Articles

ASYMMETRIC ABSORPTION PROFILES OF Lyα AND Lyβ IN DAMPED Lyα SYSTEMS

Published 2013 July 15 © 2013. The American Astronomical Society. All rights reserved.
, , Citation Hee-Won Lee 2013 ApJ 772 123 DOI 10.1088/0004-637X/772/2/123

0004-637X/772/2/123

ABSTRACT

Damped Lyα systems observed in the quasar spectra are characterized by a high neutral hydrogen column density, $N_{{\rm H\,\scriptsize{I}}}>2\times 10^{20}{\rm \ cm^{-2}}$. The absorption wing profiles are often fitted using the Voigt function due to the fact that the scattering cross section near the resonant line center is approximately described by the Lorentzian function. Since a hydrogen atom has infinitely many p states that participate in the electric dipole interaction, the cross section starts to deviate from the Lorentzian in an asymmetric way in the line wing regions. We investigate this asymmetry in the absorption line profiles around Lyα and Lyβ as a function of the neutral hydrogen column density $N_{{\rm H\,\scriptsize{I}}}$. In terms of Δλ ≡ λ − λα, we expand the Kramers–Heisenberg formula around Lyα to find σ(λ) ≃ (0.5f12)2σT(Δλ/λα)−2[1 + 3.792(Δλ/λα)], where f12 and σT are the oscillator strength of Lyα and the Thomson scattering cross section, respectively. In terms of Δλ2 ≡ λ − λβ in the vicinity of Lyβ, the total scattering cross section, given as the sum of cross sections for Rayleigh and Raman scattering, is shown to be σ(λ) ≃ σT(0.5f13)2(1 + R0)(Δλ2β)−2[1 − 24.68(Δλ2β)] with f13 and the factor R0 = 0.1342 being the oscillator strength for Lyβ and the ratio of the Raman cross section to Rayleigh cross section, respectively. A redward asymmetry develops around Lyα, whereas a blue asymmetry is obtained for Lyβ. The absorption center shifts are found to be almost proportional to the neutral hydrogen column density.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Hydrogen is the most abundant element in the universe and therefore is an important probe for the physical conditions of the intergalactic medium throughout the observable universe. High-resolution spectroscopy of quasars shows a large number of absorption lines blueward of Lyα mostly due to neutral hydrogen components in the intergalactic medium intervening along the line of sight. These quasar absorption systems due to hydrogen are classified according to the H i column densities. Quasar absorption systems with a neutral column density not exceeding $N_{{\rm H\,\scriptsize{I}}}\simeq 10^{17}{\rm \ cm^{-2}}$ constitute the Lyα forest which is attributed to the residual hydrogen atoms contained in intergalactic filamentary structures that are highly ionized (e.g., Rauch 1998; Meiksin 2009; Kim et al. 2011). The absorption systems associated with a neutral hydrogen column density in excess of $N_{{\rm H\,\scriptsize{I}}}=10^{20.3}{\rm \ cm^{-2}}$ are called damped Lyα (DLA) systems, which are distinguished from other quasar absorption systems in that they are dominantly neutral (e.g., Wolfe et al. 1986, 2005).

A catalog of 322 DLAs was provided by Curran et al. (2002), and recently 721 DLAs are listed from the Sloan Digital Sky Survey Data Release 7 (Khare et al. 2012). The exact nature of the DLAs is still controversial, but they are believed to dominate the neutral gas content in the universe, providing raw material for star formation during most of the time from the reionization era to the present (e.g., Wolfe et al. 2005; Prochaska et al. 2005). DLAs are also important to trace the chemical evolution of the universe by carefully measuring the metal abundance as a function of redshift (e.g., Calura et al. 2003; Rafelski et al. 2012). Accurate atomic physics for Lyα is essential to obtain a reliable estimate of metallicity of a DLA.

Recently, Kulkarni et al. (2012) reported a discovery of a "super-damped" Lyα absorber at zabs = 2.2068 toward QSO Q1135-0010. Their profile fit to the DLA showed a high neutral hydrogen column density of $N_{{\rm H\,\scriptsize{I}}}=10^{22.05}{\rm \ cm^{-2}}$. A little higher column density of $N_{{\rm H\,\scriptsize{I}}}=10^{22.10}{\rm \ cm^{-2}}$ was reported for the same object by Noterdaeme et al. (2012), who also suggested a significant star formation rate of 25 M yr−1 based on the strength of Hα.

Damped Lyα absorption profiles are analyzed using the Voigt function, which is defined as a convolution of a Lorentzian function with a Gaussian function (e.g., Rybicki & Lightman 1979). When $N_{{\rm H\,\scriptsize{I}}}$ is very large, the absorption at the core part is almost complete. Contributing only to the core part of the Voigt function, the Gaussian function does not play an important role in the analysis of high column density systems. In the wing part, the Voigt function coincides with the Lorentzian, and in this case the damping term in the denominator of the Lorentzian is quite negligible. Therefore, the wing profile is essentially proportional to Δλ−2, where Δλ = λ − λα is the difference in wavelength from the Lyα line center wavelength λα = 1215.671 Å.

