100% found this document useful (1 vote)
143 views30 pages

Introduction To Astrochemistry

Uploaded by

anirbanc2004
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
143 views30 pages

Introduction To Astrochemistry

Uploaded by

anirbanc2004
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 30

Astronomy and Astrophysics Library

Satoshi Yamamoto

Introduction to
Astrochemistry
Chemical Evolution from Interstellar
Clouds to Star and Planet Formation
Astronomy and Astrophysics Library

Series editors
Gerhard B€orner, Garching, Germany
Andreas Burkert, München, Germany
W.B. Burton, Mathews, Virgin Islands, USA
Athena Coustenis, Meudon CX, France
Michael A. Dopita, Weston Creek, Aust Capital Terr, Australia
Bruno Leibundgut, Garching, Germany
Georges Meynet, Versoix, Switzerland
Peter Schneider, Bonn, Germany
Virginia Trimble, Irvine, California, USA
Derek Ward-Thompson, Preston, United Kingdom
Ian Robson, Edinburgh, United Kingdom
Martin A Barstow, Heidelberg, Baden-Württemberg, Germany
More information about this series at http://www.springer.com/series/848
Satoshi Yamamoto

Introduction to
Astrochemistry
Chemical Evolution from Interstellar
Clouds to Star and Planet Formation
Satoshi Yamamoto
Department of Physics
The University of Tokyo
Tokyo, Japan

ISSN 0941-7834 ISSN 2196-9698 (electronic)


Astronomy and Astrophysics Library
ISBN 978-4-431-54170-7 ISBN 978-4-431-54171-4 (eBook)
DOI 10.1007/978-4-431-54171-4

Library of Congress Control Number: 2016962456

© Springer Japan 2017


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or
dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt
from the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained
herein or for any errors or omissions that may have been made. The publisher remains neutral with
regard to jurisdictional claims in published maps and institutional affiliations.

Cover illustration: Satoshi Yamamoto

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer Japan KK
The registered company address is: Chiyoda First Bldg. East, 3-8-1 Nishi-Kanda, Chiyoda-ku, Tokyo
101-0065, Japan
Preface

In modern astronomy and astrophysics, chemistry is becoming increasingly impor-


tant. Molecules are found everywhere in the universe owing to rapid progress of
spectroscopic observations particularly in the radio, infrared, and optical regimes.
Exploring chemical compositions of astronomical sources itself is of fundamental
importance in understanding evolution of matter in the history of the universe.
Moreover, chemical compositions are widely used as a powerful tool for investi-
gating physical conditions and formation processes of astronomical sources. Such a
chemical approach will rapidly be expanded to various areas of astronomical and
astrophysical studies in the near future. Astrochemistry is now one of the funda-
mental subfields of astronomy and astrophysics.
Astrochemistry is becoming more and more complicated as it progresses rapidly.
Hence, it is not easy even for experts to understand its whole scope in detail. When
researchers working in astronomy and astrophysics employ astrochemistry to cope
with actual problems, they may sometimes utilize a small, particular part of
astrochemistry for their studies without careful consideration of assumptions and
limitations hidden in the background. In order to make full use of the power of
astrochemistry, a deep understanding of its fundamental astrochemical concepts
based on physics and chemistry is essential.
On the other hand, astrochemistry is deeply related to molecular science, which
includes molecular physics, molecular spectroscopy, chemical dynamics, and the-
oretical chemistry. Because chemical processes in the universe are generally
ongoing under extreme physical conditions of low density and low temperature in
comparison with terrestrial conditions, astrochemistry has been posing new and
interesting research subjects in molecular science. In order for researchers in
molecular science to discover new chemical and physical phenomena in astronom-
ical studies, a deep understanding of the fundamental concepts of astrochemistry
will be very useful.
This book is organized to describe those fundamental concepts for researchers
and students in the broad fields of astronomy, astrophysics, and molecular science.
For this reason, recent topics are not always included, although some of them are

v
vi Preface

introduced as examples. Furthermore, this book does not cover the entire field of
astrochemistry but puts a particular emphasis on outlining chemical evolution from
molecular clouds to star and planet formation. For other subfields of astrochemistry
(e.g., external galaxies and late-type stars), readers should refer to available review
articles, although the basic part of the astrochemical concept in this book will still
be useful for understanding their contents.
Twenty-five years ago, I moved from the spectroscopic field to the
astrochemistry field, triggered by fortuitous discoveries of some new interstellar
molecules. Since then, I have been studying the chemical evolution of molecular
clouds toward star and planet formation with my colleagues and graduate students.
Because of my limited experience and knowledge of this field, I am afraid that some
important issues might be missing or poorly presented. However, I was encouraged
by the younger generation desiring an introductory book on astrochemistry, and I
have made up my mind to publish this work. I welcome any suggestions and
criticisms from readers.
Finally, I express my sincere gratitude to all my colleagues and collaborators,
particularly to Nami Sakai, Tomoya Hirota, Takeshi Sakai, Yoshimasa Watanabe,
Shuji Saito, Masatoshi Ohishi, Norio Kaifu, and the late Hiroko Suzuki. I also thank
the members of my group for their valuable comments.
I dedicate this book to my wife, Keiko.

Tokyo, Japan Satoshi Yamamoto


November 17, 2014
Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Matter in Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Interstellar Matter and Its Circulation . . . . . . . . . . . . . . . . . . . . 2
1.3 Interstellar Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4 Importance of Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Principal Aim of This Book . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Units Used in Astronomy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.7 Reaction Rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2 Derivation of Molecular Abundances . . . . . . . . . . . . . . . . . . . . . . . 11
2.1 Thermal Emission . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 Information Derived from Spectral Line Observations . . . . . . . 13
2.2.1 Radiative Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.2 Absorption Coefficient . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.3 Optically Thin Case: Rotation Diagram Analysis . . . . . 20
2.2.4 Optically Thick Case . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 Mechanism Determining Excitation Temperature . . . . . . . . . . . 25
2.4 Practical Issues for Deriving Molecular Abundances . . . . . . . . . 28
2.4.1 Source Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4.2 Forward and Backward Approaches . . . . . . . . . . . . . . 30
2.4.3 Molecular Parameters . . . . . . . . . . . . . . . . . . . . . . . . . 30
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3 Basic Concepts for Gas-Phase Chemical Reactions . . . . . . . . . . . . . 37
3.1 Chemical Reactions: Macroscopic and Microscopic
Viewpoints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2 Reactions in Interstellar Clouds . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3 Ion–Molecule Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4 Neutral–Neutral Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

