https://www.nrel.gov/docs/fy14osti/60886.
pdf (Accessed June 4, 2019).
Hall, I., R. Prairie, H. Anderson, and E. Boes. 1978. Generation of Typical Meteorological Years for
26 SOLMET Stations. SAND78-1601. Albuquerque, NM: Sandia National Laboratories.
Hulstrom, R. L., ed. 1989. Solar Resources. Cambridge, MA: MIT Press.
Kleissl, J. (Ed). 2013. Solar Energy Forecasting and Resource Assessment, 503 pp. Oxford, UK: Elsevier.
König-Langlo, G., R. Sieger, H. Schmithüsen, A. Bücker, F. Richter, and E.G. Dutton. 2013. The baseline
durface radiation network and its World Radiation Monitoring Centre at the Alfred Wegener Institute.
World Meteorological Organization, Geneva, Switzerland: GCOS-174, WCRP-24/2013. https://epic.
awi.de/id/eprint/34323/1/gcos-174.pdf (Accessed June 3, 2019).
Marion, W., and K. Urban. 1995. User’s Manual for TMY2s-Typical Meteorological Years Derived from the
1961–1990 National Solar Radiation Database. NREL/TP-463-7668. Golden, CO: National Renewable
Energy Laboratory.
McArthur, L. J. B. 2005. Baseline Surface Radiation Network (BSRN) Operations Manual Version 2.1. World
Meteorological Organization, Geneva, Switzerland: WCRP-121, WMO/TD-No. 1274. https://bsrn.awi.
de/other/manuals/ (Accessed June 3, 2019).
Perez, R., P. Ineichen, K. Moore, M. Kmiecik, C. Chain, R. George, and F. Vignola. 2002. A new operational
satellite-to-irradiance model. Solar Energy 73(5):307–317.
Ryberg, D. S., J. Freeman, and N. Blair. 2015. Quantifying Interannual Variability for Photovoltaic Systems
in PVWatts. NREL/TP-6A20-64880. https://www.nrel.gov/docs/fy16osti/64880.pdf. (Accessed June
4, 2019).
Sengupta, M., A. Habte, P. Gotseff, A. Weekley, A. Lopez, M. Anderberg, C. Molling, and A. Heidinger. 2014.
Physics-based GOES product for use in NREL’s National Solar Radiation Database. Preprint. Prepared
for the European Photovoltaic Solar Energy Conference and Exhibition, September 22–26. NREL/
CP-5D00-62776. Golden, CO: National Renewable Energy Laboratory.
Sengupta, M., A. Habte, C. Gueymard, S. Wilbert, D. Renné, and T. Stoffel. 2017. Best Practices Handbook
for the Collection and Use of Solar Resource Data for Solar Energy Applications, 2nd ed. NREL/
TP-5D00-68886. Golden, CO: National Renewable Energy Laboratory.
Sengupta, M., Y. Xie, A. Lopez, A. Habte, G. Maclaurin, and J. Shelby. 2018. The National Solar Radiation Data
Base (NSRDB). Renewable and Sustainable Energy Reviews 89:51–60. doi:10.1016/j.rser.2018.03.003.
Vignola, F., and R. Perez. 2004. Solar Resource GIS Data Base for the Pacific Northwest using Satellite Data,
Final Report. Eugene, OR: University of Oregon. http://solardat.uoregon.edu/download/Papers/Solar
ResourceGISDataBaseforthePacificNorthwestusingSatelliteData--FinalReport.pdf (Accessed June 4,
2019).
Vignola, F., C. Grover, N. Lemon, and A. McMahan. 2012. Building a bankable solar radiation dataset. Solar
Energy 86:2218–2229.
Wilcox, S. 2007. National Solar Radiation Database 1991–2005  Update: User’s Manual. NREL/TP-581-
41364. Golden, CO: National Renewable Energy Laboratory.
Wilcox, S. and W. Marion. 2008. Development of an updated typical meteorological year data set for the
United States. Proceedings of the Solar 2008 Conference, May 3-8, 2008, San Diego, CA (CD-ROM),
6 pp. Boulder, CO: American Solar Energy Society (ASES).
Answers to Chapter Questions

CHAPTER 1
1. Solar radiation data are used for these and other applications:
a. Agriculture—evaluating irrigation requirements
b. Atmospheric Science—understanding the Earth’s energy balance
c. Buildings—designing structures with natural daylighting and space heating
d. Materials Testing—evaluating the weathering effects of solar radiation, especially in
the ultraviolet region, on various materials
e. Renewable Energy—generating electricity and heat with photovoltaic (PV) and solar
thermal systems
f. Solar Forecasting—developing operational forecasts to improve renewable energy sys-
tem integration into the utility grid.
2. Broadband solar radiation refers to the total energy available over a wide range of wave-
lengths. Solar radiation incident on the Earth’s surface is composed of energy at different
wavelengths that interact with the changing atmosphere in complex ways that affect the rel-
ative amounts of ultraviolet, visible, and near-infrared (IR) radiation reaching the surface.
3. This personal statement could be based on the importance of solar measurements to any of
the needs listed above or to a new research topic yet to be discovered.
4. Depending on the specific financial risks associated with the scope of operations, the
banker is more likely to want more accurate solar data than the farmer before making the
long-term financial commitments needed for a successful solar concentrating power plant.
Specifically, the banker will need an accurate assessment of the direct normal irradiance
(DNI) while the farmer can use the global horizontal irradiance (GHI) information.
Accurate measurements of DNI are more demanding than the relative ease of making GHI
measurements.
5. Chapter  2 provides the terminology and basic equations  used in solar and IR radiation
measurements, including solar position. Appendix E discusses the algorithms used to
calculate the position of the Sun.

CHAPTER 2
1. According to NASA  (2018), the effective temperature of the Sun is 5772  K (5499°C or
9930°F).
2. The approximate solar declination for September 23 can be computed from Equation 2.3
using:

Day of year = 266

Day angle, β = 2π(266−1)/365 radians

= 4.561765
Declination, δ = 0.25° (Excel result is 0.24878131)
Note: The date in question is very near the Autumnal Equinox resulting in a declination angle
close to zero.
3. The currently accepted value of the Total Solar Irradiance (TSI) is 1360.8 ± 0.5 Wm−2.

469
470 Answers to Chapter Questions

4. In the Northern Hemisphere, the minimum solar zenith angle (maximum solar elevation
angle) occurs at solar noon on the summer solstice and is a function of the observer’s latitude.
The observer uses the Sun Path Charts in Appendix D to find the chart with the nearest
latitude to identify the maximum solar elevation near June 21 or uses the Solar Position
Calculator for a more accurate value (http://solardat.uoregon.edu/SolarPositionCalculator.
html). In Golden, Colorado (39.74 N, 105.18 W, 1829 m AMSL) the minimum solar zenith
angle is 16.31° at solar noon near the summer solstice.
5. First, compute the local standard time in Greenwich (GMT) at solar noon.
Solving for local standard time in Equation 2.8:

lst = st−EoT− (long sm−longob)/15

In Greenwich, the standard meridian (longsm) is the longitude of the observer (longob).
Therefore,

lst = st−EoT

Using Equation 2.4, the Equation of Time (EoT) on March 21 (DOY = 80) is −7.858 minutes.


At solar noon, solar time (st) = 12:00.
Therefore,

GMT = 12:00−(−7.858 minutes)

       = 12:08 to the nearest minute.

For a location in the Mountain Standard Time (MST) zone:

MST = GMT−7 hours
Therefore, solar noon in Greenwich is 05:08 MST
6. Daylight time adds one hour to local standard time. Therefore, 12:00 p.m. CDT is 11:00 a.m.
CST.
7. The geometric solar zenith angle at sunrise is 90° for any day of the year at any location.
8. From Equation 2.12:
Global horizontal irradiance (GHI) is the sum of the diffuse horizontal irradiance (DHI)
and the Direct normal irradiance (DNI) projected on to a horizontal surface:

GHI = DNI × cos (sza) + DHI

9. Solving Equation 2.12 for DNI:

DNI = (GHI−DHI)/cos(sza)

10. Using equations 2.4, 2.7, 2.8, 2.9, 2.10, and 2.14 and these given quantities:
Observer Latitude = 40° N
Observer Longitude = 105° W (also the standard meridian for the mountain time zone)
Local Standard Time (lst) = 14:00
Date = June 21st (DOY = 172)
Surface tilt angle above horizontal (T) = 30°
Surface azimuth angle from true north (γ) = 270°
Calculations:
Equation of Time (EoT) = −1.33 minutes on DOY = 172
Solar Time (st) = local standard time + EoT = 14:00−00:01 = 13:59 (nearest minute)
Hour Angle (ha) = 15°/hr × (st−12 hr) = 15 × (13.98−12.00)° = 29.75°
Answers to Chapter Questions 471

Equation 2.7: Cosine of the solar zenith angle [cos(sza)] = 0.8659568


Equation 2.10: Sine of the solar azimuth angle [sin(az)] = −0.99485
Equation 2.14: Cosine of the incidence angle to a tilted surface [cos(θT )] = 0.9775532
11. Using Equation 2.17 and assuming room temperature is 25°C (77°F) or (298.15 K):

E = 1.0 × 5.670367E-8 Wm−2K−4 × (298.15 K)4

   = 448.1 Wm−2

12. In thermal equilibrium, a perfect blackbody emits and absorbs energy in the same amounts,
so for a room temperature of 25°C, a blackbody absorbs 448.1 Wm−2.
13. See Figure 2.1 for extraterrestrial solar irradiance as a function of wavelength. The peak
irradiance of about 2.1 Wm−2nm−1 occurs near 450 nm.
14. Ground-based measurements of solar and atmospheric radiation are historically limited
in comparison with other meteorological parameters due to the relatively higher costs
of acquisition, installation and maintenance. Additionally, the site-specific nature of
ground-based measurements presents challenges for extrapolating the data to larger areas.
Estimating solar radiation resources from available meteorological data and satellite-based
observations can produce data with acceptable limits of estimated uncertainty. It should
be noted that model performance is limited to the quality of the measurements used to
develop and validate the model in question.

