+ All Categories
Home > Documents > Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as...

Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as...

Date post: 04-Nov-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
8
Draft version May 16, 2018 Preprint typeset using L A T E X style emulateapj v. 12/16/11 CONSISTENT MODELING OF GS 1826-24 X-RAY BURSTS FOR MULTIPLE ACCRETION RATES DEMONSTRATES THE POSSIBILITY TO CONSTRAIN RP -PROCESS REACTION RATES Zach Meisel Institute of Nuclear & Particle Physics, Department of Physics & Astronomy, Ohio University, Athens, Ohio 45701, USA Draft version May 16, 2018 ABSTRACT Type-I X-ray burst light curves encode unique information about the structure of accreting neutron stars and the nuclear reaction rates of the rp-process that powers bursts. Using the first model cal- culations of hydrogen/helium burning bursts for a large range of astrophysical conditions performed with the code MESA, this work shows that simultaneous model-observation comparisons for bursts from several accretion rates ˙ M are required to remove degeneracies in astrophysical conditions that other- wise reproduce bursts for a single ˙ M and that such consistent multi-epoch modeling could possibly limit the 15 O(α, γ ) 19 Ne reaction rate. Comparisons to the year 1998, 2000, and 2007 bursting epochs of the neutron star GS 1826-24 show that ˙ M must be larger than previously inferred and that the shallow heating in this source must be below 0.5 MeV/u, providing a new method to constrain the shallow heating mechanism in the outer layers of accreting neutron stars. Features of the light curve rise are used to demonstrate that a lower-limit could likely be placed on the 15 O(α, γ ) reaction rate, demonstrating the possibility of constraining nuclear reaction rates with X-ray burst light curves. 1. INTRODUCTION Type-I X-ray bursts, periodic thermonuclear explo- sions driven and fueled by hydrogen and/or helium rich mixtures siphoned from a binary companion, provide unique insight into the structure and dense matter of the neutron stars that host them (Lamb & Lamb 1978; Joss 1978; Schatz et al. 1998; Parikh et al. 2013; Ar- cones et al. 2017). X-ray burst models require detailed input regarding the compositional and thermal structure of the neutron star envelope, as well as well over a thou- sand reaction rates involving more than three hundred nuclides (Wallace & Woosley 1981; Woosley et al. 2004; Schatz et al. 2001; Fisker et al. 2008; Jos´ e et al. 2010). Many important calculation inputs, such as the accre- tion rate ˙ M and nuclear reactions rates of the rapid proton-capture (rp)-process powering bursts, have dis- tinctive influences on the calculation results (Woosley et al. 2004; Parikh et al. 2008, 2009; Lampe et al. 2016; Cyburt et al. 2016; Schatz & Ong 2017). This enables model-observation comparisons to determine unique so- lutions, resulting in astrophysical constraints on an X-ray bursting object (Heger et al. 2007; Galloway et al. 2017; Johnston et al. 2018). The consistency of the X-ray burster GS 1826-24 (Gal- loway et al. 2004, 2008) and its “textbook” behav- ior (Bildsten 2000) have made it the primary target of past model-observation comparisons (Heger et al. 2007; Galloway et al. 2017; Zamfir et al. 2012). To date, all of this pioneering work has been performed using the multizone astrophysical modeling code KEPLER (Weaver et al. 1978; Woosley et al. 2004), aside from initial proof- of-principle calculations performed with the open-source multizone stellar evolution code MESA (Paxton et al. 2015) and a simple ignition model use to predict burst recurrence time (Galloway et al. 2004). These KEPLER model-observation comparisons constrained the astro- [email protected] physical conditions for GS 1826-24 by reproducing the recurrence time between bursts Δt rec for several ˙ M and the average burst light curve for a single ˙ M . However, simultaneous light curve comparisons for a consistently modeled range of ˙ M that approximates the observed ˙ M variation have not yet been performed. The peril in this approach is that the light curve shape from mod- els is known to vary over the range of ˙ M similar to that inferred from observations of hydrogen/helium-burning Type-I X-ray bursts (Lampe et al. 2016). Furthermore, the sensitivity of models to varied nu- clear reaction rates has not yet been accounted for in model-observation comparisons, though some rates are known to substantially impact model calculations. In particular, the reaction rate 15 O(α, γ ) 19 Ne has been shown to alter Δt rec beyond observational uncertainties and to modify the light curve shape much more than the natural variations observed for GS 1826-24 over its reg- ular bursting epochs (Fisker et al. 2007; Cyburt et al. 2016). Here, the first consistent comparison to X-ray burst light curves for a bursting source over a range of ˙ M is used to demonstrate that multi-epoch reproduction is required to remove degeneracies in astrophysics model parameters and achieve tighter astrophysical constraints than previously possible. These calculations, the first to model hydrogen/helium-burning bursts for a large range of input conditions with MESA, demonstrate that model- observation comparisons can place tight constraints on the strength of shallow heating Q b in the accreted neu- tron star outer layers and that GS 1826-24 has a higher ˙ M than previously inferred from models. Additionally, the possibility to constrain nuclear reaction rates in the rp-process with X-ray burst light curves is demonstrated by showing that a lower-limit could likely be placed on the 15 O(α, γ ) reaction rate. A follow-up paper will com- pare results from the full grid of model calculations used for this work to results from a similar grid of calculations arXiv:1805.05552v1 [astro-ph.HE] 15 May 2018
Transcript
Page 1: Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as compared to using upper and lower rate limits based on experimental uncertainties.