vii
viii Contents

3.5 Radiative Association Reactions . . . . . . . . . . . . . . . . . . . . . . . 51


3.6 Dissociative Electron Recombination . . . . . . . . . . . . . . . . . . . . 52
3.7 Photodissociation and Photoionization . . . . . . . . . . . . . . . . . . . 54
3.8 Timescale of Chemical Reactions . . . . . . . . . . . . . . . . . . . . . . 58
3.9 Timescale for Chemical Equilibrium . . . . . . . . . . . . . . . . . . . . 60
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4 Chemistry of Diffuse Clouds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.1 Physical Conditions in Diffuse Clouds . . . . . . . . . . . . . . . . . . . 65
4.2 Molecules Detected in Diffuse Clouds . . . . . . . . . . . . . . . . . . . 67
4.3 Carbon Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.4 Oxygen Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.5 Nitrogen Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.6 Chemical Model of a Diffuse Cloud . . . . . . . . . . . . . . . . . . . . 78
4.7 CH+ Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.8 H3+ Problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.9 Diffuse Interstellar Bands . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5 Chemistry of Molecular Clouds I: Gas Phase Processes . . . . . . . . . 91
5.1 Physical Conditions in Molecular Clouds . . . . . . . . . . . . . . . . . 91
5.2 Molecules in Molecular Cloud Cores . . . . . . . . . . . . . . . . . . . . 93
5.3 H3+ Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.4 Destruction of Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.5 Carbon Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.6 Oxygen Chemistry and Its Relation to Carbon Chemistry . . . . . 112
5.7 Nitrogen Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.8 Sulfur Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5.9 Anion Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
5.10 Gas-Phase Chemical Models and Comparisons
with Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
6 Chemistry of Molecular Clouds II: Gas–Grain Processes . . . . . . . . 131
6.1 Roles of Dust Grains in Astrochemistry . . . . . . . . . . . . . . . . . . 131
6.2 Depletion of Atoms and Molecules onto Dust Grains . . . . . . . . 132
6.3 Grain-Surface Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
6.4 Formation of H2 Molecules on Grain Surfaces . . . . . . . . . . . . . 142
6.5 Formation of Various Molecules on Grain Surfaces . . . . . . . . . 144
6.6 Desorption of Molecules by Nonthermal Processes . . . . . . . . . . 147
6.7 Observation of Grain Mantles . . . . . . . . . . . . . . . . . . . . . . . . . 148
6.8 Gas–Grain Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
6.9 Deuterium Fractionation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
Contents ix

7 Chemistry of Star-Forming Regions . . . . . . . . . . . . . . . . . . . . . . . . 161


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
7.2 Formation of Low-Mass Stars . . . . . . . . . . . . . . . . . . . . . . . . . 163
7.3 Formation of High-Mass Stars . . . . . . . . . . . . . . . . . . . . . . . . . 166
7.4 Chemical Compositions of Low-Mass Star-Forming
Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
7.4.1 Hot Corino Chemistry . . . . . . . . . . . . . . . . . . . . . . . . 167
7.4.2 Warm Carbon-Chain Chemistry Sources . . . . . . . . . . . 170
7.4.3 Chemical Diversity . . . . . . . . . . . . . . . . . . . . . . . . . . 173
7.5 Chemical Compositions of High-Mass Star-Forming
Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
7.5.1 Orion KL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
7.5.2 W3(OH) and W3(H2O) . . . . . . . . . . . . . . . . . . . . . . . 176
7.5.3 Infrared Dark Clouds . . . . . . . . . . . . . . . . . . . . . . . . . 180
7.6 General Features of Chemical Processes in Star-Forming
Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
7.7 Chemical Differentiation Between Oxygen-Bearing
and Nitrogen-Bearing Species . . . . . . . . . . . . . . . . . . . . . . . . . 183
7.8 Carbon Chemistry in Low-Mass Star-Forming Regions: Hot
Corino Chemistry and WCCC . . . . . . . . . . . . . . . . . . . . . . . . . 185
7.9 Outflow Shocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
7.10 Phosphorous Chemistry in Star-Forming Regions . . . . . . . . . . . 192
7.11 Photodissociation Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
7.12 Polycyclic Aromatic Hydrocarbons (PAHs) and PDRs . . . . . . . 196
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
8 Chemistry of Protoplanetary Disks . . . . . . . . . . . . . . . . . . . . . . . . . 205
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
8.2 Basic Physical Structure of Protoplanetary Disks . . . . . . . . . . . 206
8.3 Molecules Detected in Protoplanetary Disks . . . . . . . . . . . . . . . 210
8.4 Chemical Processes in Protoplanetary Disks . . . . . . . . . . . . . . . 213
8.4.1 Disk Midplane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
8.4.2 Disk Surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
8.4.3 Molecular Zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
8.4.4 Vertical and Radial Mixing . . . . . . . . . . . . . . . . . . . . 217
8.4.5 Inner Disk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
8.5 Chemical Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
8.6 Molecules in Comets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
9 Chemical Evolution from Interstellar Clouds to Star-
and Planet- Forming Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
Further Readings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
x Contents

10 Appendix 1: Rotational Spectra of Molecules . . . . . . . . . . . . . . . . . 233


10.1 Rigid Rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
10.2 Rotational Spectra of Diatomic and Linear Molecules . . . . . . . . 234
10.3 Rotational Spectra of Symmetric-Top Molecules . . . . . . . . . . . 236
10.4 Rotational Spectra of Asymmetric-Top Molecules . . . . . . . . . . 238
10.5 Fine Structure in Rotational Transitions . . . . . . . . . . . . . . . . . . 239
10.6 Hyperfine Structure of Rotational Transitions . . . . . . . . . . . . . . 244
10.7 Internal Rotation and Inversion . . . . . . . . . . . . . . . . . . . . . . . . 246
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
11 Appendix 2: Observational Techniques . . . . . . . . . . . . . . . . . . . . . . 253
11.1 Atmospheric Transmittance . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
11.2 Antennas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
11.3 Receivers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
11.3.1 Receiver Configuration . . . . . . . . . . . . . . . . . . . . . . . . 256
11.3.2 SIS Mixers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
11.3.3 HEB Mixers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
11.3.4 Measurements of Receiver Noise Temperature . . . . . . 260
11.4 Radio Spectrometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
11.5 Aperture Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
11.6 Intensity Calibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
12 Solution to Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
Chapter 1
Introduction

1.1 Matter in Space

According to recent cosmological observations (Hinshaw et al. 2013; Planck


Collaboration 2014), 95.4 % of the total mass of the Universe is composed of
dark energy and dark matter, and the contribution of baryons (comprising atoms
and molecules, familiar to all) is only 4.6 %. Nevertheless, the existence of this
small baryonic fraction enables structures such as galaxies, stars, and planets to
form. Likewise, it is well known that 98 % of the total mass of baryons is composed
of hydrogen (H) and helium (He), whereas the total contribution of heavy elements
such as carbon (C), nitrogen (N), and oxygen (O) is only about 2 %. However, the
presence of this smaller fraction of heavy elements makes possible the enormous
variety of compounds creating the rich chemical world of materials around us.
In astronomy, exploring structure formation and physical evolution processes is
a fundamental task in understanding the history of the Universe. For instance, the
formation of the first stars, the formation of galaxies, and the formation of stars and
planets in galaxies are all currently outstanding topics, and a significant amount of
research is being conducted to understand the detailed processes involved during
formation and evolution. At the same time, exploring the chemical evolution of
matter in the Universe is also a fundamental goal of astronomy. One important
aspect of this evolution is nucleosynthesis in stars. Although only H, deuterium (D),
He, and a small amount of lithium (Li) were produced in the Big Bang, heavy
elements have been produced in stars little by little over the long history of the
Universe (1.38  1010 yr). Synthesized heavy elements in stars are distributed into
the interstellar medium in supernova explosions and mass-loss processes of late-
type stars. Increasing abundances of heavy elements significantly affect the phys-
ical process of star formation by increasing the cooling rate of interstellar gas. A
more important and interesting aspect of the evolution of matter is molecular
evolution, which is particularly important in relation to understanding star and
planet formation and eventually the origin of the Solar System. How are various

© Springer Japan 2017 1


S. Yamamoto, Introduction to Astrochemistry, Astronomy and Astrophysics
Library 7, DOI 10.1007/978-4-431-54171-4_1
2 1 Introduction

molecules formed in interstellar clouds? How large can molecules grow in these
clouds? How are molecules incorporated into stars and planets? These questions
represent current frontiers in astronomy. Studies of structure formation processes
and chemical evolution processes of matter are apparently two halves of a whole
toward a comprehensive understanding of the history of the Universe.

