XATOM: an integrated toolkit for X-ray and atomic physics · molecule imaging from small molecules...

1
atom fundamental interaction between atoms and XFEL pulses molecule imaging from small molecules to biological macromolecules solid electronic structure in extreme conditions P (i, )= 4 3 α⇥ 2 N i l i +1 X l j =|l i -1| l > 2l i +1 Z 0 P n i l i (r)P εl j (r) r dr 2 Γ A (i,jj )= π N H i N jj 0 2l i +1 l j +l j 0 X L=|l j -l j 0 | 1 X S=0 X l i 0 (2L + 1)(2S + 1)|M LS (j, j , i, i )| 2 Γ F (i, j )= 4 3 α 3 (I i - I j ) 3 N H i N j 4l j +2 · l > 2l i +1 Z 1 0 P n i l i (r)P n j l j (r) r dr 2 d dt P I (t)= all config. X I 0 6=I [Γ I 0 !I P I 0 (t) - Γ I !I 0 P I (t)] f 0 (Q)= Z ρ(r) e iQ·r d 3 r p SO (i; I,I 0 )=1 - Z 1 0 P n i l i (r; I )P n i l i (r; I 0 ) dr 2 ˆ H = ˆ H mol + ˆ H EM + ˆ H int ˆ H EM = X k,λ ! k ˆ a k,λ ˆ a k,λ , ! k = |k|/ˆ H int = Z d 3 x ˆ (x) ˆ A(x) · r i ˆ (x)+ 2 2 Z d 3 x ˆ (x) ˆ A 2 (x) ˆ (x) |I i: initial state, |F i: final state Γ FI =2⇡δ (E F - E I ) hF | ˆ H int |I i + X M hF | ˆ H int |M ihM | ˆ H int |I i E I - E M + i+ ··· 2 |I i c i |Φ N el 0 i, |F i c a ˆ c j ˆ c j 0 |Φ N el 0 i |I i c i |Φ N el 0 i|0i, |F i c i 0 |Φ N el 0 iˆ a k F ,λ F |0i |I i = | N el 0 i|N EM i, |F i = | N el F i|N EM - 1i |I i = | N el 0 i|N EM i |F i = | N el 0 iˆ a k F ,λ F |N EM - 1i ε=0 rs ε=V(rs) outer-ionization inner-ionization Sang-Kil Son 1 and Robin Santra 1,2 1 CFEL, DESY, Germany / 2 Department of Physics, University of Hamburg, Germany Introduction Theoretical and numerical details XATOM: an integrated toolkit for X-ray and atomic physics X-ray free-electron lasers (XFEL) open a new era in science and technology, offering many unique opportunities that have not been conceivable with conventional light sources. Because XFELs produce ultrashort pulses with a very high X-ray photon fluence, materials interacting with XFEL pulses undergo significant radiation damage and possibly become highly ionized. To understand the underlying physics, it is crucial to describe detailed ionization and relaxation dynamics in individual atoms during XFEL pulses. Here we present an integrated toolkit to investigate X-ray-induced atomic processes and to simulate electronic damage dynamics. This XATOM toolkit can handle all possible electronic configurations of all atom/ion species, and calculate physical observables during/ after intense X-ray pulses. By use of XATOM, we can explore many exciting XFEL-related phenomena from multiphoton multiple ionization to molecular imaging and warm dense matter formation. Hamiltonian and perturbation theory To treat X-ray–atom interactions, we employ a consistent ab initio framework based on nonrelativistic quantum electrodynamics and perturbation theory. For implementation, we use the Hartree-Fock-Slater model with the Latter tail correction. 10 4 10 3 10 2 10 1 10 0 10 1 10 3 10 4 10 5 10 6 10 7 10 8 10 9 Number of scattered photons Fluence (photons/Å 2 ) 1 fs FWHM 10 fs FWHM 100 fs FWHM Number of scattered photons for several pulse durations, showing more scatterings for a shorter pulse due to intensity- induced X-ray transparency or frustrated absorption, SKS, L. Young & R. Santra, PRA 83, 033402 (2011). Center for Free-Electron Laser Science CFEL is a scientific cooperation of the three organizations: DESY – Max Planck Society – University of Hamburg Ionization V P A Ne Ne Ne 100 fs Time 2,000 eV 1,050 eV 800 eV Ne 8+ Ne 8+ Ne 10+ 2 s,p 1 s 2 s,p 1 s 2 s,p 1 s V V V V V V V V V A A A A A A A P P P P P P P P P V Figure from Young et al., Nature 466, 56 (2010). Diagrams of multiphoton absorption mechanisms in Ne induced by ultraintense X-ray pulses Photoabsorption Auger / Coster–Kronig decay Fluorescence Shake-off process: sudden approximation Elastic X-ray scattering Resonant elastic X-ray scattering (dispersion correction) f (Q, )= f 0 (Q)+ f 0 ()+ if 00 () f 0 ()= - 1 22 α P Z 1 0 02 02 - 2 P (0 )d0 f 00 ()= - 4⇥α P () dσ S dΩ (t)= all config. X I P I (t) dσ S dΩ I = dσ dΩ T all config. X I P I (t)|f 0 I (Q)| 2 SO P A F S RS Rate equation model To simulate electronic dynamics during intense X-ray pulses, we employ the rate equation approach with all computed cross sections and rates for all possible n-hole configurations, and calculate charge state distribution, electron/fluorescence spectra, scattering signals, etc. Diagrams of X-ray–atom interaction P A S RS P F SO single-core-hole double-core-hole “hollow atom” scattering from any configurations Ne at LCLS 1110 eV & 1225 eV Ar at SACLA 5 keV Kr at LCLS 1.5 keV & 2 keV Xe at SACLA 5.5 keV Xe at LCLS 1.5 keV & 2 keV Xe at 4.5 keV 10 -1 10 0 10 1 10 2 0.5 1 2 Normalized ratio of Ne 9+ /Ne 8+ Energy (mJ) Experiment Theory + adjusted σ (2) Theory + σ (2) from N-H 1110 eV below threshold: quadratic 1225 eV above threshold: linear The Ne 9+ /Ne 8+ ratio as a function of X-ray pulse energy 10 -3 10 -2 10 -1 10 0 10 1 10 2 10 3 1060 1070 1080 1090 1100 1110 1120 1130 1140 σ (2) (10 -53 cm 4 s) Photon energy (eV) (a) experimental value ∆ω p =20 eV ∆ω p =15 eV ∆ω p =11 eV ∆ω p = 8 eV ∆ω p = 6 eV LOPT ω p (eV) Two-photon ionization cross section of Ne 8+ G. Doumy et al., PRL 106, 083002 (2011). A. Sytcheva et al., PRA 85, 023414 (2012). Ne 1s 2s 2p Xe 2s 2p 1s 3s 3p 3d 4s 4p 4d 5s 5p K L M N O Xe +Xe (n+1)+ + ne Ne +Ne 2+ + e Decay complexity for light and heavy atoms How many coupled rate equations? Ne: 63 Xe: > 70 millions Xe + relativity: > 15 billions Xe + relativity + excitation: infinity 10 -3 10 -2 10 -1 10 0 1 1 . 0 Kr 5+ Kr 9+ Kr 10+ Kr 12+ Kr 14+ Kr 15+ Kr 20+ Ion yield per shot (arb.units) Pulse energy (mJ) Peak fluence (μJ/μm 2 ) 4 5 3 2 1 0 1 1 80 5 fs Kr ion yield as a function of X-ray peak fluence, demonstrating REXMI processes, B. Rudek et al., PRA 87, 023413 (2013). 1 2 3 4 5 6 7 8 9 10 10 -3 10 -2 10 -1 10 0 Charge state Ion yield (counts/shot) Experiment Theory (20 μJ/μm 2 ) Theory (50 μJ/μm 2 ) Theory (80 μJ/μm 2 ) Ar charge state distribution, used for XFEL fluence calibration, K. Motomura et al., JPB 46, 164024 (2013). HHG of rare gases XATOM has been used for calculations of the three-step model of high harmonic generation; Collaboration with CFEL Ultrafast Optics and X-Rays Group, S. Bhardwaj, SKS et al., PRA (in press). 0 100 200 300 400 10 -4 10 -2 10 0 Energy (eV) Absolute Recombination amplitude square Ar recombination amp. Ion, electron, and photon spectra of heavy atoms, SKS & R. Santra, PRA 85, 063415 (2012). charge state distribution Auger electron photoelectron fluorescence -10 -8 -6 -4 -2 0 0 5 10 15 20 25 30 35 40 45 50 Orbital binding energies (keV) Charge state Xe Rydberg O-shell N-shell M-shell L-shell 2.0 keV 1.5 keV 4.5 keV 5.5 keV SACLA & theory LCLS & theory theory 5.0 keV Ionization potentials of Xe as a function of charge states 0 2 4 6 8 10 12 14 0 5 10 15 20 25 Energy (keV) Charge states 1.7 fs 5.3 fs 3.6 fs 1.3 fs 41 fs Photoionization Auger/Coster-Kronig Fluorescence Deep inner-shell multiphoton multiple ionization mechanism H. Fukuzawa, SKS et al., PRL 110, 173005 (2013). Collaboration with CFEL–MPG–ASG, B. Rudek, SKS et al., Nature Photon. 6, 858 (2012). Comparison of experimental and theoretical charge state distributions, revealing new ionization channels via excitation of transient resonances (REXMI) 1.5 keV (experiment) 2 keV (experiment) 1.5 keV (theory) 2 keV (theory) 10 3 10 4 0.01 0.1 1 10 Ion yield (a.u.) 36 34 32 30 28 26 24 22 20 18 16 14 12 10 8 6 4 2 Xe charge state Hollow-atom formation XeF 2 at synchrotron MAD at high x-ray intensity HIP: High-Intensity Phasing 0.0 0.2 0.4 0.6 0.8 1.0 1.2 10 10 10 11 10 12 10 13 10 14 MAD ã Fluence (photons/μm 2 ) H C N O S Collaboration with CFEL Coherent Imaging Group, SKS, L. Galli et al., in preparation. Effective scattering strength of S at 6 keV C 60 at LCLS Comparison of charge state distributions between isolated Xe atom and XeF 2 molecule, R. W. Dunford et al., PRA 86, 033401 (2012). Atomic data calculated by XATOM have been used for molecular dynamics of C 60 ; Collaboration with Z. Jurek, B. F. Murphy et al., in preparation. Screening effect Al plasma at LCLS Two-step HFS model Atomic picture in dense matter: Muffin-tin-like potential with the Wigner-Seitz radius to describe an environment of an atom within a solid or a plasma Photoabsorption cross sections for C calculated with the unscreened and Debye-screened HFS models Model: R. Thiele, SKS et al., PRA 86, 033411 (2012). Photoelectron spectroscopy: B. Ziaja et al., JPB 46, 164009 (2013). Magnetic scattering of Co/Pt at FLASH: L. Müller et al., PRL 110, 234801 (2013). 10 100 1000 Photon energy (eV) 0.1 1 10 Cross section (Mb) Unscreened Debye λ D = 5 a 0 1s 2s 2p 0 0.1 0.2 0.3 0.4 0 2 4 6 8 10 12 Probability Charge state T = 30 eV 40 eV 60 eV 80 eV 100 eV SKS, R. Thiele et al., in preparation. Probabilities of individual charge states of Al plasma calculated by thermal HFS, selecting a specific configuration and providing a plasma electron density SKS, R. Thiele et al., submitted. K-shell thresholds of Al charge states in plasma, showing IPDs 3 4 5 6 7 Charge state 1500 1550 1600 1650 1700 1750 1800 1850 Energy (eV) experimental data unscr. HFS two-step HFS Al IV Al V Al VI Al VII Al VIII Ionization Potential Depression 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 7 8 9 10 Effective scattering strength Photon energy (keV) low intensity high intensity 10 5 0 5 10 7 8 9 10 Photon energy (keV) || Fe 0+ 1s 2 2s 2 2p 6 3s 2 3p 6 3d 6 4s 2 Fe 10+ 1s 2 2s 2 2p 6 3s 2 3p 4 Fe 10+ 1s 1 2s 2 2p 6 3s 2 3p 5 Fe 20+ 1s 2 2s 2 2p 2 Fe 20+ 1s 1 2s 2 2p 3 Dispersion corrections of Fe ions Effective scattering strength Basic theory and generalized Karle–Hendrickson equation: SKS, H. Chapman & R. Santra, PRL 107, 218102 (2011). Configurational and temporal fluctuations and MAD coefficients: SKS, H. Chapman & R. Santra, JPB 46, 164015 (2013). Phasing is crucial to determine the protein structure. Our method has been applied to femtosecond nanocrystallography with XFEL.

