Star formation and abundance gradients in the galaxy

Star formation and abundance gradients in the galaxy

Vistas in Astronomy, Vol.26.pp.159-224,1982 Printed in Great Britain, All rights reserved, 0083-6656/82 $O.00 + .50 Copyright © 1983 Pergamon Press...

4MB Sizes 0 Downloads 34 Views

Vistas in Astronomy, Vol.26.pp.159-224,1982 Printed

in Great Britain,

All rights reserved,

0083-6656/82 $O.00 + .50 Copyright © 1983 Pergamon Press Ltd.

STAR FORMATION A N D A B U N D A N C E GRADIENTS IN THE GALAXY R. Giisten and P. G. Mezger Max-Planck-Institut ffir Radioastronomie, Bonn, F.R.G.

SUMMARY

Sect. I: Chemical Evolution of the Galaxy. The closed system model and its limitations. For a closed system, where the sum of gas and matter contained in stars is constant at any time~ one derives a simple set of equations which relate the enrichment (or depletion) of the interstellar matter (ISM) to the gas-to-total mass ratio, if a power-law dependence of the star formation rate (SFR) on the total mass of gas, dM~/dt = ~ is adopted. We apply the predictions of the closed system for k=l to the galactic disk. Reasonable agreement is obtained for the chemical evolution of the ISM in the solar vicinity, but the model completely fails to explain both the observed galactic abundance gradients, and the observed time variation of the chemical evolution. Sect. II: Mass Distribution in the Galactic Disk. We discuss the distribution of the mass of the ISM and stars in the galactic plane~ and derive the gas-to-total mass ratio as a function of galactocentric distance R. Special emphasis is given to the distribution of moleccular gas. We discuss the problems involved in converting observed emissivities of the J=l-O line of CO into H 2 column densities, and rederive average molecular densities, considering an increase of the C/H abundance towards the galactic center. Sect. III: Star Formation Rates in the Galactic Disk. The Lyman continuum (Lyc) photon production rate of all 0 stars in the galactic disk can be estimated from the free-free emission produced by compact and extended low-density (ELD) HII regions. SFRs can be inferred from Lyc photon production rates if the initial mass function (IMF) of a newly formed generation of stars is known. After a review of earlier work, we derive Lye photon production rates, assuming radial symmetry in the galactic plane only. It is found that the ridge line flux density S(Z,b=O °) can be reproduced correctly only if the spiral structure of the galactic disk is taken into account. For a quantitative estimate of SFRs we discuss the characteristics of IMFs derived from star counts in the solar vicinity. With a constant IMF and upper and lower mass limits of 60 and 0.1 mQ we derive a total present-day galactic SFR of ~(t o) ~13 m G yr-1 and find that more than one half of all O stars must be formed in spiral arms. We compare this value of the present-day SFR with the time-averaged SFR ~]O m~ yr -I and with the pres~ntrday SFR predicted by the gas-to-total mass ratio of the.~alactic dis~, ~=0.04, which is ~ t o ) ~! 3 m^ yr -I There is a gross inconsistency between ~t~o) and ~(to). We discuss pre . • -~ • . . • oos . pre . the relatlon of the diffuse galactlc far and near infrared emlsslon wlth star formatzon. It is shown that the observed NIR emission can be reproduced only if one superimposes on the old disk population a new population of M giants, whose distribution is similar to that of 0 stars. Sect. IV: Abundances and Abundance Gradients in the Galactic Disk. In Table IV.! we have compiled abundances of the most abundant elements and their isotopes as derived for the solar system and the ISM in the solar vicinity. Also given in this Table is the abundance variation of these elements with galactocentric distance R, the so-called abundance gradient. Some problems related to the observational determination of abundance gradients are discussed in the text. Sect. V: A Self-Consistent Model of Star Formation and Chemiaal Evolution of the C~laxy. We i ~ o d u c e as modifications of the simple closed system model continual infall of halo gas and bi-modal star formation, where the low-mass end of the IMF in spiral arms is truncated at masses m c 22 m G (i.e. only medium and high mass stars form) while the IMF in the interarm region has a low-mass cut-off at O.1 mQ. Otherwise the shape of both IMFs is the same. We show that the truncation of the low-mass end of the IMF increases the bulk yield of element production. We derive the basic equations of chemical evolution in a rigorous form and compare 159

160

Re Gbsten and P.G. Mezger

them with the instantaneous recycling approximation. We introduce a new IMF adapted for himodal star formation and use it to compute the effective bulk yield as a function of galactoeentric distance R, low-mass cut-off, mG, of the spiral arm IMF, and the ratio ~ of SFRs in spiral arm and interarm region, respectively, at R=IO kpc. We show that a yield which increases with decreasing R and a considerably reduced net SFR (for a given Lye photon production rate) are natural consequences of hi-modal star formation. We compare observations, related to the chemical evolution of the galactic disk, with the predictions of our model of hi-modal star formation, which includes infall of halo gas. For the solar vicinity, we find good agreement between the observed and predicted cumulative metal abundance distribution of G and M dwarfs, and between observed and predicted a~e metallicity relation8 of the ISM (where the observations relate to F dwarfs (Fig. V.3)).For the galactic disk we can reproduce the observed abundance variations (abundance gradient8) of 160, ~He as well as of the isotope ratios 12C/13C and lS0/170 (Fig. V.4). For the depletion of deuterium (2H) in the solar vicinity we find values of XTsM/X. ~0.05 and XQ/Xp ~0.]7, respectively, which yield primordial deuterium abundances of (s~/TH)_~(I-2).]0 -~ ( b y number). These values are considerably higher than earlier estimates. E One effect of bi-modal star formation is that the Lyc photon production rate per unit mass of ISM converted into stars is increased. The reason is that relatively more massive stars are formed. SFRs, derived from observed Lye photon production rates, are therefore expected to be lower than SFRs derived on the assumption of a constant IMF. The discrepancy between (on the basis of the mass distribution in the galactic disk) predicted SFRs and (for a constant IMF) observed SFRs disappears, if observed SFRs are derived for hi-modal star formation (Fig. V.5). For the total galactic disk the SFR is ~6 me yr-I on the assumption of bi-modal star formation, as compared to ]3 t~m~eyr -~ obtained on the assumption of a constant IMF. Even more dramatic is net the difference in SFRs, i.e. the mass of gas converted into stars which stays permanently locked up. For a constant IMF this mass is ~9 mQ yr-I, while it is only ~].5 m~ yr-I in the case of bi-modal star formation. The distribution of M-giants which is required to explain the diffuse galactic NIR emission is also explained and is a natural consequence of bi-modal star formation. Appendix A: Production Rate of Lyman Continuum Photon8 from Early-Type Stars. Characteristics of O stars during their main sequence (MS) lifetime are compiled and these characteristics are used together with the initial mass function (IMF) to relate the Lye photon production rate of an ensemble of O stars with the formation rate of stars of all masses.

Appendix B: Deconvolution of the Diffuse Galactic Free-free Emission assuming Azimuthal Symmetry. We give an analytical solution of the problem applicable to high-resolution observations of the ridge line intensity; and we describe an approximate numerical solution applicable to observations with low angular resolution, which also considers the large uncertainties introduced by the separation of thermal and non-thermal diffuse emission.

Appendix C: Sites of Nucleosynthesis. We derive IMF weighted production efficiencies for some of the most abundant primary and secondary elements (Fig. C.I). These quantities are used to compute the net bulk yields required for model computations of the chemical evolution of the Galaxy. These net bulk yields together with some nuclear synthesis processes and their characteristics are compiled in Table C.I.

Star Formation and Abundance Gradients

I.

CHEMICAL EVOLUTION OF THE GALAXY:

At an early epoch mass) hydrogen

161

THE CLOSED SYSTEM MODEL AND ITS LIMITATIONS

galaxies condensed out of a primordial

gas which consisted of -78% (by

(IH) and -22% helium (~He) and traces of Deuterium

chemical composition

of the primordial

gas is quantitatively

(=H), SHe and Lithium.

The

explained by nuclear reactions

taking place about three minutes after the expansion of the universe has started, provided the baryon density of the universe was -10% of the critical mass density which would close the universe.

Today,

in the interstellar matter

tion of hydrogen has decreased

(ISM) of the solar vicinity,

the mass frac-

to 70%, while the fraction of tHe has increased

to 28% and in

addition a mass fraction of ~2% consists of elements heavier than ~He. This process of the enrichment of the ISM with additional evolution of the Galaxy".

Burbidge,

~He and heavier elements

Burbidge,

Fowler and Hoyle

showed that this chemical evolution can be explained especially

the evolution of the more massive

and in their final evolutionary

(1957),

to as "chemical

in a classical paper,

as a byproduct of nuclear reactions

stars, which form continually

and

out of the ISM

stages, return a certain fraction of their total mass, en-

riched with ~He and heavier elements,

I.I

is referred

back to the ISM.

CHEMICAL EVOLUTION OF A CLOSED SYSTEM

The simplest model relates to the chemical evolution of a closed system, moment the fraction of matter in the form of interstellar

gas and dust, Mg, and the fraction

of matter locked up in the form of stars, M , is constant,

Mg

+ M

]J

+ S

ffi

~

in which at any

i.e.

Mto t

or

(I. 1) ffi 1

with ~ ffiMg/Mto t and s = M~/Mto t. With the assumption of an "instantaneous stellar matter processed

in stars it is found that the chemical

fraction ~ only. One can obtain its time dependence

return" of inter-

evolution depends on the gas

if one assumes a functional

the star formation rate ~ on the amount of gas available

to be transformed

dependence

of

into stars, such

as

dM~-ffi (1-r).v ffiTI M kg -a~

(I.2)

Here, r ~I/3 is the fraction of matter which is returned by a generation of stars with masses between O. 1 and IOO M 8. With Eq.

(I.I) this yields exp(-t/T}

D(t)

=

l

k = 1 (I.3) 1/(l-k)

~[l+(k-l)t/T]

k # l

where T is the time scale in which roughly two thirds locked up in stars. With increasing monotonically

(k=l) of the gas has been permanently

time more and more gas is locked up and hence ~ decreases

with time.

For primary elements,

i, such as 12C,

160 etc. which are synthesized

~He in the stars which eject them, the enrichment

from primordial

of the ISM is found to be

IH and

162

R. G~sten and P.G. Mezger

zp(t ) = (|_r) Pi

in(~ -I) = yPln(~ -I)

(I.4)

with Z~z the mass fraction of this element in the ISM. Here Pi is the fraction of mass returned by a generation of evolving stars in the form of newly synthesized element i and Yi = Pi/(!-r) is referred to as the dimensionless net bulk yield (a derivation of this and subsequent results is given in Sect. V). For secondary elements, j, such as 13C, I~N, IVO etc., which require for their synthesis the presence of certain primary "seed nuclei" the corresponding expression is

!

zS(t)j = ~

]

ySyP[In(~-1)]= = ~

s

y-~-Y(zP) 2

(1.5)

with the dimensionless yields y~, yjs defined as above and Z~z the abundance of the primary seed nuclei. For light elements, such as mH, which cannot be synthesized but only be destroyed during astration, the depletion of their primordial abundance X

is P

X(t) = Xp r / ( ] - r )

(1.6)

The significance of this simple model for the chemical evolution of a closed system with continual star formation lies in the fact that its results are often rough asymptotic approximations to more sophisticated models (see Sect. V). But in particular this simple approach allows us to discuss the properties of the chemical evolution models in analytical terms, and thus make the dependence on various model parameters more transparent. In the remaining part of this section we compare the predictions of the closed system with observations related to our Galaxy to see where these predictions fail and hence where the basic assumptions of this model of chemical evolution have to be changed.

1.2

APPLICATION TO SOLAR SYSTEM AND sOLAR VICINITY

Observations and model computations lead to the following picture for the evolution of a spiral galaxy (see Tinsley 1980 for recent review). At least one half (but probably more) of the mass of primordial gas of the (nearly spherical) protogalaxy condensed into stars before the remaining gas collapsed into a flat galactic disk. The gas out of which the galactic disk was formed was already somewhat enriched in ~He and heavier elements (rough estimates yield about Z h _<0.2 Z®) through the evolution of the first generations of halo stars. Once the disk has formed, stars are formed continually out of the gas, primarily in the main spiral arms of the disk. Gas and the disk population of stars which formed out of this gas, move approximately on circular orbits around the galactic center. Therefore, not only the disk as a whole, but also stars and gas in concentric rings can be considered as closed systems. Here, and in the remaining part of this paper we restrict our considerations to the galactic disk between galactocentric radii D R = 3 and ]3 kpc,respectively. The chemical evolution of the central region of the Galaxy appears to be much more complex than that of the disk. Outside |3 kpc the star formation rate is so low that the element abundance appears to be strongly influenced by the chemical evolution of the gas in the halo before it collapsed into the disk, so that the chemical evolution in this region will not be considered in this paper. The mass of the disk and its radial distribution is inferred from the galactic rotation curve.

Star Formation and Abundance Gradients

163

The distribution of gas in the disk is derived from observations of the H%2] cm line (atomic hydrogen) and the rotational transitions of the CO molecule (molecular hydrogen). For the solar vicinity one finds from Table 11.2 for R = IO kpc interstellar helium abundance is Y(tHe) = Y mated primordial abundance Y

+ AY

~Io = Mg/Mto t -0.04. The local

= 0.28. From this follows, with an esti-

p ev ~0.22, a stellar contribution,

AY = zP(tHe) -0.06. For the p ev most abundant isotope of oxygen (160) the observed abundance is zP(160) -O.Ol. From model computations of stellar evolution (see Appendix C) yields Yi are derived for these two elements, and with Eq. (I.4) the predicted abundances given in Table I.| (~Io = 0.04) are obtained. They are higher than the observed abundances by a factor 1.5 for the and 52 for 160. Table l.l

Predicted and observed enrichment of the ISM of the solar vicinity for the primary elements tHe and 160 Element i tlle le 0

P Yield Yi 2.8

Z~(predicted)

lO-2

0.090

(6-8) 10-3

O.019-0.O26

Z~(observed) 0.06 O.Ol

With Eq. 1.3 the time dependence of the enrichment of the ISM can be predicted. For this, the exponent "k" has to be specified, which relates the SFR with the amount of gas available. From the small scale height of the young stellar population, k ~2 is suggested, but as stars probably form out of molecular clouds, which themselves are confined to a thin layer in the galactic plane, k=l may be feasible, too (see our discussion in Sect. II1). For the sake of simplicity we adopt here k=l. Substitution of (I.3) into the relations (I.4) to (1.6) then yields P Z~(t) = Yi(t/T)

(1.7)

i.e. a linear increase with time of the enrichment of the ISM with primary elements, and

Zj(t)/zP(t) = ~I y~(tlT)

(1.8)

i.e. a linear increase with time of the abundance of a secondary element j relative to a primary element i. Finally, for primordial elements which are destroyed while ISM circulates through stars, one obtains

X(t)/Xp

=

[e-t/T] r/(|-r)

(I.9)

i.e. an exponential depletion with time. One way to check these predictions for the time dependence of primary, secondary and depleted elements is a comparison of the abundances of the solar system (which formed out of the ISM t@ "5 109 yr ago) with the ISM abundances in the solar vicinity, adopting an age for the galactic disk of roughly to -2.5 t@. From Eq. (1.3) we find with ~Io = 0.04 a ratio t /T -3.2. Substitution of these values in the above relations o yields ZIsM/Z P P@ ~1.7, (ZIsM/ZIsM)/(Z®/Z s p s ®) p -1.7 and X(~H)IsM/X@(2H) ~0.55. In Table 1.2 are compiled observationally determined isotopic abundances of the solar system and the ISM in the solar vicinity (from Table IV.I). For primary elements the predicted ISM abundances should be about twice the solar system abundances. The

He abundances estimated for the solar

system, Y®(tHe) ~O.17-O.28, suggest in fact a somewhat lower tHe abundance in the solar system, but the uncertainties are too large to warrant a meaningful comparison.

In the case of the

164

Ro Gusten and P.G. Mezger

other most abundant primary elements, whose abundances are compiled in Table IV.l, it appears as if the solar system abundances are higher than those of the ISM. This effect can be qualitatively explained by depletion of elements other than tHe by forming cores and ice mantles of interstellar dust grains, but it is quantitatively hard to assess. This failure to detect systematic differences between Z@ and ZIS M had cast severe doubts on the whole concept of a continual chemical evolution.

Table 1.2

Comparison of isotopic abundances in the solar system and in the ISM in the solar vicinity

Isotopes

Isotopic abundances solar system ISM

IH/= H 12C/13 C

3.6 IO-s

leo/IV O

89 5.4

5.|O -6 75 3.7

(ISM)/(solar system) observed predicted O. 14

0.55

0.78

0.6

0.67

0.6

First indications of systematic differences in the chemical composition of the solar system and the present-day ISM come from observations of isotopic abundances, which can be determined with high accuracy for the solar system. Through both UV spectroscopy and radiospectroscopy, isotopic abundances of the ISM have been determined for three isotopes with sufficient accuracy to warrant a meaningful comparison with the corresponding solar system values, compiled in Table IV.l. In the case of carbon and oxygen the more abundant isotope is a primary element and the less abundant isotope a (mainly) secondary element, so that for their ratio as a function of time Eq. (I.]) applies. In the case of deuterium we deal with a light element which has been formed in the initial phase of the universe, and subsequently is only destroyed when passing through stars. For this case Eq. (1.9) applies. A comparison between predicted and observed isotopic abundances in Table 1.2 shows at least a qualitative agreement, which lends support to the concept of a continual chemical evolution.

1.3

APPLICATION TO THE GALACTIC DISK

In the preceding subsection we have compared abundances in the solar system and the ISM of the solar vicinity, and found reasonable agreement with the predictions of the model of closed system chemical evolution, if the various uncertainties of observational abundance determinations and model parameters are taken into account. There is a clear incompatibility, however, if the predictions of the closed system evolution are compared with observations of the present-day physical state of the galactic disk and to its evolution with time. As an example let us consider the variation of primary elements as a function of galactocentric distance, R. Observations related to abundance gradients are discussed in Sect. IV. There we found a strong gradient for the primary element 160. In Table 11,2 we have compiled values of the gas-to-total mass ratio ~(R); this quantity increases slowly from .03 at R=4.5 kpc to .05 at R=II.5 kpc. Substituting these values into Eq. (1.4) for the closed system, we find that the constant yield model predicts an abundance increase of primary elemants

zP(4.5)/zP(]I.5) of

1.2. Observations on the other hand yield a ratio for the COO

abundance of Z(4.5 kpc)/Z(ll.5 kpc) ~2.5 (see also Fig. V.4). As shown in Sect. III, the present-day star formation rate (SFR), ~(to), can be determined observationally by connecting the Lyman continuum photon production rate of O stars with the

Star Formation and Abundance Gradients

165

total SFR through the initial mass function (IMF). This SFR, as a function of R, is shown in Fig. lll.5. By combining Eqs. (1.2 and 3) the closed system model yields a relation between average SFR, <~(t)> and Present-day SFR,

~(to)

~.In ~-I

<~(to)>

(1--~)

(1.1o)

l

With <~> = t .(]-'-------~)" M~ and ~0.04 this yields a predicted galactic star formation rate of o ~pr~to) ~|.3 m e yr -I. This rate is a factor of ten below the presently observed rate -13 m8 yr -I =<~>. The same discrepancy is found between observed and (from Eq. I. ]O) predicted radial distribution of the star formation rates in Fig. 111.5. Another inconsistency is found if the chemical evolution of the galactic disk, as derived from the iron abundances in stars of different age is compared with the predictions of the closed system model. While the observed enrichment (Fig. V.3) increases roughly exponentially with time during the first two thirds of the age of the galactic disk, it nearly saturates during the last third (i.e. during roughly the time the solar system exists). The simple closed system model on the other hand predicts for k=! a linear increase with time of the abundances of primary elements. There are refinements of this simple model. The most obvious deviation from a closed system model of the galactic disk is a gradual build-up of material in the disk by a continuous infall of halo gas. This idea is corroborated by both observations of infalling atomic

infall model of chemical evolution, which can alleviate some of the problems found when comparing hydrogen and dynamical collapse computations (Sect. V.|). Such a scenario leads to the

observations with the closed system model. However, infall models alone can explain neither the strong galactic abundance gradients nor the very high present-day SFR, derived on the assumption of a constant IMF. In this paper we develop a new model of star formation and chemical evolution. The main characteristic of our model is that star formation occurs in both spiral arms and the interarm region, but that in spiral arms primarily medium and high mass stars are formed, with low-mass star formation proceeding nearly exclusively in the interarm region. As we will show, the model of bi-modal star formation yields in fact reasonable agreement with most observations related to chemical evolution and star formation.

II.

THE PHYSICAL STATE OF THE GALACTIC DISK

The goal of this section is the determination of the quantity ~(R) = Mg/Mtot, defined in the preceding section. This requires a determination of the distribution in the disk of both the gas, Mg(R,z), and the total mass, Mtot(R,z) = Mg + M~, as a function of the galactocentrlc radius R and the coordinate z perpendicular to the galactic plane. M.(R,z) is the distribution of the stars, which account for most of the mass in the galactic disk.

II.].|

DISTRIBUTION OF THE INTERSTELLAR GAS

The interstellar matter (ISM) is a mixture of dust particles ranging in size from O.01 to 0.25 um, which appear to consist primarily of graphite and silicates. In dense clouds these

166

R. Gusten and P.G. Mezger

cores may have mantles of molecular Md/Mg ~.005-0.01.

The "standard"

ices. Estimates

description

of the dust-to-gas mass ratio range from

of the principal

gas phases is summarized in

Table ll.l. Most of the ISM is contained in cold clouds of high density, which are embedded

TABLE I1.1 State

CHARA[TERISTIf'SOF 5AS PHASES IN THE fiALAfTI[ DISK Gas Phase

Number Density (cm-3)

Kinetic Temp. (K)

Cloud Mass (m8)

103

10

lO s

102-103

IO

~103

Fraction Scale

Volume Filling Factor

of total Gas Mass

Height (pc)

~50%

60

~25%

I00

~25%

140

~%

120

Giant Molecular

Molecular Clouds

Molecular

Dark Clouds

Atomic

HI Clouds

Atomic

Intercloud

HI Gas Ionized

HII Regions "Coronal"

Ionized

in a less dense intercloud phase. jects with hydrogen dominantly

clumps

50-100

0.1-1

103-10 ~

1-10 5

I0 ~

IO-~-IO-a

Gas

with a significant

30

<_2%

I0 =

50%?

-

10-1-10 ~

~%

J

20-80%

10~-I0 s

The cloud parameters

>I000

range from medium-dense

in its atomic phase to giant molecular

fraction of their estimated

which is either warm, partially

low-mass ob-

(n(H2) >>n(Hl))

total mass of ~105 M® confined

[n(H~) =104-5 cm-3]. Most of the interstellar

"stratum",

0.5

clouds

to high-density

space is filled by the intercloud

ionized atomic hydrogen or hot coronal gas. Both

the volume filling factor and the mass of ionized gas in HII regions around OB-stars

is small.