The Lorentzian function, invoked to describe the wing parts of the absorption profile, is obtained when the scattering atom is regarded as a two-level atom. Instead of being a two-level system, the hydrogen atom has infinitely many energy levels, and therefore the scattering cross section is expected to deviate from the Lorentzian function that is symmetric with respect to the line center wavelength. Peebles (1993) discussed the resonance line shape that deviates from the simple Lorentzian and mentioned the marginal possibility of detecting this deviation in the absorption line profiles of DLA systems (e.g., Peebles 1993, p. 573). The formula for the scattering cross section he introduced is derived in a heuristic way to illustrate the behavior of the Lorentzian profile near line resonance and the classical ω4-dependence in the low-energy regime. However, this formula is inaccurate because it fails to include the contributions from the infinitely many p states that participate in the electric dipole interaction.

The exact scattering cross section is computed by summing all the probability amplitudes contributed from the infinitely many bound and free p states. The result is summarized in the Kramers–Heisenberg formula (e.g., Bethe & Salpeter 1967; Sakurai 1967). An expansion of the Kramers–Heisenberg formula around Lyα in frequency space was given by Lee (2003), in which the redward asymmetry of the scattering cross section around Lyα was briefly illustrated. However, spectroscopy is often presented in wavelength space rather than in frequency space so that it will also be useful to express the Kramers–Heisenberg formula in wavelength space. In this paper, we present the same expansion in wavelength space and quantify the redward shift of the center wavelength as a function of $N_{{\rm H\,\scriptsize{I}}}$. In addition, we also expand the Kramers–Heisenberg formula in the vicinity of Lyβ in order to investigate the asymmetry around Lyβ.

In the case of Lyβ, there is an additional scattering channel, which results from radiative de-excitation into the 2s state re-emitting an Hα line photon. This inelastic or Raman scattering branch is proposed to be important in the formation of broad Hα wings observed in young planetary nebulae and symbiotic stars (e.g., Isliker et al. 1989; Lee 2000; Schmid 1989). In this paper, we show that the cross section around Lyβ is asymmetric blueward, which is in high contrast with the behavior around Lyα. A brief discussion on the observational consequences is presented.

2. CALCULATION

2.1. Scattering Cross Section Around Lyα in Wavelength Space

The scattering cross section is given by the Kramers–Heisenberg formula that is obtained from a second-order time-dependent perturbation theory (e.g., Sakurai 1967; Merzbacher 1970). In terms of the matrix elements of the dipole operator, the Kramers–Heisenberg formula can be written as

Equation (1)

where me is the electron mass and r0 = e2/mec2 is the classical electron radius. The polarization vectors of incident and scattered radiation are denoted by e(α) and ${\bf e}^{(\alpha ^{\prime })}$, respectively. Here, ωI1 is the angular frequency between the intermediate state I and the ground 1s state. The intermediate state I that participates in the electric dipole interaction of Lyman photons consists of bound np states and free n'p states. In the atomic units adopted in this work, the bound np state has the energy eigenvalue En = −1/(2n2) and correspondingly ωI1 = ωn1 = (1 − n−2)/2. Similarly, for the n'p state, $E_{n^{\prime }}=1/(2n^{\prime 2})$ and $\omega _{I1}=\omega _{n^{\prime }1}=(1+n^{\prime -2})/2$. The term denoted by (r · e(α))I1 represents the matrix element of the position operator between the intermediate state I and the ground 1s state.

Being a two-body system, the hydrogen atom admits analytically closed expressions of the wave functions, which enables one to compute explicitly the matrix elements of the dipole operator in terms of the confluent hypergeometric function. A typical matrix element (r · epsilon(α))I1 for an intermediate state with I = |np, m > is explicitly written as

Equation (2)

where $Y_l^m(\theta,\phi)$ is a spherical harmonic function and Rnl(r) is the radial wave function given by an associated Laguerre function. The angular integration followed by summing over magnetic substates m = ±1, 0 of each np state and averaging over polarization states of incident and outgoing radiation results in a numerical factor of 8π/3, which is discussed in more detail in the Appendix.

The radial matrix elements of the dipole operators 〈rn1 and $\langle r\rangle _{n^{\prime },1}$ are readily found in many textbooks on quantum mechanics (e.g., Berestetskii et al. 1971; Bethe & Salpeter 1967). The radial matrix elements for the bound np states are given by

Equation (3)

where aB = ℏ2/me2 = 0.5292 Å is the Bohr radius. For the continuum n'p states, the corresponding values are given by

Equation (4)

which is obtained through analytic continuation into the complex plane (Bethe & Salpeter 1967; Saslow & Mills 1969).

