You are on page 1of 34

MNRAS 000, 1–22 (2018) Preprint 17 April 2019 Compiled using MNRAS LATEX style file v3.

Modern stellar spectroscopy caveats

Sergi Blanco-Cuaresma1?
1 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA

Accepted XXX. Received YYY; in original form ZZZ


arXiv:1902.09558v2 [astro-ph.SR] 15 Apr 2019

ABSTRACT
Multiple codes are available to derive atmospheric parameters and individual chemical
abundances from high-resolution spectra of AFGKM stars. Almost all spectroscopists
have their own preferences regarding which code and method to use. But the intrinsic
differences between codes and methods lead to complex systematics that depend on
multiple variables such as the selected spectral regions and the radiative transfer code
used. I expand iSpec, a popular open-source spectroscopic tool, to support the most
well-known radiative transfer codes and assess their similarities and biases when using
multiple set-ups based on the equivalent-width method and the synthetic spectral-
fitting technique (interpolating from a pre-computed grid of spectra or synthesizing
with interpolated model atmospheres). This work shows that systematic differences
on atmospheric parameters and abundances between most of the codes can be re-
duced when using the same method and executing a careful spectral feature selection.
However, it may not be possible to ignore the remaining differences, depending on
the particular case and the required precision. Regarding methods, equivalent-width-
based and spectrum-fitting analyses exhibit large differences that are caused by their
intrinsic differences, which is significant given the popularity of these two methods.
The results help to identify the key caveats of modern spectroscopy that all scientists
should be aware of before trusting their own results or being tempted to combine
atmospheric parameters and abundances from the literature.
Key words: stars: fundamental parameters – stars: abundances – stars: atmospheres
– techniques: spectroscopic

1 INTRODUCTION analysis (e.g., MATISSE Recio-Blanco et al. 2006, FERRE


Allende Prieto et al. 2006, ULySS Koleva et al. 2009, Starfish
The automation of high-resolution stellar spectral analy-
Czekala et al. 2015, sick Casey 2016), most of them base the
sis for AFGKM stars has become a necessity in recent
analysis on certain spectral features.
years owing to the enormous increase of publicly available
observations. Large surveys such as APOGEE (Eisenstein The derivation of atmospheric parameters and abun-
et al. 2011; Majewski et al. 2017) or the Gaia-ESO Public dances from stellar spectra can generally be accomplished
Spectroscopic Survey (GES; Gilmore et al. 2012; Randich by following either of two possible strategies: the equivalent-
et al. 2013), complemented by smaller surveys like OCCASO width method or the synthetic spectral-fitting technique.
(Casamiquela et al. 2016, 2017) plus other independent stud- The equivalent-width method first requires the measurement
ies and observational proposals, have contributed to this of the equivalent width of a selection of neutral and ionized
golden period of stellar spectroscopy. iron absorption lines. This is generally achieved by fitting a
Several research groups have developed codes to analyse Gaussian profile and then computing the width of the spec-
all these data, and some have made their work openly avail- tral continuum that has the same area as the absorption line.
able (e.g., SME Valenti & Piskunov 1996, GALA Mucciarelli Next, a radiative transfer code is used to derive the individ-
et al. 2013, FAMA Mucciarelli et al. 2013, StePar Tabernero ual line abundances for a given set of initial atmospheric
et al. 2013, iSpec Blanco-Cuaresma et al. 2014b, The Cannon parameters. The stellar parameters are found by flattening
Ness et al. 2015, ZASPE Brahm et al. 2017, FASMA Tsan- the abundance trends with respect to the reduced equiva-
taki et al. 2018). Except those methods that use the whole lent width, the lower excitation potential, and the ionization
spectrum (or a single continuous spectral region) for their stage. In the case of the synthetic spectral-fitting technique,
the observed spectrum is compared with theoretical spec-
tra that are synthesized on-the-fly or interpolated from pre-
? E-mail: sblancocuaresma@cfa.harvard.edu computed grids (in both cases, a radiative transfer code is

© 2018 The Authors


2 S. Blanco-Cuaresma
also necessary), and a minimization algorithm is executed. of different radiative transfer codes. For synthetic methods,
Frequently, a selection of spectral features is used instead of different line masks (i.e. the spectral region that includes the
the full spectrum in order to reduce the computation time target absorption lines) do not seem to lead to any signif-
and to concentrate on the more informative spectral regions. icant differences, but differences in broadening parameters
The above-mentioned codes generally follow one of these two do.
strategies. These studies have provided the first glimpses into the
In addition to the significant diversity of codes avail- discrepancies currently found in the literature for spectro-
able, all of them can be set up with many different com- scopic analysis. However, they are limited to a very low num-
binations of the necessary ingredients, such as the grid of ber of spectra, and the execution involved several different
model atmospheres (where interpolations are needed to de- research groups using their own codes with, sometimes, man-
rive the right model for the desired atmospheric parameters), ual operations. Despite the excellent coordination, it is easy
the reference solar abundances (these can be scaled up or to make mistakes when the analysis is not fully automatic
down following the desired metallicity, or certain elements or when a completely homogeneous analysis cannot be guar-
can be enhanced/depleted to follow certain patterns such anteed. In order to reveal the essential caveats of modern
as the enhancement of alpha elements observed for metal- spectroscopy it is necessary to have fully automatic and re-
poor stars), the radiative transfer code, the atomic data (e.g. liable tests that analyse homogeneously a higher number of
wavelengths, oscillator strengths, line-broadening parame- spectra covering a wider range of stars.
ters such as the radiative/Stark/van der Waals damping pa- In this work, I have extended iSpec1 (Blanco-Cuaresma
rameters), the selection of spectral features used (e.g. some et al. 2014b) by: (1) including a large number of radia-
regions may carry more information than others, or it may tive transfers codes widely used for spectral synthesis and
be that some codes and models are better at reproducing equivalent-width analysis (Blanco-Cuaresma et al. 2017;
certain regions), or the continuum normalization procedure. Blanco-Cuaresma 2017); (2) adding spectral interpolation
These inhomogeneities have led to large discrepancies in the capabilities, which allows the user to use/compare a different
atmospheric parameters and abundances present in the lit- spectroscopic approach (i.e. spectral interpolation instead of
erature (Hinkel et al. 2014). model atmosphere interpolation). iSpec has become a useful
The source of this problem was first explored in Hinkel tool for spectroscopic analysis and also a very convenient
et al. (2016), where four spectra were analysed using six framework in which to discover and assess the caveats of
different codes with a subsequent re-analysis with common modern stellar spectroscopy. Using this tool, I designed sev-
atmospheric parameters and atomic line lists to determine eral fully automatic experiments that compare the impact of
chemical abundances. The study showed that homogeniz- using distinct radiative transfer codes, and different set-ups
ing the atomic data and atmospheric parameters led to an and spectroscopic techniques.
improvement in the agreement between abundances derived
by each method, although the dispersion remained high for
several elements. The authors concluded that it is necessary
2 DATA
to investigate further the inherent different results between
spectroscopic techniques. The Gaia FGKM Benchmark Stars (Jofré et al. 2014, 2015;
The same problem was tackled again even more thor- Heiter et al. 2015b; Hawkins et al. 2016) constitute a set of
oughly by Jofré et al. (2017), who determined four abun- very well-known stars covering a wide range in effective tem-
dances for four Gaia Benchmark Stars (Jofré et al. 2014, perature (3 500 to 6 600 K), surface gravity (0.50 to 4.60 dex)
2015; Heiter et al. 2015b; Hawkins et al. 2016) using six dif- and metallicity (-2.70 to 0.30 dex). They are especially con-
ferent methods and one representative line for each of the venient for spectral analysis assessments because they are
four elements with fixed atmospheric parameters (i.e. effec- accompanied by reference atmospheric parameters obtained
tive temperature, surface gravity and metallicity). The study from methods independent of spectroscopy.
showed that equivalent-width methods are less affected by For this work, I used the high-resolution spectra pro-
shifted absorption lines than synthesis methods, which need vided by the public library2 (Blanco-Cuaresma et al. 2014a).
to implement mechanisms to detect the shift and perform a The original non-normalized spectra came from different
correction. The agreement between methods improved when instruments with different resolutions, spectral ranges and
a common normalization was applied (see also the brief dis- signal-to-noise ratios (S/Ns). Some pre-processing was exe-
cussion about normalization effects in section 4.1 of Blanco- cuted to homogenize the data set, including merging sep-
Cuaresma et al. (2015)) and when the same microturbulence arate wavelength regions from the same observation, co-
was used. This parameter is very sensitive to the method adding spectra to increase the S/N, cleaning areas affected
used, however, which makes it impossible of have a good cri- by telluric lines (i.e. setting fluxes to zero), correcting radial
terion to define a good common value for all the techniques. velocities, estimating the S/N and error fluxes (a procedure
The neglect of hyperfine structure (i.e. shifts and splittings that strongly influences the derived parameter errors), se-
in the energy levels of atoms, molecules and ions owing to the lecting the spectral range of 480 to 680 nm (optical range)
interaction between the state of the nucleus and the state of and degrading the resolution to 47000, which matches the
the electron clouds), and different assumptions on the abun- resolution and range of the UVES set-up used in the Gaia-
dances of blending elements contribute to increasing the dis- ESO Survey.
crepancies between analysis. Different atmospheric model in-
terpolation methods lead to very small differences, which are
only relevant if very high-precision abundances are required. 1 The iSpec version used in this work was released as v2019.03.02
For equivalent-width methods, differences arise from the use 2 http://www.blancocuaresma.com/s/

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 3
3 PIPELINE overcome local thermalizing ones (i.e. collisions). Neverthe-
less, the approximation is good enough for AFGKM stars,
3.1 Methods
depending on the type of analysis and scientific goals.
iSpec can derive atmospheric parameters using the synthetic Most of the codes use model atmospheres calculated
spectral-fitting technique and the equivalent-width method. assuming plane-parallel geometry, and hence the radiative
The former compares the observed fluxes (weighted by the transfer is solved while neglecting the curvature of the at-
flux errors if present) with synthetic spectra for a selected set mosphere and only considering one depth variable. This is
of spectral features, and then a least-squares algorithm min- a valid approximation for most stars but it breaks down
imizes the differences (i.e., by computing the χ2 ) by varying when the size of the stellar atmosphere starts to be relevant
the atmospheric parameters until convergence is reached. compared with the stellar radius (e.g. cold giants and super-
The spectral features can be absorption lines or any other giants). Only some codes can consider the curvature when
spectral region. For instance, it is common to use the wings providing model atmospheres that were calculated assuming
of the H-α/β and Mg triplet to help break degeneracies, a spherical geometry.
given that these regions are highly sensitive to the effective All the codes were integrated in iSpec in the most ho-
temperature and surface gravity, respectively. The synthetic mogeneous way possible; this implies that certain default
spectra can be computed on demand by interpolating from behaviours were overriden. For instance, codes such as SYN-
a grid of model atmospheres and using a radiative trans- THE extract the abundances from the model atmosphere
fer code, or now also by interpolating from a grid of pre- input file, while others, such as MOOG, have default hard-
computed spectra with iSpec or another tool (this also gives coded values that can be modified if the user issues the right
the possibility of using grids of synthetic spectra for stars commands. In addition, certain functions provided by these
cooler or hotter than AFGKM, or even grids of observed codes were not used, such as resolution degradation, macro-
spectra). In both cases, the input grid is used to construct turbulence and rotational effects. To guarantee compara-
convex hulls, and a linear barycentric interpolation is exe- ble results, these effects are directly implemented in iSpec
cuted at each necessary triangle. No specific code was writ- and they are homogeneously applied to all synthetic spec-
ten for this: I used widely tested methods present in the tra independently of what radiative transfer code is used.
Qhull and SciPy packages (Barber et al. 1996; Jones et al. Researchers that make use of these tools outside the iSpec
2001). framework should expect differences that were minimized for
In the case of the equivalent-width method (which this work.
also requires model atmosphere interpolations), the analy-
sis starts with a selection of absorption lines produced by
neutral and ionized iron, for which their equivalent width is
measured. Usually this is done by fitting Gaussian profiles 3.2.1 SPECTRUM
and determining the area of each absorption line. Then the SPECTRUM version 2.76e3 (Gray & Corbally 1994) is a
equivalent width can be transformed to abundances by using radiative transfer code written in C (compatible with the
a radiative transfer code, and the atmospheric parameters gcc compiler) that can synthesize spectra and derive abun-
are varied until there is no correlation between abundances dances from equivalent widths. However, the latter function-
and equivalent widths, and excitation equilibrium plus ion- ality is done by fully synthesizing each absorption line and it
ization balance is reached (i.e. there is no correlation with is computationally expensive (i.e. significantly slower) com-
excitation potential, and the average iron abundances from pared with MOOG or WIDTH9, which use a faster direct
neutral and ionized lines are equal). In this case, no spec- computational analysis (Gray 2008, chapter 16). I did not
tral features other than iron absorption lines are used, and use SPECTRUM for the tests based on the equivalent-width
the analysis is fast because the amount of information to be method, but it is an ideal code for the synthetic spectral-
computed is small compared with that in synthesis methods fitting technique because it is one of the fastest.
(i.e. the full line profiles are not considered - only their area).
In this work I compare: (1) the synthetic spectral-fitting
technique using a grid of atmospheric models; (2) the syn-
thetic spectral-fitting technique using a grid of pre-computed 3.2.2 Turbospectrum
synthetic spectra; (3) the equivalent-width method.
Turbospectrum version 15.14 (written in Fortran and com-
patible with the gfortran compiler Alvarez & Plez 1998; Plez
3.2 Radiative transfer codes 2012) is similar to SPECTRUM in terms of usage, and I
For all the method described in Section 3.1, iSpec offers a again excluded it from tests based on the equivalent-width
broad variety of radiative transfer codes (a summary can be method. In contrast to SPECTRUM, which only works with
found in Table 1). It is worth noting that all of these radia- plane-parallel model atmospheres, Turbospectrum can use
tive transfer codes assume local thermodynamic equilibrium spherical models (which offer a better approximation for gi-
(LTE), which means that the mean free path of photons is ant stars), in which the stellar radius and the depth of each
smaller than the scale over which thermodynamic quanti- layer have to be provided (as shown in Table A1).
ties vary, and thus the atmospheric state (e.g. temperature)
at a given depth is affected by radiation below or above
that point. This approximation is not valid for OB stars or
extremely metal-poor stars, which have optically thin lay- 3 http://www.appstate.edu/ grayro/spectrum/spectrum.html
ers where non-local, non-thermal influences (i.e. radiation) 4 http://www.pages-perso-bertrand-plez.univ-montp2.fr/

MNRAS 000, 1–22 (2018)


4 S. Blanco-Cuaresma
SPECTRUM Turbospectrum SME MOOG WIDTH9/SYNTHE

1D plane-parallel model atmosphere geometry X X X X X

1D spherical model atmosphere geometry X X

Non-LTE

Grids of Non-LTE departure coefficient X

Customizable chemical abundances X X X X X

Customizable isotopes X X X

Customizable molecular dissociation constants X X X

Re-computed model atmosphere electron density X X X X

Continuum scattering X X

Radiative damping parameter due to natural broadening X X X X X

Stark broadening due to collisions with charged particles X X X

Classical van der Waals damping parameter X X X X

Anstee and O’Mara van der Waals broadening theory X X X X X

Hydrogen broadening AG BPO BPO BPO AG

Customizable hydrogen lines parameters X X X X

Base line profiles Voigt Voigt Voigt Voigt Voigt

Average synthesis time in seconds (480 - 680 nm) ∼123 ∼56 ∼222 ∼68 ∼360

Table 1. Summary of radiative transfer code features. ’Customizable’ denotes the possibility of changing values without recompiling the
program; ’BPO’ stands for Barklem-Piskunov-O’Mara (Barklem et al. 2000); and AG stands for Ali-Griem (Ali & Griem 1965, 1966).

3.2.3 SME 3.2.5 WIDTH9/SYNTHE


SME version 4.235 (Valenti & Piskunov 1996) is the only ra- WIDTH9 version 9 March 1993 and SYNTHE version 20
diative code considered that is closed-source, which makes July 20017 (Kurucz 1993; Sbordone et al. 2004) are pro-
its debugging and scientific assessment more difficult. It is grams that share the same radiative transfer code, but the
distributed with IDL scripts that call a pre-compiled binary former is used to transform equivalent widths into abun-
library that performs the spectral synthesis; iSpec only uses dances, while the latter computes synthetic spectra. Both
this library. SME only does synthesis and, equivalent to Tur- are written in Fortran and require the Intel compiler, which
bospectrum, it can work with spherical model atmospheres. is not open-source (in contrast to gfortran), and thus pre-
The code is ready to consider departure coefficients for non- compiled executables are included in iSpec by default. More
LTE effects, although this has not been considered for this details about this code can be found in Cowley & Castelli
work. (2002).

