[1,2]\fnmKurtis \surBorne [2,3]\fnmTaran \surDriver

1]\orgdivJ.R. Macdonald Laboratory, Department of Physics, \orgnameKansas State University, \orgaddress\cityManhattan \postcode66506, \stateKansas, \countryUSA 2] \orgnameSLAC National Accelerator Laboratory, \orgaddress \cityMenlo Park \postcode94025, \stateCalifornia, \countryUSA 3]\orgdivStanford Pulse Institute, \orgnameSLAC National Accelerator Laboratory, \orgaddress \cityMenlo Park \postcode94025, \stateCalifornia, \countryUSA

4]\orgdivDepartment of Physics, \orgnameStanford University, \orgaddress, \cityStanford, \postcode94305, \stateCalifornia, \countryUSA

5]\orgdivDepartment of Applied Physics, \orgnameStanford University, \orgaddress, \cityStanford, \postcode94305, \stateCalifornia, \countryUSA

6]\orgdivDepartment of Physics, \orgnameUniversity of Connecticut, \orgaddress \cityStorrs, \postcode, \stateConnecticut, \countryUSA

Design and Performance of a Magnetic Bottle Electron Spectrometer for High-Energy Photoelectron Spectroscopy

kborne@phys.ksu.edu    \fnmJordan T. \surO’Neal    \fnmJun \surWang    \fnmErik \surIsele    \fnmRazib \surObaid    \fnmNora \surBerrah    \fnmXinxin \surCheng    \fnmPhilip H. \surBucksbaum    \fnmJustin \surJames    \fnmAndrei \surKamalov    \fnmKirk A. \surLarsen    \fnmXiang \surLi    \fnmMing-Fu \surLin    \fnmYusong \surLiu    \fnmAgostino \surMarinelli    \fnmAdam \surSummers    \fnmEmily \surThierstein    \fnmThomas \surWolf    \fnmDaniel \surRolles    \fnmPeter \surWalter    \fnmJames P. \surCryan    tdriver@stanford.edu [ [ [ [ [ [
Abstract

We describe the design and performance of a magnetic bottle electron spectrometer (MBES) for high-energy electron spectroscopy. Our design features a 2similar-toabsent2{\sim 2}∼ 2 m long electron drift tube and electrostatic retardation lens, achieving sub-electronvolt (eV) electron kinetic energy resolution for high energy (several hundred eV) electrons with close to 4π𝜋\piitalic_π collection efficiency. A segmented anode electron detector enables the simultaneous collection of photoelectron spectra in high resolution and high collection efficiency modes. This versatile instrument is installed at the TMO endstation at the LCLS x-ray free-electron laser (XFEL). In this paper, we demonstrate its high resolution, collection efficiency and spatial selectivity in measurements where it is coupled to an XFEL source. These combined characteristics are designed to enable high-resolution time-resolved measurements using x-ray photoelectron, absorption, and Auger-Meitner spectroscopy. We also describe the pervasive artifact in MBES time-of-flight spectra that arises from a periodic modulation in electron detection efficiency, and present a robust analysis procedure for its removal.

keywords:
ultrafast, attosecond, x-ray free electron laser, photoionization delay

1 Introduction

Electron spectroscopy is a powerful tool for probing the electronic and molecular structure of gaseous and condensed phase systems. When used in combination with ultrashort light pulses, the technique offers a route to tracking ultrafast electronic and nuclear dynamics. Time-resolved photoelectron spectroscopy (TRPES) of valence electrons is an established technique for probing photo-initiated molecular dynamics on the femtosecond timescale[1, 2, 3, 4]. The binding energies of core orbitals can be particularly sensitive to the local chemical environment, so the nascent ability to time-resolve small binding energy shifts in inner shells promises to be a useful tool for tracking dynamics in molecular systems [5, 6, 7, 8, 9]. Resonant x-ray absorption measured using Auger-Meitner spectroscopy also offers a powerful probe of the transient localized valence electron density in molecules [10, 11, 12].