Lee (2003) provided the expansion of the Kramers–Heisenberg formula for Rayleigh scattering in the vicinity of Lyα in terms of Δω = ω − ω21. The result is summarized as

Equation (5)

where $\sigma _T = 8\pi r_0^2/3=0.6652\times 10^{-24}{\rm \ cm^2}$ is the Thomson scattering cross section and f12 = 0.4162 is the oscillator strength for the Lyα transition. The coefficients ak were numerically computed by Lee (2003), who gave

Equation (6)

In Table 1, we show these coefficients up to a5.

Table 1. Expansion Coefficients of the Rayleigh Scattering Cross Section around Lyα and Rayleigh and Raman Scattering Cross Sections around Lyβ

Lyα Lyβ Lyβ (Raman)
a1 = −8.961 × 10−1 b1 = 1.621 × 101 c1 = −2.776 × 101
a2 = −1.222 × 101 b2 = −4.299 × 102 c2 = −2.128 × 102
a3 = −5.252 × 101 b3 = −2.176 × 103 c3 = −3.231 × 103
a4 = −2.438 × 102 b4 = −6.005 × 104 c4 = −1.098 × 105
a5 = −1.210 × 103 b5 = −8.414 × 105 c5 = −4.032 × 106

Download table as:  ASCIITypeset image

Up to the first order of Δω/ω12, the scattering cross section can be expressed as

Equation (7)

Here, we introduce the characteristic cross section σα for Lyα defined as σα ≡ σT(f12/2)2 = 2.880 × 10−26 cm2. The coefficient a1, being less than zero, is responsible for the redward asymmetric deviation of the scattering cross section in frequency space.

Astronomical spectroscopy is often presented in wavelength space, which makes it necessary to express the scattering cross section in terms of Δλ = λ − λα, the difference in wavelength from the Lyα line center. From the following relation,

Equation (8)

we may notice that expansion in wavelength space will yield different coefficients from those obtained in expansion in frequency space. Substituting Equation (8) into Equation (7), we obtain

Equation (9)

Thus, up to first order, we have

Equation (10)

from which it is found that the coefficient 3.792 is significantly different from the coefficient −1.792 in frequency space.

The red asymmetry in the Lyα scattering cross section can be explained by noting the denominator ωn1 − ω in Equation (1), where the dominant contribution comes from n = 2. The contributions from all the excited states with n > 2 interfere positively with the dominant contribution from n = 2 for photons on the red side whereas the interference is negative for photons on the blue side. Therefore, the red wing is strengthened relative to the pure Lorentzian profile by the positive interference of scattering from all other levels.

2.2. Rayleigh Scattering Cross Section Around Lyβ

In a similar way, we define the difference in frequency from Lyβ by

Equation (11)

and expand the Kramers–Heisenberg formula in terms of Δω231. We note that the terms appearing in the Kramers–Heisenberg formula include

Equation (12)

and

Equation (13)

For n ⩾ 2, the second term in the summation in Equation (1) can be written as

Equation (14)

Similar expressions for the continuum states are obtained in a straightforward manner.

We expand the Kramers–Heisenberg formula in frequency space for Rayleigh scattering near Lyβ, which is written as

Equation (15)

Here, angular integration has been performed and the atomic unit system is adopted, in which the Bohr radius aB = 1.

An algebraic rearrangement of Equation (15) can be made to express σRay(ω) as

Equation (16)

The coefficients B0 and B1 are determined through the following relations:

Equation (17)

where f13 = 0.07910 is the oscillator strength for the Lyβ transition. The coefficients Bk for k ⩾ 2 are given as

Equation (18)

The numerical values of the coefficients up to k = 5 are computed as follows:

Equation (19)

In particular, the coefficient b1 is positive due to the predominant contribution from the 2p state.

Therefore, up to the first-order approximation in Δω231, the Rayleigh scattering cross section around Lyβ is given by

Equation (20)

In wavelength space, the Rayleigh scattering cross section in the vicinity of Lyβ with the line center λβ = 1025.722 Å can be expanded as

Equation (21)

where Δλ2 ≡ λ − λβ is the wavelength deviation from the Lyβ center.

The negative value of the coefficient 2(1 − b1) implies that the Rayleigh scattering cross section near Lyβ is asymmetric to the blue of Lyβ, which is in high contrast with the behavior around Lyα. In the case of Lyβ, the main contributor to the Kramers–Heisenberg formula is the 3p state and the residual contribution comes from the 2p state and all the p states lying higher than the 3p state. The contribution to σ(λ) of a given np or n'p state is measured roughly by the oscillator strength inversely weighted by the energy difference from Lyβ. Therefore, the contribution of the 2p state is more important than that from all the p states lying higher than the 3p state. Hence, in the case of Lyβ, the n = 2 contribution has lower frequency and the interference with the principal n = 3 scattering contribution is negative on the red side and positive on the blue side, which explains the blue asymmetric scattering cross section.

