Using SI Units in Astronomy (PDFDrive)
Using SI Units in Astronomy (PDFDrive)
RICHARD DODD
Victoria University of Wellington
c a m b r i d g e u n ive r s i t y p r e s s
Cambridge, New York, Melbourne, Madrid, Cape Town,
Singapore, São Paulo, Delhi, Tokyo, Mexico City
Cambridge University Press
The Edinburgh Building, Cambridge CB2 8RU, UK
Published in the United States of America by Cambridge University Press, New York
www.cambridge.org
Information on this title: www.cambridge.org/9780521769174
© R. Dodd 2012
A catalogue record for this publication is available from the British Library
Preface page ix
Acknowledgements xii
1 Introduction 1
1.1 Using SI units in astronomy 1
1.2 Layout and structure of the book 2
1.3 Definitions of terms (lexicological, mathematical
and statistical) 3
1.4 A brief history of the standardization of units in general 7
1.5 A brief history of the standardization of scientific units 8
1.6 The future of SI units 11
1.7 Summary and recommendations 11
2 An introduction to SI units 12
2.1 The set of SI base units 12
2.2 The set of SI derived units 12
2.3 Non-SI units currently accepted for use with SI units 13
2.4 Other non-SI units 14
2.5 Prefixes to SI units 14
2.6 IAU recommendations regarding SI units 20
2.7 Summary and recommendations 23
3 Dimensional analysis 24
3.1 Definition of dimensional analysis 24
3.2 Dimensional equations 25
3.3 Summary and recommendations 29
4 Unit of angular measure (radian) 30
4.1 SI definition of the radian 30
4.2 Commonly used non-SI units of angular measure 30
v
vi Contents
ix
x Preface
1 The story of the Tower of Babel is set out in the Hebrew Bible in the book of Genesis, chapter 11, verses 1–9,
and relates to problems caused by a displeased God introducing the use of several different, rather than one
spoken language, to the confusion of over-ambitious mankind. A depiction of the Tower of Babel that appears
on the thirteenth-century Mappa Mundi in Hereford Cathedral Library is shown in the reproduction above.
Preface xi
However, stern, but sensible, comments from the reviewers of the outline of the
proposed book, plus a great deal more reading of relevant astronomical texts on my
part, has led to a better understanding of why some astronomers would be reluctant
to move away from non-standard units. This applies particularly in the field of
celestial mechanics and stellar dynamics, where the International Astronomical
Union approved units include the astronomical unit and the solar mass. However,
this in itself should not act as a deterrent from adding SI-based units alongside the
special unit used, with suitable error estimates to illustrate why the special unit is
necessary.
In a recent book review in The Observatory, Trimble (2010) admitted to append-
ing an average of about two corrections and amplifications per page in not only the
book she had just reviewed but also in her own book on stellar interiors (Hansen
et al., 2004), and Menzel (1960) completed the preface to his comprehensive work
Fundamental Formulas of Physics by stating: ‘In a work of this magnitude, some
errors will have inevitably crept in.’
Whilst, naturally, I hope that this particular volume is flawless, I must confess
I consider that to be unlikely! The detection and reporting of mistakes would prove
of considerable value and, likewise, comments from readers and users of the book
on areas in which they believe it could be improved would be welcome. My own
experience using various well-known reference works and textbooks, to some of
which I had previously assigned an impossibility of error, was that they all contained
mistakes; some travelled uncorrected from one edition to the next and others in
which correct numerical values or terms in an algebraic equation in an earlier
edition were incorrectly transcribed to a later.
The most radical suggestions in this book are probably: a simple way of describ-
ing and dealing with very large and very small numbers; the use of a number pair of
radians rather than a combination of three time and three angular measures to locate
the position of an astronomical body; and the replacement of the current ordinal
relative-magnitude scheme for assigning the brightness of astronomical bodies by
a cardinal system based on SI units in which the brighter the object the larger the
magnitude.
Writing a book such as this takes time. Time during which new values of astro-
nomical and physical constants may become available. I have referenced the various
sources of constants published before the end of 2010 that were used in the prepara-
tion of tables and in the worked examples presented. Readers are invited to substitute
later values for the constants, as a valuable exercise, in the worked examples should
they so wish.
In conclusion, it is important to bear in mind that the primary purpose of this book
is to act as a guide to the use of SI units in astronomy and not as an astronomical
textbook.
Acknowledgements
2 Please note that theses extracts are reproduced with permission of the BIPM, which retains full internationally
protected copyright.
xii
Acknowledgements xiii
Edinburgh, provided me with COSMOS measuring machine data of the double clus-
ter of galaxies A3266. The desk staff at Victoria University of Wellington library
and the librarian of the Martinborough public library were of great assistance in
sourcing various books and articles. The proofreading was bravely undertaken by
Anne and Eric Dodd. To all these people I express my thanks.
My aim was to write a book that would prove of use to the astronomical commu-
nity and persuade it to move towards adopting a single set of units for the benefit
of all. I hope it succeeds!
1
Introduction
1
2 Introduction
the variable object of interest and a standard non-varying star is plotted against
time or phase (see, e.g., Yang, 2009).
An X-ray survey carried out by Albacete-Colombo et al. (2008) of low-mass
stars in the young star cluster Trumpler 16, using the Chandra satellite, gives
the median X-ray luminosity in units of erg . s−1 .
The integral γ -ray photon flux above 0.1 GeV from the pulsar J0205 + 6449 in
SNR 3C58, measured with the Fermi gamma-ray space telescope, is given in
units of photons . cm−2 . s−1 by Abdo et al. (2009).
These are just a few examples of the many different units used to
specify flux. Radio astronomers and infrared astronomers often use janskys
(10−26 W . m2 . Hz−1 ), whilst astronomers working in the ultraviolet part of
the electromagnetic spectrum have been known to use flux units such as
(10−9 erg . cm−2 . s−1 . Å−1 ) and (10−14 erg . s−1 . cm−2 . Å−1 ). So it would seem
not unreasonable to conclude that whilst astronomers may well be mindful of SI
units and the benefits of unit standardization they do not do much about it.
Among reasons cited in Cardarelli (2003) for using SI units are:
1. It is both metric (based on the metre) and decimal (base 10 numbering system).
2. Prefixes are used for sub-multiples and multiples of the units and fractions
eliminated, which simplifies calculations.
3. Each physical quantity has a unique unit.
4. Derived SI units, some of which have their own name, are defined by simple
expressions relating two or more base SI units.
5. The SI forms a coherent system by directly linking the mechanical, electrical,
nuclear, chemical, thermodynamic and optical units.
A cursory glance at the examples given above shows numerous routes to possi-
ble mistakes. Consider the different powers of ten used, especially by ultraviolet
astronomers. Some examples use wavelengths, some frequencies, and some ener-
gies to define passbands. One uses a form of temperature to record the flux detected.
In short, obfuscation on a grand scale, which surely was not in the minds of the
astronomers preparing the papers. For this book to prove successful it would need
to assist in a movement towards the routine use of SI units by a majority, or at the
very least a large minority, of astronomers.
Descriptions of the base and common derived SI units, plus acceptable non-SI units
and IAU recommended units, are listed in Chapter 2 with Conférence Général des
Poids et Mesures (CGPM) approved prefixes and unofficial prefixes for SI units
with other possible alternatives.
Given the importance of the technique known as dimensional analysis to the
study of units, an entire chapter (3) is allocated to the method, including worked
examples. There are further examples throughout the book that illustrate the value
of dimensional analysis in checking for consistency when transforming from one
set of units to another.
Eight chapters (4–11) cover the seven SI base units plus the derived unit, the
radian. Each includes the formal English language definition published by the
Bureau International des Poids et Mesures (BIPM) and possible future changes
to that definition. Examples of the uses of the unit are given, including transforma-
tions from other systems of units to the SI form. Derived units, their definitions, uses
and transformations are also covered, with suitable astronomical worked examples
provided. Each chapter ends with a summary and a short set of recommendations
regarding the use of the SI unit or other International Astronomical Union (IAU)
approved astronomical units.
The book ends with a chapter (12) on astronomical taxonomy, outlining various
classification methods that are often of a qualitative rather than a quantitative nature
(e.g., galaxy morphological typing, visual spectral classification).
The subject matter of the book covers almost all aspects of astronomy but is not
intended as a textbook. Rather, it is a useful companion piece for an undergradu-
ate or postgraduate student or research worker in astronomy, whether amateur or
professional, and for the writers of popular astronomical articles who wish to link
everyday units of measurement with SI units.
3 See Carroll L. (1965). Through the Looking Glass. In The Works of Lewis Carroll. London: Paul Hamlyn, p. 174.
4 Introduction
volumes I and II of Funk & Wagnalls New Standard Dictionary of the English
Language (1946).
Standard
Any measure of extent, quantity, quality, or value established by general usage and
consent; a weight, vessel, instrument, or device sanctioned or used as a definite
unit, as a value, dimension, time, or quality, by reference to which other measuring
instruments may be constructed and tested or regulated.
The difference between a unit and a standard is that the former is fixed by
definition and is independent of physical conditions, whereas a standard, such as
the one-metre platinum–iridium rod held at the Bureau International des Poids et
Mesures (BIPM) in Sèvres, Paris, is a physical realization of a unit whose length
is dependent on physical conditions (e.g., temperature).
Quantity (Specific)
(1) Physics: A property, quality, cause, or result varying in degree and measurable
by comparison with a standard of the same kind called a unit, such as length,
volume, mass, force or work.
(2) Mathematics: One of a system or series of objects having only such relations,
as of number or extension as can be expressed by mathematical symbols; also,
the figure or other symbol standard for such an object. Mathematical quantities
in general may be real or imaginary, discrete or continuous.
Measurement
The act of measuring; mensuration; hence, computation; determination by judge-
ment or comparison. The ascertained result of measuring; the dimensions, size,
capacity, or amount, as determined by measuring.
The mathematical definition of a quantity Q, is the product of a unit U , and a
measurement m, i.e.,
Q = mU (1.1)
1.3 Definitions of terms (lexicological, mathematical and statistical) 5
Dimension
Any measurable extent or magnitude, as of a line, surface, or solid; especially one of
the three measurements (length, width and height) by means of which the contents
of a cubic body are determined; generally used in the plural. Any quantity, as length,
time, or mass, employed or regarded as a fundamental factor in determining the
units of other physical quantities (see Chapter 3); as, the dimensions of velocity are
length divided by time. The dimension of a physical quantity is the set containing
all the units which may be used to express it, e.g., the dimension of mass is the set
(kilogram, gram, pound, ton, stone, hundredweight, grain, solar mass . . .).
Accurate
Conforming exactly to truth or to a standard; characterized by exactness; free from
error or defect; precise; exact; correct.
Accuracy
The state or quality of being accurate; exactness; correctness.
Precise
Having no appreciable error; performing required operations with great exactness.
Precision
The quality or state of being precise; accuracy of limitation, definition, or
adjustment.
There is a tendency to use accuracy and precision as though they had the same
meaning, this is not so. Accuracy may be thought of as how close the average value
(see below) of the set of measurements is to what may be called the correct or actual
value, and precision is a measure (see standard deviation below) of the internal
consistency of the set of measurements. So if, for example, a measuring instrument
is incorrectly set up so that it introduces a systematic bias in its measurements, these
measures may well have a high internal consistency, and hence a high precision,
but a low accuracy due to the instrumental bias.
Error
The difference between the actual and the observed or calculated value of a quantity.
6 Introduction
Mistake
The act of taking something to be other than it is; an error in action, judgement or
perception; a wrong apprehension or opinion; an unintentional wrong act or step;
a blunder or fault; an inaccuracy; as a mistake in calculation.
1.3.2 Statistical
Statistics is an extensive branch of mathematics that is regularly used in astronomy
(see Wall & Jenkins, 2003). As a very basic introduction to simple statistics, defi-
nitions are given for the terms mean and standard deviation, which are commonly
used by astronomers and an illustration (Figure 1.1) of the Gaussian or normal
distribution curve showing how a set of random determinations of a measurement
are distributed about their mean value.
Mean
Consider a set of N independent measurements of the value of some parameter x
then the mean, or average, value, μ, is defined as:
1
N
μ= xi (1.2)
N
i=1
0.4
0.3
σ σ
f (x)
0.2
0.1
0.0
–3 –1 μ 1 3
x
Standard deviation
Given the same set of N measurements as above, the standard deviation, σ , is
defined as:
1 √
N
σ= (xi − μ)2 (1.3)
N
i=1
Gaussian distribution
For large values of N , the expression describing the Gaussian or Normal distribution
of randomly distributed values of x about their mean value μ is:
1 −1
(x−μ)2
f (x) = √ e 2σ 2 (1.4)
σ (2π)
Figure 1.1 shows the typically bell shaped Gaussian distribution with a mean value,
μ = 0, and standard deviation, σ = 1.
Occasionally published papers may be found that use expressions such as stan-
dard error or probable error. If definitions do not accompany such expressions then
they should be treated with caution, since different meanings may be attributed by
different authors.
his outstretched arm), trading from one village to another could be fraught with
difficulties and even lead to violent altercations.
One of the earliest records of attempted standardization to assist in trade is set
out in the Hebrew Bible in Leviticus, 19, 35–36 (Moffatt, 1950): ‘You must never
act dishonestly, in court, or in commerce, as you use measures of length, weight, or
capacity; you must have accurate balances, accurate weights, and an honest measure
for bushels and gallons.’
Around 2000 years later, King John of England and his noblemen inserted a
clause in the Magna Carta (number 35 on the Salisbury Cathedral copy of the
document) that stated (in translation from the original abbreviated Latin text): ‘Let
there be throughout our kingdom a single measure for wine and for ale and for corn,
namely: the London quarter,4 and a single width of cloth (whether dyed, russet or
halberjet)5 namely two ells within the selvedges; and let it be the same with weights
as with measures.’
It would appear to be very difficult to introduce a new set of standard units
by legislation. Even the French, under Emperor Napoleon I, preferred a mainly
non-decimal system, which had more than 250 000 different weights and measures
with 800 different names, to the elegant simplicity of the decimal metric system.
This preference caused Napoleon to repeal the act governing the use of the metric
system (passed by the republican French National Assembly in 1795, instituting
the Système Métrique Décimal) and allowing the return to the ancien régime in
1812. The metric system finally won out in 1837 when use of the units was made
compulsory. One hundred and sixty years later it was the turn of the British to object
to the introduction of the metric system, despite such a change greatly simplifying
calculations using both distance and weight measurements.
4 The London quarter was a measure that King Edward I of England decreed, in 1296, to be exactly eight striked
bushels, where ‘striked’ implied the measuring container was full to the brim.
5 ‘halberjet’ is an obsolete term for a type of cloth (Funk et al., 1946).
1.5 A brief history of the standardization of scientific units 9
adding prefixes to milliare. The premature death of Mouton prevented him from
developing his work further.
The English architect and mathematician Sir Christopher Wren proposed in 1667
using the length of the seconds pendulum as a fixed standard, an idea that was sup-
ported by the French astronomer Abbé Jean Picard in 1671, the Dutch astronomer
Christiaan Huygens in 1673 and the French geodesist Charles Marie de la Con-
damine in 1746. Neither the milliare nor the length of the seconds pendulum was
chosen to be the standard of length however, with that honour going to a measure-
ment of length based on a particular fraction of the circumference of the Earth.
The metre, as the new unit of length was named, was originally defined as
one ten millionth part of the distance from the North Pole to the Equator along
a line that ran from Dunkirk through Paris to Barcelona. The survey of this line
was carried out under the direction of P. F. E. Méchain and J. B. J. Delambre. Both
were astronomers by profession, who took from 1792 to 1799 to complete the
task. A comparison with modern satellite measurements produces a difference of
0.02%, with the original determined metre being 0.2 mm too short (Alder, 2004). A
platinum rod was made equal in length to the metre determined from the survey and
deposited in the Archives de la République in Paris. It was accompanied by a one-
kilogram mass of platinum, as the standard unit of mass, in the first step towards
the establishment of the present set of SI units. Following the establishment of the
Conférence Général des Poids et Mesures (CGPM) in 1875, construction of new
platinum–iridium alloy standards for the metre and kilogram were begun.
The definition of the metre based on the 1889 international prototype was
replaced in 1960 by one based upon the wavelength of krypton 86 radiation, which
in turn was replaced in 1983 by the current definition based on the length of the path
travelled by light, in vacuo, during a time interval of 1/(299 792 458) of a second.
The unit of time, the second, initially defined as 1/(86 400) of a mean solar
day was refined in 1956 to be 1/(31 556 925.974 7) of the tropical year for 1900
January 0 at 12 h ET (Ephemeris Time). This astronomical definition was super-
seded by 1968, when the SI second was specified in terms of the duration of
9 192 631 770 periods of the radiation corresponding to the transition between two
hyperfine levels of the ground state of the caesium 133 atom at a thermodynamic
temperature of 0 K.
The unit of mass (kilogram) is the only SI unit still defined in terms of a manu-
factured article, in this case the international prototype of the kilogram which, with
the metre, were sanctioned by the first Conférence Général des Poids et Mesures
(CGPM) in 1889. These joined the astronomically determined second to form the
basis of the mks (metre–kilogram–second) system, which was similar to the cgs
(centimetre–gram–second) system proposed in 1874 by the British Association for
the Advancement of Science.
10 Introduction
1. Establish fundamental standards and scales for the measurement of the principal
physical quantities and maintain the international prototypes.
2. Carry out comparisons of national and international prototypes.
3. Ensure the coordination of corresponding measuring techniques.
4. Carry out and coordinate measurements of the fundamental physical constants
relevant to these activities.
1.7.2 Recommendations
For astronomers who do not routinely use SI units, a simple approach to doing
so would be to convert whatever units are being used into their SI equivalents
and prepare research papers, presentations and lecture notes with the final results
given in both forms. This would lead to SI units becoming familiar, acceptable and,
hopefully, the universal system of choice.
2
An introduction to SI units
The name Système International d’Unités (International System of Units), with the
abbreviation SI, was adopted by the 11th Conférence Générale des Poids et Mesures
(CGPM) in 1960.
This system includes two classes of units:
- base units
- derived units,
which together form the coherent system of SI units.
12
2.3 Non-SI units currently accepted for use with SI units 13
Table 2.4. Some SI derived units whose names and symbols include SI derived
units with special names and symbols (BIPM, 2006)
Table 2.6. Non-SI units accepted for use with the International
System whose values in SI units are obtained experimentally
(BIPM, 2006)
are a 102 m2
hectare ha 104 m2
bar bar 105 Pa
angstrom Å 10−10 m
barn b 10−28 m2
in progress due to a lack of suitable theories), up to ∼1041 kg for the mass of the
Milky Way galaxy. In astronomy, the range of official SI prefixes is insufficient to
cover the range of numbers, which has lead to the imaginative invention of many
others (not yet approved by the CGPM).
Whilst being a worthy attempt to produce at least one prefix derived from a lan-
guage originating from each of the six continents there is no obvious connection
between the prefix and the multiplying factor it is meant to represent. A possible
way around this problem is simply to use the straightforward and commonly used
names (e.g., million, billion, trillion, quadrillion etc.) assigned to the numbers. For
numbers smaller than 1 adding ‘th’ after the name (e.g., millionth, billionth, tril-
lionth, quadrillionth etc.) suffices. Table 2.12 sets out a list of multipliers for SI
units that run from 10−48 (one quindecillionth) to 1048 (one quindecillion).
Note that there is a simple relationship between the power, n, to which 1000 is
raised and the Latin prefix of the ending ‘-illion’. In Table 2.12, each line, starting
at the top-left-hand side of the table, is 1000 times smaller than the values in the
next line in the table. Unfortunately, the prefixes tend to become cumbersome and
2.5 Prefixes to SI units 19
the displacement of the value of n by one relative to the Latin prefix offers the
opportunity for mistakes to be made.
angstrom Å 10−10 m
micron μ 10−6 m
fermi 10−15 m
barn b 10−28 m2
cubic centimetre cc 10−6 m3
dyne dyn 10−5 N
erg erg 10−7 J
calorie cal 4.1868 J
bar bar 105 Pa
standard atmosphere atm 101.325 kPa
gal Gal 10−2 m . s−2
eotvos E 10−9 s−2
gauss G 10−4 T
gamma 10−9 T
oersted Oe (1000/4π) A . m−1
2.6.1 Angle
Currently a sexagesimal-based system of units is used to specify the positions of
celestial bodies in astronomy. Declination values are typically given as degrees,
minutes and seconds of arc north or south of the celestial equator, and right ascen-
sion as hours, minutes and seconds of time increasing eastwards from zero at the
intersection point of the celestial equator and the ecliptic, marking the position
of the Sun at the Vernal Equinox (where the declination of the Sun moves from
negative values south of the equator to positive north of the equator). This use of a
unit of time to represent an angle is a possible source of confusion and should be
avoided. Detailed relationships between the SI unit of angle (the radian) and the
various sexagesimal systems are given in Chapter 4. If, for some reason, the radian
is not considered a suitable unit in a particular circumstance, then the degree with
decimal subdivision should be used.
The use of the ‘mas’ meaning milliarcsecond for angular resolution or angular
separation of astronomical objects should be replaced by a more appropriate SI
unit, such as the nrad (nanoradian).
2.6.2 Time
Other than the base SI unit of time, the second, astronomers use longer lengths of
time, such as the minute (60 seconds), the hour (3600 seconds), the day (86 400
seconds) and the Julian year consisting of 365.25 d or 31.557 6 × 106 s. There are,
however, several different kinds of day and year that relate to particular problems
in astronomy (see Chapter 5).
The variability of the Earth’s rotation rate means that time based on that rate
varies with respect to the SI second. Hence sidereal, solar and Universal Time
should be considered as measurements of hour angle expressed in time measure
and not suitable for precise measures of time intervals.
The parsec is a distance (∼3.086 × 1016 m) equal to that at which the astronom-
ical unit subtends an angle of 1 arcsec (π /(180 × 3600) rad).
The light year is the distance (∼9.461 × 1015 m) travelled by light, in vacuo, in
one Julian year.
The astronomical unit of mass is the solar mass (∼1.989 1 × 1030 kg) denoted
by M .
2.6.4 Wavenumber
The reciprocal wavelength or wavenumber is used mainly by infrared astronomers
and is normally based on the cm−1 . If used it should be in the SI unit form of m−1 ,
but in either case the unit must be given as it is not dimensionless.
2.6.5 Magnitude
Magnitude may be defined as the ratio of the logarithm of the signal strength of
the celestial object of interest to that of a standard star or object. As such it is a
dimensionless quantity. Some magnitude scales have been calibrated in terms of SI
units (see Chapter 8).
2.7.2 Recommendations
Astronomers should follow the recommendations proposed by the IAU with par-
ticular reference to dropping the use of the cgs-based units so prevalent at present.
Given that the current set of official SI prefixes is inadequate, consistency may be
maintained by quantities being presented as the product of the measurement times
the basic unit (i.e., the unit without any prefix). For example, the astronomical unit
should be given as 1.496 × 1011 m or 1.496 d 11 m and not 149.6 Gm. Examples of
all three forms are given in different tables throughout the book.
3
Dimensional analysis
By way of example, the derived quantity force, P , has the derived SI unit
kg . m . s−2
24
3.2 Dimensional equations 25
Table 3.2 gives examples of the dimensions of some derived SI units. Note that
both the radian and steradian are examples of dimensionless units, as in both cases
the value of the dimension is 1. Some derived quantities have the same dimension,
e.g., the hertz and the bequerel both have the dimension [T ]−1 .
v = d/t (3.4)
so the equation is homogeneous. Note that the requirement is for the dimensions
to be consistent, so that any set of consistent units within a particular dimen-
sion may be used and converted to any other by means of a constant factor. If
physical quantities have the same dimensions, they may only be combined by
addition or subtraction. For example, [L] + [L] or [L] − [L] but not [L] × [L]
(which is a measure of area), nor [L]/[L] (which is a ratio and a dimensionless
number).