Transcript of XATOM: an integrated toolkit for X-ray and atomic physics · molecule imaging from small molecules...

Page 1: XATOM: an integrated toolkit for X-ray and atomic physics · molecule imaging from small molecules to biological macromolecules solid electronic structure in extreme conditions ⇤

atom fundamental interaction between atoms and XFEL pulses

molecule imaging from small molecules to biological macromolecules

solid electronic structure in extreme conditions

⇤P(i,⌅) =4

3�⇥2⌅Ni

li+1X

lj=|li�1|

l>2li + 1

����Z ⇥

0Pnili(r)P�lj (r) r dr

����2

�A(i, jj⇥) = �

NH

i Njj0

2li + 1

lj+lj0X

L=|lj�lj0 |

1X

S=0

X

li0

(2L+ 1)(2S + 1)|MLS(j, j⇥, i, i⇥)|2

�F(i, j) =4

3�3(Ii � Ij)

3NH

i Nj

4lj + 2· l>2li + 1

����Z 1

0

Pnili(r)Pnj lj (r) r dr

����2

d

dtPI(t) =

all config.X

I0 6=I

[�I0!IPI0(t)� �I!I0PI(t)]

f0(Q) =

Z�(r) eiQ·r d3r

pSO(i; I, I0) = 1�

����Z 1

0

Pnili(r; I)Pnili(r; I0) dr

����2

H = Hmol + HEM + Hint

HEM =X

k,�

!ka†k,�ak,�, !k = |k|/↵

Hint = ↵

Zd3x

†(x)

A(x) · r

i

� (x) +

↵2

2

Zd3x

†(x)A2(x) (x)

|Ii: initial state, |F i: final state

�FI = 2⇡�(EF � EI)

�����hF |Hint|Ii+X

M

hF |Hint|MihM |Hint|IiEI � EM + i✏

+ · · ·

�����

2

|Ii = ci|�Nel0 i, |F i = c†acj cj0 |�

Nel0 i

|Ii = ci|�Nel0 i|0i, |F i = ci0 |�Nel

0 ia†kF ,�F|0i

|Ii = | Nel0 i|NEMi, |F i = | Nel

F i|NEM�1i

|Ii = | Nel0 i|NEMi

|F i = | Nel0 ia†kF ,�F

|NEM�1i

ε=0

rs

ε=V(rs)

outer-ionization

inner-ionization

Sang-Kil Son1 and Robin Santra1,2 1CFEL, DESY, Germany / 2Department of Physics, University of Hamburg, Germany

Introduction Theoretical and numerical details

XATOM: an integrated toolkit for X-ray and atomic physics

X-ray free-electron lasers (XFEL) open a new era in science and technology, offering many unique opportunities that have not been conceivable with conventional light sources. Because XFELs produce ultrashort pulses with a very high X-ray photon fluence, materials interacting with XFEL pulses undergo significant radiation damage and possibly become highly ionized. To understand the underlying physics, it is crucial to describe detailed ionization and relaxation dynamics in individual atoms during XFEL pulses. Here we present an integrated toolkit to investigate X-ray-induced atomic processes and to simulate electronic damage dynamics. This XATOM toolkit can handle all possible electronic configurations of all atom/ion species, and calculate physical observables during/after intense X-ray pulses. By use of XATOM, we can explore many exciting XFEL-related phenomena from multiphoton multiple ionization to molecular imaging and warm dense matter formation.