3.2.4 MOOG 3.3 Model atmosphere


MOOG version February 20176 (Sneden et al. 2012) is a For the grid of model atmosphere grid I used MARCS8
radiative transfer code written in Fortran (compatible with (Gustafsson et al. 2008), which was computed with solar
the gfortran compiler) that can synthesize spectra and de- abundances from Grevesse et al. (2007). Note that iSpec also
rive abundances from equivalent widths efficiently. Unfortu- support ATLAS/Kurucz and many other solar abundances,
nately, MOOG depends on the non-free SM package (for- but MARCS includes models computed with plane-parallel
merly SuperMongo) for plotting results. Given that iSpec and spherical geometries. The latter allow for spherical ra-
already has its own free python interface, I developed a SM diative transfer (although only with codes that support it,
package mock with the same functions but empty implemen- as explained in Section 3.2) and ensure a more realistic tem-
tation that allows MOOG to be compiled without the official perature structure of the model atmospheres because the
non-free SM package. This is the only code that does not re- spherically symmetric radiative transfer scheme takes into
compute electron densities, but it keeps them fixed them as account the geometric dilution of flux. The spherical radia-
provided by the input model atmosphere. tive transfer is generally not important for line formation,

5 http://www.stsci.edu/ valenti/sme.html 7 http://atmos.obspm.fr/


6 http://www.as.utexas.edu/ chris/moog.html 8 http://marcs.astro.uu.se/

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 5
but is very important for the model atmosphere structures 3.5 Line selection
(Heiter & Eriksson 2006). iSpec uses the model atmosphere
No matter whether the analysis method is based on the
grid to construct convex hulls, where linear barycentric in-
equivalent-width technique or on the synthetic spectral-
terpolations at each necessary triangle can be executed to
fitting technique, the spectral ranges used in the study are
generate the required model with the necessary atmospheric
going to have an impact on the final derived atmospheric pa-
parameters (always within the grid ranges). Regarding the
rameters. Given the nature of the methods used in this work,
radiative transfer codes, not all of them require the same
most spectral ranges will correspond to absorption lines be-
model atmosphere input values: the differences are shown in
cause they carry key information related to the atmospheric
Table A1.
parameters of the star.
It is common to find studies in the literature in which
authors use a line selection carried out by other authors. But
3.4 Atomic data this approach carries risks. A line might be reliable when us-
ing a specific spectroscopic pipeline with a concrete set-up
The atomic data used in this work corresponds to the line list
(e.g. normalization process, atmospheric models, radiative
version 5 from the Gaia-ESO Survey line list (Heiter et al.
transfer codes, atomic data) and observed spectra with a
2015a). The line list format is automatically transformed
particular resolution, but very bad when any of these com-
by iSpec to fit the requirements from every radiative trans-
ponents change. For instance, a line selection carried out
fer code. Furthermore, some lines are not used for certain
with high-resolution spectra might not be convenient for
codes if they are not compatible or necessary. For instance,
lower resolutions because lines can be blended: equivalent-
SPECTRUM has several hard-coded strong absorption lines
width methods will overestimate the abundance, and syn-
(which should not be included in the atomic line list or
thetic spectral-fitting techniques might produce inaccurate
SPECTRUM would generate strong lines twice): hydrogen-
results if the nearby lines have poor-quality atomic data.
line series (Lyman, Balmer, Paschen, Brackett, Pfund and
A strategy that minimizes some of these difficulties is
Humphreys), 31 helium I lines, 11 iron II lines, 1 magne-
to follow a purely line-by-line differential approach. For in-
sium I line and 2 magnesium I lines, 1 calcium I line and 2
stance, this could be done by calibrating the absorption
calcium II lines, 1 scandium II line, and 1 strontium II line.
lines’ log(g f ) value to better reproduce each line profile
Turbospectrum and MOOG also include their own data for
in a reference star with very well-known atmospheric pa-
hydrogen and helium lines. All these included lines will not
rameters (typically, the Sun) and then using this calibrated
be taken from the GES line list when using these codes.
atomic data to derive atmospheric parameters. In the case
In terms of isotopes, the SPECTRUM documentation of equivalent widths, a different but equivalent approach is
indicates that it supports 311 atomic isotopes plus 40 molec- to measure the abundance of all the lines in a reference star,
ular isotopes, and their relative abundances can be fine- subtract the result for all the lines measured in the target
tuned using an input file. The rest of the codes do not seem star, and use these differential abundances (instead of abso-
to offer this possibility (they have hard-coded values) and lute abundances) to reach ionization balance and excitation
are less well documented, which makes homogenization and equilibrium.
comparison difficult. Apart from filtering out isotopes not Nevertheless, given the goal of this study, I preferred to
supported by SPECTRUM, no other atomic data selection avoid the calibration of log(g f ) values and to use the same
has been performed based on isotopes. atomic data for all the different radiative transfer codes. This
Regarding molecules, not all the codes support the same avoids introducing another degree of freedom that might
molecules, and SPECTRUM is again the best-documented make the comparison more difficult, although it makes the
code, while Turbospectrum seems to be the code that sup- line selection process particularly important.
ports the most molecules. Furthermore, every code includes I used the NARVAL solar spectrum with the highest
its own dissociation energies for molecules, and only SPEC- signal-to-noise ratio in the Gaia Benchmark Stars library for
TRUM (via the input solar abundance), MOOG (via the the line selection process described in the following subsec-
input line list) and Turbospectrum (via specific molecule in- tions. The spectrum was convolved to a resolution of 47 000,
put files) allow the user to override them without modifying corrected from its radial velocity and normalized following
the source code. In any case, the public version of the GES the same procedure as any other spectra in this study.
line list does not include molecules (which are relevant only
for the coolest Benchmark Stars).
In all the cases, I discarded atomic lines for second or 3.5.1 Matching absorption lines to atomic data
higher ionized atoms (e.g. Fe III), lines with a lower state
excitation potential higher than 15 eV (corresponding to The first required step in the line selection process is to
only 20 lines) and auto-ionizing transitions for metals (cor- identify which lines from the GES atomic line list are the
responding to only 17 lines) to reduce computation time main contributors to the observed absorption lines in the
because their contribution is small for FGKM stars and fur- solar spectrum. For this, first I only considered lines from the
thermore they are not supported by all the codes. GES line list that have a theoretical depth greater than 0.01
Regardless of starting with a common atomic line list, and a reduced equivalent width9 greater than -7 for solar
there are differences that arise owing to the intrinsic func- atmospheric parameters. Then, I used iSpec to fit Gaussian
tioning of each code. Moreover, not all codes use the same
input values, as shown in Table A2. Sometimes the differ-  
ences are just a matter of units or format, but in other cases 9log10 EλW where EW is the equivalent width and λ is the
there are values that are not required at all by some codes. wavelength position

MNRAS 000, 1–22 (2018)


6 S. Blanco-Cuaresma
profiles for all the lines in this subset and discarded those line shifts. I then determined the abundances by letting only
whose fit failed or that had a depth greater than 1 or lower the corresponding element be a free parameter. In order to
than 0.01. There may be more than one close atomic line be able to later assess the quality of each line (i.e. filter-
(closer than 0.001 nm) blended in a single unique observed ing lines with differences larger than certain limits), I also
absorption line. In these cases I discarded all the atomic lines derived abundances for that line when the metallicity is ar-
except the one with the greatest theoretical equivalent width tificially increased by 0.10 dex, when a new realization of
(i.e. the one that has the highest probability of being the the spectrum is created (fluxes are drawn from a Poisson
main contributor to the observed line). From this process, a distribution using the fluxes as mean values and errors as
total of 2 496 atomic lines were selected. the sigma) and when the atomic line list contains only the
target atomic data and no other blended lines.

3.5.2 Deriving solar abundances


The second step is to determine abundances with all the 3.5.5 Synthetic spectral-fitting technique when
codes and methods for each of the selected lines in the solar interpolating pre-computed spectra
spectrum by fixing the following reference solar parameters In the previous two methods, iSpec interpolated model at-
(based on the Gaia Benchmark Stars recommended values): mospheres using the MARCS grid and provided the model
• effective temperature (Teff ): 5771 K, to the corresponding radiative transfer code. In this method,
• surface gravity (log(g)): 4.44 dex, the atmosphere is not interpolated, but a grid of synthetic
• metallicity ([M/H]): 0.00 dex, spectra was pre-computed using SPECTRUM and match-
• microturbulence velocity (Vmic ): 1.07 km/s, ing the exact atmospheric parameters that the MARCS grid
• macroturbulence velocity (Vmac ): 4.21 km/s, provides but with two different alpha abundance variations
• projected rotational velocity (v sin(i)): 1.60 km/s, ([α/Fe] = ±0.40 dex, where the alpha elements correspond
• limb-darkening coefficient: 0.6. to neon, magnesium, silicon, sulphur, argon, calcium and
titanium) and four microturbulences (0.00, 1.00, 2.00 and
I used solar abundances from Grevesse et al. (2007) to 4.00 km s-1 ). Thus, the dimensions of the grid are effective
be consistent with the MARCS model atmosphere. I imple- temperature, surface gravity, metallicity, alpha enhancement
mented the macroturbulence broadening using the radial- and microturbulence. The grid is computed with a very high
tangential formalism as described in Niemczura et al. (2014) resolution (R > 300 000). This allows iSpec to interpolate a
(adapted from SME), and applied the projected rotational very high-resolution spectrum that can then be degraded to
velocity plus the limb-darkening coefficient following equa- the target resolution, and effects such as macroturbulence,
tion (17.12) from Gray (2008) (adapted from SYNSPEC, rotation and limb-darkening can be applied.
Hubeny & Lanz (2011)). All these effects are directly im- For each line, I performed the same radial velocity cor-
plemented into iSpec and are applied independently of the rection as detailed in the previous section and I derived the
selected radiative transfer code. abundance for each line. In this case, I let the metallicity pa-
rameter be free because the grid does not have a dimension
for every possible chemical element. This strategy is also fol-
3.5.3 Equivalent width lowed by other authors and surveys (e.g. APOGEE) when
Using MOOG and WIDTH9 radiative transfer codes, I de- using pre-computed grids. To be able to assess the quality of
termined the abundances for all the selected lines using the the line, an abundance is also derived for a new realization
equivalent width (EW) derived from the previously fitted of the spectrum (as explained in the previous section).
Gaussian profiles. In addition, in order to assess the quality
of each line I also derived abundances when the metallicity
is artificially increased by 0.10 dex (in order to assess the im- 3.5.6 Selecting lines
pact of errors in the metallicity) and abundances when the I created two line selections for each code: one optimized
EW is drawn from a random distribution using the fitted to be used for determining atmospheric parameters, and a
EW as the mean and its error (computed following Voll- second less strict one that can be used for a line-by-line
mann & Eversberg (2006)) as the sigma (in order to test the determination of individual chemical abundances. To make
signal-to-noise ratio influence). these line selections, I defined the following simple criterion
to evaluate if an absorption line should be selected: a line
can be considered good when I am able to derive an accurate
3.5.4 Synthetic spectral-fitting technique when
solar abundance. In practice, this means that the derived
interpolating model atmospheres
abundance with respect to the solar abundance of reference
In this method, the codes SPECTRUM, Turbospectrum, (understood as [X/H]) should be close to zero within a cer-
SME, MOOG and SYNTHE are used. For each line, I first tain margin.
computed a small synthetic spectrum that includes the tar- After several tests, I found that the optimal mar-
get line and used it to adjust the line mask (the spectral gin for the determination of atmospheric parameters with
region used by the minimization algorithm). This way, if equivalent-width methods is ±0.10 dex (a sufficient number
nearby lines are present in the synthetic spectra but not in of neutral and ionized iron lines need to pass this filter),
the observed one, their impact can be reduced by excluding while for the synthetic spectral-fitting technique the margin
them from the mask. I cross-correlated the same synthetic can be ±0.05 dex. For the former, only iron lines are consid-
spectrum with the observed one to detect and correct small ered, while for the latter, I considered only lines not affected

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 7
by hyperfine structure splitting and that belong to iron peak ize the spectra (each full iteration includes a normalization
elements (iron, chromium, nickel) and alpha elements (sili- and the determination of parameters executed by the mini-
con, calcium, titanium). mization algorithm, which goes through multiple iterations
It is worth noting that these selected lines will not be exploring the parameter space until convergence). The first
blindly used with all the target spectra. Before executing the full iteration will normalize the spectra (before deriving any
determination of atmospheric parameters, I fitted Gaussian parameter) by applying a median and a maximum filter with
profiles to all the selected lines using the target spectrum different window steps (0.05 and 1.0 nm, respectively) and
and I discarded those that do not contain valid fluxes, are fitting the continuum with a B-spline of 2 degrees every 5
affected by telluric lines, have a bad line mask, have a Gaus- nm (ignoring strong lines automatically detected by iSpec).
sian profile fit that failed or too large a root mean square It is worth noting that the same normalization process was
(i.e. rms error >= 1.00), have a reduced equivalent width used in the line selection described in Section 3.5. The sec-
greater than -4.2 or lower than -6 (to avoid saturated or ond full iteration, if enabled, will synthesize a spectrum (i.e.
too weak lines), or have an excitation potential too extreme template) with the atmospheric parameters found in the first
(e.g. greater than 6 eV, where there are almost no lines and full iteration. The observed spectrum is divided by the tem-
an outlier can deeply affect trend computations). Hence, the plate, and I apply median and Gaussian filters with different
line selection will be further fine-tuned and adapted to each window steps (0.05 and 10 nm, respectively) to find the con-
target spectrum to be analysed. tinuum. The advantage of using a synthetic spectrum as a
Once the atmospheric parameters of a star are found, template is that areas with strong lines and many blended
the determination of individual chemical abundances can be lines (e.g. the blue part of the visual range for cooler stars
calculated differentially line-by-line, and it is not necessary tends to be very crowded and blended) will be better nor-
to be so strict with the selection criteria, especially if we malized, and differences between spectra from the same star
want to include elements that have only a few difficult lines. but with different noise levels will be normalized more simi-
Thus, if an element has more than 10 lines, the strict mar- larly; hence the determination of parameters in that second
gin is applied (±0.10 and ±0.05 dex for equivalent-width and full iteration may improve with respect to the first full iter-
synthesis, respectively) but if not, a more generous margin of ation. The risks are that the first full iteration may have led
±0.50 dex is enforced (a value also determined from experi- to inaccurate parameters (i.e. the template will be synthe-
mental tests). With this strategy, we can maximize the num- sized with bad parameters), and that lines existing in the
ber of elements for which we can derive abundances without synthetic spectrum but not in the observed one can create
affecting the quality of the elements that already have many normalization artefacts.
lines. To accelerate the convergence process during the deter-
Some more quality controls are applied to the line se- mination of atmospheric parameters, it is optimal to start
lection for atmospheric parameters and the line selection for with initial parameters as close as possible to the type of
individual chemical abundances. Absorption lines close to star we are analysing. I pre-computed with each code a very
the known strong lines H-α (652-660 nm), H-β (483.5-489.5 limited grid of synthetic spectra that covers only four tem-
nm) and the Mg triplet (514-521 nm) are discarded. Weak peratures (3500, 4500, 5500 and 6500 K), two surface grav-
lines are impacted to a greater extent by continuum place- ities (1.5 and 4.5 dex) and three metallicities (-2.0, -1.0 and
ment: to reduce this effect I discarded absorption lines with 0.0 dex). The normalized observed spectrum is compared
depths lower than 0.05. The determination of line shifts can with all the spectra in the grid, and the parameters of the one
be inaccurate for very weak and noisy lines or for extremely with the lowest χ2 are selected as initial values. This process
blended lines. Based on some manual tests, the best indica- allows me to quickly distinguish between metal-poor/rich
tor to identify these cases is the error on the radial velocity dwarfs and giants and start the minimization algorithm with
that comes from the cross-correlation process, which should values closer to the final solution, thereby speeding up the
be lower than 100 km s-1 (from visual inspection, errors convergence.
larger than this are correlated with problematic lines, while For the equivalent-width method, iSpec lets the effective
slightly lower errors can be caused by problematic lines or temperature, surface gravity and/or microturbulence veloc-
by overestimated errors). Abundance errors are also good ity be set as free parameters, and a maximum of 20 iter-
indicators for identifying good fits. I discarded any line with ations are allowed. The synthetic spectral-fitting technique
an error greater than 0.25 dex. Abundances derived from uses the effective temperature, surface gravity, metallicity,
new realizations of the equivalent-width or the spectrum alpha enhancement, microturbulence velocity and resolution
fluxes should not be more than 0.10 dex away from the as free parameters, with a maximum of six iterations (sev-
main derived abundance; otherwise, the line is too sensitive eral tests showed that these are reasonable maximums to
to the noise. Finally, in the case of the synthetic spectral- obtain accurate results in an optimal computation time; see
fitting technique using atmospheric model interpolation, I Blanco-Cuaresma et al. 2014b). The resolution, macrotur-
discarded lines with abundances more different than 0.10 dex bulence and projected rotational velocity are degenerate pa-
when the metallicity was artificially set 0.10 dex higher, and rameters, which are very difficult to disentangle by relying
lines for which the abundance was 0.50 dex different when only on spectroscopy. After several tests, the most accurate
using atomic line lists without blended lines. results were obtained by fixing the rotation to 1.6 kms-1 and
letting the macroturbulence follow an empirical relation es-
tablished by GES (although setting this parameter to zero
3.6 Atmospheric parameters and abundances
leads to similar results). This empirical relation was built by
The determination of parameters can take place in one or GES considering the effective temperatures, surface gravities
two full iterations, depending on how we want to normal- and metallicities from their data set.