26 Dimensional analysis
A dimensional equation that is to be used for converting one set of units within a
dimension to another (e.g., inches to metres) must also include a conversion factor,
k (to convert inches to metres, k = 0.0254). So Equation (3.1) above would be
rewritten in the more general form:
7
dim(X) = [ki dim(Bi )]αi (3.6)
i=1
3.2 Dimensional equations 27
where the ki relate to the conversion factors required to convert the units in which
the measurements were carried out to base units (e.g., the multiplicative conversion
factor from inches to metres (Kaye & Laby, 1959) is 1/39.370 147 = 0.025 399 96).
Quite evidently, even the application of multiplicative scaling factors would not
make the two monochromatic flux units compatible in a dimensional sense. The
difference in derived dimensions in this case is due to the different variable used to
specify the bandwidth of the observation. For the jansky, the bandwidth is measured
in hertz ([T ]−1 ) and for the TD1 unit the bandwidth is measured in angstroms ([L]).
To compare the measures directly, the wavelength bandwidth in angstroms needs to
be converted to a frequency bandwidth in hertz. The algebraic relationship between
frequency (ν) and wavelength (λ) is given by:
c
λ= (3.11)
ν
where c is the velocity of light. The relationship between the bandwidths in wave-
length form (
λ) and frequency form (
ν) is obtained by differentiating Equation
(3.11) to give:
c
λ = − 2
ν (3.12)
ν
The dimensional form of (c/ν 2 ) is ([L] . [T ]−1 /[T ]−2 ) or [L] .[T ], so multiply
dim(fλ(TD1) ) by [L] . [T ] to obtain:
The negative sign in Equation (3.12) is cancelled because the value for the fre-
quency decreases with increasing wavelength, so if
λ is positive then
ν must
be negative. Now all that has to be done is evaluate the multiplying factor to convert
the TD1 unit to janskys.
Example: how to convert TD1 cgs flux units to janskys (derived SI unit) for
the B5V star HD74071
The ESRO satellite TD1 measured ultraviolet flux of HD74071 is fλ = 24.73 ×
10−11 erg . s−1 . cm−2 . Å−1 at a wavelength of λ = 1565 Å with a bandwidth of
λ = 330 Å. The frequency ν, corresponding to the wavelength 1565Å derived by
using equation (3.11) is 1.916 × 1015 Hz, with a frequency bandwidth derived from
Equation (3.12) of 4.043 × 1014 Hz.
The TD1 flux unit may be converted to SI base units, with appropriate factors, as:
−1
TD1 flux unit = erg . s−1 . cm−2 . Å
≡ (10−7 ) W . (10−2 )−2 . m−2 . (10−10 )−1 . m−1
= 107 . W . m−2 . m−1 (3.16)
3.3 Summary and recommendations 29
Note that the powers to which the conversion factors are raised match those of
the SI units. These individual conversion factors are multiplied together to form a
single conversion factor for the entire derived SI unit. Now convert the bandpass
from wavelength,
λ, to frequency,
ν, to give the TD1 units in janskys:
λ
fν(TD1) = fλ(TD1) . . 1026 (3.17)
ν
2.998 × 108
= 107 . (24.73 × 10−11 ) . . 1026
(1.916 × 1015 )2
= 20.2 Jy
So the measured monochromatic flux density from the star HD74071 may be
given as either 24.73 × 10−11 erg . cm−2 s−1 . Å−1 or 20.20 Jy.
3.3.2 Recommendations
No matter how much care is taken in the transformation of units, mistakes do occur.
Particular problems are associated with the powers to which the base units are raised
and the signs and values of powers of ten. So always check any unit transformation
that has been carried out and when satisfied that it is correct, check it again!
4
Unit of angular measure (radian)
9 See www-wfau.roe.ac.uk/sss
30
4.2 Commonly used non-SI units of angular measure 31
B
O
2π
1 = = 0.000 290 888 209 rad
60 × 360
2π
1 = = 0.000 004 848 137 rad (4.1)
60 × 60 × 360
Similarly, there are 2π rad in 24 h or:
2π
1h= = 0.261 799 387 799 149 rad
24
2π
1m= = 0.004 363 323 129 985 rad
60 × 24
2π
1s= = 0.000 072 722 052 166 rad (4.2)
60 × 60 × 24
Example: calculate the position of the bright northern star Capella in radians
Capella has the following coordinates in the FK5 catalogue for equinox (J2000.0)
and epoch (J2000.0):
where α is the right ascension and δ is the declination of Capella (see Section 4.3.2
below). The right ascension of Capella in radians to six decimal places is:
so
α = 1.381818 rad
and the declination of Capella in radians to six decimal places is:
45 × 0.017 453 + 59 × 0.000 290 + 52.8 × 0.000 005
so
δ = 0.802 816 rad
If, instead of measuring declination from the celestial equator (CE) northwards
from 0 rad to + π2 rad and southwards from 0 rad to − π2 rad, we measure it from
the south celestial pole (SCP) as 0 rad to the north celestial pole (NCP) as +π rad
(see Zacharias et al., 2000, in which declination is given as a south polar distance
in units of milliarcseconds (mas)), then the declination in radians becomes:
π
δ(SCP0 ) = δ(CE0 ) + (4.3)
2
Using the south celestial pole zero point (SCP0 ), the revised declination of Capella
becomes 2.373 612 rad. Expressed in milliradians (mrad), the coordinates of Capella
become:
α = 1381.817 893 mrad
δSCP0 = 2373.612 876 mrad
Given that, e.g., The Astronomical Almanac lists its bright star positions to the
nearest tenth of a second of time and nearest whole second of arc, what number of
decimal places would yield a similar precision using milliradians? The answers are,
approximately, three decimal places for right ascension and two for declination. So
for the worked example using Capella above, a catalogue position in milliradians
with similar precision to that given in The Astronomical Almanac would be:
α = 1381.818 mrad
δSCP0 = 2373.61 mrad
By way of another example, the equatorial coordinates listed in the table of 86
bright stars given in the Handbook of the British Astronomical Association for 2008
were converted to milliradians and the V magitudes to janskys and ln(janskys). The
following equation was used to convert from magnitudes to monochromatic flux
densities10 in janskys (10−26 Wm−2 Hz−1 ):
VJ = 3600(10−0.2V ) (4.4)
where the conversion factor of 3600 is from an absolute calibration of a zero mag-
nitude star in the Johnson V band given by Bessell (1992), V is the listed value
of the apparent V magnitude and VJ is the apparent monochromatic flux density
of the star in janskys. A plot of the stars in Table 4.1 is shown in Figure 4.2 using
Mollweide’s projection.11
Obvious advantages of the SI radians angular measurement system are that only
two numbers are required to specify the location of the celestial body, rather than
six, and that all the declination values are positive when the declination of the South
Celestial Pole is set equal to 0 mrad.
NCP
3 Polaris
URSA MAJOR
2
Arcturus
Altair ORION
equator
Υ
Sirius
6 5 4 3 2 1
1
Canopus
CRUX
SCP
Figure 4.2. A star chart for the 86 bright stars in Table 4.1, plotted in radians,
with declination measured from 0 at the south celestial pole (SCP) northwards
to π rad at the north celestial pole (NCP) and right ascension increasing in an
easterly direction from 0 at the First Point of Aries, ϒ, to 2π rad. Right ascensions
in radians from 1 rad to 6 rad are printed to the left of the relevant circle and
declinations are printed from 1 rad to 3 rad immediately adjacent to the relevant
circle. The dashed declination lines are at the 0.5-rad points. The size of the star
images are proportional to ln(fVν ) in janskys.
a
C
is called the terrestrial equator. Other great circles that are at right angles to the
terrestrial equator and pass through the north and south poles are called terrestrial
meridians.
Geographical longitude is the angle between a terrestrial meridian circle and
an arbitrarily chosen reference meridian circle called the prime meridian. This
great circle passes through a telescope known as the Airy transit circle located
at the Royal Observatory, Greenwich, in the United Kingdom. All longitudes are
commonly measured from this prime meridian, either east or west from 0◦ to 180◦ .
Geographical latitude is the angular distance normally measured north or south
of the terrestrial equator from 0◦ to 90◦ along the meridian circle through the point
on the Earth’s surface of interest (see Figure 4.4).
Using the radian equivalents given in Section 4.2.1 above, it is a trivial exer-
cise to convert latitudes (φ) and longitudes (λ) from degrees, minutes and seconds
to radians. Retaining the prime meridian in Greenwich but measuring longitude,
increasing eastwards, from 0 to 2π rad, and measuring latitude from 0 at the south
pole to π rad at the north pole, we may rewrite the longitude and latitude of Mauna
Kea Observatory in the Hawaiian Islands (remembering that its given sexagesi-
mal longitude is in the western hemisphere, so λ = 360 − λ) from 155◦ 28 18W;
19◦ 49 36 N to:
λ = 3586.157 103 mrad φ = 1916.836 940 mrad
Note that three decimal places corresponds to locating the observatory on the surface
of the Earth to approximately ±7 m if the Earth were a perfect sphere.12
12 The length of an arc of 1 radian is equal to the radius of the Earth, 6738.136 km at the Equator according to
Cox (2000), so an arc length of 1 mrad = 6.738 136 km and 0.001 mrad = 6.738 m.
38 Unit of angular measure (radian)
NP
n
idia
mer
P
p r i me
φ
G λ R
equat
or
φSP
SP
Figure 4.4. The geographical coordinates of the point P are longitude λ (the thick
grey arc GR) and latitude, either φ (the thick black arc PR), when measured north
or south from the equator, or φSP (the thick black arc SPRP) when measured from
the south pole northwards. The labelled points are: NP (north geographic pole),
SP (south geographic pole), G (the intersection of the prime meridian with the
equator), and R (the point of intersection of the great circle through the poles and
the point P with the equator).
W
z
O
h
S N
A
horizon
H E
Figure 4.5. Altazimuth coordinates: azimuth angle, A is the angle NOH measured
(dark grey arc) from north (N) through east (E) to the intersection point on the
horizon (H) of the great circle through the zenith point (Z) and the star (P). The
altitude, h, is the angle POH (thick black arc). The zenith distance, z, is the angle
ZOP (light grey arc).
between the horizon and the body measured along the vertical circle through the
zenith point. Hence altitude and zenith distance would lie between 0 and 500π mrad.
The time and date of the observation and the location at which the observation was
made must also be given, as both altitude and azimuth vary with time, date and
geographical position.
equa Υ
tor
α B
δ SCP
SCP
Figure 4.6. Equatorial coordinates: right ascension, α, measured from the first
point of Aries (ϒ) to point B in an easterly direction (thick grey arc ϒB), decli-
nation, δ, measured from either the celestial equator at point B (thick black arc
BA) or δSCP , measured from the south celestial pole northwards through B to
A along the great circle through the celestial poles and point A (thick black arc
SCPBA).
β
D
equ
ato λ
r β SEP
ϒ
ε E
ecliptic
SEP
Figure 4.7. Ecliptic coordinates: celestial longitude, λ, measured from the first
point of aries (ϒ) eastwards along the ecliptic (thick grey arc ϒD) and celestial
latitude, β, measured either north or south of the ecliptic along a great circle
including the north and south ecliptic poles and the celestial object of interest, S,
(thick black arc DS) or northwards from the south ecliptic pole (thick black arc
SEPEDS). The angle DϒE, is the obliquity of the ecliptic – the angle between
the plane of the Earth’s equator and its orbit about the Sun.
of observations a fundamental epoch (the precise moment in time for which the
observations are specified) is adopted, to which all appropriate observations are
reduced. The resulting coordinates are referred to as mean coordinates. Star posi-
tions are listed in catalogues and plotted on star charts and atlases reduced to a
fundamental epoch. It is usual for a particular fundamental epoch to be used for
around 50 years; that currently in use (2011) is 2000.0.
These equatorial coordinates were calculated from the original epoch 1950.0
values agreed to by the International Astronomical Union when setting up the new
system of galactic coordinates in 1958 (Blauuw et al., 1960).
The values of the right ascension (α0 ) and galactic longitude (l0 ) of the ascending
node of the galactic plane on the J2000 equator and the inclination of the galactic
equator to the celestial equator (γ ), derived from Cox (2000), are:
α0 = 4936.829 mrad
l 0 = 574.770 mrad
γ = 1097.319 mrad
Although the coordinate frame in the galactic coordinate system does change
with time, the effect is extremely small in comparison with either the equatorial or
ecliptic coordinate systems. The galactic coordinate system is of particular value
in galactic structure studies.
4.3 Spherical astronomy 43
S
NGP
r
ato
b
qu
ator
equ
ic e
bSGP
t
lac
l
ga
γ
l0
galactic SGP
O
centre
Figure 4.8. Galactic coordinates: galactic longitude, l, measured along the galactic
equator from the galactic centre (dark grey arc OLE), galactic latitude, b, measured
either north or south of the galactic equator (thick black arc ES) or from the south
galactic pole northwards, bSGP , (thick black arc SGPES). The longitude of the
ascending node of the galactic plane, l0 , is the dark grey arc OL and the inclination
of the galactic equator to the celestial equator is the angle γ .
Equatorial to ecliptic
To convert the equatorial coordinates (α, δSCP ), measured in radians, using the south
celestial pole as the zero point of the declination measures, to ecliptic coordinates
(λ, β), also measured in radians, with the zero point of celestial latitude being the
ecliptic plane and being the obliquity of the ecliptic, the following expressions
may be used:
π
δ = δSCP −
2
sin β = sin δ cos − cos δ sin α sin
cos δ sin α cos + sin δ sin
sin λ = (4.8)
cos β
cos δ cos α
cos λ =
cos β
cos δ sin α cos + sin δ sin
tan λ =
cos δ cos α
π
βSEP = β +
2
To assign the correct value for λ in four quadrants, the following conditions for
the sign of each of the values of sine, cosine and tangent determine which expression
is to be used to calculate the value of λ. (∧ is the Boolean symbol for ‘and’.)
sin−1 λ if sin λ ≥ 0 ∧ cos λ ≥ 0 ∧ tan λ ≥ 0
π − sin−1 λ if sin λ ≥ 0 ∧ cos λ < 0 ∧ tan λ < 0
λ= π − sin−1 λ if sin λ < 0 ∧ cos λ < 0 ∧ tan λ ≥ 0 (4.9)
2π + sin−1 λ if sin λ < 0 ∧ cos λ ≥ 0 ∧ tan λ < 0
Using as an example, the equatorial coordinates of the star Capella given above
in radians, (α, δSCP ) = (1.381 817, 2.373 61), may be transformed to ecliptic coor-
dinates using equations (4.8) and conditions (4.9) with = 0.409 092 6 (value at
J2000.0) to yield: (λ, β) = (1.428 690, 0.399 06), where the celestial latitude is given
with reference to the ecliptic; relative to the south ecliptic pole, the value would be
βSEP = 1.979 889, obtained simply by adding π /2 to β.
There is perhaps a stronger case with ecliptic latitude than with equatorial dec-
lination to present the value of the coordinate with reference to the ecliptic plane,
as this is of physical importance when considering Solar System bodies.
( l , b ), also measured in radians, with the zero point of galactic latitude being the
galactic equator and γ being the angle of inclination between the galactic plane
and the equatorial plane, the following expressions may be used:
π
δ = δSCP −
2
sin b = sin δ cos γ − cos δ sin( α − α0 ) sin γ
cos δ sin( α − α0 ) cos γ + sin δ sin γ
sin (l − l0 ) = (4.10)
cos b
cos δ cos ( α − α0 )
cos (l − l0 ) =
cos b
cos δ sin( α − α0 ) cos γ + sin δ sin γ
tan (l − l0 ) =
cos δ cos( α − α0 )
π
bSGP = b +
2
where α0 is the right ascension of the galactic centre at J2000.0 and l0 is the
galactic longitude of the point of intersection of the celestial equator and the galactic
equator, known as the longitude of the ascending node (i.e., where b = 0 mrad or
bSGP = 1570.796 (= π/2) mrad).
To assign the correct value for l in four quadrants, the following conditions for
the sign of each of the values of sine, cosine and tangent determine which expression
is to be used to calculate the value of l. (∧ is the Boolean symbol for ‘and’.)
sin−1 (l − l0 ) if sin (l − l0 ) ≥ 0 ∧ cos (l − l0 ) ≥ 0 ∧ tan (l − l0 ) ≥ 0
π − sin−1 (l − l0 ) if sin (l − l0 ) ≥ 0 ∧ cos (l − l0 ) < 0 ∧ tan (l − l0 ) < 0
l − l0 =
π − sin−1 (l − l0 ) if sin (l − l0 ) < 0 ∧ cos (l − l0 ) < 0 ∧ tan (l − l0 ) ≥ 0
2π + sin−1 (l − l0 ) if sin (l − l0 ) < 0 ∧ cos (l − l0 ) ≥ 0 ∧ tan (l − l0 ) < 0
(4.11)
thence
l = (l − l0 ) + l0 (4.12)
Equation (4.12) can produce galactic longitudes greater than 2π , which conditional
statement (4.13) removes.
l if l < 2π
l= l − 2π if l ≥ 2π (4.13)
Using once again as an example the equatorial coordinates of the star Capella
given above as (α, δ) = (1.381 817 rad, 2.373 61 rad), which may be transformed
to galactic coordinates using Equations (4.10) and conditions (4.11) with γ =
1.097 319 rad (value at J2000.0) to yield: (l, b) = (2.837 704 rad, 0.079 699 rad),
46 Unit of angular measure (radian)
where the galactic latitude is given with reference to the galactic equator. Relative
to the south galactic pole the value would be bSGP = 1.650 496 rad, obtained simply
by adding π/2 to b.
The use of galactic coordinates is of greatest value when studying the distribution
of galactic objects with reference to the galactic plane and centre.
−→
AB = cos−1 (sin δA sin δB + cos δA cos δB cos (αB − αA )) (4.14)
P
α π/2 − δ B
Β −α
Α
π/2 − δA
B
A r
equato
Table 4.2. Further examples of radian measures: apparent sizes and separations
Sexagesimal Radian
Celestial object system system Notes, Reference
Comet Tebbutt (1861 II) >100◦ >1.75 rad length of tail Chambers (1889)
LMC 10◦ 45 .7 187.83 mrad galaxy Cox (2000)
M31 3◦ 10 .5 55.41 mrad galaxy Cox (2000)
Pleiades 2◦ 34.91 mrad galactic cluster Cox (2000)
Sun 31 59 .26 9.30 mrad star Cox (2000)
Moon 31 05 .2 9.04 mrad satellite Cox (2000)
Helix nebula 12 3.49 mrad planetary nebula Cox (2000)
ω Cen 8 .36 2.43 mrad globular cluster Cox (2000)
Jupiter 48 .9 237.07 μrad planet White (2008)
Saturn’s rings 44 .9 217.68 μrad planetary rings White (2008)
α Cen 7 .53 36.51 μrad double star separation White
(2008)
Mira 0 .060 290.89 nrad stellar diameter Karovska &
Sasselov (2001)
Aldebaran 0 .020 96.96 nrad stellar diameter Harris et al.
(1963)
Aur 0 .00227 11.01 nrad stellar diameter Stencel et al.
(2008)
Example: determine the plate scale of the AAO Schmidt telescope (formally
known as the UK Schmidt telescope)
The measured distance r, on an AAO Schmidt telescope photographic plate,
between Ori and ζ Ori is:
r = 72.654 mm
48 Unit of angular measure (radian)
The angular distance a between the two stars is calculated using Equation (4.14),
substituting for the spherical coordinates, in radians:
4.5 Steradian
The steradian is the derived SI unit of solid angle that has its apex at the centre of
a sphere and subtends an area at the surface of the sphere equal to the square of the
radius of the sphere.
It is a dimensionless quantity with the symbol sr, which may be expressed in
terms of SI base units as ([L]2 . [L]−2 ).
The number of steradians on the entire sky is given by the following expression
(4.15), where α is right ascension measured from 0 rad at the first point of Aries
increasing eastwards and δSCP is declination measured northwards from 0 rad from
the south celestial pole:
2π π
A= sin δSCP d δSCP d α = 4π (4.15)
0 0
4π = 12.566 371
A = π(4.65)2
= 67.93 mrad 2 or 67.93 μsr (4.18)
50 Unit of angular measure (radian)
For the Moon, the mean apparent angular radius derived from Table 4.2 is 4.57
mrad and hence its area is
Example: determine the visible angular area of the planet Venus on 2009
February 28
On 2009 February 28, the angular radius of the planet Venus was 107.6 μrad and
the percentage, p, of the disc illuminated by the Sun was 20%. Hence, the angular
area of the illuminated portion of the disc was:
AVenus = p π(107.6)2
= 7278 μrad 2 or 7 278 psr (4.20)
Example: determine the angular area of the disc of the F supergiant star
Aurigae
The recently measured radius (Stencel et al., 2008) of the F-type supergiant pri-
mary of the eclipsing binary star Aurigae determined using the Palomar Testbed
Interferometer was 5.5 nrad. Assuming a uniform stellar disc, the corresponding
apparent angular area of the Aur primary is:
AAur = π (5.5)2
= 95 nrad 2 = 95 asr (attosteradians) (4.21)
Example: determine the angular area of the Local Group Galaxy M31
The measured angular diameter of M31 is given (Cox, 2000) as 8.7 arcmin or
522 arcsec within the boundary set by the B = 25 mag . arcsec−2 isophote, and the
axial ratio ab as 0.32. The area (πab) of a fitted ellipse with semi-major axis a and
semi-minor axis b is:
A = π ab
A = 6.848 × 104 arcsec2
= 1.610 × 10−6 sr (4.22)
= 1.610 μsr
= 1.610 mrad 2
To convert the isophote unit from magnitudes per arcsecond squared to janskys
per steradian, the following method may be used: the B = 25 mag . arcsec−2 is
10−10 times fainter than the B = 0 mag . arcsec−2 isophote. B = 0 is equivalent to a
4.5 Steradian 51
4 × 10−7
= 17 018 Jy . sr −1
2.350 443 × 10−11
= 1.701 8 × 10−8 Jy . psr −1 (4.23)
= 1.701 8 Jy . μrad −2
So the apparent angular area of M31 to the limiting isophote of 17018 Jy . sr−1 is
1.610 × 10−6 sr.
For each image detected by COSMOS, 13 parameters relating to the image centroid,
its size, brightness, orientation and shape were measured, which enabled separate
lists of galaxy images, star images and non-standard images to be compiled. In total,
4328 images were measured, of which 1256 were classified as galaxies and three
as bright galaxies. The field was divided into a grid of 256 (16 × 16) equal-sized
(1 mrad × 1 mrad = 1 μsr) areas, and the number of galaxies in each counted (see
Table 4.3). These numbers were processed by a contour-plotting routine, the output
of which is shown in Figure 4.10.
The mean of the cell count numbers and the standard deviation about that mean
are 4.3 ± 3.9 galaxies per cell, the median cell count is 3.5 galaxies per cell and the
mode is 2 galaxies per cell.
52 Unit of angular measure (radian)
Table 4.3. Cell counts (per mrad2 ) for galaxies in the neighbouring field and
cluster of galaxies A3266
0 4 1 1 0 3 4 6 2 0 3 3 6 3 2 6
5 4 6 4 0 3 2 2 4 2 2 1 8 6 4 2
3 1 4 2 0 1 1 5 4 2 0 8 4 11 17 3
2 2 2 2 3 2 1 4 2 1 3 4 4 4 6 18
1 1 1 7 3 3 5 8 4 4 2 12 10 3 7 1
2 3 3 1 6 2 7 5 8 8 11 5 9 8 2 0
4 3 7 2 4 7 7 16 7 5 7 2 6 8 3 2
2 0 3 5 1 9 8 27 20 8 6 7 13 9 0 1
3 3 3 5 3 1 5 13 27 12 5 4 6 3 6 4
4 3 0 2 4 7 8 8 17 7 17 2 2 6 4 4
1 4 1 2 2 2 6 5 9 5 5 7 9 5 2 3
5 2 3 3 0 3 3 3 7 9 6 4 3 4 2 4
2 1 1 4 1 5 2 3 2 8 1 6 5 6 1 5
5 3 2 1 4 5 0 5 6 4 3 2 4 1 2 1
4 3 6 4 1 4 2 1 2 4 2 3 1 1 3 1
4 4 2 0 2 5 0 5 5 4 0 2 4 3 2 5
10
.8
5
8
10.