2.3. Raman Scattering Cross Section Around Lyβ

Interaction with a hydrogen atom of electromagnetic radiation around Lyβ has another channel, which is inelastic or Raman scattering. The scattering hydrogen atom de-excites into the 2s state re-emitting an Hα photon into another line of sight, which provides an important contribution to the absorption profile around Lyβ. The astrophysical importance of Raman scattering can be appreciated in the emission features at 6830 Å and 7088 Å that appear in the spectra of about a half of symbiotic stars. These are formed through Raman conversion of the resonance doublet O vi λλ1032,1038 (Schmid 1989; Nussbaumer et al. 1989). Another example of Raman scattering by atomic hydrogen is provided by far-UV He ii emission lines in symbiotic stars and young planetary nebulae (e.g., Birriel 2004; Lee et al. 2006; Lee 2012). It has also been proposed that broad Hα wings often found in planetary nebulae and symbiotic stars are formed through Raman scattering of far-UV continuum around Lyβ (e.g., Lee 2000; Arrieta & Torres-Peimbert 2003).

As is illustrated in Sakurai (1967), the term corresponding to the "seagull graph" is absent in the Kramers–Heisenberg formula for the case of Raman scattering. This difference allows an alternate expression of the Kramers–Heisenberg formula given in terms of the matrix elements of the momentum operator (see also Saslow & Mills 1969; Lee & Lee 1997). In a manner analogous to what is illustrated in the Appendix, taking angular integrations, summing over magnetic substates, and averaging over polarizations of incident and outgoing radiation, we arrive at an explicit expression of the Raman cross section given by

Equation (22)

Here ω' = ω − ω21 is the angular frequency of the Raman scattered radiation. The matrix element 〈pn1 associated with the momentum operator between the np and the 1s states is given by

Equation (23)

Here, an atomic unit system is adopted and the reality of the radial wave functions is noted.

The matrix element 〈pn2 corresponds to the transition between the np and 2s states, which is explicitly given by

Equation (24)

The contribution from the continuum n'p states is obtained by considering the matrix elements of the momentum operator given by

Equation (25)

Due to the vanishing matrix element 〈pn2 for n = 2, the sum in Equation (22) begins from n = 3 for the bound np states, which implies that the 2p state does not contribute to the cross section for Raman scattering around Lyβ. This is decisively important to the behavior of the cross section, as we discuss later in more detail.

The terms involving angular frequencies can be rearranged using the following relation,

Equation (26)

where ω32 = ω31 − ω21 is the angular frequency for Hα. Use is also made of the following relations:

Equation (27)

The Kramers–Heisenberg formula for Raman scattering near Lyβ can be expanded in frequency space as follows:

Equation (28)

where the coefficients Ck are given by

Equation (29)

and

Equation (30)

for k ⩾ 2. Here, $f_{2s,3p}={(2/3)}\omega _{32}^{-1}[\langle p\rangle _{32}]^2=0.4349$ is the oscillator strength between the 2s and 3p states (e.g., Bethe & Salpeter 1967).

Therefore, in frequency space the Raman scattering cross section is written as

Equation (31)

In wavelength space, we obtain

Equation (32)

where the lower case coefficients ci are defined by ci = Ci/C0. The numerical values of these coefficients up to c5 are shown in Table 1.

Unlike the case for Rayleigh scattering near Lyβ, the Raman scattering cross section shows a redward asymmetry with respect to the Lyβ center. This result can be traced to the fact that no contribution is made from the 2p state, which played a dominant role in the case of Rayleigh scattering near Lyβ. In the absence of the 2p contribution, all the perturbing p states are more energetic than Lyβ. This situation is exactly the same as the Rayleigh scattering around Lyα resulting in a redward asymmetry.

The ratio rb(λ) of the cross sections for Raman scattering to Rayleigh scattering in the vicinity of Lyβ is given by

Equation (33)

where the first three coefficients are explicitly R0 = (5|C0|2/32)/(f13/2)2 = 2185−9 = 0.1342, R1 = 11.26, and R2 = 535.9. This result shows discrepancy with that provided by Yoo et al. (2002), in which there is an error in their numerical calculation of the coefficients A3 and A4. The leading term can also be expressed as R0 = (f2s, 3p/f13)(ω3231)2 = 0.1342, which implies that the branching ratio is determined by a combination of the oscillator strength and the phase-space volume factor represented by ω2. It is seen that the dominant contribution is made by the phase-space volume available to scattered radiation.