Gas and dust form a flat layer around the galactic plane. As shown in Figure II.I, taken from a review paper by Downes and GHsten

(1982), the scale height of the ISM is roughly constant

inside the solar circle, but increases The distribution

toward the edge of the Galaxy.

of atomic hydrogen is determined

ture line. For an optically

is T L = TKT =N(H),

N(H). In practice,

veys in terms of the radial distribution

the interpretation

correction

(Table 11.2), as obtained by Burton and Gordon

(1981) and Liszt et al.

for this optical depth effect

The average volume density n(Hl)

seems roughly constant with galactic radius. However,

tion studies Dickey

(]978) and Baker and Burton

(1982) have pointed out the existence of cold for in earlier studies.

in the inner disk, where most of the GMCs are located,

be somewhat underestimated.

The determination

(H2) , is much more complicated

of the mass distribution

the HI density may of molecular hydrogen,

and subject to large systematic errors. As the H 2 molecule has

no transition at radio wavelengths, in the first rotational

in the

based on recent higher resolu-

atomic hydrogen associated with GMCs, that has not been corrected Thus, especially

of 21cm line sur-

of atomic hydrogen is complicated by self-absorption

the HI density in the inner disk can be inferred.

(1975),

thus yielding directly the column

however,

along the line of sight, and only after appropriate

galactic plane

struc-

thin gas of kinetic gas temperature T K the optical depth is

T =N(H)T~ I and the line temperature density of atomic hydrogen,

from surveys of the H%2]cm hyperfine

the basic observations

transition of the CO molecule

are surveys of the galactic plane

at %2.6 mm. The CO molecule

is the most

Star Formation and Abundance Gradients

0 I

J

I

I

167

I

INTERCLOUDHI

- 700

INCLINED

z

500t~ K

N

100

5oOoa~

0 r'n

100

HOLECLUOLAURDS '

I

I

O

Z

I

2

I

I

I

I

6 10 GALACTIC RADIUS R IN kpc

U.J -r-

200

I

I

12

Fi~. II.l: Sketch of the vertical structure in the galactic disk. Solid ~ines gzve the scale height of the HI-gas. Shaded boxes indicate the molecular gas component: height = scale height, width proportional to mean volume density (from Downes and G~sten, 1982). abundant molecule in interstellar space and, due to its small dipole moment, is easily excited by collisions with H a molecules at rather low gas densities(n(Ha) ~500 cm-3). While one can expect the CO line temperature to be a qualitative measure of the H a column density, the relation CO column densities to H a column densities requires a number of assumptions, which will be discussed next. The first rotational line emitted by the q2CIeO molecule, which contains the most abundant isotopes of both elements, is optically thick in nearly every molecular cloud and, since for I ~40 °, b ~0 ° the line of sight on the average intersects at least one GMC, the same statement holds for surveys of the Galaxy at these longitudes. With an isotopic abundance 12C/13C ~70 the q3CIeO molecule should be less abundant in clouds by about the same factor, and hence its first rotational transition should be more optically thin. Due to the weakness of this emission only recently systematic surveys of the galactic plane in q3CIeO lines have been made. However, in those cases

where both lines have been observed it was found, that the line pro-

files, i.e. the line temperature T L as a function of the radial velocity v, of both transitions, were rather similar and that the ratio of the integrated line profiles converges towards a value 0 = fTL(laCO)dv/fTe(13CO)dv ~(5-6)

(II.1)

The interpretation of this fact is that the gas in a GMC is not distributed homogeneously but tends to form clumps, which contain most of the mass of molecular hydrogen, but which fill only a small fraction of the cloud's volume. Although these clumps are still opaque for the 12C0 line, the line profile is determined by the velocity distribution of clumps within the beam, especially

since turbulent or systematic motions within the cloud appear to provide

enough Doppler shift

that shadowing of the clumps is small. Comparison of 13C0 emissivities

168

R. Gusten and P.G. Mezger

inferred from 12C0 surveys with direct recently made 13C0 surveys, which still are limited in their (l,b) coverage, indicates that both methods yield comparable 13C0 emissivities (Solomon et al. 1979, Liszt et al. 198l). To derive H a columa densities N(H2) , one then proceeds along the following lines. Observations yield relations between the visual extinction A N(13C0)

v

and the corresponding column densities

(Dickman, 1978)

A

v

= 4(±2) . 10-16 N(13C0)

(11.2)

and N(H 2) (Savage at al. 1977; Bohlin et al. 1978), respectively,

N(H) = N(HI) + 2 N(Hz) = 2.|0 =I . A v

(11.3)

N(H=)= 4(±2.5)

(II.4)

from which follows

10 5 N(ImC0)

For 10 ~Tex(C0)/K ~20, the typical range of observed excitation temperatures, is

N(13C0) = 1.3(±.2) 101mfTL(13C0)dv

(11.5)

and, with (ll.l) and (11.4), follows

N(H=) = 5'2(+~'s~lO2°fTL(13C0)dv~-3.s, and

(11.6)

NCHz) = l.O(~i~)102°fTL(1~CO)dv

In addition to the uncertainties in the above relations one must realize that relation (11.2) is obtained from 13C0 observations and star counts in local dark clouds, which are limited to the range I ~A ~4 mag of visual extinction, and that relation (11.3) is obtained from UV v observations of diffuse clouds (Av E2 mag). Extrapolation of these relation to GMCs, which comprise most of the molecular hydrogen in the galactic disk, is rather uncertain, although for the case of the high extinction clouds (Av ~10 mag) B 5 (Dickman 1978) and 0-Oph (Frerking et al. 1982) relation (11.2) seems still to hold. However, in the case of the Taurus cloud (Frerking et al. |982) the observed relation between A

v

and N(13C0) is found to

be too low by a factor of two. An extrapolation of relation (11.6) to clouds with A

~]0 mag, typical for GMCs, may underv estimate H2 column densities for yet other reasons: i) Condensations in GMCs have typical column densities of 1022 H= molecules cm -2, for which even the 13C0 lines become optically thick, ii) Gas in GMCs is not smoothly distributed but tends to form high-denslty clumps,

where the 13C0 abundance may be reduced by up to factors of =|0 (Wooten et al. 1978). iii) Contrary to regions with A

E4 mag, where 13C0 may be enhanced by chemical fractionation v (Langer |977; Langer et al. 1980), this should not be the case for the more shielded parts of

GMCs. It is difficult to estimate how and to which extent all these effects may affect relation (11.6). However, at least for the outer parts of GMCs with A ~4 mag, which may comprise

Star Formation and Abundance Gradients

169

most of its mass, we feel that relation (11.6) is applicable. Nevertheless, we use alternatively in the following discussion a ratio [13C0]/[H2] = 10-s (e.g. Solomon et al. 1979), which is consistent with an interstellar isotope ratio 12C/13C = 70 and a fraction of 10% of all carbon being tied up in CO. More recently attempts have been made to apply a model of radiative transfer to the observed I~C0 and 13C0, J=1-O transitions using the molecular abundance as a free parameter (Solomon and Sanders 1980; Liszt et al. 198]). However, the result of this modelling depends critically on the adopted volume density and hence, for a given column density of H~, on the clumping inside a GMC. Although this method appears to be promising, observations of the J=2-] transition should be incorporated in the model fit to derive more reliable data. H 2 densities given in Table 11.2, column (2) are computed from C0-emissivities as derived by Burton and Gordon (1978) and given in Table 11.2, using standard relation (Eq. 11.6). Contrary to the volume density of atomic hydrogen which is nearly independent of galactocentric distance R, the H 2 volume density shows a prominent peak in the inner part of the Galaxy between 4 _
[H2]/[13C0]

= 4

IO s and

106, respectively)

=

=0.8 - 1.25 cm -3 ~0.16 - 0.40 cm -s

have to be co~pared with results from the Copernicus UV-satellite (Savage et al. 1977; Bohlin et al. 1978). There the column density of HI is derived from Lye-absorption measurements against nearby stars, and that of H 2 from absorption lines in the Lyman-band of the two lowest rotational states in the ground vibrational level of the H 2 molecule. Volume densities, averaged along the llne of sight are:

=1.15

cm- a



cm- a

~0.14

Although the agreement is rather good, one should keep in mind that there are large uncertainties in both figures. The densities given in Table 11.2 are averaged over the azimuthal angle @=0-2~, i.e. they relate to both spiral arms and interarm regions, while the UV results

170

R. G~sten and P.G. Mezger

for both HI and H 2 are derived locally within heliocentric distances of only 500 and IOOO pc, and therefore, on the one hand, are biased towards physical conditions in the interarm region. On the other hand, due to the increasing interstellar extinction in the UV, these absorption measurements are also strongly biased against the more massive (i.e. molecular) clouds. The authors have attempted to correct at least for the latter bias and suggested n(H~) ~ 0 . 1 4

cm-3 as a lower limit. In fact molecular gas contained in nearby cloud complexes

like the extended Taurus region (D s ~150 pc, M ~lO~-lO 5 me, Lequeux 1981) is not considered in these measurements.

11.1.2

DISTRIBUTION OF STELLAR MASSES

The ISM accounts only for E5% of the total mass of the galactic disk and it is therefore customary to associate the surface density of the total mass of the disk, ~D ~ with the stellar mass. This quantity is derived from the rotation curve of the Galaxy, assuming a mass distribution perpendicular to the galactic plane. Values OD(R), given in Table 11.2, are taken from the mass model of Innanen (1973), however

selecting only those stellar components

which belong to the disk population. Ostriker and Caldwell

(1978) and Rohlfs and Kreitschmann

(1981) derived models, where the surface density rises somewhat faster toward the galactic center (by 10-20%). But within these uncertainties,

the mass distribution of the galactic

disk appears to be reasonably well determined by the rotation curve. It should be noted that mass estimates of the galactic halo are highly uncertain.

TABLE 11.2 ~diu.

GAS AND STELLARDENSITIESIN THE GALACTICDISK ~Z

~(IaCO)

~2

K.km s--~kpc-1

cm-3

.19

12.4

.41

(tpc)

cm-m

2- 3

°HI

OHa

Og

~D

ZH2

Zg

1.2

3.5

537

• 0010

Z

~Z

.0065

.011

( % PC- ~ ) 1.1

2.4

5.3

3- 4

.26

13.3

.44

1.5

2.3

5.6

1.2

4.0

440

.012

.009

.015

4- 5

.40

2~7

.89

2.3

5.1

11.2

2.5

7.3

360

.030

.019

.033

5- 6

.49

28.5

.94

2.8

5.4

12.0

3.1

8.7

294

• 039

.029

.052

6- 7

.41

17.6

.58

2.4

3.4

8.3

2.3

6.8

240

.033

.027

.048 .O61

7- 8

.49

14.8

.53

2.8

3.1

8.4

2.4

7.4

194

.041

.036

8- 9

.44

8.2

.27

2.5

1.5

5.7

1.4

5.5

156

.035

.033

.051

9-10

.46

5.5

.18

2.7

1.0

5.2

1.0

5.2

125

.040

.039

.055

10-11

.41

3.9

.13

2.6

.7

4.6

.7

4.5

98

.045

,044

.058

11-12

.36

.9

.03

2.7

.3

4.1

.3

4.1

76

.051

.051

.059

12-13

.45

3.9

5.3

5.3

56

• 086

.086

.086

13-14

.28

2.8

3.8

3.8

38

.091

.091

.091

14-15

.11

1.2

1.6

1.6

29

.052

.052

.052

2.8

81

13 2~r. [~Y(R) .R. dR/109 m_

1.4

0.87

3.2

0.61

Star Formation and Abundance

II.I.3

Gradients

171

SURFACE DENSITIES AND THE GAS TO TOTAL ~ASS RATIO ~ AS A FUNCTION OF GALACTOCENTRIC RADIUS R

Surface densities of atomic and molecular integrating

the corresponding

hydrogen given in Table 11.2, are obtained by

volume densities

in column (1) and (2) over the scale heights =

given in Table 11.1. The total surface density of gas is obtained with X the mass fraction of hydrogen. galactocentric

Since X=]-Y-Z,

distance R due to the galactic Y,Z gradients

on R is considered

(Sect. IV). The dependence

of X

in computing ~ . g

The Z gradient enters into the determination

of the surface and volume densities of molecular

gas in yet another way, viz. through relation of 160 on R is representative the ratio

from ~g (aHI+~H)X-I, this mass fraction decreases with

(11.4). Assuming

for all primary elements

[H2]/[C0] =Z -n with n

such as 12C, one would expect that

~I-2. While this effect has not been considered

derivation of the volume densities and surface densities with n=l has been applied

to all quantities

of CO on Z is to be expected

that the observed dependence

which have a superscript

if the abundance

in the

given in Table 11.2, a correction Z. Such a dependence

of C-atoms in the gas phase is much less than

that of O. In addition to a variation of the [H2]/[CO] must be a negative

gradient

ratio Blitz and Shu (1980) have argued that there

in the A /N(ISco)

ratio as a function of R, to reconcile

the

V

observed high 13C0 column densities

(corresponding

to A =200 mag) with an extinction of only

30 mag derived from IR observations

for the visual extinction between the sun and the galac-

V

tic center. We do not accept their arguments, over a ] arcminute galactic

center may yield wrong results,

the dust) distribution

considering

in GMCs as mentioned

sion lines sample all molecules

the clumpiness

but the central

198]). Emis-

(GHsten and Henke~ 1982). The very massive belt

]0-20 arcseconds

to be in essence free of GMCs. As a result,

of the gas (and hence of

including the massive clouds thought

gas, which is located in front of the center,

CO observations,

emission lines sampled

10 arcsecond of the

above (see also Federman and Evans,

along the line of sight,

to be located behind the center of the Galaxy of molecular

since comparing molecular

beam with high resolution IR data of the central

appear

a comparison

fills part of the beam used for

(by a happy chance for astronomy) of the IR absorption

towards the

very galactic center, with CO emission related to the general direction of the galactic center does not appear to be appropriate.

In fact, we estimate

that the total visual extinction

between R=]O and 4 kpc is only ~20 mag, the value which follows from the gas densities

given

in Table 11.2. Of these 20 mag 8 mag would come from dust mixed with atomic hydrogen while ]].5 mag (corresponding with molecular

to N(13C0)

= 3 1016 cm -2) would be contributed

from dust associated

hydrogen.

Three different values of the gas-to-total

mass ratio ~(R) are given in Table 11.2: For ^

the computation

of ~ we use ~g, for z

we use ~ zg and for ~ z we use a surface density z g' which has been computed with an adopted [H2]/[CO] ratio of lO 6 rather than the standard value 4 105 (Eq. 11.4). As shown in Fig. 11.2, the gas fraction ~ is (within the uncertainties) constant over the disk 4 _
gas depletion

in the inner disk ( z ) .

with at most moderately

172

R. Gusten and P.G. Mezger

Integration over the galactic disk 2 ~;R ~|3 kpc yields a total (stellar) mass of

13 kpc 2~

Joe(R)RdR

(II.7a)

= 8.].]0 I° m e

2

and a total mass of gas

13 kpc 2w

~/f°g(R)RdR = (2.8-3.2).I09

(ll. Tb)

me

depending on if we use o zg or G g from Table 11.2.

~]

I

'

I

'

(5; 10 300 l(.j

®

® O

]GAS ---J (]~A$ l--]

_ j _ _L _r_ .n__r_

I

'

I

R

x.~i

~

iI

I

',

100

L.

'

.08 Q=:

.04

4

B RADIUS R IN kpc

12

Fig. IT. 2: Surface densities of the stars and the gas versus galactic radius. The gas-to-total mass ratio V is shown for gas masses derived for three different molecular abundances; viz • "p" is for [H2]/[CO] = 4.10 5, "V Z" for

[H2i>[CO] c~4.105 Z-1,

"tJZ" f o r

[H=]/[CO] =IOS.Z -1.

Star Formation and Abundance Gradients

III.

173

STAR FORMATION RATES IN THE GALACTIC DISK

The procedure used in this section for estimating star formation rates was first developed by Mezger and Smith (1976) and was refined in a subsequent paper by Smith, Biermann and Mezger (1978). The basic idea is to determine in HII regions the production rate of Lyman continuum (Lyc) photons~ i.e. the number of NLy c of Lyc photons emitted per sec by the ionizing O stars. Knowing both the mass spectrum of a newly formed generation of stars, the initial mass function (IMF), ~(m) and the Lyc photon production rate as a function of the stellar mass, one computes the mass to Lyc-photon ratio / for a newly formed generation of stars. Here < > means a weighting of m and NLyc(m) with the IMF. As can be seen from Fig. (111.4) the ionization of HII regions is dominated by O stars which have a MS lifetime TMS of some I0e yr. Summing up over all HII regions in the galactic disk then yields the total presentday (t=t o) star formation rate [me yr -I] T(to) =

NLy c

i

(llI.l)

all O stars

III.I

REVIEW OF EARLIER WORK

For a determination of the Lyc photon production rate in an HII region one measures the radio flux density S~ of its free-free emission and estimates its (kinematic) distance D from the radial velocity obtained from its radio recombination line emission. The "radio luminosity" S D = of the HII region is proportional to its

Volume Emission Measure

f n.n dV I e V(HII) =

with ni, n e the ion and electron density and V(HII) the volume of the HII region. This volume emission measure is determined by an equilibrium between ionization (i.e. the number N' of Lyc Lye photons absorbed per sec by the HII region) and the corresponding number of recombinations of ions with electrons, which is proportional to fninedV. In this way one obtains a relation between S D 2 and N' which is given by Eq. (lll.2a) To obtain from N' the Lyc photon Lyc' " Lyc production rate NLy c of the ionizing stars one has to correct for the fraction of Lyc photons which are absorbed directly by dust grains within the HII region (important for compact HII regions), and for the fraction of Lyc photons, which escape from a density bounded HII region (can be important for extended low density (ELD) HII regions). While this method of determining NLy c yields very reliable results for an individual HII region, the problem in determining star formation rates rests with the completeness of the sample of HII regions used to derive the quantity ENLy c for the whole galactic disk. To understand the implications of this problem let us briefly review the evolution of an O star and its surrounding HII region as it emerges from a variety of observations. An O star forms by accretion on a previously existing starlike core. This process is eventually halted by radiation pressure acting on the dust, which is coupled to the infalling gas. In this way first a dust cocoon is formed, which intercepts all stellar radiation, which is subsequently reemitted in the far infrared (FIR). This cocoon is destroyed in a relatively short time and Lyc radiation from the 0 star forms a compact HII region of high electron density and small diameter, which appears as a point source in radio surveys at short cm wavelengths. Giant radio HII regions are usually composed of a number of such compact subcomponents ~ A

26:3

- B

174

R. Gusten and P.G. Mezger

Compact HII regions expand, and, while their size increases, their density and central emission measure decreases. Their lifetime as easily recognizable radio HII regions T(radio HII) appears to be of the order of one tenth the lifetime of O stars. Expansion of the HII regions continues, until they achieve pressure equilibrium with the surrounding neutral gas, or until the ionizing O stars move off the main-sequence. Once the electron densities have decreased to some IO cm-a, the once compact adjoining HII regions have merged and now form an extended low density (ELD) HII region which, because of its low surface brightness, is difficult to observe as individual HII region. However, the integrated emission of all ELD HII regions yield the diffuse galactic free-free emission, which accounts for 80 to 90% of the total freefree emission from the galactic disk, and which is easily observable in low angular resolution surveys of the galactic plane at decimeter wavelengths. For their estimate of the present-day star formation rate (SFR) Smith et al. (1978) used radio HII regions only. They argued, that their sample of giant HII regions in the galactic disk is complete for one half of the galactic plane. They assumed that all giant HII regions are located in the main spiral arms and that O star formation in the interarm region only produces small HII regions such as the Orion Nebula. From a comparison of the flux densities of small and giant HII regions they estimated ENLyc(small Hll)/~NLyc(giant HII) = O.175, a ratio which they assumed to hold for O star formation rates in the interarm region and in main spiral arms, respectively, throughout the galactic disk. In Eq. (111.1) the ratio ENLyc(all O stars)/ has now to be replaced by ~NLyc(radio Hll)/T(radio HII). The main uncertainty in the SFR estimated by Smith et al. comes from their estimate of the lifetime of radio HII regions, T(radio HII). To improve this estimate Mezger (1978) has tried to determine the total Lyc photon production rate of all O stars in the galactic disk, ZNLyc(all 0 stars) = ENLyc(radio HII) + ENLyc(ELD HII), by deriving the Lyc photon production rate of O stars associated with ELD HII regions from radio surveys of the diffuse free-free emission made at ~X21 cm. This involves two major problems, viz. the separation of the diffuse (thermal) free-free emission and the diffuse (non-thermal) synchrotron emission, and subsequently the deconvolution of the separated freefree emission. The first problem, the separation of thermal and non-thermal radiation, has been assumed to be solved by the authors of the surveys, who compared their radio maps at decimeter wavelengths with radio maps made at meter wavelengths, where the synchrotron emission predominates. A first deconvolution, i.e. the transition from the distribution of freefree flux density observed as a function of galactic longitude to the distribution of the Lyc photon production rate in

the galactic plane, was made by Mezger (1978) assuming radial

symmetry of the distribution NLyc(R) in the galactic plane. However, since a large fraction of all O stars in the galactic disk is known to be formed in spiral arms, the assumption of radial syn~etry in the Lyc photon production rate and hence in the SFR could lead to wrong results.

111.2

LYC PHOTON PRODUCTION RATES DERIVED FROM RADIO OBSERVATIONS

To obtain the star formation rate in a given volume, e.g. the whole galactic disk or in concentric rings surrounding the galactic center, we have to determine the Lyc photon production rate of all O stars contained in this volume. As mentioned in Sect. 111.1 we make use of the fact, that there exists a relation between radio luminosity S D 2 and Niyc, the number of Lyc photons absorbed per sec by gas in the HII region, which emits the free-free radio flux density S . We take the quantitative relation from Mezger et al. (1974)

Star Formation and Abundance Gradients

175

L Here y i s the tHe abundance by number, R _
production rate NLy c one has to correct for (l-fne t), the fraction of Lyc photons directly absorbed by dust , and for fesc' the fraction of Lyc photons which escape from the galactic disk. Hence

NLyc = (l-fesc)-I

f-~net N'Lyc

(lll.2b)

In this subsection we apply this procedure to the observed diffuse free-free emission, using an improved analysis of observations

III.2.1

and taking into account the spiral structure.

CONTRIBUTION TO ZNLy C FROM GIANT RADIO HII REGIONS

To obtain ENLyc, the total Lyc photon production rate, we have to consider both O stars ionizing radio HII regions, NLyc (I), and O stars ionizing the more evolved ELD HII regions, NLy c(4). For giant radio HII regions we use NLyc(R), as estimated by Smith et al. (1978). Their values for N'Lyc(1), NLy c(I) for giant radio HII regions are given in Table III. I.

Table III. l: The Lyc photon production rate in I051 s-I in the galactic disk as function of the galactocentrlc radius R. Radius

N~yc(

I)

NLyc (I)

NLyc (2)

N~yc(3)

NLyc(3)

N~yc(4)

NLyc

(4)

NLyc (5)

kpc

Radio H I I - R e g i o n s

ELD Hll-Reglons

2-3 3-4

<0.04


8. l ± 3 . 7

lO.1

0.08

4.9

11.9

14.6±6.8

18.3

4-5

1.8

3.6

3.7

38.2

54.6

40.9±8.2

51.1

5-6

4.1

11.2

3.2

34.6

49.4

45.5±6.3

56.9

6-7

2.3

6.1

1.3

20.4

31.8

23.9±11.7

29.9

7-8

4.4

9.3

5.8

14.2

25.0

20.8±9.4

26.0

8-9

2.1

4.5

0.6

10.7

20.9

17.3±6.1

21.6

9-10

1.5

3.2

2.2

lO.5

21.4

7.9±2.7

9.9

I0-11

2.9

5.2

5.0

13.3

11-12

2.2

3.8

3.6

10.7

12-13

0.5

0.73

3.2

10.9

145.3

249.9

Z(3-]3)

21.8

47.6

~(3-]0)

179.0±55

223.8

lO.l 18.4 54.7 68. I 36.0 35.3 26.l 13.1

271.4

51.0

(I) Smith et al. (1978). (2) This paper, for galactocentric segment 0 ~0 ~120° only. To correct for incompleteness Z(3-10) ~3 E(8 = 0-120°). (3) Mezger (1978).