1.2 Interstellar Matter and Its Circulation

In our Galaxy (the Milky Way), about 90 % of the baryon mass resides in stars, and
the remaining is in interstellar matter. Because the total mass of the Galaxy is of the
order of 1011 MJ, the total mass of interstellar matter is about 1010 MJ. Interstellar
matter consists of gas and dust particles; the latter are composed of silicate and
carbonaceous compounds, with sizes typically about 0.1 μm in diameter. Gas
dominates by mass, and the dust-to-gas mass ratio is about 0.01 in the solar
neighborhood. Because this ratio depends on metallicity—the abundance of
heavy elements—it varies among galaxies and even within our Galaxy.
Elemental abundances in interstellar matter near the solar neighborhood are
listed in Table 1.1 as number ratios relative to hydrogen nuclei. Hydrogen is the
most abundant element in the Universe, and helium is the second most abundant.
Second raw elements such as C, N, and O have abundances of 103–104 relative to
H. Heavier elements generally show lower abundances, and abundance signifi-
cantly varies within the same raw elements, reflecting the nucleosynthetic processes
occurring in stars. In particular, iron (Fe), being the element with the highest
binding energy per nucleon, is relatively abundant despite its high mass. The
elemental abundances given here do not correspond to abundances in the gas
phase. Because heavier elements tend to exist in dust grains, their gas-phase
abundances are generally lower.
Interstellar matter is not uniformly distributed over the galaxy, but is rather
concentrated in clumps and filaments of various scales, called “clouds.” Figure 1.1
shows the density and temperature for some representative classes of clouds. The
density and temperature of extended clouds such as diffuse clouds, intercloud gas,
coronal gas, and some molecular clouds lie almost on a straight line with constant
pressure, indicating that these clouds are in pressure equilibrium. In contrast, the
densest parts of molecular clouds are gravitationally contracting and hence deviate
from the line of constant pressure. Moreover, H II regions around high-mass stars,
where hydrogen mostly exists as H+, are not on the line.
The coronal and intercloud components of the interstellar medium are both
diffuse and ionized and are often referred to as hot ionized medium (HIM) and
warm ionized medium (WIM), respectively (McKee and Ostriker 1977). Such
media will form diffuse clouds after their protons recombine with electrons. In
diffuse clouds, atomic hydrogen and molecular hydrogen coexist. As the tempera-
ture decreases and the density increases, the diffuse clouds gradually become
opaque to interstellar ultraviolet (UV) radiation, and hydrogen molecules become
1.2 Interstellar Matter and Its Circulation 3

Table 1.1 Cosmic Element Relative abundance


abundance of elementsa
H 1.00E 0
He 8.51E-02
Li 1.12E-11
Be 2.40E-11
B 5.01E-10
C 2.69E-03
N 6.76E-05
O 4.90E-04
F 3.63E-08
Ne 8.51E-05
Na 1.74E-06
Mg 3.98E-05
Al 2.82E-06
Si 3.24E-05
P 2.57E-07
S 1.32E-05
Cl 3.16E-07
Ar 2.51E-06
K 1.07E-07
Ca 2.19E-06
Sc 1.41E-09
Ti 8.91E-08
V 8.51E-09
Cr 4.37E-07
Mn 2.69E-07
Fe 3.16E-05
a
Asplund et al. (2009)

the dominant form of hydrogen. Such clouds are called molecular clouds. Molec-
ular clouds are the formation sites of new stars. Protostars are formed by the
gravitational contraction of molecular cloud cores. In general, molecular clouds
harbor rich molecules formed in the gas phase and on dust grains. Protostars further
evolve into main-sequence stars, during which time planetary systems are formed.
Main-sequence stars are stable for relatively long times during hydrogen burn-
ing. The lifetime of a 1 MJ main-sequence star is 1010 yr. After hydrogen burning
is completed, main-sequence stars evolve into late-type star stages. Some late-type
stars show considerable mass loss, which supplies gas and dust to the interstellar
medium. Massive stars (>8 MJ) finally end their lives via supernova explosions,
which are the most energetic events in the Universe. Through supernova explo-
sions, heavy elements formed during nucleosynthesis in stars revert to interstellar
medium. Thus, interstellar matter circulates throughout the galaxy over the entire
life cycle of stars.
4 1 Introduction

Fig. 1.1 Temperature-density diagram of interstellar clouds. The dashed line indicates the
constant-pressure line. The diffuse interstellar medium is in pressure equilibrium

In the large-scale circulation of interstellar matter, the richest phase of molecules


occurs from the diffuse cloud stage to the star- and planet-formation stage. In this
book, we concentrate on chemical processes in this phase. Envelopes of evolved
stars are important places for the formation of dust grains and also exhibit rich
molecular chemistry. Because excellent reviews and books are dedicated to these
processes (e.g., Glassgold 1996; Ziurys 2006; Agundez et al. 2008), we do not
repeat these discussions in this book.

1.3 Interstellar Molecules

It is more than 75 years since the existence of molecules in interstellar clouds was
first established. In 1940, McKellar identified sharp optical absorption lines
(387.46 nm and 430.03 nm) in spectra toward bright stars detected using the
Mount Wilson 100-inch telescope as the electronic transitions of CN (B-X) and
CH (A-X), respectively. Subsequently, Douglas and Herzberg (1941) found CH+ in
interstellar clouds. Although these discoveries were just the tip of the iceberg of
molecular astronomy, they did not receive much attention from astronomers; these
molecules were thought to be fragments of large molecules or grains.
In the interim, rapid advances in microwave technology in the 1940s enhanced
the sensitivity of radio astronomy. Muller and Oort (1951) and Ewen and Purcell
(1951) discovered emissions from atomic hydrogen (H I) at 1.42 GHz (21 cm) in
interstellar clouds. This finding established the existence of cold atomic clouds in
interstellar space. Because radio waves penetrate through our Galaxy, the rotation
1.3 Interstellar Molecules 5