In Figure 1, we show the branching ratio rb(λ) in the neighborhood of Lyβ. The solid line shows the result from a direct numerical computation of the Kramers–Heisenberg formula. The dotted line shows the linear fit and the dot-dashed line shows the second-order fit using Equation (33). Because the coefficient R2 is large, the nonlinearity of rb(λ) is quite conspicuous in the figure. This behavior leads to a redward shift in broad Hα wings observed in young planetary nebulae and symbiotic stars which are also attributed to Raman scattering of Lyβ (Jung & Lee 2004).

Figure 1.

Figure 1. Ratio of cross sections of Raman scattering and Rayleigh scattering around Lyβ. The solid line shows the result from the full numerical calculation of the Kramers–Heisenberg formula. The dotted line shows the linear fit and the dashed line shows the quadratic fit.

Standard image High-resolution image

2.4. Total Scattering Cross Section Around Lyβ

In this subsection, we combine the results of previous subsections to provide the expansion of the total scattering cross section around Lyβ. The total scattering cross section around Lyβ is the sum of Equations (20) and (31), which is, to the first order of Δω231, given by

Equation (34)

Here, we introduce another parameter σβ defined by

Equation (35)

In wavelength space, we may combine Equations (21) and (32) to express the total scattering cross section around Lyβ as

Equation (36)

From this result, it is seen that the Lyβ absorption profiles tend to shift blueward of the Lyβ line center. In Figure 2, we show the total scattering cross section obtained from a numerical evaluation of the Kramers–Heisenberg formula around Lyα and Lyβ in wavelength space. The vertical axis shows the logarithm to the base 10 of σ(λ) in units of cm2.

Figure 2.

Figure 2. Scattering cross section around Lyα (upper panel) and Lyβ (lower panel). The horizontal axis shows the wavelength difference from the line center and the vertical axis represents the logarithm of the cross section in units of cm2.

Standard image High-resolution image

In order to take a clear view of the asymmetry of the scattering cross section, we plot the same quantities in Figure 3 as a function of the absolute value of the wavelength deviation. The cross sections redward of Lyα and Lyβ are shown with solid lines in the upper panel and the lower panel, respectively. The dotted lines show the cross sections blueward of Lyα and Lyβ. The dotted lines are mirror images of the curves blueward of Lyα and Lyβ shown in Figure 2. In Figure 3, we see that red Lyα photons have a larger scattering cross section than blue counterparts and that the opposite is the case for Lyβ.

Figure 3.

Figure 3. Scattering cross section around Lyα (upper panel) and Lyβ (lower panel) as a function of the absolute value of the wavelength difference. The solid lines show the cross section redward of Lyα (upper panel) and Lyβ (lower panel). The cross sections blueward of Lyα and Lyβ are shown in dotted lines in the upper panel and the lower panel, respectively.

Standard image High-resolution image

In Figure 4, we show the transmission coefficient $t(\lambda,N_{{\rm H\,\scriptsize{I}}})$ defined by

Equation (37)

for various neutral hydrogen column densities. The upper panel is for Lyα and the lower panel is for Lyβ. The solid line shows the result for $N_{{\rm H\,\scriptsize{I}}}=10^{20}{\rm \ cm^{-2}}$, the dotted line for $N_{{\rm H\,\scriptsize{I}}}=10^{21}{\rm \ cm^{-2}}$, and the dashed line for $N_{{\rm H\,\scriptsize{I}}}=5\times 10^{21}{\rm \ cm^{-2}}$. In the case of $N_{{\rm H\,\scriptsize{I}}}=5\times 10^{21}{\rm \ cm^{-2}}$, the asymmetry is quite noticeable in the scale shown in the figure.

Figure 4.

Figure 4. Absorption profiles around Lyα (upper panel) and Lyβ (lower panel) for various neutral hydrogen column densities.

Standard image High-resolution image

3. ASYMMETRY IN THE ABSORPTION PROFILES OF Lyα AND Lyβ

3.1. Absorption Center Shift

In this subsection, we investigate the shift of the absorption line center near Lyα and Lyβ as a function of the neutral hydrogen column density. Denoting by λc the line center wavelength of Lyα or Lyβ, the scattering cross section is approximated to the first order of dimensionless wavelength deviation from the line center x = Δλ/λc by a function

Equation (38)

The equation f(x) = k > 0 has two solutions x1, x2, of which the mean is xm = a/(2k). This implies that the absorption center can be meaningfully defined when we fix the value of cross section. The sign of the coefficient a determines the direction of asymmetry, where a positive a results in a red asymmetry.