(4) This paper, assuming azimuthal symmetry. (5) NLyc (5) = NLyc(4) + NLyc (I).

176

R. G~sten and P.G. Mezger

SgD2and y+ were taken from earlier radio surveys, fnet was computed individually for each giant HII region and f e s c =

0 was adopted, since most radio HII regions are ionization

bounded or, if Lyc photons escape from a radio HII region, they will contribute to the ELD HII regions. The production rate NLy c (m) , which is given for comparison purposes, is from the survey by Downes et el. (|980). This survey is complete for giant HII regions in the galactocentric segment O ° ~@ ~120 ° (@=0 ° toward the galactic center, and counting clockwise). We do not consider small radio HII regions, because beyond a few kpc they would not be detected as individual sources even in surveys with the IO0-m telescope, but they would merge with the background formed by ELD HII regions, and therefore should be unresolved in low resolution surveys and hence counted as part of the diffuse galactic free-free emission.

111.2.2

CONTRIBUTION TO ZNLy C FROM ELD HII ~ G I O N S ,

ASSUMING R A D I ~

SYMMETRY IN T ~

GALACTIC

PLANE Mezger (1978) used the distribution of the diffuse galactic free-free emission as derived by Westerhout (1958) and Mathewson et al. (1962), and estimated

NLye,' assuming a r a d i a l

symmetry

of the volume emissivity of the free-free emission in the galactic plane. This procedure assumes implicitely that star formation does not depend on the azimuthal angle. Since we know that 0 stars form preferentially in spiral arms this assumption can only be a first approximation of reality. Here we proceed in two steps. In Appendix B we deconvolve the diffuse freefree emission assuming radial symmetry and considering the propagation of the rather large uncertainties inherent in the separation of diffuse free-free and synchrotron radiation. This devonvolution yields a free-free emissivity c(R), which we use to estimate N~yc(R) , the number of Lyc photons absorbed by the gas at galacto-centric distance R. The volume emissivity c produces at a distance D the flux density

4~DmS

= ~cdV

(111.3)

Substitution of this relation in Eq. (lll.2a) yields, with T

= 7000 K (Mezger, 1978), an e + estimated average ionized He-abundance in the ELD HII regions of RoY = y = 0.08, with = 1.4 GHz and a gaussian z-distribution of the emissivity of scale height Zo, z Rou t N' = 4.4.10~s.~°'1.T - ° ' ~ f f c.R dR dz Lyc e o R. znn

2

= 4.6.10~6.Zo.C ° (Rou t

with Zo,Rout, Rin n in kpc and go = c(z=O)

_

2

Rin n)

in erg s-I Hz-lkpc -3. Substituting go(R) from

Fig. B.2 yields N' (~)(R) in Table III.I. For comparison the corresponding values N' (s) Lyc Lye ' derived by Mezger, are given also. The deconvolution in Appendix B yields generally higher values for the inner annuli and is set to zero outside the solar circle. Here we rediscuss the values of fnet and fesc as adopted by Mezger (1978) in the light of more recent results. It was found that the average density of ELD regions, n e ~I0 em -3 (Mezger et al., 1982), is considerably higher and hence the distribution of ionized gas much patchier than assumed earlier. That leads to smaller optical absorption depth for Lyc photons

Star Formation and Abundance Gradients

177

and hence we adopt fnet = 0.8 rather than 0.7, the value used by Mezger. Since ELD HII regions now appear as distinct entities, whose lifetimes are much shorter than that of the associated molecular cloud, they are likely to be at least partly surrounded by relatively dense gas. In addition, recent estimates of the volume filling factor of HI

clouds (Knude,

198]) show, that with a scale height of ~]20 pc most of the solid angle 4w seen from the galactic plane will be covered by these clouds, and hence only few Lyc photons can escape from the galactic disk. Therefore we put fesc=O. Substituting

these values in Eqs. (III.2) we

obtain NLy c (~)(R) in Table III.]. The total Lye photon production rate ZNLy c (s) (R) is then obtained as the sum of the Lyc photon production rate of O stars located in giant radio HII regions and in ELD HII regions, respectively. 13 ZN = (giant radio HII) R=4Lyc ZNLyc + ZNLy c (ELD HII) ÷O.as) ]OS3 Lyc photon S-1 2.7(_O.

=

(111.4)

The corresponding number derived by Mezger (1978) is 3.0 1053 s-I, while Mezger and Pauls (1979) estimated a Lyc photon production rate of 2.7 1052 for the galactic center region R
III.2.3

in (111.4) are those given for N' (~) in Table III.1 Lyc

DECONVOLUTION OF THE DIFFUSE FREE-FREE EMISSION CONSIDERING GALACTIC SPIRAL STRUCTURE

Most recent attempts to determine the spiral structure of the Galaxy are based on the distribution of giant HII regions

(Mezger,

]970). These attempts have not led to a clearcut picture.

Georgelin et al. (1980) suggested that a four-armed spiral fits the distribution of giant HII regions best. Wilson (1980), considering the large distance uncertainties, an overinterpretation

criticized this as

of observational data. Here we adopt the simpler picture of a two-armed

trailing spiral with variable pitch angle tg(i)=to+tlR , which rotates around the galactic center with a constant pattern velocity ~p, which is lower than the angular velocity ~R" with which stars and ISM rotate. In cylindrical coordinates scribed by the relation

(Lin et al., ]969; Burton,

(R,0,O) the spiral pattern can be de-

]97])

toRoexp~-(G-Oo)t o } R(O) = tlRo+to-tlRoexp{-(@-@o)to }

The parameters

to=O.|936 , t]=-O.O]O6,

R ° =6.846 kpc and 0 o =1.549 rad are chosen to fit observa-

tions which suggest, that the galactic longitudes 1=29 ° and 49 ° intersect tangentially the two spiral arms (in Fig. lll.l labelled as Scutum and Sagittarius arm, respectively). speed is

= ]3.5 km s -] kpc -1 P and the angular velocity of stars and interstellar matter at a distance R is 250+4.05(10-R) ~(R)

in units of km s -I kpc -I.

=

R-~



- 1.62(lO-R) a

885.4 R -°'s - 3 104 R -3

4
The pattern

178

R. G~sten and P.G. Mezger

Fig. III.] shows the adopted two-armed • U )150 crn-Zpc o 125
spiral pattern on which is superimposed the distribution of "supergiant HII regions" (i.e. HII regions with SsD ~ ~800 Jy kpc ~) as derived by Downes et al. (1980). While the supergiant HII regions tend to follow the spiral pattern this correlation would disI

appear, if all radio HII regions were included in the diagram, a fact already noted by Mezger (1970). This is to be considered as an indication, that the surface density of newly formed 0 stars is high in spiral arms and low in the interarm region, but that the total rate of O star formation in the interarm region is certainly not negli-

Fig. III. l: The two-a~ned trailing spiral adopted to describe the spiral structure of the Galaxy compared with the observed distribution of '~upergiant HII regions". Note that excitation parameters u >125 cm-2pc corresponds to radio luminosities $5D2>800 Jy

condense out of giant molecular clouds (GMCs)

kpc

in the main spiral arms. However, the role

gible. Observations suggest that the bulk of O stars

of molecular clouds as an intermediate stage between the diffuse ISM and star formation and the mechanism that triggers star formation in spiral arms is far from being understood. Either GMCs have long lifetimes (compared with the travel time between adjacent passages through spiral arms), and when entering the density wave are compressed and induced to collapse by the sharp pressure increase in the intercloud medium (Shu et al., 1972).

How-

ever, more recent investigations (Shu, 1978) have shown, that such a compression will not take place if the hot "coronal" gas (whose characteristics are given in Table ll.l) fills most of the space. Or GMCs form in spiral arms out of small molecular clouds (SMC's) through agglomeration (Combes,

1981), possibly with the help of Parker instabilities (Norman and Silk, 1980). In

this case their lifetime would be short as compared to the travel time of the ISM between two subsequent passages through spiral arms and most GMCs would decay into SMCs with considerably longer lifetimes. Such a scenario is supported by observations, which show a strong concentration of GMCs towards spiral arms and a rather uniform distribution of SMCs throughout the interarm region (Stark, 1982, and references therein). The increased formation d GMC rate of GMCs, ~-{ ~ H 2 }, for which we assume that it is proportional to the amount of molecular gas that flows through a spiral arm, leads to an increased surface density ~

~

sa

dF d

of the SFR

GMC ~ H 2 }dF = ~H V(R) 6{@-@m(R)}dRdO

Here ~H is the surface density of gas contained in SMCs which have managed to survive the travel through the interarm region, ~(x) is the delta function

with ~(x)=l for x=O and zero

elsewhere, @m(R) is the azimuthal angle of the m-th spiral arm and V(R)=R (~R-~p) is the relative velocity with which stars and IMS stream into the spiral arms. With ~R and ~p as given above, ~O=~R(R = IO kpc) and

Star Formation and Abundance Gradients

179

>I for R
=

~R-~p ~®_~p

l for R = I0 kpc

(III.5)

I0 kpc we rewrite

opsa= OH~(QO-~p)~{O-0(R)}

~(R)RdRdO

which, after integration, yields the SFR in spiral arms within the annulus R m

~sa(R) = MH2(R). ~-~ (Qe-Qp) ~(R) Star formation in the interarm region will take place in a small number of GMCs which have escaped from spiral arms; but primarily in SMCs. This means the SFR in the interarm region is proportional to the total amount of molecular gas

2~ Rou t ~ia(R )

,xf f ~HRdRdO= MH=(R) o

The total S F R ~ e n

Rin n

the sum of the two star formation processes m ~(R) = ~ia(R) + ~sa(R) ~ M H (R) [I+~' ~-~ (~@-~p) ~(R)]

with a' a proportionality factor, which is to determine observationally. E.g., if we know ~sa.~ia for the~sglar vicinity the ratio ~=T 8 IT® , of SFRs, in spiral arms and in the interarm re-

\\ ~k

gion, respectively, then ~0=MH=(10)

[1+~] and

MH2(R) P(R) = P® ~

where ~ is related to a' by ~ =

[]+~(R)] [l+~]

(III.6)

e' ~F m (~8_~p)

Let us now return to the problem of deconvolution of the diffuse galactic free-free emission. This distribution should follow the distribution of 0 stars whose formation - owing to the short MS lifetimes of these massive stars - should be adequately described by

o~(R,O) =

o~a + o~ a

= ~H (R) [I+ 2 ~ m ~)(R)~[@-Om(R)}]

(III.7)

The ridge line intensity of the diffuse free-free emission as derived by Westerhout (1958) with a telescope HPBW 0A=34 arc min and converted into Jy per beam area, is shown in Fig. 111.2. We assume eff(R,O) = ~ ( R , O ) .

To account for the finite beamwidth we assume a

gaussian distribution of the free-free emission perpendicular to the galactic plane, exp{-z/l.2 Zo }' with Zo=120 pc (Mezger, 1978). For ~H2(R) = nHa(R), with the latter quantity taken from Table 11.2 and substituting eff(R,0,z) =nH2(R)ex p {-z/].2 zo } • [ I + ~ 2-~~ ~(R) ~[O-Om(R)}]

(III.Sa)

180

R. Gusten and P.G. Mezger

in Eq. (111.3) yields

Sv =

with ~m = | . 1 3 3

~ !

fgffdV = ~ m fEffdrd~

(lll.8b)

@A the antenna solid angle and r a polar coordinate centered at the sun. To

fit Eq. (lll.Sb) to the ridge line intensity shown in Fig. 111.2 we smear out the delta function in Eq. (Sa) to a width A=! kpc [~(~-~p)-R.Tms]

to account for the fact that 0 stars

emit Lyc photons during all of their lifetime. It has been suggested that the SFR may have k a non-linear dependence on n H. Hence we generalized Eq. (lll.Sa) by putting gff = n H. In Fig. (111.2) are shown fits for a linear dependence on the molecular density (i.e. k=l, Eq. lll. Sa), A=I kpc, ~=0 (i.e. no spiral structure)

and 1, and molecular densities nH2(R)

Z (R) as given in Table 11.2. We also varied A,k and substituted for nH (R) recent denand nHm sities derived by Liszt et ai.(1982). Although we obtained with either k>! or the Liszt et al. densities slightly better fits, the accuracy of the observational

data, in our opinion,

does not justify to deviate from our set of consistent gas densities given in Table 11.2, or to introduce a non-linear dependence on the molecular gas density. In summary a consideration of galactic spiral structure in fitting the observed ridgeline intensity of the diffuse free-free emission leads to the following conclusions: i) The introduction of spiral structure

(i.e.~>O) is needed to fit the observations. sa ia ii) The fit becomes relatively insensitive to the parameter ~=P@ /~® for ~ O . 5 . Thus we can only give a lower limit for this parameter, which describes the fraction of stars formed in spiral arms relative to the fraction of stars formed in the interarm region, at R=IO kpc.

I

_

I

160

• . =

I

/kUoc(nHl+2".nHZ) z

/

I

811

g

/ ~

q 2 ~)

=

:

(

n

,* H

.

:..." .

-~ -

-

o . - ~ . .'° ". . . : ~

.

.." .

.° • ~

.'.

."

. . . . ".

.'

.



.."

._.1 u._



°

I

I

I

I

50

C,0

30

20

GALACTIC

LONGITUDE

(DEG)

Fi.g. III. 2: Ridgeline flux density per beam area of the diffuse galactic free-free emission a8 observed at 1.4 GHz. Curves relate to fits of Eqs. (III. Sa, b) with i) no spiral structure (~=0) and the SFR proportional to the average density of molecular gas (~); ii) no spiral structure (a=O) and the SFR proportional to the square of the average ~otal gas density (~<~Hi+2.nHz>2);iii) spiral structure (~=1) and ~=. o ~ a / ~ a is the ratio of the SFR in spiral arms to that in the interarm region at R ~1~ kpc. The upper limit of the hatched ar~a is obtained from z in Table II. 2, while the lower limit is based on Ha density e8timates by Liszt et al. 2

Star Formation and Abundance Gradients

181

iii) Although a higher than linear dependence of the SFR on ~H~ cannot be excluded from the data, there seems to be no need for this

to obtain reasonable fits. Even more, the similar

scale height of 50-60 pc for both the molecular gas (Stark, 1982; Sanders et al., 1979) and radio HII regions (Lockman, 1979), favors a linear relation. iv) However, models of the chemical evolution of the ISM require a relation between SFR and the total mass of gas, both in molecular and atomic form. Formally, we can substitute in Eq. (111.7) o~ =(OH +oH)k[l+...] and find, that this dependence requires values I
111.3

of cff only.

QUANTITATIVE VALUES FOR GALACTIC STAR FORMATION RATES(SFRs)

Eq. (III.I) relates the Lyc photon production rate of 0 stars to the total SFR. Since most of the mass of a newly formed generation of stars resides in stars less massive than O stars, a quantitative value of the SFR will depend heavily on the adopted initial mass function (II~F), After a discussion of observationally determined IMFs we reformulate relation III.I and subsequently evaluate SFRs adopting a constant IMF. While the relation between Lyc photon production rate and the formation rate of O stars is rather straightforward the separation of diffuse free-free and synchrotron emission is rather uncertain, especially in external galaxies. Most galaxies emit a strong diffuse IR radiation and we discuss the relation between this diffuse IR emission and star formation.

III.3.1

INITIAL MASS FUNCTIONS (IMF)

The value of the mass to Lyc photon production ratio / in Eq. (III.I) is in essence determined by the IMF, ¢(m), i.e. the number of stars in the mass interval (m,m+dm) of a newly born generation of stars. The IMF is derived from star counts in the solar vicinity, which for stars with MS-lifetimes shorter than the age of the disk have to be corrected for stellar evolution. Moreover, one has to take into account the change of the SFR during the galactic history (see sect. V). It is customary in a quantitative description of star formation to separate the creation function, C(t,m), i.e. the number of stars formed per unit time in the mass range, (m,m+dm), into a time-dependent and a mass-dependent function

C(t,m)dm = ~(t)~(m)dm. ! Here, ~(t) is what we refer to as SFR, i.e. the amount of mass transformed per unit time into stars, if the IMF is normalized to m up f ~(m)m-dm = 1 mlow

182

R. Gusten and P.G. Mezger

with mlo w and mup the lower and upper termination of the stellar mass spectrum. Smith et al. (]978) used the IMF derived by Salpeter (1955) for the mass range 0.4 ~an/m@~]O, with mup =

100 m0 @sal(m) = 0.35 [mT°'35-m~'3s] -I m -2"as low up

(111.9)

More recently Miller and Scalo (1978) have reanalyzed star counts which they consider to be complete for m ~O.I m@. They derived a log-normal functional dependence, which, with the above normalization, yields the IMF

@ms = 3.] [4.44-2.5 m°'~] -I m -~ exp{-l.O9 (log m+l.O2) m} low

(lll.]Oa)

For analytical computations we use a power law fit to four segments of the Miller-Scalo IMF

l.O

qbms

O. I
I . I m- 2 " #

1.0


6.3

m- 3 " 2 #

lO

~mlm 0 <50

27.2

m- 3 " 6 ~

30


[4.44-2.5 m°'W] -I. low

=

m- I . e

(III.lOb)

As shown in Fig. 111.3 the two IMFs agree well in the mass range for which Salpeter has derived his IMF; outside this mass range, however, where it is common to use an extrapolation of the Salpeter IMF, the Miller-Scalo IMF favors lower values. This reduces the influence of both very massive stars on the IMF-averaged Lyc photon production rate and of low mass stars on the total amount of matter which is permanently locked up in stars. All following quantities, which depend on the IMF, have been evaluated for both the Salpeter and Miller-Scalo IMF, but for quantitative results we quote mainly those related to the Miller-Scalo IMF. Here we use mlow=O.l m® and mup=60 m® for the solar vicinity. It appears that m

up

decreases with decreasing galactocentric radius R. From observations of radio recombination lines of H and He it has been found, that the He + StrSmgren sphere in galacI

I

10 z' 10 2

I

INITIAL •,,,

I

MASSSPECTRUM(;O(m)-

tic HII regions shrinks relative to the size of the corresponding H + StrSmgren sphere

~ SALPETER (a': 2.35)

from the edge towards the center of the galactic disk

.~ 10 0 '0. E

(see, e.g. Mezger, ILLER-SCALO(~': Z+LOG M )

]980).

Panagia (1980) explains this

10-2 I0-~

I

I

I

0.1

I

10

STELLAR MASS m IN m o

100

Fig. III. 3: Normalized initial mass function ~(m) as derived by Miller and Scalo (1978) and Salpeter (1955), respectively, from star counts in the solar neighborhood. The hatched area gives the uncertainties as estimate by Miller and Scalo. T describes the mass dependence of the function 91og~(m)/Dlog m.

Star Formation and Abundance Gradients

183

effect by a softening of the Lyc radiation (i.e. a depletion of He ionizing Lyc photons), caused mainly by a decrease of mup and hence of , the IMF weighted mean effective temperature of 0 stars with decreasing galactocentric distance R. Such an effect is to be expected, if with increasing heavy element abundance Z(R) the dust-to-gas mass ratio increases, too, and if the mass of an accreting O star is determined by radiation pressure, which acts on the dust and stops the infall of the accreting gas. For this ease Kahn (1974) has predicted a dependence m

up

=Z -°'5, and in the following we use

mup(R)

= 60

(Z(R)/Ze)-°'s

me

( I I I . 11)

with Z(R) as given in sect. IV, Table IV.I.

III.3.2

A NEW FORMULATION OF THE RELATION BETWEEN LYC PHOTON PRODUCTION RATE AND STAR FORMATION RATE

The Lyc photon production rate of a star, NLyc(m,T), depends on both its mass and its mainsequence (MS) age T. The characteristics of 0 stars are discussed in Appendix A and the quantitative relations needed in this subsection are derived. Stars in the mass range m,m+dm contribute to the present (t=t o) Lyc photon production rate to NLyc(m)dm

=

dm .f ~(t')~(m)NLyc(m,to-t')dt'

t o -T(m)

T(m) =dmf~(to-T)~(m)NLy c (m,T)dT o with (to-t')

:

T.

I

,m

I

I

1.0

i

I

/~ I---o

'

~

I

=

"/NLYc (m''~)d'~

z

~--

.8

"\

/

///

e~

//

i

~i

\

°\/SALPETER

-%

.2J

1O

I

I

I

30 50 STELLARMASSm IN m®

I

I

/

70

FiE. III. 4: The IMF-weighted Lyc photon production of stars of mass m, integrated over the stars MS-lifetime. Number of photons may be obtained by multiplication with 3.1059(~ sal) and 9.3 105s(~). Note that due to the normalization ~(m)/~(1), the numbers are independent on mup, mlo w"

184

R. Gusten and P.G. Mezger

TMS The quantity [qb(m)/@(1)]f NLyc(m,T)dT , shown for two IMFs in Figure 111.4, is practically zero for m <10 m@. The tOtal Lyc photon production rate is obtained by integrating over the mass range m >|0 m e (spectral types BO or earlier) mu up _TMS _ (m) NLyc(t o) = 7 - d m J d T ~(to-T) ~(m) NLyc(m,T) d~ ]0

o

With a SFR ~(t)=~(to) , which is constant within the MS lifetime of 0 stars, and after summation of the Lyc photon production rate over the area, whose SFR shall be determined, this relation can be rewritten in a form which is equivalent with Eq. (Ill.l)

ZNLyc(to) ~(to) = ~(1)P(#)

with

mup

(III.12)

TMS (m)

I)] P(*)= fdm fdT NLyc(m,T)[*(m)/*( I Om®

o

Analytical expressions for $(])P(~) are given for both ~sal and ~ms in Appendix A. In Table 111.2 we give ~(1)P(~) for mup=60 m® and three different values of mlo w. Table 111.2:

111.3.3

~(1)P(~) in Lyc photon s-I/[m® yr -I]

mlow=]O m®

mlow=].O m®

mlow=O.l m®

~sal

5.3

1.4

5.3

~ms

3.2 lO sa

10 s3

10 s3

4.3 10 s2

10 s2

2.5 10 sa

STAR FORMATION RATES FOR CONSTANT INITIAL MASS FUNCTIONS

13 (5) (s) from Eq. (111.4) in Eq. (111.12) Substitution of NLye(R) from Table lll. l and R~ 3 NLyc(R) yields the SFR in annuli of width AR=l kpc around the galactic center and in the galactic disk, respectively. For a given Lye photon production rate the corresponding SFR depends on both the form of the IMF, ~(m), and on the upper and lower mass cut-off, mup and mlo w. To show the influence of the various parameters we have evaluated Eq. (III.12) for ZNLy c from Eq. (III.4) and different parameters of the IMF. mlo w ~O.OI-O.l m@ is a theoretical limit imposed by opacity-limited fragmentation (Silk, 1978), with the upper value close to the fragment mass (~0.08 m@), below which nuclear burning can not be ignited, mup ~IO0 m@ corresponds to 03.5 stars, mup ~60 m® corresponds to 05 stars being the most luminous stars which form at a statistically relevant scale in the galactic disk. mup=60 (ZR/Z@)-°'m m G corresponds to a variable upper mass limit imposed by a variable metal abundance. For a given Lyc photon production rate the Salpeter IMF yields a SFR which is about one half of that obtained with the Miller-Scalo IMF. As one recognizes from Fig. 111.4 this is due to the fact that there are, relative to lower mass stars, fewer 0 stars in the Miller-Scalo IMF than in the Salpeter IMF. Deviations of the Miller-Scalo IMF from the extrapolated Salpeter

Star Formation and Abundance

Gradients

185

Table 111.3 Star formation rate ~(t o ) in the galactic disk between R=3-13 kpc

Mass spectrum (IMF)

[

¢(ra)

Scar formation rate

mLow[m@]

mup[m@]

0.1

3 3

I0.8_"1-3a

60

001 ....