Hamiltonian and perturbation theory To treat X-ray–atom interactions, we employ a consistent ab initio framework based on nonrelativistic quantum electrodynamics and perturbation theory. For implementation, we use the Hartree-Fock-Slater model with the Latter tail correction.

10−4

10−3

10−2

10−1

100

101

103 104 105 106 107 108 109

Num

ber o

f sca

ttere

d ph

oton

s

Fluence (photons/Å2)

1 fs FWHM10 fs FWHM

100 fs FWHM

Number of scattered photons for several pulse durations, showing more scatterings for a shorter pulse due to intensity-induced X-ray transparency or frustrated absorption, SKS, L. Young & R. Santra, PRA 83, 033402 (2011).

Center for Free-Electron Laser Science CFEL is a scientific cooperation of the three organizations:

DESY – Max Planck Society – University of Hamburg

followed by simultaneous multiphoton absorption, as energeticallyrequired to reach the next higher charge state17, is one proposed mech-anism, although the excitationof spectral features such as a giant atomicresonance may modify this simple picture18. Studies of high-intensityphotoabsorptionmechanisms in this wavelength regime have also beenconducted onmore complex targets3,19. For argon clusters, it was foundthat ionization is best described by sequential single-photon absorp-tion19 and thatplasmaeffects suchas inverse bremsstrahlung, importantat longer wavelengths (.100nm; refs 20, 21), no longer contribute. Forsolid aluminium targets, researchers recently observed the phenom-enon of saturated absorption (that is, a fluence-dependent absorptioncross-section) using 15-fs, 13.5-nm pulses and intensities up to1016Wcm22 (ref. 3).

In the short-wavelength regime accessible with the LCLS, singlephotons ionize deep inner-shell electrons and the atomic response toultra-intense, short-wavelength radiation (,1018W cm22, ,1 nm)can be examined experimentally. In contrast to the studies at longerwavelengths, all ionization steps are energetically allowed via single-photon absorption, a fact that makes theoretical modelling con-siderably simpler. We exploit the remarkable flexibility of the LCLS(photon energy, pulse duration, pulse energy) combined with highresolution electron and ion time-of-flight spectrometers, to monitorand quantify photoabsorption pathways in the prototypical neonatom.

X-ray ionization of neon using LCLS

We chose to study neon because notable changes in the electronicresponse occur over the initial operating photon energy range ofLCLS, 800–2,000 eV (l5 1.5–0.6 nm), as shown schematically inFig. 1. There and in the following, V, P and A refer to the ejectionof valence, inner-shell and Auger electrons, respectively. In all cases,sequential single-photon ionization dominates, although the differ-ing electron ejection mechanisms lead to vastly different electronicconfigurations within each ionization stage. The binding energy of a1s electron in neutral neon is 870 eV. For photon energies below this,the valence shell is stripped, as shown at the top of Fig. 1 in a VV…sequence. Above 870 eV, inner-shell electrons are preferentiallyejected, creating 1s vacancies that are refilled by rapid Auger decay,a PA sequence. For energies above 993 eV, it is possible to create‘hollow’ neon, that is, a completely empty 1s shell, in a PP sequenceif the photoionization rate exceeds that of Auger decay. For energiesabove 1.36 keV, it is possible to fully strip neon, as shown at thebottom of Fig. 1.

Figure 2a shows experimental ion charge-state yields at three dif-ferent photon energies, 800 eV, 1,050 eV and 2,000 eV. These photonenergies represent the different ionization mechanisms—valenceionization, inner-shell ionization and ionization in the regime farabove all edges of all charge stages of neon. Despite the relativelylarge focal spot for these studies, ,1 mm, the dosage at 2,000 eV forneon (dosage5 cross-section3 fluence) is comparable to that pro-posed for the biomolecule imaging experiment where a 0.1-mm focalspot was assumed2. At the maximum fluence of,105 X-ray photonsper A2, we observe all processes that are energetically allowed viasingle-photon absorption. Thus, at 2,000 eV, we observe Ne101 andat 800 eV we find charge states as high as Ne81 (a fractional yield of0.3%), indicating a fully-stripped valence shell. We note that valencestripping up to Ne71 was previously observed in neon for 90.5-eV,1.83 1015W cm22 irradiation18,22. At this intermediate photonenergy, 90.5 eV, the highest charge state can not be reached by asequential single-photon absorption process.

Figure 2b compares the experimental ion charge-state yields withtheoretical calculations based on a rate equation model that includesonly sequential single-photon absorption and Auger decay pro-cesses12. For simulations, two parameters are required, the X-rayfluence and pulse duration. The fluence (pulse energy/area) on targetmay be calculated from measured parameters for pulse energy andfocal spot size. The X-ray pulse energies quoted throughout this

paper were measured in a gas detector23 located upstream of thetarget; the actual pulse energy on target is reduced by five reflectionson B4C mirrors (for details, see Methods). The focal spot size wasestimated from measurements done during the commissioningperiod (J. Krzywinski, personal communication) using the methodof X-ray-induced damage craters imprinted in solid targets24.

The fluence calculated from these pulse-energy and spot-size mea-surements is corroborated by in situ ion-charge-state measurements,both at 800 eV, where ionization is dependent only on fluence andnot on intensity, and at 2,000 eV, where the observed ratio of Ne101/Ne91 resulting from photoionization of hydrogen-like neon (a pro-cess with a well-known cross-section) serves as a reliable calibrationtool. The fluence that matches the Ne101/Ne91 ratio agrees to within30% with that derived from the measured pulse energy (2.4mJ) andestimated focal spot size (,13 2mm2 full-width at half-maximum,FWHM) at 2,000 eV. This fluence predicts not only the ratio Ne101/Ne91, but also the absolute values of the fractional charge-state yield,as shown in the bottom panel of Fig. 2b. At 2,000 eV, the calculationspredict the overall trend of the charge-state yields well, but there areobvious differences—particularly at the lower charge states. Theodd–even charge-state alternation is much more pronounced inthe calculation than in the experiment. This is due to the fact thatthe calculation ignores shake-off25 and double-Auger processes26, andpredicts that 1s one-photon ionization produces charge states up toNe21 only. Experimentally, one observes a yield of,75% Ne21 and25% Ne31 from simple 1s ionization27. At 1,050 eV, the generaltrends are reproduced although differences due to the simplicity ofthe model are evident.

At 800 eV, the simulations, which include only valence-shell strip-ping, are in excellent agreement with the observed charge-state dis-tribution. The fluence, determined in situ by the 800-eV data andsimulation, is within 10% of that predicted by a ,2.13 increase infocal area when going from 2,000 eV to 800 eV (ref. 28). Here, thesimulation is more straightforward as no inner-shell processes areoperative. We note that nonlinear two-photon processes29, which

Ionization

V

P

A

Ne

Ne

Ne 100 fs

Time2,000 eV

1,050 eV

800 eV

Ne8+

Ne8+Ne10+

2 s,p

1 s

2 s,p

1 s

2 s,p

1 s

V

VV

VV

VV

VV

A

A

A

A

A

A

AP

P

P

P

P

P

P

PP

V

Figure 1 | Diagram of the multiphoton absorption mechanisms in neoninduced by ultra-intense X-ray pulses. X-rays with energies below 870 eVionize 2s,p-shell valence electrons (V, red arrow). Higher energy X-rays giverise to photoemission from the 1s shell (P, purple arrow), and in theconsequent Auger decay the 1s-shell vacancy is filled by a 2s,p-shell electronand another 2s,p electron is emitted (A, black arrow). These V, P and Aprocesses are shown inmore detail in the inset; they all increase the charge ofthe residual ion by one. Main panel, three representative schemes ofmultiphoton absorption stripping the neon atom. The horizontal directionindicates the time for which atoms are exposed to the high-intensity X-rayradiation field, and vertical steps indicate an increase in ionic charge due toan ionization step, V, P or A. Horizontal steps are approximately to scalewith a flux density of 150X-ray photons per A2 per fs, and indicate the meantime between photoionization events or Auger decay.