Given a value of the H i column density $N_{{\rm H\,\scriptsize{I}}}$ we define the mean wavelength $\lambda _{m1}^\alpha$ of the two wavelengths λ1, 2 at which $\tau (\lambda _1)=\tau (\lambda _2)=\sigma (\lambda)N_{{\rm H\,\scriptsize{I}}}=1$ around Lyα and in a similar way we define $\lambda _{m1}^\beta$ for Lyβ. We also introduce $\lambda _{m2}^{\alpha }$ and $\lambda _{m2}^\beta$ as the mean value of the two wavelengths $\lambda _{1,2}^{\prime }$ where we have $\tau (\lambda _{1}^{\prime })=\tau (\lambda _2^{\prime })=0.5$ around Lyα and Lyβ, respectively. Corresponding to these wavelengths λm1, we define the velocity shift ΔV1 by the relation

Equation (39)

for Lyα and in a similar way $\Delta V^\beta _1$ is defined for Lyβ. Here, c is the speed of light.

In Table 2, we show the values of $\lambda _{m1}^\alpha$ and $\lambda _{m2}^\alpha$ for various neutral hydrogen column densities. Also in Table 3 we show the quantities corresponding to the Lyβ transitions. At $N_{{\rm H\,\scriptsize{I}}}=10^{21}{\rm \ cm^{-2}}$ we obtain a redward center shift in the amount of Δλ = +16 km s−1 for Lyα and a blueward shift of Δλ = −3.9 km s−1 for Lyβ.

Table 2. Absorption Center Shifts Around Lyα for Various Neutral Hydrogen Column Densities

log $N_{{\rm H\,\scriptsize{I}}}$ $\lambda _{m1}^\alpha$ $\Delta V_1^\alpha$ $\lambda _{m2}^\alpha$ $\Delta V_2^\alpha$
(Å) (km s−1) (Å) (km s−1)
19.0 6.10E-04 0.151 1.34E-03 0.331
19.7 3.30E-03 0.813 6.47E-03 1.60
20.0 6.47E-03 1.60 1.32E-02 3.25
20.7 3.26E-02 8.04 6.51E-02 16.0
21.0 6.51E-02 16.0 1.30E-01 32.1
21.7 3.25E-01 80.2 6.51E-01 1.60E+02
22.0 6.51E-01 1.60E+02 1.31 3.20E+02
22.7 3.25 8.01E+02 6.48 1.60E+03

Download table as:  ASCIITypeset image

Table 3. Absorption Center Shifts Around Lyβ for Various Neutral Hydrogen Column Densities

log $N_{{\rm H\,\scriptsize{I}}}$ $\lambda _{m1}^\beta$ $\Delta V_1^\beta$ $\lambda _{m2}^\beta$ $\Delta V_2^\beta$
(Å) (km s−1) (Å) (km s−1)
19.0 −1.22E-04 −0.03568 −2.44E-04 −0.07136
19.7 −7.32E-03 −0.214 −1.34E-03 −0.392
20.0 −1.34E-03 −0.392 −2.69E-02 −0.785
20.7 −6.59E-03 −1.93 −1.34E-02 −3.92
21.0 −1.34E-02 −3.92 −2.67E-02 −7.81
21.7 −6.64E-02 −19.4 −1.32E-01 −38.5
22.0 −1.32E-01 −38.5 −2.60E-01 −75.9
22.7 −6.22E-01 −1.82E+02 −1.16 −3.38E+02

Download table as:  ASCIITypeset image

In Figure 5, we show $\Delta V_1^\alpha$ and $\Delta V_2^\alpha$ as a function of $N_{{\rm H\,\scriptsize{I}}}$ in the cases of Lyα (upper panel) and $\Delta V^\beta _1$ and $\Delta V^\beta _2$ for Lyβ (lower panel). The dotted line shows a fit to the data, which implies that both $\Delta V_{1}^{\alpha,\beta }$ and $\Delta V_{2}^{\alpha,\beta }$ are proportional to $N_{{\rm H\,\scriptsize{I}}}$. The linear fit shown by the dotted line for Lyα in the figure is given by

Equation (40)

and similarly for Lyβ it is given by

Equation (41)
Figure 5.

Figure 5. Center shift of the absorption profile around Lyα (upper panel) and Lyβ (lower panel). The horizontal axis shows the logarithm of the H i column density. The squares show the absorption center defined by the mean values of the wavelengths, where τ(λ) = 1. The triangles show the absorption center similarly defined by the condition τ(λ) = 0.5.

Standard image High-resolution image

3.2. Profile Fitting by Shifting the Lorentzian

Another way of quantifying the asymmetry is provided by fitting the absorption profiles. In this subsection, we compare the transmission coefficient $t(\lambda,N_{{\rm H\,\scriptsize{I}}})$ derived from the Kramers–Heisenberg formula with that obtained from the Lorentzian shifted by a finite amount. For simplicity, we fix the H i column density $N_{{\rm H\,\scriptsize{I}}}=5\times 10^{21}{\rm \ cm^{-2}}$. This procedure may illustrate an error estimate in determining the redshift of a DLA system with $N_{{\rm H\,\scriptsize{I}}}=5\times 10^{21}{\rm \ cm^{-2}}$.