of our Galaxy was studied in detail by observations of the H I 21-cm line (e.g., Oort
et al. 1958; Van de Hulst et al. 1954; Kalberla et al. 2005; McClure-Griffiths et al.
2009; Peek et al. 2011). In 1955, Townes presented thorough discussions of the
possibilities of detecting atoms and molecules using radio observations (Townes
1957). Townes pointed out the importance of rotational spectral lines of molecules.
However, this proposal also did not receive much attention from astronomers.
In 1963, Weinreb et al. detected the Λ-type doubling transition (1.67 GHz) of
OH in absorption spectra toward the bright radio continuum source Cassiopeia
A. OH was the first interstellar molecule detected by radio astronomical observa-
tions. The radio emissions from OH often show anomalous features, including
maser action. Strong OH masers have been the subject of astrophysical studies of
interstellar clouds, as well as late-type stars (Robinson and McGee 1967; Elitzur
1992). Above all, the importance of molecules was first recognized with the
detection of NH3. In 1968, Cheung et al. detected the inversion transition of NH3
at 23.6 GHz in emission spectra toward the center of our Galaxy. The existence of a
familiar molecule such as NH3 in interstellar clouds clearly demonstrated the
existence of a new world of molecular astronomy. Cheung et al. (1969) also
detected extremely bright rotational emission from H2O at 22.2 GHz toward the
Orion Nebula and W49, which was later found to be due to maser action. Addi-
tionally, Snyder et al. (1969) detected the first organic interstellar molecule, H2CO.
Following these discoveries, HCN and CO were successively reported in interstel-
lar clouds (Snyder and Buhl 1971; Wilson et al. 1970). The original papers
reporting the discoveries of these molecules clearly convey the vivid surprise and
excitement of their authors.
To date, we know of about 170 interstellar molecules; see Table 1.2. Molecules
containing up to 13 atoms have so far been found in interstellar media. Most of
these molecules have been detected by radio observations of their rotational
spectral lines; some have been identified by optical or infrared observations of
their electronic or vibration-rotation spectra. A few characteristic features of
interstellar molecules can be deduced from Table 1.2. First, many chemically
reactive species such as free radicals and molecular ions are listed. This fact reflects
the very low density and very low temperature conditions of interstellar clouds
compared with laboratory conditions. Given the extreme conditions, chemically
reactive species can survive for a long time. Second, the existence of ionic species
such as HCO+, H3+, and C6H means that interstellar clouds are weakly ionized
plasmas. As will be discussed in later chapters, these ions play central roles in the
formation of various molecules in interstellar clouds. Third, a number of highly
unsaturated hydrocarbon molecules such as carbon-chain molecules and their
geometrical isomers exist despite the hydrogen-dominated conditions of interstellar
clouds. The presence of unsaturated hydrocarbon molecules in interstellar clouds is
the most characteristic feature of interstellar chemistry and is one of the important
subjects of this book. Finally, saturated complex organic molecules such as
HCOOCH3 and (CH3)2O can be found in some sources, meaning that chemical
evolution in interstellar clouds produces fairly complex molecules even under
6 1 Introduction

Table 1.2 Molecules found in interstellar clouds


Simple neutral molecules
H2, CH, CN, CO, HCl, NH, NO, NS, OH, PN, SO, SiO, SiS, CS, HF, O2, SH, CH2, HCN, HCO,
H2O, H2S, HNC, HNO, N2O, OCS, SO2, CO2, NH2, HO2, NH3, H2CO, H2CS, CH3, H2O2, CH4
Ionic species
(Cation)
CH+, CO+, SO+, CF+, OH+, SH+, HCl+, ArH+, HCO+, HCS+, HOC+, N2H+, H3+, H2O+, H2Cl+,
OH3+, HCNH+, HCO2+, C3H+, H2COH+, NH4+, H2NCO+, HC3NH+
(Anion)
C4H, C6H, C8H
Carbon-chain molecules and their isomers
C2, C3, C2H, C2O, C2S, c-C3H, l-C3H, C3N, C3O, C3S, C2H2, C5, C4H, l-C3H2, c-C3H2, HC3N,
HCCNC, HNC3, C5H, l-C4H2, C5N, C6H, CH3CCH, HC5N, CH3C3N, C6H2, CH2CCHCN,
CH3C4H, HC7N, CH3C5N, HC9N, CH3C6H, HC11N
Complex organic molecules
HCOOH, CH2CO, CH3CN, CH3NC, CH3OH, CH3SH, HC2CHO, c-C3H2O, CH2CNH,
HNCHCN, CH2CHCN, CH3CHO, CH3NH2, c-C2H4O, H2CCHOH, HCOOCH3, CH3COOH,
CH2OHCHO, CH2CHCHO, NH2CH2CN, CH3CHNH, CH3CH2CN, (CH3)2O, CH3CH2OH,
CH3CONH2, C3H6, CH3CH2SH, (CH3)2CO, (CH2OH)2, CH3CH2CHO, C2H5OCHO,
CH3OCOCH3, C2H5OCH3, n-C3H7CN
Other molecules
FeO, HNCO, HNCS, H2CN, HCNO, HOCN, HSCN, CH2CN, H2CNH, NH2CN, HCOCN,
HNCNH, CH3O, NH2CHO
Note: Based on the Cologne Database for Molecular Spectroscopy (CDMS) (Muller et al. 2001,
2008). The classification of molecules is arbitrary

extreme conditions. These saturated organic molecules are thought to be produced


mainly by grain-surface reactions.
In addition to gas-phase molecules, some molecules are found on dust grains as
solid molecules by infrared observations. Although water ice (H2O) is a dominant
species on dust grains, other species such as CO, CO2, CH4, NH3, and CH3OH are
known to be contained in H2O ice. The interplay between the gas-phase species and
solid-phase species is a central issue in current astrochemical studies.

1.4 Importance of Chemistry

After the discoveries of various molecules in interstellar clouds, the important and
interesting goal became the understanding of the chemical processes that produce
these species. Molecular synthesis through gas-phase ion–molecule reactions was
proposed and successfully explained the basic molecular abundances observed in
interstellar clouds (e.g., Klemperer 1970; Herbst and Klemperer 1973; Watson
1974). This mechanism of molecular formation was verified via the identification
of molecular ions such as HCO+ and HN2+ (Woods et al. 1975; Saykalley et al.
1976). Later, the importance of neutral–neutral reactions was recognized, even in
1.4 Importance of Chemistry 7