Draft version May 16, 2018Preprint typeset using LATEX style emulateapj v. 12/16/11

CONSISTENT MODELING OF GS 1826-24 X-RAY BURSTS FOR MULTIPLE ACCRETION RATESDEMONSTRATES THE POSSIBILITY TO CONSTRAIN RP -PROCESS REACTION RATES

Zach MeiselInstitute of Nuclear & Particle Physics, Department of Physics & Astronomy, Ohio University, Athens, Ohio 45701, USA

Draft version May 16, 2018

ABSTRACT

Type-I X-ray burst light curves encode unique information about the structure of accreting neutronstars and the nuclear reaction rates of the rp-process that powers bursts. Using the first model cal-culations of hydrogen/helium burning bursts for a large range of astrophysical conditions performedwith the code MESA, this work shows that simultaneous model-observation comparisons for bursts fromseveral accretion rates M are required to remove degeneracies in astrophysical conditions that other-wise reproduce bursts for a single M and that such consistent multi-epoch modeling could possiblylimit the 15O(α, γ)19Ne reaction rate. Comparisons to the year 1998, 2000, and 2007 bursting epochs

of the neutron star GS 1826-24 show that M must be larger than previously inferred and that theshallow heating in this source must be below 0.5 MeV/u, providing a new method to constrain theshallow heating mechanism in the outer layers of accreting neutron stars. Features of the light curverise are used to demonstrate that a lower-limit could likely be placed on the 15O(α, γ) reaction rate,demonstrating the possibility of constraining nuclear reaction rates with X-ray burst light curves.

1. INTRODUCTION

Type-I X-ray bursts, periodic thermonuclear explo-sions driven and fueled by hydrogen and/or helium richmixtures siphoned from a binary companion, provideunique insight into the structure and dense matter ofthe neutron stars that host them (Lamb & Lamb 1978;Joss 1978; Schatz et al. 1998; Parikh et al. 2013; Ar-cones et al. 2017). X-ray burst models require detailedinput regarding the compositional and thermal structureof the neutron star envelope, as well as well over a thou-sand reaction rates involving more than three hundrednuclides (Wallace & Woosley 1981; Woosley et al. 2004;Schatz et al. 2001; Fisker et al. 2008; Jose et al. 2010).Many important calculation inputs, such as the accre-tion rate M and nuclear reactions rates of the rapidproton-capture (rp)-process powering bursts, have dis-tinctive influences on the calculation results (Woosleyet al. 2004; Parikh et al. 2008, 2009; Lampe et al. 2016;Cyburt et al. 2016; Schatz & Ong 2017). This enablesmodel-observation comparisons to determine unique so-lutions, resulting in astrophysical constraints on an X-raybursting object (Heger et al. 2007; Galloway et al. 2017;Johnston et al. 2018).

The consistency of the X-ray burster GS 1826-24 (Gal-loway et al. 2004, 2008) and its “textbook” behav-ior (Bildsten 2000) have made it the primary target ofpast model-observation comparisons (Heger et al. 2007;Galloway et al. 2017; Zamfir et al. 2012). To date, allof this pioneering work has been performed using themultizone astrophysical modeling code KEPLER (Weaveret al. 1978; Woosley et al. 2004), aside from initial proof-of-principle calculations performed with the open-sourcemultizone stellar evolution code MESA (Paxton et al.2015) and a simple ignition model use to predict burstrecurrence time (Galloway et al. 2004). These KEPLERmodel-observation comparisons constrained the astro-

[email protected]

physical conditions for GS 1826-24 by reproducing therecurrence time between bursts ∆trec for several M andthe average burst light curve for a single M . However,simultaneous light curve comparisons for a consistentlymodeled range of M that approximates the observed Mvariation have not yet been performed. The peril inthis approach is that the light curve shape from mod-els is known to vary over the range of M similar to thatinferred from observations of hydrogen/helium-burningType-I X-ray bursts (Lampe et al. 2016).

Furthermore, the sensitivity of models to varied nu-clear reaction rates has not yet been accounted for inmodel-observation comparisons, though some rates areknown to substantially impact model calculations. Inparticular, the reaction rate 15O(α, γ)19Ne has beenshown to alter ∆trec beyond observational uncertaintiesand to modify the light curve shape much more than thenatural variations observed for GS 1826-24 over its reg-ular bursting epochs (Fisker et al. 2007; Cyburt et al.2016).

Here, the first consistent comparison to X-ray burstlight curves for a bursting source over a range of M isused to demonstrate that multi-epoch reproduction isrequired to remove degeneracies in astrophysics modelparameters and achieve tighter astrophysical constraintsthan previously possible. These calculations, the first tomodel hydrogen/helium-burning bursts for a large rangeof input conditions with MESA, demonstrate that model-observation comparisons can place tight constraints onthe strength of shallow heating Qb in the accreted neu-tron star outer layers and that GS 1826-24 has a higherM than previously inferred from models. Additionally,the possibility to constrain nuclear reaction rates in therp-process with X-ray burst light curves is demonstratedby showing that a lower-limit could likely be placed onthe 15O(α, γ) reaction rate. A follow-up paper will com-pare results from the full grid of model calculations usedfor this work to results from a similar grid of calculations

arX

iv:1

805.

0555

2v1

[as

tro-

ph.H

E]

15

May

201

8

Page 2: Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as compared to using upper and lower rate limits based on experimental uncertainties.

2

performed with KEPLER (Lampe et al. 2016). A previouspaper featured MESA X-ray burst ash abundances (Meisel& Deibel 2017).