5.4
10.
8
.8
10
5.4
Δδ mrad
.8
10 16
.6
.2
21
0
21.6
4
5.
5.4
–5 5. 5.4
5.4
–5 0 5
Δα mrad
Figure 4.10. Number of galaxies per microsteradian, from data obtained using the
COSMOS measuring machine, of the cluster of galaxies A3266 and its surrounding
area from which all the star images have been removed.
Figure 4.10 clearly shows that A3266 is in fact two adjacent clusters of galaxies
and not just one, which Henriksen & Tittley (2002), using the X-ray space-based
CHANDRA Observatory, have recently shown to be in the process of merging.
4.6 Summary and recommendations 53
4.6.2 Recommendations
The ease with which the location of points on the terrestrial and celestial spheres
may be made is mentioned, with a pair of numbers replacing the current six. The step
from sexadecimal to radian measure is considered too large by some astronomers
so, as a first step in this process, using decimal degrees rather than degrees, minutes
and seconds may make the final changeover to radians easier.
5
Unit of time (second)
The general idea, relation, or fact of continuous or successive existence; or the abstract con-
ception of duration as limitless, capable of division into measurable portions, and essentially
comprising the relations of present, past and future.
54
5.3 Systems of time or time scales 55
UT, Universal Time, based on the mean solar day, is the time system used for all civil
timekeeping. UT0 is rotational time for a particular location on the Earth, uncor-
rected for shifts in longitude due to polar motion. UT1 is the rotational time corrected
for such shifts, though it is still non-uniform, being subject to the irregularities in
the Earth’s rotation.
UTC, Coordinated Universal Time, is the time scale distributed by means of radio
time signals, satellites, radio and TV broadcasts. UTC differs from TAI by an
14 www.usno.navy.mil/USNO/time/master-clock/systems-of-time
15 See http://tycho.usno.navy.mil/clockdev/cesium.html
56 Unit of time (second)
integral number of seconds and is kept to within 0.9 s of UT1 by means of the
irregular introduction of integer leap seconds. UTC has replaced GMT (Greenwich
Mean Time) as an international time standard.
TT or TDT, Terrestrial Dynamic Time, is the idealized time based on the rotation
period of the geoid. The time unit is 1 day consisting of 86 400 s in SI units. It is
approximately related to TAI by the expression:
TCG = TDT + 6.969 290 4 × 10−10 (JD − 2 443 144.5) × 86 400 s (5.5)
where JD is the julian day number equal to the number of days that have elapsed
since Greenwich Mean Noon on 1 January 4713 BCE, Julian proleptic calendar
(the Julian Date is the julian day plus the elapsed time since the preceding noon).
TCB = TDB + 1.550 506 × 10−8 (JD − 2 443 144.5) × 86 400 s (5.6)
ST, Sidereal Time, is defined to be the hour angle of the First Point of Aries, ϒ,
the point of intersection of the celestial equator and the ecliptic (see Chapter 4). In
other words, the right ascension in transit over the prime meridian (the north–south
line) at a given terrestrial longitude (λ). The relationship between the local sidereal
time (LST) and the Greenwich sidereal time (GST) is:
where λ is measured eastwards from Greenwich. The local hour angle (LHA) of a
celestial object S at right ascension, αS , is:
The date of any such measurement of the local hour angle has to be given, as the
location of ϒ is not fixed due to the effects of precession and nutation.
MJD, Modified Julian Date, which begins and ends at midnight, is defined in terms
of the julian day number (see above) as:
1 minute = 60 s
1 hour = 60 m = 3 600 s
1 day = 24 h = 1 440 m = 86 400 s
−33 s from 2006 January 01, 0 h UTC to 2009 January 01, 0 h UTC;
−34 s from 2009 January 01, 0 h UTC until further notice.
16 See http://tycho.usno.navy.mil/bulletinc2008.html
58 Unit of time (second)
5.3.6 Relationships between mean solar time and mean sidereal time
The mean sidereal rotation period of the Earth is equivalent to (The Astronomical
Almanac, 1995):
86 164.090 54 s (SI units) or
23 h 56 m 04.090 53 s (mean solar time) or
0.997 269 566 33 d (mean solar time)
One mean solar rotation period of the Earth is equivalent to:
1.002 737 909 35 d (mean sidereal time) or
24 h 03 m 56.555 37 s (mean sidereal time)
These values relate to the year 1995.
The lengths of some other types of year are given in Table 5.2.
2. Time of the day uses the 24-hour system in the form hh:mm:ss starting with 00 for
the first hour after midnight and ending with 23 as the last before the following
midnight. Hence, 21:56:16 is equivalent to 9 h 56 m 16 s p.m. or 3 m 44 s before
10 p.m.
3. The representation of both date and time as a single composite number uses the
capital letter T to separate the date component from the time component.
60 Unit of time (second)
The SI derived unit of angular acceleration has a dimension of [1] . [T ]−2 , its unit
is the radian per second per second and its symbol is rad . s−2 .
Table 5.3. Rotation and revolution periods and rates for the planets and some
dwarf planets in the Solar System
Table 5.3 gives the sidereal period of revolution in julian years, the sidereal rotation
period in days, the mean orbital angular velocity in nrad . s−1 and the rotational
angular velocity in μrad . s−1 .
The negative values for the mean rotational periods and angular motions of
the planets Venus and Uranus and the dwarf planet Pluto are due to their retro-
grade motions. The sources used for Table 5.3 were: Cox (2000); The Astronomical
Almanac (1995); Duffard et al. (2008) and the JPL Small-Body Database Browser.18
18 See http://ssd.jpl.nasa.gov/sbdb.cgi
62 Unit of time (second)
S1
NCP
μ
θ
S0
μδ μα cos δ
equato ϒ
r B
α
δSCP
SCP
Figure 5.1. The proper motion, over a period of time t years, of the celestial body
at S0 , with equatorial coordinates (α0 , δ0 ), takes it to position S1 with coordinates
(α1 , δ1 ).
equatorial coordinates (α1 , δ1 ) relative to the Sun. The components of the annual
proper motion in the equatorial coordinate system are (μα , μδ ) where:
α1 − α0
μα =
t
and
δ1 − δ 0
μδ = (5.13)
t
and the magnitude, μ, is given by:
μ = μ2α cos2 δ + μ2δ (5.14)
the factor cos δ being due to the radius of the small circle of declination being
a function of δ. The position angle θ, as shown in Figure 5.1, is related to the
proper-motion components by:
μ cos θ = μδ
μ sin θ = μα cos δ (5.15)
year (mas . y−1 ) and by using radio interferometry, the microarcsecond per Julian
year (μas . y−1 ).
π
1 arcsec . y−1 = rad . y−1
180 × 3 600
= 4.848 136 811 × 10−6 rad . y−1 (5.19)
1 arcsec . (century)−1 = 4.848 136 811 × 10−6 rad . (century)−1 (5.20)
Table 5.4 gives the proper motions of the ten brightest stars in the night sky,
taken from values obtained by the HIPPARCOS astrometric satellite and listed on
the ESO website database.19 The basic SI unit of angular velocity (column 3 in Table
5.4) could be rendered in alternative forms as, e.g., −83.88 frad . s−1 (femtoradians
per second) or −83.88 quadrillionths rad . s−1 or −8.388 d 14 rad . s−1 , which is
equivalent to −8.388 × 10−14 rad . s−1 .
19 See http://archive.eso.org/skycat/servers/ASTROM
64 Unit of time (second)
Table 5.4. Proper motions of the ten brightest stars in the night sky
Star name μα cos δ μδ mas . y−1 μα cos δ μδ rad . s−1 μα cos δ μδ nrad . y−1
Table 5.5. HD73904, a member of the galactic cluster IC2391; data from the
revised HIPPARCOS/TYCHO catalogue by van Leeuwen (2007)
α δ μα μδ
Unit HD HIP h m s ◦ mas . y−1 mas . y−1
ID number number mrad mrad nrad . y−1 nrad . y−1
IAU 73904 42374 08 38 23.94 −53 43 18.6 −23.49 21.86
SI 2261.942 −937.623 −113.88 105.98
Table 5.6. A partial entry for the star id13222 from the OGLE-II proper-motion
catalogue by Sumi et al. (2004)
Unit α δ μα μδ μl μb
ascension and declination are presented in radians, though not the proper-motion
components.
On a much larger scale is the USNO B1.0 all sky catalogue of magnitudes,
colours, positions and proper motions for more than one billion celestial objects
to a limiting monochromatic flux density of 14.3 μJy in the V waveband (V mag-
nitude ∼21).20 This catalogue (Monet et al., 2003) was produced from automated
measuring-machine scans of photographic plates taken for the POSS (Palomar
Observatory Sky Survey) and SERC-I, the UK Science and Engineering Research
Council’s first Southern Sky Survey. The proper motions are given in units of
arcsec . y−1 with a precision of 0.002 arcsec . y−1 .
By way of comparison, the equatorial coordinates of HD73904 given in the
USNO-B catalogue are (08h 38m 23s .91; −53◦ 43 18 .39) and its proper-motion
coordinates (−26, 22) mas . y−1 , which are the equivalent in SI units of (2261.940,
−937.622) mrad and (−126.05, 106.66) nrad . y−1 .
An example of a proper-motion catalogue derived from sequential observa-
tions over a period of three years, obtained primarily to search for stellar and
exoplanetary microlensing, is that published by Sumi et al. (2004). Their cata-
logue, covering 49 selected fields, totalling approximately 0.003 4 sr (11 square
degrees) of the galactic bulge surrounding the galactic centre, lists over 5 million
stars with I-band monochromatic flux densities between ∼100 mJy and ∼150 μJy
(11 < I < 18). Each entry in the catalogue includes an internal identity number, mean
proper motion components in mas . y−1 for both equatorial and galactic coordinates,
with standard deviations about the mean values, right ascension and declination
(J2000.0) in decimal degrees, V magnitudes and (V − I) colours.
A partial entry from the proper-motion catalogue with angular measures as
published and as converted to SI units is given in Table 5.6.
20 See http://www.nofs.navy.mil/data/fchpix/
66 Unit of time (second)
5.6.1 Nucleocosmochronology
Nucleocosmochronology is defined by Schramm (1990) as ‘. . . the use of the abun-
dance and production ratios of radioactive nucleides coupled with information
on the chemical evolution of the Galaxy to obtain information about the time
scales over which the solar system elements were formed’. The basic method for
determining ages using this method was first set out by Rutherford (1929).
It is well known that some atomic nuclei are stable and some are not. Those that
are not exist for a certain average length of time and then eject particles sponta-
neously (radioactivity). The new nucleus so formed, known as the daughter nucleus,
may or may not be stable. For each radioactive nucleus there exists a transition prob-
ability λ, such that the nucleus will decay into its daughter nucleus within the next
small interval of time dt. The inverse of this transformation probability is the aver-
age lifetime of the parent nucleus and is measured in seconds (base SI unit) or
Julian years. λ is measured in bequerels (a derived SI unit of dimension [T ]−1 and
symbol Bq), equal to the number of radioactive decays per second. A non-SI unit
that is sometimes used to express the activity of a radioactive sample is the curie,
equivalent to the mean rate of decay of 1 g of radium (∼3.7 × 1010 Bq).
Radioactivity has no ‘memory’, so it is not possible to predict when an individual
nucleus will decay, but it is possible to make predictions about the average behaviour
of large numbers of nuclei.
If a large number, N, of radioactive nuclei of a particular element exists at time t,
then the number, dN, that will decay in the next small time interval dt, is given by:
dN = − N λ dt (5.21)
or
dN
= − λ dt
N
Integrating with respect to t from t0 to (t + t0 ) yields the law of exponential decay:
N(t) = N(t0 ) e−λ t (5.22)
Figure 5.2 illustrates the shape of a typical radioactive decay curve.
5.6 The determination of the ages of celestial bodies 67
1.0 N(t0)
0.8
0.6 N(t1 )
2
N(t)
0.5
0.4
0.2
0.0
0.139
Figure 5.2. The radioactive decay of a large number of parent atomic nuclei into
daughter atomic nuclei follows an exponentially shaped curve with time.
The half-life, t 1 , of a radioactive nucleus is the time it takes, on average, for half
2
of the original number of nuclei to decay into daughter nuclei:
ln 2
t1 = (5.23)
2 λ
The mean lifetime of a state τ is defined by:
1
τ= (5.24)
λ
After a time interval τ has elapsed, on average, approximately (1 − 1/e)% of
the original nuclei have decayed into daughter nuclei. The half-life and the mean
lifetime of the state are related by:
t 1 = τ ln 2 (5.25)
2
If at time t = 0, the number of parent nuclei is Np0 and the number of daughter
nuclei Nd0 then, at time t, the number of daughter nuclei is:
Ndt
= eλ t − 1 (5.26)
Npt
From this equation, if λ is known and the ratio Ndt : Npt can be measured observa-
tionally or experimentally, then a value can be determined for the age of the sample
being investigated.
68 Unit of time (second)
5.6.2 Pulsars
Pulsars, named by contraction from pulsating stars, were discovered in 1967 by
Bell and Hewish using a newly designed and built radio telescope in Cambridge
(Thorsett, 2001). Radio signals from pulsars take the general form of a series of
extremely regular short bursts of radiation. It was shown that pulsars are neutron
stars, a late stage of evolution of stars whose masses, from model stellar computa-
tions, are expected to lie between 0.2 and 2 solar masses (4 × 1029 − 4 × 1030 kg).
Pulsars have high rotational inertia that leads to an extremely stable rotation. The
individual pulses are stable to around 1 part in 1014 , which is not far removed from
the current stability of caesium fountain atomic clocks at 5 parts in 1016 . Compar-
ison between the pulse arrival time and the reference atomic standard time can be
used to determine very precise orbital parameters of binary pulsars. Observations
of the pulsation period and the rate of change of the period can also provide data on
the magnetic field strength and age of the pulsar (Burke & Graham-Smith, 2002).
The slowing of the rotation period provides the energy for the pulses.
If the original period of pulsation at the time the pulsar was formed is P0 , the
present pulsation period is P , and the rate of increase in the period is Ṗ , then the
current age of the pulsar, t, is:
n − 1
P P0
t= 1− (5.27)
(n − 1) Ṗ P
21 See http://nvo.stsci.edu/vor10/getRecord.aspx?id=ivo://nasa.heasarc/crabtime
22 http://www.jb.man.ac.uk/pulsar/crab.html (provided by A.G. Lyne, C.A. Jordan & M.E. Roberts in Jodrell Bank
Crab Pulsar Timing Results, Monthly Ephemeris).
23 See http://heasarc.gsfc.nasa.gov/
5.6 The determination of the ages of celestial bodies 69
30.05
30.00
29.95
ν Hz
29.90
29.85
29.80
Figure 5.3. Crab pulsar rotation-rate slowdown from 1982 to 2002. The abscissa
is given as modified julian day numbers and the ordinate as frequency in hertz.
Over the time covered in this data set the mean pulsation period, P, was
0.033 271 ± 0.002 106 s and the mean rate of change of the pulsation period, Ṗ ,
was 4.192 877 × 10−13 ± 0.265 206 × 10−13 s . s−1 . Substituting these values into
Equation (5.28) and using an observational value of n = 2.51 ± 0.01 for the braking
index determined by Lyne et al. (1993), the age of the Crab pulsar is found to be
1665 y. This may be compared with the known age of 956 y at the present time
(2010 CE). Burke & Graham-Smith (2002) warn that, at best, ‘present day slow-
down rates are only indications of actual age and not infallible measures of it’. One
possible cause of this discrepancy, the effect of a large transverse motion on the
observed period derivatives, was first pointed out by Shklovskii (1970). Applying
a suitable correction for this (see Manchester et al., 2005) produces a better value
for the pulsar slow-down rate.
A catalogue of pulsar data compiled by Taylor et al. (1993) forms the basis of
an online database, the ATNF Pulsar Catalog.24 Some 1640 pulsars in the database
have had their characteristic ages computed (using a braking index of 3), which
24 See http://www.atnf.csiro.au/research/pulsar/psrcat/
70 Unit of time (second)
0.3
0.2
0.1
0.0
2 3 4 5 6 7 8 9 10
log (age)
vary from the youngest at 218 y (PSR J1808-2024) to the oldest at 6.75 × 1010 y
(PSR J0514-4002A)(see warning above!). An histogram of the logarithms of these
pulsar ages is shown in Figure 5.4. The median age from the database is 5.47 × 106 y.
ages of pulsars may be found from a detailed study of the rate at which their rotation
rate slows; an example of such a study is shown for the Crab pulsar.
5.7.2 Recommendations
The IAU recommends using the TAI system when precise measures of time intervals
are required.
The commonly used unit for angular rotation, degrees per day, should be replaced
by the SI unit, rad . s−1 . Similarly, stellar proper motion should use the same unit,
though, given that its numerical value will be very small, rad . y−1 or submultiples
thereof (e.g., mrad . y−1 , μrad . y−1 or nrad . y−1 ) could be used. For very distant
or slow moving bodies it is usual, at present, to quote the proper motion as angular
measure per julian century.
The inverse unit of time can be either the hertz, when used as a derived SI unit
of frequency, or the bequerel, when used as the number of radioactive decays per
second. In the case of the latter, since the age of the celestial body is the unknown
being sought, then the unit of interest becomes the inverse of the bequerel, which
is the SI unit, the second.
6
Unit of length (metre)
The original definition of the metre was one ten millionth of the distance from
the Earth’s north pole to its equator, determined along a meridian arc that ran
from Dunkirk in the north to Barcelona in the south. Observations were begun in
1792 by J. B. J. Delambre, who worked from Paris northwards and P. F. A. Méchain
who made measurements from Paris to Barcelona. They completed their task in
seven years and the metre thus determined was modelled in pure platinum as a
one-metre-long bar (Alder, 2004).
72
6.2 Linear astronomical distances and diameters 73
r = a (1 − f . sin2 φ) (6.2)
where φ is measured northwards from the geodetic equator. Now substitute for φ,
and for a and f (from McCarthy & Petit, 2004) in Equation (6.2) to determine the
radius at mean sea level, rmsl :
To this value add the height of the Paris Observatory above mean sea level (= 67 m)
to produce a final result for the geocentric radius at the location of the Paris
Observatory of:
25 McCarthy & Petit cite the 1999 report of E. Groten to the IAG Special Commission SC3, Fundamental
Constants, XXII IAG General Assembly.
26 Woolard & Clemence (1966) define geodetic latitude as ‘the angle between the geodetic vertical and the plane
of the geodetic equator’.
74 Unit of length (metre)
or
1
ρ = (a 2 c) 3 (6.4)
Substituting the above values for a and c in Equation (6.4) produces a value for
the mean radius of the geoid of ρ = 6.371 000 37×106 m.
π M
C
Figure 6.1. Horizontal parallax: point C is the centre of the Earth, angle MCT is a
right angle, CT is the equatorial radius of the Earth, CM is the geocentric distance
of the Moon (or Sun; for solar parallax, see Section 6.2.4) and π is the parallax
angle.
The horizontal parallax of the Moon πMoon , is the angle subtended by the equato-
rial radius of the Earth at the distance of the Moon. The mean value of the equatorial
horizontal parallax is 16.593 271 8 mrad ( = 3422 . 608).
P12 a13
2
= (6.5)
P2 a23
If one of the planets is the Earth, with a period of revolution about the Sun of one
sidereal year and a mean solar distance of one astronomical unit, then Equation
(6.5) may be rewritten as:
P2
=1 (6.6)
a3
where P is the period of revolution about the Sun of a gravitationally bound astro-
nomical body other than the Earth and a is its mean distance from the Sun in
astronomical units.
76 Unit of length (metre)
For any other planet with sidereal period P years Equation (6.5) may be
rewritten as:
2
a =P 3 (6.7)
Newton expressed Kepler’s third law in a more general form, which involved
the constant of gravitation k, the mass of the planet m, and the mass of the Sun M :
k 2 P 2 (M + m) 1
3
a= 2
(6.8)
4π
or when expressed in IAU astronomical units (see below, version 3):
k 2 P 2 (1 + m) 1
3
a= (6.9)
4π2
28 See http://www.iau.org/public-press/themes/measuring/
29 See http://asa.usno.mil/index.html
6.2 Linear astronomical distances and diameters 77
Also, given the precision with which the astronomical unit can now be defined,
the assumption that the astronomical unit of mass (one solar mass) is a constant is
no longer acceptable. It is known that the Sun converts mass into energy resulting
in a loss of mass equal to 4.3×109 kg . s−1 , as well as further particulate mass loss
through the solar wind of approximately 1.3×109 kg . s−1 . A total of some 5.5 mil-
lion tonnes per second, or, in IAU astronomical unit terms: 2.405×10−16 M . d−1 ,
or in units of solar masses per century: 8.797×10−12 M . century−1 . According
to Noerdlinger (2008), such a loss will result in the orbits of the planets expanding
at the same relative rate and their periods of revolution to increase at twice that
relative rate, causing the planet Mercury to drift away from its predicted position
by more than 5 km in an interval of 200 years.
Another way in which the astronomical unit may be increased is via the total
conservation of angular momentum law, proposed by Miura et al. (2009).
It is apparent that further revision of the definition of the astronomical unit
of length is not only desirable but necessary. However, given that such revisions
may take many years to carry out, the only presently available solution to convert
astronomical units to metres is to use the currently accepted best conversion value.
Table 6.1. Sidereal period in days and distances in IAU astronomical units and SI
metres for the planets and some dwarf planets in the Solar System
astronomical unit, when the astronomical unit itself was defined to be the mean
Earth–Sun distance:
a⊕
π = (6.10)
1 au
From Section 6.2.1, the equatorial radius of the Earth is 6.378 136 60×106 m and
from the previous section the value of the astronomical unit in metres is given, so
substituting these numerical values into Equation (6.10) gives the solar parallax
to be: 42.635 209 56 μrad (equivalent to 8 .794 143 240). As the Earth’s orbit is
elliptical, the value of the solar parallax varies from approximately 43.342 μrad
(8 .94) at perihelion on January 4, when the distance to the Sun is about 0.983 au,
to 41.936 μrad (8 .65) at aphelion on July 4, when the distance to the Sun is about
1.017 au.
πΣ
E
S
Figure 6.2. Trigonometric parallax: point S is the Sun, SE is the mean radius of
the Earth’s orbit about the Sun, S is the heliocentric distance of the celestial body
and π is the parallax angle of . NB: this figure is NOT to scale since the
distance S
SE.
30 See http://archive.eso.org/skycat/servers/ASTROM
80 Unit of length (metre)
6.2.7 Parsec
The parsec is the distance, D, at which one astronomical unit subtends an angle of
one arc second. It is the reciprocal of the parallax π, expressed in arcseconds of a
celestial body, i.e.:
1
D= (6.14)
π
Table 6.2. Parallaxes and distances in astronomical units, metres, parsecs and
light years for the ten brightest stars in the night sky
Parallax Distance
πmas astronomical Distance (SI) Distance parsecs
Star name πnrad units metres ×1015 light years
extragalactic distances, the megaparsec, equal to 1 000 000 pc, is commonly used
with, for very large distances, the gigaparsec, equal to 1 000 000 000 pc.
In SI units, the parsec is equal to, 3.085 677 6×1016 m, and in IAU units is equal
to 206 264.806 265 au.
The parsec is calculated directly from a measurement of the parallax of an
astronomical object.
32 http://www.univie.ac.at/webda/
33 http://nedwww.ipac.caltech.edu/
6.3 Linear motion 83
Table 6.3. Some examples of distances given in metres, parsecs and light years to
astronomical bodies lying outside the Solar System
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
The simple prefixes work well for the parsec unit and are easy to remember, and
those for the metre also work well, though are less easy to remember and may be
difficult to associate immediately with the relevant power of 10. The d notation does
not have this problem. Prefixes are not commonly used with light years, but with
numbers of light years not exceeding tens of billions this is not an insurmountable
difficulty (e.g., the Coma cluster of galaxies would be said to be at a galactocentric
distance of approximately 310 million light years).