0.1

13 -+a.e -2 •

0.01

6O

0.1

1 O0

0.1

60

0.01

60

13. ] - 2 . ,

! O0

3.3_0.e

O.l

IMF at both the low-mass and the high-mass rived with the Miller-Scalo

~(t o) in [ ~ yr -1 ]

12.6 +3.s -2.3 +2

s.s_+~ 'e.~ +3.

6

+0.9

limit also account for the fact, that SFRs de-

IMF are much less sensitive

view of the fact, that the Salpeter

• 5

8.9_1.s

to changes

in both mlo w and mup. In

IMF was derived only for the mass range 0.4-]0 mo, we

adopt in the following the Miller-Scalo

IMF as the correct IMF for the whole mass range of

stars. We further consider stars with 60 m e (i.e. 05 stars) as the most massive and luminous stars, which form at a statistically

significant

rate at least in the vicinity of the sun,

although we are aware of the fact that stars of spectral (e.g. Humphreys, mlow=O.Ol

1978). We further adopt mlow=O.l

type 03 have actually been observed

m®, although the choice of a lower value,

m® would increase the formation rate by only 15%. With these parameters

adopted we

derive a total SFR in the galactic disk (SFR) as given in the upper two rows of Table 111.3, ~(to) =I]-13 m G yr -I. For the following discussions

13 ~(t o) = R~ 3 ~(to,R)

-I

(ZR/Zo) -° s m®.

of ±3 m e yr -I is from our statistical

(see Appendix B), and systematic

(111.13)

= [13 ± 3] m e yr

derived for a variable upper mass limit mup=60 The uncertainty

we adopt the value

analysis of the deconvolution

effects which we will discuss briefly, may change the in-

ferred SFR in either direction.

At first there is the uncertainty

tion of thermal and non-thermal

components

based on low-frequency

(Hirabayashi, production

introduced by the separa-

of the diffuse galactic radio emission,

surveys, where the non-thermal

vations at higher frequencies,

only,

radiation

still dominates.

which is

But obser-

e.g. at 2.7 GHz (Altenhoff et al., 1960) and ]5.5 GHz

]974), tend to confirm these earlier estimates on TB(1) and thus the Lyc photon

rates derived here on the basis of the Westerhout

Such a reliable

separation of free-free

and synchrotron

and Mathewson

emission

et al. surveys.

is only possible

in our

Galaxy, because of our position close to the galactic plane and the sma]l scale height of the free-free emission

(z

~120 pc as compared to 5OO-10OO pc for the synchrotron

emission).

O

This makes the ridge line brightness that of the synchrotron

temperature

of the free-free

emission at b ~0 ° exceed

emission for % <6 cm. The situation is different

systems, where radio observations emission so that the synchrotron

with single dish telescopes emission,

due to its larger scale height,

the total flux density even at 20 GHz (Klein et al.,

in extragalactic

do not resolve the free-free seems to dominate

1982). Similar problems are encountered

if one tries to estimate SFRs from the far infrared emission of dust heated by absorption of

186

R. G~sten and P.G. Mezger

starlight. Here the main problem is to separate heating due to early-type stars and heating due to the older stellar population (see Sect. 111.5.|). The SFR derived in this paper would be overestimated, if sources other than OB stars contribute to the ionization of the ELD HII regions. The most likely alternative sources, viz. the hot central stars of Planetary Nebula appear, however, to account only for a minor fraction of NLy c (Mezger, 1974; Salpeter, 1977). The sharp rise of NLy c towards the central part (R ~4 kpc) of the Galaxy supports this conclusion, since the distribution of Planetary Nebula should follow that of the old disk population II, and not that of the young population I. The SFR depends sensitively on the number of OB stars per stellar generation, and IMFs which fall off less rapidly in the upper mass range would reduce the ratio NLyc/~ (see Table III.3) Lequeux (1979) estimates an uncertainty of ~0.2 in the slope of the upper IMF, and if we reevaluate Eq. (III.13) with a flattened Miller-Scalo IMF

CmS(m >|0 m®)= m -3"°, the SFR is re-

duced by ~20%. On the other hand, we have neglected the possible escape of Lyc photons into the galactic halo (Sect. 111.3), and the effect of mass loss on the evolution of massive stars (App. A); effects which, taken into account, could increase the SFR by a similar amount We therefore feel that the absolute error in the SFR cannot exceed the statistical error as given in Eq. (III.]3) by much, provided the assumption of a constant IMF i8 correct. Having determined the total SFR in the galactic disk we want to know its radial dependence, i.e. the quantity ~(to,R). The straightforward solution is to substitute

'

I

I

'

NLyc(R)(5) from Table III.1

I

I

I

I

(a)

[rn®yr -I'.

(b)

r-.1

3

["

:

..

T"3

I---

""

.°..I

2

z

F---

L.I--.

q

F--0

0

I LOfAL SFR ~pre I

........

,

/,

I

8

,

,

I

I

I

,

/, 12 GALA[TIE RADIUS R IN kpc

,

8

I

12

Fig. III. 5: Observed and predicted SFRs vs. galactocentric radius R. Filled and open dots (connected by a vertical bar) refer to SFRs derived on the assumption of azimuthal syn~netry • and for upper mass limits mu3~ o: Z- 0 •5 (o) and mup = const. (). Histograms refer to SFR8, whose dependence on R, gas density nil, and the parameter a=~ 8s a /~8l a is given by Eqs. III. 6 and 7.<~> is the time averaged SFR according to Eq. (III.15), ~Dre i8 predicted by the closed system model [EqZ (I" 10)]" The histograms renresent. In ~+ia 5a SFRs with a=l, ~nT,£ 1 2 c • ~ • (dashed curve) and ~ ~n,, (solid curve) In Fig 5b SFR8 with ~ =n z, , a=O 5 dashed curve, ~=1 solzd curve, a=2 dash-dotted curve. •

£I

"

"

212

"

Star Formation and Abundance Gradients

187

in Eq. (III.]2),

~(to'R) =

N(S) -. Lyc (K) #(1)P(~)

(ln.]4)

and evaluate this relation for mup=60 (ZR/Ze) -°'s m G and constant mup=60 m®, respectively, yielding the open and filled dots in Fig. 111.5, which are connected by a vertical bar. Re(s) member that NLyc(R) has been obtained from a deconvolution assuming radial symmetry. However, in subsection 111.2.3 we have shown, that to reproduce the observed ridge line intensity of the free-free emission the galactic spiral structure has to be incorporated

in the fit, using

a radial and azimuthal dependence according to Eq. (111.7). Fixing the number of spiral arms, sa ia m=2, there remains only one free fit parameter, the ratio of SFRs c~=~O /~s) in spiral arms and interarm region, respectively,

in the solar annulus. Acceptable fits were obtained for

~>O.5. The fitted relation ~(t ,R)=~sa+~ sla according to (111.6) is shown for ~=O.5,1,2 and o

gas densities , in Fig. 111.5. The fraction of stars formed in spiral arms ~ssa/(~sa+~za)=e-x)(R)/[l+e'~)(R)]

is given in Table 111.4 for selected galactic regions and

R=0.5,1,2. Table 111.4:

Star Formation in Spiral Arms, assuming constant IMF

Fraction psa/(psa+~xa)

~sa(4-6 kpc) 3-10 kpc

m@ yr -I

R = 4-6 kpc

]0 kpc

~=0.5

0.60

0.33

0.55

3.5

1

0.75

0.50

0.70

4.3

2

0.85

0.67

0.80

4.9

5

0.95

0.83

0.90

5.5

In the last column of Table 111.4 we give the star formation rate in spiral arms for the two annuli with the highest star formation rate. Obviously,

the star formation rate in spiral arms

~sa(R=4-6 kpc) =3.5-4.9 m@ yr -I can not be higher than the mass flow of molecular gas through the spiral arms, 2(~R-~p)RAR~g %6.5-4.5 m@ yr-l,so that the high SFR is in conflict with our pre sent knowledge regarding gas consumption in spiral arms and the efficiency of star formation. The derived present-day SFRs, ~(t o) and ~(to,R), time-averaged

should be compared with the corresponding

SFRs

<~(to'R)>

-

1 t o

27 (l-r)

Rout f ~.(R)RdR R. inn

(111.15)

Here r is the fraction of matter returned to the ISM by a generation of newly formed stars. We adopt r = 0.4. Substitution of Tdisk=to ~1.2 101° yr for the galactic disk and 8 101° m® for the total stellar mass between R=2 and 13 kpc yields a time-averaged

galactic

SFR <~(t)> ~II m® yr -I. Taken at face value the result

~(to)/<~(t)>~1.2

would mean that the SFR in the galactic disk

was rather constant during the lifetime of the Galaxy. In comparing this result with the predictions of the simple closed system model for k=], Eq. (I.|O) yields a ratio of present-day to time-averaged N=O.O4,

8FR which depends on the

the value appropriate

gas-to-totalmass

for the galactic disk, yields

ratio ~ only. Substitution of

~(to)/<~(to)>~0.13,

factor of ten lower than the above ratio based on observationally duction rates and a constant IMF. Similar discrepancies

which is a

determined Lyc photon pro-

hold for predicted and observed SFRs

188

R. G~sten and P.G. Mezger

for individual galactocentric annuli, as shown by the curves ~ and

in Figs. lll.Sa pre and b. Moreover, the functional dependences on galactocentric distance R of predicted and observed SFRs is quite different. While the observed SFR attains a very steep peak between R=4 and 6 kpc the predicted SFR is flat and nearly independent on R. How accurate are SFRs based on estimates of Lyc photon production rates? We can try to answer this question by comparing SFRs estimated by our method with SFRs estimated by other independent methods. It is therefore of interest to note that the star formation rate derived by us for the annulus between R=9-]O kpc assuming azimuthal symmetry, ~(9-]O) ~0.5 m e yr -I, is rather similar to an independent estimate by Miller and Scalo (]978), who derived for the solar vicinity ~=(3-7).]O -9 m e yr -1 pc -2, corresponding to ~(I0)=.2-.45 m e yr -I.

111.4

DIFFUSE INFRARED EMISSION AND ITS RELATION TO STAR FORMATION

From the Lyc limit at h=O.O9 ~m to the sub~m region the Galaxy emits a diffuse radiation. From 0.09 ~m to 8 ~m this diffuse emission is dominated by direct stellar radiation; at longer wavelengths, the diffuse emission comes primarily from dust grains which are heated by absorption of stellar radiation. Fig. 111.6, from Mezger (1982), shows the overall spectrum of the Galaxy from the radio range through the UV as seen by an extragalactic 'observer'. The integrated flux density of the diffuse free-free emission is always less than the flux density of the diffuse synchrotron emission, which makes it so difficult to reliably

I

I

I

0"88

I

I

C÷157 0,o63 I II

1

<

-

i",..4Yn

10-4[ lcm

CO610

~/~

/I / / \d~

I lmm

I lO01Jm

"

\

I lOtJm

I llJm k

I 0.11Jm

Fi~. III. 6: The composite continuum spectrum of our Galaxy from the radio to the optical region. The position of the four strongest atomic and ionic lines is indicated. "syn" refers to synchrotron emission, emitted by relativistic electrons gyrating in the galactic magnetic field, "f-f" refers to the free-free emission from galactic ELD HII regions. The dust continuum is composed of contributions from cold dust associated with quiescent molecular clouds (dl) which probably also emit most of the C°610#m line; medium warm dust associated with diffuse atomic hydrogen (d2), which emits the C+157pm line and possibly (i.e. the hot intercloud gas) the O°63pm line; warm dust associated with primarily the ionized gas in ELD HI.[ regions (d3), which is expected to emit the 0++88~m line, too; and hot dust associated with circumstellar shells (d~). At wavelengths <~Spm the stellar radiation dominates the spectrum (from Mezger, 1982).

Star Formation and Abundance Gradients

189

separate the two components. In the mm-range the thermal radiation from dust, heated by stellar radiation, begins to dominate. The stellar spectrum peaks at wavelengths of ~1 ~m. The near IR emission is dominated by the radiation from cool stars (Tef f ~3 0OO K). In the following we discuss the diffuse galactic FIR and NIR emission. While the far IR radiation of the Galaxy seems to be a useful tracer of the OB-star distribution and the presentday SFR, the near IR emission comes from stars now evolving as red giants (with lower-mass progenitors) and thus provides some insight into the history of star formation in the galactic disk. 111.4.1 THE THERMAL FAR INFRARED (FIR) EMISSION FROM DUST The origin of the diffuse galactic sub,m/FIR emission has been discussed in two papers by Mezger et al. (1982) and Mathis et al. (1983). An extrapolation of their results to the FIR emission from external galaxies can be found in Mezger (1982). In normal spiral galaxies we deal primarily with medium warm dust (labelled component d2 in Fig. 111.6) and with warm dust (component ds). Medium warm dust is associated with the diffuse atomic intercloud gas; the dust is heated by the general interstellar radiation field (ISRF). Warm dust is associated with ELD HII regions where most of the heating is provided by O stars. It is this latter dust component which is directly linked to the free-free emission emitted by ELD HII regions and thus allows an estimate of the SFR. Based on the IR-excess = LiR/N~ych~ ~ of our Galaxy, which expresses the IR luminosity of an HII region in terms of the energy available in form of Lyman alpha photons, Mezger (1982) derives a relation between the free-free flux density and the integrated FIR spectrum of dust component d3 ~°'IS (ff) = 1.1 10-16 fS (da) d~ FIR v with ~ in GHz, dv in Hz and S

(lll.16a)

in Jy. Integration of a typical 'ds'-spectrum relates the free-

free flux density to the maximum FIR flux density of component d3 ~°'IS (ff) = 5.5 IO-~Smax(d3)

(Ill.16b)

These relations have been derived using observational results obtained for our Galaxy. To show their usefulness, we apply them to an external galaxy, NGC 253, which is an Sc galaxy at a distance of 3.4 Mpc, with a very high activity of star formation within its central 500 pc. The maximum FIR flux density, at ~1OO ~m, is Sma x ~1OOO Jy. From Eq. (III.16b) follows ~°'ISv(ff) = 0.55 and, substituted in Eq. (III.2a)

N' =5.3 IO~3 s-I. Because of the known ' Lyc high star formation activity it is safe to assume that most of the dust is heated by O stars. We correct for the loss of Lyc photons due to absorption by dust grains or due to escape by writing in Eq. (III.2b) NLy c ~2 N'Lyc' which, substituted in Eq. (III.12) yields a SFR for the central 500 pc of NGC 253 of P(to) ~20 m@ yr -I, for a constant IMF with m ~O.I m , correlov sponding to twice the total present-day SFR of our Galaxy or about twenty tlmes the SFR in its central part. However, as shown by Rieke et al. (1980) for a star burst in M82, various observations suggest mlo w ~3.5 mo, a value which would decrease considerably the SFR in NGC253, i.ei the total mass of gas transformed into stars.

111.4.2

THE DIFFUSE NEAR INFRARED (NIR) EI~SSION

The ridge line intensity 1(£,6=O o) of the diffuse galactic emission at 2.4 ~m as observed by Oda et al. (1979) and Ito et al. (i977) is shown in Fig. III.7. While it resembles the ridge line intensities of both diffuse galactic free-free and FIR emission its interpretation in ~ A

26:a

- C

190

R. Gusten and P.G. Mezger

I 141

I

l

I

I

I

I

I

IT o,. . . . , Ito

et

al

~o e

!

e=6

t

t i

t

+r
20

30

_,__

40

50

GA[. LONGITUDES(DEG)

+ 60

70

BO

Fi~. III. 7: The ridge line intensity at 2.4 ~m as observed by Oda et al. (1979) and Ito et al. (1977) with 0~6x0~6 and 2°x2 ° square beams. Curves and shaded regions refer to model computations with different distributions of the volume emissivity E2.~um in the galactic plane and the extinction model given by Eq. (III.17) and shown in Fig. ~II. 8. The solid curve for £ ~20 ° relates to a scalin~ of the 2.4 pm volume emissivity observed in the vicinity of the sun with Innanen's mass model. The solid curve for £ <10 ° comes from a model by Serra et al. (1980) which also includes contributions from stars in the nuclear bulge. The shaded area refers to our model, where the excees ridge line intensity between £ ~5~ and 50 ° is provided by evolved stars (M giants) with masses 3 ~m/m@ %11, whose distribution follows that of 0 stars and whose total luminosity is consistent with the present-day SFR. The upper envelope of the hatched area is obtained from a slightly modified extinction model, which takes the patchiness of the CO-distribution into account. Giving less weight to the molecular gas component Eq. III. 17 is changed to 2.9 nHI.Z(R)/Z(IO) + O.07.E(CO). terms of volume emissivity as a function of

.~

1,0



position in the galactic plane is complicated by extinction due to interstellar dust. Considering the contribution of dust associated with both the diffuse ISM and with molecular clouds, respectively, we use the relation

'

2.0

'

I

'

, •

T2.&pm=1.6~ ~ " - . - . . . , . . ~

.G.C.

/,"

kv(R)

-

= 1.875 nHI(R)Z(R)/Z(IO)+O. I1ER(12CO) (iii.17)

1.2\\,

"~~ ' ~

_

between visual extinction k v i n magnitudes per kpc, mean density of atomic hydrogen nHl in cm -3 and 12C0 emissivity g in K km s-1 kpc -I.

\

\--

OA

This relation assumes that the dust-to-CO -

--

-

SUN

ratio is independent on the gas metallicity, while the dust-to-HI ratio increases as md/mg =Z(R). Substitution of nHl and E(I~CO) from Table 11.2 yields an extinction at 2 . 4 ~ m

Fi~. III. 8: Curves of constant extinction optical depth, T 2 ~,_ in the galactic plane, as seen from the posztzon of the sun. Computed with Eq. (III.17) and T2.~m=Tv/13.6. •

~

in the galactic plane as shown in Fig. 111.8.



To convert visual extinction into extinction at 2.4 ~m we use the relation T2.~=Tv/13.6.

Star Formation and Abundance Gradients

191

The diffuse galactic NIR emission in the solar neighbourhood is mainly attributed to normal M-type giants (Tef f ~3000 K) with low-mass progenitors m ~<1-2 m®

(Hayakawa et al., 1978).

Scaling their emissivity at 2.4 pm with the galactic stellar mass distribution (e.g. Innanen (]973)), and considering the extinction in the galactic plane according to Eq. (111.17) and Fig. 111.8, yields the ridge line intensity shown as solid curve in Fig. 111.7. This predicted curve fits the observations only for galactic longitudes larger than 50 ° and less than 5 ° , but fails completely to account for the observed ridge line intensity between these longitudes. This fact was already noted by Maihara et al. (1978) and Hayakawa et al. (1977; 1978), who suggested the existence of an excess-population of M type giants and supergiants, distributed in a ring between galactic radii 4
~I0 s yr ago - was not much different from the present-day SFR.

With Tg the time which a star of mass m spends in the red giant (g) phase (Tg ~0.1 TMS) and ik(m,Tg)

the corresponding luminosity at ~=2.4 lJm (referred to by the letter k), the present

2.4 ~m luminosity per stellar mass is TMS-Tg Lk(m)dm = dm f #(m)~(to-t)Ik(m,t)dt , J

TMS

or, with Tg <
Lk(m)dm = dm ~(m)~(to-TMS(m))/~(m,t)dt

One sees immediately that the 2.4 ~m luminosity per unit stellar mass depends on (i) the galactic history, which determines the number of stars now evolving along the red giant branch, (ii) the "slope" of the mass spectrum, which may alter the contributions from different populations (e.g. giant versus supergiants) (iii) Lk(m) =Tflk(m,t)dt, the emission of a star of mass m, integrated along its post-main sequence

g evolutionary track.

The 2.4 ~m luminosity contained in the excess component (i.e. the difference between full and hatched curve) in Fig. 111.7 is ~109 L 0 ~m -I , assuming radial symmetry in the stellar distribution. According to Becker's (1982) closed evolution tracks from the zero-age main sequence up to the asymptotic giant branch, the post-main sequence integrated 2.4 pm luminosity Lk(m) is roughly independent on stellar mass (i.e. within a factor of two) for 3
192

R. G~sten and P.G. Mezger

Lk(m) =

/

ik(m,T)dT ~IO 9 L@ ~m -I yr.

post-MS

The integrated luminosity of a given stellar population is then determined mainly by #(m)dm, the number of stars in the mass intervall (m,m+dm). From Eq. (III.|O) follows

[~] Lk

~60 =/Lk(m)dm =2.5.]OS.~(to). (m~ I"" -0.02)

(iii. 18)

me

for I Smc S I0 m e . With ~ ~13 me yr -I and mc=2(3) me the above equation yields L k ~10(6)

]Os L® ~m -I and a total luminosity of L(red giants) ~6(4) ]09 L@, if these red

giants emit like a 3000 K blackbody. This means that a "medium mass" red giant population (i.e. a population of red giants which have evolved from MS stars with masses m ~mc=2(3) m@), whose SFR about 109 ago was similar to the present-day SFR, can account for the luminosity of the 2.4 ~m excess ridgeline intensity shown in Fig. 111.7. Moreover, and as shown in this Figure by the shaded curve, such a population of medium mass red giants gives a reasonable fit to the observed 2.4 ~m ridgeline intensity, too, if i) the radial distribution of the medium mass red giants in the galactic plane resembles that of the present-day 0 stars; and ii) the 2.4 ~m emission from this medium mass red giants is superimposed on a contribution due to old disk population M giants (shown in Fig. 111.7 as (interrupted) solid curve).

IV.

ABUNDANCE GRADIENTS IN THE GALAXY

Abundance gradients were first detected in nearby spiral galaxies from optical spectroscopy of HII regions (see Pagel and Edmunds (1981) for recent review). Naturally, due to extinction by interstellar dust, such observations are much more difficult to make for HII regions located in the galactic plane. Therefore attempts to establish galactic abundance gradients in the galactic plane by means of optical spectroscopy are limited to "local" sources. Observations of radio recombination lines, which are of relevance for the determination of He and O abundance gradients, do not suffer from extinction and therefore allow surveys of the entire galactic plane. In the future, FIR fine structure lines will certainly play an important role in determining galactic abundance gradients. The main purpose of this paper is the development of a model for the chemical evolution of the Galaxy, which can explain the evolution of abundance gradients in the galactic disk within the constraints of gas-to-total mass ratios and present-day SFR's as derived by radioastronomical observations and discussed in the previous sections. Early work related to both optical and radio observations has been reviewed by Peimbert (1978). The most recent data on galactic abundance gradients are giyen by Shaver et al. (]983). In this section we discuss briefly the galactic abundances of tHe and leO as examples of primary elements, and the gradients of isotopic abundances of H, C and O.