low-temperature conditions (Herbst et al. 1994). The rate coefficients of the


gas-phase reactions are often available from laboratory experiments. Even if the
rate coefficients are not measured, they can be estimated based on theoretical
considerations and/or quantum chemical calculations. Hence, a reaction network
considering a number of reactions can be developed quite readily, and numerical
simulations using the network are applied to interpret the observational data (e.g.,
Herbst and Leung 1989; Prasad and Huntress 1980; Millar and Nejad 1985; Graedel
et al. 1982; Lee et al. 1996).
In addition to the gas-phase processes, molecular synthesis through grain-sur-
face reactions was also considered just after the discovery of interstellar molecules
(Watson and Salpeter 1972a, b; Watson 1976; Iguchi 1975; Allen and Robinson
1977; Tielens and Hagen 1982). A reaction network similar to the gas-phase
network was also constructed, and numerical calculations were performed to
simulate the observed abundances. However, reaction probabilities on grain sur-
faces are much more uncertain, mainly because the grain surface itself is an
inhomogeneous system. Laboratory experiments are more difficult, and theoretical
estimates are less certain than for gas-phase processes. Hence, many assumptions
have to be made in deriving the reaction probabilities when constructing a reaction
network. For this reason, grain-surface models received less attention than the
gas-phase models for comparison with observations in the 1970s and 1980s.
Although the gas-phase chemical model succeeded in explaining the observed
abundances in cold quiescent molecular clouds, it failed to reproduce the abun-
dances of saturated complex organic molecules found in high-mass, star-forming
regions. Furthermore, the depletion of molecules onto dust grains was observation-
ally established (e.g., Caselli et al. 1999). Under these circumstances, the impor-
tance of gas-grain interactions was recognized as critical. Large chemical models
considering gas-grain interactions were developed to simulate chemical composi-
tions (e.g., Hasegawa et al. 1992; Garrod et al. 2007; Aikawa et al. 2005). Surface
reactions were studied in the laboratory, proffering better assumptions in the
chemical models. The gas-grain chemical model is now a standard model for
studying the chemical compositions of molecular clouds.
Detections of various interstellar molecules and relevant studies pertaining to
their production mechanisms stimulated molecular science researchers and gave
rise to various new investigations of molecular spectroscopy, molecular dynamics,
surface science, and quantum chemical calculations. Many new interstellar mole-
cules, mostly free radicals, were identified with the aid of laboratory molecular
spectroscopy. Above all, it should be remembered that an experiment aimed at
understanding the formation processes of carbon-chain molecules eventually led to
the discovery of fullerene C60 (Kroto et al. 1985). C60 itself was also discovered in
space many years later (Cami et al. 2010). Even now, astrochemical studies are
continually posing many interesting problems in the field of molecular science.
Given the steady progress of observational studies, source-to-source variations
of chemical compositions were recognized as an interesting line of research.
Exploring the origin of the chemical diversity in dense molecular cloud cores led
scientists to establish the concept of chemical evolution (e.g., Suzuki et al. 1992).
8 1 Introduction

Because chemical compositions systematically vary as clouds evolve, chemical


compositions can be used as an age tracer of dense cores. On the other hand, some
molecules preferentially exist in special physical conditions such as shocked
regions, photodissociation regions, and hot molecular cores around protostars.
Thus, chemical compositions are now being used to diagnose the physical condi-
tions and evolutionary stages of various astronomical sources. A chemical approach
is indispensable in modern astrophysics, and it will only become more so in the
future.

1.5 Principal Aim of This Book

This book describes the basic aspects of astrochemistry for researchers and graduate
students who are investigating molecules in interstellar media or who are purely
interested in the subject. This book is also for researchers in fields of molecular
science and planetary science. The main aim of this book is to provide the
fundamental physical and chemical concepts that astrochemical concepts are
entirely based on. Astrochemistry is a rapidly growing field that is constantly
pushing its frontiers. Researchers in this field need to remain abreast of rapid
developments and also have to produce new ideas and concepts from new obser-
vational results. Furthermore, chemical models are becoming more and more
complex, which makes it difficult for nonexperts to understand their applicability
and limitations. To handle such situations, basic concepts of astrochemistry are
essential and hence are covered in this book. Nevertheless, current frontiers and
topics of astrochemistry that are still controversial are not expanded on, although
some recent works are introduced as examples of applications. The reader is
encouraged to seek out the excellent review articles on these topical issues. Radio
astronomy is the most important observational technique of astrochemistry; its
basic principles are briefly introduced in the Appendix. Additional details of
radio astronomy are comprehensively presented in other books, and we hope that
the reader will refer to these resources.

1.6 Units Used in Astronomy

Because astrochemistry is an interdisciplinary field, the participation of researchers


outside of astronomy is essential for the development of the field. Hence, some
readers may not be familiar with the units used in astronomy. Before proceeding,
we briefly summarize here several of these important units:
1. Astronomical unit (au): One au corresponds to the average distance between the
Sun and the Earth, 1.496  1011 m.
References 9

2. Parsec (pc): One pc corresponds to the distance to an object whose annual


parallax is 2 arc seconds, the object being on the line perpendicular to the orbital
plane of the Earth. One pc is 3.086  1016 m, and hence an apparent source size
of 1 arc second at a distance of 100 pc corresponds to a size of 100 au.
3. Visual extinction (Av): Visual extinction describes the loss in magnitude of
visible light with a wavelength of 550 nm attenuated because of absorption
and scattering (extinction) by interstellar dust. Extinction is represented in units
of magnitudes; an extinction of 1 magnitude means that the intensity is reduced
by a factor of 100.4 (~2.512). Visual extinction is approximately proportional to
the column density of the hydrogen atom. For our local galactic neighborhood:

NðHÞ ¼ 2  1021 Av cm2 : ð1:1Þ

Alternatively, it is related to the column density of H2 in molecular clouds:

NðH2 Þ ¼ 1  1021 Av cm2 : ð1:2Þ

Note that the coefficient of proportionality may be different for galaxies of


different metallicity.
4. Solar mass (MJ): A mass unit, where 1 MJ corresponds to 1.988  1030 kg.
5. Solar luminosity (LJ): A luminosity unit, where 1 LJ corresponds to
3.845  1026 W.

1.7 Reaction Rates

Coefficients of reaction rates are often essential in discussing chemical processes in


interstellar clouds. The values of such coefficients used in this book are mostly
taken from the University of Manchester Institute of Science and Technology
(UMIST) Database for Astrochemistry (McElroy et al. 2013).

References

W.S. Adams, Annual Report of the Director of the Mount Wilson Observatory 1938–1939, pp. 23
M. Agundez, J. Cernicharo, J.R. Pardo, J.P. Fonfria Exposito, M. Guelin, E.D. Tenenbaum,
L.M. Ziurys, A.J. Apponi, Astrophys. Space Sci. 313, 229 (2008)
Y. Aikawa, E. Herbst, H. Roberts, P. Caselli, Astrophys. J. 620, 330 (2005)
M. Allen, G.W. Robinson, Astrophys. J. 212, 396 (1977)
M. Asplund, N. Grevesse, A.J. Sauval, P. Scott, Ann. Rev. Astron. Astrophys. 47, 481 (2009)
J. Cami, J. Bernard-Salas, E. Peeters, S. Malek, Science 329, 1180 (2010)
P. Caselli, C.M. Walmsley, M. Tafalla, L. Dore, P.C. Myers, Astrophys. J. 523, L165 (1999)
A.C. Cheung, D.M. Rank, C.H. Townes, D.D. Thornton, W.J. Welch, Phys. Rev. Lett. 21, 1701
(1968)
10 1 Introduction