MNRAS 000, 1–22 (2018)


8 S. Blanco-Cuaresma
The determination of individual chemical abundances 3.8 The one variable at a time experiment
follows the same structure as described in Section 3.5.2,
To understand what each code does and how they differ in
where individual line shifts are detected by cross-correlating
detail, there are mainly two broad strategies: (1) read the
the spectrum region of the target line with a synthetic tem-
documentation, ask the author(s) and invest a large amount
plate, and then abundances are derived using the corre-
of time interpreting the thousands of lines present in each
sponding method and some additional controls are executed
source code (if available); (2) design an experiment in which
as explained in Sections 3.5.3, 3.5.4 and 3.5.5. I discarded
one output variable is measured while all the input variables
lines for which it was not possible to derive the main abun-
remain constant except one. These are not exclusive strate-
dance or any of the quality-control abundances, and lines
gies, and following both of them would be instructive, but
with extreme abundances that fall outside the metallicity
given the complexity of the codes (and that fact that they
range considered in the atmospheric models grid (i.e. abun-
are written in different programming languages, and one of
dances with respect to the Sun greater/lower than 1.0/-
them is not public) and the limited amount of resources, I
5.0 dex). From the remaining data set, I discarded lines that
mainly followed the second one.
do not pass the quality controls following the same criteria
In the case of the equivalent-width method, I consid-
as explained in Section 3.5.6, except for elements that have
ered the lines in common in the two codes (MOOG EW and
only one line (I relaxed the criteria to maximize the number
WIDTH) plus the equivalent widths measured in the solar
of measured elements).
spectrum used in Section 3.5, and I measured the median
To partially compensate for modelling errors, it is use-
abundance. In the case of the spectral-fitting technique, I did
ful to perform a differential abundance analysis. For certain
not use any observed data and I measured the synthetic flux
studies it could be convenient to use more than one refer-
depth around 556.45 nm (a region with practically no blend,
ence star, which are at different evolutionary stages (Blanco-
and thus close to the continuum) and the flux depth at the
Cuaresma & Fraix-Burnet 2018; Blanco-Cuaresma & Soubi-
line peaks for the common selection. In both cases, I used the
ran 2016), but for this work I used the Sun as the only refer-
solar parameters detailed in Section 3.5.2, all the default val-
ence and included seven solar spectra from the Gaia Bench-
ues for the atomic line list and atmospheric model, and the
mark Stars library, which were analysed using the same pro-
results from MOOG (EW) and SPECTRUM, respectively,
cess as for the rest of the spectra (i.e. pre-processing, nor-
as reference points.
malization, determination of atmospheric parameters and
In the experiment, I tracked the changes to the reference
abundances). For most of the selected lines, I obtained seven
equivalent-width abundance, continuum and absorption line
different measurements (one per spectrum) and I used them
peak depths while (1) changing (one at a time) the effec-
to compute an averaged abundance and a dispersion (to be
tive temperature, surface gravity, metallicity, alpha enhance-
used as the error). To ensure good reference values, I fil-
ment, microturbulence, number of layers in the atmospheric
tered lines that were not measured in more than three solar
model; (2) multiplying by a factor (between 0.25 and 1.75)
spectra (some lines can fail because of quality issues in the
the values of column mass (rhox), temperature, gas pressure
observed spectrum or missing fluxes). For the rest of stars,
(pgass), electron density (xne), Rosseland mean absorption
the final differential abundances were derived by subtract-
coefficient (abross), radiation pressure (accrad), microturbu-
ing the reference abundance line by line, while errors were
lence velocity (vturb), optical depth (logtau5) and electro
added quadratically.
pressure (pelectron) on each atmospheric model layer; (3)
multiplying by a factor (between 0.25 and 1.75) the values
of the oscillator strength (loggf), radiative damping param-
eter (rad), Stark damping parameter (stark), van der Waals
damping parameter (waals). The factor multiplication was
applied as a logarithmic addition for parameters expressed
3.7 The non-observed data set experiment in logarithmic terms such as logtau5, loggf, rad, stark and
waals.
As described in Section 2, the main analysis in this work uses
the high-resolution spectra from the Gaia FGKM Bench-
mark Star public library. Observed data may be affected
by many different variables that depend on the instrument 4 RESULTS
used, the night conditions, the treatment of the raw data,
4.1 Line selection
etc. In addition, there is no model that can perfectly re-
produce all the physical processes that take place in a star The total numbers of selected lines for each code (as de-
(e.g. theoretical assumptions are made to make the prob- scribed in Section 3.5) are shown in the diagonal of the table
lem tractable with our current computer resources and time drawn in Fig. 1. The rest of the table shows the number of
constraints). To remove the possibility that any of these vari- lines in common between each pair of codes. MOOG EW
ables is playing a role in the main analysis done in this work, (i.e. MOOG using equivalent width) is the code with which
I created a purely theoretical data set by synthesizing spec- the greatest number of accurate solar abundances were ob-
tra using the Gaia FGKM Benchmark Star reference values, tained, followed by Grid (synthetic spectral-fitting using a
the signal-to-noise ratios from the public library, and all the grid of pre-computed spectra with SPECTRUM) and SYN-
synthesis codes used in this study plus interpolation from a THE. Recall that lines from the equivalent-width codes were
pre-computed grid method. This produced a data set of 672 selected using a looser criterion than the rest (as described
normalized synthetic spectra, which was analysed following in Section 3.5.6) and these numbers are not directly compa-
the same procedure as for the observed data set. rable.

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 9
:: Selected solar lines in common systematics for any particular element (see Fig. 3). When

Turbospectrum

MOOG Synth
comparing line-by-line equivalent-width code results, some

SPECTRUM
MOOG EW
WIDTH9

SYNTHE
systematics are observed for certain elements, and major

SME
Grid
disagreements appear with larger reduced equivalent widths
MOOG EW 381 258 165 168 166 159 165 175 (see Fig. 4). Similar systematics are also observed for syn-
1256 350 thesis codes, but the size of the reduced equivalent width
WIDTH9 258 269 107 110 111 102 109 111 does not seem to have a major impact on the results be-
908
Grid 165 107 321 280 254 264 251 261 300 tween different codes (see Figs 5 and 6). This shows the
1582
importance of executing line-by-line differential analysis to
SPECTRUM 168 110 280 305 257 273 256 275
1619 250 minimize different systematics between codes.
Turbospectrum 166 111 254 257 296 244 257 257 In total, considering the abundances within ±0.05 dex
1546
SME 159 102 264 273 244 302 242 265 200 with respect to the solar abundance, there are only 45 ab-
1549 sorption lines in common between all the codes (of which
MOOG Synth 165 109 251 256 257 242 288 242 26 correspond to neutral iron and one to ionized iron). It
1522 150
SYNTHE 175 111 261 275 257 265 242 311 would not be possible to determine atmospheric parameters
1586
with this limited number of lines. Instead, given the differ-
ent nature of the equivalent-width method and the synthetic
spectral technique, I created one line selection for each ap-
Figure 1. Absorption lines in common for which a ±0.05 dex proach (hereafter, the common line selection). The common
abundance (±0.10 for MOOG EW and WIDTH9) was derived line selection is composed of 258 lines (where 146 correspond
with a particular code when analysing the NARVAL solar spec- to neutral iron and 11 to ionized iron) for equivalent-width
trum with the highest signal-to-noise ratio. The number below
methods (i.e. MOOG EW and WIDTH9), and of 205 lines
each code name correspond to the sum of all the values in the
for Grid plus the rest of the synthesis codes. The numbers
row minus the lines that correspond to the same code. The label
Grid corresponds to the results obtained when interpolating from are higher for the former because a less strict limit was re-
a pre-computed grid of synthetic spectra. quired for this method (i.e. the limit was set to ±0.10 instead
of 0.05, as explained in Section 3.5.6; otherwise, not enough
ionized iron lines would be left).
Merely by looking at the colour-coding of the results Regarding lines selected for chemical abundance deter-
we can see two separate islands: equivalent-width codes mination where the constraints were more relaxed, as de-
(MOOG EW and WIDTH9) and synthesis codes (Grid, scribed in Section 3.5.6, an average of ∼1 200 lines were
SPECTRUM, Turbospectrum, SME, MOOG Synth AND selected for all the codes, with the exception of WIDTH9
SYNTHE). These two groups have a greater number of lines and Grid, for which ∼900 and ∼1 400 lines were selected.
in common within themselves but not so many across each This is coherent with the original distribution of derived
other. These differences can be intrinsic to how these meth- abundances shown in Fig. 7, where WIDTH9 underperforms
ods work: one only considers the area of an absorption line, compared with MOOG EW. Equivalent-width codes show a
while the other takes into account the full shape of the line larger variance with a skewed distribution favouring larger
profile including blends (see also section 5.1 in Casamiquela abundances, and Grid has the largest number of lines around
et al. 2017). When comparing line-by-line abundances for zero.
each element from the two methods, it can be seen that the
equivalent-width method provides larger abundances than
synthesis, because the latter can reproduce and account for 4.2 Impact on atmospheric parameters
blends (if the synthesis is forced to ignore blends, the agree-
4.2.1 Full Gaia Benchmark Stars data set
ment with equivalent-width results increases, as shown in
section 5.1 in Casamiquela et al. 2017). At the same time, Using the common line selection, I compared the derived
saturated lines (those with a greater reduced equivalent atmospheric parameters for the Gaia Benchmark Stars by
width) depart from a Gaussian profile (used to determine the computing the mean difference between each radiative trans-
equivalent width), and the equivalent-width method derives fer code and calculating the robust standard deviation10 (i.e.
smaller abundances. The example shown in Fig. 2 compares dispersion) of these differences (Fig. 8). Ideally, we would like
codes that use the same radiative transfer core code; thus, both quantities to be as close as possible to zero, meaning
these differences do not arise from major differences in their that the precision is high between different pairs of radia-
implementation but from the intrinsic differences between tive transfer codes. The dispersion found in the three atmo-
the two methods. spheric parameters again shows two islands that separate
In a previous discarded analysis (not included in this equivalent-width methods from synthesis methods with pre-
work), when interpolating from a grid of spectra that was cisions higher within each group but not lower between the
computed with only two microturbulences (0.00 and 4.00 km two groups.
s−1 ), a lower number of lines in common with pure synthesis In terms of median differences, a clear bias is observed
was found. Hence, increasing the number of data points to
cover four microturbulences (0.00, 1.00, 2.00, and 4.00 km
s−1 ) led to a higher agreement between Grid and the rest 10Function mad std from the astropy.stats package (The Astropy
of the synthesis codes. Collaboration et al. 2018; Astropy Collaboration et al. 2013):
A line-by-line comparison between Grid and SPEC- σ ≈ MAD
−1 ≈ 1.4826 MAD where Φ−1 (P) is the normal inverse
Φ (3/4)
TRUM (the code used to pre-compute the grid) shows no cumulative distribution function evaluated at probability P = 3/4

MNRAS 000, 1–22 (2018)


10 S. Blanco-Cuaresma
-0.08±0.18 0.03±0.20 -0.16±0.15 -0.14±0.17 -0.04±0.18
0.4 Fe Ca Co Cr Mg
MOOG Synth - MOOG EW

0.2
0.0
0.2
0.4
-0.25±0.21 -0.10±0.16 -0.09±0.17 -0.15±0.17 -0.12±0.14
0.4 Mn Ni Si Ti V
MOOG Synth - MOOG EW

0.2
0.0
0.2
0.4
6 5 6 5 6 5 6 5 6 5
-0.09±0.19 0.02±0.25 -0.16±0.15 -0.17±0.19 No lines in common
0.4 Fe Ca Co Cr Mg
SYNTHE - WIDTH9

0.2
0.0
0.2
0.4
-0.29±0.20 -0.10±0.16 -0.12±0.23 -0.17±0.18 -0.13±0.13
0.4 Mn Ni Si Ti V
SYNTHE - WIDTH9

0.2
0.0
0.2
0.4
6 5 6 5 6 5 6 5 6 5

Figure 2. Solar abundance difference between equivalent-width and synthesis codes as a function of reduced equivalent width for various
elements. The median and absolute median deviation are indicated in the upper left of each subplot.

for the surface gravity and a less significant systematic in periment described in Section 3.8, the results of which are
effective temperature, where the equivalent-width methods presented in Section 4.5.
provide consistently lower and higher values, respectively. To be able to visually compare all the parameters for
This effect may be driven by the differences in the microtur- all radiative transfer codes at the same time, I normalized11
bulence velocity, as shown in Fig. 9. The microturbulence all the values from Fig. 8 and added them together as shown
parameter represents ensemble velocity fields that are not in Fig. 10 (left plot). In addition, I repeated the same op-
available in 1D model atmospheres (in 3D models, the micro- eration for all the results obtained when the best line se-
turbulence parameter is not necessary), and these velocity lection (hereafter the ’own lines’ selection) for each code is
fields have broadening effects (depth-independent) on the used (right plot). Using the best line selection improves the
line opacity (the parameter serves to desaturate the line). statistics by increasing the number of lines, but it introduces
The differences shown for the microtubulence velocity, es- more inhomogeneities into the analysis. The former effect
pecially between the equivalent-width method and the syn- dominated for synthesis codes because the agreement among
thetic spectral-fitting technique, could be caused by a com- them slightly increased (mainly for MOOG SYNTH), while
pensatory effect on differences in the derived effective tem-
peratures and surface gravities or by real differences between
11 All the values were scaled to unit norm (vector length) us-
the methods and codes. The latter is explored with the ex-
ing the sklearn.preprocessing.normalize function (Buitinck et al.
2013; Pedregosa et al. 2011).

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 11
-0.00±0.03 -0.01±0.03 0.00±0.06 -0.00±0.05 0.00±0.01
0.2 Fe Ca Co Cr Mg
Grid - SPECTRUM

0.1
0.0
0.1
0.2
-0.01±0.09 0.00±0.05 -0.00±0.06 -0.00±0.08 -0.01±0.07
0.2 Mn Ni Si Ti V
Grid - SPECTRUM

0.1
0.0
0.1
0.2
6 5 6 5 6 5 6 5 6 5

Figure 3. Solar abundance difference between synthesis and interpolation from pre-computed grid of spectra (i.e., Grid) as a function of
reduced equivalent width for different elements. Median and absolute median deviation are indicated on the upper left of each subplot.

0.01±0.02 -0.00±0.01 0.04±0.01 0.02±0.02 No lines in common


0.2 Fe Ca Co Cr Mg
WIDTH9 - MOOG EW

0.1
0.0
0.1
0.2
0.01±0.01 0.01±0.02 0.02±0.09 0.04±0.01 0.02±0.01
0.2 Mn Ni Si Ti V
WIDTH9 - MOOG EW

0.1
0.0
0.1
0.2
6 5 6 5 6 5 6 5 6 5

Figure 4. Solar abundance difference between equivalent width codes as a function of reduced equivalent width for different elements.
Median and absolute median deviation are indicated on the upper left of each subplot.

the latter was more significant for equivalent-width methods, set-ups using the selection of common lines or the best lines
where WIDTH9 results separated from MOOG and they got for each code plus enabling/disabling the following options.
slightly more similar to the synthesis results.
• In addition to the selected lines, consider the wings of H-
Assessing the precision between pairs of codes allows us α/β and the Mg triplet. Enabling this option is not possible
to verify what codes lead to the most similar results, but for the equivalent-width methods, for which the results in
does not verify which code and/or set-up obtains the results Fig. 11 are just duplicated.
closest to the expected reference parameters (e.g. a pair of • Run a second full iteration as described in Section 3.6,
codes may be very imprecise because only one of them is where the normalization is repeated but using a synthetic
very accurate). The accuracy of the results (the difference spectrum as a template (which matches the atmospheric pa-
with respect to the Gaia Benchmark Stars reference values) rameters found in a first full iteration) and re-determine the
is shown in Fig. 11. I ran the analysis with eight different atmospheric parameters.

MNRAS 000, 1–22 (2018)


12 S. Blanco-Cuaresma

-0.01±0.04 -0.01±0.04 0.01±0.05 -0.03±0.03 -0.01±0.04


Fe Ca Co Cr Mg
Turbospectrum - SPECTRUM

0.2
0.1
0.0
0.1
0.2

-0.02±0.04 0.00±0.03 -0.03±0.05 0.01±0.03 0.01±0.03


Mn Ni Si Ti V
Turbospectrum - SPECTRUM

0.2
0.1
0.0
0.1
0.2
6 5 6 5 6 5 6 5 6 5
0.01±0.01 -0.02±0.05 0.01±0.02 0.00±0.04 -0.02±0.02
0.2 Fe Ca Co Cr Mg
SME - SPECTRUM

0.1
0.0
0.1
0.2
-0.02±0.03 -0.01±0.02 -0.02±0.02 -0.01±0.03 -0.00±0.02
0.2 Mn Ni Si Ti V
SME - SPECTRUM

0.1
0.0
0.1
0.2
6 5 6 5 6 5 6 5 6 5

Figure 5. Solar abundance difference between synthesis codes as a function of reduced equivalent width for different elements. Median
and absolute median deviation are indicated on the upper left of each subplot.

The equivalent-width method presents a higher disper- malized and added all the values from Figs 11 and 12, as
sion for all the atmospheric parameters. It also has the low- shown in Fig. 13. As a general rule, using the best line se-
est level of agreement when analysing several spectra cor- lection instead of the common line selection leads to a bet-
responding to the same star, as shown in Fig. 12. Metal- ter accuracy for all the codes, thanks to the increase in the
licity is not included in that figure because the results are statistics without sacrificing quality.
very similar across codes: the median robust standard devia-
The codes MOOG EW and WIDTH lead to similar re-
tion per star is about 0.03 dex for equivalent-width methods
sults for the equivalent-width method, with MOOG EW be-
and 0.01 dex for synthesis methods. The synthetic spectral-
ing the best of the two when using its own line selection and
fitting technique performs better in this test mainly because
executing a second full iteration normalizing with a synthetic
the Gaia Benchmark Stars include a wide range of FGKM
spectrum matching the atmospheric parameters found in the
stars and the equivalent-width method is not the best option
first full iteration. This second full iteration does not have
for all of them (see Section 4.2.2). For instance, the accuracy
the same positive effect for all the codes. Its major contribu-
of the equivalent-width method degrades more strongly with
tion is improving the dispersion per star for most synthesis
cooler stars owing to blends and with metal-poor stars owing
codes, as shown in Fig. 12, but sometimes it slightly wors-
to the lack of iron lines.
ens the overall results. The effect of this second full iteration
In order to visually compare all the parameters for all could be due to the template-based normalization or to the
radiative transfer codes and set-ups at the same time, I nor- execution of an extra batch of iterations until convergence

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 13
-0.00±0.04 -0.01±0.02 -0.00±0.07 -0.04±0.05 -0.00±0.02
0.2 Fe Ca Co Cr Mg
MOOG Synth - SPECTRUM

0.1
0.0
0.1
0.2
-0.01±0.04 -0.00±0.05 -0.02±0.06 -0.01±0.05 -0.01±0.04
0.2 Mn Ni Si Ti V
MOOG Synth - SPECTRUM

0.1
0.0
0.1
0.2
6 5 6 5 6 5 6 5 6 5
0.01±0.02 -0.00±0.01 0.04±0.04 -0.02±0.03 -0.00±0.01
0.2 Fe Ca Co Cr Mg
SYNTHE - SPECTRUM

0.1
0.0
0.1
0.2
0.00±0.03 0.01±0.01 0.01±0.02 0.03±0.02 0.02±0.02
0.2 Mn Ni Si Ti V
SYNTHE - SPECTRUM

0.1
0.0
0.1
0.2
6 5 6 5 6 5 6 5 6 5

Figure 6. As Fig. 5

for the determination of atmospheric parameters is reached. agreement is outstanding. Synthetic solar spectra for each
I executed a validation test with Grid using only one full code are also shown in the bottom subplots of the same fig-
iteration with the best line selection but allowing the pro- ure. SME and Turbospectrum are close together with lower
cess that determines the atmospheric parameter to run for normalized fluxes, while MOOG is on the other extreme with
a greater maximum number of iterations (12 instead of 6), higher normalized fluxes. Grid, SPECTRUM and SYNTHE
and the results did not change significantly (rms decreased have a large region of overlap with each other and they are
by less than 0.01 for 11 spectra, and the rest remained at generally closer to the observed spectrum. These discrepan-
roughly the same level). This is a strong indication that the cies reflect differences in how the broadening of the hydrogen
effects of adding the second full iteration are caused mainly lines is computed by each code.
by the template-based normalization. Determining the effective temperatures using the wings
For the synthetic spectral-fitting methods, adding the of the hydrogen lines is a recognized strategy (Niemczura
wings of H-α/β and the Mg triplet generally improves the et al. 2014; Cayrel et al. 2011), and is understood to be
results for Grid, SPECTRUM and SYNTHE, while it wors- very difficult (Barklem et al. 2002). In this context, Giribaldi
ens the results for Turbospectrum, SME and MOOG Synth. et al. (2018) showed how normalization plays a major role
To rule out that the normalization process is not favouring and identified a systematic of 28 K for the Sun when using
some of the codes, a comparison of an observed NARVAL H-α. Given the differences found in this work for the consid-
solar spectrum and an observed solar ATLAS (Hinkle et al. ered radiative transfer codes, I executed an extra analysis in
2000) is shown in the top subplots of Fig. 14, where the which I determined the effective temperature for all the Gaia

MNRAS 000, 1–22 (2018)


14 S. Blanco-Cuaresma
higher disagreements for giant stars and the coolest star.
When one code from each method is compared, as shown in
MOOG EW the right subplots in Fig. 18, the highest discrepancies are
300 WIDTH9 EW found for stars with low temperatures, gravities or metallic-
Grid ities.
250 SYNTHE
Number of lines

200 4.2.2 Limited Gaia Benchmark Stars data set

150 The accuracy of the equivalent-width method can be affected


by the presence of strongly blended lines and by the lack
100 of enough observed iron lines, while the synthetic spectral-
fitting technique is more lenient. Hence, the cooler and/or
50 more metal-poor Gaia Benchmark Stars are challenging tar-
gets for the equivalent-width method, and to account for this
0
1.00 0.75 0.50 0.25 0.00 0.25 0.50 0.75 1.00 I repeated the previous assessment but selecting only the de-
[Element/H] rived atmospheric parameters for the Gaia Benchmark Stars
that have a reference effective temperature greater than
4 500 K and a metallicity higher than -1.0 dex. In Fig. 20,
Figure 7. Original distribution of derived abundances for the
I show the added normalized median differences and robust
solar spectrum before any filtering was applied.
standard deviation between codes (i.e. precision) using the
common and their own line selections. Interestingly, the sim-
ilarities between codes are in line with what was observed in
Grid SPECTRUM Turbospec. SME MOOG SYNTHE
Fig. 10, but in this case it is MOOG EW that comes closer
H-α -30±27 -6±30 -105±29 -138±26 239±22 -21±31 to the synthesis codes when using their own line selection
instead of WIDTH9.
H-β -93±45 -79±46 -185±46 -237±49 367±85 -87±46
Regarding the accuracy of the codes depending on their
H-α+H-β -217±69 -185±69 -322±65 -381±59 795±128 -188±71 set-up, Fig. 21 shows a similar pattern to Fig. 13, except
that the equivalent-width results are significantly improved
Table 2. Median and absolute median deviation for differences and become more accurate when this limited subset of the
between effective temperatures derived using the wings of hydro- Gaia Benchmark Stars is considered, with MOOG EW the
gen lines and the solar reference value. best code for equivalent width.