CHAPTER 3
1. In 1837, the French physicist Claude Servais Mathias Pouillet first used the term pyrhe-
liometer to name his invention for measuring the solar constant (Crova 1876), now more
accurately termed the total solar irradiance (TSI).
2. W. J. H. Moll, a physicist at the University of Utrecht, designed a thermopile detector used
by Ladislas Gorezynski at the Polish Meteorological Institute to measure solar irradiance.
Available by the mid-1920s, the Moll-Georezynski pyranometer was often called a solar-
imeter and addressed the measurement performance characteristics still sought in modern
pyranometers and pyrheliometers (Moll 1923).
3. In 1979, based on the results of electrically self-calibrating absolute cavity radiometers at
the Fourth International Pyrheliometer Comparisons (IPC-IV) in Davos, Switzerland, the
WMO adopted the WRR as the replacement for the existing International Pyrheliometric
Scale of 1956 for the measurement of direct normal irradiance (Zerlaut 1989).
4. Based on measurements from the total irradiance monitor (TIM) as part of the SORCE
launched in January  2003, the presently accepted value of the total solar irradiance is
1360.8 ± 0.5 Wm−2 (Kopp and Lean 2011).
5. Spectral response. Because a silicon photodiode is less responsive near the 400 nm spec-
tral region (i.e., blue), errors of up to 30 percent can occur for measurements of clear-sky
diffuse irradiance (Vignola 1999; Wilbert et al. 2016).
6. A pyrgeometer is designed to measure infrared irradiance (Dorno 1923; Gröbner and Los
2007).
7. In  1901, the U.S. Weather Bureau began the first measurements of solar radiation by
deploying three Ångström’s electric compensation pyrheliometers in Washington, DC,
Baltimore, Maryland, and Providence, Rhode Island, to measure direct normal irradiance
(Coulson 1975).
8. At cryogenic temperatures, the copper cavity sensors used in the CSAR become thermally
super-conductive, limiting the errors introduced by the electrical substitution principle
(Walter 2017).
472 Answers to Chapter Questions

CHAPTER 4
1. Equation 4.1. Approximately 25.84°.
2. Scattering and/or absorption by atmospheric aerosols, water vapor, ozone, clouds and other
atmospheric gases (e.g., oxygen and carbon dioxide).
3. The distance between the sun and earth changes by ±1.5 percent because the Earth is in an
elliptical orbit around the Sun. This leads to ±3 percent change over the year because the
intensity of solar radiation at Earth is inversely proportional to the square of the distance.
Approximately 0.9904.
4. 3, 1, 2.
5. Pyranometer angular (cosine) response, spectral response, or temperature response.
6. By maintaining calibration traceability with side-by-side measurements with a cavity that
has been calibrated at an IPC or NPC to obtain a valid WRR transfer factor.
7. Michalsky, J., E. G. Dutton, D. Nelson, J. Wendell, S. Wilcox, A. Andreas, P. Gotseff, D.
Myers, I. Reda, T. Stoffel, K. Behrens, T. Carlund, W. Finsterle, and D. Halliwell. 2011.
An  extensive comparison of commercial pyrheliometers under a wide range of routine
observing conditions. Journal of Atmospheric and Oceanic Technology 28:752−766.
8. No, yes.
9. From https://www.esrl.noaa.gov/gmd/grad/neubrew/SatO3DataTimeSeries.jsp  one  finds:
316 Dobson units (DUs); NASA’s OMI provides daily averaged ozone data on a 0.25 × 0.25
degree grid.
10. As with any measurement, an estimate of the measurement uncertainty.

CHAPTER 5
1. The cosine of 45 degrees is 0.707. The DNI on a horizontal surface is 1,000 Wm−2 times
0.707  or 707  Wm−2. The  GHI is the DNI on a horizontal surface plus the diffuse or
807 Wm−2.
2. The irradiance received at a surface is proportional to the cosine of the incident angle if the
surface has a Lambertian response.
3. A secondary standard pyranometer is more precise than a first-class pyranometer. New ISO
9060 pyranometer classification terminology now calls a secondary standard pyranometer
Class A and a first-class pyranometer Class B. Class A pyranometers have a better cosine
response, linearity, and are more stable over a wider range of temperatures than a Class B
pyranometer.
4. A black disk pyranometer has a thermopile with one side connected to the black disk detec-
tor that is exposed to the sunlight (hot junction) and the other side connected to a heat sink
(cold junction). The difference in temperature between the two junctions causes a voltage
signal that is proportional to the temperature difference created by the incident solar radia-
tion on the detector.
5. Pyranometers with two glass domes reduce convective and advective heat losses from the
receiver.
6. Thermal offset is related to the infrared radiation emitted from the pyranometer to the sky.
The black disk radiates to the inner dome, the inner dome radiates to the outer dome, and
the outer dome radiates to the sky. This reduces the temperature difference between the
black disk detector and the instrument body heat sink. The process occurs at night as well
as during the day. The thermal offset can be measured at night when incident solar radia-
tion is zero. This information along with knowledge of the sky temperature or net infrared
irradiance can be used to estimate the pyranometer’s thermal offset during the day.
Answers to Chapter Questions 473

7. A  ventilator is a motorized fan and enclosure that blows air across the surface of the
pyranometer dome. This inhibits dust, moisture, and ice from accumulating on the dome
and, when the air is artificially heated, reduces the pyranometer’s thermal offset.
8. Non-linearity is the deviation from the sensitivity at 500 Wm−2 over 100−1,000 Wm−2 range.
The deviation of the responsivity at 500 nm is ±0.05 µV/Wm−2  at 100 and 1,000 nm or
0.625 percent.
9. The responsivity of a pyranometer can decrease from year to year due to changes in the
material characteristics even when the instrument is kept in storage although the rate of
change in storage can be much less. For  field instruments, much of the degradation in
absorption, and hence responsivity, is related to changes caused by UV radiation. By cali-
brating the pyranometer each year, the change in responsivity can be measured and appro-
priate action can be taken to ensure the relative irradiance measurements from year to
year are the same. With modern sensor coatings most manufacturers now  recommend
pyranometer recalibration after two years.
10. Advective heat loss is when air blows against a surface and transfers heat to or from the
surface to the surrounding air.
11. A  black surface absorbs more radiation than a white surface because of the differing
absorptance characteristics. A pyranometer design based on a “black-and-white” thermo-
pile detector exposes the blackened “hot” junction and white “cold” junction to solar radia-
tion. As with a single-black thermopile detector, the radiative flux is proportional to the
voltage produced by the temperature difference between the two junctions. This design
results in almost no thermal offsets because in the infrared range greater than 3,000 nm,
the absorptivity and emissivity of the black surface is about the same as the white surface.
The relatively large detector size presents challenges for the angular response due to the
difficulty of generating a completely planar detector surface and the instrument’s long time
response due to the large thermal mass.
12. The absorptivity is the same as the emissivity or 95 percent.
13. Yes, the surface reflectivity or absorptivity is dependent on the wavelength of the radiation.
14. The absorptivity (emissivity) of the white surface (cold junction) at wavelengths of 3000 nm
and greater is approximately the same as the absorptivity (emissivity) of the black sur-
face (hot junction). Therefore, they both radiate about the same to the sky at wavelengths
3,000 nm and greater and hence the relative temperature difference between them is inde-
pendent of this long wavelength radiation exchange with the sky.
15. Photodiode-based pyranometers are designed to measure the available solar irradiance and
are based on the fact that the short-circuit current of a solar cell is proportional to the
incident radiation. The angular or cosine response of a PV module is significantly affected
by the glazing and photodiode-based pyranometers are designed to have a fairly accurate
angular response. The  photodiode-based pyranometer measures the short-circuit current
that has a significantly different temperature dependence than a PV module that operates at
the maximum power point (MPP). The PV module’s MPP is dependent on several factors
other than the short circuit current.
The spectral response of a photodiode-based pyranometer is different than a PV module
because it has a diffuser that limits the range of wavelengths seen by the detector, and the
photodiode can have a different spectral responsivity than the PV module.
16. No. The  solar cell in a silicon photodiode requires a photon with sufficient energy
to separate the electron from the atom. In  physics terms, that energy has to be enough
to excite the electron into the conduction band. This limits the response of photodiodes to
wavelengths less than around 1100 nm. Shorter wavelength photons have higher energy.
474 Answers to Chapter Questions