20 γ - rays
X - rays
log(frequency Hz)
16 ultraviolet
visible
infrared
12
radio frequencies
ν=c/λ
8
–13 –11 –9 –7 –5 –3 –1 1
log(wavelength m)
34 See http://simbad.u-strasbg.fr/simbad/sim-fid
35 In radial velocity searches for exoplanets (planets which orbit stars other than the Sun).
86 Unit of length (metre)
Table 6.4. Radial velocities and brief classifications for the ten brightest stars in
the night sky
vr B
Σ
θ vt
ight
of s v
line
S μ C
Figure 6.4. A plot showing the relationship between the transverse and radial
velocity components and the space velocity of a star. The proper-motion angle is
labelled μ, the Sun S, the star , and the line of sight SB. The angle BC is a
right angle.
and
vt = v sin θ (6.19)
Radial and transverse components are combined to yield the space velocity v of the
star, thus:
v = (vt2 + vr2 ) (6.21)
For radial velocities, v c, the relationship between z and v may be written as:
v=c.z (6.26)
Hubble (1929) showed that the velocity v, with which a galaxy is receding (its
recession velocity), is directly proportional to its distance D, i.e.:
v = H0 . D = c . z (6.27)
88 Unit of length (metre)
For very distant galaxies, the recession velocity is no longer very small in compari-
son with the velocity of light and the relationship has to be written in the relativistic
form:
(z + 1)2 − 1
v=c. (6.28)
(z + 1)2 + 1
There are three component parts to a red shift: that due to the actual or peculiar
motion of the celestial object itself; that due to the gravitational field of the object
and known as the gravitational red shift; and that due to the expansion of the
Universe and known as the cosmological red shift.
Since, in general, celestial objects are detected by the electromagnetic radiation
they emit, and given that the speed of such radiation is finite, then the further
away an object appears to be, the longer ago the detected radiation was emitted.
This leads to yet another unit in common use the look-back time or sometimes
its complement, the time elapsed since the Big Bang, which occurred some 13.7
billion years ago. Note that the sum of these two numbers is the age of the Universe
at the time the observation was made. The dimension of look-back time and time
elapse since the Big Bang is [T ] and its SI unit is the second and its IAU unit the
Julian year.
Wright (2006) has constructed an online calculator36 which, for a given cos-
mological model and input value of red shift, computes distances in parsecs and
light years, light travel times, the age of the Universe at the departure time of the
electromagnetic radiation and other variables of use in cosmological studies.
6.4 Acceleration
The dimension of the SI derived unit of acceleration is [L] . [T ]−2 , its unit is the
metre per second per second and its symbol is m . s−2 .
The acceleration at the surface of a celestial body due to its gravity is known as
its surface gravity.
In Cox (2000), the values of the surface gravity of the bodies in Table 6.5 are
given in cgs units. To convert to SI units, divide the cgs value by 100, e.g., the
surface gravity of Mars:
Mars(cgs) 371
Mars(SI) = = = 3.71 m . s−2
100 100
Whilst cgs units are still in common use for surface gravity measurements, SI
units are now readily accepted and recommended by the IAU.
36 See http://astro.ucla.edu/∼wright/DlttCalc.html
6.6 Volume 89
6.5 Area
The dimension of the SI unit of area is [L] . [L] = [L]2 , its name is the square metre
and its symbol is m2 .
This unit is in common use in astronomy, as are the prefixed units cm2 and km2 ,
and surface areas of planets relative to that of the Earth and the stars relative to that
of the Sun.
Celestial bodies such as the planets and stars generally have a regular shape
and are mainly either spheres or oblate spheroids.37 The surface area, A of such
bodies, with an equatorial radius 38
a −c of √
a, a polar
radius of c, and values of oblateness
a −c
ω = c and eccentricity e = c are given by:
4 π a2 if c=a
A= 2 (6.29)
2 π a 2 + πec . ln 11 +
−e
e
if a>c
6.6 Volume
The dimension of the SI unit of volume is [L] . [L] . [L] = [L]3 , its name is the cubic
metre, and its symbol is m3 .
This unit is in common use in astronomy, as are the prefixed SI units cm3 and
km3 , and volumes of planets relative to that of the Earth and the stars relative to
that of the Sun.
37 See http://mathworld.wolfram.com/OblateSpheroid.html
38 Note that the mathematical expressions for oblateness and flattening are identical.
90 Unit of length (metre)
Table 6.6. Radii, surface areas and volumes of planets and one dwarf planet in
the Solar System
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
Volumes V of spheres and oblate spheroids are computed from one or other of
the following expressions:
4 π a3 if c = a
V = 34 (6.30)
π a c if a > c
2
3
Table 6.6 was compiled by substituting data on the radii (listed in km and Earth
radius units) and oblateness from the JPL D405 ephemeris and The Astronomi-
cal Almanac Online39 into Equations (6.29) and (6.30). The planets Mercury and
Venus, and the dwarf planet Pluto, were treated as spheres, whilst Earth to Neptune
inclusive were treated as oblate spheroids. In the table, column 2 lists the planetary
radii in both m (SI unit) and km; column 3 gives the radii in Earth radius units;
columns 4 and 5 the planetary surface areas in m2 and Earth surface area units,
and columns 6 and 7 planetary volumes in m3 and relative to the volume of the
Earth.
6.7.2 Recommendations
Wherever possible the SI units, m, m . s−1 , m . s−2 , m2 and m3 should be used.
Until more accurate distance measurements in metres are available, particularly
for objects in the outer Solar System, it is appropriate and sensible to go on using
astronomical units that depend solely on measures of angle and time. The IAU
recommends the use of the parsec, though the light year may prove easier to define,
given that it does not rely on the astronomical unit but only on the defined speed
of electromagnetic radiation and a defined period of time (one Julian year).
SI units of velocity are routinely used in either the m . s−1 form or, for greater
velocities, the km . s−1 form, and this should continue. For cosmological purposes,
red shifts and look-back times can be of more use than simple distance measures.
7
Unit of mass (kilogram)
92
7.1 SI definition of the kilogram 93
F = m×a (7.1)
where F is the force in newtons (kg . m . s−2 ), m the mass in kg and a the acceleration
in m . s−2 . Applying a force of 1 N to a mass of 1 kg will result in an acceleration
of 1 m . s−2 .
Defined by Planck’s constant, using the relationship between energy E in joules
(m2 . kg . s−2 ), frequency ν, in hertz (s−1 ) and Planck’s constant h, in J . s:
E = hν (7.2)
E = m c2 (7.3)
Now rearrange Equation (7.2) and substitute the values for h, E, m (= 1 kg) and c
(= 299 792 458 m . s−1 ) to find ν:
E (299 792 458)2
ν= =
h 6.626 068 96 × 10−34
= 1.356 × 1050 Hz (7.4)
41 www.french-metrology.com/en/feature/watt-balance.asp
94 Unit of mass (kilogram)
In a paper on redefining the kilogram, Mills et al. (2005) have proposed six
separate revised definitions, three that fix the value of the Planck constant and three
that fix the value of the Avogadro constant. An example of each type follows.
The kilogram is the mass of a body at rest whose equivalent energy corresponds to
frequency of exactly [(299 792 458)2 / (6 626 069 311)] × 1043 Hz.
The kilogram is the mass of a body at rest such that the value of the Avogadro
constant is exactly 6.022 141 527 × 1023 mol−1 .
Until a final decision is made about the revised formal definition by the CIPM
the definition of the kilogram remains as set out at the beginning of this chapter.
and its value in SI units is 6.673 × 10−11 m3 . kg−1 . s−2 . To convert G to other
systems of units it is only necessary to determine the appropriate constant of
proportionality between the systems.
G(other units) = [kL . L]3 . [kM . M]−1 . [kT . T ]−2 × GSI (7.10)
Note that the accuracy with which the values of μ may be expressed far exceeds that
of the constant of gravitation as determined in the laboratory. The best current values
for G have standard deviations about the mean measured value of approximately
±0.01 % or 1 part in 104 , whereas the heliocentric gravitational constant is known
to approximately 1 part in 1010 . Hence it is customary when computing the values
for semimajor axes of Solar System bodies to use IAU units rather than SI units
via the relationship (Kepler’s third law):
3
G . M . P 2 3
2
k . M . P 2 3
μ.P 2
a= = = (7.17)
4.π2 4.π2 4.π2
where a is the semimajor axis of the orbit, P is the orbital period, M is the mass of
the Sun, G is the Newtonian gravitational constant, k is the Gaussian gravitational
constant and μ is the standard gravitational parameter. Planetary ephemerides are
currently computed using k with distances in astronomical units and masses in
inverse mass units M / MP .
so
μIAU = [kL . L]3 . [kT . T ]−2 . μSI (7.19)
4 . π 2 . a3
M (7.21)
G.P 2
where a is the semimajor axis of a planetary orbit in metres, P is the period of revolu-
tion in SI seconds, and G is the Newtonian gravitational constant in m3 . kg−1 . s−2 .
Substituting the values for a and P that approximate for the Earth and making the
assumption that the mass of the Earth is small in comparison with that of the Sun
98 Unit of mass (kilogram)
gives:
4 . π 2 . (1.495 978 714 64 × 1011 )3
M
6.673 × 10−11 . (365.25 × 86 400)2
1.989 × 1030 kg (7.22)
Note that if IAU units are used, then P = 365.25 d, a = 1 (A), and G = 2.968 ×
10−4 (A)3 . (M )−1 . (d)−2 , so that:
4 . π 2 . 13
M
2.968 × 10−4 . 365.252
0.997
1 (7.23)
The mass of the Sun may also be determined from observations of the orbital
velocity and position of a planet at a particular time. If the measured velocity of
the planet is vP , and its radius vector is RP , then the approximate mass of the Sun
may be calculated from:
v 2 . RP
M = P (7.24)
G
where G is the Newtonian constant of gravitation.
IAU units plus the values for RP = R and vP = v in Equation (7.24) yields a value
for the mass of the Sun of M 1.004 (in solar mass units).
To determine the mass of the Sun in SI units multiply R by
1.495 978 714 64 × 1011 (the number of metres in an astronomical unit) and divide
by the number of SI seconds in a julian day (86 400). Insert the adjusted numbers
into Equation (7.24) to give M 2.002 × 1030 kg.
dim(G) . dim(M )
dim((M )T ) =
dim(c3 )
[L]3 . [M]−1 . [T ]−2 . [M]
=
[L]3 . [T ]−3
= [T ] (7.30)
the mass of the Earth. Laboratory-based determinations later used torsion balances
and gravity balances (Spencer-Jones, 1956).
The advent of artificial satellites in orbit about the Earth, which obey Kepler’s
third law, allow the same dynamical method to be employed in the determination
of the Earth’s mass as is used for the Sun. If, in Equation (7.21), the mass of the
Earth M⊕ , is substituted for the mass of the Sun, and the semimajor axis aS and
the orbital period PS of the artificial satellite for those of the Earth, then:
4 . π 2 . aS3
M⊕ = (7.31)
G . PS2
Example: The International Space Station and the mass of the Earth
By way of an example it is possible to determine approximately the mass of the Earth
using observations of the International Space Station. Measurements of the orbital
period in minutes and the height of the apogee point of the orbit above the surface
of the Earth in kilometres are to be found on the satellite database of astrosat.net.43
The semimajor axis length of the satellite’s orbit is taken to be the sum of the
Earth’s equatorial radius and the satellite’s apogee distance and the listed orbital
period multiplied by 60 to convert to seconds of time. The value GSI is substituted
in Equation (7.31) for G to give:
4 . π 2 . (6.732 × 106 )3
M⊕ =
6.673 × 10−11 . (5.491 × 103 )2
= 5.97 × 1024 kg (7.32)
which, allowing for the very basic analysis performed, is a reasonable approxi-
mation to the mass of the Earth given in The Astronomical Almanac Online as
5.972 198 6 × 1024 kg.44
For a more detailed account of how to determine the mass of the Earth or any
other planet taking into account the oblateness of the planet see, e.g., Blanco &
McCuskey (1961).
43 See www.satview.com.br/us/track_lista_sat.php
44 http://asa.usno.mil/index.html, page K7, 2010.
7.3 Masses of astronomical bodies 101
Table 7.1. Masses, in kilograms, solar mass units, reciprocal solar mass units,
Jovian mass units and Earth mass units of the Sun, planets, dwarf planets,
satellites and a comet in the Solar System
Name Mass
Star kg M 1 / M MJ M⊕
Sun 1.9884 d 30 1 1 1.0473 d 3 3.3294 d 5
Planet
Mercury 3.3010 d 23 1.6601 d − 7 6.0236 d 6 1.7387 d − 4 0.0553
Venus 4.8673 d 24 2.4478 d − 6 4.0852 d 5 2.5637 d − 3 0.8150
Earth 5.9721 d 24 3.0035 d − 6 3.3295 d 5 3.1457 d − 3 1
Mars 6.4169 d 23 3.2272 d − 7 3.0987 d 6 3.3800 d − 4 0.1074
Jupiter 1.8985 d 27 9.5479 d − 4 1.0473 d 3 1 317.891
Saturn 5.6846 d 26 2.8589 d − 4 3.4979 d 3 2.9942 d − 1 95.184
Uranus 8.6818 d 25 4.3662 d − 5 2.2903 d 4 4.5730 d − 2 14.5371
Neptune 1.0243 d 26 5.1514 d − 5 1.9412 d 4 5.3953 d − 2 17.1512
Dwarf Planet
Ceres 9.3455 d 20 4.7 d − 10 2.1 d 9 4.9 d − 7 1.6 d − 4
Pluto 1.4707 d 22 7.3964 d − 9 1.3520 d 8 7.7467 d − 6 2.4626 d − 3
Eris 1.9884 d 20 1 d − 10 1 d 10 1.0 d − 7 3.3294 d − 5
Satellite
Moon 7.3458 d 22 3.6943 d − 8 2.7068 d 7 3.8693 d − 5 1.230 d − 2
Ganymede 1.4818 d 23 7.4522 d − 8 1.3419 d 7 7.805 d − 5 2.48 d − 2
Europa 4.7994 d 22 2.4137 d − 8 4.1430 d 7 2.528 d − 5 8.04 d − 3
Comet
Halley 5 d 14 2.4 d − 16 4 d 15 2.6 d − 13 8.4 d − 11
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
Examples of the masses of the planets, dwarf planets, satellites and Halley’s
comet are given in Table 7.1. The values were computed from data given in Cox
(2000) (for the Moon), Brown & Schaller (2007) (for Eris) and The Astronomical
Almanac Online.
It should be remembered that until the value of the solar mass in kilograms is
greatly improved, the relative masses of the planets in solar mass units are much
more accurate. In Table 7.1, the masses in kilograms of the various objects were
derived by multiplying the mass of the body in solar mass units by one solar mass
in kilograms which, being given to five significant figures, immediately restricts
all the other derived values to being no better than five significant figures, and
generally rather worse. The values given for the reciprocal solar masses of these
same objects range from seven significant figures for Uranus and Neptune up to 12
significant figures for the Earth, which explains why IAU units are to be preferred
for the computations of ephemerides at present.
102 Unit of mass (kilogram)
B⬙
B⬘
A
C A⬘
A⬙
Figure 7.1. The orbits of a pair of stars, A and B, in a binary system about their
centre of mass, C, the barycentre. Star A moves from A through A to A whilst
star B moves from B through B to B . The barycentric distances BC and CA are
equal to rB and rA , respectively.
7.3 Masses of astronomical bodies 103
values into Equation (7.33) to obtain the total mass of the binary system. To obtain
the individual masses, consider the binary star orbits shown in Figure 7.1. The
distance from star A to the barycentre C, is rA , and likewise star B is distance rB
−→
from the barycentre. Distance AB = r, so r = rA + rB . The total mass of the system
is M, where M = MA + MB and MA . rA = MB . rB . Using Newton’s second law of
motion balancing gravitational and centripetal force gives:45
G . MA . MB MA . ṙA2
= (7.34)
r2 rA
where ṙA is the orbital speed of star A, which may be measured spectroscopically
or simply taken as the mean value defined by:
2 . π .rA
ṙA = (7.35)
P
Combining the various expressions above produces the relationship:
M . (r − rA )
MA = (7.36)
r
The value for the mass of star B is derived from:
MB = M − MA (7.37)
If IAU units are used with mass in solar mass units, distance in astronomical
units and time in days of 86 400 SI seconds, Equation (7.33) may be rewritten as:
3
4 . π 2 . aIAU
MIAU = 2
(7.38)
GIAU . PIAU
which gives the mass of the binary system in solar mass units. Equations (7.34) to
(7.37) may similarly be reworked.
Example: determine the total mass of the Sirius binary system in both IAU
and SI units
(1) IAU units
Parallax: 379 mas
Distance: 206 264.806
0.379
25
= 5.442 343 × 105 au
7.5
Orbital semimajor axis (in radians): 206 264.806 25 rad
7.5×5.442 343×105
Orbital semimajor axis (in astronomical units): 206 264.806 25 = 19.789 au
Asteroseismology
Asteroseismology may be defined as the study of the internal structure of individual
stars through the analysis of their pulsation periods and the interpretation of the
derived frequency spectra. Stellar masses and radii can be determined to a much
higher precision than by any other method.47
This branch of astronomy has matured in recent times, with the development of
high-speed photometers and the establishment of networks of observational sites
throughout the world (e.g., the WET (Whole Earth Telescope) collaboration). The
46 See http://ad.usno.navy.mil/wds/
47 See http://kepler.nasa.gov/Science/about/RelatedScience/Asteroseismology/
7.3 Masses of astronomical bodies 105
analysis of the power spectrum of the light curve (Winget et al., 1994) allows direct
estimations to be made of the masses of individual stars. For the pulsating DB white
dwarf GD358, Bradley (1993) derived a mass of 0.61 ± 0.03 M .
Asteroseismological reduction techniques have also been applied to observations
made of relatively bright main-sequence and subgiant branch stars using spectro-
graphs. Bedding et al. (2006) used such observations to determine the mass of the
metal-poor subgiant star ν Indi to be 0.85 ± 0.04 M . Ferdman et al. (2010) used
radio telescopes in the USA, Australia and France to make extensive observations
of the intermediate-mass binary pulsar PSR J1802-2124, which yielded masses
for the pulsar component of 1.24 ± 0.11 M and for the white-dwarf component
of 0.78 ± 0.04 M . Such measurements and analysis provide important indepen-
dent estimates of stellar masses using a method other than those of Kepler/Newton
dynamics.
Some examples of stellar masses determined from dynamical and asteroseis-
mological techniques are given in Table 7.2, which was compiled, in part, from
material to be found in Cox (2000), Ferdman et al. (2010), Winget et al. (1994)
and Bedding et al. (2006). It is worth noting that currently there is no prefix in SI
that is large enough for stellar masses, even the smallest mass in Table 7.2 would
be given as 3.182 × 108 Yg (3.182 × 108 yottagrams). The additional, unofficial,
prefixes proposed by Mayes (1994) would describe the mass as 318.2 Sg (318.2
sansagrams). (Again, remembering that the prefix is applied to the mass unit gram
and not to the SI unit kilogram.)
Example of mass gain: According to Love & Brownlee (1993), the infall of dust
and small meteoroids on to the Earth amounts to some 4 × 107 kg . y−1 , equivalent
to approximately 1.3 kg . s−1 .
Examples of mass loss: in the region near the centre of the Sun, energy is generated
by the conversion of hydrogen nuclei into helium nuclei. This reaction results in a
loss of mass amounting to 4.3 × 109 kg . s−1 , with a further 1.3 × 109 kg . s−1 being
lost via the solar wind. A total mass loss of 5.6 × 109 kg . s−1 , equivalent in IAU
units to 2.4 × 10−16 M . d−1 , or 8.8 × 10−14 M . y−1 .
106 Unit of mass (kilogram)
Mass
Star identifier Spectral type kg M
Sirius A A1V 4.535 d 30 2.28
Procyon A F5IV–V 3.361 d 30 1.69
PSR J1802-2124 A Pulsar 2.466 d 30 1.24
α Cen A G2V 2.148 d 30 1.08
SUN G2V 1.989 d 30 1.00
Sirius B DA 1.949 d 30 0.98
70 Oph A K0V 1.790 d 30 0.90
α Cen B K0V 1.750 d 30 0.88
ν Ind G0IV 1.691 d 30 0.85
70 Oph B K4V 1.293 d 30 0.65
GD358 DBV 1.213 d 30 0.61
Procyon B WD 1.193 d 30 0.60
Krüger 60 A dM4 5.370 d 29 0.27
Krüger 60 B dM6 3.182 d 29 0.16
7.4 Density
Density may be defined as the mass or quantity of matter of a substance per unit
volume.
The dimension of density is [M ] . [L]−3 , its derived unit is the kilogram per
cubic metre and its symbol, kg . m−3 .
48 Reported by D. Valencia, M. Ikoma, T. Guillot and N. Nettlemann at the 41st Lunar and Planetary Sci-
ence Conference (2010) in a paper entitled ‘Composition and fate of short-period super-Earth’s: The case
of CoRoT-7b.’
7.4 Density 107
Mean density
Celestial object kg . m−3 g . cm−3 M . pc−3 Notes
Sun 1409 1.409 – mean density
Earth 5514.8 5.515 –
Jupiter 1330 1.33 –
Saturn 700 0.70 –
Moon 3341 3.341 – mean density
Asteroids 2250 2.25 – mean of estimated
range of densities
Halley’s comet 650 0.65 – mean of estimated
range of densities
Interstellar matter 2.71 d − 21 2.71 d − 24 0.04
Main sequence stars 3.38 d − 21 3.38 d − 24 0.050 mass density in the
solar neighbourhood
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
Some examples of the densities of astronomical bodies are given in Table 7.3,
with data from Cox (2000) recalculated where necessary.
A form of density unit that is often used is the column mass density, whose unit
is the M . pc−2 . Column mass density is normally accompanied by a range within
which the quoted value is valid. For example, the observed column mass density to
108 Unit of mass (kilogram)
so in SI units the column mass density of the local interstellar medium would be
0.027 kg . m−2 within |z| = 3.394 × 1019 m.
7.5 Force
Funk et al. (1946) define force as: ‘Any cause that produces, stops, changes or tends
to produce, stop or change the motion of a body.’
The dimension of force is [M ] . [L] . [T ]−2 , its derived SI unit is the newton with
symbol, N, or, kg . m . s−2 .
Another unit still used by astronomers is the dyne, a cgs unit equivalent to
1 g . cm . s−2 . The conversion coefficient between the two systems is derived as
follows:
7.5.1 Energy
The energy of a body is its capacity for doing work, where work is defined as the
product of the force that moves the body and the distance through which it is moved.
The SI units for energy and work are the same.
The dimension of energy is [M ] . [L]2 . [T ]−2 , its derived SI unit is the joule with
symbol, J, or N . m or kg . m2 . s−2 .
The cgs unit of energy, the erg, is regularly used by astronomers. Measures in
ergs are converted to joules via 1 erg = 10−7 J.
The potential energy of a system is the energy that the system has due to the
relative positions of its component parts.
7.6 Moments of inertia and angular momentum 109
The kinetic energy of a moving body or system is the energy it possesses due to
its motion.
The dimensions, units and symbols of both potential and kinetic energy are the
same as those for energy.
7.5.2 Power
Power is the rate at which energy is transferred per unit time.
The dimension of power is [M ] . [L]2 . [T ]−3 , its derived SI unit is the watt with
symbol W, or J . s−1 , or kg . m2 . s−3 .
The cgs unit of power, the erg . s−1 , is still used as, e.g., a measure of luminosity
in model stellar atmosphere calculations (see Table G.10 in Irwin, 2007).
7.5.3 Pressure
Pressure is a force applied perpendicularly to a unit area.