2. MODEL CALCULATIONS

Type-I X-ray burst model calculations were performedwith the one-dimensional stellar evolution code MESA ver-sion 9793. The numerical approach and physics mod-els adopted in MESA are detailed in the associated in-strumentation papers (Paxton et al. 2011, 2013, 2015,2018). Here the most pertinent details for this workare summarized. The neutron star envelope is 0.01 km-thick, with an inner boundary of neutron star massMNS = 1.4 M� and radius RNS = 11.2 km, comprisedinitially of 70% hydrogen, 28% helium, and 2% met-als, by mass, using the solar metallicity Z of Grevesse& Sauval (1998). The envelope is discretized into ∼1000 zones, which adapts during the calculation, wherethe local gravity in a zone is corrected for general rel-ativity effects using a post-Newtonian correction. Con-vection is approximated using the mixing length theoryof Henyey et al. (1965) and is time-dependent (Pax-ton et al. 2011). Accretion is achieved by addinga small amount of mass to the model’s outer layersand re-adjusting the stellar structure (Paxton et al.2011). The spatial and time resolution are adaptive,where the MESA settings varcontrol target=1d-3 andmesh delta coeff=1.0 (Paxton et al. 2013) were cho-sen after tests for convergence of the light curve shapeand ∆trec in which varcontrol target from 10−4−10−2

mesh delta coeff from 0.5−2.0 were investigated. Hereconvergence means that the mean light curve and re-currence time changed << 1σ for finer spatial and/ortime resolution settings. The nuclear reaction networkincludes the 304 isotopes of Fisker et al. (2008) usingreaction rates from the REACLIB V2.2 library (Cyburtet al. 2010).

Sequences of X-ray bursts were simulated for 84 dif-ferent sets of initial conditions, where models differedin M , Qb, Z, hydrogen mass fraction X, and a re-duction factor for the 15O(α, γ) reaction rate R. Thenumber of bursts N belonging to a sequence varies,mostly due to numerical and practical challenges. Ex-ample burst sequences are shown in Fig. 1. Simula-tions were performed in sets of three for M , with alow, medium, and high multiple of the Eddington ac-cretion rate ME = 1.75 × 10−8M�/yr (Schatz et al.

1999). A low set, M = 0.05, 0.07, 0.08ME, matched

observed M for the year 1998, 2000, and 2007 burst-ing epochs of GS 1826-24 (Galloway et al. 2008). A

high set, M = 0.11, 0.15, 0.17ME, employed M used inthe proof-of-principle X-ray burst calculations of Pax-ton et al. (2015) for the highest M and then reduced

this by the ratio of M for the observation epochs. Forexample, 0.17 ME × (0.05)/(0.08) ≈ 0.11 ME. Qb =0.1, 0.5, 1.0 MeV/u were used to mimic the shallow heat-ing of unknown origin that is thought to operate in theouter layers of accreting neutron stars (Brown & Cum-ming 2009; Keek & Heger 2017), where the lower limit ison the order expected from accretion-induced reactionsin the accreted neutron star crust (Gupta et al. 2007;Meisel et al. 2016) and the upper limit is on the order of

Time [hr]0 10 20 30 40

]-1

erg

sec

38 [1

0BL 0

1

2

3 0.05,0.1,0.01,1

Time [hr]0 10 20 30 40

]-1

erg

sec

38 [1

0BL

0

1

2

30.11,0.1,0.01,1

Time [hr]0 10 20 30 40

]-1

erg

sec

38 [1

0BL

0

1

2

30.11,0.5,0.01,1

Time [hr]0 10 20 30 40]

-1 e

rg s

ec38

[10

BL0

1

2

30.11,1.0,0.01,1

0 10 Time [hr]20 30 40

]-1

erg

sec

38 [1

0BL 0

1

2

30.11,0.1,0.01,10

Time [sec]0 20 40 60 80 100

]-1

erg

sec

38 [1

0BL 0

1

2

3

Fig. 1.— Bolometric luminosity over time (not redshifted) for

example MESA calculations. The legends indicate, in order, M , Qb,Z, and R. X = 0.7 for each of these models. The bottom panelshows the average light curve, excluding the first burst, for allbursts in a sequence for a set of conditions, where the color of theband matches the color of the corresponding burst sequence above.

typical shallow heating inferred from observations of neu-tron star cooling after accretion turnoff (Brown & Cum-ming 2009; Turlione et al. 2015). In MESA this is achievedby fixing the luminosity of the base of the envelope, sothat the base luminosity depended on Qb and M of themodel. Z = 0.01, 0.02 were used to investigate the solarZ favored by previous investigations of GS 1826-24 (Gal-loway et al. 2004; Heger et al. 2007) and a slight reduc-tion from that value. X = 0.7 and helium mass fractionY = 0.28 were used for most simulations, while X = 0.75and Y = 0.23 were employed for a set of simulationswith Qb = 0.1 and 1.0 MeV/u for each M . R = 1, 5, 10were used as R = 10 is roughly the experimental lower-limit for the 15O(α, γ) rate uncertainty (Tan et al. 2007;Davids et al. 2011) and rate increases have not beenfound to impact the burst light curve (Cyburt et al.

Page 3: Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as compared to using upper and lower rate limits based on experimental uncertainties.

3

2016). Strictly speaking, using a rate scaling factor isa simplification as compared to using upper and lowerrate limits based on experimental uncertainties. How-ever, the present rate uncertainty is generally a factor of10 or larger in the temperature range of interest (Davidset al. 2011), namely from the onset of hot CNO cycleburning to break-out, ∼ 0.1 − 0.5 GK (Wiescher et al.1999). Furthermore, a rate scaling factor enables a moredirect comparison to similar calculations performed inthe past with the codes KEPLER and AGILE (Cyburt et al.2016; Fisker et al. 2007).