17. The  structure of the diffuser and the rim of the pyranometer reduce the responsivity
of the pyranometer at angles greater than 82°. This construction is designed to give the
pyranometer a better overall response at large angles and results in the cat ears affect
(i.e., a relatively predictable change in responsivity with incidence angles greater than 75°).
Without these features, unreasonably high measurements (percentage wise) would result at
the extreme angles of incidence.
18. From Figure 5.18, there should be no change in the responsivity because it is specified at
25°C. At 0°C, the responsivity decreases by 0.1 percent from the reference value at 25°C.
19. Under clear skies, the Rayleigh scattering is the predominant source of DHI. The magnitude
of the Rayleigh scattering is proportional to λ−4. This means that the short wavelengths are
preferentially scattered from the DNI irradiance much of which ends up as DHI. Therefore,
the DHI irradiance distribution peaks in the short-wavelength (blue) portion of the spectrum
and the DNI spectrum peaks in the mid visible range with a considerable near-infrared
component. Since the spectral responsivity of the photodiode pyranometer is skewed to the
near-infrared, and this pyranometer is calibrated using the full solar spectrum, the average
DHI responsivity is much less (up to 30%) than the average DNI responsivity.
During the middle of the day under clear skies, the DHI contribution is small and there-
fore the average GHI responsivity is very close to the average DNI responsivity. Therefore,
the GHI measurements would show little spectral effect.
20. The shade–unshade method of calibration compares the direct irradiance on a horizontal
surface (DrHI) measured by the pyranometer under test with the reference DNI irradiance as
measured by a reference pyrheliometer. When the pyranometer is unshaded, the instrument
measures the GHI irradiance. When the pyranometer is shaded, it measures the DHI irradi-
ance. By subtracting the DHI from the GHI value, the DrHI value is obtained. By dividing
the DrHI by the cosine of the sza, the DNI value is obtained. By comparing this DNI value
against a reference DNI value, the calibration of the pyranometer can be obtained.
21. The summation calibration methodology compares the measured GHI value in volts from
the pyranometer under test with a corresponding DNI value from a reference pyrheliom-
eter (traceable to the WRR) projected onto a horizontal surface by multiplying the DNI by
the cosine of the sza and adding a reference DHI value from a shaded pyranometer that has
been previously calibrated. The uncertainty in the shade-unshade calibration is typically
smaller than the uncertainty of the summation method because one only uses the reference
DNI for comparison instead of a reference DNI and a reference DHI instrument. Also,
the shade-unshade method practically eliminates any thermal offsets in the pyranometer
under calibrations as the GHI and DHI offsets are assumed to be nearly the same.
22. Type A uncertainties are associated with uncertainties obtained from a series of measure-
ments, such as repeated comparison of measurements by an instrument under test with mea-
surements from a reference instrument. The cosine response of a pyranometer is checked by
binning the calibrations in sza bins. The standard deviation of these measurements in each bin,
is a Type A uncertainty. Similar analyses of repeated measurements for responses to changing
azimuth, temperature, spectral, and linearity can also contribute to Type A uncertainties.
23. Type B uncertainties are non-statistical and are often inferred from effects that are not mea-
sured. For example, a voltmeter has an uncertainty that is specified by the manufacturer.
This is a Type B uncertainty. Separately one can measure the uncertainty of the voltmeter,
but when using the instrument to measure the pyranometer, this is still a Type B uncer-
tainty. Uncertainties in the reference instruments are Type B uncertainties. See Table 5.10
for more examples.
24. The combined uncertainty is obtained by quadrature, the square root of the sum of the
squares. For the example, square root of ±(3% × 3% + 4% × 4%) is ±5 percent. The com-
bined uncertainty at the 95 percent level of confidence is twice the combined uncertainty
or ±10 percent.
Answers to Chapter Questions 475

CHAPTER 6
1. Molecular or Rayleigh scattering and scattering from aerosols. Also, acceptable is
scattering/reflection from obstructions in the field of view.
2. Use a tracking ball or disk that continuously blocks the sun from a thermopile-based
pyranometer. An enhanced answer would include calculating and subtracting the offset or
using a zero-thermal offset pyranometer.
3. A very clear sky can produce a large negative thermal offset in many single-black detector
pyranometers making a measured value of DHI smaller than a pure Rayleigh sky possible.
4. Band-alignment errors and insufficient correction for the excess skylight blocked by the
large, fixed band.
5. Black and white pyranometers have very small, or negligible thermal offsets because the
hot (black) and cold (white) thermopile junctions are exposed equally to the sky.
6. The horizon will be brighter because of multiple scattering by molecules and aerosols.
7. Spectral response of a photodiode pyranometer only covers a limited portion of the solar
spectrum (~400 to ~1,100 nm) and is not uniform with wavelength. This type of pyranometer
also has known problems with angular (cosine) response and ambient temperature response.
8. Never. Solar and thermal radiation interact in many ways with the ever-changing atmo-
sphere to create non-uniform radiance distributions. The closest it gets to this is for over-
cast skies.

CHAPTER 7
1. Gueymard (2018) recommends a new value of 1,361 Wm−2  for the integrated extraterres-
trial solar spectrum. This value is also consistent with the more recent broadband measure-
ments from space of the total solar irradiance (TSI).
2. Rayleigh scattering varies approximately as λ−4; therefore, blue sunlight is scattered more
than red sunlight giving a clear sky with little aerosol a blue look. For the same reason a
setting sun looks red because nearly all the blue light has been preferentially scattered
from the direct beam.
3. The term single scattering albedo (ϖ0) is a measure of the probability that a photon encoun-
tering an aerosol particle will scatter from it. The lower the single scattering albedo the
more the absorption.
4. Water vapor (see Figure 2.15).
5. Around 555 nm. Green or greenish-yellow. Photopic response.
6. UVC is absorbed in the upper atmosphere by O3, O2, and H2O, so it does not reach the
Earth’s surface.
7. Plot the natural log of the measurements versus the air masses at which those measure-
ments were taken and extrapolate the linear fit to this to zero air mass. For stable conditions
this will yield the Top of the Atmosphere (TOA) value that would be measured for this
instrument at the solar distance for that day (see Figure 7.9).
8. From Table  7.1, at 555  nm the ozone absorption coefficient is 3.505e-21  molecule−1.
A Dobson unit is 2.687 × 1016 molecules·cm−3 (see text after Eq. 7.6). Therefore, following
Equation 7.7 we get

308 ⋅ 2.687 ⋅1016 ⋅ 3.505 ⋅10 −21 = 0.0290(ozone optical depth)

9. Using Equation 7.15 we get


1
mRayleigh = = 6.1577.
cos ( 81° ) + 0.50572 ⋅ ( 96.07995 − 81° )
−1.6364
476 Answers to Chapter Questions

10. One sun radiometer points toward the sun with a narrow field of view for measuring DNI;
the other is the rotating shadowband radiometer that measures GHI and DHI and calcu-
lates the equivalent DNI.
11. Using a diffraction grating or a prism.
12. The amount and distribution of atmospheric water vapor are highly variable.
13. Integrating water vapor measurements from a radiosonde or using GPS measurements.
A sun radiometer can also be used, but requires a good deal of effort to calibrate.

CHAPTER 8
1. Using two matched pyranometers mounted horizontally with one pointed up and  the
other pointed  down over a uniform surface that represents the general surroundings.
The NOAA SURFRAD network has installed the downward facing pyranometer at 9 m
above ground level. The WMO recommends 1–2 m AGL for the upward facing pyranom-
eter (CIMO 2017).
2. Generally, surfaces reflect radiation differently at different wavelengths. For  example,
green vegetation shows a great deal of structure. Snow is very neutral in spectral albedo
between 400 and 700 nm, but shows structure in the near infrared (see Figure 8.6).
3. White sky albedo produced by all sky radiation incident on the surface does not change
appreciably with sun angle. Black sky albedos change dramatically with solar angle with
the largest albedo at the lowest solar elevation.
4. Green alfalfa shows much spectral structure in albedo with a green peak near 550 nm and
a sharp infrared increase starting just before 700 nm. Wheat stubble shows a monotonic
increase with wavelength (see Figure 8.4).
5. A calm dark ocean shows a very low albedo, but wind-whipped white-capped waves will
cause this to increase.
6. Albedo can be measured accurately using a single sensor that can be alternately pointed up
and down if sky conditions are stable.
7. The albedo of the ocean is very small and the albedo of ice is large. As the polar icecap
melts, the albedo of the Arctic decreases dramatically meaning that much less light is
reflected. This results in increased absorption of solar radiation and higher surface tem-
peratures (a positive feedback mechanism).

CHAPTER 9
1. A thermopile-based pyranometer with a single black disk.
2. A black and white type pyranometer.
3. A photodiode pyranometer uses a silicon photodiode with a spectral response similar to a
silicon cell photovoltaic (PV) module. Because the spectral distribution of solar radiation
will change with air mass, the photodiode pyranometer’s responsivity will also change.
Photodiode pyranometer measurements require corrections for changing spectral condi-
tions, angular response, and ambient temperature to reduce the biases and uncertainties
when compared to reference broadband solar irradiance measurements.

CHAPTER 10
1. Manual reduction of the data from strip chart recorders.
2. The Yankee Environments Systems, Inc. MFRSR makes only four strategic measurements
for each cycle of the shadowband; band positioned out of the field of view, one block-
ing direct radiation and two symmetrically positioned near the Sun, but not blocking the
Answers to Chapter Questions 477

diffuser from seeing the direct beam. The latter two are used to correct diffuse irradiance
blocked by the Sun during the blocked beam measurement.
The Irradiance, Inc. RSR2 measures continuously during the band sweep and processes
the data to measure GHI and to measure and correct DHI.
3. The RSR2, RSI, and MFRSR provide the three major components of solar radiation GHI,
DNI, and DHI from a single pyranometer without the need for separate pyranometers and
a pyrheliometer mounted on an expensive solar tracker.
The RSR designs require post-measurement corrections for pyranometer temperature
response, angular response, and spectral response. The temperature-controlled MFRSR is
characterized for angular response and requires only spectral corrections to the output.
4. Traditional thermopile pyranometers are not typically used in rotating shadowband radi-
ometers because their slow response times do not  allow for rapid sampling required to
measure the shaded and unshaded outputs.
5. Typically, corrections for temperature, angular response, spectral response are made. Also,
a correction for the diffuse skylight blocked by the band during the beam-blocked mea-
surement should be made.
6. The GHI is equal to the lowest reading plus the highest reading. The DrHI is equal to the
highest reading minus the lowest reading. The DHI is equal to twice the lowest reading
because the unique shade mask blocks 50 percent of the dome.
7. The DNI value is equal to the DrHI value divided by the cosine of the solar zenith angle at
the time of measurement.

CHAPTER 11
1. A blackbody has an emissivity of unity. For a gray body the emissivity is less than 1.
2. The main atmospheric window is near 10 µm.
3. The windows of a pyrgeometer begin transmitting between 3 and 5 µm.
4. The working standard for broadband infrared measurements is the interim World Infrared
Standard Group (WISG) consisting of the average output of four pyrgeometers (two Eppley
PIRs and two Kipp & Zonen CG4s) maintained by the World Radiation Center in Davos,
Switzerland.
5. Because the exposed (hot) junction of the pyrgeometer thermopile sees the dome that cools
to the sky, it is most often at a lower temperature than the body temperature giving rise
to a small negative voltage. A positive thermopile signal indicates the pyrgeometer body
temperature is less than the effective sky temperature—typically the result of a transient
condition caused by a sudden change in the weather.
6. The dome and body temperature are measured in the Eppley PIR. Only the body tempera-
ture is measured for the Kipp & Zonen CG4/CGR4 because body and dome temperatures
are considered the same for this pyrgeometer.
7. 273.15 K + 37 C ° 310 K
8. The sun shade minimizes heating of the dome by sunlight, which decreases the noise asso-
ciated with the changing dome-case temperature correction. It also minimizes the effect of
possible solar leaks because of the short wavelength cut-on of the filter or pinholes in the
“solar blind” filter.