A magnetic bottle electron spectrometer (MBES) combines a strong (similar-to{\sim}1 T), inhomogenous magnetic field at the interaction point with a weaker (similar-to{\sim}1.5 mT) uniform magnetic field through a long drift region (or flight tube). This configuration achieves a high-collection efficiency for electrons while maintaining a good kinetic energy resolution. Electrons emitted in different directions from the interaction point are guided via𝑣𝑖𝑎viaitalic_v italic_i italic_a the Lorentz force along the magnetic field lines and into the flight tube, where they undergo betatron oscillations about the flight tube axis as they travel towards the detector. The original MBES design afforded 2π𝜋\piitalic_π solid angle collection volume, by collecting all electrons emitted in the hemisphere directed toward the flight tube [13, 14]. Later developments increased the collection efficiency to 4π𝜋\piitalic_π by exploiting the magnetic mirror effect [15]. Some designs [16, 17, 18] make use of an electrostatic retardation lens to reduce the electron velocity through the flight tube and so increase kinetic energy resolution. The high detection efficiency of the MBES is particularly advantageous for electron-electron coincidence or covariance measurements [19, 20, 21, 22], for which the signal-to-noise is critically dependent on collection efficiency [23]. Finally, MBES has proven particularly useful for electron spectroscopy at x-ray free-electron lasers (XFELs) [20, 11, 24, 22, 25].

In this work, we describe the design and performance of a MBES, which is installed at the time-resolved molecular and optical sciences (TMO) instrument at the Linac Coherent Light Source (LCLS) [26]. Our design features a 2similar-toabsent2{\sim}2∼ 2 m long flight tube, an electrostatic retardation lens and a segmented anode electron detector, which enables the simultaneous acquisition of electron spectra in two different collection modes: high resolution and high collection efficiency. We show that the high collection efficiency enables covariance mapping of photoelectron/Auger-Meitner electron emission in the core ionization of nitrous oxide. We employ spectral domain ghost imaging [27, 28, 29] in combination with the single-shot spectral diagnostics available at the beamline to isolate the resolution of our spectrometer from the inherent spectral width associated with a self-amplified spontaneous emission (SASE) XFEL. We achieve an electron energy resolution of δEE1/100similar-to𝛿𝐸𝐸1100\frac{\delta E}{E}\sim\nicefrac{{1}}{{100}}divide start_ARG italic_δ italic_E end_ARG start_ARG italic_E end_ARG ∼ / start_ARG 1 end_ARG start_ARG 100 end_ARG.

2 Instrument Design

Refer to caption
Figure 1: (A) Schematic of the MBES. Not to scale; the length of the flight tube has been truncated in this depiction. (B) To-scale cross-section of MBES taken from the graphical representation in the SIMION software package. Contour lines show regions of equivalent magnetic field strength.

A schematic of the MBES is shown in Fig. 1. The inhomogenous magnetic field at the interaction point is generated by a nickel-coated neodynium permanent magnet. A soft iron cone is mounted on the permanent magnet and acts as a magnetic shunt to focus the magnetic field lines at the interaction point. The permanent magnet is mounted to a remotely controlled three-axis manipulator to facilitate spectrometer alignment. The tip of the iron cone sits 6 mm opposite a copper nose cone, which is mounted on a 1.951.951.951.95 m long drift tube. The interaction point sits directly between the tip of the iron cone and the tip of the copper nose cone. The key elements of our design are modeled using the SIMION software (version 8.1) to simulate the trajectories of electrons through the MBES. These simulations are detailed throughout the results of the article.

The drift tube has an outer diameter of 76767676 mm and is wrapped by kapton-insulated, 22222222 AWG wire, which acts as an in-vacuum solenoid, with 1.2similar-toabsent1.2{\sim}1.2∼ 1.2 turns/mm. To achieve a uniform magnetic field of 1.5similar-toabsent1.5\sim 1.5∼ 1.5 mT in the drift tube, a current of 1similar-toabsent1{\sim}1∼ 1 A is applied to the coil. The drift tube is mounted inside a stainless-steel vacuum tube with DN160 CF flanges on each end, which is surrounded by a sleeve of high-magnetic-permeability metal to reduce the effects of external magnetic fields on electrons traveling through the tube by a factor of roughly 50.