Star Formation and Abundance Gradients

IV.l

193

ABUNDANCES IN THE INTERSTELLAR MATTER

Results of observationally determined number abundances and abundance gradients (excluding the inner 3 kpc region) are compiled in Table IV.I, which otherwise should be self-explanatory. In the following we discuss briefly some problems related to observations of abundance gradients.

IV.].]

ABUNDANCE GRADIENTS OF aH AND tHe

From our present knowledge of nuclear synthesis all aH and most of the tHe have been synthesized in the big bang about three minutes after the expansion of the universe has started. Since the fusion of IH into tHe is the main source of energy of stars, we expect an enrichment of the ISM in tHe as a consequence of the chemical evolution of the Galaxy. At any position in the galactic disk the total He abundance (in mass fraction) is

Y(~He) = Y

p

+ bY

ev

(IV.I)

with Y

the primordial abundance and AY the enrichment of the ISM due to chemical evolution. p ev Reasonable estimates of the primordial He production fall in the range Y ~0.22-0.25. From P the estimated abundance in the solar vicinity, Y1o ~0.28, follows AY ~0.06-0.03. ev First, due to the fact that AYev is much smaller than YP , determination of the variation in the stellar contribution to the present helium abundance requires an accuracy in the measure-

ment of the total abundance better than 0.005 (abundance by number). Secondly, in the HII regions, which are used for most He abundance determinations, He is in part neutral and in part singly ionized. Both radio and optical spectroscopy yield, however, only the ionized He abundance, while the fraction of neutral He has to be estimated indirectly. It has been found that (probably as a result of an increase in the metal abundance) the fraction of ionized helium in HII regions decreases towards the galactic center. At galactocentric distances R ~8 kpc, this latter effect begins to dominate over a possible increase in the total He abundance as discussed in Mezger (1981) and Mezger and Wink (1983). In Table IV.l and subsequent discussions, ~e use the the abundance gradient as derived by Thum et al. (1980) for 8 SR ~12 kpc. Deuterium (2H), on the other hand, is a very weakly bound nucleus and is destroyed if exposed to temperatures of ~IO e K for sufficiently long times. One expects therefore that all ~H, which has been circulated with the ISM through stars ("astrated") since the formation of the Galaxy has been destroyed. Since there is no astrophysical site known (other than the big bang, of course), where 2H could be produced in substantial quantity, one would expect that the depletion of the ISM in =H is proportiQnal to the enrichment in tHe and heavier elements. Hence, a SH gradient should have a different sign from that of the primary elements. The interstellar deuterium abundance is determined locally only by UV-observations of the diffuse gas. Due to the strong chemical fractionation of the deuterated species in cold molecular clouds, no quantitative D/H abundance ratio can be determined from molecular line studies. Recently, Vidal-Madjar et al. (1983) re-interpreted DI UV-absorption lines toward nearby luminous stars, and concluded that (due to contamination of the Dl-line with blueshifted circumstellar HI-lines) the local interstellar D/H abundance is probably ~5 I0-e, rather than ~2 IO-s as previously discussed.

194

IV.I.2

Ro G~sten and P.G. Mezger

ABUNDANCE GRADIENTS OF 160 AND OTHER ABUNDANT PRIMARY ELEMENTS

One of the basic problems in determining absolute abundances of the elements of the C,N,0 group in the ISM is posed by depletion. It is estimated that about l/4 to l/2 of the total mass fraction of these elements is used to form dust grains. While the cores of these dust grains appear to be made primarily of graphite and silicates, mantles of molecular ices tend to form on these cores within molecular clouds. As an effect of chemical enrichment during galactic evolution one would expect that the abundances of C,N,O in the ISM in the solar vicinity should be higher than the corresponding abundances in the solar system (which decoupled from the ISM about 5 lO 9 years ago). The fact that the observations compiled in Table IV.l show the reverse trend is probably due to depletion. Large-scale oxygen abundance variations have been determined from gradients in electron temperatures in HII regions, measured by radio recombination lines on the assumption that oxygen is the main coolant of the plasma (see e.g. the review by Churchwell, 1980). A strong increase in the metallicity of the ISM with steepening gradients towards the inner Galaxy is implied by these studies (Mezger et al. 1979). A recent combined optical and radio-based study of Hll-regions confirms these results, with

dlog(0/H)/dR~O.07(±O.OI5) dex kpc -1 for R >7 kpc,

by Shaver et al. (1983). The gradient in I~N/H, found in the latter investigation, appears to be considerably steeper. Due to the lack of suitable transitions in the optical range, there is little information on the large-scale distribution of the carbon abundance.

IV.I.3

ISOTOPIC ABUNDANCE GRADIENTS

Relat~ue,

isotopic abundances can be inferred from molecular line studies of the neutral gas

component. However, although probably not affected by ("selective") depletion onto dust grains, the line formation process depends strongly on the physics of the molecular cloud. Line saturation and chemical fractionatlon necessitates sometimes large corrections to the observed line ratios.

Observations of carbon and oxygen isotope ratios, as inferred from several molecular species (mainly CO and H2CO) have been reviewed recently by Penzias (]980) and Wannier (1980). Since in particular the CO line formation suffers from optical depth and fractionation effects, in the following we compare our model predictions with [12C/1sC] ratios obtained from H~CO. The formaldehyde absorption lines mainly trace higher density regimes of molecular clouds, where fractionation may be less effective. The centimeter lines of H2C0 have relatively small optical depths, which (in principle) allows a reliable derivation of the column densities. For a discussion of uncertainties (mainly due to photon trapping in the optically thick mm lines) see Henkel et al. (1980, 1983). According to these data, the mean

[12C/13C]

abundance ratio

in the local disk is 70-75, only slightly below the solar system value of 89. Excluding the galactic center sources, a small gradient in [12c/1Sc] is indicated. In any case this gradient is smaller than the large-scale variations in 160 and I~N. Due to the weakness of the rare isotope line, very few sources have been measured so far in the H2clS0 line, yielding a somewhat lower than solar system [160/lS0] abundance ratio. A positive galactic gradient is suggested by the low galactic center value. [16c/lS0] ratios based on CO observations (namely the double-isotope line ratio [12C180/13C160]) are inconclusive, as it is difficult to correct for carbon fractionation and saturation effects.

(2)

4.2(~;~).I0 -4

References:

I~N/IS N

(6) (7) (8) (9) (10) (11)

Henkel eL a i . (1979) P e n z l a s (1981) Wannier eL a l . (1981) P e n z i a s (1980) B i n e t t e eL a l . (1982) Vidal-Madjar eL a l . (1983)

saturation, fractionaL±on, use of double isotope ratios (HCN) molecular (HCN)

(14)

>(500)

(12) (13) (14) (15) (16) (17)

unobserved ionisation states, depletion shock excitation model (SNR)

optical ([NIl])

(11) (16)

?

molecular (H2CO,CO)

inferred indirectly only from T e depletion onto grains

r a d i o recomb. (HII) optical ([O11],[O111])

molecular (CO)

(12)

(1o) (11)

(9)

saturation, fractlonation

depletion onto grains, unobserved ionisatlon states

optical (ell)

(8)

m o l e c u l a r (CO,H2CO)

correction for unobserved neutral Hel contamination from PN-progenitor

r a d i o recomb. ( H e l l ) optlcal (Hell)

(8)

(7)

(6,15) (17)

strong chemical fraetionation blending with blueshifted circumstellar lines

Difficulties in Interpretation

molecular (HCN,DCN) atomic (H,D)

Observed Lines

3.5(±0.2) (13)

15o(±50)

?

20±5

?

?

?

Galactic Center

Bruston eL a l . (1981) Thum eL a l . (1980) Pelmbert (1978) Henkel e t a l . (1982) Mezger eL a l . (1979) Shaver eL a l . (]983)

(positive)

4(300)

(I)

-O.O9(±O.O15) -0.095(±0.03)

4-IO -s

(3) (2)

_<0.01

3.7

positive

- 0 08 _O[O7 (±0.015)

5 ( ± 1 ) . 1 0 -~

~3OO

<0.05

75(±5)

(i)

(I)

(3) (2)

(I)

Cameron (1980) Ross + Aller (1976) Lambert (1978) Hirayama (1971), Heasley + Milkey (1978) Kunde eL a l . (1982)

272

l.O(t.2).lO -~ . 9 ( ± . 3 ) . 1 0 -~

5.4

lSO/lV O

~NIH

490±25

8.3(+2).10 -~ 6.9(±1.3).10 -4

1sO~leO

160/H

12C/13c

(I) (2) (3) (4) (5)

(3.10-")

(3)

4.7(±1.I),10 -~

I~C/H

89±2

-0.015 -0.020

~0. I0

(4)

0.05-0.10

~He/H


positive

~2 10-s 5 10-6

(5)

3.6(~:~).I0 -s

2H/H

Disk-Gradient [dex kpc -1 ]

[OMPILATION OF SOLAR SYSTEM AND INTERSTELLAR ABUNDAN[ES

Local ISM

Isotopic Species

Solar System

TABLE I V . 1

Ln

e~

,.~ o

196

R. G~sten and P.G. Mezger

This is different for the determination of the [ISo/IVO] ratio, which according to singleisotope measurements in the optically thin 12C17(Ia) 0 lines is significantly below the solar system value and is rather constant over the galactic disk (R ~6 kpc). The intrinsic scatter in these data is small

compared with carbon-isotope measurements. This emphasizes the im-

portance of carbQn fractionation in the CO lines; for the oxygen isotopes no such fractionation process is expected (Watson,

IV.2

1979).

THE SOLAR sYSTEM ABUNDANCE AND ITS RELEVANCE FOR GALACTIC EVOLUTION MODELS

For comparison in Table IV.I, the composition of the pre-solar nebula is given. The abundances of the C,N,O-group are derived from measurements of the solar photosphere and of quiet prominences. Their isotopic ratios are from terrestrial or from meteoritic samples. The C] carbonaceous chrondrites are believed to be representative for the early solar system material (Meyer, ]979). There is considerable uncertainty in the photospheric abundances, mainly due to uncertain oscillator strengths and deviations from LTE. A reliable deuterium abundance is difficult to derive from the solar atmosphere (astration) or from terrestrial determinations (gravitational fractionation). In the Table, we give the value for D/H determined recently in Jupiter's atmosphere. Are the solar system abundances representative of the ISM composition 4.6 Gyrs ago, the time when the solar system decoupled from the interstellar gas? This question relates to the homogeneity of both the solar system material and the general ISM. Despite its well-known heterogeneities, as observed e.g. in the oxygen isotope ratios in various types of meteoretic material (Wasserburg et al. ]979), within a few percent the relevant composition of the primitive solar nebula seems rather homogeneous and well-mixed. In our chemical enrichment models of the Galaxy we shall thus assume that the bulk composition of the solar system is representative for the composition of the ISM 4.6 Gyrs ago. A matter of controversy is still the question of whether there exist chemical inhomogeneities in the interstellar medium, i.e. if element abundances vary from cloud to cloud

and from star to star. There is ~25% disper-

sion in the metallicity distribution of stars with the same age (Twarog, ]980), and obviously there are source-to-source variations in the molecular abundance between CO and H2CO and their isotope substituted species. But the latter may well be attributed to errors in correcting for line saturation and chemical fractionation. In the following

we assume that within a galactic

annulus R, R+AR the ISM is well mixed. (For a qualitative discussion of elemental mixing in the ISM see Chevalier (]979)).

V. A SELF-CONSISTENT MODEL OF STAR FORMATION AND CHEMICAL EVOLUTION OF THE GALACTIC DISK In Sect. I. we compared predictions of the closed system model with constant IMF with observations related to both element abundances and SFRs. We found that this model yields reasonable agreement with observations if applied to abundances in the solar system and the ISM in the solar vicinity. Severe disagreement between this model and observations arises, however, if the predictions are compared with observations related to: i) the radial element abundance distribution in the galactic plane (Sect. IV); ii) present-day SFRs and especially its radial dependence in the galactic plane compared to the mass distribution (Sect. 111.3.3); iii) the time evolution of abundances. Several attempts have been made to construct models of the chemical evolution of the Galaxy in order to explain the observed abundance gradients. These models are, to our knowledge, not

Star Formation and Abundance Gradients

197

consistent with the observed SFRs, net SFRs, mass distributions and abundance gradients. In this section we present a model, which we believe, avoids these difficulties.

V.l

MODIFICATIONS OF THE CLOSED SYSTEM MODEL

We introduce two modifications viz. a variable yield and a slow build-up of the galactic disk by accretion of halo gas (which leads to the well-known class of infall models).

V.l.l

A YIELD WHICH VARIES WITH GALACTOCENTRIC DISTANCE

The existence of a galactic abundance gradient in the C,N,O elements is well established (Sect. IV). However, as mentioned in Sect. 1.3 and shown in Fig. V.5, the closed system model fails to explain these gradients if the gas-to-total mass ratio, ~(R), from Table 11.2 is substituted in Eq. (1.4). As discussed in Sect. II there are relatively large uncertainties in the estimated molecular abundances, and other investigations yield higher densities . This tends to flatten ~(R) and hence the predicted metal enhancement in the inner disk would even be decreased. There is no way that closed system models with constant yield can produce the observed abundance gradients. Can infall models explain these gradients? Tinsley and Larson (1978) have shown that infall models ~an produce a modest gradient of primary element abundances during periods in the evolution of galactic disks. This case, however, requires high accretion rates which must vary strongly with galactocentric distance. In systems in highly evolved (saturated) stages with reduced infall rates, these gradients tend to flatten. If applied to the present-day state of evolution of the galactic disk, with small accretion rates and a rather constant degree of gas depletion ~(R), these models can explain neither the observed abundance gradients nor the SFRs as derived for constant IMF in Sect. III. But nevertheless any realistic attempt to follow the galactic enrichment with time has to incorporate the dynamical history of the disk, with its slow build-up by accretion of settling halo gas (see below). Therefore we include in the following discussion of the chemical evolution of the galactic disk infall of halo gas, too. We want to reiterate, however, that it is not the infall which is used to explain the observed abundance gradients across the disk. The only way to account for e.g. the strong leo gradient is to introduce a variable yield Yi(R) which increases towards the inner part of the disk. To make this modification more transparent, we discuss its consequences first in terms of the simple closed system model. The yield used in Eq. (1.4) is given by (Eq. C.l) as yi=Pi/(l-r), with m up Pi = f $(m)Am(Zi)dm

(V. la)

mlow the fraction of mass of a newly synthezised element Z. returned by a stellar generation to i the ISM, and m up r

=

f

m ~(m)(m-mf)dm

=

~P~(m)r(m)dm

(V. lb)

the dimensionless mass fraction returned by a stellar generation to the ISM, with mf the final mass of the stellar remnant, which can be a white dwarf, a neutron star or, possibly, a black hole. r(m)=(m-mf)

is the mass returned by a dying star of mass m to the ISM.

198

R. Gusten and P.G. Mezger

Results derived in the previous

sections were based on the assumption of the universality

the local IMF, which has been derived from star counts in the solar vicinity. ever, a number of observational hypothesis,

indications which have led Mezger and Smith (1977) to the

that medium and high mass stars form predominantly

mass stars form predominantly

of

There are, how-

in the interarm region.

lished which supports this hypothesis.

in spiral arms, while low-

Since then further work has been pub-

Miller and Scalo

(1978) conclude that in OB associa-

tions (which are the later stages of giant radio HII regions and, like these, are well-known spiral arm tracers)

probably stars with m 52-5 m O form only. A deficiency of low-mass

there is found by Eggen (1977) and Larson

(1982). T-Tauri stars and Herbig-Haro

are known as tracers of low-mass star formation. in T-Tauri associations,

as witnessed

by the presence of early B stars in the 0-Ophiuchus

sites of formation of low-, intermediate-

cloud.

and high-mass

of the different

In fact, there may be different stars and the local IMF would then

star formation processes.

the details we try as a first and certainly crude approximation scenario of star formation a quantitative

formulation

to in the following as hi-modal star formation:

Without knowing

to this rather complicated

of the above hypothesis, which we refer

There are two different

which describe the integrated star formation processes respectively.

objects

Only a few medium mass stars appear to form

cloud or even some 0 stars in the Orion molecular

be the result of a superposition

stars

IMFs, ~sa and ~ia,

in spiral arms and interarm regions,

We assume that both IMFs have the same slope (for which we adopt the Miller-

Scalo IMF as given in Sect. III.3.1)with a lo~mmass cut-off for the interarm region IMF at mlow=O.l m®. The deficiency of low-mass

stars in spiral arm star formation

by an increased low-mass cut-off of the IMF, mlow--mc >] m®. Some theoretical tend to support the picture of bi-modal therein).

Obviously,

spiral arms and in the interarm region. et al.

star formation

the physical conditions

(]980), in an observational

is described

investigations

(see e.g. Silk (1980) and references

in star forming clouds are quite different

in

In this context it is of interest to note that Rieke

investigation

of the physical

conditions of a star forma-

tion burst in the center of the galaxy M82, find that the IMF has to be truncated at mlow~3.5 m@ in order to match both the total IR luminosity,

the Lyc photon production rate

and the absolute magnitude at 2.2 ~m. It thus appears to be a characteristics like" star formation

of "burst-

(and as "burst-like" we can also consider the increased SFR in spiral

arms) that it produces primarily medium and high mass stars. The net effect of bi-modal low-mass

stars relative

hence results

star formation is an increase of the ratio of massive

stars to

to the local IMF, which increases the effective bulk yield Yi and

in a higher enrichment

of the ISM. Since the ratio of massive

spiral arms to massive stars formed in the interarm regions

stars formed in

increases with decreasing galac-

tocentrlc distance R, the effective bulk yield Yi(R) increases with decreasing R, too. In this way abundance gradients will form even if the gas-to-total increases

mass ratio ~(R) is constant or

slightly towards the galactic center. A quantitative

yield due to bi-model

star formation

description of the variable

is given in Sect. V.2.3. Here we want to show in rather

a qualitative way why the bulk yield Pi Yi = 1----~

increases with increasing synthesized

(V. 2)

low-mass cut-off of the spiral arm IMF. P i, the fraction of newly

element Z.I ejected into the ISM will not be affected until mlo w approaches

Stars with lower mass have MS lifetimes do not contribute

~I m®.

exceeding the age of the galactic disk and therefore

to the chemical evolution of the ISM. Fo r mlo w >I m®, as can be seen from

Fig. C.I, the production of secondary elements

like ISC and I~N will first be reduced. With

Star Formation and Abundance Gradients

199

increasing mlow, Pi of primary elements will be reduced, too. (l-r), the fraction of lockedup matter, on the other hand will steadily decrease With increasing mlo w. Evaluation of Eq. (V. lb) yields for mlow=O.l,

I and 5 m® the corresponding fractions (l-r)=0.6 , 0.23 and 0.13,

respectively. The net effect on the bulk yield (Eq. V.2) is that Yi rises sharply with increasing mlo w for primary elements from massive progenitors, but tends to flatten out for secondary elements provided predominantly by medium mass stars with m !5 m®, once mlo w >] mQ. Another desirable effect of bi-modal star formation is an increase of the ratio of Lyc photon production rate NLy c to gas consumption rate ~ for a new generation of stars. This ratio NLyc/~=~(1)P(#)

is given in Eqs.(A.3a and b) as a function of mlo w. An increase of this com-

puted quantity means that for an observationally determined Lyc photon production rate both the SFR and the net SFR (i.e. the mass of gas converted and permanently locked-up stars) is decreased. This alleviates the problem of the high SFRs derived in 111.3.3 assuming a constant IMF.

V.l.2

A REALISTIC MODEL OF THE GRADUAL BUILD-UP OF THE GALACTIC DISK BY ACCRETION OF HALO GAS

Though fragmentary, our knowledge of the history of the early Galaxy~ as obtained from the analysis of older population objects (globular clusters of the halo-, and open cluster of the diskpopulation) yield strong chronological constraints on the early dynamical and chemical evolution. Despite rather large uncertainties, nearby globular clusters(6 = o~/tdisk ~8 mOPC ~= Gyr-I, as compared to a present-day rate A(t o) ~2 m®pc -a Gyr-1.This is derived from a study of highvelocity HI clouds (Oort, 1970). But even this value is disputed since these high-velocity clouds may not relate to halo gas at all, but rather represent cooled and condensed debries of hot coronal disk matter (see Sect. II.I.I), which has been ejected into the galactic halo as suggested by the galactic "fountain model" (see York (1982) and references therein). Based on X-ray observations of the galactic corona Cox and Smith (1974) derived a present-day accretion rate which is about ten times lower than the above value. The infall pattern, or more precisely the ratio between net SFR and accretion rate

k(t) = (]-r)~(t)/A(t)

(v.3)

determines the main characteristics of infall models. Owing to the weak observational constraints on k there exists in the literature a wide range of solutions. These are summarized in Table V.I.

200

R. Gusten and P.G. Mezger

TABLE V.1

CLASSESOF INFALL MODELS Cheracteristic~ of model

~ccretiou £ 1 ~

Ys

A~

mate1 ~ r i ~ t

e

esyaptotic solution

u*O Xscc=to

Z'Yi [ I - e (1-~°*)]

Larson (1972)

constant

TwerOK (1980)

(y,A)conit,

slow, still c o n t i n u i n 8 accretion

Tosl (1982)

Tact=t°/2

slow, but l a r s e l y completed disk formation process

Zt(Yi.k÷~)" IO

$

e~p. decrease xeec~5 G~rs

~3

~2

Gdst~÷Hezger (1983)

uon~xp, decrease kl
~5

e.g. Tinsiey (1980)

closed system ("ei~le model'*)

Vader+deJong (1981~

Yi'~+Zh (Zh'O)

1.6(±5).y i

Chiosi (1980)

! &cc kl-O

rapid disk formation

~.2

Equstion (V.ll)

yi[l+ln ~]+K.zh

Z=Zo*yi ln~-~

(S. Yi)

t f o r primary non-depleted nttclel like leO. ~scc " tlmescsle of disk formation by accretion o£ halo p s , t o - "age" of the disk, ~I(A,) present-day local 8tar f o ~ t i o n

(secretion rata) i n [ ~ pc -z GYm-I], ~ - ( ] - r ) ' ¥ / A

These models range from slow and still continuing accretion (Tac c ~tdisk) with the accretion time scale comparable to the age of the disk, to a rapid (instantaneous) formation (Tacc << 0.5 tdisk )which leads to the initial conditions of the closed system model discussed in Sect. I. Intermediate are models that take into account a slow accretion process which, however, is presently completed and which have typical accretion times of ~5 Gyrs (T

acc

%0.5 tdisk ).

Another model parameter which has to be determined from observations is Zh, the metallicity of the infalling halo gas. Models of neighbourhood (Fig. V.3 ) constrain

the age-metallicity relation of stars in the solar the pre-enrichment of the halo gas to Zh S0.2 Z O. Thus

we neglect this effect in discussing the present-day abundance gradients in the disk. In realistic infall models, ~(t) is time dependent. We are doubtful about the reality of infall models which assume both constant SFRs and constant k (and hence a constant infall rate), as used by Twarog (1980) and Tosi (1982). In our model, we use a convenient analytical description of the mass flow pattern first suggested by Lynden-Bell (1975) (Moo-M~) (M~+M o) Mg(t) =

(V.4) (No+M)

It

is based on the assumption, that in a typical accretion model the

gas mass in the disk, Mg(t), rises from an initial threshold mass M ° < O . 1 M formation begins, to a maximum

, for which star

and finally declines to zero, if the mass converted into stars,

M~, approaches Moo, the asymptotic mass of the disk

{MD=CMg+M~)+Mco}.