CHAPTER 12
1. The most accurate measurement of net radiation is the four-component measurement.
The very best measurement uses infrared radiometers that have been calibrated using the
WISG, one looking down and one looking up, the latter shaded from direct sunlight; the
478 Answers to Chapter Questions

solar radiometers should be calibrated using a transfer from the WSG with the down-
welling shortwave component a summation of DNI and DHI.
2. A single sensor with transparent windows to all radiation with hot junction looking up and
cold junction looking down; two sensors with transparent windows to all radiation—one look-
ing up and one looking down; two sensors with one measuring solar net and one measuring
infrared net; four component sensors with two solar and two infrared radiometers for indepen-
dently measuring the downwelling and upwelling shortwave and longwave irradiances.
3. Polyethylene transmits all solar and infrared wavelengths.
4. Net radiometers are less accurate than any of the four components because the uncertain-
ties of each instrument add in quadrature.
5. Net radiation is important because it provides a measure of what drives the weather:
Sensible heating, latent heating, and surface heating.

CHAPTER 13
1. Define metrology.
The science of measurement.
2. What is the estimated measurement uncertainty of the reference measurement standard for
solar irradiance?
The WRR has an estimated uncertainty of ±0.3 percent.
3. When is instrument calibration considered to be a characterization?
Never. A calibration is conducted under specific conditions and traceable to a recognized
reference measurement standard. Instrument characterization requires a series of measure-
ments under varying conditions affecting the instrument performance.
4. Describe three instances when an instrument needs calibration.
Any three of these conditions:
• Installing a new instrument
• After an instrument has been repaired or modified
• After a specified time of operation has elapsed
• Before and/or after a critical measurement period
• After an instrument has been exposed to a shock, excessive vibration, or physical dam-
age, which may have compromised the calibration integrity
• When radiometer readings appear questionable or are not consistent with other mea-
surements or general atmospheric conditions
• As specified by a requirement (e.g., customer specification, radiometer manufacturer
recommendations, or an applicable standard).
5. True or False:
A radiometer calibration provides the instrument-specific responsivity (Rs) specifying the
value for converting the radiometer output signal to irradiance (µV/Wm−2) and its esti-
mated measurement uncertainty.
TRUE
6. Describe the World Radiometric Reference and its application.
The WRR is the internationally recognized Reference Measurement Standard for measur-
ing solar irradiance. It is used to harmonize the calibration of pyrheliometers for compar-
ing measurements of direct normal irradiance above 700 Wm−2.
7. How can one acquire a copy of the publicly available Broadband Outdoor Radiometer
Calibration reports?
BORCAL reports are available from https://aim.nrel.gov/borcal.html.
8. What is the Stefan-Boltzmann law and why is it important to radiometry?
Answers to Chapter Questions 479

The law relates the temperature of an object to the radiative power. Pyrgeometer designs,
operation, and calibration depend on the heat transfer associated with known temperature
differences.
9. What is the estimated expanded measurement uncertainty for pyrgeometer calibrations
based on the World Infrared Standard Group?
For  a 95 percent level of confidence, the expanded uncertainty of the WISG is
±2.6 Wm−2.
10. Which of these instrument types for measuring infrared irradiance not  suited for all-
weather operation: WISG, IRIS, or ACP?
Considered absolute instruments, the IRIS and ACP lack protective windows needed for
all-weather operation.
11. Of all physical measurement domains, which is the least reliable?
Spectroradiometric measurements place severe demands on instrument design and
therefore have the greatest estimated uncertainties among measurement and test
equipment.
12. What is the source of wavelength calibrations for spectroradiometers?
Known atomic emission lines as determined by quantum mechanics provide the wave-
length calibrations of spectroradiometers in the laboratory.
13. Name the three National Metrology Institutes (NMI) providing reference measurement
standards for spectroradiometer calibration.
National Institute of Standards and Technology (NIST) in the United States
National Physical Laboratory (NPL) in the United Kingdom
Physikalisch-Technische Bundesanstalt (PTB) in Germany
14. What parameter can be determined from clear-sky measurements using a sun photometer/
radiometer?
The wavelength-dependent calibration factor (Vo) for each channel of the radiometer can
be determined by extrapolation of the measurements at various solar zenith angles during
stable atmospheric conditions.

CHAPTER 14
1. The World Meteorological Organization
2. 0°C is 273.15 K. 300 K is therefore 26.85°C. This is equivalent to 80.33°F.
3. Thermocouple, Resistive temperature detector (RTD)
4. An aspirator is a fan that blows air at a steady rate across a sensor. This flow of air prevents
cold or warm air from stagnating around the sensor and thus provides better exposure to air
around the sensor.
5. The dew-point temperature is the temperature at which dew or ice, if it is cold enough,
forms on the surface. A chilled-mirror hygrometer is used to measure dew-point tempera-
tures. The surface is cooled until dew forms on the surface.
6. Cup anemometers, propeller anemometers, sonic anemometers
7. Response time is the time to register 63.2 percent of a step change. For an anemometer,
the Response length is passage of wind (in meters) required for the output of a wind-speed
sensor to indicate a 63.2 percent of a step change in the input wind speed.
8. Wind blowing from the north has a wind direction of zero. A wind direction of zero is also
recorded for wind speeds less than 0.5 ms−1.
9. Turbulence, wind blowing perpendicular to the axis of rotation, low wind speeds, when the
propeller is not pointed in the direction of air flow. Old bearings need replacement and this
also can affect the measured airflow.
480 Answers to Chapter Questions

10. A wind rose is a plot of wind speed verses direction in a polar plot. The plot goes from 0°
to 360°. The length of the arm is related to the intensity of the wind speed in that direction
for the period under study.
11. A sling psychrometer measures relative humidity. It has two thermometers, one with a
wet wick and the other exposed to air (wet bulb and dry bulb temperatures). The wet
bulb either swings around or is stationary (depending on design). The  ambient and
dew-point temperatures are then measured and a psychrometric table gives the relative
humidity.
12. The  response of these instruments to water vapor is non-linear. Mathematical tables or
formula are provided by the manufacturer to adjust the measurements.
13. Aneroid Displacement Transducers use materials that expand or contract in response to
changes in pressure. Piezoresistive barometers run a current through a crystal whose resis-
tance changes with changes in pressure. Piezocapacitance barometers use a diaphragm and
substrate that has a gap that changes with pressure and the deformation of the diaphragm
increases or decreases the capacitance.
14. Resolution relates to the smallest difference distinguishable between measured values.
Uncertainty relates to the confidence assigned to a measured value. For example, a mea-
sured value is x ± y with a 95 percent level of confidence. In this case, the measured value
is within the range of x−y and x + y 95 percent of the time. Resolution is one contribution
to uncertainty so that uncertainty is always larger than resolution.
15. The unit for pressure is the Pascal (Pa); typically, units are expressed in hectopascals where
1 hPa = 100 Pa.
16. A horizon survey is a survey of the height of objects above the horizon. A horizon survey
is useful in positioning a solar monitoring station or a solar energy system by identifying
obstructions that will block direct sunlight during part of the day. The goal is to site the
station at a location with the minimal interference between the direct sunlight and the
obstructions around the station.
17. If a cloud is moving with the wind at 5 miles per hour, it will travel 1/4th the distance in
15 minutes. The information given does not allow the determination of the movement of
the shadow because the relative position of the sun to the cloud is not known.
18. If the cloud is 15 degrees above the horizon with respect to the sensor and the Sun is 45
degrees above the horizon, the cloud does not  block the Sun. However, as the day pro-
gresses, it is necessary to know the projected movement of the cloud and the geometry of
the Sun, sensor, and cloud. A sky imager can provide this information.
19. Satellite observations can map the movement of the clouds relative to the location of inter-
est. By measuring the movement and evolution of the clouds, it is possible to calculate
when the cloud will pass between the area of interest and the Sun.
20. Instruments measuring DNI irradiance also measure part of the circumsolar irradiance.
Concentrating solar power stations are designed to concentrate the irradiance coming from
the solar disk and maybe a little of the circumsolar irradiance. Circumsolar irradiance can
increase the measured DNI irradiance by 2 to 10 percent or more. In order to accurately
estimate the DNI that will be concentrated by the power station, the circumsolar irradiance
has to be subtracted from the DNI.

CHAPTER 15
1. What are a few issues to consider when selecting a solar monitoring station?
• The instruments at a solar-monitoring station should have minimal shading, especially
by other instruments at the station.
• The instruments should be easy to access for cleaning and maintenance.
Answers to Chapter Questions 481

• Wiring should be shielded from the environment and animals. Exposed wiring should
be kept to a minimum.
• Access to the station should be limited to those who need to maintain or calibrate the
instruments.
• Growth around the station should be manageable.
• Station layout should allow for side-by-side calibrations when possible.
2. How will sun path charts help in this process?
Sun path charts can be used to identify objects on the horizon that will block the direct
sunlight. The location of the site should be such that such obstructions should be mini-
mized (ideally, not to exceed 5 degrees elevation above the horizon).
3. Why are site logs useful?
Site logs serve many purposes:
• Site logs are useful when analyzing and archiving data. They tell what is happening at
the site if there is a question in the data.
• Help ensure that instruments are being properly maintained.
• Indicate if problems are occurring at the site such as alignment problems.
4. Why do some networks prefer field calibrations as opposite to changing out instruments for
calibrations?
Field calibrations provide a more consistent measurement at the site by using the same
instrument over a long period of time. Degradation of the responsivity over time can
be tracked and this allows for better year to year comparison of the data. Each time an
instrument is switched out, the absolute accuracy of the calibration is added to the relative
comparison from year to year.
In addition, field calibrations are done under circumstances under which the field mea-
surements are made.
5. What is the advantage of having all three primary solar measurements, GHI, DNI, and DHI?
The relationship between the global, diffuse, and direct horizontal radiation (DNI times
the cosine of the sza) can be used to identify problems with the measurement. For example,
if there is soiling of the pyrheliometer window, the calculated GHI will show a marked
decrease from the measure GHI.
6. What is the primary cleaning fluid for solar instruments?
Water and pressurized air
7. Explain why the calculated DHI is larger than the measured DHI in Figure 15.6.
The  rain scatters the sunlight incident on the pyrheliometer and reduces the measured
DNI. When the DNI is projected onto the horizontal surface and subtracted from the GHI,
the calculated DHI will be higher than typical.