There is an electrostatic lens in the drift tube to create a decelerating electric field to slow the electrons, which increases their time-of-flight, improving the energy resolution of the spectrometer. The position of the lens was chosen such that the magnetic field lines created by the combined fields of the permanent magnet and solenoid are parallel to the electric field lines of the lens. The lens begins 183similar-toabsent183{\sim}183∼ 183 mm from the interaction point and consists of two pairs of three electrostatic lens plates with 6.1similar-toabsent6.1{\sim}6.1∼ 6.1 mm spacing, separated by 30similar-toabsent30{\sim}30∼ 30 mm. Each set of three plates is electrically connected and a retarding effect is produced on the electrons by applying a potential between the two triplets.

Electrons are detected at the end of the drift tube by a 40404040 mm diameter micro-channel plate (MCP) detector coupled to a conical anode (Surface Concept GmbH). The anode is segmented into two concentric sections of 3 mm and 40 mm diameter, with a 1 mm spacing ring between them. In operation, a similar-to{\sim}1.8 kV potential difference is applied across the chevron MCP stack with a further 300similar-toabsent300{\sim}300∼ 300 volts to the anode. The voltages on the two anodes are decoupled from the high voltage source and amplified (Ortec 9306, 1111 GHz preamplifier), digitized with 168168168168 picosecond precision (Abaco Systems FMC134 analog-to-digital card), and read into our data recording system.

The segmented anode design was chosen to enable the simultaneous collection of spectra with high energy resolution (lower collection efficiency) and high collection efficiency (lower energy resolution). In a MBES, the time-of-flight distribution of monoenergetic electrons features a characteristic long tail, consisting of electrons emitted perpendicular to the time-of-flight axis but still directed down the flight tube by the inhomogenous magnetic field [13]. In simulation, we find that the electrons with longer time-of-flight impinge on the detector at larger radii (i.e. further from the center of the detector), as shown in Fig. 2 (A)-(C). The segmented anode enables the exclusion of this long time-of-flight tail by selectively detecting electrons which land close to the center of the detector. For measurements requiring higher detection efficiency, it is possible to also include electrons detected on the outer anode. Inclusion of these electrons causes a broadening of the peaks in time-of-flight, to the detriment of the kinetic energy resolution.

3 Instrument Performance

The MBES is currently installed in the LAMP [30] chamber at the TMO endstation [26] at the Linac Coherent Light Source (LCLS) XFEL. In this work we characterize the performance of the MBES for x-ray photoelectron spectroscopy with an XFEL source, making use of SASE pulses with a duration of <50absent50<50< 50 fs and a median pulse energy of approximately 50505050 μ𝜇\muitalic_μJ. X-rays were focused to a 1similar-toabsent1{\sim}1∼ 1 μ𝜇\muitalic_μm diameter at the interaction point of the MBES using a pair of Kirkpatrick-Baez mirrors [31]. The base pressure of the chamber housing the MBES was below 2×1092superscript1092\times 10^{-9}2 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT Torr and the sample gas was introduced to the interaction point using an effusive gas needle. The TMO endstation also features a transmissive Fresnel zone plate-based x-ray spectrometer which enables a shot-to-shot characterization of the incoming x-ray spectrum [32]. This measured x-ray spectrum is used for the spectral-domain ghost imaging analysis described in section 3.6.

3.1 Segmented Anode

Refer to caption
Figure 2: Motivation and performance of the segmented anode detector. (A) Simulated spatial distribution at the detector plane of 10,000 electrons. The electrons’ time-of-flight (ToF) is encoded in the color of the point. The initial kinetic energy of the electrons is 812similar-toabsent812{\sim}812∼ 812 eV and the retardation voltage is 400 V. (B) Time-of-flight spectrum for the electrons shown in panel (A). The blue curve shows the ToF spectrum discriminating on electrons landing within 1.5 mm of the center of the detector. (C) Dependence of ToF on radius at which electrons impinge on the detector for multiple kinetic energies for a retardation of 100 V. (D) Experimental ToF spectrum for neon KLL𝐾𝐿𝐿KLLitalic_K italic_L italic_L Auger-Meitner electrons produced following ionization by similar-to{\sim}1.35 keV x-rays, recorded with a retardation voltage of 400400400400 V. The red curve is for electrons detected on the outer anode and the blue curve is for electrons detected on the inner anode. The signal for the inner anode has been scaled by a factor of 30. (E) Kinetic energy representation of spectra in panel (D), compared to the spectrum produced by similar-to{\sim}1.5 keV photons obtained using an electrostatic analyzer  [33].