Benchmark Stars using the wings of H-α and H-β separately 4.3 Impact on chemical abundances
and together, while the rest of parameters were fixed to their
4.3.1 Full Gaia Benchmark Stars data set
reference values. The results are shown in Fig. 15. HΚ is the
worse modelled line of the two, as shown in the middle sub- Individual chemical abundances were derived by fixing the
plots; however, when combined with H-α (right subplots) atmospheric parameters to the reference values and using
the results improve or remain similar for all of the codes the best line selection for each method. The iron abundances
except MOOG. Turbospectrum, SME and MOOG show the tend to be used as a proxy for metallicity, and indeed Fig. 22
largest systematics, and this may be the reason why adding shows very similar patterns to Fig. 8 (bottom plot), where
these regions to the different analyses presented in this work equivalent-width codes and synthesis codes form separate
does not improve the overall results. If I limit the validation islands. However, the dispersion is worse in Fig. 22 for the
to the solar spectra, the closest effective temperature to the equivalent-width codes. Imposing the same atmospheric pa-
reference value is obtained when using only H-α with Grid, rameters on all the spectra probably worsens the results be-
SPECTRUM and SYNTHE, as shown in Table 2. cause, as shown in Section 4.2, the microturbulence velocity
The best global results (i.e. considering all the meth- does not have exactly the same effect for all the methods.
ods/codes) are obtained using each code’s own line selec- The rest of the parameters may play a role too, and the
tion. I used these results to assess the accuracy and preci- abundance determination may compensate for the discrep-
sion as a function of effective temperature, surface gravity ancies from other parameters when enforcing a certain tem-
and metallicity. Figs 16 and 17 illustrate accuracies by com- perature or gravity that does not match what our analysis
paring the results to the reference values, and they show would have found with our models, codes and set-up.
that the biggest disagreements tend to happen with cold In order to compare visually the precision for all the
and/or metal-poor stars, which is especially significant for analysed elements simultaneously among all the codes, I
the equivalent-width method, as explained in the next sec- added the normalized differences and dispersion and rep-
tion. In terms of agreement between the two equivalent- resent them in Fig. 23. The results indicate that the pattern
width codes (see the left subplots in Fig. 18), discrepancies observed for iron abundances can be generalized for the rest
seem to be mainly influenced by the stellar metallicity, al- of the elements.
though higher discrepancies also appear for lower and higher Regarding the accuracy of all the derived individual
effective temperatures and surface gravities. Metallicity also chemical abundances with fixed atmospheric parameters
affects the level of agreement between synthetic spectral- (Fig. 24), equivalent-width methods outperform synthesis
fitting codes (see Fig. 19), and certain code pairs also show methods for calcium abundances but not for the rest of

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 15
:: teff (median) :: teff (dispersion)

Turbospectrum

Turbospectrum
Common lines Common lines

MOOG Synth

MOOG Synth
SPECTRUM

SPECTRUM
MOOG EW

MOOG EW
WIDTH9

WIDTH9
SYNTHE

SYNTHE
SME

SME
Grid

Grid
250
MOOG EW 0 -23 -22 -46 -50 -54 -10 -71 MOOG EW 0 30 191 171 184 157 172 199
34.50 100 171.50
WIDTH9 23 0 -18 -34 -41 -52 1 -46 WIDTH9 30 0 203 201 224 197 201 203
28.50 201.00 200
Grid 22 18 0 15 -12 -9 39 8 50 Grid 191 203 0 24 40 33 37 24
4.50 35.00

teff (dispersion)
150

teff (median)
SPECTRUM 46 34 -15 0 -11 -23 25 -3 SPECTRUM 171 201 24 0 30 14 40 9
7.00 27.00
0 Turbospectrum 184 224 40
Turbospectrum 50 41 12 11 0 -6 37 11 30 0 31 27 32
8.50 31.50 100
SME 54 52 9 23 6 0 46 18 SME 157 197 33 14 31 0 46 16
7.50 50 32.00
MOOG Synth 10 -1 -39 -25 -37 -46 0 -29 MOOG Synth 172 201 37 40 27 46 0 50 50
13.00 43.00
SYNTHE 71 46 -8 3 -11 -18 29 0 100 SYNTHE 199 203 24 9 32 16 50 0
28.00
1.50 0
:: logg (median) :: logg (dispersion)
Turbospectrum

Turbospectrum
Common lines Common lines
MOOG Synth

MOOG Synth
SPECTRUM

SPECTRUM
MOOG EW

MOOG EW
WIDTH9

WIDTH9
SYNTHE

SYNTHE
SME

SME
Grid

Grid
0.20 0.40
MOOG EW 0.00 -0.03 -0.04 0.03 -0.02 0.00 0.09 -0.01 MOOG EW 0.00 0.04 0.32 0.33 0.32 0.33 0.32 0.31
0.01 0.32
0.15 0.35
WIDTH9 0.03 0.00 -0.01 0.04 0.03 0.02 0.09 0.03 WIDTH9 0.04 0.00 0.34 0.36 0.33 0.34 0.29 0.35
0.03 0.34
0.10 0.30
Grid 0.04 0.01 0.00 0.02 -0.03 0.01 0.08 0.02 Grid 0.32 0.34 0.00 0.09 0.09 0.09 0.21 0.07

logg (dispersion)
0.01 0.09
0.05 0.25
logg (median)

SPECTRUM -0.03 -0.04 -0.02 0.00 -0.04 0.00 0.07 0.00 SPECTRUM 0.33 0.36 0.09 0.00 0.10 0.04 0.10 0.01
0.01 0.10
0.00 0.20
Turbospectrum 0.02 -0.03 0.03 0.04 0.00 0.02 0.11 0.03 Turbospectrum 0.32 0.33 0.09 0.10 0.00 0.08 0.13 0.07
0.01 0.10
0.05 0.15
SME 0.00 -0.02 -0.01 0.00 -0.02 0.00 0.08 0.01 SME 0.33 0.34 0.09 0.04 0.08 0.00 0.11 0.03
0.01 0.08
0.10 0.10
MOOG Synth -0.09 -0.09 -0.08 -0.07 -0.11 -0.08 0.00 -0.08 MOOG Synth 0.32 0.29 0.21 0.10 0.13 0.11 0.00 0.09
0.08 0.12
0.15 0.05
SYNTHE 0.01 -0.03 -0.02 0.00 -0.03 -0.01 0.08 0.00 SYNTHE 0.31 0.35 0.07 0.01 0.07 0.03 0.09 0.00
0.01 0.07
0.20 0.00
:: MH (median) :: MH (dispersion)
Turbospectrum

Turbospectrum

Common lines Common lines


MOOG Synth

MOOG Synth
SPECTRUM

SPECTRUM
MOOG EW

MOOG EW
WIDTH9

WIDTH9
SYNTHE

SYNTHE
SME

SME
Grid

Grid

0.10
MOOG EW 0.00 0.01 0.04 0.04 0.03 0.05 0.05 0.03 0.06 MOOG EW 0.00 0.01 0.07 0.09 0.09 0.10 0.09 0.10
0.04 0.09
WIDTH9 -0.01 0.00 0.03 0.04 0.04 0.03 0.06 0.02 WIDTH9 0.01 0.00 0.08 0.08 0.09 0.09 0.07 0.09
0.03 0.04 0.08 0.08
Grid -0.04 -0.03 0.00 0.00 -0.01 0.01 0.02 0.00 Grid 0.07 0.08 0.00 0.01 0.03 0.03 0.04 0.03
0.01 0.02 0.03
MH (dispersion)

0.06
MH (median)

SPECTRUM -0.04 -0.04 0.00 0.00 -0.01 0.00 0.01 0.00 SPECTRUM 0.09 0.08 0.01 0.00 0.03 0.01 0.03 0.01
0.00 0.02
0.00 Turbospectrum 0.09 0.09 0.03 0.03 0.00 0.04 0.03 0.03
Turbospectrum -0.03 -0.04 0.01 0.01 0.00 0.01 0.02 0.01
0.01 0.03 0.04
SME -0.05 -0.03 -0.01 0.00 -0.01 0.00 0.00 0.00 0.02 SME 0.10 0.09 0.03 0.01 0.04 0.00 0.03 0.01
0.01 0.03
MOOG Synth -0.05 -0.06 -0.02 -0.01 -0.02 0.00 0.00 -0.01 0.04 MOOG Synth 0.09 0.07 0.04 0.03 0.03 0.03 0.00 0.03 0.02
0.01 0.03
SYNTHE -0.03 -0.02 0.00 0.00 -0.01 0.00 0.01 0.00 0.06 SYNTHE 0.10 0.09 0.03 0.01 0.03 0.01 0.03 0.00
0.03
0.00 0.00

Figure 8. Median and robust standard deviation of the difference in effective temperature, surface gravity or metallicity between different
radiative transfer codes when analysing the Gaia Benchmark Stars and using the common line selection (subtraction sense: column minus
row).

the elements. It is worth remembering that the reference group analysed, how results were combined, and how out-
chemical abundances were determined by combining spec- liers were treated or removed (Adibekyan et al. 2015), the
troscopic results obtained by different groups using their own reference values may be biased towards one method or code,
techniques (abundances cannot be obtained independently depending on the element. For instance, when considering
from spectroscopy, in contrast to the effective temperature calcium abundances derived with Turbospectrum but using
and surface gravity). Thus, depending on how many differ- only lines in common with the groups ULB and GAU from
ent methods were used, how many stars and elements each Jofré et al. 2015 (which used synthesis with Turbospectrum)

MNRAS 000, 1–22 (2018)


16 S. Blanco-Cuaresma
:: vmic (median) :: vmic (dispersion)

Turbospectrum

Turbospectrum
Common lines Common lines

MOOG Synth

MOOG Synth
SPECTRUM

SPECTRUM
MOOG EW

MOOG EW
WIDTH9

WIDTH9
SYNTHE

SYNTHE
SME

SME
Grid

Grid
0.6 0.6
MOOG EW 0.00 -0.10 -0.29 -0.34 -0.34 -0.32 -0.26 -0.33 MOOG EW 0.00 0.06 0.33 0.47 0.47 0.40 0.39 0.47
0.30 0.40
WIDTH9 0.10 0.00 -0.22 -0.28 -0.28 -0.26 -0.25 -0.27 0.4 WIDTH9 0.06 0.00 0.36 0.43 0.46 0.42 0.40 0.45 0.5
0.26 0.41
Grid 0.29 0.22 0.00 0.02 0.02 0.02 0.06 0.02 Grid 0.33 0.36 0.00 0.04 0.04 0.06 0.09 0.05
0.2 0.4

vmic (dispersion)
0.02 0.06

vmic (median)
SPECTRUM 0.34 0.28 -0.02 0.00 0.00 0.00 0.03 0.00 SPECTRUM 0.47 0.43 0.04 0.00 0.01 0.02 0.04 0.01
0.00 0.03
0.0 0.3
Turbospectrum 0.34 0.28 -0.02 0.00 0.00 0.00 0.04 0.00 Turbospectrum 0.47 0.46 0.04 0.01 0.00 0.04 0.04 0.02
0.00 0.04
SME 0.32 0.26 -0.02 0.00 0.00 0.00 0.04 0.00 0.2 SME 0.40 0.42 0.06 0.02 0.04 0.00 0.07 0.01 0.2
0.00 0.05
MOOG Synth 0.26 0.25 -0.06 -0.03 -0.04 -0.04 0.00 -0.03 MOOG Synth 0.39 0.40 0.09 0.04 0.04 0.07 0.00 0.04
0.01 0.4 0.06 0.1
SYNTHE 0.33 0.27 -0.02 0.00 0.00 0.00 0.03 0.00 SYNTHE 0.47 0.45 0.05 0.01 0.02 0.01 0.04 0.00
0.00 0.03
0.6 0.0

Figure 9. Median and robust standard deviation of the difference in microturbulence velocity between different radiative transfer codes
when analysing the Gaia Benchmark Stars and using one common line selection for equivalent width methods plus another common one
for synthetic spectral-fitting technique.

:: Common lines :: Own lines


Turbospectrum

Turbospectrum
MOOG Synth

MOOG Synth
SPECTRUM

SPECTRUM
MOOG EW

MOOG EW
WIDTH9

WIDTH9
SYNTHE

SYNTHE
SME

SME
Grid

Grid
(Normalized | Teff| + | logg| + | MH|)

(Normalized | Teff| + | logg| + | MH|)


MOOG EW 0.00 0.43 1.33 1.47 1.39 1.48 1.56 1.52 2.00 MOOG EW 0.00 1.09 1.60 1.81 2.06 1.75 1.41 2.12 2.00
9.18 11.83
WIDTH9 0.43 0.00 1.18 1.48 1.52 1.44 1.53 1.39 1.75 WIDTH9 1.09 0.00 0.97 1.12 1.58 1.17 1.26 1.31 1.75
8.98 8.51
Grid 1.33 1.18 0.00 0.31 0.48 0.38 1.03 0.31 1.50 Grid 1.60 0.97 0.00 0.32 0.46 0.32 0.49 0.36 1.50
5.03 4.53
SPECTRUM 1.47 1.48 0.31 0.00 0.51 0.22 0.72 0.07 1.25 SPECTRUM 1.81 1.12 0.32 0.00 0.38 0.12 0.42 0.23 1.25
4.78 4.40
Turbospectrum 1.39 1.52 0.48 0.51 0.00 0.42 1.03 0.45 1.00 Turbospectrum 2.06 1.58 0.46 0.38 0.00 0.37 0.52 0.41 1.00
5.80 5.77
SME 1.48 1.44 0.38 0.22 0.42 0.00 0.83 0.22 0.75 SME 1.75 1.17 0.32 0.12 0.37 0.00 0.38 0.13 0.75
4.99 4.25
MOOG Synth 1.56 1.53 1.03 0.72 1.03 0.83 0.00 0.79 0.50 MOOG Synth 1.41 1.26 0.49 0.42 0.52 0.38 0.00 0.54 0.50
7.50 5.03
0.25 0.25
dispersions

dispersions
SYNTHE 1.52 1.39 0.31 0.07 0.45 0.22 0.79 0.00 SYNTHE 2.12 1.31 0.36 0.23 0.41 0.13 0.54 0.00
median

median
4.76 5.10
0.00 0.00

Figure 10. Sum of the normalized absolute median differences and normalized robust standard deviation for effective temperature,
surface gravity and metallicity when analysing the Gaia Benchmark Stars. Lower numbers indicate the codes lead to more similar results
(higher precision).

then the median difference and robust standard deviation go the reference macroturbulence velocity and rotation) while
down to −0.03 ± 0.08 and −0.01 ± 0.06, respectively, from the fixing the rest of the parameters to their reference values,
original −0.09±0.11 dex. Similarly, there are several elements and then calculating the iron abundance using the reference
for which equivalent-width codes strongly underperform. For parameters plus the microturbulence and resolution found. I
instance, if I consider cobalt results with MOOG EW and I obtained 0.00±0.03 and −0.02±0.02 dex for Grid and SPEC-
use the same spectra and absorption lines in common with TRUM respectively, which show lower dispersions than the
the groups EPI, POR and UCM from Jofré et al. 2015 (which values −0.05 ± 0.10 and −0.06 ± 0.12 dex obtained with all
used equivalent-width with MOOG) then the median differ- the parameters fixed.
ence and robust standard deviation go to 0.00 ± 0.05 and
0.00 ± 0.03 and 0.00 ± 0.02, respectively, from the original Finally, when considering derived abundances instead
0.22 ± 0.30 dex. of line-by-line differential ones (i.e. the derived abundance
I also showed in Section 4.2 that the microturbulence for a particular absorption line and target star minus the de-
velocity does not have exactly the same effect for all the rived abundance for the same absorption line in the reference
methods, thus imposing the same reference value affects each star, here the Sun), the accuracy of the results and level of
code differently. To assess this effect, I used Grid and SPEC- agreement between codes significantly worsens. Line-by-line
TRUM (i.e. interpolating from a grid of spectra and syn- differential analysis helps to reduce systematics, such as the
thesizing) with their best line selection, and I repeated the ones presented in Section 4.1, although the more different
analysis by first computing the microturbulence velocity and the target star is from the reference star, the less effective
resolution for each star (which will compensate for errors in this strategy is.