NATURE |Vol 466 | 1 July 2010 ARTICLES

57Macmillan Publishers Limited. All rights reserved©2010

Figure from Young et al., Nature 466, 56 (2010).

Diagrams of multiphoton absorption mechanisms in Ne induced by ultraintense X-ray pulses

Photoabsorption

Auger / Coster–Kronig decay

Fluorescence

Shake-off process: sudden approximation

Elastic X-ray scattering

Resonant elastic X-ray scattering (dispersion correction)

f(Q,⌅) = f0(Q) + f 0(⌅) + if 00(⌅)

f 0(⌅) = � 1

2⇥2�PZ 1

0

⌅02

⌅02 � ⌅2⇤P(⌅

0)d⌅0

f 00(⌅) = � ⌅

4⇥�⇤P(⌅)

d�S

d�(t) =

all config.X

I

PI(t)d�S

d�

���I=

✓d�

d�

T

all config.X

I

PI(t)|f0I (Q)|2

SO

P

A

F

S

RS

Rate equation model To simulate electronic dynamics during intense X-ray pulses, we employ the rate equation approach with all computed cross sections and rates for all possible n-hole configurations, and calculate charge state distribution, electron/fluorescence spectra, scattering signals, etc.

Diagrams of X-ray–atom interaction

P

AS RS

P

F

SOsingle-core-hole

double-core-hole “hollow atom”

scattering from any configurations

Ne at LCLS 1110 eV & 1225 eV

Ar at SACLA 5 keV

Kr at LCLS 1.5 keV & 2 keV

Xe at SACLA 5.5 keV

Xe at LCLS 1.5 keV & 2 keV

Xe at 4.5 keV

10-1

100

101

102

0.5 1 2

Nor

mal

ized

ratio

of N

e9+/N

e8+

Energy (mJ)

ExperimentTheory + adjusted σ(2)

Theory + σ(2) from N-H

1110 eVbelow threshold:quadratic

1225 eVabove threshold:

linear

The Ne9+/Ne8+ ratio as a function of X-ray pulse energy

SYTCHEVA, PABST, SON, AND SANTRA PHYSICAL REVIEW A 85, 023414 (2012)

10-4

10-3

10-2

10-1

100

101

102

103

104

1100 1105 1110 1115 1120 1125 1130 1135 1140

σ(2) (

10-5

3 cm

4 s)

Photon energy (eV)

experimental value

TDCIS, ∆ωp=20 eVTDCIS, ∆ωp=15 eVTDCIS, ∆ωp =11eVTDCIS, ∆ωp = 8 eVTDCIS, ∆ωp = 6 eV

LOPT

FIG. 2. (Color online) Effective two-photon ionization crosssection for Ne8+. The TDCIS results are given by Eq. (7) for severaldifferent pulse bandwidths (FWHM). The LOPT result is obtainedusing Eq. (9). The point at 1110 eV corresponds to the experimentalvalue of 7 × 10−54 cm4 s reported in Ref. [6].

field E0 = 0.03 a.u. was used. Also shown is the cross sectionσ

(2)LOPT(ω) given by Eq. (9). For the latter, we use the HFS

model [29], implemented within the XATOM code [30,31]. TheHFS model positions the intermediate resonances at lowerenergies than those obtained in TDCIS, therefore we shiftedthe curve σ

(2)LOPT(ω) such that the 1s2-1s4p resonance is at

the right position of 1127.1 eV. Doumy et al. [6] noticed thatin a similar perturbative calculation [11] the authors did notaccount for the 1s2-1s4p resonance. We have included thisresonance in both TDCIS and LOPT calculations. However, aswe see from Fig. 2, neither the inclusion of this resonance northe finite bandwidth of the radiation pulse taken into accountin TDCIS can explain the discrepancy of several orders ofmagnitude between the theoretical and experimental values.

Now, we use Eq. (8) to convolve the monochromatictwo-photon ionization cross section obtained with Eq. (9),with the spectral distribution function given by Eq. (10),and show the results in Fig. 3(a). One can see that withinthe bandwidth, off from the resonances, the cross sectionis substantially enhanced, because the main contribution tothe convolution in Eq. (8) comes from the resonance peaks.Indeed, for a bandwidth of 11 eV the cross section at 1110 eVis 1.6 × 10−55 cm4 s, thus is enhanced by at least one andone-half orders of magnitude with respect to the perturbativeresult (4 × 10−57 cm4 s). In Fig. 3(b), we show the relationbetween the pulse bandwidth #ωp and mean photon energyωin, which is needed for the calculated two-photon ionizationcross section to reach the experimentally found value of7×10−54 cm4 s. For a bandwidth of 17 eV, the calculated crosssection increases up to this value at the photon energy of1110 eV used in the experiment. Thus, our findings suggest thatthe main reasons for the enhanced two-photon ionization crosssection of Ne8+ at 1110 eV originate from the proximity ofthe 1s2-1s4p resonance, the chaoticity of the LCLS radiation,and the finite bandwidth of its pulses.

In connection with the study of two-photon ionizationof core electrons, it is worth mentioning another recent

10-3

10-2

10-1

100

101

102

103

1060 1070 1080 1090 1100 1110 1120 1130 1140

σ(2) (

10-5

3 cm

4 s)

Photon energy (eV)

(a)

experimental value

∆ωp=20 eV∆ωp=15 eV∆ωp=11 eV∆ωp = 8 eV∆ωp = 6 eV

LOPT

0

5

10

15

20

1110 1115 1120 1125

∆ ω

p (e

V)

Photon energy (eV)

(b)

FIG. 3. (Color online) (a) Two-photon ionization cross sectionfor Ne8+, given by Eq. (8). The perturbative result σ

(2)LOPT of Eq. (9)

(dotted line) is taken as a reference signal for averaging over differentbandwidths (FWHM) of the pulses. The point at 1110 eV correspondsto the experimental value of 7 × 10−54 cm4 s reported in Ref. [6].(b) Relation between the bandwidth #ωp and the mean photon energyωin for which the two-photon ionization cross section σ

(2)incoh is fixed

at 7 × 10−54 cm4 s.

experiment of Young et al. [5], where direct multiphotonionization of neon was completely shadowed by a sequenceof one-photon ionization events. One of the measurementshas been done at the photon energy of 800 eV, just belowthe K edge, 870 eV, of neutral neon. In this case, one x-rayphoton carries enough energy to ionize valence electrons,and therefore the valence-shell electrons are stripped in asequence of one-photon absorption processes. Creation of a1s-shell vacancy is possible only through the absorption oftwo photons. No evidence for this process was detected.

10-10

10-8

10-6

10-4

10-2

100

100 101 102 103 104 105

Ioni

zatio

n pr

obab

ility

Intensity (1015 W/cm2)

1s valence

FIG. 4. (Color online) Intensity dependence of the ionizationprobability of neutral neon, given by the IDM corrections δρ1s for1s electrons and δρ2s +

!m δρ2pm for valence electrons, at a photon

energy of 800 eV (below the one-photon ionization threshold for theK shell, but above the one-photon ionization threshold for the valenceshells). A deterministic pulse of 6-eV bandwidth (FWHM) is used.