CHAPTER 16
1. Weather satellites in geosynchronous orbits are used to model irradiance data between
60° S and 60° N; polar orbiting satellites are used for higher latitudes.
2. First, clear-sky irradiance is modeled using aerosol optical depth, ozone, and water vapor
data as input parameters. Next, the cloud index is determined from satellite images. Clear-sky
data and the cloud index information are then combined to general the estimated irradiance.
3. Satellite-derived datasets use values of reflected light obtained from satellite images. It is
looking at broad areas and estimates are averaged over a variety of cloud patterns. Satellite
images are taken at discrete times, at best, no more frequently than every 15 minutes.
Ground-based irradiance measurements are taken at one point and directly measure the
solar quantity of interest. The temporal interval for ground-based measurements can be
much shorter than satellite-derived datasets.
482 Answers to Chapter Questions

APPENDIX B
1. The clearness index is the irradiance divided by the equivalent extraterrestrial value.
2. DHI values are small when GHI values are small. DHI values are also small during clear
sky periods when the GHI values are large. Dividing DHI by GHI removes the duel rela-
tionship between GHI and DHI.
3. A clear sky model estimates the irradiance when there are no clouds in the sky.
4. The maximum hourly kt is about 0.8 and the maximum kb is about 0.75.
5. If the instrument used to derive the model or the instruments that are the source of data
used by the model have a consistent, large systematic biases that are different, then these
systematic biases need to be considered during the calculations. Otherwise, the systematic
biases are incorporated into the results of the analysis.
6. When the average of the cosine of the sza is fifteen minutes or longer, better results can be
obtained by weighting the cosine with the clear-sky DNI model because the DNI value is
largest when the cosine(sza) is largest. If the average value of the cosine(sza) over the time
period is used to multiply the average DNI value, the product will systematically under-
estimate the DNI on a horizontal surface. The correlation of high DNI values with high
cosine(sza) values means that a weighted cosine should be used. This is also true when
projecting DNI onto tilted surfaces.

APPENDIX F
1. Average energy incident on a one square meter PV panel is 5.6 kWh per day. If the panel is
16 percent efficient, then 1 m2 would produce an average of 0.90 kWh of electricity per day.
For a month that is 30 days long, 1 m2 of PV panels would produce 27 kWhr per month. If
the house uses 1200 kWh per month, 44 m2 of PV panels would be needed to supply the
electricity for the house. 44 m2 is approximately the size of a 7 by 7 m2. 1 m2 is 11 ft2 and
44 m2 is 484 ft2. These answers are rounded to the nearest integer.
2. Energy is the product of power and time. Energy is the ability to do work. For example
energy can be measured in Joules. A joule (energy) is a watt (power) second (time). Power
is the rate that energy is supplied. Think of power as the amount of muscle that has to be
worked while energy is the amount of work that is done per unit time. The energy supplied
by a utility is recorded in kW (power) hr (time).

APPENDIX K
1. Self-documenting data formats provide access to supporting metadata such as station loca-
tion, instrument serial numbers and calibration factors, measurement interval(s), and other
information important for the proper use and interpretation of the measured data.
2. The location of the measurement station is required to compute solar position, an impor-
tant input to solar data quality assessment analyses.
3. Theoretically, irradiance on a horizontal surface obeys the Beer-Lambert law that says
irradiance on a horizontal surface is proportional to the irradiance times the cosine of the
incident angle. There are many factors that can affect irradiance received on a pyranome-
ter’s surface. For example, the surface may not be exactly flat or horizontal. In addition, the
dome or filter over the detecting surface may not transmit the light uniformly. This non-
uniform transmission can be caused by different densities of material or the shape of the
covering. Soiling also can affect the uniform transmission of light.
4. Pyrheliometer used for DNI measurements should show little change in responsivity as the
solar zenith angle changes because the pyrheliometer is always pointed at the Sun. Some
change in pyrheliometer responsivities have been observed, but they are much smaller
Answers to Chapter Questions 483

than the cosine response observed in pyranometers. Therefore, there is only a small devia-
tion from true cosine response when the DNI measurements are projected onto a horizon-
tal surface. DHI measurements made with pyranometers with very small thermal offsets
and mounted on an automatic tracker have a deviation from true cosine response similar
to pyranometers that measure GHI irradiance. However, as explained in Section  6.4 of
this book, we can minimize the cosine response issue for DfHI by calibrating the diffuse
pyranometer at 45° solar zenith angle. Furthermore, the DfHI irradiance is only a fraction
(10 to 20%) of the DrHI irradiance under clear skies. Therefore, the GHI calculated from
DrHI plus DfHI has a smaller deviation from a Lambertian response than the GHI values
measured with a pyranometer.
5. The  NetCDF libraries support a machine-independent format for representing scientific
data. The core library is written in C, and provides an API for C, C++ and two APIs for
Fortran applications, one for Fortran 77 and one for Fortran 90. An independent imple-
mentation, also developed and maintained by Unidata, is written in 100 percent Java,
which extends the core data model and adds additional functionality. Interfaces to NetCDF
based on the C library are also available in other languages including R (ncdf, ncvar, and
RNetCDF), Perl, Python, Ruby, Haskell, Mathematica, MATLAB, IDL, and Octave.
Index
Note: Page numbers in italic and bold refer to figures and tables, respectively.

A Ångström
electrical compensation pyrheliometer, 38–40, 39
Abbot, C. G., 34, 41–43, 92 pyranometers, 44, 44
Abbot silver-disk pyrheliometer, 41–42, 42 pyrgeometers, 41
Absolute Cavity Pyrgeometer (ACP), 236, 261, 262 angular measurement, 10–12, 12
absolute cavity radiometers (ACRs), 74, 77, 77–78 angular response, 105
absolute sky-scanning radiometer (ASR), 259–260 Antarctic circle, 17
absorption, 163–165 AOD, see aerosol optical depth (AOD)
and scattering, clear-sky conditions, 25 aphelion, 8, 12, 344
absorptivity, 28, 47–48, 54, 107 Arctic circle, 17
accuracy, 337 ASCII data files, 445, 447
accurate diffuse measurements, 158–159 ASI-16 All-Sky Imager, EKO Instruments B.V.,
ACP, see Absolute Cavity Pyrgeometer (ACP) 289, 290
acquisition, 157, 219, 259, 299 ASR, see absolute sky-scanning radiometer (ASR)
acrylic diffuser, 118, 127, 306 ASTM Standard G173-03, 8
actinograph, 45–46, 46, 92 Astronomical Almanac, 385–386, 388
adjustment, 337 astronomical unit (AU), 344
advection, 97 asunpos, 385
AERONET, 177 asymmetric diurnal profiles, 438
aerosol optical depth (AOD), 84, 175–177, 265–266 asymmetry parameter, 164–165, 202
aerosol scattering, 153, 163–165 atmosphere transmission, 169–170
Aerovane anemometer, 277 atmospheric interactions
air mass, 104 aerosol scattering/absorption, 163–165
albedo, 3, 193 atmosphere transmission, 169–170
BRDF, 199, 199–200 gas absorption, 165
broadband, 193–194, 194 Rayleigh scattering, 163
measurements, 200–202 atmospheric pressure, see pressure
soils, 198 Atmospheric Radiation Measurement (ARM), 87, 435
spectral, 194–198, 195, 196, 198 atmospheric scattering concepts, 151–153
alcohol, 306, 310, 438 atmospheric water vapor, 282–283
all-wave net radiometers, 242–244 GPS satellites, 283
alt-azimuth, 15, 17 Vaisala, 283, 284
altitude, 15, 15, 176, 282, 313, 365, 444 Automated Surface Observing System (ASOS), 272
ambient temperature, 101, 271 automatic Hickey–Frieden (AHF) cavity radiometer,
measurements, 272–274 61–63, 62
response times, 272 auxiliary pyranometer, 219
temperature sensors types, 272 average monthly daily total (AMDT), 86
analemma, 13, 14 azimuth, 14–15, 111, 199, 206, 235, 365, 385
Anasazi, 7 azimuth response, 113, 137
ancillary measurements
ambient temperature, 271–274
atmospheric water vapor, 282–284 B
circumsolar instruments, 292–295
minimum accuracies, 295 baffle, 62, 72
pressure, 284–286 bankable, 2, 319, 329
relative humidity, 281–282 barometers
sky-imaging systems, 286–292 piezocapacitance, 286
wind speed and direction, 274–281 piezoresistive, 284–285
anemometers, 275 barometric pressure, 102, 136
cup, 275–276 Baseline Surface Radiation Network (BSRN), 86–87, 247,
installing, 278 455, 458
propeller, 277 beam–global correlations, 349–350, 350
sonic, 278 beam irradiance, see direct normal irradiance (DNI)
Aneroid displacement transducers, 284 beam radiation, see direct normal irradiance (DNI)
angle of incidence response, 104–107 Beer-Lambert law, 175, 343