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 17

:: teff (median) :: teff (dispersion)

Common+wings [R]
Common+wings [R]

Common+wings
Common+wings

Own+wings [R]
Own+wings [R]

Common [R]
Common [R]

Own+wings
Own+wings

Common
Common

Own [R]
Own [R]
325

182.00

135.00

151.50

129.00

159.00

155.00

145.00

127.50
26.50

Own
Own
6.50

5.00

1.00

1.00
20.50

100
8.00

2.00
MOOG EW 260 327 260 327 236 283 300
MOOG EW 30 120 30 120 69 139 69 139 236 283
94.50 271.50
WIDTH9 18 87 18 87 5 81 WIDTH9 275 322 275 322 243 332 243 332 275
5 81

Absolute median difference


49.50 50 298.50
Grid 16 -43 18 -22 -2 -29 Grid 168 105 147 82 135 123 250

Difference dispersion
-20 -16 129.00
149 98
18.00
SPECTRUM 44 -28 -3 -6 7 -6 -3 0 SPECTRUM 196 128 118 102 156 147 124 115 225
3.00 126.00
Turbospectrum -6 -25 -49 -69 -9 0 Turbospectrum 149 129 156 119 119 119
34.00
-29 -39 -54 130 115 200
124.00
SME 23 -38 -69 -92 0 -4 -44 -36 SME 205 161 213 143 162 163 196 140
37.00 162.50 175
MOOG Synth 69 6 131 97 2 -7 111 90 50 MOOG Synth 142 133 126 139 178 152 141 148
79.50 141.50
SYNTHE 0 -41 -2 -7 2 SYNTHE 162 137 122 80 139 158 150
2.00
-4 1 -2 106 106
129.50
100 125

:: logg (median) :: logg (dispersion)

Common+wings [R]
Common+wings [R]

Common+wings
Common+wings

Own+wings [R]
Own+wings [R]

Common [R]
Common [R]

Own+wings
Own+wings

Common
Common

Own [R]
Own [R]

0.15

Own
Own
0.07

0.10

0.07

0.04

0.05

0.04

0.04

0.03

0.29

0.28

0.21

0.19

0.26

0.26

0.18

0.16
0.40
MOOG EW -0.10-0.05-0.10-0.050.02-0.020.02-0.02 MOOG EW 0.45 0.44 0.45 0.44 0.41 0.33 0.41 0.33
0.04 0.10 0.42
WIDTH9 -0.11-0.09-0.11-0.09-0.08-0.10-0.08-0.10 WIDTH9 0.41 0.42 0.41 0.42 0.36 0.37 0.36 0.37 0.35
Absolute median difference

0.10 0.39
Grid -0.04-0.10-0.03-0.03-0.09-0.12-0.04-0.04 Grid 0.24 0.28 0.16 0.16 0.22 0.24 0.12 0.18

Difference dispersion
0.05
0.04 0.20
SPECTRUM -0.03-0.10-0.04-0.03-0.05-0.03-0.03-0.01 SPECTRUM 0.26 0.25 0.13 0.11 0.25 0.24 0.10 0.09 0.30
0.03 0.18
Turbospectrum -0.09-0.18-0.15-0.16-0.16-0.17-0.16-0.14 0.00 Turbospectrum 0.27 0.25 0.25 0.13 0.27 0.27 0.20 0.13
0.16 0.25 0.25
SME -0.06-0.13-0.12-0.13-0.01-0.03-0.10-0.08 SME 0.28 0.27 0.21 0.22 0.22 0.22 0.17 0.15
0.09 0.05 0.22
MOOG Synth 0.01-0.010.22 0.19 0.00-0.010.14 0.15 MOOG Synth 0.40 0.39 0.21 0.22 0.40 0.40 0.23 0.24 0.20
0.08 0.32
SYNTHE -0.08-0.10-0.03-0.04-0.05-0.04-0.03-0.02 0.10 SYNTHE 0.29 0.21 0.12 0.10 0.24 0.21 0.15 0.10
0.04 0.18 0.15
0.15

:: MH (median) :: MH (dispersion)
Common+wings [R]
Common+wings [R]

Common+wings
Common+wings

Own+wings [R]
Own+wings [R]

Common [R]
Common [R]

Own+wings
Own+wings

Common
Common

Own [R]

0.13
Own [R]

Own
Own
0.04

0.04

0.02

0.04

0.01

0.04

0.10

0.08

0.10

0.06

0.08

0.07

0.07

0.07
0.01

0.01

0.04 0.12
MOOG EW -0.03-0.05-0.03-0.050.04 0.01 0.04 0.01 MOOG EW 0.10 0.13 0.10 0.13 0.12 0.07 0.12 0.07
0.01 0.11
WIDTH9 -0.02-0.03-0.02-0.03-0.01-0.02-0.01-0.02 WIDTH9 0.13 0.12 0.13 0.12 0.13 0.11 0.13 0.11 0.11
Absolute median difference

0.02 0.02 0.12


Grid 0.00-0.030.02-0.03-0.02-0.04-0.01-0.05 Grid 0.08 0.07 0.10 0.06 0.06 0.04 0.07 0.04
Difference dispersion

0.03 0.07 0.10


SPECTRUM 0.01-0.030.02-0.02-0.02-0.03-0.01-0.02 SPECTRUM 0.10 0.09 0.11 0.06 0.06 0.06 0.06 0.07
0.02 0.07
Turbospectrum 0.02-0.040.02-0.04-0.01-0.04-0.02-0.04 0.00 Turbospectrum 0.10 0.07 0.07 0.04 0.07 0.07 0.07 0.05 0.09
0.03 0.07
SME 0.01-0.030.00-0.04-0.02-0.05-0.02-0.05 SME 0.10 0.08 0.10 0.06 0.08 0.09 0.11 0.07 0.08
0.03 0.08
MOOG Synth 0.03-0.040.00-0.05-0.02-0.04-0.01-0.04 0.02 MOOG Synth 0.07 0.07 0.06 0.07 0.08 0.10 0.07 0.07
0.03 0.07 0.07
SYNTHE 0.01-0.040.02-0.03-0.03-0.05-0.02-0.04 SYNTHE 0.10 0.09 0.11 0.06 0.08 0.07 0.07 0.06
0.03 0.08
0.04 0.06

Figure 11. Median and robust standard deviation of the difference in effective temperature with respect to the reference values when
analysing the Gaia Benchmark Stars and using different radiative transfer codes with several set-ups: using lines in common within
equivalent-width and synthesis methods (labelled as ’Common’); using the best lines for each code (i.e. their own lines, which are not
necessarily good for other codes, labelled as ’Own’); using the wings of H-α/β and the Mg triplet (labelled as ’Wings’); and repeating the
normalization (labelled as ”[R]”) but using a synthetic spectrum that matches the atmospheric parameters found in the first iteration.
MNRAS 000, 1–22 (2018)
18 S. Blanco-Cuaresma
4.3.2 Limited Gaia Benchmark Stars data set different pattern shown for metallicities between -2.0 and
0.0 dex is particularly puzzling. The effect of the microturbu-
Similarly to in Section 4.2.2, in order to explore a more lim-
lence velocity shows a small but increasing systematic when
ited region of the parameter space than the Gaia Bench-
going towards higher values (i.e. it is not just a constant
mark Stars cover, I selected individual chemical abundances
offset between codes). Alpha enhancement does not have an
for stars with reference effective temperature higher than
impact on MOOG results, probably because MOOG does
4 500 K and metallicity greater than -1.0 dex. Then I added
not re-compute the electron density but directly uses the
the normalized median difference and robust standard dis-
values from the model atmosphere (i.e. changing the alpha
persion between codes for all the analysed elements, as
parameter does not have an effect unless a full new model
shown in Fig. 25. Filtering out stars that are less convenient
atmosphere is computed with the changed alpha parame-
for the equivalent-width method does not erase the system-
ter). WIDTH9 leads to more stable results, despite signif-
atic differences already observed in Fig. 25 and described in
icantly reducing the number of layers in the model atmo-
Section 4.3.1.
sphere, while MOOG median abundances show an offset of
When comparing the results to the reference values as
0.10 dex when only 12 layers are used instead of the original
shown in Fig. 26, the median differences and robust standard
56
deviations improve for all the codes compared with Fig. 24,
Regarding synthesis (see Figs 29 and 30), lower effective
but more significantly for the equivalent-width methods,
temperatures lead to increasing discrepancies in the contin-
with MOOG EW obtaining better results than WIDTH9.
uum and absorption line core depths. At the coolest end
(2 800 K), Turbospectrum and SME are relatively close, with
4.4 The non-observed data set experiment the deepest continuum among all the codes. MOOG is at the
other end of the range, with a continuum depth similar to
The analysis of the data set of 672 normalized synthetic spec- the Sun. In terms of surface gravity, all the codes compute
tra (112 spectra synthesized with the five codes included in a continuum depth that is generally in better agreement for
this analysis plus 112 interpolated with the pre-computed giant stars, while line core depths show the opposite pat-
grid) described in Section 3.7 represents a major computa- tern, with the exception of MOOG, which is systematically
tional effort, involving hundreds of CPU hours. Apart from deeper than the rest of the codes (the opposite is true for
the determination of atmospheric parameters, it has led line depths). Changes in metallicity show a high level of
to the determination of more than two million abundances agreement for absorption line depths, although small dis-
from more than 60 000 absorption lines analysed with each crepancies in continuum depth are present for solar values.
method and code. Interestingly, variations in alpha abundances lead to very
The outcome of this experiment, summarized in Fig. 27, different continuum depth patterns among the codes. The
reproduces very closely the same main results as obtained highest agreement is found for absorption line depths with
using the observed data set. This ensures that the conclu- solar alpha abundances, while MOOG and Turbospectrum
sions from this work are not affected by observational biases deviate from the rest for negative values. Microturbulence
such as instrument effects, night conditions, raw data pro- effects are only evident on the continuum depths, where all
cessing, or the inability of models to perfectly reproduce the codes keep an extremely small constant difference for
physical processes. all tested values. Finally, when the number of layers in the
model atmosphere is reduced, SME is the code that deviates
4.5 The one variable at a time experiment the most from the reference point.

The zero-point of this experiment (described in Section 3.8)


is set by the atmospheric parameters of the Sun, the MARCS 4.5.2 Model atmosphere
model atmosphere, the GES atomic line list, and the code From the model atmosphere, equivalent-width codes are
MOOG for the equivalent-width approach and SPECTRUM only affected by changes in column mass, temperature, and
for synthesis approach. In all the figures referenced in this gas pressure, plus electron density in the case of MOOG,
section, the zero-point is found at the intersection of the hor- as shown in Fig. 31. The highest disagreements occur with
izontal axis and the thick light grey vertical line. The x-axis variation of temperature values, followed by reductions in
corresponds to the independent variable being changed (e.g. gas pressure.
effective temperature, electron density), and the y-axis is the Regarding the results from synthesis codes, shown in
outcome being evaluated, which is very simply computed as Figs 32 and 33, no effects are observed when changing the
Rosseland mean absorption coefficient, radiation pressure,
f (x) = median (v − vref ) (1) and microturbulence velocities, plus electron density except
in the case of MOOG. Turbospectrum is the only code that
where v is the abundance for equivalent-width codes or uses optical depth instead of column mass, and electron pres-
the depth for synthesis codes, and vref is the abundance or sure instead of gas pressure. In general terms, all the codes
depth of reference. The dispersion is computed using the show differences that remain constant, with the exception of
absolute median deviation the variation of temperature and gas pressure.

4.5.1 Atmospheric parameters 4.5.3 Atomic line list


As shown in Fig. 28, the equivalent-width codes show a All the equivalent-width codes (see Fig. 34) present very
systematic discrepancy for lower surface gravities, and the similar results when varying the oscillator strength, but dis-

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 19
crepancies appear for the atomic broadening parameters. sphere and atomic line list does not imply that the synthetic
MOOG does not implement the Stark damping parameter, spectrum is going to be computed using the same data, be-
and hence no effect is expected when changing this value. cause each code may use different values (also discussed in
However, MOOG also does not show any change when vary- Section 4.5), there are differences in the implementation of
ing the radiative damping parameter. Regarding the van der the line-broadening parameters (see Section 4.5.3), and in
Waals damping parameter, MOOG and WIDTH9 show an the case of synthesis the effect of these differences increases
almost constant systematic. because blends are also taken into account (e.g. nearby lines
With respect to synthesis codes (see Figs 35 and 36), can negatively influence the results if the atomic data are
the effect of varying oscillator strength parameters shows inaccurate). In addition, it is remarkable how much the av-
small changing continuum differences between all the codes. erage percentage of matching lines decreases when compar-
MOOG and Turbospectrum do not implement broadening ing results between an equivalent-width code and a synthesis
using the Stark damping parameter, and thus no effect is ob- one (∼50 per cent). The reason for this can again be related
served when it is modified. MOOG does not show any effect to the intrinsic difference already discussed between each
when varying the radiative damping parameter either. The approach.
rest of the codes show very small variations in line depths, For the different set-ups evaluated in this work, the
and changes are more noticeable in continuum depth, where codes that lead to more similar results are SME, SYNTHE
different patterns emerge. These results reveal clear differ- and SPECTRUM, as shown in Figs 10 and 23, although
ences in the implementation of line-broadening effects for the first one performs worse when using the wings of H-
each code. α/β and/or the Mg triplet, as described in Section 4.2. For
studies in which only the wings of hydrogen lines are used
to determine effective temperature, SPECTRUM and SYN-
THE produce the results with the smallest systematics, and,
5 DISCUSSION
for all the codes, it is preferable to use H-α over H-β because
From the 2496 absorption lines studied in the optical range the latter is not well reproduced. In general terms, the most
(480-680 nm) of the solar spectrum, only an average of ∼300 accurate results are obtained with SPECTRUM closely fol-
abundances are within the range ±0.10 dex when using the lowed by SYNTHE, as seen in Fig. 13, but the former is
equivalent-width method, and within ±0.05 dex when using faster in terms of computation time. Furthermore, it is worth
the synthetic spectral-fitting technique. This represents less remembering that all the radiative transfer codes considered
than 15 per cent of the total number of analysed absorption in this work use 1D models and assume LTE: overviews of
lines, despite the fact that the Sun is the star of reference the present and future prospects of 3D non-LTE models can
that we know best and that the same observed data, nor- be found Barklem (2016); Nissen & Gustafsson (2018); Jofré
malization process, model atmosphere, solar abundances and et al. (2018).
atomic data were used in all the codes. When considering a Interpolating from a grid of pre-computed synthetic
larger margin of ±0.25 dex, the percentage increases to 50 spectra (i.e. Grid) also provides a remarkably good accu-
per cent, 65 per cent and 75 per cent for equivalent-width racy, despite being slightly lower than in SPECTRUM and
methods, synthesis, and interpolation from a pre-computed SYNTHE. It is worth considering, given that for a single
grid of spectra (i.e. Grid), respectively. The equivalent-width spectrum it can take an average of 30 min to derive atmo-
method suffers from blends, skewing the abundance distribu- spheric parameters with synthesis and only 5 min with in-
tion towards greater abundances (i.e. overestimating abun- terpolation. Grid requires having an existing pre-computed
dances from blended lines), and, consequently, presenting grid, which is computational work that only has to be done
a lower number of lines for narrower margins. Grid is the once, and a grid of ∼30 000 spectra covering the region from
strategy with which a higher number of lines are close to 480 to 680 nm can be completed in less than a day on a
the reference abundance. The reasons for this are not obvi- modern machine with 32 CPUs.
ous, and there could be a combination of causes, such as the When comparing all the codes in an even more con-
following. (1) Model atmospheres are computed with a cer- trolled experiment (see Sections 4.4 and 4.5), differences
tain chemical composition (typically directly related to the in abundance determinations (equivalent-width methods) or
solar abundances). Grid scales the metallicity and does not synthetic flux computation emerge when any of the input
alter the individual chemical pattern, and thus its coherence parameters (atmospheric parameters, model atmosphere or
with the model atmosphere is maximal compared with the atomic data) are varied. This is because of implementa-
synthesis codes. (2) For blended absorption lines, synthesis tion differences such as continuum calculations and line-
approaches have a wider range of allowed abundances for a broadening effects, and it shows how varying discrepancies
particular element, while Grid has more constraints, given appear when using the same method (i.e. equivalent-width
that all the abundances are being increased/decreased at the or synthetic spectral-fitting) but different codes.
same time via the metallicity parameter. As shown in Sec- In addition to individual differences between codes, we
tion 4.1, line-by-line comparisons between all the codes show can also note that there are some clear systematics between
systematics for certain elements and disagreements that de- equivalent-width methods and the synthetic spectral-fitting
pend on the reduced equivalent width of each line (except technique. Limiting the analysed data set by discarding the
for synthesis codes). cooler and metal-poorer stars, which are very challenging
The number of good lines in common between types of stars for the equivalent-width method, does not fully
equivalent-width codes is high, while the average percent- erase the systematics, as shown in Figs 10 and 20. MOOG
age of matching for synthesis codes is around 85 per cent. As EW and WIDTH9 lead to extremely similar results when
explained in Section 3.3 and 3.4, using the same model atmo- they use the common line selection, but they deviate when

MNRAS 000, 1–22 (2018)