023414-4

Two-photon ionization cross section of Ne8+

G. Doumy et al., PRL 106, 083002 (2011).

A. Sytcheva et al., PRA 85, 023414 (2012).

Ne

1s

2s2p

Xe

2s2p

1s

3s3p3d

4s4p4d

5s5p

K

L

M

N

O

Xe+✽ → Xe(n+1)+ + ne–

Ne+✽ → Ne2+ + e–

Decay complexity for light and heavy atoms

How many coupled rate equations? Ne: 63 Xe: > 70 millions Xe + relativity: > 15 billions Xe + relativity + excitation: infinity

BENEDIKT RUDEK et al. PHYSICAL REVIEW A 87, 023413 (2013)

becomes more prominent (such as in the 2.5 mJ data shown inFig. 1).

C. Fluence dependence of the multiphoton ionization process

Finally, the fluence dependence of the ionization processeswas investigated for both pulse lengths. The fluence iscalculated as the measured pulse energy multiplied by thetransmission and divided by the focal area; due to an assumedGaussian fluence distribution the result is multiplied by anotherfactor of 4 ln(2)/π . Because intensity is defined as the fluencedivided by the pulse length, the intensity is considerably higherfor the shorter pulses. In Fig. 5 the ion yield per shot inlogarithmic units is plotted against the peak fluence, likewise inlogarithmic units. Because the unsaturated ion yield dependson the fluence to the power of n, with n as the number ofabsorbed photons, the slope in a double-logarithmic graphgives the average number of photons that were absorbed inorder to reach a certain charge state. Resonant transitions caninfluence the yield curves in two ways: first, they increase thecross section so that a given yield is already found at lowerfluences than expected without resonances; second, they cansaturate a transition so that it loses its fluence dependence.Then the slope decreases by the number of saturated transitionsinvolved in the multiphoton ionization. For instance, this effectwas observed at the soft-x-ray free-electron laser the SPring-8Compact SASE Source for resonances in helium [32].

In Fig. 5, the long-pulse data are illustrated by dots, theshort-pulse data by stars, calculated yields are drawn as solid

10-3

10-2

10-1

100

11.0

Kr5+

Kr9+

Kr10+

Kr12+

Kr14+

Kr15+

Kr20+Ion

yiel

dpe

rsho

t(ar

b.un

its)

Pulse energy (mJ)

Peak fluence (µJ/µm2)

4 532

1

011

80 5 fs

FIG. 5. (Color online) Krypton ion yield for selected charge statesKrq+ as a function of x-ray fluence (Kr5+ on top; higher charge statesconsecutively follow beneath). Measurements at a photon energy of2.0 keV are compared for nominal FEL pulse lengths of 80 fs (dots)and 5 fs (stars). The fluence at 2.0 keV is calculated from the x-raypulse energy measured by the LCLS gas detectors (top axis) assuminga 3 × 3 µm2 focus and 35% beamline transmission. To correct forgas detector nonlinearities at FEL pulse energies below 0.8 mJ, thegas detector readings were recalibrated using a linear ion signal (thatis, H+ ions created from residual gas). Calculated ion yields (withoutinclusion of REXMI) are shown as solid lines. To guide the eye,dashed lines with integer slopes are drawn through the experimentaldata points before they reach saturation. Error bars in the experimentaldata reflect statistical error only.

lines, and dashed black lines with integer slopes are addedto guide the eye. For short pulses, the LCLS pulse energyfluctuation is between 0.1 and 0.5 mJ. Long pulses had apulse energy between 2 and 3 mJ at full beam and wereattenuated to distributions centered around 1.2, 0.5, 0.3, 0.15,and 0.07 mJ. Because the LCLS pulse energy monitors seemedto be disturbed by attenuation pressures required to reach pulseenergies below 0.8 mJ, the pulse energy was recalibrated withthe proton yield stemming from residual gas which is assumedto have a linear pulse energy dependence.

Seven exemplary krypton charge states are depicted inFig. 5: the ion yield of Kr5+ is biggest so it is depicted as thetopmost line, and higher charge states consecutively followbeneath. For Kr5+, the slope is 1 before the ion yield saturates,corresponding to single-photon ionization accompanied bya long Auger cascade. Other charge states up to Kr7+ (notdepicted) also have a slope of 1 and saturate at similar fluence.Charge states such as Kr8+, Kr11+, or Kr13+ have a nonintegerslope which lies between the slopes of the next lower and nexthigher charge states that are depicted here. The long-pulsedata saturate at higher ion yield than the theory line, indicatingthat the experimental focus has more low-fluence areas thanare simulated. The early saturation of the Kr5+ signal forshort pulses is most likely an artifact of the MCP detectorsaturation appearing at high ion yield combined with high(2 kV) extraction voltage used.

Two photons need to be absorbed to reach Kr9+ or Kr10+;Kr12+ requires the absorption of three photons according tothe slopes in Fig. 5. The yield for Kr12+ at high fluence islower than predicted, as the ionization to even higher chargestates is enhanced by resonances. Also, the next higher chargestates show signatures of resonant processes. Compared to theresonance-free calculations, the yield curve of Kr14+ is shiftedby about 3 µJ/µm2 to lower fluences due to the enhanced crosssection, the curve of Kr15+ is shifted by about 20 µJ/µm2.The slopes for those charge states lie between 4 and 5. Nosignificant differences in the slopes were found for short andlong pulses.

V. CONCLUSIONS

In summary, we investigated the sequential inner-shellmultiple ionization of krypton at photon energies above(2 keV) and below (1.5 keV) the L edge. We determinedthe number of sequentially absorbed photons to be one forKr5+, two for Kr9+, three for Kr12+, and four for Kr14+. Belowthe L edge, the experimental charge state distribution wasadequately reproduced by theoretical calculations based ona rate-equation model. The same theoretical approach fails,however, at a photon energy above the L edge, where theexperimental spectrum shows higher charge states. We explainthis enhancement by a resonance-enhanced x-ray multipleionization mechanism, i.e., resonant excitations followed byautoionization at charge states higher than Kr12+, where directL-shell photoionization at 2 keV is energetically closed.Interestingly, in xenon, where we recently studied multiphotonx-ray ionization at the same two photon energies, we observeddramatic ionization enhancement due to REXMI at 1.5 keV,and achieved good agreement with theory at 2 keV. We readilyunderstand this behavior since both photon energies are above

023413-6

Kr ion yield as a function of X-ray peak fluence, demonstrating REXMI processes, B. Rudek et al., PRA 87, 023413 (2013).

1 2 3 4 5 6 7 8 9 1010-3

10-2

10-1

100

Charge state

Ion

yiel

d (c

ount

s/sh

ot)

Experiment Theory (20 µJ/µm2) Theory (50 µJ/µm2) Theory (80 µJ/µm2)

Ar charge state distribution, used for XFEL fluence calibration, K. Motomura et al., JPB 46, 164024 (2013).

HHG of rare gases XATOM has been used for calculations of the three-step model of high harmonic generation; Collaboration with CFEL Ultrafast Optics and X-Rays Group, S. Bhardwaj, SKS et al., PRA (in press).

11

0 100 200 300 400

10-4

10-2

100

Energy (eV)

Abs

olut

e R

ecom

bina

tion

ampl

itude

squ

are

Argon

100 200 300 400

10-4

10-2

100

Energy (eV)

Krypton

Figure 1: Square of absolute values of the Recombination Amplitude of Argon and Krypton for Plane Wave (PW) and Scattering Eigenstate (SC) in Length Form (LF) and Acceleration Form (LF): blue dashed (PW-LF), red dashed (PW-AF), blue solid (SC-LF) and red solid (SC-AF). In order to compare the length and the acceleration form, the former has been multiplied by square of the pre-factor in Eq. (2).

Ar recombination amp.