Combining Eq. (V.4) with a relation between SFR and available gas mass in a form similar to Eq. (1.2) yields a relation,

| 1-r

dM~ dt

~(t) = T] k ( t )

(V.5)

Star Formation

!

i

and Abundance

1,0

M~

/./,/

M.

A I

I

I

~

d N

d%

.3

/ .// ' - X -

2

"\

-

\'x

/"

.1

= which connects

~

.0

X

I

I

3

a

I

6

accretion with

star formation and hence determines

R

I

I

201

Fig. V.I: Application of Lynden-Bell's '~est Accretion Model" to the gradual build-up of the galactic disk. A linear dependence of the SFR on the gas mass i8 assumed (i.e. k=1 in Eq. V. 5). In the upper diagram are shown the relative variations with time of total disk mass (MD), matter transformed into stars (M~) and mass of gas (Ma). As dotted line is shown the present-da~ local gas-to-total mass ratio u(t o) = 0.04. In the lower diagram are shown accretion rate (d/dt)MD and net star formation rate (d/dt)M~. i8 the star formation efficiency. From the intersection point of v(t o) with Ma/ M and with an adopted age of the galaDt~c disk of to=12 Gyr we determine ~= 5 10 -I°. In the right hand corner of the lower diagram are shown observational estimates of the accretion rate (HVC stands for high velocity clouds; XR stands for X-ray observations). The shaded area refers to the SFR in the solar vicinity.

i

M,,,/ ."

Gradients

5

,

the

time-scale of the build-up of the disk.

,t

In Eq.

I

12 t ISYRS]

(V.5), ~ measures

tion efficiency";

AGE OF THE OISK

the "star forma-

k, the power law ex-

ponent may vary between

I and 2 (Sect.

III). For k=l Fig. V.! shows the variation of the main parameters Bell's as function of a normalized present gas-to-total tdisk=to=]2

time Nt.

of Lynden-

"Best Accretion Model"

(Eq. (V.4))

The dotted curve in the upper diagram refers to the

mass ratio N(to) of the local galactic disk. Adopting an age of the disk

Gyr we determine

from this diagram h=5 ]0 -I° yr -I . The model shows satisfactory

agreement with observed constraints

on the local SFR as well as the X-ray limit on the accre-

tion rate. The major peak in the SFR occurred only 6 Gyr ago. This is consistent with the mean age of the stars in the solar neighbourhood.

For k=2 the phase of major disk accretion and

star formation would shift toward earlier epoches,

but still yields a reasonable description

of observations.

V.2

THE BASIC SET OF EQUATIONS

V.2.!

DESCRIBING CHEMICAL EVOLUTION

A RIGOROUS FORMULATION

In Sect. I we gave relations and instantaneous

for the chemical

evolution of a closed system with constant

return rate. Here we drop these simplifying

and derive equations, which describe

assumptions

IMF

and approximations

the chemical evolution of a galaxy for the general case

with infall, delayed mass return from evolving stars and bi-modal

star formation.

The net

SFR is 60m 8 d d--~ M~(t) = ~(t) - f ~{t-T(m)} m(t=r)

with ~(m) the mass of stellar material

*(mlr(m)dm

that is re-ejected

(V.6)

into the ISM at the star's death,

and T(m) >TMS the total lifetime of the star (Fig. A.2). m(t=T)

is the mass of those stars

202

R. Gusten and P.G. Mezger

which reach the end of their evolution at t=T. In the following we assume that stars with MS masses m ~5-8 m 0 end as white dwarfs with typical final masses ~0.6 m O (K~ster and Weidmann, 1980) while the remnants of higher mass stars are neutron stars with ~1.4 m®. The he/ g ~

consumption rate is

ddt Mg(t) = A(t) - ~t M~(t)

(V.7)

In the general case A(t)=dMD/dt allows for any type of mass exchange; in the following A(t) represents the accretion rate of halo gas on the galactic disk. The net enrichment of the ISM with an element Z. is i

d d--{ [Zi (t)Mg(t)] = -Zi(t)P(t) + Zh(t)A(t) 60m0 + f P{t-T(m)} ~(m)~r(m)-d(m)}Zi{t-T(m)}+ m(t=T) =

(v.8)

Am(Zi~

for primary element i

{ AmP~i)

with Am(Zi) AmS(zj)'Zi(t-T (m))

for secondary elements j with seed elements i.

Here the superscrlpt'p' refers to primary elements and 's' refers to secondary elements, while the subscript h refers to the halo gas, i.e. Z h is the metal abundance by mass of the infalling halo gas. The material returned from an evolved star can be enriched in newly synthesized element i (Am(Zi)~)

but it can also be depleted if the element j has been destroyed in some of the

nuclear reaction chains in the stellar interior.

In this latter case the depleted mass is

d(m)'PZ.~t-T(m)} with Z~(t-T) the abundance of the ISM at the time of the star's formation. 3 As a secondary element requires for its synthesis a seed element j its production is =Z.(t-T). 3 V.2.2

THEINSTANTANEOUS

RECYCLING APPROXIMATION

The delay T between the formation of a star and its final evolutionary stage where it returns its outer shells to the ISM

complicates Equations (V.6 through 8) in such a way,that exact

solutions can be obtained numerically only. Neglecting this time delay leads to the "instantaneous recycling approximation" and to approximate analytical solutions which are often very useful.

In this approximation Eqs. (V.6 through 8) become

dM~

dt dMg dt

=

( l-r)V

(V.9a)

=

A-(l-r)~

(V.9b)

d d-~ (ZiMg) = Pi ~-(1-r)Zi~+ZhA

(V.9c)

with (l-r) the mass fraction locked-up permanently in low-mass stars and stellar remants and the return rate r, as defined by Eq. (V. Ib). Pi is given by Eq. (V. la). Without considering accretion (A=0),and ~(t=O)=l,this set of equations has the simple solutions given in Sect. I for the closed system model. Retaining accretion, A(t) leads, after substitution of Eq. (V.9b) into Eq. (V.ga), to a "tlme-independent" formulation of Eq. (V.9c)

Star Formation and Abundance Gradients

d d d--~ (Z Mg) = yi-Zi+Zh~-~

(MD)

203

(v. 10)

With the accretion law Eq. (V.4) and Zh = const, this equation has the solution

= (M°o~M°)

IMP+ (M+Mo) in( I-

]

M. ---~--o,

(W.l la)

In late phases of the evolution of the galactic disk, when infall has stopped, Mg÷Moo-M~ and Eq. (V. 1 la) approaches asymptotically

Z(M~)-~y i

[l+ln~a-1] +Z h

(V. llb)

In Fig. V.2 solutions using the instantaI

I

I

I

PRIMARY NON-OESTROYED ELEMENT ENRICHMENT

neous recycling approximation are compared with numerical solutions of Eqs. (V.6-8)

k--2

1.0

_~- .

in Eq. (V.5). Solutions shown in Fig. V.2 are normalized to present-day abundances

'l-S/

///

for two values of the powerlaw exponent k

(i.e. ~=O.O5, to=12 Gyr and adopting Zh=O)

j

and show the relative enrichment of the

w

galactic disk with a primary non-destroyed

.S

z

iI

#.

element (such as, e.g. 160) as a function

f~

of time. We see that the instantaneous re-

INSTANT. RECYCLING

cycling approximation first tends to underestimate the enrichment of the ISM. The

I 3

I I 6 9 AGE (GYRS)

t 12

reason is that in this approximation it is assumed, that all stars with mass m 51 m® reeject the mass r(m) instantaneously after their formation. In reality, however, a

Fig. V. 2: Gomparison of numerical (exact) and analytical solutions of Eq. (V. 11) for the enrichment of the ISM as a function of time with a primary non-depleted element. For the analytical solution the instantaneous recycling approximation is used. Parameter k relates to Eq. (V. 5). Computed abundances are normalized for their Values at to=12 Gyr.

significant fraction of gas is locked up in medium-mass stars during the early evolutionary phases of the galactic disk. Hence the rigorous solution first yields a higher enrichment of the ISM. Later, when low-mass stars begin to eject their enve-

lopes of virtually unprocessed matter to the ISM, further enrichment of the ISM proceeds much more slowly than in the instantaneous recycling model.

V.2.3

A VARIABLE YIELD AS A CONSEQUENCE OF BI-MODAL STAR FORmaTION

Bi-modal star formation and its effect on the yield have been discussed qualitatively in Sec. V.I.I. For an incorporation in our model of the chemical evolution of the galactic disk we derive a quantitative formulation of the yield Yi(R) as a function of galactocentric distance R by defining a new IMF,~, whose shape is that of the Miller-Scalo IMF ~ms, but whose lowmass cut-off is mlow=O.! m 8 in the interarm region and mlow=mc 51 m@ in the spiral arm region. The spiral arm IMF will be weighted with the factor av(R), which gives an adequate description

204

R. Gusten and P.G. Mezger

of the observed ratio of Lye photon production rates (and hence of the SFR of massive stars) in spiral arms and interarm regions, respectively.

(See Sect. 111.2.3).

~(R,mc,m) ffiN-I {~ia(m ~O.I me) + ~v(R)~Sa(m ~mc) }

(V.12)

with the normalization factor m up N(R,mc) = f @(R,m)mdm mlow

Note that ~ in (V. 12) is the normalized Miller-Scalo IMF as given in Eq. (lll.lOa,b). As previously we use mupffi60 m 8 in all numerical expressions. Substitution of Eq. (V.12) into Eq. (V. la) yields

P(R,m c) = N-I

60 60 1 f @(m)Am(Zi)dm + ev(R) f @(m)Am(Zi)dm 0.1 me

(v

+

Herein, the yields Pi are those derived for the normalized continuous or truncated Miller-Scalo

I I ~ , r e s p e c t i v e l y and given, f o r continuous I I ~ , in Table C . I . For primary elements i

with

massive progenitors m ~mc, p~a becomes independent on m e and hence pSa=piaffip. . . (Table C. I) i i i

(V.13b)

P(R'mc) = N-I Pi {I+~v(R)}

Substitution of Eq. (V.12) in Eq. (V. lb) yields the fraction of matter locked-up permanently in stars (I ~mc/m e !6)

l-Vt(R,mc) ffiN-I

f ~(m)mdm + 0.6 IO.l

~(m)dm + ~v(R) f ~(m)d mc

6o

+ ].4(l+otV(R)) f ~(m)dm 6

1

= N-I F(R,mc)

(v.14) r(R,mc) should not be confused with r(m) = m-mf used in Sect. V.2.1. Here 0.6 m e relates to the remnants of medium mass stars and 1.4 me relates to the remnants of high mass stars, respectively (see Sect. V.2.1). Substitution of Eqs. (Vl3a and 14) in Eq. (V.2) gives the effective yield as a function of R and m

Yi(R,mc) =

P(R'mc) ]_~(R,mc )

=

c

ia sa Pi +~v(R)Pi F(R,mc )

(V.15a)

which, with Eq. (V.]3b) and an analytical approximation of the numerically evaluated function F(R,mc) , becomes ]+~v(R) =

Yi(R'mc )

(V.]5b)

Pi

{0.6 +~v(R) [O.14m-I"~+O.O08]} e

Here, numerical values, of Pi can be taken from Table C. ]. For cxffiO(i.e. no star formation in spiral arms, ~=~bla) yi=Pi/O.6 , the relation used in Sect. I for a constant Miller-Scalo IMF.

Star Formation and Abundance Gradients

205

Table V.2: The variable yield Yi(R,mc) for the case of bimodal star formation (with e=1). Pi is the production efficiency, as calculated for a continuous standard IMF (Table C.]). Yi/Pi R/kpc

=~(R)

3

5

1

2.7

3.1

3.2

3.14

3.9

5.8

6.2

! .46

! .84

I .92

10

4.5

mc/m e = I

Yi(4.5; m c) =

Yi(10; m c) Values of Eq. (V.15b) are given in Table V.I, and have to be compared with yi(a=O)/Pi=(l-r) -I = 1.7, the value that applies for a constant (continuous) IMF. The modified yield increases with both ~ and m c , and relative abundance enhancements by factors of J2 between the inner disk and the solar vicinity are easily obtained, for m ~3 and a ~]. Further increasing these c parameters, however affects the yield only moderately, which indeed is rather insensitive on m

~3 and ~ ~1. C

V.3

PREDICTIONS OF THE MODEL OF BI-MODAL STAR FORMATION COMPARED ~ T H

OBSERVATIONS

As discussed in Sect. 1.3 the predictions of the closed system model with constant IMF (or constant yield, respectively) are in severe disagreement with observations related to 1) the chemical evolution of the ISM as a function of the age of the galactic disk; 2) abundance variations with galactocentric distance (abundance gradients); 3) SFRs and the stellar mass distribution.

In this section we compare predictions of the model of bi-modal star formation

and infall of halo gas with observations related to these three points and obtain good agreement in all cases.

V.3.1

THE CHEF~CAL EVOLUTION OF THE ISM IN THE SOLAR VICINITY

Time variation of the metal abundance ("metallicity") can be derived from observations of stars (G and M dwarfs) with MS lifetimes longer than the age of the galactic disk. Metal abundances are derived from line-blanketing indices such as the ultraviolet excess 6(U-B) or the StrSmgren parameter Am I. Naturally, these observations are limited to the solar vicinity. In the cumulative metallicity distribution of stars, the fraction S(Z)/S I of a sample of S I stars with metal abundance SZ is determined. This fraction is related to the SFR ~(t) by t S(Z)/S I ~ f ~ ( t ' ) d t '

o with t=t'(Z)

the time at which t h e ISM had t h e m e t a l l i c i t y

B e l l (1975) a l r e a d y ,

Z. I t has been shown by Lynden-

t h a t h i s "Best A c c r e t i o n Model" d i s c u s s e d in S e c t . V.1.2 and i n c o r p o -

rated in our model, can well reproduce the observed cumulative metallicity distribution. Twarog (1980) has refined this type of analysis by determining in addition to the metallicity of F dwarfs their ages from four-color and H8 photometry. His observed age-me~llici~y

re-

lation (AMR) in Fig. V.3 shows a sharp rise during the early phase of galactic evolution, starting ~12 Gyrs ago, but tends to flatten out during the last ~5 Gyrs, about the time the solar system was formed. As discussed by Twarog this time dependence is typical for (unJPv^ 26:3 - v

R,

206

I

I

I

I

Gusten and P.G. Mezger

I

0.1

Fig. V. 3: Age-metallicity relation (AMR) as obtained by Twarog(1980) for F dwarfs in the solar vicinity (filled squares and error boxes). Curves are A ~ s as predicted by our model for the chemical evolution of the ISM in the solar vicinity. The shaded curve has as its limits solutions for Zh=O and 0.2 Z@, respectively. saturated) infall models with constant ratio

-0.1

of net SFR to accretion rate, ~(t)=const. (Eq. V.3). In this case Eq. (V.9c) has the -o.~

simple solution

Z

-05

=

(V.16)

(yik+Zh) [I-(~/~o )(k-')-1]

with ~ the present-day gas-to-total mass ratio and ~o the initial ratio at the time 0

,~

8 12 STELLAR A6E IN GYRS

16

accretion started. The fact, however) that this relation fits the observed AMR can not be used as an argument for ~(t)fficonst.

In

Fig. V.3 are shown two AMRs, calculated from our model and two different values of the metal abundance of the infalling halo gas, Zh. For these numerical solutions of the set of equations (V.4,6,7 and 8) an exponent k=].5 (not to be confused with k in Eq.

V.16) has been

adopted, intermediate between the (extreme) cases k=] and 2 discussed in Sect. III. In spite of the strong time variations of both accretion rate and SFR (see Fig. V.], lower diagram), our predicted AMR for Zh=0.2 Z8 gives an excellent fit to the observed relation. But regarding the uncertainties in the age-scaling of the stars, satisfactory agreement may be obtained also for the case of accretion of unenriched halo gas. It should be stressed that this good agreement between observed and predicted time variation of the metal abundance of stars in the solar vicinity is a characteristic of infall models, and does not require bi-modal star formation.

V.3.2

THE CHEMICAL EVOLUTION OF THE ISM IN THE GALACTIC DISK

As shown in Sect. V.2.3 a yield which varies with galactocentric distance R is a consequence of bi-modal star formation. In this section we compare the abundance variations predicted by the model of bi-modal star formation for a number of primary and secondary elements with observations. Model parameters are k=l.5, me=3 m 8 and a=l. Production efficiencies P(R,mc) , analogue to F4~V. 13a) but without the instantaneous recycling approximation, have been computed with the IMF weighted production of elements Z. as given in Fig. C.I. Predicted and obi served abundances are shown in Fig. V.4. It should be noted that we have adjusted all predicted abundances to agree with abundances observed for the ISM in the solar vicinity. Absolute abundances of 160 (and other primary elements) predicted by our model are usually by a factor of two higher than the observed abundances. As shown in Table 1.1 this disagreement exists already for the closed system model, but is aggravated for the bi-modal star formation model as shown in Table V.I. We do not attach importance to this point, as several uncertain parameters enter the calculation of absolute abundances such as stellar yields, the state of gas depletion ~(R), the galactic history etc.

Star Formation and Abundance Gradients

[i0-~,] '

1

I

[,60/HI

I

'

I

1

IwO/HII

207

I

I

I

(b)

i~'He/HI

(a)

x

-÷-

(+) ~ b l e

10

I

'

x

yield

0.12

xl x M x

x

0.08

constant,ie~7- - - "~T--'"- ~ + 0.0/* I

,

[

I

90 60

,

I

[12Z/13[1

I

i

J I

I

(c)

[]

~

I

,

I

I

,

I

[180/1701

(d)

e

rl

m

consi'anf Y~mld~. -. -

I

1 k

30 I

i

2

I

i

6

I

10

,

I

I

,

1/, 2 GALACTIC RADIUS R IN kpc

I

,

6

I

10

I

1~

Fig. V. 4: Observed and predicted present-day element abundances as a function of galactocentric distance R. Solid lines refer to predictions of the hi-modal star formation model with parameter ~=1, mc=3m @. Shaded curves have as their limits solutions for ~z (from Table II. 2) and ~=const.=O.04, respectively. Dashed curves are the predictions of our model for ~ 0 (i.e. no star formation in spiral arms and continuous I~). Symbols o and ® refer to predicted and observed solar system abundances. Predicted abundances have been adjusted to agree with abundances observed for the ISM in the solar vicinity. a) Symbols + (Mezger et al., 1979) and o (Shaver etal., 1983) refer to observed 160 abundances. b) Shaded curve shaws the variation of the ionized abundance y+ with R of a group of giant radio HII regions. Open dots refer to high-resolution radio recombination line observations yielding lower limits to the number abundance y(ISM) ~y+ (Thum et al., 1980). Crosses refer to He-abundances in planetary nebulae yielding upper limits y(ISM) ~y(PN)=y++y ++. Solid curve is the He abundance Y=Y~+~Yev, with yD=O.075 the primordial abundance by number and AYev, the contribution from stellar nucleos~nthesi8, predicted by our model. c) Crosses refer to observations by Henkel et al. (1980, 1983) of H2CO isotopic radio lines. d) Crosses refer to observations by Penzias (1981) of CO isotopic radio lines. Triangles show changes in the predicted variation for standard depletion (&, see Table C.I) and increased depletion ( A and dotted square) of ISo.

Figure V.4, together with its figure caption, should be self-explanatory. limit our discussion to a few comments. abundance is excellent.

Therefore, we

The agreement between predicted and observed 160

The stellar contribution to the He abundance, AYev , is that pre-

dicted for a non-depleted primary element and added to a primordial abundance of yp=O.075 (by number) or Y =0.23 (by mass). IZC follows closely the 160 enrichment but is somewhat less P effectively enriched (~-]0%) due to depletion into 14N during the CNO cycle. The isotope ratio 12C/13C is the ratio of a primary to a secondary element, where 13C is burned on 12C. Eqs.

(1.4 and 5) for the closed system evolution predict zP/z s = | / y S

and we

therefore expect that the gradient of 12C/13C should decrease toward the inner disk. Predicted and observed isotope ratios are in good agreement.

That the decrease of I~c/Iac is

less steep than the increase of 160 is due to the fact, that the yield of a secondary element from lower mass stars (~m c) does not fully benefit from the increase due to the bimodal star formation,

as discussed in Sect. V.].].

208

R. GSsten and P.G. Mezger

In the case of the isotope ratio lSo/IVO, both nuclei are probably secondary elements, so that no gradient at all would be exprected across the galactic disk. Observations seem in fact to confirm this, but the well established difference in the isotopic ratio between the solar system (5.4) and the ISM (3.7) is in conflict with this simple analysis. The observed strong variation of the ISO/IvO abundance ratio

may however in part be explained by strong

depletion of leO due to astration in lower mass (longer-living) stars. While the enrichment of a secondary (non-depleted) element proceeds =~y~y~ (in~) 2 (in terms of the simple model, Eq. 1.5), depletion leads to a less steep increase with ~ (and hence with time) during advanced evolution stages. Let the (negative) yield y~ describe the depletion of the secondary element j during subsequent astration, then Eq. (V. IO) becomes (for closed system evolution) d Zs d in~ = -ySzp + y-Zs

which is solved by zS = - ySyp (y_) 2

ll+y-lm/ - ~Y- 1

(V. 17)

Obviously in late phases and for effective depletion Z s =In~, and thus even the ratio between (here) two secondary elements (e.g. Is0/170) can show strong variations (=in~) if depletion for one of the isotopes is important.

The solid line in Fig. V.4 is calculated with y-(leO)

from Dearborn et al. (Table C.I). In case of leO this flattening is further enhanced by dilution of the ISM with the (leO-depleted) ejecta of low-mass long-livlng stars, which at present provide major contributions to the amount of gas in the ISM. Here we do not want to discuss details of the ISo production, for which we refer to Glisten and Mezger (1983). As a last point we want to discuss present-day and solar system deuterium abundances and their relation to the primordial abundance of this isotope. Deuterium is not produced but only destroyed during astration and therefore the primordial 2H abundance X

should decrease with P increasing metal enrichment of the ISM. With our model we derive for the present-day local

ISM a depletion of XIsM/X p ~0.05 (for ~=O.04) and for the local ISM at the time when the solar system decoupled a depletion of X@/Xp ~O.]7. With [2H/IH]@= 3.6 ]0-s for the solar -6 system and ~5 10 for the ISM in the solar vicinity (Table IV.I) we estimate with these depletion factors primordial deuterium abundances of [2H/IH]

~(I-2) 10-4. These values are P considerably higher than earlier estimates based on the closed system model with constant IMF, which yields according to Eq. (1.6) a depletion of XIsM/Xp=O.2 only. The predicted de-

crease of the solar system deuterium abundance to the present-day ISM composition,

[Xe/XIs M]

~3.5, is consistent with recent observations by Vidal-Madjar et al. (]983). The enhanced astration in the inner Galaxy due to the increased number of short-living stars per stellar generation in our bi-modal star formation model leads to a decrease in XIS M by a factor of 1.5-2.O between R=9.5 kpc and 4.5 kpc.