20 S. Blanco-Cuaresma
using their own line selection. It is particularly interesting sion of 0.1 cm, while a different selection would correspond
that depending on whether the full or the limited data set to a ruler in inches with a precision of 0.1 inches (i.e. 0.254
of Gaia Benchmark Stars is considered, it is WIDTH9 or cm). If we wanted to compare people’s heights, we would ob-
MOOG EW that obtains results more similar to synthe- viously not choose to measure them using rulers in different
sis codes. This shows that the systematics between codes units and with different precisions: homogeneity is neces-
and methods is not independent from the stellar type, and sary unless we know how to transform one unit into another
discourages the practice of blindly combining results from very accurately and the precision of each measurement is
multiple methods and/or codes. similar. We would not also gain much benefit from blindly
In general terms, the equivalent-width method can only combining measurements for each person with several differ-
compete with the synthetic spectral-fitting technique when ent rulers of different units, especially if some of these rulers
the spectral range is limited and the cooler and metal-poorer work worse for a particular type of person (e.g. because of
stars (higher number of blends) are not considered, with their body structure): who we measure, what/how we mea-
MOOG EW leading to better results than WIDTH9. The sure and how we combine the results may introduce biases
analysis of one spectrum using equivalent widths can take that are not easy to account for. It is generally preferable to
less than a minute, and this constitutes a strong argument have a controlled and homogeneous measurement process.
for this approach if the target stars are expected to be in The combination of results from different sources would
the optimal range of parameters. be valuable if it were shown that one method is particu-
It is worth mentioning that the high level of precision larly good for a type of star for which the other methods
between codes shown in this work is higher than the level of are less reliable (e.g. as claimed in Smiljanic et al. 2014).
agreement that one could find between studies from differ- Then the difficult task of combining measurements obtained
ent authors. The tests presented here used a line selection from different rulers may make sense, although the process
that was executed in a very homogeneous way for all the should correctly account for the strengths of each approach,
codes, and it can be expected that heterogeneous line se- each method should be assessed using benchmark objects,
lections lead to higher differences between codes. Moreover, the strategy should be described in detail, and the origi-
the strictly line-by-line differential analysis for the deter- nal non-combined individual values (spectrum per spectrum
mination of individual chemical elements had a significant and line by line if it applies) should also be included in the
effect, increasing the precision among codes and methods. publication. For instance, the current work does not provide
This work also shows how the selection of absorption lines any reason to support mixing results from the equivalent-
and other spectral regions plays a very important role in width method and the synthetic spectral-fitting technique,
terms of accuracy, and this selection should be done using because, apart from the identified systematics, the former is
the same pipeline and criteria that are going to be used less reliable for stars with many blends while the latter leads
for the target spectra. Re-using line selection done by other to more robust results for a wider range of parameters. Thus,
authors using different model atmospheres, atomic data, ra- combining results would not lead to a better overall outcome
diative transfer codes and normalization processes will not but will introduce biases and complicate comparisons.
guarantee the best results. These results also question the
practice of blindly using the full spectrum to derive atmo-
spheric parameters, given that with the existing models it is
6 CONCLUSIONS
not possible to completely reproduce even the Sun, the star
that we know the best. In this work, I expanded the capabilities of iSpec to de-
The dependence of the microturbulence velocity on the rive atmospheric parameters and abundances using several
spectroscopic method seen in this work clearly suggests new radiative transfer codes. The user can choose between
that re-using microturbulence results from different meth- MOOG and WIDTH9 for equivalent-width methods, and
ods should not be a recommended practice. In the case of between SPECTRUM, Turbospectrum, SME, MOOG and
synthesis, the same recommendation could be extended to SYNTHE for the synthetic spectral-fitting technique. In ad-
the macroturbulence velocity, for which different recipes ex- dition, I included the possibility of interpolating from a grid
ist (e.g. Gaussian and radial-tangential broadening), and it of pre-computed/observed spectra (i.e. Grid), which reduces
tends to be degenerated together with the projected rota- the analysis time for the determination of atmospheric pa-
tional velocities and the resolution. Moreover, it is some- rameters of a single spectrum down to 5 min or less with a
times common practice to fix some of the atmospheric pa- modern computer.
rameters to values obtained by other independent (occasion- By designing a completely automatic spectroscopic
ally more accurate) means, but, for instance, forcing the pipeline and analysing the Gaia Benchmark Stars, I exe-
model to use a certain effective temperature can lead to cuted an experiment that explores the key pitfalls of mod-
biases in the rest of the free parameters, as described in ern spectroscopy by comparing the spectroscopic results (at-
Section 4.3. There are advantages in fixing parameters to mospheric parameters and individual chemical abundances)
values derived by more accurate methods, but the effects of obtained when using several radiative transfer codes and dif-
this practice should be carefully assessed and controlled. ferent spectroscopic techniques with multiple set-ups.
Given the results presented in this work, the situation The results showed that the synthetic spectral technique
can be illustrated with the following analogy. We consider has a higher accuracy when considering the full range of at-
a given selection of model atmospheres, solar abundances, mospheric parameters that the Gaia Benchmark Stars cover,
atomic data, radiative transfer codes, normalization pro- while the equivalent-width method is competitive only when
cedures, general data treatments and spectral wavelength the data set is limited by discarding cooler and metal-poorer
ranges as a ruler in centimetres (for instance) with a preci- stars. When the right set-up is selected, interpolating from a

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 21
grid of pre-computed spectra has an accuracy almost com- Anstee S. D., O’Mara B. J., 1991, MNRAS, 253, 549
parable to the results obtained when using synthesis with Anstee S. D., O’Mara B. J., 1995, MNRAS, 276, 859
an interpolated model atmosphere. To obtain the best re- Astropy Collaboration et al., 2013, A&A, 558, A33
sults, an appropriate line selection should be executed, using Barber C. B., Dobkin D. P., Huhdanpaa H., 1996, ACM Trans.
the same model atmosphere, atomic data, radiative transfer Math. Softw., 22, 469
Barklem P. S., 2016, Astronomy and Astrophysics Review, 24, 9
code and normalization process that will be applied to the
Barklem P. S., Piskunov N., O’Mara B. J., 2000, A&A, 363, 1091
target spectra. In the case of synthesis, including the wings Barklem P. S., Stempels H. C., Allende Prieto C., Kochukhov
of H-α/β and the Mg triplet can help to increase the accu- O. P., Piskunov N., O’Mara B. J., 2002, A&A, 385, 951
racy when using Grid, SPECTRUM and SYNTHE codes. Blanco-Cuaresma S., 2017, in EWASS Special Session 4
Despite using a homogeneous process to select the (2017): Star-planet interactions (EWASS-SS4-2017. ,
best common absorption lines, there are clear systemat- doi:10.5281/zenodo.1056914
ics between the equivalent-width method and the synthetic Blanco-Cuaresma S., Fraix-Burnet D., 2018, A&A, 618, A65
spectral-fitting technique. In addition, for the latter, there Blanco-Cuaresma S., Soubiran C., 2016, in SF2A-2016: Proceed-
are also differences when interpolating from a pre-computed ings of the Annual meeting of the French Society of Astron-
grid of spectra or synthesizing with an interpolated atmo- omy and Astrophysics. pp 333–336
Blanco-Cuaresma S., Soubiran C., Jofré P., Heiter U., 2014a,
spheric model. This demonstrates that blindly combining at-
A&A, 566, A98
mospheric parameters and chemical abundances measured
Blanco-Cuaresma S., Soubiran C., Heiter U., Jofré P., 2014b,
using heterogeneous set-ups and methods is not a recom- A&A, 569, A111
mended procedure, especially when a high precision is key Blanco-Cuaresma S., et al., 2015, A&A, 577, A47
for the scientific goals of the study. Blanco-Cuaresma S., et al., 2017, in Highlights on Spanish Astro-
This study has uncovered code-to-code differences that physics IX. pp 334–337
can affect the scientific interpretation of spectroscopic anal- Brahm R., Jordán A., Hartman J., Bakos G., 2017, MNRAS, 467,
ysis; this suggests that it would be a good practice to assess 971
if the conclusions still hold when deriving atmospheric pa- Buitinck L., et al., 2013, in ECML PKDD Workshop: Languages
rameters and/or abundances with different codes. for Data Mining and Machine Learning. pp 108–122
There are plenty of models and tools freely accessible Casamiquela L., et al., 2016, MNRAS, 458, 3150
Casamiquela L., et al., 2017, MNRAS, 470, 4363
today to analyse the growing number of high-quality spec-
Casey A. R., 2016, ApJS, 223, 8
tra available in the public archives or to execute our own Cayrel R., van’t Veer-Menneret C., Allard N. F., Stehlé C., 2011,
observations and studies. This ease of access represents an A&A, 531, A83
unprecedented opportunity in the history of stellar spec- Cowley C. R., Castelli F., 2002, A&A, 387, 595
troscopy to investigate the wonders of the stars, but it carries Czekala I., Andrews S. M., Mandel K. S., Hogg D. W., Green
with it the need to consider carefully the caveats of modern G. M., 2015, ApJ, 812, 128
spectroscopy exposed here. Eisenstein D. J., et al., 2011, AJ, 142, 72
Gilmore G., et al., 2012, The Messenger, 147, 25
Giribaldi R. E., Ubaldo-Melo M. L., Porto de Mello G. F.,
Pasquini L., Ludwig H.-G., Ulmer-Moll S., Lorenzo-Oliveira
ACKNOWLEDGEMENTS D., 2018, arXiv e-prints, p. arXiv:1811.12274
This research has made use of NASA’s Astrophysics Data Gray D. F., 2008, The Observation and Analysis of Stellar Pho-
tospheres
System. This work would not have been possible without the
Gray R. O., Corbally C. J., 1994, AJ, 107, 742
invaluable contribution from all the authors that developed Grevesse N., Asplund M., Sauval A. J., 2007, Space Sci. Rev.,
the considered radiative transfer codes: Richard O. Gray, 130, 105
Robert L. Kurucz, Luca Sbordone, Nikolai Piskunov, Jeff Gustafsson B., Edvardsson B., Eriksson K., Jørgensen U. G.,
A. Valenti, Bertrand Plez, Chris Sneden plus any other ma- Nordlund Å., Plez B., 2008, A&A, 486, 951
jor contributor that I may have forgotten. Likewise, Bengt Hawkins K., et al., 2016, A&A, 592, A70
Gustafsson and contributors should be acknowledged for Heiter U., Eriksson K., 2006, A&A, 452, 1039
their major contribution to stellar astrophysics with their Heiter U., et al., 2015a, Phys. Scr., 90, 054010
MARCS model atmosphere. Additionally, I am very grateful Heiter U., Jofré P., Gustafsson B., Korn A. J., Soubiran C.,
to Thomas Nordlander, Ulrike Heiter, Paula Jofré, Thomas Thévenin F., 2015b, A&A, 582, A49
Hinkel N. R., Timmes F. X., Young P. A., Pagano M. D., Turnbull
Masseron, Laia Casamiquela, Hugo M. Tabernero, Andrew
M. C., 2014, AJ, 148, 54
R. Casey, Jorge Meléndez, and Ivan Ramı́rez for their help
Hinkel N. R., et al., 2016, ApJS, 226, 4
integrating and testing all these radiative transfer codes in Hinkle K., Wallace L., Valenti J., Harmer D., 2000, Visible and
iSpec. I thank the anonymous referee, whose great ideas and Near Infrared Atlas of the Arcturus Spectrum 3727-9300 A
suggestions significantly improved the quality of this work. Hubeny I., Lanz T., 2011, Synspec: General Spectrum Synthesis
Program (ascl:1109.022)
Jofré P., et al., 2014, A&A, 564, A133
Jofré P., et al., 2015, A&A, 582, A81
REFERENCES
Jofré P., et al., 2017, A&A, 601, A38
Adibekyan V., et al., 2015, A&A, 583, A94 Jofré P., Heiter U., Soubiran C., 2018, arXiv e-prints, p.
Ali A. W., Griem H. R., 1965, Physical Review, 140, 1044 arXiv:1811.08041
Ali A. W., Griem H. R., 1966, Physical Review, 144, 366 Jones E., Oliphant T., Peterson P., et al., 2001, SciPy: Open
Allende Prieto C., Beers T. C., Wilhelm R., Newberg H. J., Rock- source scientific tools for Python, http://www.scipy.org/
osi C. M., Yanny B., Lee Y. S., 2006, ApJ, 636, 804 Koleva M., Prugniel P., Bouchard A., Wu Y., 2009, A&A, 501,
Alvarez R., Plez B., 1998, A&A, 330, 1109 1269

MNRAS 000, 1–22 (2018)


22 S. Blanco-Cuaresma
Kurucz R., 1993, SYNTHE Spectrum Synthesis Programs and :: teff (dispersion) per star

Common+wings [R]
Line Data. Kurucz CD-ROM No. 18. Cambridge, Mass.:

Common+wings

Own+wings [R]
Smithsonian Astrophysical Observatory, 1993., 18

Common [R]
Majewski S. R., et al., 2017, AJ, 154, 94

Own+wings
Common
Mucciarelli A., Pancino E., Lovisi L., Ferraro F. R., Lapenna E.,

Own [R]
2013, ApJ, 766, 78

Own
18.50

12.50

13.00

10.50

11.50
8.50

7.00

9.00
Ness M., Hogg D. W., Rix H.-W., Ho A. Y. Q., Zasowski G., 2015, 40
ApJ, 808, 16
Niemczura E., Smalley B., Pych W., 2014, Determination of
MOOG EW 32 40 32 40 27 20 27 20
29.50
Atmospheric Parameters of B-, A-, F- and G-Type Stars, WIDTH9 38 48 38 48 23 43 23 43 35

Median dispersion per star


doi:10.1007/978-3-319-06956-2. 40.50
Nissen P. E., Gustafsson B., 2018, Astronomy and Astrophysics Grid 10 13 12 8 7 8 12 6 30
Review, 26, 6 9.00
Pedregosa F., et al., 2011, Journal of Machine Learning Research,
SPECTRUM 23 12 13 7 10 7 14 10
11.00 25
12, 2825 Turbospectrum 12 9 13 9 12 6 10 9
Plez B., 2012, Turbospectrum: Code for spectral synthesis, As- 9.50
trophysics Source Code Library (ascl:1205.004) SME 25 13 11 9 9 7 10 8 20
Randich S., Gilmore G., Gaia-ESO Consortium 2013, The Mes- 9.50
senger, 154, 47 MOOG Synth 10 9 14 8 10 7 11 8
9.50 15
Recio-Blanco A., Bijaoui A., de Laverny P., 2006, MNRAS, 370, SYNTHE 14 12 10 8 11 6 10 9
141 10.00
Sbordone L., Bonifacio P., Castelli F., Kurucz R. L., 2004, Mem- 10
orie della Societa Astronomica Italiana Supplementi, 5, 93
Smiljanic R., et al., 2014, A&A, 570, A122
Sneden C., Bean J., Ivans I., Lucatello S., Sobeck J., 2012, :: logg (dispersion) per star

Common+wings [R]
MOOG: LTE line analysis and spectrum synthesis, Astro-

Common+wings
physics Source Code Library (ascl:1202.009)

Own+wings [R]
Tabernero H. M., González Hernández J. I., Montes D., 2013, in
Common [R]

Own+wings
Guirado J. C., Lara L. M., Quilis V., Gorgas J., eds, Highlights
Common

Own [R]
of Spanish Astrophysics VII. pp 673–673

Own
The Astropy Collaboration et al., 2018, preprint,
0.03

0.01

0.02

0.01

0.02

0.01

0.03

0.02
(arXiv:1801.02634) 0.08
Tsantaki M., Andreasen D. T., Teixeira G. D. C., Sousa S. G., MOOG EW 0.06 0.07 0.06 0.07 0.06 0.03 0.06 0.03
0.06
Santos N. C., Delgado-Mena E., Bruzual G., 2018, MNRAS, WIDTH9 0.09 0.10 0.09 0.10 0.08 0.07 0.08 0.07 0.07

Median dispersion per star


473, 5066 0.08
Valenti J. A., Piskunov N., 1996, A&AS, 118, 595 Grid 0.03 0.01 0.02 0.03 0.01 0.01 0.01 0.01 0.06
Vollmann K., Eversberg T., 2006, Astronomische Nachrichten, 0.01
327, 862 SPECTRUM 0.04 0.02 0.02 0.01 0.02 0.01 0.01 0.01
0.01 0.05
Turbospectrum 0.03 0.02 0.02 0.02 0.02 0.02 0.02 0.02
0.02
SME 0.03 0.01 0.02 0.01 0.02 0.01 0.02 0.02 0.04
APPENDIX A: MODEL ATMOSPHERE AND 0.02
ATOMIC DATA MOOG Synth 0.03 0.01 0.01 0.01 0.02 0.01 0.03 0.02 0.03
0.01
SYNTHE 0.03 0.01 0.03 0.01 0.01 0.01 0.03 0.01
0.01 0.02
This paper has been typeset from a TEX/LATEX file prepared by
the author.
0.01

Figure 12. Median robust standard deviation of the effective


temperature and surface gravity per star (when multiple spectra
were available) when analysing the Gaia Benchmark Stars and
using different radiative transfer codes with several set-ups: using
lines in common within equivalent-width and synthesis methods;
using the best lines for each code (i.e. their own lines, which are
not necessarily good for other codes); using the wings of H-α/β
and the Mg triplet; and repeating the normalization (labelled as
”[R]”) but using a synthetic spectrum that matches the atmo-
spheric parameters found in the first iteration.

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 23
SPECTRUM Turbospectrum SME MOOG WIDTH9/SYNTHE

Column mass above each point [g cm− 2] X X X X

Temperature [K] X X X X X

Gas pressure [dyn cm−2 ] X X X X X

Electron density [cm−3 ] X X X X

Rosseland mean absorption coefficient [cm2 g−1 ] X X

Radiation pressure [dyn cm−2 ] X X

Microturbulence velocity [m s−1 ] X X

Optical depth [log τ at 5000 Å] X

Depth [cm] X X

Electron pressure [cm2 g−1 ] X

Turbulence pressure [dyn cm−2 ] X

Table A1. Model atmosphere fields required as input values for each radiative transfer code. This list respects the expected input order
as required by the used version of iSpec.