Ion, electron, and photon spectra of heavy atoms, SKS & R. Santra, PRA 85, 063415 (2012).

charge state distribution

Auger electron

photoelectron

fluorescence

-10

-8

-6

-4

-2

0

0 5 10 15 20 25 30 35 40 45 50

Orb

ital b

indi

ng e

nerg

ies

(keV

)

Charge state

Xe

RydbergO-shellN-shellM-shellL-shell

2.0 keV1.5 keV

4.5 keV

5.5 keVSACLA & theory

LCLS & theory

theory

5.0 keV

Ionization potentials of Xe as a function of charge states

0

2

4

6

8

10

12

14

0 5 10 15 20 25

Ener

gy (k

eV)

Charge states

1.7 fs5.3 fs

3.6 fs

1.3 fs

41 fsPhotoionizationAuger/Coster-KronigFluorescence

Deep inner-shell multiphoton multiple ionization mechanism

H. Fukuzawa, SKS et al., PRL 110, 173005 (2013).

Collaboration with CFEL–MPG–ASG, B. Rudek, SKS et al., Nature Photon. 6, 858 (2012).

Comparison of experimental and theoretical charge state distributions, revealing new ionization channels via excitation of transient resonances (REXMI)

occurring at this photon energy. Within the expectation from asimple model of purely sequential single-photon absorption,charge states up to Xe32þ can potentially be reached with 2.0 keVphotons via sequential removal of 3d electrons, as can be seenfrom the binding energies in Fig. 2.

In striking contrast to such a simple consideration, we findcharge states as high as Xe36þ for the lower photon energy of1.5 keV. To the best of our knowledge, this is the highest ionizationstage ever created in an atom with a single electromagnetic pulse(that is, both by photon impact26,33 and by ion impact34). At1.5 keV photon energy, sequential removal of electrons from therespective ionic ground state ends at Xe26þ, where direct ionizationcloses as the ground-state ionization energy rises above the photonenergy (Fig. 2). This is in qualitative agreement with our simulationin Fig. 1b, which predicts a maximum charge state of Xe27þ (with astrong decrease beyond Xe26þ) for the X-ray fluence achieved in theexperiment. In the simulations, the charge states above Xe26þ stemfrom Auger decay of multiple-core-hole states, which are createdwith significant abundance towards the end of the ionizationsequence when the Auger lifetime of 3d holes starts to be

comparable to or even exceed (at Xe25þ) the average inversephoto-ionization rate of !9 fs (Supplementary Fig. S1). It shouldbe noted that, within our model, significantly higher charge statescannot be produced, even when assuming considerably higher X-ray fluences. Thus, simulations using a straightforward rate equationapproach, which have successfully described earlier experiments onNe and N2 in a broad wavelength range (including hollow atom cre-ation)2,3 and yield good agreement with the xenon data at photonenergies of 850 eV (ref. 13) and 2.0 keV, fail dramatically for ourexperimental results at 1.5 keV. At this photon energy, another effi-cient ionization process must play a role, boosting multiple ioniz-ation far beyond the limit intuitively expected for sequential one-photon absorption.

We therefore propose and provide evidence that the highlycharged ionic states produced at 1.5 keV are reached via resonantpathways, as described in the following and schematically illustratedin Fig. 2. These resonances, which occur in highly charged xenonions produced during the course of a single femtosecond X-raypulse, are not included in our simulations, which only take intoaccount bound-free transitions. Inclusion of the additional

1.5 keV (experiment)2 keV (experiment)

1.5 keV (theory)2 keV (theory)

10−3

10−4

0.01

0.1

1

10

Ion

yiel

d (a

.u.)

Ion

yiel

d pe

r sho

t (a.

u.)

35+

31+27+

22+21+

20+19+

18+17+

16+15+

14+

13+

12+ 11+

10+

9+

128Xe

129Xe

130Xe

131Xe

132Xe

134Xe

136Xe3+

36+

35+

34+

30+

33+

32+

29+31+

25+

28+26+27+

24+

0.05a

b

0.04

0.03

0.02

0.01

0.006,000 8,000 10,000 12,000 14,000 16,000 18,000

5,200 5,400 5,600 5,800 6,000

Xe8+

Xe7+

Xe6+

Xe5+

Xe4+

Ion time of flight (ns)

5,000 6,200

23+

1.5 keV, 2.4–2.6 mJ

2 keV, 2.4–2.6 mJ

36 34 32 30 28 26 24 22 20 18 16 14 12 10 8 6 4 2Xe charge state

Figure 1 | Comparison of experimental and simulated xenon charge state yields. a, Xenon ion TOF spectra at photon energies of 1.5 keV (black) and2.0 keV (red) for (nominally) 80 fs pulses with 2.4–2.6 mJ pulse energy as measured by the LCLS gas detectors upstream of the target. Assuming a3 × 3 mm2 X-ray focus and 35% beamline transmission at 2.0 keV, this corresponds to a peak fluence of !82–89 mJ mm22 at the target. At 1.5 keV, thispeak fluence is reduced by a factor of two (see Methods). b, Experimental xenon charge state distribution (bars) after deconvolution of overlapping chargestates and comparison to theory (circles with lines) calculated for an 80 fs X-ray pulse with a pulse energy of 2.5 mJ and integrated over the interactionvolume. The theoretical charge state distributions are scaled such that the total ion yield integrated over all charge states agrees with the total ion yield inthe experiment. Error bars for experimental data reflect the statistical error only. a.u., arbitrary units.

NATURE PHOTONICS DOI: 10.1038/NPHOTON.2012.261 ARTICLES

NATURE PHOTONICS | VOL 6 | DECEMBER 2012 | www.nature.com/naturephotonics 859

Hollow-atom formation

XeF2 at synchrotron

MAD at high x-ray intensity

HIP: High-Intensity Phasing

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1010 1011 1012 1013 1014

MAD

ã

Fluence (photons/µm2)

H

C

N

O

S

Collaboration with CFEL Coherent Imaging Group, SKS, L. Galli et al., in preparation.

Effective scattering strength of S at 6 keV

C60 at LCLS

R. W. DUNFORD et al. PHYSICAL REVIEW A 86, 033401 (2012)

FIG. 5. Charge-state distributions for XeF2 (squares with solidcurve) compared with Xe (circles with dashed curve) in coincidencewith Kα photons. (a) XeF2 breakup pattern for the subset of eventsinvolving two fluorine ions in the final state with charge states F3+

and F2+. We plot the fraction of events (in %) vs the total charge(i.e., the Xe charge state plus 5). (b) The distribution of total chargeaveraged over all XeF2 breakup modes.

charge than the atomic Xe case. In Fig. 5(b), the distributionof the total charge of all three ions, summed over all XeF2breakup modes, is compared with the charge-state distributionfor atomic Xe. The mean value of the charge-state distributionincreases from +8.1 in atomic Xe to +8.6 in molecular XeF2;the total charges of +11 and +12 have significantly higherprobabilities in comparison to the atomic case. Nonresonantphotoionization, in contrast to excitation of a K-shell electroninto the lowest unoccupied molecular orbital, would lead toan even larger charge enhancement. Note that an incrementor decrement of decay rates will by itself not change the finalcharge production. Also charge migration within the moleculedoes not affect the total charge state. Therefore, the fact thatXeF2 reaches higher charge states than Xe does is evidence thatICD-type channels exist in a deep inner-shell-decay cascade.Besides ICD, direct impact ionization by Auger electrons fromXe can lead to a charge enhancement by knockout of anotherelectron from the two F atoms. We estimate, based on electronimpact ionization cross sections, that the charge enhancementdue to direct impact ionization is at most +0.1 and thereforecannot account for our experimental observation.