A.C. Cheung, D.M. Rank, C.H. Townes, D.D. Thornton, W.J. Welch, Nature 221, 626 (1969)
A.E. Douglas, G. Herzberg, Astrophys. J. 94, 381 (1941)
M. Elitzur, Ann. Rev. Astron. Astrophys. 30, 75 (1992)
H.I. Ewen, E.M. Purcell, Nature 168, 356 (1951)
R.T. Garrod, V. Wakelam, E. Herbst, Astron. Astrophys. 467, 1103 (2007)
A.E. Glassgold, Ann. Rev. Astron. Astrophys. 34, 241 (1996)
T.E. Graedel, W.D. Langer, M.A. Frerking, Astrophys. J. Suppl. 48, 321 (1982)
T.I. Hasegawa, E. Herbst, C.M. Leung, Astrophys. J. Suppl. 82, 167 (1992)
E. Herbst, W. Klemperer, Astrophys. J. 185, 505 (1973)
E. Herbst, C.M. Leung, Astrophys. J. Suppl. 69, 271 (1989)
E. Herbst, H.-H. Lee, D.A. Howe, T.J. Millar, Mon. Not. R. Astr. Soc. 268, 335 (1994)
G. Hinshaw, D. Larson, E. Komatsu, et al., Astrophys. J. Suppl. 208, 19 (2013)
T. Iguchi, Publ. Astron. Soc. Japan 27, 515 (1975)
P.M.W. Kalberla, W.B. Burton, D. Hartmann, E.M. Arnal, E. Bajaja, R. Morras, W.G.L. Roppel,
Astron. Astrophys. 440, 775 (2005)
W. Klemperer, Nature 227, 1230 (1970)
H.W. Kroto, J.R. Heath, S.C. O’Brien, R.F. Curl, R.E. Smalley, Nature 318, 162 (1985)
H.-H. Lee, R.P.A. Bettens, E. Herbst, Astron. Astrophys. Suppl. 119, 111 (1996)
N.M. McClure-Griffiths et al., Astrophys. J. Suppl. 181, 398 (2009)
C.F. McKee, J.P. Ostriker, Astrophys. J. 218, 148 (1977)
D. McElroy, C. Walsh, A.J. Markwick, M.A. Cordiner, K. Smith, T.J. Millar, Astron. Astrophys.
550, A36 (2013). http://udfa.ajmarkwick.net/
A. McKellar, Publ. Astron. Soc. Pac. 52, 187 (1940)
T.J. Millar, L.A.M. Nejad, Mon. Not. R. Astr. Soc. 217, 507 (1985)
C.A. Muller, J.H. Oort, Nature 168, 357 (1951)
H.S.P. Muller, S. Therwirth, D.A. Roth, G. Winnewisser, Astron. Astrophys. 370, L49 (2001)
H.S.P. Muller, F. Schloder, J. Stutzki, G. Winnewisser, J. Mol. Struct. 742, 215 (2008)
J.H. Oort, F.J. Kerr, G. Westerhout, Mon. Not. R. Astr. Soc. 118, 379 (1958)
J.E.G. Peek et al., Astrophys. J. Suppl. 194, 20 (2011)
Planck Collaboration, Astron. Astrophys. 571, A16 (2014)
S.S. Prasad, W.T. Huntress Jr., Astrophys. J. Suppl. 43, 1 (1980)
B.J. Robinson, R.X. McGee, Ann. Rev. Astron. Astrophys. 5, 183 (1967)
R.J. Saykalley, T.A. Dixson, T.G. Anderson, P.C. Szanto, R.C. Woods, Astrophys. J. 205, L101
(1976)
L.E. Snyder, D. Buhl, Astrophys. J. 163, L47 (1971)
L.E. Snyder, D. Buhl, B. Zuckerman, P. Palmer, Phys. Rev. Lett. 22, 679 (1969)
H. Suzuki, S. Yamamoto, M. Ohishi, N. Kaifu, S. Ishikawa, Y. Hirahara, S. Takano, Astrophys.
J. 392, 551 (1992)
A.G.G.M. Tielens, W. Hagen, Astron. Astrophys. 114, 245 (1982)
C.H. Townes, in The Fourth International Astronomical Union Symposium, Manchester 1955,
ed. by H.C. Van de Hulst, (Cambridge University Press, Cambridge 1957), paper 16
Van de Hulst, C.A. Muller, J.H. Oort, Bull., Astron. Inst. Netherlands 452, 117 (1954)
W.D. Watson, Astrophys. J. 188, 35 (1974)
W.D. Watson, Rev. Mod. Phys. 48, 513 (1976)
W.D. Watson, E.E. Salpeter, Astrophys. J. 174, 321 (1972a)
W.D. Watson, E.E. Salpeter, Astrophys. J. 175, 659 (1972b)
S. Weinreb, A.H. Barrett, M.L. Meeks, J.C. Henry, Nature 200, 829 (1963)
R.W. Wilson, K.B. Jefferts, A.A. Penzias, Astrophys. J. 161, L43 (1970)
R.C. Woods, T.A. Dixon, R.J. Saykalley, P.C. Szanto, Phys. Rev. Lett. 35, 1269 (1975)
L.M. Ziurys, Proc. Nat. Acad. Sci. 103, 12274 (2006)
Chapter 2
Derivation of Molecular Abundances

2.1 Thermal Emission

Here, we consider the relation between radiation emitted by matter and the tem-
perature of matter. Imagine a gas burner heating a brick. The brick appears first
tinged with red; later, as it continues to be heated, it shines as white light. This
experiment indicates that a brick emits visible light at high temperature and that its
color depends on its temperature. Then, what about a brick at room temperature?
When a room-temperature brick is set in a closed room and the light in the room is
turned off, we cannot see the brick. That is, the brick at room temperature does not
emit visible light detectable by human eyes. We usually recognize a brick at room
temperature by the light that it reflects from sunlight or lights in the room.
Nevertheless, a room-temperature brick emits electromagnetic radiation in the
infrared part of the spectrum. Even without any illumination, we can indeed detect
such bricks using an infrared camera. Matter at different temperatures emits
electromagnetic waves of different wavelengths. For lower-temperature objects,
the emitted wavelengths become longer. Interstellar matter, whose temperature is as
low as 10 K, does not emit even infrared radiation but instead emits radio waves.
Cold interstellar matter can accordingly be traced by radio observations.
Radiation from matter at a nonzero temperature is called thermal emission. In
general, its intensity depends on the composition of material as well as its temper-
ature, to be described below. However, we can suppose as an extreme case that
matter is completely opaque and absorbs all radiation at any wavelength without
reflection. Such matter is called a blackbody. Among all matter, a blackbody gives
the maximum thermal emission at a given temperature. The above condition
defining a blackbody is too idealized, however, and it is rarely fulfilled in reality.
Most blackbodies can be regarded as blackbodies in a particular range of wave-
lengths. For instance, let us consider a brick and an iron block of similar shape. At
room temperature, these two objects can be readily distinguished in terms of how
they reflect visible light. Hence, they cannot be regarded as blackbodies. At high