3. LIGHT CURVE CONSTRUCTION

In order to compare to observational data, bursts ina simulated sequence needed to be stacked (as is donewith observed light curves) so that an average light curveand uncertainty band could be calculated. The firstburst in a sequence was excluded, as the first simulatedburst is typically far more energetic than subsequentbursts (Woosley et al. 2004). The burst start time t = 0was defined as the point when the luminosity crossed athreshold indicating thermonuclear runaway. Individualbursts were mapped onto the same time grid with a linearspline and, to mitigate numerical noise, each burst lightcurve was smoothed by averaging over the luminosity fora ±1 s time window. This smoothing is frequently donefor multi-zone numerical calculations of X-ray burst lightcurves, e.g. as described for the KEPLER models of Cy-burt et al. (2016)1. Using luminosity data from all bursts(after the first) in a sequence, an average luminosity andupper and lower 1σ uncertainties were computed for eachtime point. An example of this process is shown in Fig. 2.

4. OBSERVATIONAL DATA

The observed light curve data for GS 1826-24 are cour-tesy of the Multi-Instrument Burst Archive (MINBAR)2.The data analysis is described in Galloway et al. (2017)and is briefly rehashed here. Observational data fromthe Rossi X-ray Timing Explorer (Galloway et al. 2004,2008) were stacked and averaged in a similar manneras described above for the simulated light curves of thiswork. Since uninterrupted observation was not possibledue to periodic occultation by the Earth, observed recur-rence times were determined using an iterative approach.Each burst was assigned a trial integer indicating whichburst it was in the sequence and a fit was performed toquantify how well the set of assignments matched thedata if one assumes a regular recurrence time. The ob-served M were determined using a distance of 6.1 kpc, abolometric correction cbol between ≈ 1.75− 1.8, and theaverage persistent flux Fp over the burst sequence, where

M = 4πd2Fpcbol/ME. These observed M are potentially

systematically shifted by some factor from the true Mbased on the fact that no burst anisotropy ξ (Fujimoto1988; He & Keek 2016) is assumed. Anisotropy accountsfor the fact that X-ray flux can be beamed toward or

1 An example of when this has not been done are the light curvesof Jose et al. (2010), where (inconsequential) sharp, unphysicalfeatures are seen in some calculation results.

2 https://burst.sci.monash.edu/minbar/

Time [sec]0 50 100 150

]-1

sec

-2 e

rg c

m-9

Flu

x [1

0

0

20

40

60

(a) Original.

Time [sec]0 50 100 150

]-1

sec

-2 e

rg c

m-9

Flu

x [1

00

20

40

60

(b) Smoothed.

Time [sec]0 50 100 150

]-1

sec

-2 e

rg c

m-9

Flu

x [1

0

0

20

40

60

(c) Averaged.

Fig. 2.— Light curve processing steps for the 20 bursts of the0.111 ME, X = 0.7, Z = 0.02, Qb = 0.1 MeV/u, R = 1 simulationfor an arbitrary distance and redshift. The first burst (black linein (a)) is not included in the averaged light curve and uncertaintyband in (c).

away from the observer, where the effect depends on theaccretion disk geometry and source inclination angle andcan be different for the burst flux and the persistent flux.This is described in more detail in the discussion.

5. MODEL-OBSERVATION COMPARISONS

Comparison to observations required adjusting thesimulated light curve for the distance and surface gravita-tional redshift (1+z). The burst anisotropy ξb is included

in the distance, so that distance is dξ1/2b . The luminosity

L to flux F conversion is F = L/(4πξb(1 + z)d2) (Gal-loway et al. 2017). Time is redshifted by multiplying thesimulation time by (1 + z). In principle, the neutron

Page 4: Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as compared to using upper and lower rate limits based on experimental uncertainties.

4

Distance [kpc]4 5 6 7 8

Red

shift

(1+

z)

1.2

1.3

1.4

1

2

3

4

5

6

7

8

9

10

Fig. 3.— χ2red, indicated by the color, for a distance dξ

1/2b and

redshift, using the optimum δt, for the MESA calculation that is thebest-fit to the year 2007 burst epoch for GS 1826-24. χ2

red>10 areincluded in the upper bin.

star mass and radius adopted for the simulations corre-spond to (1 + z) = 1.26, so choosing other redshifts isinconsistent. However, in practice burst properties areinsensitive to modest changes in MNS and RNS (Ayasli& Joss 1982; Zamfir et al. 2012).

Model-observation comparisons were performed by cal-culating F (t) for the averaged light curve for the highest

M in a set of three and comparing to the year 2007 burst

epoch of GS 1826-24. For each point in a grid of dξ1/2b ,

(1 + z), and time-shift δt, χ2red was calculated using data

from t = 0 − 50 s. δt is necessary, so that neither the

burst rise nor tail dominates χ2red. dξ

1/2b varied in steps

of 0.2 kpc from 4 to 8 kpc, based on observational lim-its (Galloway et al. 2004). (1 + z) varied in steps of 0.02from 1.18 to 1.44, in order to roughly stay within therange RNS ∼ 8 − 15 km determined by Steiner et al.(2010) for MNS = 1.4M�. δt varied from 0.5 to 1.5 s insteps of 0.1 s, as the best-fit was located in this rangefor each of the 84 models. The results for the best-fitout of all models (which also roughly reproduced the ob-served ∆trec, see Fig. 5), is shown in Fig. 3. The tight

constraints on dξ1/2b are due to its strong impact on the

peak F , where as (1+z) is poorly constrained due to thecompetition between fitting the burst rise and burst tail(and is sensitive to the range over which time is fit) (Zam-fir et al. 2012).