V.3.3

STAR FORMATION RATES DERIVED FOR BI-MODAL STAR FORMATION

As discussed earlier and shown in Fig. 111.4, SFRs derived from observed Lye photon production rates adopting a constant IMF are about ten times higher then SFRs predicted from time averaged SFRs. As mentioned in Sect. V.I.1 one effect of bi-modal star formation is to increase the Lye photon production rate per unit mass, NLyc/~ , since the bi-modal star formation process produces more massive stars relative to low mass stars. Hence we expect that SFRs derived from Lye photon production rates adopting bi-modal star formation will be lower than the corresponding SFRs derived for a constant IMF.

Star Formation and Abundance Gradients

I

'

I

'

I

,t._.

I I/

o

I~obsLl-r]konst,tlF)

.,..o

/ ! i

I

.,

~

~)obs II-~lbimd.

IHF)



~!~!~!~!~!~..~.~!~!~ii~!~!~:i~t~irell-r] I

4.

i

I

Fig. V. 5: Observed and predicted present-day net SFRs, i.e. the amount of matter permanently locked-up in stars. ~obsEl-r] (const. IMF) refers to net SFRs obtained for a constant IMF from the observed Lye photon production rates (see also Fig. III. 5 and Sect. III. 3.3). ~re[1-r] are the net SFRs which are c~nsistent with the present stellar mass distribution (see text). ~obs[1-~] (bi-mod. IMF) refers to the net SFR obtained for the case of bimodal star formation from the observed Lyc photon production rates. This is demonstrated in Fig. V.5. Shown

.

/

209

I

l

8 12 GALACTIC RADIUS R IN kpc

here are the net SFRs (i.e. that part of the ISM which is converted permanently into stars, normalized to the total present-day mass of stars to<~(l-a)> (with to=tdisk the age of the galactic disk and <~(l-r)> the

average net SFR). In the case of continual star formation observed and predicted net SFRs should coincide. The gross discrepancy between predicted and observed SFRs, if the observed SFRs are derived for a constant IMF from the observed Lyc photon production rates, has already been mentioned in Sect. 111.3.3 and Fig. 111.5 and has been noted before by Cass& et al. (1979) and Talbot (]980). This discrepancy between observed and predicted net SFRs disappears, however, if net SFRs are derived for hi-modal star formation from the observed Lyc photon production rates. The reason is that in spiral arm star formation the fraction of mass that is permanently locked-up in stars is much smaller (due to the increased mlow=mc >I m@) than through star formation in the interarm region with ~low ~0.I m@. As a quantitative demonstration we compare SFRs and net SFRs for the galactic disk as derived from the observed total Lyc photon production rate given in Eq. (111.4). We obtain for the present-day SFR in the galactic disk ~(const. IMF) ~]3 m@ yr -I and

~(bi-mod. IMF) ~6 m e yr.

For the net SFR (i.e. the mass of ISM which per year is permanently locked-up in stars) we derive ~[l-r]

(const. IMF) ~9 m@ yr -I as compared to ~[l-r] (bi-mod. IMF) ~1.5 m e yr -I. The

specific parameters of the model of hi-modal star formation, ~=l and m c ~3 m@, give reasonable fits to the observations discussed in Sect. V.3.2 and this section. At this point we would llke to come back to the galactic diffuse near infrared (NIR) emission discussed in Sect. 111.4.2. There we encountered a similar problem as with SFRs derived for a constant IMF: Both NIR emission and SFRs do not scale with the preseNt-day mass distribution of stars in the disk. To obtain a fit for the ridge line intensity of the observed NIR emission we had to add a hypothetical population of M-giants with masses m Z(2-3)m®, whose distribution follows that of O stars. We showed that the luminosity of these M-giants, integrated over the galactic disk, is consistent with a SFR for stars m ~(2-3)m8 which ~IO s yr ago (the time, when the progenitors of the present-day M-giants formed) was not much different from the present-day SFR for the O stars. This result is easily explained by the model of bi-modal star formation, where the spatial distribution of stars with m Z(2-3)m@ (and hence of O stars and the progenitors of M-giants) is similar. The IMF for bi-modal star formation, Eq. (V.12), yields the same spatial distribution for O stars and the progenitors of M giants (m ~2-3 m e ) but a different spatial distribution for lower-mass stars, which determine the mass distribution of stars in the galactic disk.

210

R. GSsten and P.G. Mezger

ACKNOWLEDGEMENTS It is our pleasure to thank C.M. Walmsley and T.L. Wilson for reading the manuscript. Our special thanks go to Mrs. Breitfeld and Mrs. M~rz who typed this manuscript on a very tight time schedule.

Star Formation and Abundance Gradients

APPENDIX A:

211

PRODUCTION RATE OF LYMAN CONTINUUM PHOTONS FROM EARLY-TYPE STARS

In Sect. III we use observationally determined Lyc photon production rates to estimate the SFR in the galactic disk. The Lyc photon production rate of a star of mass m and main sequence (MS) age T is

NLyc(m,%) = 4~-R~(T)-nLyc(Te(T),g(l~,m)) , with R~ the stellar radius, g the surface gravity and T photon flux

Fig.

nLyc(Cm-as-1)

while R,, T

spheres,

at

the stellar

as function of m,T are derived from models of stellar evolution. In

e temperatures

A.I effective

surface

the effective temperature. The Lyc e is obtained from models of stellar atmo-

and stellar

radii

are compiled for the zero-age

('ZAMS')

and terminal-age ('TAMS') MS from recently computed stellar evolution tracks. The MS lifetime TMS ( F i g .

A.2)

is defined

here as the phase of core-hydrogen

burning;

it

ends after

the

core exhaustion of hydrogen, when the star begins to contract (first T -minimum) and shell e b u r n i n g d e v e l o p e s . F o r m o s t t r a c k s Y ~ 0 . 3 a n d Z=O.03 h a s b e e n a s s u m e d a n d no m a s s l o s s h a s been taken into account. Stars with less enriched initial composition

Z ~0.02 Z e are

I

I



I zAMs .08 06

% ±TAMS

-BO -BI

=

-B2 -BS

2

I 25

~z4 Z3

TAMS.~

T

p

Fi~. A.I: Effective temperatures and stellar radii for zero-age ('ZAMS') and terminal-age ('TAMS') main-sequence stars according to stellar evolution tracks from Alcock and Paczynski (1978, e), Becker (1981,x), Brunish and Truran (1982,o, Z=O. 02), de Loore et al. (1977,A), Ezer and Cameron (1967,A, Z=O.02), Maeder (1981,+), and Stothers (1972,o). If not indicated otherwise, an initial composition of Z=O.03 and Y ~0.3 has been assumed. Spectral type (Te) and stellar mass (m) are connected following Cont i (1975), Hutchings (1980) and Remie and

illlili] I

6O

20 40 STELLARMASSm IN me

I

109

"\

I

I~~MS=

I

1

Fig. A.2: Main-sequence lifetime versus stellar mass. For references see Fig. A.1. Simple analytical fits for m ~ 8 m@ and Z=O.03 are given. Tracks computed for a less enriched initial composition (i.e. Z =0.02) yield shorter MS lifetimes.

5.109. m-Z'7*l.Z.107

~I0 B t-o

~',

g

/ I~Ms=1.Z.109.m-t8%3.106

systematically hotter, more luminous and evolve faster. The simple analytical ap-

:E

10~

proximations of TMS(m) given in this figure refer to stellar evolution tracks with Z=O.03. For reasonable, viz. obserI

I

I

I

I

I

3

ID STELLARMASSmINme

30

60

vatlonally well-determined mass loss rates of lO-(v tO6)m® yr-1 for 30-40 m@

212

R. GUsten and P.G. Mezger

MS stars (Abbott et al., 1981, Garmany et al., 1981) T and R~ differ only moderately from e values obtained for constant-mass tracks (Brunish and Truran, 19821Maeder, 1981). Hence, as shown below, the IMF weighted Lyc photon production rate of a given stellar population is underestimated by only ~10-20% and thus is comparable to the uncertainty introduced by assumptions of the initial composition. Therefore, we neglect mass loss effects in the following discussion. At any rate, we feel that the biggest uncertainties in our computations of the Lyc photon production rate come from the adopted relation between stellar atmospheres and stellar masses, which is for massive stars m ~30 mQ entirely based on theoretical investigations. Only for less massive objects have observations of wide binary systems shown satisfactory agreement (Hutchings,

1980). Other uncertainties in the computed Lyc photon

production rates come from the models of stellar atmospheres. Since the time of earlier compilations of Lyc photon production rates (Panagia, 1973; Mezger et al., 1974) results of more realistic llne-blanketed LTE model atmospheres have been published (Kurucz, 1979) which predict moderately reduced Lyc photon production rates and less He ionizing Lyc photons. For surface gravities g, appropriate for massive stars, values of nLyc(Te,g) are shown in Fig. A.3. Kurucz's models are computed for solar system abundances only and with increasing metal abundance (and hence line opacity) a lower Lyc photon production rate is to be expected for a star of given mass. While this effect is not considered here, the (probably much stronger) effect of a decrease of the upper mass limit (mup) of the IMF with increasing Z is taken into account.

In Fig. A.4 we give the Lyc photon production rate as a function of stellar mass for both the ZAMS and TAMS

points on the evolution tracks. The total number of Lyc photons produced

by a star of mass m during its lifetime is obtained by integration

NLyc(m) = j" NLyc(m,T)dT TMS

SPECTRAL TYPE B5 I

I

B2

B1

I

I

BO I I

(A. 1)

and given in Fig. A.5. For two

08 i

06 I

mass ranges analytical approxi-

I

mations are also given in this Figure. We see that only stars A

.t-~/

with masses >10 m@ contribute significantly to the integrated Lyc photon production rate and after weighting with the IMF -

1022

the "typical" ionizing 0 star is

u

found to be in the mass range 30

102o x

C=, 10~8

, ~

~4.5° x I 2

I I 3 4 TEMPERATURETeff (lOZ'K)

I 5

Fig. A.3: Lye photon flux densities, given as function of surface temperature (Te) and gravity (g), according to the LTE line-blanketed models of Kurucz (1979), and solar system abundances (Z ~0.02) of the stellar atmospheres. For comparison Panagia's flux densities are given, derived from earlier stellar model atmospheres and for log g=4.

Star Formation and Abundance Gradients

I

u

U

213

I

u

n

48

It.x-~ 46 u

~ /,4 ,,,,It

42 o

I I0

,

I 30

ZAMS '( TAMS o I 50

,

,

STELLAR MASS m IN me

Fi~. A.4: Lyc photon production rate NL (m,T) in s -I for zero-age (ZAMS) and terminal-age (TAMS) YCain-sequence stars as function of stellar mass. I

I

I

I

I

01

NLYC('l:,m)d1:

/ /

MS

/ / / I I

C~

/

3.0

i I

8

/o I

2.75 106z(m-261°'6~,,,, / .ii / £'/

Z.O

/ / iI

~

iI le ii

1 . 0

/ I

lO

,- .,t

,J ~

3.5 10ss ,m 5 I

I

I

30

50

70

STELLAR MASS m IN me

Fi~. A.5: The total number of Lye photons emitted by a star of mass m durin E its main-sequence evolution. Simple analytical approximations are given for m ~ 30 m@.

214

R. GHsten and P.G. Mezger

to 40 me, corresponding to a spectral type 06. Relation (A.]) weighted with the IMF, yields the quantity

e(~)

= f ~(m) .NLyc(m)dm

(A.2a)

which in this form is independent on the normalization of the IMF (especially on mlow) but depends on both the shape of the IMF, ~(m), and on m up . We have evaluated eq. (A.2a) for both the Salpeter (sal) and Miller-Scalo (ms) IMF. For mup 235 m 8 P(@) may be approximated by =

pt~ms)-.~ .-

p(~sal)

3.5.106o.[1-l.l.lOa.m

=

-1-97

+

up

1.2.10O.m-2.av+4.0.10".m

up

-s-97]

up

2.9.1061.[l_13.5.m-O.V + 94.m-1.7 + 261.m-2.v] up up up

(A.2b)

Substitution of eqs. (A.2b) in eq. (111.12) yields a relation between Lyc photon production rate NLy c and SFR NLyc/~ = ~(1).P(~)

(A.3a)

which for Salpeter and Miller-Scalo IMF, respectively, has the form

~sal ms

• A54(

NLye/~ = 3.2 lu

--0,35_

mlo w

--0.3Sx--1

mup

)

(0.I056-1.43 m-~7+9"g4up m-l"7)up

NLyc/~ = f-12.5 I0S6(4.71410-4-0.508 m-l"gV+5"69up m-2"SV)up

with 0.1
4.44 - 2.5 mlo w°'4

i2.75 f(mlow,mup) =

m~-O.808

- 16.8 m -1"62 up

1.O
5.1 mZl"2~-71ow.210-3-16.8 m -1"62up I0 --1.62

16.8 (mlow

--1,62

-mup

)

(A.3b) 30

As usual all stellar masses are expressed in solar masses. APPENDIX B:

DECONVOLUTION OF THE DIFFUSE GALACTIC FREE-FREE EMISSION ASSUMING AZIMUTHAL SYMMETRY

After separation of the thermal and non-thermal components, one obtains the diffuse galactic free-free emission as a function of galactic longitude I and latitude b. The ridge line intensity S(~,b=O °) is shown in Fig. 111.2 in units of Jy/beam area. Eq. (111.3) relates the flux densities at a distance D to the volume emissivity e. The telescope beam with which the diffuse emission e was observed had HPBW's of AlxAb, corresponding to a beam solid angle = ].]33AZ.Ab. As outlined in Sect. III the free-free volume emissivity is directly rem

lated to the Lyc photon production rate and hence to the distribution of 0 stars in the galactic disk. Substitution in Eq. 111.3 of the volume emissivity eff(R,@,z), given as a function of the cylinder coordinates (R,@,z) (see Fig. B.]) yields for the ridge line flux density

Star Formation and Abundance Gradients

S(Z,b=O) ~m

1

@

215

Z

m m Rout ~-- f dO f dR R ~ dz ~(R,C),z)/4~T(D2+z 2) mo R.inn -z m

(B.I)

Note that z is the coordinate perpendicular to the galactic plane and 6 = arc tan (z/D). The integration limits O ~z ~z

= D tan(Ab/2) max

:0

o

Rin~ R(O,/-AZ/2) ~R ~Rou t = R(O,/+AZ/2) 0 ~@ ~@max :

G.C.

@(/'Rmax)--

depend on both Rmax, the cut-off radius of the freefree emission, and on the telescope HPBW's, AZ and ~b. If Ab is small as compared to the latitude distribution of the free-free emission,

(i.e. if

Dmax'tan(Ab/2) < Zo, with z° the scale height of the

\

free-free emission) and if we assume azimuthal symmetry for the distribution of 0 stars (i.e, Eff(R,Q,z)÷Eff(R,z))

SUN

then eq. (B.I) reduces to D max

~

:

Definition of coordinate

S (/'b=o°) m

I

~

I

J

I

x'" "x

12 •

i rn

x-.

~-9

,

,.-,,

i

Y

i=~

=--6

l xl \

i +-+,

/f

I-=-

I=~ I----

l ,,_~I~

i

I

/

'."



E (R)dD.

'

(B.2)

I

x WESTERHOUT ~,0+ MAIHEWSONet oL.~"O

X~

A

f o

'~

',

\ ,,

1,, \ \,..-~.

x%%

÷

%,

i..,..i z

,-3 ZXZ

r,,,."

co

I

I0

i

I

~

I

30 50 LONGITUDE,~ IN DEG

,

+

70

Fig. B. 2: Ridge line brightness temperature of the diffuse thermal background according to Westerhout (1958) and Mathewson et al. (1962). Observing frequencies and angular resolution were (1.39 GHz, 34') and (1.44 GHz, bO'), respectively.

216

R. G{isten and P.G. Mezger

We neglect free-free emission outside the solar circle. Hence R

max

= I0 kpc and

D = I0 cos(l) ±[(R 2 - lOesina(/)]-°'s d D = ±(R 2 _ 10 2 sin2(/))_o.s R dR

with ± corresponding

to @ ~ 90 ° . Substitution

in the above relation yields

R

S(l'b=o°)

2"~axe(R)

~m and, with new variables

R (R 2 - 10 2 sin2(/)) -°'s dR

10sin(l)

u = R2 - R 2 and Q = R 2 - 102. sin2(/) max max Q S(Q) = Se(u ) (Q _ u)_O.S du m o

(B. 3a)

This integral equation can be solved by a Laplace transformation,

yielding the solution

u

s



(B.3b)

(u

m o

Obviously,

the accuracy of the deconvolution

depends critically on the accuracy with which

the derivative of the ridge line intensity S(/), i.e. the quantity dS(Q)/d~

in eq.

(B.3b),

can be determined.

Fig. B.2 shows the ridge line intensities diffuse galactic free-free emission,

(expressed

in brightness

as derived from observations

Mathewson et al. (1962). The uncertainties

in these low-resolution

temperatures) by Westerhout observations,

of the (]958) and which hardly

fulfill condition B.2 are large. To follow how errors in the ridge line intensities into the deconvolved

radial volume emissivities,

cedure. Herein e(R,z=O)

is approximated

and for 'm' independent

line-of-sights

we developed a numerical

by 'n' radial

propagate

deconvolution

intervals of constant

pro-

emissivity el'

along the galactic plane eq. (B.l) becomes a system

of m times n linear equations. n S(/j) = E e "F i=] i ij '

! ~j Sm

which is inverted numerically.

The radial emissivity and Mathewson

e(R) as derived from the average of the measurements

of Westerhout

et al. is shown in Fig. B.3a. The error bars measure the uncertainty

radial distribution

as estimated from the differences

in the

between the two sets of brightness

temperatures.

The corresponding Lyc photon production rates, derived from g(R) as described 111.2.2 are shown in Fig. B.3b.

in Sect.

Star Formation

and Abundance

I

I

I

Gradients

I

217

I

b 50 • •

.:.:..'.:.:.

......

::::::::::~:~

.':.;...,,.,.;.~,..'

Z

~...-,~

~- 30

: .'!:~:::...~

r-~

~%'~ ~

r~

:.,. ,~.:::. "•":i

I

I0 :~. ..:.

I

I

I

Z

4

I

I

_6

I0

8-

N -rI

4

>

- -

m

LZD

m

I

I

2

4

I

I

I

6 8 RADIUS(kpc}

I0

Fi~. B. 3a : shows the radial distribution of the free-free emissivity ~(R) in the galactic plane, as deconvolvedfrom the ridge line brightness temperatures shown in Fig. B.2. For a gaussian z-distribution, the Lye photon production rates, integrated over 1 kpc wide annuli, are derived in Fi E. B. 3b.

APPENDIX C:

SITES OF NUCLEOSYNTHESIS

In this appendix we give a short summary of the stellar nucleosynthesis elements

and isotopes, which are incorporated

sented before by e.g. Audouze et al. mechanisms

Such discussions

for the

have been pre-

(]975); ours is an update which includes new production

and more accurate estimates of yields. While there is general agreement

nuclear processes controversy

in our model.

processes

that synthesize

the elements and isotopes discussed here, there is some

about which astrophysical

for general references). Perhaps more surprising the predictions

about the

objects are the main sources

For example the origin of rare isotopes

(see e.g. Truran

is the fact that there has been considerable

of nuclear astrophysics

and models of chemical

(1977)

such as leo is unclear. disagreement

between

evolution of the Galaxy,

in

218

R. GUsten and P.G. Mezger

particular concerning the nucleosynthetic sites of the conm~on species of the CNO-group (and Helium).In addition to the physical conditions in the synthesis regions (temperature, density, reaction rates, etc.), a major uncertainty is how to dredge-up the freshly made elements from their production sites deep in the star's interior to the envelope. From the envelope they may be ejected into the ISM - via (steady) mass loss, planetary nebula-like ejection and super nova explosion.

In the following we briefly summarize the astrophysical environment in which the nucleosynthesis of various elements and isotopes and their liberation into the ISH may occur. To obtain the timescale of the enrichment, special attention must be paid to the mass (and hence age) of the stellar progenitor and to the question whether the synthesis process is primary or not. Species like 160, that in a given star can be synthesized from primordial hydrogen and helium (perhaps in a network of processes) are termed primary elements. Those like ISc, which are built up from a pre-existing abundance of elements such as 12C, which are produced in a past generation of stars, are of s e c o n d l y

origin. Consequently the en-

richment in secondary relative to that in primary nuclei is delayed (see Sect. V). The production rates, compiled in Figure C.I and Table C.I from recent yield calculations, are based on single-object tracks only, and may have to be modified if interaction in a binary system is taken into account. Although 2H is a basic step in normal hydrogen burning, its destruction proceeds much faster than its formation. Hence in stellar nucleosynthesis, 2H is not produced. Furthermore, the reaction 2H(p,y)3Hestarts at temperatures ~(0.5-].0)]0 s K. This temperature is generally exceeded in stellar interiors, even during pre main-sequence evolution, so that the pre-stellar (primordiaD Deuterium cannot survive. As mentioned before, the amounts of freshly made ~He (the end product of simple H-burning), 12C and 16C (products of He-burning) as well as 13C, I~N, and IVO (from the CNO-cycle) that are liberated into the ISM are rather uncertain. Hence these production rates depend on details of the late evolution of the stellar progenitor (i.e. mass loss, mixing efficiencies along the giant branch etc.). Earlier investigations of the pre-supernova composition of massive He-cores ~3 m e (corresponding to total mass ~I0 m~), without mass loss have been made by Arnett (]978). More recent closed evolution tracks (including the hydrogen envelope) with mass loss yield somewhat lower metal-productlon

(Maeder, ]98]). It is shown that various rates of (mainly post main-

sequence) mass loss affect the net yields of He and "metals" only slightly. Although with increasing mass loss rates, stellar winds from evolved cores (WN, WC-stars) become an important ejection mechanism, their contribution to the enrichment of the interstellar gas with 12C, I~N and IVO is only a fraction of that expected from medium mass ~ stars (see Fig.

C.1). 160

is ejected into the ISM from stars m ~I0 m@ only. Otherwise it is kept locked in a stel-

lar carbon-oxygen core that never exceeds the Chandrasekhar-limlt.

For most of the other

elements however, their major contribution is from medium mass stars. As pointed out by

Stars m >8 m@ ignite carbon burning in a non-degenerate core, and develop the well-known shell-structure during their subsequent burning phases. Thus their evolution is quite different from that of stars of lower mass, and we follow the c o l o n classification into massive (>8 m 8) and medium mass (!
hot CN

ISO(a,v)laC

incomplete CNO

15N

le0

170

(CHO)

References:

(I) (2) (3) (4) (5)

A r n e t t (1978) R e n z i n i + Voll (1981) Woosley + Weaver (1980) Dearborn e t a l . (1978) Thlelemann + Arnold (1978)

I~N

Is O

laC

12C,I~N

CNO

a-capture CNO

~ac(~eO)

~aC

~He

1~N(a,y)InF

a-capture

CNO

CNO

l~N(~,T)leF(8+)leO(a,y)a2Ne

CNO

I~N

Is 0

CNO

13C

incomplete

triple'a'

laC

a-capture

a-capture

pp,CNO

"He

Seed-nuclei

Nucleosynthesis Process

Destruction

aD(p,¥)SHe

Production

(6) (7) (8) (9) (IO)

IHS

MS

IMS

(Ms)

MS

IMS/MS?