485
486 Index

Beer’s law, 224 C


bias, 337
bidirectional reflectance distribution function (BRDF), calibration, 233–235, 249–252, 337
199, 199–200 diffuse pyranometers, 157–158
black-and-white pyranometers, 208 improved, 235–236
characteristics, 111–114 pyrheliometer, 78–80, 79
thermal offset, lack, 114, 114–115 ratio data, Eppley PSP pyranometer, 123
blackbody calibrators, 258 traceability, 4
blackbody irradiances, 230 Callendar pyranometer, 40, 40–41
black-disk thermopile pyranometers, 97–100, 126 Campbell Scientific, 397
angle of incidence response, 104–107 Campbell-Stokes Sunshine Recorder, 37–38, 38
ice/snow on dome, 108–109 card storage module, 438
nonlinearity, 103 Cartesian sun path charts, 365–375
optical anomaly, 109 cat ears effect, 125, 220
response degradation, 107 caustic, Eppley 8–48 pyranometer, 109, 110
spectral response, 103–104 celestial sphere, 9–11, 10
temperature dependence, 108, 108 cellular modem, 300
thermal offsets, 101–103 Celsius, 41, 125, 130, 271, 394
black-sky albedo, 199 centigrade, 271
bolometric sunshine receiver, 40 Chappuis, 165
boom-mounted anemometers, 278 charge-coupled device (CCD) spectrometers, 186
BORCAL, see Broadband Outdoor Radiometer chilled-mirror hygrometer, 281
Calibration (BORCAL) circumsolar instruments, 292
Bouguer-Lambert-Beer law, 264–266 automatically and continuously monitor, 292–294
BRDF, see bidirectional reflectance distribution function RSR, 294–295
(BRDF) SAM, 292
broadband albedo, 193–194, 194, 200–201 sky imager usage, 294
broadband global irradiance circumsolar radiation, 153
black-and-white pyranometers, 111–115 cleaning, 305–307
black-disk thermopile pyranometers, 97–109 clearness indices KT, K DF and K B, 344–345
GHI measurements, 91–97 clear-sky
photodiode-based pyranometers, 115–131 conditions, VCPC, 75, 76–77
pyranometer calibration, 131–145 models, 343–344
Broadband Longwave Radiometer Calibration, clock, 15–16, 70, 282, 302, 440
257–262 cloud cover, 206, 286, 289, 319, 324, 354–355
Broadband Outdoor Radiometer Calibration (BORCAL), cloud-index, 323
141, 253–254 cloud particles, 153
Broadband Outdoor Radiometer Calibration Longwave co-albedo, 164
(BORCAL-LW), 425–426 cold junction, 26, 26, 53, 74, 78, 93, 97–98, 111, 237, 272
control instrument history, 427 collimating tube, 41, 42, 183
environmental and sky conditions, 433–434 combined standard uncertainty, 337
instrument details, 426 combined uncertainty, 140
results summary, 426 comma separated variable (CSV) format, 445
30681F3 Eppley PIR, 430–432 Commission for Instruments and Methods of Observation
traceability, 428 (CIMO), 271
Broadband Outdoor Radiometer Calibration Shortwave communications, 302–303
(BORCAL-SW), 402–403 compass, 300
control instrument, 404 concentrating photovoltaic (CPV), 70–71, 205
environmental and sky conditions, 423–424 concentrating solar power (CSP), 70–71, 205, 292
instrument details, 404–406 conduction, 26–27, 27, 29
140043 Kipp & Zonen CMP22, 407–409 confidence level, 338
140108 Kipp & Zonen CHP1, 413–415 connections, 301, 439
results summary, 403 convection, 27
traceability, 411, 417 conversion factors, 393–394, 393–394
21096 Eppley 8–48, 419–421 coordinated universal time, 302, 316
Broadband Shortwave Radiometer Calibration, copper-constantan thermopile, 94
255–257 corn snow, 198
broadband spectral, 161 corrective maintenance actions
broad filter radiometry, 170 check/adjust orientation, 438–439
PAR, 173, 173–174 clean optics, 438
photometry, 170–172, 172 confirm ventilator, 439
UVA/UVB, 174 data acquisition system ground connections, 439
BSRN, see Baseline Surface Radiation Network (BSRN) data logger clock, 440
bucking thermopiles, 93 data logger power, 439
Index 487

instantaneous data, 439 surface-measured data, 86–88


instrument serial number, calibration factors, 439 tilted surface, 345–347
correlation, 338 TSI, 69
cosine response, Kipp & Zonen CM 10 pyranometer, 105, 106 uncertainty analysis, 78–84
Cotton Region Shelter, 273 direct solar transmission, 170
coverage factor, 338 discrete spectral measurements, 188
CPV, see concentrating photovoltaic (CPV) DNI, see direct normal irradiance (DNI)
critical damping factor, 275 Dobson spectrophotometer community, 176
cryogenic radiometer, 262, 264 dome
Cryogenic Solar Absolute Radiometer (CSAR), 64, 255 ice/snow on, 108–109
CSP, see concentrating solar power (CSP) polyethylene, 244
cubic spline, 142 downsizing, 321
cup anemometers, 275–276 downwelling solar radiation, 201
drip loop, 301
D
E
data logger, 302–303
clock, 440 Earth–Sun distance, 344
power, 439 ecliptic plane, 9, 10
data review, 309–311, 310, 311 EKO Instruments Trading Co., LTD., 74
Davis Instruments, Corp., 397 EKO MS-711 Specifications, 183, 185
daylight savings time, 16, 303 EKO spectroradiometers
declination, 10, 10–11, 14 DNI spectral measurements, 185
degradation, 47–48, 63, 107 GHI spectral measurements, 184
degrees of freedom, 338 electron volt, 29
Delta-T Devices Ltd, 225, 397 elevation, 15, 18, 102, 163, 323, 365
dependence elliptical path, Earth’s orbit, 9, 9, 12–13, 13
solar zenith angle, 81–82, 82 emissivity, 28, 229, 233, 241, 258
temperature, 108, 108 empirical models, 322
detector-based reference, 131 Engineering Correction data, 48
dew-point temperature, 178, 281 Eppley 8–48 pyranometer, 55–56, 56, 110
DIAL radiometer, see rotating shadowband radiometer specifications, 112
(RSR) Eppley 180° pyrheliometer, 46–48
diffraction, 33, 78, 181, 263 Eppley black-and-white type pyranometer, 111
diffuse fraction, 343, 347–349, 348 Eppley Laboratory, Inc., 154
diffuse horizontal irradiance (DHI), 91, 93, 151, 158, 193, Eppley Light Bulb, 47, 47–48
215, 289, 321, 343, 437, 447–449 Eppley PIR pyrgeometers, 237
component, forecasting, 354–355 Eppley PSP pyranometer, 101
diffuse fraction, 347–349 caustic on, 110
diffuse irradiance, 153–157, 324 distorted disk on, 107
shaded-disk measurement, 155 Eppley SPP pyranometer, 99
spectral data, 352 specification, 100
tilted surface, 345–347 equation of time (EoT), 11, 13, 14
direct beam radiation (DNI), 152 equinox, 7, 9–11, 10, 18, 19, 74, 154
direct current (DC), 435 equipment sources, 397–399
direct horizontal irradiance (DrHI), 92 error, 338
direct normal irradiance (DNI), 2–3, 20–21, 22, 23, 37, ETR, see extraterrestrial radiation (ETR)
91, 92, 118, 193, 213, 215, 224, 292–294, 309, expanded uncertainty, 339
323–324, 331, 332, 363, 436 exposed junction, thermopile, 232, 236
ACR, 77–78 extraterrestrial radiation (ETR), 8–9, 69
beam–global models, 349–350 extraterrestrial solar spectrum, 161–162, 162
clear-sky models, 343–344
component, forecasting, 354 F
daily integrated, 71
direct irradiance, 323 Fahrenheit, 271
ground-based modeling, 84 failure modes
issues, 88 asymmetric diurnal profiles, 438
operational thermopile pyrheliometers, 74–77 bad data time sequence, 438
partitioning, GHI, 69, 70 corrective maintenance actions, 438–440
pyrheliometer geometry, 72–74 data collection, 437–438
satellite model estimates, 71, 84–86 GHI, DNI and DHI, 436–437
shaded/un-shaded pyranometer measurements, 71–72 no signal, 435
solar energy conversion, 70 signal greater/less than physical limits, 436
spectral data, 352 unstable signal, 435–436
488 Index

FEL lamp, 263–264, 264 hot junction, 26, 26, 49, 53, 74, 78, 81, 97, 242
field calibrations, 313–314 hour angles, 14–15, 15
fixed shadowband measurements, 153–155 Hukseflux Thermal Sensors B.V., 74, 247, 398
forecasting solar irradiance, 353–356 humidity, 281–282
Fortran program, 13, 387–390 hydroelectric generation, 1
forward-scattered radiation, 153
Foster–Foskett sunshine recorder, 364 I
Fourier transform spectrometers, 229
four-sensor net radiometers, 245, 246 ideal gas, 271
Fresnel, 193, 212 ideal measurement, 201
illuminance, 25, 171, 292
G incident solar irradiance, 205
incident solar radiation, 212
gas absorption, 165, 166–169 information sources, 7
gas-discharge lamps, 262 infrared (IR)
gauge, 153, 272, 274, 285, 301, 335 irradiance, 3
Gaussian distribution, 339 measurements, 229–230
Geostationary Operational Environmental Satellite radiation, 445
(GOES), 84, 303, 320 spectrum, 229
geostationary satellites, 320–321 Infrared Integrating Sphere (IRIS), 236
geosynchronous satellites, 320 radiometer, 260, 261
Gershun, 72, 74 instantaneous data, 439
GHI, see global horizontal irradiance (GHI) interference filters, 60, 174, 180–181, 231,
glazing’s refractive index, 213 237, 245
glint, 193 Interim World Infrared Standard Group (WISG), 260–261
global horizontal irradiance (GHI), 3, 21, 22, 34, 151, internally consistent, 436–437
215, 309, 323, 331, 333, 334, 343, 345–352, International Pyrheliometer Comparisons (IPCs), 55, 61,
354–355, 436–437, 447–449 63, 78–79
daily integrated, 71 International Satellite Cloud Climatology Project
EKO MS-711 and MS-712 spectrometers, 186 (ISCCP), 321
global irradiance, 91, 322 iron absorption, 197
measurements, 91–97 Irradiance, Inc. RSR version 2 (RSR2), 217
partitioning, 69, 70 irradiance data, 311
pyranometer installations, 70 IR radiative loss/thermal offset, 98
RSR, 83 ISIS, 86
Global Positioning Satellite (GPS), 282–283, 302–303 isotropic, 151, 154, 158, 199–200, 225, 346–347
global tilted irradiance (GTI), 3, 91, 205 isotropic diffuse model, 346
GMT, see Greenwich mean time (GMT)
gnomon, 17, 19 J
GOES, see Geostationary Operational Environmental
Satellite (GOES) Josephson effect, 249
GOES-16 satellite, 321
Greek, 1, 7, 33–34, 91 K
Greenwich mean time (GMT), 16, 178, 440
ground-based modeling, DNI, 84 K B, 344–345
ground-based solar radiation databases, 455–457, 458–459 K DF, 344–345
ground-reflected radiation, 346 Kelvin (K), 271
Guide to the Expression of Uncertainty in Measurements Kipp & Zonen CG 4, 234
(GUM), 78–79, 137–138, 141–144, 330–331 Kipp & Zonen CM 22 pyranometers, 98
procedure, 141–144, 143 and CM 21 pyranometers, 99
protocols, 137 Kipp & Zonen model CNR1, 245
Kipp & Zonen NR Lite2, 243
H Kipp & Zonen solarimeter, 44–45
Kipp & Zonen SP Lite pyranometer, 121, 122
Hartley-Huggins band, 165 Kirchhoff’s law, 114
heat transfer KT, 344–345
conduction, 26–27, 27
convection, 27 L
radiative transfer, 28
hectopascal, 284 Lambertian response, 212
Henyey–Greenstein function, 164 Lambert’s cosine law, 104, 158
Hickey-Frieden (HF) cavity radiometer, 120 Laminar flow, 27
Horicatcher, 288 lamp vs. Langley calibrations, 176
horizon brightening, 151 Langley plots, 175–177, 180, 264–266
Index 489