To move beyond previous studies, a consistent com-parison to the year 1998, 2000, and 2007 burst epochswas performed by calculating F (t) for each model us-

ing dξ1/2b , (1 + z), and δt obtained for the best fit to

the year 2007 outburst for the highest M model in a setof three. I.e. F (t) for M = 0.05 and 0.07 ME models

were calculated using the best-fit dξ1/2b , (1 + z), and δt

found for M = 0.08 ME models, whereas M = 0.11 and0.15 ME F (t) were calculated based on M = 0.17 ME

models, for the same Qb, Z, X, and R. The justificationfor comparing simulations with M higher than observedvalues to observed light curves is that ξb can differ fromthe persistent anisotropy ξb in between bursts, so that

the true M could be different than inferred from obser-vations (Fujimoto 1988; Galloway et al. 2017). Examplemodel-observation comparisons are shown in Fig. 4, in-cluding a comparison to the best-fit found with KEPLER,as reported by Galloway et al. (2017), using light curves

calculated by Lampe et al. (2016). The lowest M KEPLERmodel shown is not as low as would be required to matchthe observed M ratio, but is the closest available. ∆trecis the average time between thermonuclear runaways us-ing the best-fit (1 + z) for the highest M in the set ofthree. Comparisons to the observed ∆trec, normalizingM so that the highest M of a set of three (i.e. 0.08

or 0.17 ME) matches M for the year 2007 epoch of GS1826-24, are shown in Fig. 5.

6. DISCUSSION

6.1. Model Parameter Impacts

Prior to discussing the results from model-observationcomparisons, the impact of model parameters on the X-ray burst light curve and recurrence time are briefly dis-cussed.

Increased M decreases ∆trec. This is because the shal-low heating scales with the accretion rate, increased heat-ing speeds up the CNO cycle, which results in an earlierarrival at the temperature and He-abundance requiredto trigger the 3α reaction for burst ignition. Lower Mtherefore requires more H to be burned prior to burstignition, resulting in a smaller H/He ratio at burst ig-nition, and therefore a relatively He-rich burst. He-richbursts burn fuel more rapidly, with higher peak lumi-nosities and shorter tail decay times as compared to lessHe-rich conditions (Weinberg et al. 2006).

Increased Qb decreases ∆trec for a given M due to theinfluence of shallow heating on burst recurrence discussedabove for M . Similarly, for a given M , increased Qb

preserves H prior to burst ignition, extending the bursttail, which is powered by H-burning. When considering arange of M , increased Qb increases the curvature in thetrend for the M -∆trec relationship. This is because theshallow heating in the model results from the product ofM and Qb, and therefore increasing both has a nonlinearinfluence on ∆trec.

Increased Z corresponds to increased CNO abun-dances, increasing the amount of H-burning prior toburst ignition and leaving less H to burn during the burst.Z was varied over a relatively small range here, so theonly obvious impact is a slightly extended burst tail forZ = 0.01 relative to Z = 0.02. Increased X, at the ex-pense of Y , naturally results in a reduced He abundanceat burst ignition, and therefore a decreased peak lumi-nosity, and more H left to burn during the burst, andtherefore an extended burst tail.

The influence of R on the burst properties derives fromthe nature of 15O(α, γ) as a “valve” controlling the flowof material out of the hot CNO cycle during interburstburning (Fisker et al. 2006; Cyburt et al. 2016). Decreas-ing the 15O(α, γ) reaction rate (increasing R) reducesthe amount of material escaping the hot CNO cycle dur-ing quiescent burning, enabling more He to be producedprior to burst ignition, shortening ∆trec and resulting ina more He-rich burst.

Page 5: Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as compared to using upper and lower rate limits based on experimental uncertainties.

5

Time [sec]20 40 60 80

]-1

sec

-2 e

rg c

m-9

Flu

x [1

0

10

20

300.11,0.1,0.02,1 (19)0.15,0.1,0.02,1 (20)0.17,0.1,0.02,1 (20)

Time [sec]20 40 60 80

]-1

sec

-2 e

rg c

m-9

Flu

x [1

0

10

20

300.11,0.1,0.01,1 (18)0.15,0.1,0.01,1 (19)0.17,0.1,0.01,1 (2)

Time [sec]20 40 60 80

]-1

sec

-2 e

rg c

m-9

Flu

x [1

0

10

20

30

0.11,0.1,0.02,10 (6)0.15,0.1,0.02,10 (18)0.17,0.1,0.02,10 (18)

Time [sec]20 40 60 80

]-1

sec

-2 e

rg c

m-9

Flu

x [1

0

10

20

300.11,0.5,0.01,1 (22)0.15,0.5,0.01,1 (3)0.17,0.5,0.01,1 (8)

Time [sec]20 40 60 80

]-1

sec

-2 e

rg c

m-9

Flu

x [1

0

10

20

30

0.05,0.5,0.01,1 (10)0.07,0.5,0.01,1 (18)0.08,0.5,0.01,1 (17)

Time [sec]20 40 60 80

]-1

sec

-2 e

rg c

m-9

Flu

x [1

0

10

20

30Ka019: 0.07,0.2,0.02,1 (19)Ka028: 0.08,0.2,0.02,1 (18)Ka05d: 0.09,0.2,0.02,1 (30)

Fig. 4.— Comparison of light curves calculated with MESA (red bands) to GS 1826-24 light curves observed for the years 1998 (brown),2000 (green) , and 2007 (purple). Legends indicate the same information as the legend in Fig. 1, where the additional number in parenthesesis N . The red bands in the lower-right panel are light curves calculated with KEPLER from Lampe et al. (2016) using the optimum distanceand redshift determined in Galloway et al. (2017).