© Springer Japan 2017 11


S. Yamamoto, Introduction to Astrochemistry, Astronomy and Astrophysics
Library 7, DOI 10.1007/978-4-431-54171-4_2
12 2 Derivation of Molecular Abundances

temperatures, both the brick and iron block similarly radiate white light; hence, we
cannot distinguish them. In this situation, radiation from the environment is
absorbed, and emissions from the two objects are seen. Hence, both the brick and
iron block can be regarded as blackbodies in this instance. It should be noted that
the emission from blackbodies does not depend on its composition, but only on its
temperature. The importance of this characteristic was first noted by Gustav
Kirchhoff (Kirchhoff 1860). Based on his findings, Max Planck later established
the theory of blackbody radiation and introduced the Planck constant (Planck
1901). This theory was one of the important milestones in the development of
quantum theory.
Energy emitted from the blackbody per area dS, time dt, frequency dν, and solid
angle dΩ is written as:

dE ¼ Bν ðTÞdSdtdνdΩ, ð2:1Þ

where Bv(T) is the intensity of the blackbody radiation at temperature T. The


intensity is represented by the Planck function as:

2hν3 1
Bν ðTÞ ¼ , ð2:2Þ
c2 expðhν=kB TÞ  1

where h denotes the Planck constant, c the speed of light, v the frequency of
radiation, and kB the Boltzmann constant. The units of Bv(T) are erg s1 cm2 Hz
1
sr1. It should be noted that, for a given frequency, Bv(T) only depends on
temperature, irrespective of composition and structure of the blackbody. Figure 2.1
shows a plot of Bv(T) as a function of frequency for various temperatures. The
frequency corresponding to the peak intensity rises at higher temperatures, the
relationship being known as Wien’s displacement law. For instance, the surface
temperature of the Sun is 6000 K, and hence the Sun can be regarded as a blackbody
with that temperature. Its radiation has an intensity peak in the visible wavelengths.
In contrast, the cosmic microwave background radiation can be well approximated
as a blackbody with a temperature of 2.73 K, whose intensity peak appears in the
microwave regime.
If hν/kBT1, then by expanding the exponential part of Eq. (2.2), the following
approximate relation known as the Rayleigh–Jeans law is derived:

2v2
Bν ðTÞ ffi kB T: ð2:3Þ
c2

In this case, the intensity is proportional to temperature. Note that this equation does
not involve the Planck constant and is regarded as the classical limit of blackbody
radiation. In contrast, for hν/kBT  1, Wien’s law holds:
2.2 Information Derived from Spectral Line Observations 13

Fig. 2.1 Spectral lines for blackbody radiation at various temperatures (Note the log–log scale of
the plot)

 
2hν3 hν
Bν ðTÞ ffi 2 exp  : ð2:4Þ
c kB T

In this case, the intensity exponentially decreases with increasing frequency.


As noted, blackbody radiation only depends on the temperature of the emitting
source. Nevertheless, thermal radiation generally contains rich information about
the nature of matter (e.g., atoms and molecules). For instance, we can determine the
abundance of atoms and molecules in interstellar clouds from the intensities of their
spectral lines. In the following sections, we describe the thermal radiation from
atoms and molecules.

2.2 Information Derived from Spectral Line Observations

2.2.1 Radiative Transfer

Consider the propagation of an electromagnetic wave with a frequency v in inter-


stellar clouds. Here, we assume that radiation travels in straight lines. This
14 2 Derivation of Molecular Abundances

assumption is justified when the beam diameter is much larger than the wavelength
(λ ¼ c/v). Similar to Eq. (2.1), the intensity Iv of the wave is defined from:

dE ¼ I ν dSdtdνdΩ: ð2:5Þ

The unit of intensity is erg s1 cm2 sr1 Hz1. Although intensity is conserved in
the vacuum of space (Problem 2.1), it varies due to absorption and emission by
atoms and molecules in the intervening clouds. We define the x axis as the direction
of propagation. A change in intensity, dIv, for an infinitesimal propagation of dx is
represented as:

dI ν ¼ αν I ν dx þ jν dx, ð2:6Þ

where αv is an absorption coefficient and jv stands for emissivity. Hence, the


following radiation transfer equation holds:

dI ν
¼ αν I ν þ jν : ð2:7Þ
dx

In general, αv and jv vary from position to position and hence are functions of x.
Dividing both sides of this equation by αv and defining a new variable τv:

dτν ¼ αν dx, ð2:8Þ

the equation simplifies:

dI ν
¼ I ν þ Sν , ð2:9Þ

where,

Sν ¼ jν =αν : ð2:10Þ

Here, Sv is called the source function, whose meaning is described below. By


solving the differential equation, the intensity at any position can be obtained, if
Sv is known. As a boundary condition, the intensity at x ¼ 0 (τv ¼ 0) is set as Bv(Tb),
meaning that the source radiation is blackbody radiation at Tb. If the cloud only
exists for x  0, the background radiation should be the cosmic microwave back-
ground radiation (Tb ¼ 2.73 K). For simplicity, we assume that the cloud is phys-
ically and chemically homogeneous. Then, Sv is constant over all positions x  0,
and we obtain the intensity at x ¼ L as follows:

I ν ¼ Sν þ expðτν Þ½Bν ðT b Þ  Sν , ð2:11Þ


2.2 Information Derived from Spectral Line Observations 15

where,

τν ¼ αν L: ð2:12Þ

Here, τv is a dimensionless quantity referred to as the optical depth or optical


thickness. This quantity is related to the physical thickness of the cloud, L, by the
above relation. Even for large physical thicknesses, the optical depth can be very
small if the absorption coefficient is small.
We consider here a source function, Sv. Suppose that the above cloud is
completely enclosed with a blackbody at a temperature of T. After sufficient
time, the cloud will be in thermal equilibrium at this temperature. If we make an
infinitesimally small hole to look at the cloud inside, the intensity that we observe is
Bv(T ), which is because the system (the cloud enclosed by the blackbody) remains a
blackbody. Furthermore, we can set Tb ¼ T. Hence, Eq. (2.11) is written as:

Bν ðTÞ ¼ Sν þ expðτν Þ½Bν ðTÞ  Sν : ð2:13Þ

This relation must hold for any value of τv, because blackbody radiation is inde-
pendent of the internal structure of the emitting body. To satisfy this condition, the
source function must be the Planck function. Specifically, we obtain:

Sν ¼ Bν ðTÞ: ð2:14Þ

This result means that the absorption coefficient and the emissivity are not inde-
pendent, but their ratio (the source function) is constrained by the temperature
through the Planck function. This relation is known as Kirchhoff’s law of thermal
radiation and is based on the physical requirement that emission and absorption
have to be balanced in thermal equilibrium.
By substituting Sv with Bv(T) in Eq. (2.11), we obtain:

I ν ¼ Bν ðTÞ þ expðτν Þ½Bν ðT b Þ  Bν ðTÞ: ð2:15Þ

In actual observations, the target position and a nearby position without any cloud
emission and absorption are alternatively observed, and their difference is taken to
reduce the effects of gain variation of the telescope system and changes in atmo-
spheric conditions. Because the intensity toward the off-source position is given as

I ν ¼ Bν ðT b Þ, ð2:16Þ

the difference is written as:

ΔI ν ¼ ½Bν ðTÞ  Bν ðT b Þf1  expðτν Þg: ð2:17Þ

This value represents the intensity actually observed by radio telescopes.