NARVAL Grid Turbospectrum MOOG Synth


ATLAS SPECTRUM SME SYNTHE
Common+wings [R]

1.0 H- 1.0 H-
Common+wings

Normalized flux Normalized flux


Own+wings [R]

0.8
Common [R]

Own+wings

0.9
Common

Own [R]

2.0 0.6
0.8
Own
8.04

8.82

7.93

8.36

6.69

7.37

7.03

7.27

1.0 1.0
(Normalized | Teff| + | logg| + | MH|)

MOOG EW 1.47 1.85 1.47 1.85 1.29 1.11 1.29 1.11 1.8
11.44 0.9 0.8
WIDTH9 1.58 1.78 1.58 1.78 1.30 1.67 1.30 1.67 1.6 0.6
12.66
0.8
Grid 0.66 0.84 0.69 0.65 0.64 0.80 0.58 0.65 655.5 656.5 657.5 485.6 486.1 486.6
Wavelength (nm) Wavelength (nm)
5.50 1.4
SPECTRUM 0.92 0.86 0.66 0.49 0.67 0.62 0.50 0.49
5.21
Turbospectrum 0.83 0.96 0.93 0.93 0.79 0.94 0.86 0.88 1.2
7.11
SME 0.91 0.92 0.90 0.99 0.61 0.75 0.88 0.85
6.80 1.0 Figure 14. Hydrogen line regions from a NARVAL-observed solar
MOOG Synth 0.90 0.72 1.05 1.13 0.70 0.77 1.01 1.06 spectrum and the solar ATLAS (top), plus the same NARVAL-
7.35 observed solar spectrum and the corresponding synthetic spectra
SYNTHE 0.79 0.89 0.65 0.54 0.69 0.72 0.60 0.57 0.8 generated with each code (bottom).
dispersions
median

5.44
0.6

Figure 13. Sum of the normalized absolute median differences


and normalized robust standard deviation and normalized median
robust standard deviation for effective temperature, surface grav-
ity and metallicity when analysing the Gaia Benchmark Stars.
Lower numbers indicate results closer to reference values and
lower dispersion (higher accuracy).

MNRAS 000, 1–22 (2018)


24 S. Blanco-Cuaresma
Grid :: teff -69 +/- 208 Grid :: teff 34 +/- 412 Grid :: teff -74 +/- 200
1000 1000 1000
500 500 500
0 0 0
500 500 500
1000 1000 1000
SPECTRUM :: teff -93 +/- 219 SPECTRUM :: teff 41 +/- 414 SPECTRUM :: teff -78 +/- 196
1000 1000 1000
500 500 500
0 0 0
500 500 500
1000 1000 1000
Turbospectrum :: teff -136 +/- 210 Turbospectrum :: teff -92 +/- 430 Turbospectrum :: teff -149 +/- 204
1000 1000 1000
500 500 500
0 0 0
500 500 500
1000 1000 1000
SME :: teff -162 +/- 221 SME :: teff -154 +/- 407 SME :: teff -174 +/- 203
1000 1000 1000
500 500 500
0 0 0
500 500 500
1000 1000 1000
MOOG Synth :: teff 131 +/- 313 MOOG Synth :: teff 696 +/- 612 MOOG Synth :: teff 304 +/- 423
1000 1000 1000
500 500 500
0 0 0
500 500 500
1000 1000 1000
SYNTHE :: teff -60 +/- 200 SYNTHE :: teff 55 +/- 397 SYNTHE :: teff -47 +/- 170
1000 1000 1000
500 500 500
0 0 0
500 500 500
1000 1000 1000
3796
3927
4197
4286
4474
4587
4858
4983
5059
5123
5308
5522
5792
5868
5902
6083
6121
6223
6554

3796
3927
4197
4286
4474
4587
4858
4983
5059
5123
5308
5522
5792
5868
5902
6083
6121
6223
6554

3796
3927
4197
4286
4474
4587
4858
4983
5059
5123
5308
5522
5792
5868
5902
6083
6121
6223
6554
Figure 15. Difference between the derived effective temperature and its reference value for each spectrum analysed with the synthetic
spectral-fitting codes when using only the wings of H-α (left subplots), H-β (middle subplots) and both of them together (right subplots).
The vertical thick grey line denotes the Sun. The colour-coding represents the metallicity of each star (see right subplots in Fig. 16 for
colour-code interpretation). All the subplots are sorted taking into account the reference effective temperature. The median and absolute
median deviation are indicated in the upper right of each subplot.

MOOG EW :: teff 69 +/- 236 MOOG EW :: logg 0.02 +/- 0.41 MOOG EW :: MH 0.04 +/- 0.12
1000 2 1.0
500 1 0.5
0 0 0.0
500 1 0.5
1000 2 1.0
WIDTH9 :: teff 5 +/- 243 WIDTH9 :: logg -0.08 +/- 0.36 WIDTH9 :: MH -0.01 +/- 0.13
1000 2 1.0
500 1 0.5
0 0 0.0
500 1 0.5
1000 2 1.0
3796
3927
4197
4286
4474
4587
4858
4983
5059
5123
5308
5522
5792
5868
5902
6083
6121
6223
6554

0.68
1.05
1.43
1.64
2.51
2.77
2.9
3.58
3.79
4.0
4.1
4.2
4.27
4.3
4.41
4.44
4.53
4.6
4.63

-2.43
-1.46
-1.31
-1.23
-1.0
-0.82
-0.62
-0.5
-0.45
-0.37
-0.33
-0.1
-0.05
0.01
0.06
0.13
0.21
0.24
0.3

Figure 16. Difference between the derived parameter and its reference value for each spectra analyzed with the equivalent method
width when using their best line selection (i.e., own lines). The vertical thick gray line denotes the Sun. The color coding represent the
metallicity of each star. All the subplots are sorted taking into account the reference value of the corresponding atmospheric parameter.
Median and absolute median deviation are indicated on the upper right of each subplot.

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 25
Grid :: teff -3 +/- 135 Grid :: logg -0.09 +/- 0.22 Grid :: MH -0.05 +/- 0.06
1000 2 1.0
500 1 0.5
0 0 0.0
500 1 0.5
1000 2 1.0
SPECTRUM :: teff 7 +/- 157 SPECTRUM :: logg -0.05 +/- 0.25 SPECTRUM :: MH -0.04 +/- 0.07
1000 2 1.0
500 1 0.5
0 0 0.0
500 1 0.5
1000 2 1.0
Turbospectrum :: teff -10 +/- 120 Turbospectrum :: logg -0.16 +/- 0.27 Turbospectrum :: MH -0.05 +/- 0.07
1000 2 1.0
500 1 0.5
0 0 0.0
500 1 0.5
1000 2 1.0
SME :: teff -1 +/- 162 SME :: logg -0.01 +/- 0.22 SME :: MH -0.05 +/- 0.09
1000 2 1.0
500 1 0.5
0 0 0.0
500 1 0.5
1000 2 1.0
MOOG Synth :: teff 2 +/- 178 MOOG Synth :: logg 0.00 +/- 0.40 MOOG Synth :: MH -0.05 +/- 0.08
1000 2 1.0
500 1 0.5
0 0 0.0
500 1 0.5
1000 2 1.0
SYNTHE :: teff 2 +/- 140 SYNTHE :: logg -0.05 +/- 0.24 SYNTHE :: MH -0.05 +/- 0.07
1000 2 1.0
500 1 0.5
0 0 0.0
500 1 0.5
1000 2 1.0
3796
3927
4197
4286
4474
4587
4858
4983
5059
5123
5308
5522
5792
5868
5902
6083
6121
6223
6554

0.68
1.05
1.43
1.64
2.51
2.77
2.9
3.58
3.79
4.0
4.1
4.2
4.27
4.3
4.41
4.44
4.53
4.6
4.63

-2.43
-1.46
-1.31
-1.23
-1.0
-0.82
-0.62
-0.5
-0.45
-0.37
-0.33
-0.1
-0.05
0.01
0.06
0.13
0.21
0.24
0.3
Figure 17. As Fig. 16, but for spectra analyzed with the synthetic spectral-fitting technique.

MNRAS 000, 1–22 (2018)


26 S. Blanco-Cuaresma
SPECTRUM Turbospectrum SME MOOG WIDTH9/SYNTHE

Element name X X X X

Wavelength [Å] X X X X

Wavelength [nm] X

loggf X X X X X

Lower state [eV] X X X

Lower state [cm−1 ] X X

Lower j or total angular momentum quantum number X

Upper state [eV]

Upper state [cm−1 ] X

Upper j or total angular momentum quantum number X

Upper g or statistical weight X

Lower Landé g-factor X

Upper Landé g-factor X

Transition type: X X

10 to the power of the radiative damping parameter X X

Radiative damping parameter X X X

Stark damping parameter X X X

van der Waals damping parameter (σ.α format for AO theory) X X

van der Waals damping parameter (classic) X X

Fudge factor (common for the same atomic number and ion)
or van der Waals damping parameter (classic or AO theory) if present X

Fudge factor (always set to 1.0) X

Lower orbital type X

Upper orbital type X

Line due to molecular absorption (True/False) X X X X X

Isotope in spectrum format X X

Ion (e.g., 0 for neutral lines, 1 for ionized) X X

Species code: ”atomic number” + ”.” + ”ion state - 1” X X

Species code: ”atomic number” + ”.” + ”isotope code” X

Species code: ”atomic number” + ”.0” + ”ion state - 1” X

Table A2. Atomic line list fields required as input values for each radiative transfer code. Transition type indicates whether the α and
σ parameters used in the Anstee and O’Mara broadening theory are provided (Anstee & O’Mara 1991, 1995, coded as AO type;) or the
classic van der Waals broadening should be used (GA type) as described in SPECTRUM documentation. Fudge factors are arbitrary
non-physical values used to increase the line broadening to compensate for unknowns. This list respects the expected input order as
required by the used version of iSpec.

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 27
WIDTH9 - MOOG EW -43 +/- 105 SPECTRUM - MOOG EW -90 +/- 160
1000 1000
500 500
0 0

teff
teff

500 500
1000 1000

3807
4197
4374
4587
4954
5059
5231
5522
5810
5902
6099
6223
6635
3807
4197
4374
4587
4954
5059
5231
5522
5810
5902
6099
6223
2
6635
WIDTH9 - MOOG EW -0.09 +/- 0.16
2
SPECTRUM - MOOG EW -0.17 +/- 0.38
1 1
0 0
logg
logg

1 1
2 2
1.05
1.43
2.09
2.77
3.52
3.79
4.06
4.2
4.3
4.41
4.49
4.6
4.67
1.05
1.43
2.09
2.77
3.52
3.79
4.06
4.2
4.3
4.41
4.49
4.6
4.67

WIDTH9 - MOOG EW -0.03 +/- 0.06 SPECTRUM - MOOG EW -0.04 +/- 0.07
1.0 1.0
0.5 0.5
0.0 0.0
MH
MH

0.5 0.5
1.0 1.0
-2.09
-1.31
-1.16
-0.82
-0.53
-0.45
-0.34
-0.1
-0.04
0.06
0.14
0.24
0.33
-2.09
-1.31
-1.16
-0.82
-0.53
-0.45
-0.34
-0.1
-0.04
0.06
0.14
0.24
0.33

Figure 18. Differences between the two equivalent widths codes


(left subplots), and one code from each method (right subplots)
when using their best line selection (i.e. own lines). The vertical
thick grey line denotes the Sun. The colour-coding represents the
metallicity of each star. All the subplots are sorted taking into
account the reference value of the corresponding atmospheric pa-
rameter. The median and absolute median deviation are indicated
in the upper right of each subplot.

MNRAS 000, 1–22 (2018)


28 S. Blanco-Cuaresma
Grid-SPECTRUM -14 +/- 47 Turbo.-SPECTRUM -11 +/- 19 SME-SPECTRUM -9 +/- 21 MOOG S.-SPECTRUM 15 +/- 31 SYNTHE-SPECTRUM -2 +/- 29
1000
500
0
teff

500
1000
3807
4197
4374
4587
4954
5059
5231
5522
5810
5902
6099
6223
6635

3807
4197
4374
4587
4954
5059
5231
5522
5810
5902
6099
6223
6635

3807
4197
4374
4587
4954
5059
5231
5522
5810
5902
6099
6223
6635

3807
4197
4374
4587
4954
5059
5231
5522
5810
5902
6099
6223
6635

3807
4197
4374
4587
4954
5059
5231
5522
5810
5902
6099
6223
6635
Grid-SPECTRUM -0.04 +/- 0.07 Turbo.-SPECTRUM -0.04 +/- 0.14 SME-SPECTRUM 0.00 +/- 0.04 MOOG S.-SPECTRUM 0.03 +/- 0.09 SYNTHE-SPECTRUM 0.00 +/- 0.04
2
1
0
logg

1
2
1.05
1.43
2.09
2.77
3.52
3.79
4.06
4.2
4.3
4.41
4.49
4.6
4.67

1.05
1.43
2.09
2.77
3.52
3.79
4.06
4.2
4.3
4.41
4.49
4.6
4.67

1.05
1.43
2.09
2.77
3.52
3.79
4.06
4.2
4.3
4.41
4.49
4.6
4.67

1.05
1.43
2.09
2.77
3.52
3.79
4.06
4.2
4.3
4.41
4.49
4.6
4.67

1.05
1.43
2.09
2.77
3.52
3.79
4.06
4.2
4.3
4.41
4.49
4.6
4.67
Grid-SPECTRUM 0.00 +/- 0.02 Turbo.-SPECTRUM 0.01 +/- 0.01 SME-SPECTRUM 0.00 +/- 0.01 MOOG S.-SPECTRUM 0.01 +/- 0.03 SYNTHE-SPECTRUM -0.01 +/- 0.02
1.0
0.5
0.0
MH

0.5
1.0
-2.09
-1.31
-1.16
-0.82
-0.53
-0.45
-0.34
-0.1
-0.04
0.06
0.14
0.24
0.33

-2.09
-1.31
-1.16
-0.82
-0.53
-0.45
-0.34
-0.1
-0.04
0.06
0.14
0.24
0.33

-2.09
-1.31
-1.16
-0.82
-0.53
-0.45
-0.34
-0.1
-0.04
0.06
0.14
0.24
0.33

-2.09
-1.31
-1.16
-0.82
-0.53
-0.45
-0.34
-0.1
-0.04
0.06
0.14
0.24
0.33

-2.09
-1.31
-1.16
-0.82
-0.53
-0.45
-0.34
-0.1
-0.04
0.06
0.14
0.24
0.33
Figure 19. As Fig. 18, but only for the synthetic spectral-fitting technique.

:: Common lines :: Own lines


Turbospectrum

Turbospectrum
MOOG Synth

MOOG Synth
SPECTRUM

SPECTRUM
MOOG EW

MOOG EW
WIDTH9

WIDTH9
SYNTHE

SYNTHE
SME

SME
Grid

Grid
(Normalized | Teff| + | logg| + | MH|)

(Normalized | Teff| + | logg| + | MH|)


MOOG EW 0.00 0.43 1.60 1.14 1.32 1.13 1.72 1.20 2.00 MOOG EW 0.00 1.56 1.53 1.39 1.29 1.31 1.11 1.33 2.00
8.54 9.52
WIDTH9 0.43 0.00 1.63 1.52 1.77 1.52 2.02 1.63 1.75 WIDTH9 1.56 0.00 1.43 1.57 1.71 1.55 1.93 1.56 1.75
10.53 11.31
Grid 1.60 1.63 0.00 0.30 0.19 0.28 0.95 0.34 1.50 Grid 1.53 1.43 0.00 0.31 0.26 0.37 0.65 0.41 1.50
5.30 4.95
SPECTRUM 1.14 1.52 0.30 0.00 0.22 0.34 0.53 0.10 1.25 SPECTRUM 1.39 1.57 0.31 0.00 0.33 0.14 0.52 0.16 1.25
4.15 4.42
Turbospectrum 1.32 1.77 0.19 0.22 0.00 0.19 0.74 0.20 1.00 Turbospectrum 1.29 1.71 0.26 0.33 0.00 0.34 0.59 0.55 1.00
4.64 5.06
SME 1.13 1.52 0.28 0.34 0.19 0.00 0.69 0.23 0.75 SME 1.31 1.55 0.37 0.14 0.34 0.00 0.43 0.25 0.75
4.39 4.38
MOOG Synth 1.72 2.02 0.95 0.53 0.74 0.69 0.00 0.55 0.50 MOOG Synth 1.11 1.93 0.65 0.52 0.59 0.43 0.00 0.42 0.50
7.21 5.65
0.25 0.25
dispersions

dispersions
SYNTHE 1.20 1.63 0.34 0.10 0.20 0.23 0.55 0.00 SYNTHE 1.33 1.56 0.41 0.16 0.55 0.25 0.42 0.00
median

median
4.25 4.69
Teff > 4500 K and [Fe/H] > -1.0 dex
0.00 Teff > 4500 K and [Fe/H] > -1.0 dex
0.00

Figure 20. As Fig. 10, but considering only Gaia Benchmark Stars with effective temperatures greater than 4 500 K and metallicities
greater than -1.0 dex.

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 29

Common+wings [R]
Common+wings

Own+wings [R]
Common [R]

Own+wings
Common

Own [R]
2.0
Own
8.82

8.93

8.47

8.38

7.18

7.72

7.60

7.62

(Normalized | Teff| + | logg| + | MH|)


MOOG EW 1.15 1.24 1.15 1.24 0.97 0.79 0.97 0.79 1.8
8.32
WIDTH9 1.20 1.39 1.20 1.39 1.06 1.35 1.06 1.35
10.01 1.6
Grid 0.87 1.03 0.88 0.83 0.77 0.95 0.74 0.79
6.87
SPECTRUM 1.21 1.06 0.83 0.62 0.85 0.78 0.65 0.62 1.4
6.63
Turbospectrum 1.03 1.13 1.14 1.11 0.96 1.10 1.04 1.05
8.56 1.2
SME 1.17 1.11 1.14 1.20 0.80 0.91 1.09 1.03
8.44
MOOG Synth 1.18 0.91 1.28 1.33 0.92 0.97 1.28 1.29 1.0
9.16
SYNTHE 1.00 1.08 0.84 0.66 0.85 0.87 0.77 0.70
dispersions

0.8
median

6.76

:: Teff > 4500 K and [Fe/H] > -1.0 dex

Figure 21. As Fig. 13, but considering only Gaia Benchmark


Stars with effective temperatures greater than 4 500 K and metal-
licities greater than -1.0 dex.