layout, solar-monitoring station, 315–317 NASA Surface meteorology and Solar Energy (SSE)
lensing effect of clouds, 21, 22 website, 85
LI-COR 200R pyranometer specifications, 122 National Aeronautics and Space Administration (NASA), 69
LI-COR model LI-200 pyranometer, 121, 125, 126, 127, SSE approach, 85
128, 130, 218 uncertainty in, 325
LI-COR model LI-200SA, 56–57, 57 National Climatic Data Center1 (NCDC), 445
LI-200SA pyranometer measurements, 209 National Institute of Standards and Technology (NIST), 103
light, spectral transmission, 104 National Oceanic and Atmospheric Administration
light detection and ranging (LIDAR), 275 (NOAA), 87, 234
limit angle, 73 National Physical Laboratory (NPL), 64, 181
Local Solar Time (LST), 16 National Renewable Energy Laboratory (NREL), 78, 219,
local standard time (LST), 15–16, 20, 303–304 307, 451
Loschmidt’s number, 165 National Weather Service (NWS), 363
LOWTRAN (LOW resolution atmospheric NetCDF (Network Common Data Format), 452
TRANsmission), 352 net radiation measurements
LST, see Local Solar Time (LST) accuracy, 246
luminosity function, 170 four-sensor, 245, 246
overview, 241–242
M single-sensor (all-wave), 242–244
two-sensor, 244–245, 245
maintenance, solar monitoring stations, 305–307 net radiation standard, 246–247
Marvin pyrheliometer, 43–44 net radiometer sources, 247
MBE, see mean bias error (MBE) Network Common Data Format (NetCDF), 452
mean, 339 nonlinearity, 103
mean bias error (MBE), 85–86, 324 normal distribution, 339
measurand, 329, 330, 339, 342 normal incidence pyrheliometer (NIP), 51, 51–52, 52, 72,
measurement 72–74, 436
equation, 79, 137, 339 novel pyranometer, 225
interval, 303–305, 304–305 NREL pyrheliometer comparison (NPC), 78
result, 340 numerical weather prediction (NWP), 355
standard, 339
measurement uncertainty, 329 O
budget, 341
estimating, 331–334 180° pyrheliometer, 91
measurable difference, determining, 335–337 opening angle, 73
reducing, 334–335 operational considerations, 236–238
Measure the Integral Transmittance of Windows (MITRA), 64 operational thermopile pyrheliometers
mercury-argon source emission, 263 solar zenith angle dependence, 81–82, 82
meridional plane, 11 temperature sensitivity, 81, 82
meteorological data, 102 uncertainties, 80, 80–81
metrology, 339 window transmittance, 81
MFRSRs, see multi-filter rotating shadowband operator error, 339
radiometers (MFRSRs) optical depth, 163
Middleton Solar, 236 optical radiation measurement, 34, 34–36
mismatch factor (MMF), 115 optics, 438
MODTRAN® (MODerate resolution atmospheric orientation, 438–439
TRANsmission), 352 outdoor calibration, 251
molecular scattering, 163 ozone absorption, 165, 166–169
Moll thermopile, 44–45, 45 Ozone Monitoring Instrument (OMI), 165
monitoring station, 300
monochromators, 262, 263 P
Moon/Spencer, 153
multi-filter rotating shadowband radiometers (MFRSRs), PAR, see photosynthetically active radiation (PAR)
179, 179, 188, 215, 223–225 Peltier effect heater, 281
multivariate correlation, 130 percent uncertainties, 144, 144–145, 145
photodiode-based pyranometers, 115–120, 116, 119,
N 212–213, 353, 442–444, 443, 444
characterizing, 121–124, 123
n- and p-type doped silicon, 29 reference solar cells, 130–131, 131
narrow-band filter radiometry removing biases, 124, 124–130
aerosol optical depth, 175–176 photodiodes, 28–29
sun radiometers, 179–181 photometry, 170–172, 172
water vapor, 177–179 photopic response, 25
narrowband photometry, 161 photosynthesis, 173
490 Index

photosynthetically active radiation (PAR), 173, R


173–174, 194
photosynthetic photon flux density (PPFD), 174 radiant energy, 97
photovoltaic (PV), 385, 395 density, 395
cells, 3 radiation shields, 272
photodiode pyranometers, 212–213 radiative heat transfer, 28
physical models, 321 radiative transfer modeling, 352
Physikalisch-Meteorologisches Observatorium Davos Radiometer Calibration Certificates, 257
(PMOD), 78, 108, 182, 183, 249, 259–260 radiometer calibrations, 249–252
Physikalisch-Technische Bundesanstalt (PTB), 250 broadband longwave, 257–262
piezocapacitance barometers, 286 broadband shortwave, 255–257
piezoresistive barometers, 284–285 frequency of, 253–255
PIR, see precision infrared radiometer (PIR) purpose, 252–253
pixel-to-cloud index conversion, 322–323 spectral calibrations, 262–266
Planck’s law, 28 standard methods for, 256
plane of array (POA), 159 traceability, 250
platinum resistance thermometer (PRT) radiometers
detectors, 272 Ångström and Tulipan pyrgeometers, 41
PMOD, see Physikalisch-Meteorologisches Observatorium Ångström electrical compensation pyrheliometer,
Davos (PMOD) 38–40, 39
polar sun path charts, 356, 375–384 Callendar pyranometer, 40, 40–41
polycrystalline silicon (Si), 285 Campbell-Stokes Sunshine Recorder,
polyethylene domes, 244 37–38, 38
potential locations, 300 pyrheliometer, 37, 37
precipitable water vapor (PWV), 101, 283 radiometry, 393, 393
precision, 339 random effects, 339
Precision Filter Radiometers (PFR), 266 random error, 339
precision infrared radiometer (PIR), 53, 53, 259 range, 339
Precision Solar Spectroradiometer (PSR), 182 Rayleigh scattering, 151, 163
precision spectral pyranometer (PSP), 48, 49, 50 reading, 339
Prediction Of Worldwide Energy Resource (POWER) receiver, bolometric sunshine, 40
project website, 85 record, 307–309
pressure rectangular distribution, 339
Aneroid displacement transducers, 284 reference
piezocapacitance barometers, 286 diffuse pyranometer, 100
piezoresistive barometers, 284–285 diffuse values, 155
primary absolute cavity radiometer (PACRAD), 54, instruments, 249, 251
54–55 measurement standards, 251
Process Control Instruments, 251 scale, 339
propeller anemometer, 277 solar cells, 130–131, 131, 211
PSP, see precision spectral pyranometer (PSP) standard lamps, 249
pyranometer calibration, 158 reflectance, 47, 63, 85, 114, 194–195,
shade–unshade calibration method, 132–135, 198, 205
133, 134 reflected radiation, 199
summation method calibration, 135–136, 136, 137 reflectivity, 187–188, 193–194, 197–198, 209,
uncertainties, 137–139 212, 213
pyranometers, 20–21, 34, 206, 307 relative humidity, 281–282
classification, specifications, 95–96 relative spectral response, 117
pyrgeometers, 230–233, 257–261 repeatability, 339
body temperatures, thermistor circuit, 238 reproducibility, 339
calibrations, 258–261 resistance temperature detector (RTD), 272
Eppley PIR, 237 resolution, 340
manufacturers, 236 response
thermopile detector, 231 degradation, 107
transmission curves, 231 length, 275
pyrheliometers, 20, 33, 37, 37, 306, 307 times, 272
calibration, 78–80, 79, 254, 255 responsivity, 74, 80, 82, 95, 105, 107, 113, 115, 117, 120,
NIP, 51, 51–52, 52, 72, 72–74, 436 123, 124–126, 128–130, 132, 136, 139–140,
operational thermopile, 74–77, 75 142–143, 340
optical geometry, 73, 73 result (of a measurement), 340
right ascension, 10–11, 15
Robitzsch bimetallic actinograph, 45–46, 46
Q root mean square errors (RMSEs), 85–86, 142
quality control, 311–313, 312, 313 rotating shadowband pyranometer (RSP), 218
Index 491