The dimension of pressure is [M ] . [L]−1 . [T ]−2 , its derived SI unit is the pascal
with symbol, Pa, or N . m−2 , or kg . m−1 . s−2 .
Astronomers tend to use the cgs unit of pressure, the dyn . cm−2 , or the bar,
equal to 106 dyn . cm−2 . The conversion factors from these units to the SI unit of
pressure are:
1 bar = 105 Pa 1 Pa = 10−5 bar
1 dyn . cm−2 = 10−6 bar 1 Pa = 10 dyn . cm−2 (7.47)
The atmospheric surface pressures on the Sun, some planets, a dwarf planet and
a satellite, listed by Cox (2000), are converted to SI units using the appropriate
factors from Equations (7.47) and set out in Table 7.4.
The atmospheric pressure given for the Sun refers to a point where the radial
optical depth at a wavelength of 500 nm is equal to 1.
The dimension of moment of inertia is [M ] . [L]2 , its unit is the kilogram metre
squared and its symbol is kg . m2 .
110 Unit of mass (kilogram)
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
Moments of inertia are often presented in cgs units as g . cm2 and, for Solar
System objects, as M . R 2 , where M is the mass of the objects in units of the
Earth’s mass and R is the object’s radius in units of the Earth’s radius.
To convert from cgs to SI units: the conversion factor is derived in the following
way:
1 1 2
ISI = . . Icgs
1000 100
= 10−7 . Icgs (7.49)
In Cox (2000), the moment of inertia of the Sun is given as 5.7 × 1053 g . cm2 ,
which, using the conversion factor in Equation (7.49), becomes 5.7 × 1046 kg . m2
in SI units.
To convert from M⊕ . R⊕ 2 units: the conversion factor is derived by substituting
values for the Earth’s mass in kilograms and the Earth’s radius in metres:
ISI = (5.972 198 6 × 1024 ) . (6.378 136 6 × 106 )2 . IM⊕ . R 2
⊕
38
= 2.429 527 × 10 . IM⊕ . R 2 (7.50)
⊕
2
In Cox (2000), the moment of inertia of the Earth is given as 0.333 5 M⊕ . R⊕
which is readily converted into SI units via Equation (7.50) to equal
8.10 × 1037 kg . m2 .
7.7 Summary and recommendations 111
The angular momentum, Iω, of a celestial body about its rotation axis is defined
as the product of its moment of inertia and its angular velocity.
The dimension of angular momentum is [M ] . [L]2 . [T ]−1 , its unit is the
kilogram metre squared per second, and its symbol is kg . m2 . s−1 .
In Cox (2000), the angular momentum of the Sun is given in cgs units as
1.63 × 1048 g . cm2 . s−1 . The time unit is the same in both cgs and SI units. The
conversion factor from cgs to SI units is the same as that given in Equation (7.49),
so that in SI units the solar angular momentum is 1.63 × 1041 kg . m2 . s−1 .
7.7.2 Recommendations
The probable change from the IPK to a fundamental particle or energy definition
of the SI kilogram should have no effect on the use of the kilogram as the base unit
of mass in astronomy.
112 Unit of mass (kilogram)
113
114 Unit of luminous intensity (candela)
radiometry rather than photometry as defined (it should be noted that astronomical
observations cover an even larger total bandwidth than that of radiometry).
Equivalent photometric, radiometric and astronomical units and the quantities
they represent are set out in Table 8.1.
The jansky is a non-SI unit that is recognized by the IAU for use in astronomy.
Formally:
The jansky was introduced by radio astronomers and is also used by infrared
astronomers. Absolute calibrations of the many magnitude systems (see below)
in common use often use the jansky or a similar SI-based or cgs-based unit.
Other than the generally very small measurements encountered in astronomy for
monochromatic flux density there would seem to be no good reason for not adopt-
ing the SI derived radiometric unit for spectral irradiance, the W . m−2 . Hz−1 .
As an example, the monochromatic flux density of a zero-magnitude A0V star
in the V waveband according to Bessell (2001) is 3636 Jy. Using the radiometric
unit for spectral irradiance, this is equivalent to 3.636×10−23 W . m−2 . Hz−1 or
3.636 d −23 W . m−2 . Hz−1 .
8.2 Radiometry and photometry 115
Monochromatic flux density and total luminosity are the quantities most com-
monly used in astronomy. An in-depth description of these and other radiometric,
photometric and astrophysical terms is given by Sterken & Manfroid (1992).
The surface area of a sphere with a radius r, equal to the mean Sun–Earth
distance, is:
where νhfb is the cut-off frequency at the high-frequency boundary of the FWHM
location of the filter and νlfb is the cut-off frequency at the low-frequency boundary
of the FWHM location of the filter. Figure 8.2 illustrates the differences between
bandwidths in wavelength and frequency spaces.
λ ν
SW LW HF LF
0.5 Δλ
Δλ Δν
Compute the corresponding frequencies νhfb and νlfb using Equations (8.14) and
(8.15) to give νhfb = 5.966 0×1014 Hz and νlfb = 5.102 9×1014 Hz. The frequency
bandwidth is simply the difference between the two:
ν = 8.631 × 1013 Hz. (8.20)
Note that the wavelength FWHM is symmetrically placed with respect to the
mean wavelength but the frequency FWHM is not (see Figure 8.2).
λ (nm) ν (Hz)
Filter
λ (nm)
ν (Hz) νlfb (Hz) νhfb (Hz)
Johnson Cousins Glass
U 367 8.169 d 14 7.495 d 14 8.976 d 14
66 1.481 d 14
B 436 6.876 d 14 6.207 d 14 7.707 d 14
94 1.500 d 14
V 545 5.501 d 14 5.103 d 14 5.966 d 14
85 8.632 d 13
R 638 4.699 d 14 4.175 d 14 5.373 d 14
160 1.197 d 14
I 797 3.762 d 14 3.440 d 14 4.149 d 14
149 7.094 d 13
J 1220 2.457 d 14 2.260 d 14 2.692 d 14
213 4.323 d 13
H 1630 1.839 d 14 1.681 d 14 2.030 d 14
307 3.495 d 13
K 2190 1.369 d 14 1.257 d 14 1.503 d 14
390 2.457 d 13
L 3450 8.690 d 13 8.133 d 13 9.328 d 13
472 1.194 d 13
M 4750 6.311 d 13 6.020 d 13 6.633 d 13
460 6.126 d 12
Geneva
U 350 8.565 d 14 8.027 d 14 9.182 d 14
47 1.155 d 14
B 424 7.071 d 14 6.489 d 14 7.767 d 14
76 1.278 d 14
B1 402 7.458 d 14 7.121 d 14 7.827 d 14
38 7.065 d 13
B2 448 6.692 d 14 6.399 d 14 7.013 d 14
41 6.137 d 13
V 551 5.441 d 14 5.129 d 14 5.793 d 14
67 6.640 d 13
V1 541 5.541 d 14 5.325 d 14 5.776 d 14
44 4.514 d 13
G 578 5.187 d 14 4.984 d 14 5.407 d 14
47 4.225 d 13
Strömgren
u 349 8.590 d 14 8.236 d 14 8.976 d 14
30 7.398 d 13
v 411 7.294 d 14 7.129 d 14 7.467 d 14
19 3.374 d 13
b 467 6.420 d 14 6.298 d 14 6.546 d 14
18 2.475 d 13
120 Unit of luminous intensity (candela)
λ (nm) ν (Hz)
Filter
λ (nm)
ν (Hz) νlfb (Hz) νhfb (Hz)
y 547 5.481 d 14 5.368 d 14 5.598 d 14
23 2.306 d 13
βw 489 6.131 d 14 6.038 d 14 6.226 d 14
15 1.881 d 13
βn 486 6.169 d 14 6.150 d 14 6.188 d 14
3 3.808 d 12
Walraven
W 323.3 9.273 d 14 9.057 d 14 9.499 d 14
15.4 4.420 d 13
U 361.6 8.291 d 14 8.037 d 14 8.561 d 14
22.8 5.233 d 13
L 383.5 7.817 d 14 7.600 d 14 8.047 d 14
21.9 4.468 d 13
B 427.7 7.009 d 14 6.630 d 14 7.435 d 14
49.0 8.057 d 13
V 540.6 5.546 d 14 5.207 d 14 5.931 d 14
70.3 7.242 d 13
DDO
35 349.0 8.590 d 14 8.143 d 14 9.089 d 14
38.3 9.455 d 13
38 381.5 7.858 d 14 7.532 d 14 8.213 d 14
33.0 6.810 d 13
41 416.6 7.196 d 14 7.125 d 14 7.269 d 14
8.3 1.434 d 13
42 425.7 7.042 d 14 6.982 d 14 7.103 d 14
7.3 1.208 d 13
45 451.7 6.637 d 14 6.582 d 14 6.693 d 14
7.6 1.117 d 13
48 488.6 6.136 d 14 6.021 d 14 6.255 d 14
18.6 2.337 d 13
Thuan Gunn
u 353 8.493 d 14 8.037 d 14 9.003 d 14
40 9.654 d 13
v 398 7.532 d 14 7.172 d 14 7.931 d 14
40 7.589 d 13
g 493 6.081 d 14 5.678 d 14 6.546 d 14
70 8.678 d 13
r 655 4.577 d 14 4.283 d 14 4.915 d 14
90 6.319 d 13
Vilnius
U 345 8.690 d 14 8.213 d 14 9.224 d 14
40 1.011 d 14
P 374 8.016 d 14 7.747 d 14 8.305 d 14
26 5.579 d 13
8.2 Radiometry and photometry 121
λ (nm) ν (Hz)
Filter
λ (nm)
ν (Hz) νlfb (Hz) νhfb (Hz)
X 405 7.402 d 14 7.207 d 14 7.609 d 14
22 4.024 d 13
Y 466 6.433 d 14 6.259 d 14 6.618 d 14
26 3.592 d 13
Z 516 5.810 d 14 5.694 d 14 5.931 d 14
21 2.365 d 13
V 544 5.511 d 14 5.382 d 14 5.646 d 14
26 2.635 d 13
S 656 4.570 d 14 4.501 d 14 4.641 d 14
20 1.394 d 13
2MASS
J 1235 2.427 d 14 2.278 d 14 2.598 d 14
162 3.198 d 13
H 1662 1.804 d 14 1.677 d 14 1.951 d 14
251 2.740 d 13
K 2159 1.389 d 14 1.309 d 14 1.478 d 14
261 1.685 d 13
SDSS A0 star
u 356 8.421 d 14 7.727 d 14 9.253 d 14
64 1.526 d 14
g 475 6.311 d 14 5.526 d 14 7.357 d 14
135 1.831 d 14
r 620 4.835 d 14 4.354 d 14 5.436 d 14
137 1.082 d 14
i 761 3.939 d 14 3.577 d 14 4.383 d 14
154 8.055 d 13
z 907 3.305 d 14 3.058 d 14 3.597 d 14
147 5.392 d 13
Washington
C 391 7.667 d 14 6.722 d 14 8.922 d 14
110 2.201 d 14
M 509 5.890 d 14 5.339 d 14 6.567 d 14
105 1.228 d 14
T1 633 4.736 d 14 4.455 d 14 5.056 d 14
80 6.010 d 13
T2 805 3.724 d 14 3.407 d 14 4.107 d 14
150 7.000 d 13
ANS
ANS1 154.5 1.940 d 15 1.910 d 15 1.972 d 15
5.0 6.281 d 13
ANS2 154.9 1.935 d 15 1.847 d 15 2.033 d 15
14.9 1.866 d 14
ANS3 179.9 1.666 d 15 1.600 d 15 1.738 d 15
14.9 1.383 d 14
122 Unit of luminous intensity (candela)
λ (nm) ν (Hz)
Filter
λ (nm)
ν (Hz) νlfb (Hz) νhfb (Hz)
ANS4 220.0 1.363 d 15 1.303 d 15 1.428 d 15
20.0 1.241 d 14
ANS5 249.3 1.203 d 15 1.167 d 15 1.240 d 15
15.0 7.242 d 13
ANS6 329.4 9.101 d 14 8.965 d 14 9.241 d 14
10.0 2.764 d 13
TD1
TD1 156.5 1.916 d 15 1.733 d 15 2.141 d 15
33 4.085 d 14
TD2 196.5 1.526 d 15 1.407 d 15 1.666 d 15
33 2.580 d 14
TD3 236.5 1.268 d 15 1.185 d 15 1.363 d 15
33 1.777 d 14
TD4 274.0 1.094 d 15 1.036 d 15 1.160 d 15
31 1.242 d 14
IRAS
12 μm 12 000 2.498 d 13 1.934 d 13 3.527 d 13
7 000 1.593 d 13
25 μm 25 000 1.199 d 13 9.805 d 12 1.543 d 13
11 150 5.628 d 12
60 μm 60 000 4.997 d 12 3.932 d 12 6.852 d 12
32 500 2.921 d 12
100 μm 100 000 2.998 d 12 2.590 d 12 3.558 d 12
31 500 9.684 d 11
HST
HST336 334 8.976 d 14 8.386 d 14 9.655 d 14
47 1.269 d 14
HST439 430 6.972 d 14 6.440 d 14 7.599 d 14
71 1.159 d 14
HST450 451 6.647 d 14 5.942 d 14 7.542 d 14
107 1.600 d 14
HST555 532 5.635 d 14 4.951 d 14 6.539 d 14
147 1.587 d 14
HST675 667 4.495 d 14 4.104 d 14 4.968 d 14
127 8.636 d 13
HST814 788 3.804 d 14 3.480 d 14 4.196 d 14
147 7.159 d 13
HIPPARCOS TYCHO
BT 421 7.121 d 14 6.574 d 14 7.767 d 14
70 1.192 d 14
VT 526 5.699 d 14 5.205 d 14 6.298 d 14
100 1.093 d 14
HP 517 5.799 d 14 4.744 d 14 7.458 d 14
230 2.714 d 14
8.2 Radiometry and photometry 123
λ (nm) ν (Hz)
Filter
λ (nm)
ν (Hz) νlfb (Hz) νhfb (Hz)
MACHO
B 519 5.776 d 14 5.073 d 14 6.707 d 14
144 1.634 d 14
R 682 4.396 d 14 3.888 d 14 5.056 d 14
178 1.167 d 14
UBV(CCD)
B 436 6.876 d 14 6.207 d 14 7.707 d 14
94 1.500 d 14
V 545 5.501 d 14 5.103 d 14 5.966 d 14
85 8.632 d 13
R 641 4.677 d 14 4.158 d 14 5.344 d 14
160 1.186 d 14
I 791 3.790 d 14 3.476 d 14 4.167 d 14
143 6.908 d 13
Z 909 3.298 d 14 3.133 d 14 3.482 d 14
96 3.493 d 13
USNO-B(I)
I 807.5 3.713 d 14 3.331 d 14 4.193 d 14
185 8.619 d 13
EROS
BE1 485 6.181 d 14 5.557 d 14 6.964 d 14
109 1.407 d 14
BE2 539 5.562 d 14 4.729 d 14 6.752 d 14
190 2.023 d 14
RE1 657 4.563 d 14 3.984 d 14 5.339 d 14
191 1.355 d 14
RE2 767 3.909 d 14 3.342 d 14 4.706 d 14
260 1.364 d 14
Spitzer
24 μm 23700 1.265 d 13 1.151 d 13 1.404 d 13
4700 2.533 d 12
70 μm 71000 4.222 d 12 3.724 d 12 4.875 d 12
19000 1.151 d 12
160 μm 156000 1.922 d 12 1.728 d 12 2.165 d 12
35000 4.367 d 11
DENIS
i 791 3.790 d 14 3.476 d 14 4.167 d 14
143 6.910 d 13
J 1228 2.441 d 14 2.265 d 14 2.647 d 14
191 3.820 d 13
KS 2145 1.398 d 14 1.306 d 14 1.503 d 14
302 1.970 d 13
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
124 Unit of luminous intensity (candela)
λ
fν = fλ . (8.21)
ν
49 See http://www.irsa.ipac.caltech.edu/data/SPITZER/docs/mips/mipsinstrumenthandbook/
50 See http://irsa.ipac.caltech.edu/IRASdocs/
8.2 Radiometry and photometry 125
So the units used by Bessell were essentially janskys expressed in a hybrid cgs and
SI system.
126 Unit of luminous intensity (candela)
Example 2
Convert the Vilnius U-band monochromatic flux density fλ(U =0) (Straižys, 1992)
expressed in units of W . cm−2 . μm−1 into janskys, for a hypothetical OV-type
star.
From Table 8.2, extract the appropriate values for bandwidth in wavelength (
λU )
and frequency (
νU ) space and substitute these values into Equation (8.22) to
determine the value of the monochromatic flux density in janskys:
Example 3
Convert the TD1-S68 ultraviolet satellite measure at λ = 274 nm (Thompson et al.,
1978) expressed in erg . cm−2 . s−1 . Å−1 into janskys for the A0V star Vega.
−1
f274 = 3.123 × 10−9 erg . cm−2 . s−1 . Å
= 3.123 × 10−9 (10−7 J) . s−1 . (10−2 m)−2 . (10−10 m)−1
= 3.123 × 10−9 × 10−7 × 104 × 1010 J . s−1 . m−2 . m−1
= 3.123 × 10−2 W . m−2 . m−1 (8.28)
From Table 8.2,
λ274 = 3.1 × 10−8 and
ν274 = 1.242×1014 Hz, so, using
Equation (8.22):
3.1 × 10−8
fν274 = 3.123 × 10−2 . . 1026 Jy
1.242 × 1014
= 779.5 Jy (8.29)
Example 4
HST spectrophotometry of the star Vega from 170 nm to 1010 nm. Bohlin &
Gilliland (2004) used observations obtained with the Space Telescope Imaging
Spectrograph (STIS) to determine the flux density in units of mW . m−2 . (0.1 nm)−1
(where 0.1 nm = 1 Å) against wavelength in angstroms. A plot of their data con-
verted to SI units (i.e., wavelength in metres and monochromatic flux densities in
W . m−2 . m−1 ) is shown in Figure 8.2.
The conversion from Bohlin & Gilliland (2004) units to janskys is carried out
for the flux measurement of 3.04×10−9 mW . m−2 . (0.1 nm)−1 determined at a
wavelength of 5800 Å and an FWHM measurement of 11.6 Å:
0.08
0.06
0.04
0.02
0.00
2 3 4 5 6710–6 2 3 4 5 67 –5 2 3 4 5 6710–4 2 3 4
10–7 10
wavelength (m)
frequency monochromatic flux density is derived from Equation (8.22). For the
numerical example, entry 2436 in the Bohlin & Gilliland (2004) tabulation is used
and firstly converted to SI units:
c = 299 792 458 m . s−1
λ = 5.80 × 10−7 m
λ = 1.16 × 10−9 m (8.31)
Convert from wavelength to frequency:
c
ν = = 5.169 × 1014 Hz
λ
c
νhfb = = 5.174 × 1014 Hz
λ − 0.5
λ
c
νlfb = = 5.164 × 1014 Hz
λ + 0.5
λ
ν = νhfb − νlfb = 1.034 × 1012 Hz
λ
fν = fλ . 1026 = 3415 Jy (8.32)
ν
As an additional example that utilizes the Bohlin & Gilliland (2004) data, a subset
was extracted bounded by the bandwidth of the Johnson V filter. In total, some 157
data points are available within the V bandwidth in the HST Vega data. As above,
(λ,
λ, fλ ) were converted to (ν,
ν, fν ) and a plot drawn (Figure 8.3) of ν against
fν . The mean monochromatic flux density f(ν=Vν ) , within the V frequency band,
is the sum of the product of the individual monochromatic flux densities multiplied
by the FWHM of each measured frequency sample (ν
ν) then divided by the sum
of the FWHM values, thus:
157
(fνk ) . (
νk )
f(ν=Vν ) = k=1 157 = 3646 Jy (8.33)
k=1 (
νk )
For the Bohlin & Gilliland (2004) data, the value of f(ν=Vν ) is 3646 Jy, which
may be compared with the Bessell (2001) value of 3636 Jy for a standard A0V star
of magnitude V = 0.00 (note that Bohlin & Gilliland, 2004 determined a value of
0.026 for the V magnitude of Vega).
3900
νeff
monochromatic flux density (Jy)
3700
3500
ga
Ve
3300
5.0 5.2 5.4 5.6 5.8 6.0
frequency (x 1014Hz)
5000
4000
monochromatic flux density (Jy)
3000
2000
1000
67 2 3 4 5 67 2 3 4 5 67 2 3 4 5 67 2 3 4 5 6
1012 1013 1014 1015
frequency (Hz)
Table 8.3. Monochromatic flux densities, either measured observationally for the
A0V star Vega or determined by other means for a standard A0V star
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
Sources for Table 8.3
1. IRAS (Infrared Astronomical Satellite) photometry (iras) at 12 μm, 25 μm, 60 μm and 100 μm for the star
Vega (Cohen et al., 1992).51
2. Johnson–Cousins–Glass (jcg). Monochromatic flux densities for the nine bands (U, B, V, R, I, J, H, K, L, M)
are those given by Bessell (2001) for an A0V star with magnitude V = 0.00.
3. The 2 Micron All Sky Survey (2MASS) measured monochromatic flux densities in three infrared bands (J,
H, KS ). Cohen et al. (2003) provided the absolute calibration for a zero-magnitude A0V star.
4. DENIS (den). Fouqué et al. (2000) used a synthetic spectrum of Vega to derive monochromatic flux densities
at magnitude zero for the three bands (i, J, KS ).
5. The Vilnius (vil) seven-colour (U, P, X, Y, Z, V, S) absolute monochromatic flux densities given by Straižys
(1992) in W . cm−2 . μm−1 units for an A0V star were converted to janskys using the transformation given
in example 2 above.
6. Geneva (gen) seven-colour (U, B, B1, B2, V, V1, G) photometric bandwidths with their absolute calibration
for Vega were determined by Rufener & Nicolet (1988), who used the relationship:
where Eν is the spectral irradiance at frequency ν in W . m−2 . Hz−1 , mν is the measured magnitude at
frequency ν, Kν is the colour index of the Geneva band at frequency ν relative to that at the effective
frequency of the B filter, and C is the zero-point shift for the B filter.
7. Strömgren (strom) photometry absolute calibrations of monochromatic irradiance are given by Sterken &
Manfroid (1992) in units of 10−11 W . m−2 . nm−1 for a star of spectral type A0V and magnitude V = 0.00.
Example 2 above shows how a similar unit is converted to janskys.
8. Astronomical Netherlands Satellite (ans) photometry. Wesselius et al. (1980) related magnitudes to fluxes in
W . m−2 . nm−1 (fλ ) using:
26.1+mλ
−2.5 +9
fλ = 10 (8.35)
which may be converted to janskys using Equation (8.22). From the ultraviolet magnitudes listed by Wes-
selius et al. (1982), the following magnitudes were extracted for Vega; ans1 = −0.491, ans2 = −0.441,
ans3 = −0.462, ans4 = no measurement, ans5 = 0.046, ans6 = 0.191. Substituting these magnitudes into
Equation (8.35) give values for fλ , which may then be converted to fν in janskys via Equation (8.22).
These monochromatic flux density values are shown in Table 8.3.
9. The Thor–Delta (td1) ultraviolet catalogue (Thompson et al., 1978) the wavelength-space monochromatic
flux densities of Vega in units of 10−10 erg . cm−2 . s−1 . Å−1 , which were converted to janskys in Table 8.3
as per example 3 above.
5000
4000
monochromatic flux density (Jy)
3000
2000
1000
67 2 3 4 5 67 2 3 4 5 67 2 3 4 5 67 2 3 4 5 6
1012 1013 1014 1015
frequency (Hz)
jansky, appropriately named after one of the pioneers of radio astronomy, Karl
Jansky.
The monochromatic flux density per unit solid angle (Jy . rad−1 ) is called the
brightness of the (extended) source.
In short, radio astronomers already use derived SI units that meet with the
approval of the IAU.