The Fq+ breakup kinetic energies can be determined fromthe time-of-flight splittings between the faster and slowercomponents in Fig. 2(b) under the assumption of symmetricFq+ partner ions. These energies are compared in Table IIwith the Coulomb repulsion energies the ions would have ifthey were formed instantaneously at the ground-state Xe-Finternuclear separation of 1.97 A. The Coulomb energies were

TABLE II. Kinetic energies (eV) of Fq+ ions from fragmentationof XeF2 compared with Coulomb energies calculated at the ground-state Xe-F internuclear distance 1.97 A. Coulomb energies areaveraged over F and Xe partner ions weighted by charge-distributionprobabilities and ion detection efficiencies. Results for symmetricbreakup modes are listed in column 3 and for all breakup modes incolumn 4.

Fq+ Experiment Symmetric modes All modes

F1+ 27 ± 1 36 43F2+ 58 ± 1 78 77F3+ 93 ± 2 137 116F4+ 126 ± 7 218 157

calculated by averaging over all possible Xe and F partner ionsweighted by charge-distribution probabilities and detectionefficiencies. Results are also given for averaging over onlythe symmetric breakup modes. In either case, the measuredenergies are lower than the Coulomb energies, which suggeststhat the ions begin moving away from their ground-statepositions before their charges are fully developed.

III. THEORY

The decay cascade for atomic Xe after K-shell ionizationwas simulated with the XATOM code [25]. The calculationinvolves Auger and Coster-Kronig rates, fluorescence rates,and shake-off branching ratios after K-shell ionization, basedon the Hartree-Fock-Slater (HFS) method. We analyzed theappearance of vacancies in the 5p shell as a function of time,and the results are presented in Fig. 6. The Xe 5p shell will bemostly involved in forming molecular orbitals in combinationwith the outer valence orbitals of F in XeF2 [24]. Therefore,the dynamics of the 5p electrons in Xe is used as a model toestimate the time scale for which the cascade in XeF2 wouldstart to involve molecular orbitals and for which ionizationin the F atoms should be expected. Although the atomicmodel is not capable of directly reproducing the total chargeenhancement seen in the experiment, it provides nonethelessa valid means of estimating time scales for the hole dynamicsin the molecule. The number of inner-valence and core holesin Xe reaches its final value after about 2 fs. Subsequently,holes are produced in the outer valence shell, and therefore

0

1

2

3

4

5

6

7

8

0.01 0.1 1 10 100 1000

Num

ber

of h

oles

Time (fs)

Total number of holes

Number of holes in 5p

FIG. 6. (Color online) Time evolution of the total number of holesand the population of 5p holes in the decay of a Xe K-shell vacancy.

033401-4

Comparison of charge state distributions between isolated Xe atom and XeF2 molecule, R. W. Dunford et al., PRA 86, 033401 (2012).

Atomic data calculated by XATOM have been used for molecular dynamics of C60; Collaboration with Z. Jurek, B. F. Murphy et al., in preparation.

Screening effect

Al plasma at LCLS

Two-step HFS model

Atomic picture in dense matter: Muffin-tin-like potential with the Wigner-Seitz radius to describe an environment of an atom within a solid or a plasma

Photoabsorption cross sections for C calculated with the unscreened and Debye-screened HFS models

Model: R. Thiele, SKS et al., PRA 86, 033411 (2012). Photoelectron spectroscopy: B. Ziaja et al., JPB 46, 164009 (2013). Magnetic scattering of Co/Pt at FLASH: L. Müller et al., PRL 110,

234801 (2013).

EFFECT OF SCREENING BY EXTERNAL CHARGES ON . . . PHYSICAL REVIEW A 86, 033411 (2012)

0 0.5 1 1.5 2 2.5 3

r (Å)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

|Pnl

|2

UnscreenedDebye: λ

D=5 a

0

Ion-sphere: 2.0 g/cm3, Z

i=2

1s x1/4

2s

2p

FIG. 7. (Color online) Comparison of 1s, 2s, and 2p radial wavefunctions of neutral carbon for the unscreened, Debye-screened, andion-sphere models. For the Debye-screened model, λD = 5 a0 is used,and for the ion-sphere model, the ion density of 2.0 g/cm3 and Zi = 2are used.

to the screening effect. The atomic potential for boundelectrons is weakened by screening from free electrons,and bound electrons are pulled out due to the existence ofneighboring ions. The valence electrons are more affected bythose screening effects than the core electrons. We emphasizehere that although the plots of the screening-induced absoluteenergy shifts, "E, show the largest energy shifts for coreelectrons, the respective relative shifts calculated in respectto the unscreened energy level are smallest for the deep-lyingorbitals.

B. Photoabsorption cross section

The effect of screening on the photoabsorption cross sectionis shown in Fig. 8. Here we compare the total photoabsorptioncross section for carbon in the ground state, calculated withdifferent models. The reference is the result obtained from theunscreened HFS model. Debye screening shifts the ionizationthreshold. As expected, for λD = 10.0 a0, this shift is corre-spondingly smaller than for λD = 5.0 a0. The Debye-screenedcross section is “extrapolated” towards the shifted threshold(cf. Fig. 9 for Debye-screened photoabsorption from differentshells) without much affecting the cross-section magnitude.In the case of ion-sphere screening, the thresholds are shiftedtowards the “screened” value, but the magnitude of the crosssection also increases in the vicinity of the threshold. Again,this effect is more pronounced in the case of the denser ionenvironment. The important observation is that the screeningeffects manifest themselves in the cross section only at lowerphoton energies (VUV, soft x-ray). The hard-x-ray regimeseems to be unaffected by the screening. This is significantfor CDI simulations, as it may justify using the unscreenedcross sections in such simulations.

C. Auger and fluorescence processes

Auger and fluorescence effects are relaxation processes ofan inner-shell vacancy within an atom. Photoabsorption at

10 100 1000Photon energy (eV)

0.1

1

10

100

Cro

ss s

ecti

on (M

b)

UnscreenedDebye λ

D= 5 a

0

Debye λD

=10 a0

2.0 g/cm 3 , Zi=1

2.0 g/cm 3 , Zi=2

FIG. 8. (Color online) Total photoabsorption cross section forcarbon in the ground state calculated with XATOM as a function ofthe photon energy for (i) the unscreened HFS model (black solidline), (ii) the Debye-screened HFS model (λD = 5.0 a0, black double-dot-dashed line; λD = 10 a0, black dot-dashed line), and (iii) theion-sphere screened HFS model (density of 2.0 g/cm3, Zi = 1,2, reddashed line and red dotted line).

high x-ray energy typically leads to the creation of such coreholes. The relaxation processes are then an important part ofthe electronic damage. An accurate calculation of their rates isnecessary for preparing a reliable CDI simulation.

In Tables I and II, we show the Auger and fluorescencerates calculated for the Debye-screened and the ion-spherescreened HFS models. They are compared to the unscreenedHFS calculations [32]. All possible electronic configurationsof carbon for a given charge state are considered. For theAuger rate calculation, the Debye screening effect is onlyincorporated through the screened orbitals. Thus, the sameexpression for the Auger rate as in Ref. [32] is employed.

For the Debye calculation in Table I, we have chosenthe Debye screening length of λD = 5.0 a0. The screening

10 100 1000Photon energy (eV)

0.1

1

10

Cro

ss s

ecti

on (M

b)

UnscreenedDebye λ

D= 5 a0

1s

2s

2p

FIG. 9. (Color online) Subshell photoabsorption cross sectionfor carbon in the ground state for different shells calculated withXATOM for the unscreened HFS model (solid lines) and for theDebye-screened HFS model (λD = 5.0 a0, dashed lines) as a functionof the photon energy.

033411-5

0

0.1

0.2

0.3

0.4

0 2 4 6 8 10 12

Prob

abilit

y

Charge state

T = 30 eV

40 eV

60 eV

80 eV

100 eV

SKS, R. Thiele et al., in preparation.