6.2. M and Qb Constraints

Fig. 4 demonstrates that light curve shape for theyear 2007 burst epoch of GS 1826-24 can be accommo-dated for several M and Qb, meaning reproduction of alight curve for a single observed M is insufficient to con-strain an X-ray bursting source’s conditions with model-observation comparisons. While M = 0.08 ME repro-duces the year 2007 epoch, lower M in the set result ina larger peak flux and shorter burst tail decay than seenin observations, particularly for the lowest M . The fig-ure shows that this result is not only seen with MESA, butalso for KEPLER models. Therefore, the GS 1826-24 M forobserved bursting epochs must be larger than previously

inferred.Fig. 5 demonstrates that ∆trec provides an additional

necessary discriminant, as models reproducing the lightcurve shape for all three observed epochs do not nec-essarily reproduce the observed ∆trec. Though Qb =0.5 MeV/u can accommodate the light curve shape forall bursting epochs, ∆trec is significantly shorter than forobservations. Therefore, shallow heating in GS 1826-24is limited to ≤ 0.5 MeV/u, providing an example howmulti-epoch X-ray burst modeling can be used to con-strain the shallow heating mechanism in accreting neu-tron star outer layers.

It is evident that X = 0.75 cannot be accommodatedeither, since there is a significant curvature in ∆trec for

Page 6: Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as compared to using upper and lower rate limits based on experimental uncertainties.

6

[Edd.]norm.M0.06 0.08 0.1

t [hr

]∆

0

5

10

150.17,0.1,0.01,1

0.17,0.1,0.02,1

0.17,0.1,0.02,10

0.17,0.5,0.01,1

0.08,0.1,0.01,1

0.08,0.1,0.02,10

0.08,0.5,0.01,1

0.08,1.0,0.02,1,0.75

K:a05d,a028,a019

GS 1826-24

Fig. 5.— Relationship between ∆trec and normalized accretionrate. Connected symbols indicate bursts belonging to a triplet ofsimulated accretion rates. The legend indicates the same infor-mation as in Fig. 1, where the extra column for the eighth modelindicates X = 0.75. The KEPLER simulations from Fig. 4 and ob-served properties for GS 1826-24 are shown for comparison.

decreasing M which is not seen in the observed data.One sees a similar behavior in KEPLER models, which canbe seen by comparing models a003 and a020 of Lampeet al. (2016). X less than 0.7 were not explored here asthis would move toward the conditions for helium bursts,as most or all of the hydrogen would be burned stablybefore burst ignition.

The best-fit MESA model for light-curve shape and ∆trechas M = 0.17 ME (for the year 2007 epoch), Qb =0.1 MeV/u, R = 1, X = 0.70, and Z = 0.02, thoughthe same conditions with Z = 0.01 perform nearly aswell.

6.3. Comparison to KEPLER

Fig. 5 also highlights a discrepancy between MESA andKEPLER models. While KEPLER reproduces the year 2007epoch ∆trec with M = 0.09 ME, MESA models requireM = 0.17 ME, as noted by Paxton et al. (2015). Thiscannot be explained by the slightly higher Qb employedin the best-fit for KEPLER (Galloway et al. 2017), as the

MESA model with M = 0.08 ME and Qb 0.5 MeV/u re-sults in ∆trec roughly 2/3 larger than observed for theyear 2007 epoch. Systematic comparisons between MESAand KEPLER, which are beyond the scope of this work,are necessary to resolve this discrepancy.

Nonetheless, the constraints on dξ1/2b for past KEPLER

fits and for this work are in agreement, where the bestfit here (see Fig. 3) favors 6 kpc and the most recentKEPLER results (Galloway et al. 2017) favor 6.1 kpc. Thiswork favors a much larger redshift than Galloway et al.(2017), (1+z) = 1.42 as compared to 1.23, however Fig. 3demonstrates that (1+z) down to ∼ 1.28 performs nearlyas well. As in Galloway et al. (2017), uncertainties arenot quoted here due to the large number of systematicswhich will require several further studies to quantify. Itshould be noted that (1+z) = 1.42 corresponds to RNS =8.2 km for the canonical MNS = 1.4 M�, which is smallerthan expectations (Steiner et al. 2013), though RNS =11.7 km for MNS = 2.0 M� (Lampe et al. 2016).

Time After Ignition [sec]10 20 30 40

]-1

sec

-2 e

rg c

m-9

F [1

4−

2−

0

2

4

Fig. 6.— δF = FMESA − FGS2007 over the early light curve,where the first ∼10 s are the light curve rise, for models withM = 0.17 ME, Qb = 0.1 MeV/u, Z = 0.02, X = 0.7, and, fromlight to dark bands, R = 1, 5, 10.

6.4. Anisotropies

The burst anisotropy ξb and persistent anisotropy ξpbetween bursts can differ substantially due to the burstinfluence on accretion disk geometry. The ratio ξp/ξb isdetermined by the inclination angle relative to the ob-server and accretion disk geometry. This ratio can beinferred from simulation results via ξp/ξb = Mc2(z/(1 +z))/(4πd2Fpcbol), where c is the speed of light (Heger

et al. 2007). Using this work’s best-fit M , z, and d and Fp

and cbol from Galloway et al. (2017), ξp/ξb = 3.5. Thiscould be explained (see Fig. 12 of He & Keek (2016)) bya flat accretion disk for a system with a relatively highinclination angle θ ≈ 80◦. For the same conditions, theKEPLER best-fit requires θ ≈ 65◦, whereas roughly thesame θ explains the best-fit from both codes if a curvedaccretion disk is assumed.