IbiS

IMS

IMS

7.9 ( - 4 )

-3(-7)?

?

?

~2(-5)

2(-6)

6.4-9.2(-3)

?

6=1.5 3.1 ( - 3 ) 2. I ( - 3 ) p r i m .

6ffiO

6=1.5 1.6(-4) 0.9 (-4)prim.

in m a s s i v e s t a r s

d u r i n g q u i e t c o r e H e - b u r n i n g i n m
e x p l o s i v e H e - b u r n i n g (SN?)

incomplete CNO-burning with deep mixing along AGB, no hot-bottom burning included hot CNO-burning in novae/supernovae

enriched winds from WN-stars

helium-burning

hot CN-cycle (T>2-108) in novae (or supernovae)

major sink in CNO-cycle; s e v e r a l d r e d g i n g up p h a s e s a l o n g AGB, h o t - b o t t o m b u r n i n g in c o n v e c t i v e e n v e l o p e

low-temperature 'hot-bottom t burning

6=0

4.4 (-5)

incomplete CNO-burnlng in Red Giant envelopes

in Red G i a n t s

4,5

4 6,7 6

9

1,3.8

7,9

2,4

2

1,3,10

WR-stars

Helium-burning, partly dredge-up

1.5 ( - 3 )

2

1,8.10

References

dredge-up in Red Giants 6-mixing length ratio

winds (WR-stars)

of all masses

hydrogen-burning, 50Z in s t e l l a r

completely destroyed in s t a r s

Notes

3.9 ( - 3 ) 6ffil.5 2 . 3 ( - 3 ) 8ffiO

IMS MS

1.3(-2) 6=o 1.8 (-2) 6=1.5 1 . 9 ( - 2 )

-0.6

Net Bulk Y i e l d s

MS

all

Stellar ~ss Ran~

Howard e t a l . (1971) Audouze e t a l . (1978) blaeder (1981) Woosley + Weaver (1980) Maeder ( 1 9 8 2 ) , Abbott (1982)

(ter.)

sec.

sec. (prim.)

prim.

sec.

(prim.)

sec.

(prim.)

sec.

prim.

prim.

Type

YIELDSOF NUCLEOSYNTHESIS PROCESSES

2H

Species

TABLE C.1

e~

> o"

~rj 0

220

R. GUsten and P.G. Mezger

Iben and Truran (1978), during their post main-sequence evolution (first ascent along the red giant branch) these stars develop deep convective envelopes, that even penetrate to the outer layers of the hydrogen burning shell, carrying fresh products of CN0-burning (ItN, the and some 13C, Iv0) to the surface ('first dredge-up' phase). In stars more massive than~3 me further stirring occurs after ignition of the He-burning shell ('second dredge-up' phase) and during their unstable asymptotic giant branch ('AGB') evolution (third dredge-up). In this last phase these stars undergo He-shell flashes, that stimulate periods of convective mixing between the base of the He-burning shell and the stellar envelope, dredging up fresh the (from the H-burning shell) and lmc (from the He-burning shell; see Iben and Truran (1978), Wood (1981) and references therein)). Depending on the details of the mixing process, the matter released during the AGB-evolution (via stellar winds or planetary nebula ejection) is significantly enriched in ImC and tHe. Further modification of the envelope composition could be caused by hot-bottom CN-burning during the quiescent interflash phase. This occurs when the temperature is sufficiently high at the base of the convective envelope to start the (cold) CN-cycle, and convert part of the fresh ImC, that has been dredged-up from the He-shell, into 13C and ItN (see Renzini and Voli, 1981). The efficiency of this burning process depends sensitively on the temperature (and hence the treatment of convection) in these envelope layers. For illustration in Fig. C.I, we show results for two extreme cases, i.e. for 8=0 and 1.5. The parameter 6 is the ratio of mixing length to pressure scale height. Although hard to quantify, hot-bottom burning is nevertheless of special importance for our chemical evolution models, since part of the 13C and ItN (and probably Iv0, which has not been included by Renzini and Voli) is of

primary origin in this process (Fig. C.1). Clearly, observational constraints on the 'primary' tiN-yields are needed, either from abundance determinations in evolved AGB-stars (Wood etal., 1981; Kaler, 1982) or from chemical enrichment models as described in Sect. V. The nucleosynthetic sources of 15N and leo are even more uncertain, and during normal (cold) equilibrium CNO-burning both are destroyed. A possible site of 15N (and IVO) production seems to be hot CNO-burning in novae (and supernovae), because at high temperatures the final abundances in the cycle are changed in favor of the rare 15N (Iv0) isotopes.

For example, for

T >2 10s K the 15N/tiN ratio may increase to unity, compared with 10-t in the cold cycle. There are several potential sites for the nucleosynthesis of Is0. However the problem is to protect 180 from destruction in further CNO- or He-processing [Is0(a,y)22Ne]. One promising production scheme is explosive He-burning on 1iN in the He-rich shells during supernova-explosions (e.g. Howard etal.,

1971). However, using recent results on the Is0(a,y)2mNe cross-

section, Thielemann and Arnould (1978) suggested that even during "quiet core" and/or shell He-burning in medium

mass stars (2 ~m N5 me), significant amounts of leo could be pro-

duced. What fraction of this survives further He-shell flash-burning (most likely in the lower mass objects with lower flash temperatures) and can be mixed to the surface is uncertain. In these scenarios 15N and leo are likely to be of secondary origin. They are synthesized from ItN, which itself is produced in situ (from ImC) in this stellar generation. If, however, burned on 1iN, produced in an earlier stellar generation, part of the 180 may be even a "tertiary" product. In Table C.I the net bulk yield Yi of a stellar generation is calculated, which is defined as the fraction Pi of mass of newly synthesized element 'Z.',I per fraction of matter locked up in long-lived objects Yi

Pi f~(m)Am(Z)dm (l-r) = (l-r)

(C.1)

Star Formation

Am(Z)

and Abundance

is the mass of the newly synthesized

Net bulk yields

Gradients

221

element Z that is returned

in Table C.] are given for the Miller-Sealo

into the ISM (Fig. C.I).

IMF with

(]-r) = 0.6

I

1

[1¢]

I

' 10s211

16 0

2

,

13 C

I

I I-.

E
E

.I

,

e

10

Z

I

I

I

30

--

i10-3]

T

4He

"~i i

5 z,

i l

1.0

l

(MS t MS_. O I

5

-r-

t

10

10

"'

2O

3O

I

I

i

,,



14N

3 2 1

ti i I

I

20

10

![10-4]

I

f

2.0 C)

I

70

50

-- PRIMARY

--

-

"',.

_

i

PRIMARY

)

I

10

20

12 C

[10_1, l >

I

'

• ! o • T

I

;

I

10

i

20

I

ARNETT (1978) MAEDER [1981) DEARBORN et al (1978) WOOSLEY+WEAVER (1980) RENZINI+VOLI (1981) : a = o 6

15

I

30 STELLARMASS(too)

Fi9. C.I: I ~ weighted production of some primary and secondary elements. Am(Z) is the mass of a newly synthesized element Z that is ejected into the ISM by a star of mass m. Symbols refer to model computations by different authors.

222

R. G~sten and P.G. Mezger

REFERENCES Abbott, D.C., Bieging, J.H., Churchwell, E. 1981, Astrophyso J. 250, 645 Abbott, D.C. 1982, Astrophys. J. 263, 723 Alcock, C., Paczynski, B. 1978, Astrophys. J. 223, 244 Altenhoff, W., Mezger, P.G., StraBl, H., Wendker, H., Westerhout, G. 1960, VerSffentl. Universit~tssternwarte Bonn Nr. 59 Arnett, W.D. 1978, Astrophys. J. 219, 1008 Audouze, J., Lequeux, J., Vigroux, L. 1975, Astron. Astrophys. 43, 71 Baker, P.L., Burton, W.B. 1975, Astrophys. J. 198, 281 Becket, S.A. 1981, Astrophys. J. Suppl. 45, 475 Binette, L., Dopita, M.A., D'Odorico, S., Benvenuti, P. 1982, Astron. Astrophys. 115, 315 Blitz, L., Shu, F.A. 1980, Astrophys. J. 238, 148 Bohlin, R.C., Savage, B.D., Drake, J.F. 1978, Astrophys. J. 224, 132 Brunish, W.M., Truran, J.W. 1982, Astrophys. J. 256, 247 Bruston, P., Audouze, J., Vidal-Madjar, A., Laurent, C. 1981, Astrophys. J. 243, 161 Burbidge, E.M., Burbidge, G.R., Fowler, W.A., Hoyle, F. 1957, Rev. Mod. Phys. 29, 547 Burton, W.B. 1971, Astron. Astrophys. IO, 76 Burton, W.B., Gordon, M.A. 1978, Astron. Astrophys. 63, 7 Cameron, A.G.W. 1980, in "Nuclear Astrophysics", ed. C. Barnes, D. Clayton, D. Schra~un, Cambridge Cass~, M., Kundt, D., Scalo, J.M. 1979, Astron. Astrophys. 76, 346 Chevalier, R.A. 1979, Mem. Soc. Astron. It. 50, 65 Chiosi, C. 1980, Astron. Astrophys. 83, 206 Churchwell, E. 1980, in "Radio Recombination Lines", ed. P.A. Shaver, Reidel, p. 225 Combes, F. 1981, in "Extragalactic Molecules", ed. L. Blitz, M.L. Kutner, p. 159, Green Bank Conti, P.S. 1975, in "HII-Regions and Related Topics", ed. T.L. Wilson and D. Downes, p. 207, Springer-Verlag, Heidelberg Cox, D.P., Smith, B.W. 1976, Astrophys. J. 203, 361 Dearborn, D., Tinsley, B.M., Schra~, D.N. 1978, Astrophys. J. 223, 557 de Loore, C., de Greve, J.P., Lamers, H.J.G.L.M. 1977, Astron. Astrophys. 6_.!I , 251 Demarque, P. 1980, IAU Symp. 85 "Star Clusters", ed. J.E. Hesser, Reidel, p. 281 Dickey, J. 1981, in "Extragalactic Molecules", ed. L. Blitz and M.L. Kutner, Green Bank Dickman, R.L. 1978, Astrophys. J. Suppl. 37, 407 Downes, D., GHsten, R. 1982, Astron. Gesellschaft 57, 207 Downes, D., Wilson, T.L., Bieging, J., Wink, J. 1980, Astron. Astrophys. Suppl. 40, 379 Ezer, D., Cameron, A.G.W. 1967, Canad. Journal of Physics 45, 3429 Eggen, O.J. 1976, Q.J.R. Astr. Soc. 17, 472 Federman, S.R., Evans II, N.J. 198], Astrophys. J. 248, 113 Frerking, M.A., Langer, W.D., Wilson, R.W. 1982, preprint "The Relationship between Carbon Monoxide Abundance and Visual Extinction in Interstellar Clouds" Garmany, C.D., Olson, G°L., Conti, P.S., Van Steenberg, M.E. 1981, Astrophys. J. 250, 660 Georgelin, Y.M., Georgelin, Y.P., Sivan, J.-P. 1978, IAU Symp. 84 "The Large Scale Characteristics of the Galaxy" p. 65, ed. W.B. Burton, Reidel GUsten, R. 1981, unpublished thesis, University of Bonn G~sten,~R., Henkel, C. 1983, Astron. Astrophys. in press, "H 2 Densities and Masses of the Molecular Clouds close to the Galactic Clouds" GHsten, R., Mezger, P.G. 1983, Astron. Astrophys. in preparation Guibert, J., Lequeux, J., Viallefond, F. 1978, Astron. Astrophys. 68, 1 Hayakawa, S., Ito, K., Matsumoto, T., Uyama, K. 1977, Astron. Astrophys. 58, 325 Hayakawa, S., Matsumoto, T., Murakami, H., Uyama, K., Yamagami, T., Thomas, J.A. 1979, Nature 279, 510 Hayakawa, S., Matsumoto, T., Murakami, H., Uyama, K., Thomas, J.A., Yamagami, T. 1981, Astron. Astrophys. IO0, 116 Heasley, J.N., Milkey, R.W. 1978, Astrophys. J. 22], 677 Henkel, C., Wilson, T., Dowries, D. 1979, Astron. Astrophys. 73, L13 Henkel, C., Walmsley, C.M., Wilson, T. 1980, Astron. Astrophys. 82, 41 Henkel, C., Wilson, T.L., Bieging, J. 1982, Astron. Astrophys. 109, 344 Henkel, C., Wilson, T.L., Walmsley, C.M., Pauls, T. 1983, preprint "Formaldehyde towards compact HII-Regions; Densities and Isotope Ratios" Hirayabashi, H. 1974, Publ. Astron. Soc. Japan 26, 263 Hirayama, T. 1971, Solar Physics 19, 384 Howard, W.M., Arnett, W.D., Clayton, D.D. 1971, Astrophys. J. 165, 495 Humphreys, R.M. 1978, Astrophys. J. Suppl. 38, 309 Hutchings, J.B. 1980, Space Sci. Rev. 26, 331 Iben, I. Jr., Truran, J.W. 1978, Astrophys. J. 220, 980 Innanen, K.A. 1973, Astrophys. Space Sci. 22, 393 Ito, K., Matsumoto, T., Uyama, K. 1977, Nature 265, 517 Kahn, F.D. 1974, Astron. Astrophys. 37, 149 Kaler, J.B. 1982, in IAU Symp. 103 "Planetary Nebulae" Klein, U., Beck, R., Bucszilowski, U.R., Wielebinski, R. ]982, Astron. Astrophys. 108, 176 Knude, J. 1981, Astron. Astrophys. 97, 380

Star Formation and Abundance Gradients

223

Kunde, V., Hanel, R., Maguire, W., Gautier, D., Baluteau, J.P., Martin, A., Chedin, A., Husson, N., Scott, N. 1982, Astrophys. J. 263, 443 Kurucz, R.L. 1979, Astrophys. J. Suppl. 40, 1 Kutner, M., Mead, K. 1981, Astrophys. J. (Letters) 249, L15 Lambert, D.L. 1978, M.N.R.A.S. ]82, 249 Langer, W.D. 1977, Astrophys. J. 212, L39 Langer, W.D., Goldsmith, P.F., Carlson, E.R., Wilson, R.W. 1980, Astrophys. J. 235, L39 Larson, R.B. 1972, Nature Phys. Sci. 236, 7 Larson, R.B. 1982, M.N.R.AoS. 200, 159 Lequeux, J. 1979, Astron. Astrophys. 80, 35 Lequeux, J. 1981, Cormnents on Astrophysics 9, ]17 Lin, C.C., Yuan, C., Shu, F°H. 1969, Astrophys. J. 155, 721 Liszt, H.S., Xiang, D., Burton, W.B. 1981, Astrophys. J. 249, 532 Liszt, H.S., Burton, W.B., Bania, T.M. 1982, preprint "Some Observational Consequences of Residual HI in Galactic Molecular Clouds" Lynden-Bell, D. 1975, Vistas in Astronomy ]9, 299 Maeder, A. 1981, Astron. Astrophys. 102, 401 Maeder, A. 1982, Astron. Astrophys. 120, 113 Maihara, T., Oda, N., Sugiyama, T., Okuda, H. 1978, Publ. Astron. Soc. Japan 30, I Mathewson, D.S., Healy, J.R°, Rome, J.M. 1962, Austr. J. Phys. 15, 369 Mathis, J.S., Mezger, P.G., Panagia, N. 1983, Astron. Astrophys. in press, "Interstellar Radiation Field and Dust Temperatures in the Diffuse Interstellar Matter and in Giant Molecular Clouds" McClure, R.D., Twarog, B.A. 1978, IAU Coll. 45 "Chemical and Dynamical Evolution of our Galaxy", ed. E. Basinska-Grzesik, M. Mayor, p. 193 Meyer, J.-P. 1979, in "Les Elements et leurs Isotopes dans l'Univers", XXII e Coll. Internat. d'Astrophys., Liege, p. 153 Mezger, P.G. 1970, in IAU Symp. 38 "The Spiral Structure of our Galaxy", ed. W. Becker, G. Contopoulus, Reidel, p. 107 Mezger, P.G. 1974, Proc. of the ESO/SRC/CERN Conference on Research Programs for the New Large Telescopes, Geneva, p. 79 Mezger, P.G. 1978, Astron. Astrophys. 70, 565 Mezger, P.G. 1980, in "Radio Recombination Lines", ed. P.A. Shaver, Reidel Publ. Co., p. 81 Mezger, P.G. 1982, Proc. XXVIth ESLAB-Symp., Toledo 1982, in press Mezger, P.G., Mmthis, J.S., Panagia, N. 1982, Astron. Astrophys. 105, 372 Mezger, P.G., Pankonin, V., Schmid-Burgk, J., Thum, C., Wink, J. 1979, Astron. Astrophys. 80, L3 Mezger, P.G., Pauls, T. 1978, IAU Symp . 84 "The Large Scale Characteristics of the Galaxy", ed. W.B. Burton, p. 357 Mezger, P.G., Smith, L.F. 1977, IAU Symp. 75 "Star Formation", ed. T. de Jong, A. Maeder, Reidel, p. 133 Mezger, P.G., Smith, L.F., Churchwell, E. 1974, Astron. Astrophys. 32, 269 Mezger, P.G., Wink, J.E. 1983, in "Proc. of ESO-Workshop on Primord~'al Helium", in press Miller, G.E., Scalo, J.M. 1978, Astrophys. J. Suppl. 41, 513 Miller, G.E., Scalo, J.M. 1978, P.A.S.P. 90, 506 Norman, C., Silk, J. ]980, in IAU Symp. 8~-"Interstellar Molecules", ed. B.H. Andrew, Reidel, p. 137 Oda, N., Maihara, T., Sugiyama, T., Okuda, H. 1979, Astron. Astrophys. 72, 309 Oort, J.H. 1970, Astron. Astrophys. 7, 381 Ostriker, J.P., Caldwell, J.A.R. |979, IAU Symp. 84 "The Large-Scale Characteristics of the Galaxy", ed. W.B. Burton, Reidel, p. 441 Pagel, B.E.J., Edmunds, M.G. 1981, Ann. Rev. Astron. Astrophys. 19, 77 Panagia, N. 1973, Astron. J. 78, 929 Panagia, N. 1980, in "Radio Recombination Lines", ed. P.A. Shaver, Reidel, p. 99 Peimbert, M. ]978, IAU Symp. 84 "The Large Scale Characteristics of the Galaxy", ed. W.B. Burton, Reidel, p. 84 Peimbert, M., Serrano, A. 1980, Rev. Mexicana ~, 9 Penzias, A.A. 1980, Science 208, 663 Penzias, A.A. 1981, Astrophys. J. 249, 518 Rieke, G.H., Lebovsky, A.J., Thompson, R.I., Low, F.J., Tokunaga, A.T. 1980, Astrophys. J. 238, 24 Remie, M~J, Lamers, H.J.G.L.M. 1983, preprint "Effective Temperatures and Radii of Luminous O and B stars". Renzini, A., Voli, M. 1981, Astron. Astrophys° 94, 175 Rohlfs, K., Kreitschmann, J. 1981, Astrophys. Space Sci. 79, 289 Ross, J.E., Aller, L.H. 1976, Science ]91, 1223 Salpeter, E.E. ]955, Astrophys. J. 121, 161 Salpeter, E.E. 1977, IAU Symp. 76 "Planetary Nebula", ed. Y. Terzian, Reidel, p. 333 Savage, B.D., Bohlin, R.C., Drake, J.F., Budich, W. 1977, Astrophys. J. 216, 291 Serra, G., Puget, J.L., Ryter, C.E. 1980, Astron. Astrophys. 84, 220 Shaver, P.A., HcGee, R.X., Newton, L.M., Danks~ A.C., Pottasch, S°R. 1983, preprint "The Galactic Abundance Gradient" Silk, J. 1978, in "Protostars and Planets", ed. T. Gehrels, Tucson

224

R. GHsten and P.G. Mezger

Silk, J. 1980, in "Star Formation", IOth Advanced Course S.S.A.A., Saas-Fee Shu, F.H. 1978, IAU Symp. 77 "Structure and Properties of Nearby Galaxies", ed. E.M. Berkhuijsen, R. Wielebinskl, Reidel, p. 139 Shu, F.H., Milione, V., Roberts, W.W. 1973, Astrophys. J. 183, 819 Smith, L.F., Biermann, P., Mezger, P.G. 1978, Astron. Astrophys. 66, 65 Solomon, P.M., Sanders, D.B. 1980, in "Giant Molecular Clouds in the Galaxy", ed. P.M. Solomon, A.G. Edmunds, Pergamon Press, p. 41 Solomon, P.M., Scoville, N.Z., Sanders, D.B. 1979, Astrophys. J. 232, L89 Solomon, P.M., Stark, A.A., Sanders, D.B. 1983, Astrophys. J. Letter 267, L29 Stark, A.A. 1982, in "Kinematics, Dynamics and Structure of the Milky-'W~y", ed. W.L.H. Shuter, Reidel, p. 127 Stothers, R. 1972, Astrophys. J. 175, 431 Talbot, R.J. Jr. 1980, Astrophys. J. 235, 821 Thielemann, F.-K., Arnould, M. 1979, in "Les Elements et leurs Isotopes dans l'Univers", XXII e Coll. Internat. d'Astrophys., Liege, p. 59 Thum, C., Mezger, P.G., Pankonin, V. 1980, Astron. Astrophys. 87, 269 Tholine, J.E. 1980, Astrophys. J. 239, 417 Tinsley, B.M. 1980, Fundamentals of Cosmic Physics ~, 287 Tinsley, B.M., Larson, R.B. 1978, Astrophys. J. 221, 554 Tosi, M. 1982, Astrophys. J. 254, 699 Truran, J.W. 1977, in "CNO Isotopes in Astrophysics", ed. J. Audouze, Reidel, Dordrecht, p. 145 Twarog, B.A. 1980, Astrophys. J. 242, 242 Vader, J.P., de Jong, T. 1981, Astron. Astrophys. IOO, 124 Vidal-Madjar, A., Laurent, C., Gry, C., Bruston, P., Ferlet, R., York, D.G. 1983, Astron. Astrophys. 120, 58 Wannier, P.G. 1980, Ann. Rev. Astron. Astrophys. 18, 399 Wannier, P.G., Linke, R.A., Penzias, A.A. 1981, Astrophys. J. 247, 522 Wasserburg, G.J., Papanastassiou, D.A., Typhoon Lee 1979, in "Les Elements et leurs Isotopes dans l'Universe", XXII e Coll. Internat. d'Astrophysique., Liege Watson, W.D. 1978, Ann. Rev. Astron. Astrophys. 16, p. 585 Westerhout, G. 1958, Bull. Astron. Inst. Neth. 14, 215 Wilson, T. 1980, in "Radio Recombination Lines", ed. P.A. Shaver, p. 205 Wood, P.R. 1981, Proc. ASA ~, 145 Wood, P.R., Bessell, M.S., Fox, M.W. 1981, Proc. ASA ~, 203 Woosley, S.E., Weaver, T.A. 1980, in "Nuclear Astrophysics", ed. C. Barnes, D. Clayton, D. Schramm, Cambridge Wooten, A., Evans, V.J., Snell, R., Vanden Bout, P. 1978, Astrophys. J. 225, L143 York, D.G. 1982, Ann. Rev. Astron. Astrophys. 20, 221