In Figure 6 we show the result for Lyα. The solid line in each panel shows the transmission coefficient obtained from the Kramers–Heisenberg formula. The dotted line in the top panel shows the transmission coefficient from the Lorentzian function, which provides an excellent fit near the line center. However, a considerable deviation in the wing part is quite noticeable. With the dotted line in the bottom panel, we show the quantities obtained by shifting the Lorentzian redward by an amount of +0.8 Å. Improvement of the fitting in wing parts is achieved only at the expense of a poor approximation near the line center. In the middle panel, we show the Lorentzian shifted by +0.4 Å, in which the quality of the fit is compromised between the previous two cases.

Figure 6.

Figure 6. Fit to the absorption profile around Lyα using shifted Lorentzian functions. The solid line is the transmission probability for a neutral slab of hydrogen with $N_{{\rm H\,\scriptsize{I}}}=5\times 10^{21}{\rm \ cm^{-2}}$.

Standard image High-resolution image

In the analysis by Lee (2003), the optimal wavelength shift was proposed by +0.2 Å, which is smaller than +0.4 Å suggested in this work. This discrepancy is due to the fact that the fitting procedure in Lee (2003) was confined to a rather narrow interval of |Δλ| < 34 Å excluding extreme wing parts. The procedure of fitting a DLA profile using the shifted Lorentzian tends to overestimate the line center wavelength of Lyα leading to corresponding overestimate of the redshift of the DLA. We note that Lee (2003) made a mistake in pointing out that the redshift would be "underestimated," which should be corrected to be "overestimated."

A similar analysis corresponding to Lyβ is shown in Figure 7 for the same neutral hydrogen column density $N_{{\rm H\,\scriptsize{I}}}=5\times 10^{21} {\rm \ cm^{-2}}$. In the figure, the dotted line in each panel shows the transmission coefficient from the Lorentzian function (top panel) and shifted Lorentzian functions (middle and bottom panels), whereas the solid line shows the exact transmission coefficient computed from the Kramers–Heisenberg formula. The amount of wavelength shift blueward of Lyβ is Δλ = −0.1 Å and −0.2 Å for the middle and bottom panels, respectively. As in the case of Lyα illustrated in Figure 6, the unshifted Lorentzian gives an excellent fit to the core part of the absorption profile whereas the bottom panel shows an improved fit to the wing parts with the loss of fitting quality at the core part.

Figure 7.

Figure 7. Fit to the absorption profile around Lyβ using shifted Lorentzian functions. The solid line is the transmission probability for a neutral slab of hydrogen with $N_{{\rm H\,\scriptsize{I}}}=5\times 10^{21}{\rm \ cm^{-2}}$.

Standard image High-resolution image

In Figure 8, we present the transmission coefficients using the Kramers–Heisenberg formula and the Lorentzian functions around Lyα and Lyβ in the wavelength interval between 980 Å and 1400 Å for a very thick H i medium with $N_{{\rm H\,\scriptsize{I}}}=5\times 10^{22}{\rm \ cm^{-2}}$. This kind of an extreme neutral hydrogen column density has been found toward the gamma-ray burst GRB080607 (e.g., Prochaska et al. 2009). For comparison, we show the transmission coefficients obtained from the Lorentzian around Lyα and Lyβ by the dotted line and the dashed line, respectively. In this highly thick medium, the deviation from the Lorentzian is quite severe due to the contribution from higher order terms, which prevents one from obtaining satisfactory results by fitting the absorption profiles by a Voigt or equivalently a Lorentzian function.

Figure 8.

Figure 8. Absorption profile in the wavelength region between 980 Å and 1400 Å for a slab of a neutral hydrogen column density $N_{{\rm H\,\scriptsize{I}}}=5\times 10^{22}{\rm \ cm^{-2}}$. The dotted line shows the transmission coefficient obtained from the sum of the two Lorentzian functions around Lyα and Lyβ.

Standard image High-resolution image

In particular, in the wavelength range shown in Figure 8, the local peak transmission is found at λp = 1062 Å, for which $t(\lambda, N_{{\rm H\,\scriptsize{I}}}=5\times 10^{22} {\rm \ cm^{-2}})=0.0144$. However, the sum of two Lorentzian functions around Lyα and Lyβ admits a local maximum at λ = 1070 Å. This shows the inadequacy of using a Voigt function for fitting analyses in extended wing parts in the case of very high column density systems. Furthermore, at this high $N_{{\rm H\,\scriptsize{I}}}$, the blue wing region of Lyα overlaps with that of the red Lyβ wing, for which a full quantum mechanical formula should be invoked for an accurate analysis.