6.5. Possibility to Constrain the 15O(α, γ)19NeReaction Rate

It is apparent from Fig. 5 that R has a relatively mod-est impact on ∆trec, in agreement with prior observa-tions (Cyburt et al. 2016). However, as shown in Figs. 4and 6, R significantly increases the departure from lin-earity in the light curve rise, known as the convexityC (Maurer & Watts 2008) (where C = 0 is linear). Theyear 1998, 2000, and 2007 epochs of GS 1826-24 exhibita low C, whereas MESA models with R > 1 show an in-crease in C due to a shoulder introduced in the light curverise. A similar shoulder is present in KEPLER models forR = 10 (Cyburt et al. 2016).

It is possible that this signature in the light curvecould be erased by convolving the one-dimensional re-sults presented here with a more sophisticated treatmentfor flame-spreading on the neutron star surface, whichalso impacts the light curve rise. For instance, Maurer& Watts (2008) found using a phenomenological modelthat the longitudinal dependence of the flame speed canresult in C > 0 or < 0 depending on the ignition lati-tude. Since C ∼ 0 for the GS 1826-24 1998, 2000, and2007 burst epochs, the R = 5, 10 models presented herecould potentially describe the observational data if con-volved with near-polar burst ignition. Alternatively, rel-atively slow flame-spreading from an equatorial ignitioncould smear-out any intrinsic bump-like artifacts in the

Page 7: Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as compared to using upper and lower rate limits based on experimental uncertainties.

7

light curve rise; however, Zamfir (2010) found that thiswould require a flame that takes ∼ 7 s to encompassthe neutron star surface, which is several times longerthan inferred from oscillations in the burst light curverise (Chakraborty & Bhattacharyya 2014).

Taking the current results at face value suggests thatthe 15O(α, γ) reaction rate cannot be more than 5×lower than the currently accepted rate of Davids et al.(2011). This limit is more stringent than the constraintderived from nuclear physics experiments, for which the3σ uncertainty sets a lower limit & ×10 (Davids et al.2011). This limit for the 15O(α, γ)19Ne reaction rate,> 5× lower than Davids et al. (2011), implies that the α-particle decay branching ratio for the key 4.03 MeV reso-nance in 19Ne is likely within reach of a newly developedexperimental probe using radioactive ion beams (Wredeet al. 2017). Nonetheless, it should be stressed thatmore reliable constraints will require systematic inves-tigations, beyond the scope of this work, which employvarious treatments of effects impacting the light curverise that are not included here, especially flame spread-ing on the neutron star surface. A large number of calcu-lations are underway which will examine the MESA X-rayburst model sensitivity to other nuclear reaction rates,similar to the study of Cyburt et al. (2016).

7. CONCLUSIONS

In summary, a large number of X-ray burst model cal-culations performed with the code MESA have been usedto reproduce the year 1998, 2000, and 2007 burstingepochs from GS 1826-24. It has been shown that Mfor these bursting epochs must be larger than previouslyinferred. This work also shows that model-observation

comparisons for X-ray burst light curves and ∆trec per-formed consistently for several M are necessary to re-move model degeneracies. Consistent comparisons canbe used to constrain Qb for a bursting source and canpossibly set a lower limit on the 15O(α, γ) reaction rate.The Qb < 0.5 MeV/u limit for GS 1826-24 provides avaluable constraint that can be used to investigate theorigins of the poorly understood shallow heating mecha-nism in accreting neutron stars. Furthermore, using thecase of 15O(α, γ), this work shows that it is possible forX-ray burst model-observation comparisons to constrainreaction rates of the rp-process, though the constraintsdetermined here are contingent upon astrophysical ef-fects such as flame spreading on the neutron star sur-face. It is likely that constraining other nuclear reactionrates will require comparisons to more observable prop-erties and many more model calculations due to theirmore subtle impacts on the X-ray burst light curve (Cy-burt et al. 2016). Logically, such investigations may beextended to nuclear masses as well (Schatz & Ong 2017).

I thank Duncan Galloway, Zac Johnston, and Hen-drik Schatz for useful discussions, and acknowledge theinstructive tutorials from Ed Brown, Andrew Cum-ming, and Rob Farmer located on the MESA Mar-ketplace (http://cococubed.asu.edu/mesa market/)which helped at the beginning stages of this work. Rele-vant MESA inputs for this work are available on the MESAMarketplace. This work was supported in part by theU.S. Department of Energy under grant № DE-FG02-88ER40387. This material is based on work supportedby the National Science Foundation under grant № PHY-1430152 (Joint Institute for Nuclear Astrophysics–Centerfor the Evolution of the Elements).

REFERENCES

Arcones, A., et al. 2017, Prog. Part. Nucl. Phys., 94, 1Ayasli, S., & Joss, P. C. 1982, Astrophys. J., 256, 637Bildsten, L. 2000, in Proc. 10th Astrophys. Conf., AIP Conf. Ser.,

Vol. 522, 359Brown, E., & Cumming, A. 2009, Astrophys. J., 698, 1020Chakraborty, M., & Bhattacharyya, S. 2014, Astrophys. J., 792, 4Cyburt, R. H., Amthor, A. M., Heger, A., et al. 2016, Astrophys.