By way of an example, an attempt was made by Hollis et al. (1985) to detect Vega
at radio frequencies using the VLA (Very Large Array) of the US National Radio
Astronomy Observatory at a frequency of 4.86 GHz (λ ∼ 6 cm) with a beamwidth
of 40 μrad by 17 μrad (8.5 arcsec × 3.5 arcsec). No signal significantly greater than
the background noise (30 μJy) was detected.
134 Unit of luminous intensity (candela)
Example 1: convert 1 eV to Hz
1.602 177 × 10−19
ν = 1× = 2.418 × 1014 Hz (8.37)
6.626 076 × 10−34
Example 2: convert 1 keV, 1 MeV and 1 Gev to Hz
ν(1 keV) = 2.418 × 1017 Hz; ν(1 MeV) = 2.418 × 1020 Hz;
ν(1 GeV) = 2.418 × 1023 Hz (8.38)
where 1 keV = 1000 eV, 1 MeV = 106 eV and 1 GeV = 109 eV.
8.2 Radiometry and photometry 135
A typical observational X-ray spectrum has photon energy in keV as the abscissa
plotted against the monochromatic flux density in units of photons . cm−2 . keV−1
as ordinate.
1 ph . cm−2 . s−1 . keV−1 = (1.602 177 × 10−16 J)(4.136 × 10−14 ) m−2 . s−1 . Hz−1
= 6.6266 × 10−30 J . s−1 . m−2 . Hz−1
= 6.6266 × 10−30 W . m−2 . Hz−1 (8.40)
= 6.6266 × 10−4 Jy (8.41)
photons with a monochromatic flux density as low as 10−12 erg . s−1 . cm−2
(≡ 10−15 W . m−2 ) indirectly by means of the Čerenkov radiation emitted by the
γ -ray’s passage through the Earth’s atmosphere. The use of several telescopes in the
array allows the direction from which the γ -ray photon originated to be estimated.
For lower-energy γ -ray photons, detectors are commonly used with high-altitude
balloons, rockets and satellites. The most recent project is the Fermi γ -ray space
telescope mission launched by NASA in 2008. The principal instrument on this
satellite is the Large Area Telescope (Atwood et al., 2009), which covers the energy
range from 20 MeV to 300 GeV, corresponding to frequencies from 4.8×1021 Hz
to 7.5×1025 Hz.
In the first catalogue of active galactic nuclei detected by the Fermi Large Area
Telescope, Abdo et al. (2010), in common with most γ -ray astronomers, used
photons . cm−2 . s−1 as a unit for flux density, with the electron volt being used
instead of frequency in hertz. Atwood et al. (2009) used a mixture of units for flux
density, such as photons . cm−2 . s−1 and particles . m−1 . s−1 .
The unit transformations used in the section on X-ray astronomy apply equally
well to the γ -ray units.
1 ν
ν= = (8.43)
λ c
8.3 Magnitudes 137
103
102
monochromatic flux density (Jy)
101
100
10–1
10–2
10–4 radio
10–5
10–6
109 1010 1011 1012 1013 1014 1015 1016 1017 1018
frequency (Hz)
Figure 8.6. A log-log plot of HST spectrophotometry of Vega overlaid with broad-
and intermediate-band ultraviolet, optical and infrared photometry, with diamond
symbols for measures or attempted measures in the radio, submillimetre and X-ray
bands. The radio and X-ray plots are upper limits to the flux at that frequency.
8.3 Magnitudes
A very brief history (for more detail, see Hearnshaw 1996) of the use of magnitudes
in describing the brightness of stars is given in Chapter 12 and the present chapter
includes numerous examples of different magnitude schemes used over the past
century. Magnitudes are thoroughly engrained in the minds of astronomers, both
138 Unit of luminous intensity (candela)
professional and amateur. What are the advantages and disadvantages of using
magnitudes? Is there a better way that would link magnitudes more strongly and
logically to physical measurements?
m
f
f0 log2 log10 log √
5
100
ln
zero-point adjustment that transforms the measured magnitude of the detector into
that of a standard magnitude system.
Essentially, the simple revised magnitude system proposed changes Equation
(8.46) to:
–5
III
mV magnitude
5
V
10
–1
0 V
(B – V) colour
I
III
Figure 8.8. Colour–colour diagram showing the loci of luminosity class I (dark
grey), III (light grey) and V (black) stars as listed in Cox (2000) and Straižys (1992).
The colour of the stars changes from blue to red in the direction of increasing
value of the colour index. In the currently conventional colour–colour diagram,
the ordinate is plotted in the direction of decreasing (B–V) index.
where (fν )0 is the value, in janskys, of an A0V star of zero magnitude in the
passband that has an effective frequency of ν. The 0.921 is a conversion factor
relating natural logarithms to base-10 logarithms.
15 I
10
III
emV (ln(Jy))
–5
–1.5 –1.0 –0.5 0.0 0.5
(B – V) e – colour
length and time base units of the SI system. Astronomical units are rather more
esoteric. Comparisons between the different photometries are given and worked
examples of converting from one set of units to another given. For many of the
common astronomical photometric systems, the central wavelengths and band-
widths are transformed from wavelength to frequency space. Examples are given
of determining the spectral irradiance (radiometic unit) and the monochromatic flux
density of stars from many different published non-SI unit sources. Magnitudes,
a much-used astronomical photometric unit, are converted to monochromatic flux
densities and a flux versus frequency plot of HST spectrographic observations of
Vega produced.
Some selected commonly used photometric systems from the ultraviolet to the
infrared are calibrated in radiometric terms and in janskys. Very high frequency
astronomical observations in the X-ray and γ -ray regions are also converted to
janskys, with observations from the opposite end of the electromagnetic spectrum
in the submillimetre, millimetre and radio regions. Wavenumbers of three different
144 Unit of luminous intensity (candela)
I
0
e(U – B) colour
V
–1
–2
I III
Figure 8.10. e-colour–colour diagram showing the loci of luminosity class I (dark
grey), III (light grey) and V (black) stars as listed by Cox (2000). The stars colour
change from red to blue in the direction of increasing colour index along both axes.
types, μm−1 , cm−1 and m−1 are converted to frequencies in hertz and simple
worked examples given.
A discussion on the advantages and disadvantages of astronomical magnitudes
is given along with a proposed magnitude system based on radiometric measures
in janskys.
8.4.2 Recommendations
The major shortcoming of the defined SI photometric unit, the candela, when applied
to astronomy, is due principally to the restricted frequency band covered. However,
the radiometric units defined in terms of base SI units are easily related to astronom-
ical photometry. The great variety of different units in common use in astronomical
photometry makes inter-comparison between them time consuming and prone to
the making of mistakes.
For observational work, the most logical unit to use for monochromatic flux
densities is the jansky. It has been in use for many decades by both radio and
infrared astronomers and is not difficult to calculate from other similar units, which
8.4 Summary and recommendations 145
may differ by scale factors or by being tied to wavelength rather than frequency
space. The logarithm (whether natural or to base 10) of the flux density in janskys
makes a very acceptable physically based magnitude that allows for the continued
use of colour–magnitude and colour–colour diagrams. For both theoretical and
observational studies of the total power output of all types of astronomical bodies,
the watt is ideal.
9
Unit of thermodynamic temperature (kelvin)
The kelvin is a unit of thermodynamic temperature such that the Boltzmann constant
is exactly 1.380 650 5 × 10−23 J . K−1 (joules per kelvin).
146
9.2 Temperature scales 147
constant,54 i.e.
∂S −1
Tt = (9.1)
∂U N,V
The entropy S, is a measure of the disorder of the system and increases with
increasing levels of disorder (e.g., steam has a higher entropy than water, which in
turn has a higher entropy than ice).
m < v2 >
Tk = (9.2)
3k
54 See http://hyperphysics.phy-astr.gsu.edu/hbase/thermo/temper2.html
55 See http://hyperphysics.phy-astr.gsu.edu/hbase/kinetic/kintem.html
148 Unit of thermodynamic temperature (kelvin)
9.2.1 Converting the Celsius, Fahrenheit and Rankine scales to the Kelvin
temperature scale
Temperatures measured in the Celsius and Fahrenheit scales may be either positive
or negative and both define the freezing point of water with a pre-assigned value for
temperature (0 for celsius and 32 for fahrenheit). The kelvin and rankine have as
their lowest temperature value 0, known as absolute zero, the temperature at which,
in classical mechanical systems, all translational motion of atoms and molecules
ceases. In quantum mechanical systems there remains the possibility of zero-point
energy-induced particle motion.
The temperature intervals in the Kelvin and Celsius scales are the same,
1 K = 1 ◦ C, as are those in the Fahrenheit and Rankine scales, 1 ◦ R = 1 ◦ F. The
Kelvin and Rankine scales are both absolute temperature systems.
The IAU recommends citing temperatures in kelvin. To assist with their rec-
ommendation, the appropriate conversion formulae from one temperature system
to another is given. There are many online converters such as that given in the
Engineering Toolbox.56
θK = θC + 273.16 (9.4)
56 See http://www.engineeringtoolbox.com/temperature-d_291.html
9.3 Some examples of the temperatures of astronomical objects 149
57 See http://www.princeton.edu/∼willman/planetary_systems/Sol/Ceres/
58 See http://starchild.gsfc.nasa.gov/docs/StarChild/solar_system_level2/eris.html
150 Unit of thermodynamic temperature (kelvin)
Temperature
Object K ◦C ◦R ◦F
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
where Equation (9.9) gives the value of the Stefan–Boltzmann constant in SI units.
Temperature
Object K ◦C ◦R ◦F
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
Notes:
Cosmic microwave background radiation was discovered by Penzias & Wilson (1965). The value given in Table 9.2
is that obtained from measurements made using the COBE (COsmic Background Explorer) satellite by Mather
et al. (1994).
Ferriere (2001) lists various parameters of the different components of interstellar gas for the region of space
near the Sun. Temperatures are given for the molecular (the value in the table), cold and warm atomic, and warm
and hot ionized gas.
Lucas et al. (2010) reports on the discovery of a cool brown dwarf listed as UGPS J0722-05 in the UKIDSS
Galactic Plane Survey. A provisional spectral type (see Chapter 12) T1O has been assigned, though the possibility
of a new spectral class has been suggested. The temperature in Table 9.2 is the mean value of the range given.
The effective temperature (see below) of the brightest main sequence star (spectral type O5V) is given as
42 000 K in Cox (2000).
Weaver et al. (1978) give the central temperature of a 25 M (= 4.97 × 1031 kg) supernova at the onset of core
ignition for the silicon-to-iron-burning stage.
The Planck temperature is that which existed at the Planck time (5.38 × 10−44 s) after the Big Bang (Lang,
2006).
Planck’s law
The frequency distribution of the radiation emitted by a blackbody following
Planck’s law is a function of temperature alone. The spectral radiance or specific
intensity, Bν (T ), emitted by a blackbody whose temperature is T K, at a frequency
of ν Hz, is given by:
2 h ν3 1
Bν (T ) = 2 hν
W . m−2 . Hz−1 . sr −1 (9.11)
c ekT − 1
152 Unit of thermodynamic temperature (kelvin)
2hc2 1
Bλ = 5 hc
− 1 W . m−2 . m−1 . sr −1 (9.12)
λ e λkT
10
9 K
1 0
3.4x
T= e
7 K
0
x1
log(Bν) W . m–2 . sr –1 . Hz
0 1.6
T=
9K d
5 77
–10 T=
K
2 90 c
T=
b
K
2.73 a
–20
T=
8 10 12 14 16 18 20
log(ν) Hz
Figure 9.1. Blackbody curves computed using Equation (9.11). Curve ‘a’ is for a
blackbody at the same temperature as the cosmic background radiation, curve ‘b’
for the surface of the Earth, curve ‘c’ for the surface of the Sun, curve ‘d’ for the
centre of the Sun and curve ‘e’ for the centre of a 25 M star.
9.4 Blackbody radiation 153
En En nk − n1,2,3
E ν λ
Name n J eV J Hz nm
Lyman series
L∞ ∞ 0 0 ∞−1 2.18 d − 18 3.29 d 15 91
L 6 −6.06 d − 20 −0.82 6−1 2.12 d − 18 3.20 d 15 94
Lδ 5 −8.72 d − 20 −1.19 5−1 2.09 d − 18 3.16 d 15 95
Lγ 4 −1.36 d − 19 −1.85 4−1 2.04 d − 18 3.08 d 15 97
Lβ 3 −2.42 d − 19 −3.30 3−1 1.94 d − 18 2.92 d 15 103
Lα 2 −5.45 d − 19 −7.41 2−1 1.63 d − 18 2.47 d 15 122
1 −2.18 d − 18 −13.61
Balmer series
H∞ ∞ 0 0 ∞−2 5.45 d − 19 8.23 d 14 364
Hδ 6 −6.06 d − 20 −0.82 6−2 4.84 d − 19 7.31 d 14 410
Hγ 5 −8.72 d − 20 −1.19 5−2 4.58 d − 19 6.91 d 14 434
Hβ 4 −1.36 d − 19 −1.85 4−2 4.09 d − 19 6.17 d 14 486
Hα 3 −2.42 d − 19 −3.30 3−2 3.03 d − 19 4.57 d 14 656
2 −5.45 d − 19 −7.41
Paschen series
P∞ ∞ 0 0 ∞−3 2.42 d − 19 3.66 d 14 820
Pγ 6 −6.06 d − 20 −0.82 6−3 1.82 d − 19 2.74 d 14 1094
Pβ 5 −8.72 d − 20 −1.19 5−3 1.55 d − 19 2.34 d 14 1282
Pα 4 −1.37 d − 19 −1.85 4−3 1.06 d − 18 1.60 d 14 1875
3 −2.42 d − 19 −3.30
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
CONTINUUM
0
energy levels 10–1 8J
Hα Hβ
Pα
13.6eV
–1
Lα Lβ
–2
0 1 2 3 4 5 6 7
Figure 9.2. A Grotrian diagram for the Lyman (L), Balmer (H) and Paschen (P)
series of hydrogen lines. The vertical axis is measured in units of 10−18 J. The
energy difference in electron volts between the n = 1 and the n = ∞ is shown by
the two-headed dark grey arrow labelled 13.6 eV.
158 Unit of thermodynamic temperature (kelvin)
and
where j is the excitation potential above the ground state of the j th energy level
of the atom or ion and gj is the statistical weight (gj = 2 J + 1) of the j th energy
level, J being the total angular quantum number of the energy level.
i.e.,
N1 = 47.27 N0 (9.44)
so just over 2 % of the aluminium atoms are neutral and the rest are singly ionized.
F0 continuum level
Fλ
0
λmin λ1 λ0 λ2 λmax λ
Figure 9.3. The area enclosed by the Gaussian-shaped spectral line profile between
λmin and λmax , centred on wavelength λ0 is equal to the rectangular area, F0 . (λ2 −
λ1 ), whose width, (λ2 − λ1 ), is defined to be the equivalent width of the actual
spectral line.
9.5 Spectral classification as a temperature sequence 161
temperature (K)
CaII HI
TiO HeI
HeII
CaI
FeI FeII MgII
SiIII
M0 K0 G0 F0 A0 B0 O5
spectral type
Figure 9.4. A plot showing the variation of the strength (equivalent width) of
various atoms (e.g., FeI iron) and ions (e.g., FeII singly ionized iron) against both
spectral type and temperature. This figure was adapted, reversed and updated from
the original, which appeared in Struve et al. (1959). The black lines are for neutral
atoms, the dark grey lines are for ions and the wider light grey line is for the
molecule TiO. It can readily be seen that the ratio of equivalent widths of atoms
and/or ions to other atoms and/or ions is a function of spectral type and temperature.
162 Unit of thermodynamic temperature (kelvin)
and for Tc :
Tc = 12 340 K (9.54)
Note that this value is considerably higher than the value for the effective tem-
perature of 9790 K given by Cox (2000). Selecting other pairs of photometric bands
would produce different values for the colour temperature. By way of illustration,
consider the red and near-infrared bands R and I.
Tc = 7398 K (9.56)
The colour temperature does not represent the physical temperature within the
photosphere of a star, but rather may be used to give an indication of the slope of
the energy distribution emitted by a star.
where
Teff ,i+1 − Teff ,i
κ = Teff ,i + (B − V)i (9.58)
(B − V)i+1 − (B − V)i
O5
40 000
30 000 B0
Teff K
20 000
A0
10 000 F0
G0
K0
M0
M5
0
Sun
Figure 9.5. A colour index (B − V), versus (Teff ) for main sequence stars from
type O5 to M5 (Cox, 2000).
9.6 Model stellar atmospheres 165
59 See http://www.stsci.edu/hst/observatory/cdbs/k93models.html
166 Unit of thermodynamic temperature (kelvin)
The output of the Kurucz models is in the format (λ, fλ ), where λ is in nm and
Fλ is in erg . s−1 . cm−2 . Hz−1 . sr−1 .
erg . s−1 . cm−2 . Hz−1 . sr −1 = (10−7 ) J . s−1 . (10−2 )−2 m−2 . Hz−1 . sr −1
= 10−3 W . m−2 . Hz−1 . sr −1
= 1023 Jy . sr −1 (9.59)
The total emergent monochromatic flux from the star is 4 π times the unit stera-
dian value above. If the radius of the emitting star is ρ metres and its distance is
R , then the conversion factor, conv, between Kurucz units and janskys becomes:
ρ 2
× 1023
conv = 4 π (9.60)
R
or if fJ is the monochromatic flux density in janskys and fK is the monochromatic
flux density in Kurucz units, then:
ρ 2
. 1023 × fK
fJ = 4 π (9.61)
R
Example: plot the comparison between the absolute HST spectrum of Vega
and a Kurucz model stellar atmosphere
By the absolute HST spectrum of Vega is meant the unreddened spectrum that
would be observed at a distance of 10 pc or 3.085 677 6 × 1017 m. The model stellar
9.6 Model stellar atmospheres 167
Table 9.4. Stellar radii in solar radius and SI units against spectral class and
luminosity class
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
atmosphere is also computed as though the star were at a distance of 10 pc. The
Kurucz model selected has an effective temperature of 9750 K, a surface gravity
of log g = 4.00 and solar chemical abundances (log Z = 0.00). The radius of an
A0V star is, from Table 9.4, 1.593 × 109 m. Using the conversion between Kurucz
units and janskys given in Equation (9.61) and inserting the values for the stel-
lar radius and its distance, plus a value for the monochromatic flux density of
4.93 × 106 erg . s−1 . cm−2 . Hz−1 . sr−1 from the Kurucz model at a wavelength of
3 μm (or a frequency of 1014 Hz) gives:
1.593 × 109 2
fJ = 4 π × 1023 × 4.93 × 106
3.0856776 × 1017
= 165.03 Jy (9.62)
Figure 9.6 shows segments of the observed and theoretical spectra from 1014 Hz
to approximately 2 × 1015 Hz.
104
8
7
6
5
4
monochromatic flux density (Jy)
3
a
Veg
2
103
8 Hα
7
6
el
5
4 od
c zm
3 ru
Ku
2
102
2 3 4 5 6 7 8 2
1014 1015
frequency (Hz)
Figure 9.6. Comparison between the HST spectrum of the A0V star Vega and a
Kurucz model stellar atmosphere, with parameters T = 9 750 K, log g = 4.00 and
log Z = 0.00. The effective temperature was deliberately set too low to allow for
easier comparison between the model-generated spectrum lines and the actual HST
spectrum. Note that the monochromatic flux densities of Vega and the Kurucz
model have been adjusted to appear as though they were both at the standard
distance of 10 pc.
simply a matter of summing the product of the flux elements φν and the elemental
frequency bandwidths δν and dividing by the sum of the frequency bandwidth
elements, i.e.,
φ ν δν
fν = (9.63)
δν
Since, in general, the maximum and minimum frequency values lie between tab-
ulated model frequencies, end corrections using simple linear interpolation were
made. Monochromatic flux densities were calculated for various photometric bands
from the satellite ultraviolet (the ANS and TD1 satellites), through the optical spec-
trum (Vilnius and Johnson photometry), to the infrared (USNO-B I band, 2MASS
and IRAS photometry). The number of model data points available per band var-
ied from 1 for the IRAS 60μm and 100μm bands to a maximum of 95 for the
9.6 Model stellar atmospheres 169
Table 9.5. Wavelength frequency and mean monochromatic flux density data for
Kurucz model (Teff = 9750 K, log g = 4.00, log Z = 0.00) photometric bands
Band λ
λ ν
ν Fν
ID nm nm THz THz Jy nK
ANS1 154 5 1940 63 206.4 7
ANS2 154 14 1935 187 215.9 18
ANS3 179 14 1666 138 327.9 17
ANS4 220 20 1363 124 390.0 22
ANS5 249 15 1203 72 397.7 18
ANS6 329 10 910 28 623.4 7
TD1 156 33 1916 409 216.1 38
TD2 196 33 1526 258 367.6 36
TD3 236 33 1268 178 382.4 36
TD4 274 31 1094 124 466.2 32
U 345 40 869 101 650.3 23
P 374 26 802 56 968.4 15
X 405 22 740 40 1922.5 13
Y 466 26 643 36 2005.0 15
Z 516 21 581 24 1809.1 12
V 544 26 551 26 1718.5 14
S 656 20 457 14 1322.9 12
U 367 66 817 148 955.6 36
B 436 94 688 150 1980.5 48
V 545 88 550 89 1728.5 46
I 808 185 371 86 1132.5 95
J 1235 162 243 32 725.9 34
H 1662 251 180 28 459.3 34
KS 2159 262 139 17 302.7 28
L 3600 1200 83 29 126.9 71
M 4800 800 63 11 71.2 42
12 μm 12 000 7000 25 16 15.1 42
25 μm 25 000 11 150 12 6 3.0 3
60 μm 60 000 32 500 5 3 0.5 1
100 μm 100 000 31 500 3 1 0.2 1
USNO-B I band. Table 9.5 lists the band identification, mean wavelength λ, band-
width
λ, mean band frequency ν, frequency bandwidth
ν, the mean absolute
monochromatic flux density, Fν , emitted by the model star atmosphere at a distance
of 10 pc, and number nK of Kurucz model data points used in computing the mean
monochromatic flux density. Wavelength units are nm and frequency units THz
(1012 Hz).
Figure 9.7 illustrates the output for the USNO-B I band of the Kurucz model
stellar atmosphere with Teff = 9750 K, log g = 4.00 and log Z = 0.00. The model
170 Unit of thermodynamic temperature (kelvin)
1300
1200
monochromatic flux density (Jy)
1100
1000
900
800
32 34 36 38 40 42
13
frequency x 10 (Hz)
Figure 9.7. USNO-B I band mean monochromatic flux density and frequency
bandwidth against Kurucz model stellar atmosphere output. The νmin and νmax
values for each monochromatic flux density sample computed are shown, which
creates the step-function-like appearance. The construction of the I band used
95 sample fluxes from the Kurucz model for T = 9750 K, log g = 4.00 and
log Z = 0.00. The light-grey line shows the mean monochromatic flux measured in
the bandwidth shown by the horizontal light-grey line. The black square is located
at the mean frequency and mean monochromatic flux density of the band.
104
8
7 HD73952
6
5
4 Hα
monochromatic flux density (Jy)
2
K
750
11
=
103 T
8
7
6
5
4
3
T = 9750 K
2
102
2 3 4 5 6 7 8 2
1014 1015
frequency (Hz)
Figure 9.8. A comparison between two computed model stellar atmospheres, with
log g = 4.00, log Z = 0.00 and temperatures 9750 K and 11 750 K, with actual pho-
tometric observations of the B8Vn star HD73952. The error bars attached to each
observational point represent the photometric bandwidth and an estimate of the
total error in the photometry. Circles are TD1 photometry, squares are Vilnius pho-
tometry, the inverted triangle is the USNO-B I band and the diamonds 2MASS
photometry. The monochromatic flux densities in janskys are adjusted to those that
would be observed were the models and the star at the standard 10 pc distance.
density. The error bar in the x direction being the frequency bandwidth and that
in the y direction being due to the uncertainty in the observational photometric
measurement.