Probabilities of individual charge states of Al plasma calculated by thermal HFS, selecting a specific configuration and providing a plasma electron density

SKS, R. Thiele et al., submitted.

K-shell thresholds of Al charge states in plasma, showing IPDs

3 4 5 6 7Charge state

1500

1550

1600

1650

1700

1750

1800

1850

Ener

gy (e

V)

experimental dataunscr. HFStwo-step HFS

Al IVAl V

Al VIAl VII

Al VIII

FIG. 2. (color online) K-shell thresholds for aluminum as a function of the charge state calculated

with the two-step HFS model. The experimental data are taken from Ref. [22]. The unscreened

HFS results correspond to the calculations for isolated ions.

according to the di↵erence between the inner-ionization energy (1538.1 eV) calculated with

the average-atom model at T=0 eV and the experimental binding energy (1559.6 eV) [34].

Note that the absolute accuracy of HFS binding energies is typically about 1%.

By taking the di↵erence of the ionization thresholds with and without the plasma environ-

ment, we can examine the lowering of the ionization potentials not only for K-shell electrons

but also for electrons in other subshells. For individual bound orbitals of Al ions, we calcu-

late the IPD as shown in Fig. 3(a). For isolated atoms, the ionization potential is given by

�"�j where "�j is the jth orbital energy from the unscreened HFS calculation. For atoms in

the plasma for T > 0 eV, the inner-ionization potential is calculated by "s � "j, where both

orbital energies are obtained from the two-step HFS calculation. Since the plasma screening

a↵ects each orbital di↵erently, it is expected that IPDs for individual orbitals are di↵erent.

7

Ionization Potential Depression

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

7 8 9 10

Effe

ctiv

e sc

atte

ring

stre

ngth

Photon energy (keV)

low intensity

high intensity

2

−10

−5

0

5

10

7 8 9 10Photon energy (keV)

|f ˝ |Fe0+ 1s22s22p63s23p63d64s2

Fe10+ 1s22s22p63s23p4

Fe10+ 1s12s22p63s23p5

Fe20+ 1s22s22p2

Fe20+ 1s12s22p3

FIG. 2. (Color online) Dispersion corrections of atomic form factorsfor selected configurations of several charge states of Fe.

toolkit has been extended to compute the dispersion correc-tion, f ′ + i f ′′. In Fig. 2, one can see remarkable changes ofthe dispersion correction for different configurations and dif-ferent charge states of Fe. Both f ′ and f ′′ have a singularposition at the K-shell edge, which is shifted to a higher ωby ∼1 keV as the charge state increases. The plotted curvesin Fig. 2 correspond to the configurations of the ground stateand the single-core-hole state (except for the neutral Fe) forgiven charge states. Since the MAD phasing method is basedon the dispersion correction of heavy elements, it is inevitablyrequired to take into account the electronic damage dynamicsand accompanying changes of the dispersion correction underintense x-ray pulses.In the MAD phasing method, the Karle–Hendrickson equa-

tion [21, 22] represents a set of equations of scattering crosssections at several different wavelengths (photon energies).The molecular scattering form factor is separated into normaland anomalous scattering terms and the phase information canbe derived from their interferences. In this Letter, we extendthe Karle–Hendrickson equation to intense x-ray pulses withextensive electronic damage on anomalous scatterers.Let P be any protein (or any macromolecule) whose struc-

ture we want to solve by coherent x-ray scattering. Let Hindicate heavy atoms and NH be the number of heavy atomsper macromolecule to be considered. Note that P excludes H.Our assumption is that only heavy atoms scatter anomalouslyand undergo damage dynamics during an x-ray pulse. It is jus-tified by the fact that the photon energy of interest is near theinner-shell ionization threshold of heavy atoms and the pho-toabsorption cross section σ of the heavy atom is much higherthan that of the light atom for a given range of ω . For exam-ple, σFe/σC ≈ 300 at 8 keV. The scattering intensity (per unitsolid angle) is evaluated by time-integrating over one pulse,

dI(Q,ω)dΩ

= FC(Ω)! ∞

−∞dt g(t)∑

IPI(t)

×

"

"

"

"

"

F0P (Q)+NH∑j=1

fIj (Q,ω)eiQ·R j

"

"

"

"

"

2

, (2)

where j denotes a heavy atom index and I indicates a globalconfiguration index. The global configuration for NH heavy

atoms is given by I = (I1, I2, · · · , INH ). Each I j indicates oneof the configurations for all possible charge states of the j-th heavy atom. PI(t) is the population of the I-th configura-tion at time t. It is assumed that changes in the configura-tions happen independently among heavy atoms, so the pop-ulation of I is given by a product of individual populations,PI(t)=ΠNH

j=1PIj (t). The incoherent summation over I properlyaverages all combinations of different configurations amongindividual heavy atoms. Since the summation over j is coher-ently made in Eq. (2), one can find a coherent summation overI j, which contributes to Bragg peaks as shown in the follow-ing discussion. Here F is the x-ray fluence and g(t) is thenormalized pulse envelope. Then the x-ray flux is given byFg(t), which is assumed to be spatially uniform throughoutthe sample. C(Ω) is a coefficient given by the polarization ofthe x-ray pulse.In Eq. (2), F0P (Q) is the molecular form factor for the pro-

tein (without any dispersion correction) and our purpose is tosolve its amplitude and phase, F0P (Q) = |F0P (Q)|exp[iφ0P(Q)].On the other side, fIj (Q,ω) is the atomic form factor (withthe dispersion correction) of the j-th heavy atom in its I j-thconfiguration. It is most interesting to consider only one pre-dominant species of heavy atoms in the sample. Accordingly,the heavy atoms are located at different positions of {R j} andundergo damage processes individually, but I j can be consis-tently denoted as IH . In addition, to simplify Eq. (2), one canintroduce a molecular form factor for heavy atoms of one type,

F0H(Q) = f 0H(Q)NH∑j=1

eiQ·R j , (3)

where f 0H(Q) indicates the normal scattering atomic form fac-tor for the ground-state configuration of the heavy atom.NowEq. (2) can be readily expanded to derive a generalized

Karle–Hendrickson equation,dI(Q,ω)dΩ

= FC(Ω)#

"

"F0P (Q)"

"

2+"

"F0H(Q)"

"

2 a(Q,ω)

+"

"F0P (Q)"

"

"

"F0H(Q)"

"b(Q,ω)cos$

φ0P(Q)−φ0H(Q)%

+"

"F0P (Q)"

"

"

"F0H(Q)"

"c(Q,ω)sin$

φ0P(Q)−φ0H(Q)%

+NH"

" f 0H(Q)"

"

2{a(Q,ω)− a(Q,ω)}

&

, (4)

where the coefficients depending on Q and ω are defined by

a(Q,ω) =1

'

f 0H(Q)(2∑

IHPIH | fIH (Q,ω)|

2 , (5a)

b(Q,ω) =2

f 0H(Q)∑IHPIH

'

f 0IH (Q)+ f ′IH (ω)(

, (5b)

c(Q,ω) =2

f 0H(Q)∑IHPIH f

′′IH (ω), (5c)

a(Q,ω) =1

'

f 0H(Q)(2

! ∞

−∞dt g(t)

"

" fH(Q,ω , t)"

"

2. (5d)

Here PIH =) ∞−∞ dt g(t)PIH (t) is the pulse-weighted averaged

population for the IH -th configuration. The relevant co-efficients from Eq. (5a) to Eq. (5d) are atom-specific and

Dispersion corrections of Fe ions Effective scattering strength

Basic theory and generalized Karle–Hendrickson equation: SKS, H. Chapman & R. Santra, PRL 107, 218102 (2011).

Configurational and temporal fluctuations and MAD coefficients: SKS, H. Chapman & R. Santra, JPB 46, 164015 (2013).

Phasing is crucial to determine the protein structure. Our method has been applied to femtosecond nanocrystallography with XFEL.