J., 830, 55Cyburt, R. H., et al. 2010, Astrophys. J. Suppl. Ser., 189, 240,

https://groups.nscl.msu.edu/jina/reaclib/dbDavids, B., Cyburt, R. H., Jose, J., & Mythili, S. 2011,

Astrophys. J., 735, 40Fisker, J. L., Gorres, J., Wiescher, M., & Davids, B. 2006,

Astrophys. J., 650, 332Fisker, J. L., Schatz, H., & Thielemann, F.-K. 2008, Astrophys. J.

Suppl. Ser., 174, 261Fisker, J. L., Tan, W., Gorres, J., & Wiescher, M. 2007,

Astrophys. J., 665, 637Fujimoto, M. Y. 1988, Astrophys. J., 324, 995Galloway, D. K., Cumming, A., Kuulkers, E., et al. 2004,

Astrophys. J., 601, 466Galloway, D. K., Goodwin, A. J., & Keek, L. 2017, Publ. Astron.

Soc. Aust., 34, 19Galloway, D. K., Muno, M. P., Hartman, J. M., Dimitrios, P., &

Chakrabarty, D. 2008, Astrophys. J. Suppl. Ser., 179, 360Grevesse, N., & Sauval, A. J. 1998, Space Sci. Rev., 85, 161Gupta, S., Brown, E. F., Schatz, H., Moller, P., & Kratz, K.-L.

2007, Astrophys. J., 662, 1188He, C.-C., & Keek, L. 2016, Astrophys. J., 819, 47Heger, A., Cumming, A., Galloway, D. K., & Woosley, S. E. 2007,

Astrophys. J. Lett., 671, L141Henyey, L., Vardya, M. S., & Bodenheimer, P. 1965, Astrophys.

J., 142, 841Johnston, Z., Heger, A., & Galloway, D. K. 2018, Monthly

Notices of the Royal Astronomical Society, sty757

Jose, J., Moreno, F., Parikh, A., & Iliadis, C. 2010, Astrophys. J.Suppl. Ser., 189, 204

Joss, P. C. 1978, Astrophys. J. Lett., 225, L123Keek, L., & Heger, A. 2017, Astrophys. J., 842, 113Lamb, D. Q., & Lamb, F. K. 1978, Astrophys. J., 220, 291Lampe, N., Heger, A., & Galloway, D. K. 2016, Astrophys. J.,

819, 46Maurer, I., & Watts, A. L. 2008, Mon. Not. R. Astron. Soc., 383,

387Meisel, Z., & Deibel, A. 2017, Astrophys. J., 837, 13Meisel, Z., George, S., Ahn, S., et al. 2016, Phys. Rev. C, 93,

035805Parikh, A., Jose, J., Iliadis, C., Moreno, F., & Rauscher, T. 2009,

Phys. Rev. C, 79, 045802Parikh, A., Jose, J., Moreno, F., & Iliadis, C. 2008, Astrophys. J.

Suppl. Ser., 178, 110Parikh, A., Jose, J., Sala, G., & Iliadis, C. 2013, Prog. Part. Nucl.

Phys., 69, 225Paxton, B., Bildsten, L., Dotter, A., et al. 2011, Astrophys. J.

Suppl. Ser., 192, 3, www.mesa.sourceforge.netPaxton, B., Cantiello, M., Arras, P., et al. 2013, Astrophys. J.

Suppl. Ser., 208, 4Paxton, B., Marchant, P., Schwab, J., et al. 2015, Astrophys. J.

Suppl. Ser., 220, 15Paxton, B., Schwab, J., Bauer, E. B., et al. 2018, Astrophys. J.

Suppl. Ser., 234, 34Schatz, H., Bildsten, L., Cumming, A., & Wiescher, M. 1999,

Astrophys. J., 524, 1014Schatz, H., & Ong, W.-J. 2017, Astrophys. J., 844, 139Schatz, H., et al. 1998, Phys. Rep., 294Schatz, H., Aprahamian, A., Barnard, V., et al. 2001, Phys. Rev.

Lett., 86, 3471Steiner, A. W., Lattimer, J. M., & Brown, E. F. 2010, Astrophys.

J., 722, 33—. 2013, Astrophys. J. Lett., 765, L5Tan, W. P., Fisker, J. L., Gorres, J., Couder, M., & Wiescher, M.

2007, Phys. Rev. Lett., 98, 242503

Page 8: Zach Meisel - arXiv · 2016). Strictly speaking, using a rate scaling factor is a simpli cation as compared to using upper and lower rate limits based on experimental uncertainties.

8

Turlione, A., Aguilera, D. N., & Pons, J. A. 2015, Astron. &Astrophys., 577, A5

Wallace, R. K., & Woosley, S. E. 1981, Astrophys. J. Suppl. Ser.,45, 389

Weaver, T. A., Zimmerman, G. B., & Woosley, S. E. 1978,Astrophys. J., 225, 1021

Weinberg, N., Bildsten, L., & Schatz, H. 2006, Astrophys. J., 639,1018

Wiescher, M., Gorres, J., & Schatz, H. 1999, J. Phys. G., 25, R133

Woosley, S. E., Heger, A., Cumming, A., et al. 2004, Astrophys.J. Suppl. Ser., 151, 75

Wrede, C., et al. 2017, Phys. Rev. C, 96, 032801(R)Zamfir, M. 2010, Master’s thesis, McGill UniversityZamfir, M., Cumming, A., & Galloway, D. K. 2012, Astrophys. J.,

749, 69


Recommended