300
200
100
The final section illustrates how model stellar atmospheres in conjunction with
coarse (broadband and intermediate-band photometry) and standard spectropho-
tometry may be used to calculate stellar temperatures.
9.7.2 Recommendations
The basic recommendation to most astronomers would be to carry on with what
they are doing now, as most already use the SI unit, the kelvin, as a measure of
temperature.
Some popular astronomical publications often cite temperatures in degrees Cel-
sius or Fahrenheit and should be encouraged to show the equivalent temperature
in kelvin in parenthesis. This would lead to an increasing familiarity with kelvin
amongst the interested laity.
Whilst it is possible to replace spectral class and dereddened colour indices by
a measure of temperature, there would seem to be no really good reason for doing
so, though it may be argued that for ease of comparison between observations and
theory, similar variables should be used.
10
Unit of electric current (ampere)
The ampere is a unit of electric current such that the elementary charge is exactly
1.602 176 53 × 10−19 C, where 1 C (coulomb) = 1 A . s (ampere second).
Electricity: A material agency which, when in motion, exhibits magnetic, chemical and
thermal effects, and which, whether in motion or at rest, is of such a nature that when it is
present in two or more localities within certain limits of association, a mutual interaction
of force between such localities is observed.
That branch of science that treats of this agency and the phenomena caused by it.
174
10.2 SI and non-SI electrical and magnetic unit relationships 175
q 2 = fG r 2 (10.3)
so
1
q = fG2 r (10.4)
In dimensional terms, force and length may be expressed as:
so the dimension of the unrationalized cgs Gaussian unit of charge dim[q] is:
1
dim[q] = ([L] . [M] . [T ]−2 ) 2 [L]
1 3
= [M] 2 . [L] 2 . [T ]−1 (10.7)
dim[Q] = [I ] . [T ] (10.8)
In Table 10.1, a selection of named electric and magnetic derived units with
dimensions and combinations of base units are given for both SI and unrationalized
cgs Gaussian units.
The last two entries in Table 10.1 may be defined in the following manner.
10.2 SI and non-SI electrical and magnetic unit relationships 177
Table 10.1. Names, symbols and dimensions of various electric and magnetic
units in SI and unrationalized cgs Gaussian form
where fSI is the force in joules between two charges q1 and q2 , measured in
coulombs, separated by a distance of r metres. The appearance of the factor 4 π
is due to SI units being rationalized. In general, for equations set out in rational-
ized units, 4 π appears in those involving spherical symmetry, 2 π in those with
cylindrical symmetry and none in those with plane symmetry.
The magnetic permeability of free space, μ0 , forms part of the constant of pro-
portionality ( 4μπ0 ) in the equation relating the magnetic force fM in joules, between
two current elements i1 and i2 in amperes, in conductors 1 and 2, of length dl1 and
dl2 metres, separated by distance r1,2 metres, where r1,2 is measured along a line
which makes an angle θ in radians to the direction of the flow of the current in
conductor 1.
μ0 −2
dfM = r i1 dl1 i2 dl2 sin θ (10.11)
4 π 1,2
Note that the constants 0 and μ0 are related in SI units by the expression:
1
0 μ0 = (10.12)
c2
where c is the velocity of light (Bleaney & Bleaney, 1962).
For a more complete treatment of the above material see, e.g., Coulson & Boyd
(1979), Bleaney & Bleaney (1962) and Menzel (1960).
10.2 SI and non-SI electrical and magnetic unit relationships 179
Force field dF = E + V ∧ B dF = E + V∧ B
dq dq c
Electric field D = 0 E D=E
0 μ0 = c−2 0 μ0 = 1
Polarization P = D − 0 E P = 41π D − 0 E
Magnetization M = μB − H M = 41π μB − H
0 0
Ampère’s law ∇ ∧ H = J + ∂∂D
t ∇ ∧ H = 4 πc J + 1c ∂∂D
t
Maxwell’s div D = ρ div D = 4 πρ
equations
div B = 0 div B = 0
curl E = − ∂B curl E = − 1c ∂B
∂t ∂t
curl H = σ E + ∂∂D
t curl H = 1c 4 π σ E + ∂∂D
2 t
Wave equation ∇ 2 (E , H) = μ μ0 0 ∂ 2 (E , H) μ
∇ 2 (E , H) = 2 ∂ 2 (E , H)
2
∂t c ∂t
qG2
fG = (10.13)
rG2
qG2 = rG2 fG
= 1002 × 105 = 109
9
qG = 10 2 statcoulombs (10.14)
Second, how many coulombs per charge, qSI , are required to produce a force fSI
of 1 N when the charges are rSI = 1 m apart?
180 Unit of electric current (ampere)
2 1
qSI = r 2 fSI
4 π 0 SI
1
= (10.16)
4 π 0
Now 0 is a quantity defined by:
1
0 = (10.17)
μ0 c2
where c, the speed of light in vacuo, is a defined constant, as is μ0 , the permeability
of free space, hence 0 also has a fixed value. Substituting for μ0 ( = 4 π × 10−7 )
gives:
1
0 =
4 π × 10−7 c2
1
4 π 0 = −7 2
10 c
1 1
√ = (107 c−2 ) 2 (10.18)
(4 π 0 )
Comparing the values for the identical charges in the cgs Gaussian and SI units
gives:
7 9
10 2 c−1 (coulombs) ≡ 10 2 statcoulombs
9 7
10 2 − 2
1 C = −1 (statcoulombs)
c
= 10 c
= 2.997 924 58 × 109 statcoulombs (10.19)
Magnetic flux, , is usually defined in terms of magnetic flux lines as the group
or number of such lines emitted outwards from the north pole of a magnetic body.
Its unit is the weber and its symbol Wb.
Magnetic flux density, B, is defined to be the amount of magnetic flux per unit
area perpendicular to the direction of the magnetic flux. Its unit is the tesla and its
symbol T. These quantities are related by the simple expression:
B= (10.21)
A
MAGNETIC
FLUX
Φ
θ
Next check the dimensional consistency between gauss and tesla, the dimensions
for each unit are given in Table 10.1 as:
1 3
dim[gauss] = [M] 2 . [L]− 2 (10.22)
dim[tesla] = [M] . [T ]−2 . [I ]−1 (10.23)
Also from Table 10.1, the cgs Gaussian dimension for electric current IG is given as:
1 3
dim[I ]G = [M] 2 . [L] 2 . [T ]−2 (10.24)
substitute for [I ] in Equation (10.23) to give:
1 3
dim[tesla] = [M] . [T ]−2 ([M] 2 . [L] 2 . [T ]−2 )−1
1 3
= [M] 2 . [L]− 2
= dim[gauss] (10.25)
so the gauss and the tesla are dimensionally consistent.
Mass in cgs Gaussian units is measured in grams (g) and length in centimetres
(cm); in SI units, mass is measured in kilograms (kg) and length in metres (m).
For a magnetic flux of 1 weber (Wb) passing perpendicularly through an area of
1 square metre (m2 ), the magnetic flux density is BSI = 1 T (tesla).
For a similar magnetic flux of 1 Wb (= 108 Mx (maxwells) in cgs Gaussian units)
passing through a 1 cm2 (= 10−4 m2 ) area, the magnetic flux density is given by:
BG = 10−4 T = 1 G (10.26)
and it follows that:
1 Wb = 1 T . m2
= 1 . (104 G) . (104 cm2 )
= 108 G . cm2
= 108 Mx (10.27)
A cgs Gaussian unit of magnetic flux density that was in common use in
astrophysics was the gamma γ , a submultiple of the gaussian given by:
1 γ = 10−5 G = 10−9 T (10.28)
A listing of transformations and conversions for other electric and magnetic units
from cgs Gaussian to SI is given in Table 10.3.
An example of an online magnetic unit converter is given at: http://www.
smpspowersupply.com/magnetic-unit-conversion.html.
10.3 Magnetic fields in astronomy 183
Table 10.3. Names, symbols and dimensions of various electric and magnetic
units in SI and unrationalized cgs Gaussian form
Table 10.4. Magnetic flux densities (B) of various celestial objects in cgs
Gaussian and SI units
BG BSI
Object Notes G T
Solar System
Interplanetary space 1 d −6 − 1 d −5 1 d −10 − 1 d −9
Mercury equatorial value 3.4 d −3 3.4 d −7
Venus <4 d −6 <4 d −10
Earth 0.31 3.1 d −5
Mars <5 d −6 <5 d −10
Jupiter 4.3 4.3 d −4
Saturn 0.22 2.2 d −5
Uranus 0.23 2.3 d −5
Neptune 1.4 1.4 d −4
Sun
Sunspot penumbra 0.8 d 3 − 2.0 d 3 0.08 − 0.2
umbra 2d3−4d3 0.2 − 0.4
Plage 1.4 d 3 − 1.7 d 3 0.14 − 0.17
Prominences horizontal field 2 − 40 2 d −4 − 4 d −3
Corona 1 d −5 − 1 d 2 1 d −9 − 1 d −2
Interstellar space
general field 1 d −6 − 1 d −5 1 d −10 − 1 d −9
Stars
Vega A0V star 0.6 6 d −5
FK Com cool giant 60 − 272 6 d −3 − 2.7 d −2
Be 10 − 100 1 d −3 − 1 d −2
WR 1500 0.15
Flare ≥1000 ≥0.1
T Tau 1d3 − 1d4 0.1 − 1
WD 1d5 − 1d9 10 − 1 d 5
Neutron stars 1 d 12 − 1 d 13 1d8 − 1d9
Pulsar surface field 1 d 12 1d8
SGR 1806 -20 magnetar 2.1 d 15 2.1 d 11
Gaseous nebulae
Planetary 1 d −4 − 1 d −3 1 d −8 − 1 d −7
SNR 1 d −5 − 1 d −2 1 d −9 − 1 d −6
Galaxy
z = 0.692 8.4 d −5 8.4 d −9
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
published are in the cgs Gaussian unit, the gauss, though a few are presented in the
SI unit, the tesla. Both units are given in Table 10.4.
Table 10.4 was compiled, in part, from the following sources: Cox (2000), Land-
street (2001), Gnedin (1997), Reddish (1978), Bemporad & Mancusco (2010),
10.3 Magnetic fields in astronomy 185
cgs Gaussian SI
Total magnetic flux at/in Mx Wb
Solar minimum 1.5 − 2.0 d 23 1.5 − 2.0 d 15
Solar maximum 1.0 − 1.2 d 24 1.0 − 1.2 d 16
Small active region 3 d 20 3 d 12
Large active region ≥1 d 22 ≥1 d 14
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
cgs Gaussian SI
Planet G . cm−3 T . m−3
Mercury 2 − 6 d 22 2 − 6 d 12
Venus <1 d 21 <1 d 11
Earth 7.84 d 25 7.84 d 15
Mars <1 d 22 <1 d 12
Jupiter 1.55 d 30 1.55 d 20
Saturn 4.6 d 28 4.6 d 18
Uranus 3.9 d 27 3.9 d 17
Neptune 2.2 d 27 2.2 d 17
In this table the shorthand notation m d n has been used where m d n ≡ m × 10n .
Lignières et al. (2009), Korhonen et al. (2009) and Wolfe et al. (2008). The surface
magnetic flux density of the magnetar was extracted from the McGill SGR/AXP
Online-Catalog.60
Other than magnetic flux density, Cox (2000) also lists examples of total magnetic
flux for the Sun and active solar regions (see Table 10.5) and planetary magnetic
dipole fields (see Table 10.6).
The relationship between cgs Gaussian units and SI units in Table 10.5 is simply:
1 Wb = 108 Mx.
In Table 10.6, the relationship between the cgs Gaussian unit and the SI unit is:
60 See http://www.physics.mcgill.ca/∼pulsar/magnetar/main.html
186 Unit of electric current (ampere)
Zeeman effect
The effect of a magnetic field of flux density B on the spectrum of an atom with
singlet lines (S = 0 where S is the spin value) is to split such lines into 2 J + 1
(where J is the total angular momentum) equally spaced lines either side of the
original singlet line. In cgs Gaussian units, the energies EM , in ergs, of the levels
are given by:
ehB M
EM = E0 ± (10.30)
4 π me c
where E0 is the energy level of the zero field singlet, e is the charge of the electron
in statcoulombs, where:
1 C = 10 c stat C (10.31)
me is the rest mass of the electron in grams, h is Planck’s constant in erg . s, c is the
speed of light in cm . s−1 , B is the magnetic flux density in gauss, and M takes an
integer value of 0 or ± 1 (see Lang 2006).
Since:
EM = h νM (10.32)
eB M
ν M = ν0 ± (10.33)
4 π me c
where νM is the frequency of the singlet line split by the magnetic field and ν0 is
the frequency of the zero field singlet. Now, Equation (10.34) may be rewritten in
SI units as:
eB M
νM = ν0 ± (10.34)
4 π me
10.3 Magnetic fields in astronomy 187
Example: show that the dimension of the shift due to the Zeeman splitting
of a singlet spectral line in both cgs Gaussian and SI units is the same
(a) cgs Gaussian case:
eG BG
νG = (10.35)
4 π mG cG
Use Table 10.1 to give the dimensions for each of the units:
dim[eG ] . dim[BG ]
dim[
νG ] =
dim[mG ] . dim[cG ]
3 1 3 1
([L] 2 . [M] 2 ] . [T ]−1 ) . ([L]− 2 . [M] 2 . [L] . [T ]−1 )
=
[L] . [M] . [T ]−1
= [T ]−1 (10.36)
The unit which has a dimension of [T ]−1 is the hertz, the unit of frequency.
(b) SI case:
eSI BSI
νSI = (10.37)
4 π mSI
Again, using Table 10.1:
dim[eSI ] . dim[BSI ]
dim[
νSI ] =
dim[mSI ]
([I ] . [T ]) . ([M] . [T ]−2 . [I ]−1 )
=
[M]
= [T ]−1 (10.38)
So the frequencies of the singlet line and the first pair of lines either side due to the
presence of a magnetic field of flux density B T are given by:
Example: at the rest frequency of the 21-cm HI line, what would be the
approximate amount of Zeeman splitting resulting from a background
magnetic flux density of 1 nT, similar to that found in the Milky Way Galaxy?
(a) SI case:
ν21 = 1.420 405 751 768 × 109 Hz (Cox, 2000) is the radio frequency of the HI line,
which corresponds approximately to a wavelength of 21.106 114 cm. The frequency
separation, in a magnetic field of flux density of 10−9 T, is 2
ν21 , and may be
written as:
eSI BSI
2
νSI = 2
4 π mSI
(1.602 × 10−19 ) (10−9 )
2
4 π (9.109 × 10−31 )
28 Hz (10.42)
or
λ2
λ = ∓
ν (10.47)
c
Now, in SI units, the frequency shift
νSI , in hertz, is given by Equation (10.38),
substitute
λSI for
νSI thus:
λ20
λSI = ∓
νSI
c
λ2 eSI
=∓ 0 BSI (10.48)
4 π mSI c
where λ0 is the wavelength in metres of the spectral feature in the absence of a
magnetic field and λ0 ∓
λSI are the central wavelengths, in metres, of the first
Zeeman components formed in the presence of a magnetic field of flux density BSI
in teslas, with eSI being the charge of the electron in coulombs and mSI the mass
of the electron in kilograms.
If cgs Gaussian units are used with the charge of the electron eG measured in esu
charge units (statcoulombs), the electron mass mG in grams, the wavelengths λ in
centimetres, and the speed of light in cm . s−1 , then by substituting for
νG from
Equation (10.36), Equation (10.48) gives:
λ20 eG
λG = ∓ BG (10.49)
4 π mG c 2
where BG is the magnetic flux density in gauss.
In optical astronomy, the variables that are measured are λ0 and
λ, and that
which is to be determined is B. Equations (10.49) and (10.50) may then be
rewritten as:
4 π c mSI
λSI
BSI = (10.50)
λ20 eSI
and
4 π c2 mG
λG
BG = (10.51)
λ20 eG
Example: determine the global surface magnetic flux density of the magnetic
A-type star HD94660 in both teslas and gauss
(a) SI case:
Approximate measurements were made of a small portion of the spectrum of
HD94660 between 6 144 Å and 6 154 Å (as presented by J. D. Landstreet at the
Leverhulme lecture on stellar magnetism.). The Zeeman doublet of interest in this
narrow spectral region is due to singly ionized iron (FeII), with a laboratory-
measured central wavelength of 6 149.238 Å (Moore, 1959). The separation of
190 Unit of electric current (ampere)
Since 1 T = 104 G, the global magnetic field values in each set of units are equal
to one another and may be compared with the value of − 2200 G (= − 0.22 T) for
the mean longitudinal magnetic field obtained by Bagnulo et al. (2002) using the
FORS1 spectropolarimeter at the ESO VLT.
The unit (rad . m−2 ) of the rotation measure is not dependent on which of the
cgs Gaussian, IAU/SI or SI units is used.
The unit of the dispersion measure in cgs Gaussian units is the pc . cm−3 , the
appropriate unit in IAU/SI units would be pc . m−3 and that in pure SI units is
m . m−3 . The conversions from cgs Gaussian to IAU/SI and SI units and gauss to
teslas are:
Applying these conversion factors to the cgs Gaussian dispersion measures alters
the proportionality coefficient, PG , in the following ways:
430 MHz. Included in their programme was the pulsar J2022+2854, at a distance of
around 3 kpc (Verbiest et al., 2010). The observed values for the rotational measure
and dispersion measure were: Rm = −73.7 rad . m−2 and Dm = 24.623 pc . cm−3 .
Given these measurements, the mean line-of-sight magnetic flux densities, < B|| >,
were calculated in each of the systems of units.
(a) cgs Gaussian units:
R
m
< B|| >G = PG (10.64)
Dm
where
Rm = −73.7 rad . m−2 (10.65)
so
−73.7
< B|| >IAU = 1.232 × 10−4
2.4623 × 107
= −3.688 × 10−10 T = −368.8 p T = −3.688 μG (10.70)
(c) SI units:
So, to calculate the galactic field magnetic flux density in teslas from Faraday
rotation measures, input Rm in units of rad . m−2 , Dm in units of m . m−3 and use
PSI as the value for the coefficient of proportionality.
194 Unit of electric current (ampere)
Jupiter’s lightning
Lightning on Jupiter was detected by the Cassini spacecraft whilst enroute to Saturn
in early 2001 and the observations analyzed by Dyudina et al. (2004). The most
powerful lightning storm recorded was measured to emit 800 MW of power in the
Hα line, corresponding to some 40 GW of broadband optical power.
Saturn’s aurorae
A study of the electric currents in the polar ionosphere of Saturn was carried out by
Cowley et al. (2004). They reported potentials in kV units, electron energy fluxes in
mW . m−2 , upward field-aligned current intensities in mA . m−1 and particle number
densities in cm−3 (cgs Gaussian units); in SI, the unit would be particles . m−3 .
10.5.2 Recommendations
Units of electricity and magnetism are manifold, and great care needs to be taken
in ascertaining to which of the five major systems the unit belongs (esu, emu,
cgs Gaussian, practical or SI). The conversion coefficients are not just powers of
ten, but often involve the physical and mathematical constants c, μ0 , 0 and π .
196 Unit of electric current (ampere)
Given that the IAU strongly recommends using SI units in all cases and that the
most common electric or magnetic unit used by astronomers is that for magnetic flux
density, such an outcome should be easy to accomplish. Perusal of relevant astro-
nomical literature to date indicates that SI units are becoming the unit of choice in
this area, so perhaps all that is needed is to encourage those still using cgs Gaussian
for research or teaching to change to SI. Until such a desirable situation is achieved,
a very strong recommendation is made to all astronomers when publishing results
to specify precisely which system or systems of units they are using.
11
Unit of amount of substance (mole)
The mole is an amount of substance such that the Avogadro constant is exactly
6.022 141 5 × 1023 mol−1 (per mole).
197
198 Unit of amount of substance (mole)
and that the mass of one atomic mass unit is given by:
1.992 646 54 × 10−26
1 amu =
12
= 1.660 538 78 × 10−27 kg (11.3)
In cgs units (still in common use in modern university textbooks, e.g., Chang
2005) the mass of one atom of 12 C is 1.992 646 54 × 10−23 g and that of 1 amu is
1.660 538 78 × 10−24 g. The conversion factor between cgs units and SI units is
0.001 (1 g = 0.001 kg; 1 kg = 1000 g).
of Saturn’s largest moon, Titan, to determine the stratospheric mole fractions of the
molecules CH4 (methane) and CO (carbon monoxide) to be 1.6 ± 0.5 × 10−2 and
4.5 ± 1.5 × 10−5 , respectively.
A comprehensive listing of the average atomic masses for each of the elements is
given by Wieser & Berglund (2009) and a periodic table based on the data by G. P.
Moss is available on the web.62
The average atomic mass m of an element with n isotopes, where the mole
fraction of the ith isotope of mass mi is xi , is given by:
n
m= xi mi (11.6)
i=1
62 See http://www.chem.qmul.ac.uk/iupac/AtWt/table.html
200 Unit of amount of substance (mole)
CO
molecular mass of CO = (atomic mass of C) + (atomic mass of O)
= 12.011 + 15.999
= 28.010 amu (11.8)
Note that the mass of 1 mole of CH4 is 0.016 043 kg and that of CO is
0.028 010 kg, since the molar mass in kg of a molecule is numerically equal to
0.001 times its molecular mass (there is a direct equivalence if cgs units are used,
i.e., 1 mole of CH4 has a mass of 16.043 g and 1 mole of CO has a mass of 28.010 g).
For further examples of the use of the mole, see Mills et al. (1993).
The SOHO63 spacecraft also records the speed of the solar wind and its particle
density at a given time. For example, at 23 h 44 m on 12 October 2010 the solar
63 See http://sohowww.nascom.nasa.gov/
11.2 Avogadro’s constant and atomic masses 201
wind speed was 363 km . s−1 and its particle density in cgs units, 3.89 cm−3 , which
is equivalent to 3.89 × 106 m−3 in SI units since 1 cm−3 ≡ 106 m−3 .
An example of a ground-based determination of isotopic ratios is the examination
of five silicon carbide grains by Amari et al. (1992) obtained from the Merchison
carbonaceous meteorite. The sample produced values for the 26Al / 27Al isotopic
ratio that varied from 0.20 ± 0.01 to 0.61 ± 0.04.
Q
F= (11.9)
n
The dimension of the Faraday constant is [I ] . [T ] . [N ]−1 and its unit C . mol−1 .
Since the charge carried by a single monovalent ion is equivalent to the electric
charge e of an electron, then Equation (11.9) may be rewritten as:
F = NA e (11.10)
These three gas laws may be combined to create a universal, ideal or perfect gas
law in which:
VP
R= (11.12)
nT
where R is the universal or molar gas constant that has dimension
[L]2 . [M ] . [T ] . [N ]−1 . [
]−1 and unit J . mol−1 . K−1 , and is equal to (Mohr et al.,
2007):
R = 8.314 472 J . mol−1 . K −1 (11.13)
In cgs units, R = 8.314 472 × 107 erg . mol−1 . K−1 (see, e.g., page 94 et seq.
in Aller (1963) for a complete cgs treatment of the ideal gas laws as relevant to
astrophysics).
Table A in Viala (1986) also contains an entry, dE, for variation in enthalpy
(defined by Brimblecombe et al. 1998 as the heat energy associated with a chem-
ical change) during the chemical reaction, with a negative value indicative of an
exothermic reaction and positive for an endothermic reaction. The unit used is the
kcal . mol−1 .
Example: convert the value of the variation in enthalpy dE from cgs units to
SI units for the chemical reaction C + H2 O → C O + H2
From Viala (1986), the value of dE for this reaction given in Table A is
−140.076 kcal . mol−1 and there is an exact relationship between calories and
joules, such that 1 cal ≡ 4.184 J and 1 kcal ≡ 4184 J, hence:
11.4.2 Recommendations
Wilkins (1989) makes no specific reference to the mole or its usage, though the SI
version is implied from the overriding recommendation from the IAU that all SI
units are acceptable.