MNRAS 000, 1–22 (2018)


30 S. Blanco-Cuaresma
:: [Fe/H] (median) :: [Fe/H] (dispersion)

Turbospectrum

Turbospectrum
Own lines all fixed Own lines all fixed

MOOG Synth

MOOG Synth
SPECTRUM

SPECTRUM
MOOG EW

MOOG EW
WIDTH9

WIDTH9
SYNTHE

SYNTHE
SME

SME
Grid

Grid
0.20 0.30
MOOG EW 0.00 0.00 -0.11 -0.11 -0.08 -0.10 -0.10 -0.11 MOOG EW 0.00 0.02 0.19 0.19 0.16 0.15 0.14 0.16
0.10 0.15
0.15
WIDTH9 0.00 0.00 -0.09 -0.08 -0.08 -0.09 -0.09 -0.08 WIDTH9 0.02 0.00 0.13 0.15 0.12 0.13 0.13 0.13 0.25
0.08 0.13
0.10
Grid 0.11 0.09 0.00 0.00 0.01 0.00 -0.01 -0.01 Grid 0.19 0.13 0.00 0.01 0.04 0.03 0.04 0.02
0.20

[Fe/H] (dispersion)
0.01 0.04

[Fe/H] (median)
0.05
SPECTRUM 0.11 0.08 0.00 0.00 0.01 0.00 0.00 0.00 SPECTRUM 0.19 0.15 0.01 0.00 0.03 0.01 0.04 0.02
0.00 0.03
0.00 0.15
Turbospectrum 0.08 0.08 -0.01 -0.01 0.00 -0.01 -0.01 -0.01 Turbospectrum 0.16 0.12 0.04 0.03 0.00 0.02 0.01 0.02
0.01 0.03
0.05
SME 0.10 0.09 0.00 0.00 0.01 0.00 0.00 0.00 SME 0.15 0.13 0.03 0.01 0.02 0.00 0.03 0.01 0.10
0.00 0.03
0.10
MOOG Synth 0.10 0.09 0.01 0.00 0.01 0.00 0.00 0.00 MOOG Synth 0.14 0.13 0.04 0.04 0.01 0.03 0.00 0.03
0.01 0.04 0.05
0.15
SYNTHE 0.11 0.08 0.01 0.00 0.01 0.00 0.00 0.00 SYNTHE 0.16 0.13 0.02 0.02 0.02 0.01 0.03 0.00
0.01 0.02
0.20 0.00

Figure 22. Median and robust standard deviation of the difference in iron abundance between different radiative transfer codes when
analysing the Gaia Benchmark Stars and fixing all the atmospheric parameters to the reference ones and using the best line selection for
each code.

:: Own lines all fixed


Turbospectrum

MOOG Synth
SPECTRUM
MOOG EW
WIDTH9

SYNTHE
SME
Grid

MOOG EW 0.00 1.59 7.26 6.94 6.05 6.26 6.33 6.25 7


40.67
(Normalized | [Element/H]|)

WIDTH9 1.59 0.00 6.08 5.91 5.44 5.77 5.31 5.57


35.67 6
Grid 7.26 6.08 0.00 0.64 0.93 0.94 1.05 0.83
17.73 5
SPECTRUM 6.94 5.91 0.64 0.00 0.77 0.54 0.85 0.55
16.19 4
Turbospectrum 6.05 5.44 0.93 0.77 0.00 0.72 0.53 0.68
15.11 3
SME 6.26 5.77 0.94 0.54 0.72 0.00 0.61 0.30
15.14 2
dispersions

MOOG Synth 6.33 5.31 1.05 0.85 0.53 0.61 0.00 0.60
median

15.29
1
SYNTHE 6.25 5.57 0.83 0.55 0.68 0.30 0.60 0.00
14.77
0

Figure 23. Sum of the normalized absolute median differ-


ences and robust standard deviation for iron, calcium, cobalt,
chromium, magnesium, manganese, nickel, silicon, titanium and
vanadium abundances when analysing the Gaia Benchmark Stars.
Lower numbers indicate the codes lead to more similar results
(higher precision).

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 31

0.20

[Mg/H]
[Mn/H]
[Ca/H]
[Co/H]
[Mg/H]
[Mn/H]

[Fe/H]

[Sc/H]
[Cr/H]
[Ca/H]
[Co/H]

[Ni/H]
[Fe/H]

[Sc/H]

[Si/H]
[Ti/H]
[Cr/H]

[Ni/H]

[V/H]
[Si/H]
[Ti/H]
[V/H]
0.10
0.18
MOOG EW 0.05 0.02 0.22 0.11 0.09 0.08 0.09 0.07 0.14 0.05 0.19 MOOG EW 0.08 0.08 0.30 0.16 0.17 0.27 0.13 0.10 0.19 0.10 0.22
0.09 0.16
WIDTH9 0.05 0.04 0.19 0.14 0.06 0.07 0.08 0.08 0.05 0.07 0.09 WIDTH9 0.09 0.11 0.25 0.17 0.16 0.21 0.13 0.11 0.08 0.13 0.14 0.16
0.07 0.05 0.13
Grid -0.05-0.09-0.03-0.03-0.010.01-0.03-0.04-0.03-0.03-0.03 Grid 0.10 0.10 0.08 0.07 0.05 0.08 0.10 0.10 0.06 0.07 0.04

Difference dispersion
Median difference
0.03 0.08 0.14
SPECTRUM -0.06-0.08-0.03-0.050.00 0.03-0.02-0.05-0.02-0.03-0.01 SPECTRUM 0.12 0.10 0.09 0.09 0.06 0.08 0.13 0.10 0.08 0.05 0.13
0.03 0.09
Turbospectrum -0.03-0.09-0.02-0.010.00 0.02-0.02-0.02-0.01-0.03-0.02 0.00 Turbospectrum 0.10 0.11 0.05 0.04 0.06 0.09 0.06 0.09 0.06 0.05 0.07
0.02 0.06
0.12
SME -0.07-0.08-0.02-0.06-0.02-0.01-0.03-0.020.00-0.03-0.02 SME 0.13 0.11 0.05 0.08 0.05 0.10 0.09 0.11 0.07 0.04 0.10
0.02 0.09
MOOG Synth -0.03-0.08-0.01-0.010.00-0.00-0.02-0.02-0.01-0.02-0.02 0.05 MOOG Synth 0.09 0.13 0.06 0.08 0.06 0.10 0.06 0.10 0.07 0.05 0.12 0.10
0.02 0.08
SYNTHE -0.06-0.08-0.03-0.03-0.010.01-0.03-0.02-0.01-0.03-0.03 SYNTHE 0.12 0.12 0.06 0.05 0.07 0.10 0.10 0.10 0.06 0.04 0.10
0.03 0.10 0.08
0.10
0.06
:: Median difference :: Difference dispersion
Own lines all fixed Own lines all fixed

Figure 24. Median and robust standard deviation of the difference in individual abundances between the reference values and different
radiative transfer codes when fixing all the atmospheric parameters to the reference ones when analysing the Gaia Benchmark Stars.
The best line selection for each code was used. The median dispersion per star is equal or below 0.02 dex for all the cases.

:: Own lines all fixed


Turbospectrum

MOOG Synth
SPECTRUM
MOOG EW
WIDTH9

SYNTHE
SME
Grid

MOOG EW 0.00 1.43 6.38 6.16 5.82 5.55 5.51 6.04 7


36.90
(Normalized | [Element/H]|)

WIDTH9 1.43 0.00 6.00 5.92 5.12 5.63 5.69 5.61


35.39 6
Grid 6.38 6.00 0.00 1.21 1.77 1.78 1.92 1.59
20.66 5
SPECTRUM 6.16 5.92 1.21 0.00 1.38 1.09 1.61 1.11
18.48 4
Turbospectrum 5.82 5.12 1.77 1.38 0.00 1.43 1.23 1.25
18.00 3
SME 5.55 5.63 1.78 1.09 1.43 0.00 1.28 0.54
17.30 2
dispersions

MOOG Synth 5.51 5.69 1.92 1.61 1.23 1.28 0.00 1.21
median

18.44
1
SYNTHE 6.04 5.61 1.59 1.11 1.25 0.54 1.21 0.00
17.36
Teff > 4500 K and [Fe/H] > -1.0 dex
0

Figure 25. As Fig. 23, but considering only Gaia Benchmark


Stars with effective temperatures greater than 4 500 K and metal-
licities greater than -1.0 dex.

MNRAS 000, 1–22 (2018)


32 S. Blanco-Cuaresma
* Teff > 4500 K and [Fe/H] > -1.0 dex * Teff > 4500 K and [Fe/H] > -1.0 dex
0.10

[Mg/H]
[Mn/H]
[Ca/H]
[Co/H]
[Fe/H]

[Sc/H]
[Cr/H]
[Mg/H]
[Mn/H]

[Ni/H]

[Si/H]
[Ca/H]
[Co/H]

[Ti/H]
[Fe/H]

[Sc/H]
[Cr/H]

[V/H]
[Ni/H]

[Si/H]
[Ti/H]
0.04

[V/H]
0.09
MOOG EW 0.04 0.00 0.05 0.03 0.00 0.00 0.02 0.04 0.02 0.02 0.08 MOOG EW 0.07 0.04 0.09 0.05 0.04 0.12 0.03 0.06 0.05 0.06 0.12
0.02 0.06
WIDTH9 0.02 0.00 0.04 0.04 0.00 0.04 0.01 0.05 0.02 0.02 0.03 0.02 WIDTH9 0.04 0.09 0.08 0.06 0.19 0.12 0.04 0.07 0.03 0.06 0.07 0.08
0.02 0.07
Grid 0.10 0.07 0.05 0.04 0.05 0.08 0.05 0.07 0.04 0.03 0.04

Difference dispersion
Grid -0.06-0.05-0.01-0.02-0.010.00-0.03-0.02-0.03-0.03-0.03

Median difference
0.03 0.05
SPECTRUM -0.04-0.040.00-0.020.00 0.01-0.02-0.02-0.03-0.03-0.02 SPECTRUM 0.09 0.09 0.06 0.05 0.03 0.10 0.06 0.09 0.06 0.03 0.09 0.07
0.02 0.06
Turbospectrum -0.03-0.030.00-0.020.00-0.01-0.02-0.01-0.02-0.03-0.02 0.00 Turbospectrum 0.10 0.07 0.03 0.04 0.04 0.06 0.03 0.07 0.04 0.03 0.07
0.02 0.04 0.06
SME -0.03-0.04-0.01-0.03-0.01-0.02-0.03-0.01-0.02-0.03-0.02 SME 0.09 0.07 0.02 0.04 0.04 0.07 0.06 0.10 0.05 0.03 0.04
0.02 0.05
MOOG Synth -0.02-0.03-0.00-0.030.00-0.01-0.03-0.01-0.03-0.03-0.02 MOOG Synth 0.07 0.08 0.03 0.04 0.06 0.07 0.06 0.09 0.05 0.03 0.05 0.05
0.02 0.02 0.06
SYNTHE -0.04-0.04-0.01-0.03-0.01-0.02-0.03-0.02-0.02-0.03-0.03 SYNTHE 0.10 0.07 0.03 0.06 0.04 0.09 0.06 0.10 0.04 0.03 0.04
0.03 0.06
0.04
0.04
:: Median difference :: Difference dispersion 0.03
Own lines all fixed Own lines all fixed

Figure 26. As Fig. 24 but considering only Gaia Benchmark Stars with effective temperatures greater than 4 500 K and metallicities
greater than -1.0 dex.

:: Common lines :: Own lines all fixed


Turbospectrum

Turbospectrum
Synthetic data set Synthetic data set
MOOG Synth

MOOG Synth
SPECTRUM

SPECTRUM
MOOG EW

MOOG EW
WIDTH9

WIDTH9
SYNTHE

SYNTHE
SME

SME
Grid

Grid
10
(Normalized | Teff| + | logg| + | MH|)

MOOG EW 0.00 4.35 8.17 8.48 9.78 8.98 8.75 8.34 MOOG EW 0.00 7.84 39.0938.1333.6335.8235.6837.46 40
56.86 227.65

(Normalized | [Element/H]|)
WIDTH9 4.35 0.00 8.38 7.32 8.38 8.21 9.27 7.63 WIDTH9 7.84 0.00 37.8337.3733.2836.0534.5237.21 35
53.53 8 224.10
Grid 8.17 8.38 0.00 2.13 2.96 2.48 6.02 2.54 Grid 39.0937.83 0.00 5.60 7.66 8.44 9.04 7.74 30
32.70 115.40
SPECTRUM 8.48 7.32 2.13 0.00 1.98 1.19 4.86 0.78 6 SPECTRUM 38.1337.37 5.60 0.00 6.45 5.45 7.39 5.17 25
26.74 105.56
Turbospectrum 9.78 8.38 2.96 1.98 0.00 2.46 7.20 2.56 Turbospectrum 33.6333.28 7.66 6.45 0.00 5.48 2.72 5.52 20
35.33 4 94.75
SME 8.98 8.21 2.48 1.19 2.46 0.00 4.80 0.99 SME 35.8236.05 8.44 5.45 5.48 0.00 6.43 3.48 15
29.12 101.17
10

dispersions
MOOG Synth 8.75 9.27 6.02 4.86 7.20 4.80 0.00 3.77 MOOG Synth 35.6834.52 9.04 7.39 2.72 6.43 0.00 5.77

median
2
44.66 101.56
5
dispersions

SYNTHE 8.34 7.63 2.54 0.78 2.56 0.99 3.77 0.00 SYNTHE 37.4637.21 7.74 5.17 5.52 3.48 5.77 0.00
median

26.61 102.35
0 0

Figure 27. As Fig. 10 and 23 (left and right plots, respectively), but using the results from the analysis of the synthetic spectra built
for the non-observed data set experiment.

MNRAS 000, 1–22 (2018)


Modern stellar spectroscopy caveats 33

MOOG EW WIDTH9 SPECTRUM SME SYNTHE


Turbospectrum MOOG Synth Grid
0.1 0.05
1 0.0
median(abundance - abundanceref)

0.0 0.00

median(line depth - line depthref)


0
0.5
4000 6000 8000 1 2 3 4 5 4000 6000 8000 2 4
teff logg teff logg
0.1
0.00 0.0 0.00
0.0 0.5
0.25 0.05
4 2 0 0.4 0.2 0.0 0.2 0.4 4 2 0 0.25 0.00 0.25
MH alpha MH alpha
0.00 0.1 0.0 0.2

0.25 0.1 0.0


0.0
1 2 3 10 20 30 40 50 1 2 3 20 40
vmic nlayers vmic nlayers

Figure 28. Median and absolute median deviation (error bars) Figure 30. As Fig. 29, but considering the depth at the line
difference in abundance over reference abundance (i.e. MOOG peaks for the common selection.
abundances for the Sun, zero-point represented by the intersection
of the two thick grey lines) when varying selected atmospheric
parameters and setting the rest to the reference values for the Sun. MOOG EW WIDTH9
The lower right plot shows the results when all the atmospheric
0.2 5 0.2
parameters are set to the solar reference and the number of layers
median(abundance - abundanceref)

in the atmospheric model is reduced. 0.0 0 0.0


0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5
rhox temperature pgas
0.02 0.02
SPECTRUM SME SYNTHE 0.2
Turbospectrum MOOG Synth Grid
median(continuum depth - continuum depthref)

0.01 0.001 0.0 0.00 0.00


0.000 0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5
xne abross accrad
0.00 0.001 0.02 0.02 0.02
4000 6000 8000 2 4
teff logg
0.001 0.00 0.00 0.00
0.0025 0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5
0.000 vturb logtau5 pelectron
0.0000
4 2 0 0.25 0.00 0.25
MH alpha
0.0005
0.0000 0.0025
Figure 31. Median and absolute median deviation (error bars)
0.0005 0.0000 difference in abundance over the reference abundance (i.e. MOOG
1 2 3 20 40 abundances for the Sun, zero-point represented by the intersection
vmic nlayers
of the two thick grey lines) when varying selected values of the
solar model atmosphere.

Figure 29. Median difference in continuum depth around 556.45


nm over the reference continuum depth (i.e. SPECTRUM contin-
uum depth for the Sun, zero-point represented by the intersection
of the two thick grey lines) when varying selected atmospheric pa-
rameters and setting the rest to the reference values for the Sun.
The lower right plot shows the results when all the atmospheric
parameters are set to the solar reference and the number of layers
in the atmospheric model is reduced.

MNRAS 000, 1–22 (2018)


34 S. Blanco-Cuaresma
SPECTRUM SME SYNTHE MOOG EW WIDTH9
Turbospectrum MOOG Synth
0.001 0.02
median(continuum depth - continuum depthref)

0.0000 0.5

median(abundance - abundanceref)
0.025 0.000
0.0005
0.000 0.001
0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5 0.0
rhox temperature pgas 0.00
0.0000 0.0000 0.0000 0.5 1.0 1.5 0.5 1.0 1.5
0.0005 0.0005 0.0005 loggf rad
0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5 0.02 0.1
xne abross accrad
0.002
0.0000 0.0000 0.01
0.000
0.0
0.0005 0.0005 0.00
0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5
vturb logtau5 pelectron 0.5 1.0 1.5 0.5 1.0 1.5
Stark Waals

Figure 32. Median and absolute median deviation (error bars) in


continuum depth around 556.45 nm over the reference continuum Figure 34. Median and absolute median deviation (error bars)
depth (i.e. SPECTRUM continuum depth for the Sun, zero-point difference in abundance over the reference abundance (i.e. MOOG
represented by the intersection of the two thick grey lines) when abundances for the Sun, zero-point represented by the intersection
varying different values of the solar model atmosphere. of the two thick grey lines) when varying selected values of the
atomic line list.

SPECTRUM SME SYNTHE


Turbospectrum MOOG Synth
SPECTRUM SME SYNTHE
0.0 0.5 0.00 Turbospectrum MOOG Synth
0.0005
median(continuum depth - depthref)

0.0
median(line depth - line depthref)

0.05 0.001
0.1 0.5 0.0000
0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5 0.000
rhox temperature pgas 0.0005
0.005 0.005 0.001
0.00 0.5 1.0 1.5 0.5 1.0 1.5
0.05 0.000 0.000 loggf rad
0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5 0.001
xne abross accrad 0.0000
0.000
0.005 0.00 0.01 0.0005 0.001
0.000 0.05 0.00
0.01 0.5 1.0 1.5 0.5 1.0 1.5
0.5 1.0 1.5 0.5 1.0 1.5 0.5 1.0 1.5 Stark Waals
vturb logtau5 pelectron

Figure 35. Median and absolute median deviation (error bars)


Figure 33. As Fig. 32, but considering the depth at the line difference in continuum depth around 556.45 nm over the refer-
peaks for the common selection. ence continuum depth (i.e., SPECTRUM continuum depth for the
Sun, zero point represented by the intersection of the two thick
grey lines) when varying different values of the atomic line list.

SPECTRUM SME SYNTHE


Turbospectrum MOOG Synth
median(line depth - line depthref)

0.0 0.005
0.000
0.2 0.005
0.5 1.0 1.5 0.5 1.0 1.5
loggf rad
0.005 0.01
0.000 0.00
0.01
0.005
0.5 1.0 1.5 0.5 1.0 1.5
Stark Waals

Figure 36. As Fig. 35, but considering the depth at the line
peaks for the common selection.

MNRAS 000, 1–22 (2018)

You might also like