rotating shadowband radiometer (RSR), 71–72, 156–157, sky imager, 294


202, 222, 292, 294–295 sky-imaging systems, 286
advanced commercial versions, 58 circumsolar irradiance measure, 294
angular response, 59 site surveys, 286–288
atmospheric turbidity, 59 sky conditions, 289–292
innovations, 58 skylight, 153–154, 156, 158, 199–200, 223–224
instrument design, 57–58, 58 sky radiance distribution, 152
multifilter, 60, 223–225 sky temperature, 207
overview, 215–217 sling psychrometer, 281
spectral response, 59 slope angle, 73
temperature response, 59 Smithsonian water-flow pyrheliometer, 43, 43
uncertainty analysis, 83, 83–84 solar cells, 94
using silicon detector, 217–222, 220 building, 441–442
round-robin, 259 photodiode-base pyranometer, 442–444,
RSR, see rotating shadowband radiometer (RSR) 443, 444
solar constant, 33, 41, 69
S solar coordinates, 9–14
solar day, 11, 12, 304
sampling method, 447 solar electric power plants, 30
Santa Barbara DIScreet Ordinate Radiative Transfer solar-generating facilities, 1
(SBDART), 187–188 solar/infrared spectral regions, 23
satellite-derived solar radiation databases, 460–467, solar irradiance, 395
463–467 solar irradiance forecasting, 353–356
satellites day ahead forecasts, 356
deriving irradiance, 321–324 DHI component, 354–355
geostationary, 320–321 DNI component, 354
irradiance models, 324–325 GHI, 355
model estimates, 84–86 global tilted irradiance, 355
scattering intra-hour forecasts, 355–356
aerosols, 163–165 NWP, 355
atmospheric, 151–153 solar irradiance monitor (SIM), 162
Rayleigh, 163 Solar Light Company, 399
Schenk star pyranometer, 111, 124, 208, 222, 244 solar millennium AG, 217
cosine response, 113 solar monitoring stations, 4
image, 112 cleaning and maintenance, 305–307, 307
specifications, 113 data logger/communications, 302–303
Schott WG295 glass window, 105 data review, 309–311, 310, 311
Schulze-Däke net radiometer, 244 field calibrations, 313–314
scientific equipment sources, 397–399 grounding/shielding, 301
scotopic response, 170 measurement interval, 303–305, 304–305
Seebeck Effect, 26, 93 physical layout, 315–317
sensitivity coefficient, 340 quality control, data, 311–313, 312, 313
shade–unshade calibration method, 132–135, 133, record keeping, 307–309, 308, 309
134, 184 site choosing, 299–301, 300
shadow, 59, 154, 157, 218, 309 solar noon, 11
shield, 301 Solar Pathfinder™, 286, 287
short time interval data, 450–452, 451 solar position
shortwave, 241–242, 245, 247 algorithms, 385–390, 386
SI (International System of Units), 25, 251, 255, 262 azimuth angle, 17
sidereal day, 12 latitude 35°N/45°N, 17, 18
signal LST and solar time, 16
greater/less than physical limits, 436 sundial plot, latitude 45°N, 17, 19
unstable, 435–436 three-dimensional plot, latitude 23.44°N, 17, 19
silicon photodiodes, 28–29 solar radiation
silicon solar reference cell, 117 beam–global correlations, 349–350, 350
silver-disk pyrheliometer, 41–42, 42 clearness indices KT, K DF, and K B, 344–345
Simple Model of the Atmospheric Radiative Transfer of clear-sky models, 343–344
Sunshine (SMARTS), 187–188 diffuse fraction, 347–349, 348
single black detector pyranometers, 206–207 solar zenith angle, average cosine of,
single scattering albedo, 164 351–352
single-sensor (all-wave) net radiometers, 242–244 spectral irradiance modeling, 352–353
site data system, 437 tilted surface, irradiance on, 345–347
site survey, 286–288 SOlar Radiation and Climate Experiment (SORCE)
siting, 4, 286 satellite, 69, 162
492 Index

solar radiation databases Steinhart-Hart equation, 238


ground-based, 455–457, 458–459 stepper motor, 215
satellite-derived, 460–467, 463–467 Step Solar, 286, 288
TMY, 457–460, 461 storage module, 438
Solar Radiation Monitoring Laboratory (SRML), 445–447, stratospheric aerosol layer, 163, 164
449, 451 student’s statistic, 340
SOLar RADiation (SOLRAD) network, 47, 47, 87 student’s t-distribution, 79, 137, 141, 338
Solar Radiation Research Laboratory (SRRL), 87 summation calibration method, 129, 135–136, 136, 137
solar resource, 1–2, 4 The Sun, 7–8, 8
solar spectral measurements Sun and Aureole Measurement (SAM), 292
atmospheric interactions, 163–170 sundial, 17, 19
broad filter radiometry, 170–174 SunEye, 286, 287
extraterrestrial solar spectrum, 161–162, 162 sun path charts
narrow-band filter radiometry, 175–181 Cartesian, 365–375
overview, 161 polar, 356, 375–384
spectrometry, 181–188 sun photometers, 179, 264
solar time, 15–17, 18, 20, 365, 470 sun radiometers, 179–181
solar zenith angle (SZA), 14–15, 21, 23, 351–352, 450 sunrise/sunset times, 20
solid electrical ground, 301 Sunshape Profiling Irradiometer (SPI), 292, 293, 294
Solmetric, 286, 287, 300 sunshine duration, 363–364, 364
solpos.c code, 13 sunshine recorder, 37–38
solstice, 11, 19, 135, 154, 470 sunspots, 33
sonic anemometer, 278 SURFace RADiation (SURFRAD) network, 87
SPA, 386 albedo tower, 201
spectral albedo, 194–198, 195, 196, 198, 201–202 station, 315
spectral calibrations, 262 systematic bias, 340–341
Langley plots, 264–266 systematic errors, 138
measurement equation, 262–263 SZA, see solar zenith angle (SZA)
standard lamps, 263–264
spectral correction, 224 T
spectral data, 352–353
spectral distribution Teflon, 243, 244
solar irradiance, 23–24, 24 temperature dependence, 108, 108
sunlight, 33 temperature measurements
spectral irradiance modeling, 352–353 airflow, 272
LOWTRAN, 352 Cotton Region Shelter, 273
MODTRAN, 352 temperature stability, 181
photovoltaic (PV), 352 thermal isolation, 206
radiative transfer modeling, 352 thermal offsets, 98, 101–103, 206
spectral data, 352–353 lack of, 114, 114–115
spectral models, 186–188, 187 thermal radiation, 447
Spectralon® diffuser, 224 thermistor circuit, 237, 238
spectral response, 103–104 thermocouple, 272
spectrometers, 182–184, 186 thermodynamics/heat transfer
spectrometry, 181 conduction, 26–27, 27
discrete spectral measurements, 188 convection, 27
spectral models, 186–188, 187 radiative, 28
spectrometers, 182–184, 186 thermoelectric effect, 93
spectroradiometers, 262 thermopile-based pyranometers, 441
spectroradiometry, 161 thermopile Eppley PSP pyranometer, 210, 210, 211
spectrum, 161–163 thermopile pyranometers, 108, 219
specular, 33, 85, 205 thermopile voltage, 237
SPI, see Sunshape Profiling Irradiometer (SPI) tilted irradiance, 211–212, 324
SPN1 sunshine pyranometer, 225, 225–226 tilted surfaces, 23
SRML, see Solar Radiation Monitoring Laboratory global irradiance, 345–347
(SRML) tilt effects
stable conditions, 266 black-and-white pyranometers, 208
standard deviation, 338, 340 photodiode pyranometers, 209–210
and confidence level, Gaussian distribution, 138 single black detector pyranometers, 206–207
standard lamps, 263–264 time sequence, 438
standard precision pyranometer (SPP), 99 top of the atmosphere (TOA), 69, 73, 319, 363
standards data formats benefits, 452–453 total irradiance monitor (TIM), 63, 64, 162
standard uncertainty, 340, 341 total solar irradiance (TSI), 7–8, 33, 69, 344
Stefan–Boltzmann law, 229, 258 total solar radiation, 91
Index 493

traceability, 341 vignetting, 73


tracker vinegar, 306
alt-azimuth, 17
solar, 70, 153, 349 W
Transfer Reference Instruments, 251, 256
transmission water-flow pyrheliometer, 43, 43
of atmosphere, 169–170 water vapor, 177–179
Mauna Loa Observatory, 164 watts per square meter, 395
spectral, 103, 104, 105 Weather Bureau thermopile, 47
transmittance, 81, 96, 104 Wesely-type constant-speed RSR, 216
true value, 132, 138, 337–338, 341 white-sky albedo, 199
Tulipan pyrgeometers, 41 Wien’s law, 28
turbulent flow, 27 window, 229, 306
Twin RSI, 157 wind resource, 301
two-sensor net radiometers, 244–245, 245 wind rose, 280
type A uncertainty, 79, 141, 338, 474 wind speed/direction, 274–275
type B uncertainty, 79, 138, 140, 141–143, 474 anemometers, 275
Typical Meteorological Months (TMM), 457 cup anemometer, 275–276
Typical Meteorological Year (TMY), 457–460, 461 installing anemometer, 278
propeller anemometer, 277
U sensor terminology, 275
sonic anemometer, 278
ultraviolet A (UVA), 174 wind vanes, 278–281
ultraviolet B (UVB), 174 wire-wound copper–constantan thermopile, 48, 49
ultraviolet (UV) radiation, 174 Working Reference Instruments, 251, 257
uncertainty analysis World Climate Research Program, 86
measurement uncertainty principles, 329–342 World Infrared Standard Group (WISG), 53
operational thermopile pyrheliometers, 80, World Meteorological Organization (WMO), 94
80–82 World Radiation Center (WRC), 249, 259–260
pyranometer calibration, 139–141 World Radiation Data Center (WRDC), 38, 458
pyrheliometer calibration, 78–80, 79 World Radiometric Reference (WRR), 55, 77–78, 131,
RSR/DNI, 83–84 162, 249–250, 255
uncertainty budget, 341 World Standard Group (WSG), 60, 61, 78
uniform distribution, 128 Wulf, 165
universal time (UT), 16
University of Oregon Solar Radiation Monitoring Y
Laboratory (UO SRML), 86–87
UO SRML data file structure, 445 Yanishevsky pyranometer, 49–51, 50
data synopsis, 449–450 Yankee Environmental Systems, Inc., 157, 224
header, 446, 446–449 seven-channel RSR (MFR-7), 216, 216
short time interval data, 450–452, 451
Z
V
zenith angle, 9, 11, 12, 15, 199
variable conditions pyrheliometer comparison (VCPC), definition, 14–15
75, 76–77 determination of, 17
vegetation, 194–197, 209, 315, 346, 355 determining average cosine of, 351–352
ventilator (VEN), 108–109, 109, 207 uncertainty due to dependence of, 81–82
vernal, 10–11 Z-test, 336

You might also like