The limited appearances of the mole in modern astronomical literature means
that it has to be assumed that a definition of the unit in a particular source is based on
12 C rather than the earlier usage of 16 O or 1 H. When compound units are used, there
is a tendency for some parts of the unit to come from the cgs range (e.g., cm3 . mol−1
rather than m3 . mol−1 ) and, as is not uncommon in astronomical papers, textbooks
or reference works, sometimes a mixture of both cgs and SI units. This practice
should be avoided as confusion is the likely outcome.
12
Astronomical taxonomy
The act or process of arranging by classes; a grouping into classes; the putting together of
like objects or facts under a common designation; a process based on similarities of nature,
attributes, or relations. Classification may proceed by the gathering together of similar things
into a class, or by the unfolding of general groups into narrower or more specific divisions.
206
12.3 Classification of stellar objects 207
Group Subgroup
The Universe Clusters of galaxies
Clusters of galaxies Galaxies
Intergalactic gas
Intergalactic dust
Galaxies Stars
Star clusters
Interstellar gas
Interstellar dust
Stars Double stars
Exoplanets
The Solar System
The Solar System The Sun
Solar wind
The planets
Dwarf planets
Satellites
Small Solar System bodies
taxonomy at the present time are quantitative, i.e., they depend on the measurement
or measurements of physical attributes (position, size, distance, motion, emergent
flux etc.) made using objective instrumentation. Earlier forms of taxonomy relied
on a qualitative assessment of the observational attribute, e.g., in estimating the
brightness of a star by visual comparison with others nearby of known brightness,
or the assignment of a spectral class from the examination of a spectrogram, or deter-
mining the morphological type of a galaxy from a photographic or digital image.
Whilst such subjective methods are still in use, they are far less common than
they once were. Examples of objective measurements and classifications, where
relevant, are given in the chapters on individual SI units. This chapter contains
some examples of subjective classifications that originally depended solely on a
visual estimation of some parameter or ratio, e.g., the comparison of spectral line
strengths, the apparent shape of a galaxy or the magnitude of a star. Such classi-
fications are generally dimensionless. Table 12.1 sets out one possible scheme for
subdividing the Universe into classifiable groupings and subgroupings.
Constellations
The entire sky is subdivided into 88 groups of stars of varying sizes, called constel-
lations, well-known examples of which are Orion, Ursa Major (The Great Bear),
Gemini (The Twins), Crux Australis (The Southern Cross) and Scorpio (The Scor-
pion). Some date back thousands of years and others, mainly constellations near
the south celestial pole, are more recent inventions that followed the exploration of
the southern hemisphere by European navigators. In the early seventeenth century,
just prior to the use of the telescope for astronomical purposes, Bayer produced
a star atlas (the Uranometria) in which the stars forming the constellations were
identified with a Greek letter, assigned approximately in order of decreasing star
brightness from α to ω. Thus, Sirius is also known as α Canis Majoris and Mirzam,
the second brightest star in the constellation, as β Canis Majoris. If all the Greek
letters are used, then the letters of the Roman alphabet or numbers assigned by
Flamsteed (the first Astronomer Royal, appointed in 1675) are used.
12.3 Classification of stellar objects 209
Catalogue Identifier
Name Sirius, Dog Star
Uranometria α CMa, 9 CMa
Bonner Durchmusterung BD-16 1591
5th Fundamental Catalogue FK5 257
Henry Draper Catalogue HD 48915
HIPPARCOS Catalogue HIP 32349
Bright Star Catalogue HR 2491
Infrared Astronomical Satellite IRAS 06429-1639
ROSAT (X-ray) RX J0645.1-1642
Smithsonian Astrophysical Observatory SAO 151881
TD1 (Ultraviolet Photometry) TD1 8027
64 See http://simbad.u-strasbg.fr/simbad/
210 Astronomical taxonomy
method of classifying the brightness of a star was given a little more mathematical
rigour by Pogson in 1850, when he set a factor of exactly 100 as the difference in
brightness between a first and a sixth magnitude
√ star. So a one magnitude difference
5
in brightness corresponds to a factor of 100 or approximately 2.512. It should be
noted that the human eye has a logarithmic response to stimulation by light.
The last major catalogue that relied on the human eye to estimate the brightness
of stars was the Bonner Durchmusterung or BD. This contained all the stars from
the north celestial pole to a declination of −2◦ observed by Argelander with a
72 mm Fraunhofer telescope (King, 1955). The catalogue lists some 320 000 stars
with positions to 0s .1 in right ascension and 0 .1 in declination. Argelander visually
estimated magnitudes to ±0.1 to a limiting magnitude of 9.5 (van Biesbroeck,
1963). Subsequent photographically determined measures of Argelander’s stars in
the BD catalogue are remarkably similar to his eye estimates down to magnitude 9.0.
Visual estimates of stellar magnitudes are still made by a group of dedicated
amateur astronomers whose particular interest is stars that vary in brightness over a
long period of time (∼100 d) or that vary erratically or unpredictably. An example
of one of the techniques used is the fractional method (Sidgwick, 1955), where
two non-varying comparison stars near to the variable star are selected, one of
which is slightly brighter than the variable and the other slightly fainter (ideally the
magnitude difference between the comparison stars should not exceed 0.4). The
magnitude of the variable star obviously lies between the two comparison stars.
The brightness interval between the comparison stars is expressed as an integer
number of parts (normally less than 10) and an estimate made of the fraction of that
number that best represents the brightness of the variable. If the comparison stars
are of magnitudes m1 and m2 , where m2 > m1 , and the number of steps between
them is n, then if the variable is estimated to be k steps brighter than the fainter
comparison star, its magnitude mvar , will be:
k (m2 − m1 )
mvar = m2 − (12.1)
n
Boyden telescope in South Africa to complete the coverage of the entire sky. Clas-
sification was at first limited to magnitude 6. Since these photographs, plus others
taken with a 10-inch telescope in Peru for the survey, had a much fainter limit-
ing magnitude, the catalogue was able to be extended. Altogether, Miss Cannon
classified nearly 400 000 stars by eye over a period of 45 years. The results were
published in the Henry Draper catalogue plus its extensions. This gargantuan work
formed the basis of the modern scheme of classifying stellar spectra.
The HD (Henry Draper) classifications are essentially one dimensional, with
spectral class assigned by the relative strengths of the spectral lines of ions and
atoms and the bands of molecules. The letters O, B, A, F, G, K, M, L, T, (Y) form
the spectral sequence, with each class being subdivided by a number from 0 to 9,
e.g., A1 (spectral type of Sirius), G2 (spectral type of the Sun) and M2 (spectral
type of Betelgeuse). In physical terms, the spectral sequence from O to (Y) is one of
decreasing temperature. The defining spectral features (Keenan, 1963; Kirkpatrick,
2005) and effective temperatures (Cox, 2000; Burningham et al., 2008; Leggett
et al., 2009) of each class are given in Table 12.3. Class (Y) is not yet well defined
due to a lack of suitable candidate objects.
Luminosity classification
To convert the MK classification from a one- to a two-dimensional scheme, pairs of
spectral lines were selected for standard stars of known spectral type and absolute
luminosity. By comparing the line ratios exhibited by stars of unknown luminosity
class with those of the standard stars, an estimate of the luminosity class for the
candidate star may be made. Luminosity is a function of the surface area and hence
the radius of the star. If the large and small radius stars of the same spectral type have
212 Astronomical taxonomy
similar masses, then the larger star must have a lower gas density and pressure, and
surface gravity, than the smaller star. These differences show up in the appearance
of certain spectral lines; those of a very large star are narrow and sharply defined,
whilst those of a smaller radius star show the effects of pressure broadening and
appear wider and less clearly defined.
Table 12.4 lists the spectral line pairs that are of particular value in assigning
luminosity classes to stars of spectral type O to M. The fainter types, L, T and (Y),
are for the most part not truly stars and are currently termed L dwarfs, T dwarfs and
(Y) dwarfs.
There are eight main luminosity classes: class O (hypergiants), class I (super-
giants), class II (bright giants), class III (giants), class IV (subgiants), class V (main
sequence or dwarfs), class VI (subdwarfs), and class VII (white dwarfs). Some
examples of the two-dimensional spectral type / luminosity class system are Betel-
geuse (M2Iab), CMa (B2II), Arcturus (K2III), α Cru (B0.5IV), the Sun (G2V)
and α Cen A (G2V).
1. The open or galactic clusters, which typically are located in or near to the
galactic plane of the Milky Way Galaxy, with member stars that generally are
not centrally concentrated and have less than 1 000 members.
12.3 Classification of stellar objects 213
2. The globular clusters, which are strongly spherical in shape, have upwards of
1 000 members (exceeding 100 000 stars for very large examples) and are located
in a roughly spherical distribution about the centre of the Milky Way Galaxy.
Schemes for classifying clusters generally depend on the number of members,
location relative to the galactic plane, ages, colours and spectral types of the
members.
1. Degree of concentration
I: Detached clusters with a strong central concentration
II: Detached clusters with little central concentration
III: Detached clusters with no noticeable concentration
IV: Clusters not well detached but with strong field concentrations.
2. Range of brightness
1: Most of the cluster stars are approximately the same apparent brightness
2: A medium brightness range between the stars in the cluster
3: The cluster is composed of a mixture of bright and faint stars
3. Number of stars in the cluster
p: Poorly populated clusters with less than 50 member stars
m: Medium rich clusters with between 50 and 100 member stars
r: Rich clusters with over 100 member stars.
In addition, the letter ‘n’ may be added at the end of the classification to signify
the presence of any form of nebulosity.
In the first instance, the degree of concentration and the range of brightness may
be assessed by eye, as may the counting of those stars qualitatively presumed to
be cluster members. A recent example of a Trumpler classification for the southern
hemisphere open cluster IC2391 of II3r is given by Dodd (2004).
galaxies belonging to the different classes is set out in The Hubble Atlas of Galax-
ies by Sandage (1961) and online examples may be found on the Hubble Space
Telescope website.65
Sa Sb Sc
E0 E4 E7 S0
increasing
ellipticity >>> SB0
increasing
openness >>>
65 http://hubblesite.org/newscenter/archive/images/galaxy
12.4 Classification of Solar System objects 215
The Earth’s Moon is a satellite, as are: Phobos and Deimos of the planet Mars;
Io, Ganeymede, Europa and Callisto of the planet Jupiter; Iapetus, Mimas and Titan
of the planet Saturn; Umbriel, Titania and Oberon of the planet Uranus; Triton and
Nereid of the planet Neptune; and Charon of the dwarf planet Pluto.
There are a number of virtual observatories throughout the world. Their web
addresses are given above in Table 12.6, along with that of the International
Virtual Observatory Alliance (IVOA), whose primary task is to develop agreed
interoperability standards with the member VOs.
12.6.2 Recommendations
When celestial bodies are classified by measurement (e.g., positions, distances,
motions, flux outputs, temperatures etc.) and published in an online database or
virtual observatory, then such measurements should be given in SI units. If other
units are used, then they should be accompanied by SI units.
References
Abdo, A. A., Ackermann, M., Ajello, M. et al. (2009). Discovery of pulsations from the
pulsar J0205+6449 in SNR 3C 58 with the FERMI γ -ray space telescope ApJL, 699,
L102–L107.
Abdo, A. A., Ackermann, M., Ajello, M. et al. (2010). The first catalog of active galactic
nuclei detected by the FERMI Large Area Telescope. ApJ, 715, 429–457.
Abramowitz, M. & Stegun, I. A. eds (1972). Handbook of Mathematical Functions. New
York: Dover Publications Inc., p. 67.
Albacete-Colombo, J. F., Damiani, F., Micela, G., Sciortino, S. & Harnden, F. R., Jr. (2008).
An X-ray survey of low-mass stars in Trumpler 16 with CHANDRA. A&A, 490,
1055–1070.
Alexeev, I. I., Belenkaya, E. S., Slavin, J. A. et al. (2010). Mercury’s magnetospheric
magnetic field after the first two MESSENGER flybys. Icarus, 209, 23–39.
Alder, K. (2004). The Measure of All Things. London: Abacus.
Allen, C. W. (1951). Dex. Observatory, 71, 157.
Allen, R. H. (1899). Star Names and their Meanings. New York: G. E. Stechert & Co.
Aller, L. H. (1963). Astrophysics: The Atmospheres of the Sun and Stars, 2nd edn. New
York: The Roland Press Company.
Alonso-Medina, A., Colón, C., Montero, J. L. & Nation, L. (2009). Stark broadening of
PbIV spectral lines of astrophysical interest. MNRAS, 401, 1080–1090.
Amari, S., Hoppe, P., Zinner, E. & Lewis, R. S. (1992). Interstellar SiC with unusual isotopic
compositions: Grains from a supernova? ApJ, 394, L43–L46.
Anderson, B. J., Acuña, M. H., Korth, H. et al. (2010). The magnetic field of Mercury. Space
Sci. Rev., 152, 307–339.
Astronomical Almanac for 1995, The (1994). London: HMSO.
Atkin, K. (2007). Size matters. A&G, 48, 4.7.
Atwood, W. B. Abdo, A. A., Ackermann, M. et al. (2009). The Large Area Telescope on the
FERMI γ -ray space telescope mission. ApJ, 697, 1071–1102.
Bagnulo, S., Szeifert, T., Wade, G. A., Landstreet, J. D. & Mathys, G. (2002). Mea-
suring magnetic fields of early type stars with FORS1 at the VLT. A&A, 389,
191–201.
Barlow, C. W. C. & Bryan, G. H. (1956). Elementary Mathematical Astronomy. London:
University Tutorial Press Ltd., p. 125.
Battat, J. B. R., Murphy, T. W., Jr., Adelberger, E. G. et al. (2009). The Apache Point Lunar-
ranging Operation (APOLLO): Two years of millimeter-precision measurements of
the Earth–Moon range. PASP, 121, 29–40.
219
220 References
Harris, III, D. L., Strand, K. Aa. & Worley, C. E. (1963). Empirical data on stellar masses,
luminosities and radii. In Basic Astronomical Data, ed. K. Aa. Strand. Chicago IL:
University of Chicago Press, pp. 273–292.
Hawkins, J. M. & Allen, R. eds. (1991). The Oxford Encyclopedic English Dictionary.
Oxford: Clarendon Press.
Hearnshaw, J. B. (1986). The Analysis of Starlight. Cambridge: Cambridge University
Press.
Hearnshaw, J. B. (1996). The Measurement of Starlight: Two Centuries of Astronomical
Photometry. Cambridge: Cambridge University Press.
Henriksen, M. J. & Tittley, E. R. (2002). CHANDRA observations of the A3266 galaxy
cluster merger. ApJ, 577, 701–709.
Herbst, E. (2001). In Encyclopedia of Astronomy & Astrophysics, Vol. 2, ed. P. Murdin,
Bristol: IoP. pp. 1258–1266.
Hohle, M. M. Eisenbeiss, T., Mugrauer, M. et al. (2009). Photometric study of the OB
star clusters NGC1502 and NGC2169 and mass estimation of their members at the
University Observatory Jena. AN, 330, 511.
Holland, W. S., Greaves, J. S., Zuckerman, B. et al. (1998). Submillimeter images of dusty
debris around nearby stars. Nature, 392, 788–791.
Hollis, J. M., Chin, G. & Brown, R. L. (1985). An attempt to detect mass loss from α Lyrae
with the VLA. ApJ, 294, 646–648.
Hovestadt, D., Hilchenbach, M., Bürgi, A. et al. (1995). CELIAS: Charge, Element and
Isotope Analysis System for SOHO. Sol. Phys., 162, 441–481.
Huang, T.-Y., Han, C.-H., Yi, Z.-H., & Xu, B.-X. (1995). What is the astronomical unit of
length? A&A, 298, 629–633.
Hubble, E. P. (1926). Extragalactic nebulae. ApJ, 64, 321–369.
Hubble, E. P. (1929). A relation between distance and radial velocity among extra-galactic
nebulae. PNAS, 15, 168–173.
Irwin, J. A. (2007). Astrophysics: Decoding the Cosmos. Chichester: John Wiley & Sons
Ltd.
Józsa, G. I. G., Garrett, M. A., Oosterloo, T. A. et al. (2009). Revealing Hanny’s Voorwerp:
radio observations of IC2497. A&A, 500, L33–L36.
Karovska, M. & Sasselov, D. (2001). In Encyclopedia of Astronomy & Astrophysics, vol. 4,
ed. P. Murdin. Bristol: IoP, pp. 3068.
Kaye, G. W. C. & Laby, T. H. (1959). Tables of Physical and Chemical Constants and some
Mathematical Functions, 12th edn. London: Longmans, Green and Co.
Keenan, P. C. (1963). Classification of stellar spectra. In Basic Astronomical Data, ed. K. Aa.
Strand. Chicago: University of Chicago Press, pp. 78–122.
King, H. C. (1955). The History of the Telescope. London: Charles Griffin & Co. Ltd.
Kirkpatrick, J. D. (2005). New spectral types L and T. ARA&A., 43, 195–246.
Klioner, S. A. (2007). Relativistic scaling of astronomical quantities and the system of
astronomical units. A&A, 478, 951–958.
Korhonen, H., Hubrig, S., Berdyugina, S. V. et al. (2009). First measurement of the magnetic
field on FK Com and its relation to the contemporaneous star-spot locations. MNRAS,
395, 282–289.
Kurucz, R. L. (1979). Model atmospheres for G, F, A, B and O stars. ApJS, 40, 1–340.
Kutner, M. L. (2003). Astronomy: A Physical Perspective, 2nd edn. Cambridge: Cambridge
University Press.
Landstree, J. D. (2001). In Encyclopedia of Astronomy & Astrophysics, vol. 2, ed. P. Murdin.
Bristol: IoP, pp. 1508–1514.
Lang, K. R. (2006). Astrophysical Formulae. Vols I and II: Berlin: Springer-Verlag.
References 223
Leggett, S. K., Cushing, M. C., Saumon, D. et al. (2009). The physical properties of four
600 K T dwarfs. ApJ, 695, 1517–1526.
Leschiutta, S. (2001). In Encyclopedia of Astronomy & Astrophysics, vol. 4, ed. P. Murdin.
Bristol: IoP, pp. 3313–3315.
Liddell, H. G. & Scott, R. (1996). A Greek – English Lexicon. Oxford: Clarendon Press.
Lignières, F., Petit, P., Böhm, T. & Aurière, M. (2009). First evidence of a magnetic field
on Vega. A&A, 500, L41–L44.
Longair, M. S. (1989). Royal Observatory Edinburgh: Research and Facilities Handbook.
Edinburgh: ROE, p. 79.
Love, S. G. & Brownlee, D. E. (1993). A direct measurement of the terrestrial accretion rate
of cosmic dust. Science, 262, 550–553.
Lovell, B. & Clegg, J. A. (1952). Radio Astronomy. London: Chapman & Hall Ltd.
Lucas, P. W., Tinney, C. G., Burningham, B. et al. (2010). Discovery of a very cool brown
dwarf amonst the ten nearest stars to the Solar System. arXiv:1004.0317v1[astro-
ph.SR].
Lyne, A. G., Pritchard, R. S. & Graham-Smith, F. (1993). 23 years of Crab pulsar rotational
history. MNRAS, 265, 1003–1012.
McCarthy, D. D. & Petit, G. (2003). IERS Technical Note, 32. IERS Conventions (2003).
Frankfurt am Main: Verlag des Bundesamtes für Kartographie und Geodäsie.
Manchester, R. N., Hobbs, G. B., Teoh, A. & Hobbs, M. (2005). The Australia Telescope
National Facility pulsar catalogue. AJ, 129, 1993–2006,
Mather, J. C., Cheng, E. S., Cottingham, D. A. et al. (1994). Measurement of the cosmic
microwave background spectrum by the COBE FIRAS instrument. ApJ, 420, 439–444.
Mayes, V. (1994). Unit prefixes for use in astronomy. QJRAS, 35, 569–572.
Menzel, D. H. (1960). Fundamental Formulas of Physics. Vols 1 and 2. New York: Dover
Publications Inc.
Mills, I., Cvitaš, T., Homann, K., Kallay, N. & Kuchitsu, K. (1993). Quantities, Units and
Symbols in Physical Chemistry, 2nd edn. International Union of Pure and Applied
Chemistry. Oxford: Blackwell Science.
Mills, I. M., Mohr, P. J., Quinn, T. J., Taylor, B. N. & Williams, E. R. (2005). Redefinition
of the kilogram: a decision whose time has come. Metrologia, 42, 71–80.
Miura, T., Arakida, H., Kasai, M. & Kuramata, S. (2009). Secular increase of the astronom-
ical unit: a possible explanation in terms of the total angular-momentum conservation
law. PASJ, 61, 1247–1250.
Moffatt, J. (1950). A New Translation of the Bible. London: Hodder and Stoughton Ltd.
Mohr, P. J., Taylor, B. N. & Newell, D. B. (2007). The fundamental physical constants. Phys.
Today, 60(7), 52–55.
Monet, D. G., Levine, S. E., Canzian, B. et al. (2003). The USNO-B Catalog. AJ, 125,
984–993.
Moore, C. E. (1959). A Multiplet Table of Astrophysical Interest. Washington, DC: US
Department of Commerce.
Moore, P. A. (2001). In Encyclopedia of Astronomy & Astrophysics, vol. 1, ed. P. Murdin.
Bristol: IoP, p. 464.
Muhleman, D. O., Holdridge, D. B. & Block, N. (1962). The astronomical unit determined
by radar reflections from Venus. AJ, 67, 191–203.
Murray, C. A. (1989). The transformation of coordinates between the systems of B1950.0
and J2000.0, and the principal galactic axes referred to J2000.0. A&A, 218, 325–329.
Noerdlinger, P. D. (2008). Solar mass loss, the astronomical unit, and the scale of the Solar
System. 2008arXiv0801.3807N, 1–31.
Pannekoek, A. (1961). A History of Astronomy. London: George Allen & Unwin Ltd.
224 References
2MASS (Two Micron All Sky Survey) photometry, barycentric coordinate time (TCB), 56
132 barycentric dynamic time (TDB), 56
baryonic matter, 207
absolute elliptical orbit, 102 Bayer, 208
absolute magnitude, 140 bequerel, 66
absolute zero, 148 binary star, 102
acceleration, 88 BIPM (Bureav International des Poids et Mesures), 3,
accuracy definition, 5 10
accurate definition, 5 blackbody radiation, 151
Airy transit circle, 37 Boltzmann constant, 146, 152
altazimuth coordinate system, 38 Bonner Durchmusterung, 210
altitude, 38 bound-bound transitions, 158
ampere, 174 bound-free transition, 158
angstrom unit, 27 Boyle’s law, 201
angular acceleration, 60 brief history of units, 7
angular momentum, 111 brightness, 133
angular quantum number, 159 brightness temperature, 154
angular velocity, 60
ANS photometry, 132 candela, 113
antenna temperature, 154 Cannon, 210
apparent magnitude, 140 Cavendish, 94
area, 89 celestial equator, 32, 39
Arecibo radio telescope, 192 celestial latitude, 40
Argelander, 210 celestial longitude, 40
asteroseismology, 97, 104 celestial sphere, 36
astrochemistry, 202 CELIAS, 200
astronomical classification, 206 celsius to kelvin temperature, 148
astronomical taxonomy, 206 centre of a sphere, 36
astronomical unit, 22, 75 cerenkov radiation, 136
atomic mass unit, 197 cesium fountain atomic clocks, 68
atomic time, 55 CGPM (Conférence Général des Poids et Mesures),
attribute, 206 3, 9
average atomic mass, 198 cgs Gaussian units, 175
Avogadro’s constant, 197 cgs system, 14
Avogadro’s law, 201 CHANDRA, 52
axis of a great circle, 36 characteristic age, 68
azimuth, 38 Charles’ law, 201
colour index, 140
BAAS (British Association for the Advancement of colour temperature, 161
Science), 10 column mass density, 107
barycentre, 102 Condamine, 9
226
Index 227