4. SUMMARY AND DISCUSSION

The Kramers–Heisenberg formula is expanded around Lyα and Lyβ in order to investigate the asymmetric deviation of the scattering cross section. A redward asymmetry is seen around Lyα and a blueward asymmetry is found around Lyβ. For red Lyα photons, the perturbing transitions from (n + n')p states (n ≠ 2) provide a positive contribution to the scattering cross section because they are in the same side as the 2p state in the energy space, resulting in a red asymmetry. In the case of Lyβ, Rayleigh scattering contributes more than Raman scattering by a factor 6.452. Raman scattering around Lyβ exhibits a red asymmetry like Lyα because all the perturbing transitions lie higher than the main transition. However, for Rayleigh scattering around Lyβ, the transition from the 2p state is the dominant perturbing transition, which is less energetic than Lyβ. This leads to a blue asymmetry in σ(λ) around Lyβ. In an attempt to quantify these asymmetries, we compute the mean wavelengths for which the scattering optical depth becomes unity or one-half for various values of H i column density $N_{{\rm H\,\scriptsize{I}}}$. Also we fitted the transmission coefficients for given $N_{{\rm H\,\scriptsize{I}}}$ by shifting the Lorentzian function.

Peebles (1993) introduced the formula for a resonance scattering cross section around Lyα

Equation (42)

which is often used in fitting wing profiles of Lyα (e.g., Miralda-Escude 1998). In particular, the red damping wing of Lyα is essential to probe the partially neutral intergalactic medium expected around the end of cosmic reionization (Gunn & Peterson 1965; Scheuer 1965; Mortlock et al. 2011). Neglecting the damping term in the denominator, this expression yields an expansion in frequency space

Equation (43)

In this expression, the coefficient of the first-order term is 4, which differs significantly from the correct value of −1.792. According to this formula, the scattering cross section is larger in the blue part of Lyα than in the red part, which is incorrect. The discrepancy in the expansion may be traced to the approximation adopted in the derivation of Equation (42), where the hydrogen atom is effectively treated as a two-level system.

The Lorentzian or Voigt profile matches the Kramers–Heisenberg profile excellently only in the core part. Therefore, the redshift will be measured reliably when the profile fitting is more weighted toward the deeply absorbed core part than far wing parts. With the accurate determination of the redshift and column density of the DLA, one may obtain reliable transmission coefficients using the Kramers–Heisenberg formula or its first-order approximation given in Equations (10) and (36).

In an analysis of a quasar spectrum, it is highly difficult to obtain the accurate continuum level due to intervening Lyα forest systems. Securing the quasar continuum level around the damped Lyα center with high precision is critical to verify the asymmetry presented in this work. With the advent of extremely large telescopes in the near future equipped with a high-resolution spectrometer, the accurate atomic physics will shed light on the physical conditions of the neutral hydrogen reservoir in the early universe.

The author is very grateful to the anonymous referee whose comments greatly improved the presentation of this paper. This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science, and Technology (2011-0027069).

APPENDIX: ANGLE-AVERAGED CROSS SECTION

We show a detailed angular integration of the matrix element that constitutes the Kramers–Heisenberg formula. Because of the selection rule of the electric dipole interaction, the relevant states are the np and 1s states in the case of Rayleigh scattering. In the case of Raman scattering relevant to the interaction around Lyβ, the 2s state is also involved. However, as long as the angular and polarization average is concerned, the same calculation is performed.

A typical matrix element to be summed in the Kramers–Heisenberg formula is

Equation (A1)

where I denotes an intermediate state. In particular, I can be written as I = |np, m〉, where m is the magnetic quantum number taking one of zero and ±1 in the case of a p state. The spherical harmonic functions with l = 1 are explicitly defined by

Equation (A2)

from which we may set

Equation (A3)

Therefore given an intermediate state I = |np, m〉, we have

Equation (A4)

The wave function |np, m〉 is given by the product of the radial part Rn1(r) and the angular part $Y_1^m$, whereas the 1s state is characterized by the radial part R10(r) multiplied by the trivial spherical harmonic $Y_0^0=1/\sqrt{4\pi }$. Therefore, we have

Equation (A5)

Here, δm, n is the Kronecker delta and $\langle 1s|r|np\rangle \,{=} \int _0^\infty R_{10}(r) rR_{n1}(r) r^2 dr$ is the radial expectation value between the 1s and np states. In a similar way, for the operators y and z we have

Equation (A6)

From this, we note that

Equation (A7)

Given the np states, we sum over substates with m = ±1, 0 to obtain

Equation (A8)

As is well known for Thomson scattering (e.g., pp. 51 and 52 in Sakurai 1967), a numerical factor of 8π/3 results from averaging over polarization states for both incoming and outgoing radiation.

Please wait… references are loading.
10.1088/0004-637X/772/2/123