In radio astronomy, the intensity is typically represented by a temperature scale.
As recognized by differentiating the Planck function by T, the intensity of a
blackbody monotonically increases with increasing temperature at a given
16 2 Derivation of Molecular Abundances

frequency (Problem 2.2). Hence, the intensity can be represented by the tempera-
ture at which a blackbody emits the same intensity. In particular, the intensity is
proportional to the temperature in the Rayleigh–Jeans law (hν/kBT1). Hence, the
temperature scale of the intensity, ΔT, is defined using the Rayleigh–Jeans law,
regardless of its validity. Then, we obtain:
 
hν 1 1
ΔT ¼  f1  expðτν Þg: ð2:18Þ
kB expðhν=kB TÞ  1 expðhν=kB T b Þ  1

If hν/kBT1, the equation can be further simplified as:

ΔT ¼ ðT  T b Þf1  expðτν Þg: ð2:19Þ

This temperature scale for intensity is very useful because we can avoid the rather
complex units of erg s1 cm2 Hz1 sr1 and can readily imagine a physical
temperature of an emitter. Furthermore, this scale is convenient for intensity
calibrations from blackbody radiation, as explained in Chap. 11.

2.2.2 Absorption Coefficient

As described above, the observed intensity depends on the temperature of the cloud,
the background temperature, and the optical depth. The optical depth is related to
the absorption coefficient, which reflects all electromagnetic properties of an
emitter. Here, we formulate the absorption coefficient in terms of the properties
of atoms or molecules. For simplicity, we first consider a two-level system, of lower
level l and upper level u specified by the energy levels of a molecule or an atom
(Fig. 2.2), and consider the absorption coefficient for the transition from level l to
level u. The absorption probability per unit time is represented by Blu J v , where Bij is

Fig. 2.2 Two-level system: Populations for the upper and lower levels are denoted by nu and nl ,
respectively; degeneracies by gu and gl , respectively. Spontaneous emission, stimulated emission,
and absorption are indicated by the corresponding Einstein coefficients; collisional excitation and
de-excitation are also indicated
2.2 Information Derived from Spectral Line Observations 17

the Einstein B coefficient for the transition from level i to level j, and Jv represents
the mean intensity for the molecule:
Z
1
Jν ¼ I ν dΩ: ð2:20Þ

A similar relation holds for a stimulated emission from level u to level l. Hence, the
energy extracted from an electromagnetic-wave beam in a frequency range dν, solid
angle dΩ, time dt, and volume dV is given by:


dI ν dνdΩdtdS ¼ ϕðνÞðnl Blu  nu Bul ÞI ν dνdΩdtdV, ð2:21Þ

where ni stands for a number of molecules populating level i. Here, ϕ(v) is a line
profile function satisfying the following normalization relation:
Z
ϕðvÞdv ¼ 1: ð2:22Þ

Because dV ¼ dSdx, the absorption coefficient is written as:


αν ¼ ϕðνÞðnl Blu  nu Bul Þ, ð2:23Þ

by comparing Eq. (2.21) with its definition:

dI ν ¼ αν I ν dx: ð2:24Þ

The Einstein B coefficients, Blu and Bul , are related to each other and are
proportional to the Einstein A coefficient, Aul , as:

gl Blu ¼ gu Bul , ð2:25Þ

and

c2
Bul ¼ Aul : ð2:26Þ
2hν3ul

Here, gi represents the degeneracy of level i, and νul is the resonance frequency for
the two-level system. Using these relations, the absorption coefficient becomes:
   
c 2 nu hνul
αul ¼ exp  1 Aul , ð2:27Þ
8πν2ul Δν kB T
18 2 Derivation of Molecular Abundances

Fig. 2.3 Line profile


function employed for
simplicity. The line center
frequency is νul , and the line
width is Δν

where the Boltzmann distribution at a temperature T is assumed for the populations


of level l and level u according to:
 
nu gu hνul
¼ exp  : ð2:28Þ
nl gl kB T

Here, the line profile function is approximated, as shown in Fig. 2.3:

1
ϕðνÞ ¼ , ð2:29Þ
Δν

where the line width, Δν, is assumed to be much smaller than νul . The Einstein A
coefficient is written as:

64π 4 ν3ul Sul μ2


Aul ¼ : ð2:30Þ
3hc3 gu

Here, Sul is the line strength, which is the square of the matrix elements for the
direction cosines (Sect. 2.4.3), and μ is the dipole moment responsible for the
transition. In the case of a linear (or diatomic) molecule, Sul equals J þ 1 for the
transition J þ 1 $ J. These values are tabulated in spectral line databases for many
molecules.
If a Boltzmann distribution at temperature T is assumed for the level population,
the number of molecules in level u per unit volume can be written in terms of the
total number of molecules, n:
 
gu Eu
nu ¼ n exp  , ð2:31Þ
U ðT Þ kB T

where U(T) is the partition function at temperature T, and Eu the upper-state energy.
By combining Eqs. (2.27), (2.30), and (2.31), we obtain:
2.2 Information Derived from Spectral Line Observations 19

     
8π 3 νul Sul μ2 hνul Eu
αul ¼ n exp  1 exp  : ð2:32Þ
3hcΔνUðT Þ kB T kB T

The line width is usually represented in terms of a velocity width, rather than a
frequency width, because the line width largely depends on the Doppler effect. The
velocity width is defined as:

Δν
Δv ¼ c: ð2:33Þ
νul

Considering this relation, we finally obtain an expression for the optical depth as:
     
8π 3 Sul μ2 hνul Eu
τul ¼ exp  1 exp  N, ð2:34Þ
3hΔvU ðT Þ kB T kB T

where N is the column density, N ¼ nL. Apparently, the optical depth is proportional
to the column density, the line strength, and the square of the dipole moment.
Unless the depth of the cloud (L ) along the line of sight is known, only the column
density can be derived from observations.
In derivation of Eq. (2.34), the level population is assumed to be represented by
the Boltzmann distribution at a temperature T to relate the population of the upper
level to the total number of molecules. If this assumption is satisfied, the system is
said to be in local thermodynamic equilibrium (LTE). For a rotational energy-level
system, this temperature is called the rotation temperature.
The excitation temperature, Tex, for the transition u ! l can be determined by the
populations of level l and level u using the following relation:
 
nu gl hν
¼ exp  , ð2:35Þ
nl gu kB T ex

In LTE conditions, the excitation temperature is the same as the rotation temper-
ature. In interstellar clouds, molecules are excited and de-excited in collisions with
H2 molecules, although collisions with H atoms and electrons are also important in
some cases. That is, the H2 molecules form a heat bath. If the collisions are
sufficiently frequent when compared with radiation probabilities, the rotation
temperature is close to the kinetic temperature of H2, and the LTE condition is
satisfied. If the collisions are less frequent, cooling by radiation processes becomes
more important. The level population then diverges from LTE and cannot be
represented by a single Boltzmann distribution. In this case, the excitation temper-
ature differs from transition to transition and is usually lower than the gas kinetic
temperature. We describe this situation in Sect. 2.3.

You might also like