As described in section 2, the MBES features a segmented anode to discriminate the time-of-flight spectrum based on the position of electron impact on the MCP detector. The central part of the anode has a diameter of 3 mm, and is separated from the outer anode (diameter 40 mm) by a ring of 1 mm thickness. Panels (A)–(C) in Fig.  2  illustrate the motivation for this design. Panels (A) and (B) show simulation results for 10,000 electrons sampled from an isotropic distribution of emission angles, with an initial kinetic energy of 812812812812 eV and 400400400400 V retardation applied to the electrostatic lens. The coupling between electron time-of-flight and arrival position at the detector plane is shown in Fig. 2 (A). The simulations show that electrons contributing to the prompt peak in the time-of-flight distribution, which are initially emitted along the axis of the flight tube, impinge close to the center of the detector. The slower electrons, which produce the tail in the time-of-flight distribution and are emitted with a significant velocity component perpendicular to the time-of-flight axis, impinge on the detector at larger radii. This effect is made clear in Fig. 2 (B), which shows the simulated time-of-flight spectrum for all electrons (black dotted line), and only those which arrive within a 1.5 mm of the detector center (blue line). In panel (C) we explore the coupling between time-of-flight and detector position for different kinetic energies and find that it is a general effect: electrons which impinge at larger radii have a longer time-of-flight. The retardation voltage for the simulations in panel (C) was 100 V but we have verified that this behavior persists for all retardation voltages.

We experimentally investigate the performance of the segmented anode detector in panels D and E of Fig. 2, which show the Auger-Meitner electron spectrum of neon ionized by 1060106010601060 eV x-rays. The plot shows the measured electron distributions for the same shots, measured by the inner and outer anodes. While the outer anode has a significantly higher electron count rate, which is attributed to its much larger surface area, the energy resolution is degraded compared to the spectrum measured on the inner anode. In contrast, the resolution of the Auger-Meitner spectrum measured by the inner anode is improved but has a reduced overall count rate. We note that the performance of the segmented anode strongly depends on the alignment of the spectrometer. We achieve good alignment by maximizing the electron yield on the inner anode, by translating the permanent magnet using the motorized three-axis stage on which it is mounted.

3.2 Energy Calibration

We calibrate the MBES by the numerical calculation of electron trajectories using SIMION. We simulate the trajectories for electrons emitted over an isotropic distribution of emission angles, and fit the known initial kinetic energy to the simulated time of flight for different retardation voltages. To account for potential deviations from the standard KE1/(tt0)2proportional-to𝐾𝐸1superscript𝑡subscript𝑡02KE\propto\nicefrac{{1}}{{\left(t-t_{0}\right)^{2}}}italic_K italic_E ∝ / start_ARG 1 end_ARG start_ARG ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG behavior of a regular time-of-flight spectrometer, and an additional dependence of the calibration on the retardation voltage, we fit for a mapping from time-of-flight to kinetic energy over retardation which takes the form:

KE(t)=i=05pn(1(tt0)2)n,𝐾𝐸𝑡superscriptsubscript𝑖05subscript𝑝𝑛superscript1superscript𝑡subscript𝑡02𝑛\displaystyle KE(t)=\sum_{i=0}^{5}p_{n}\left(\frac{1}{\left(t-t_{0}\right)^{2}% }\right)^{n},italic_K italic_E ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( divide start_ARG 1 end_ARG start_ARG ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , (1)

where tt0𝑡subscript𝑡0t-t_{0}italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the time-of-flight of the electron referenced to its arrival time of the x-ray pulse at the interaction point, t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and pnsubscript𝑝𝑛p_{n}italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is a function of retardation voltage VRsubscript𝑉𝑅V_{R}italic_V start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT:

pn=i=06qi(n)(VR)i.subscript𝑝𝑛superscriptsubscript𝑖06superscriptsubscript𝑞𝑖𝑛superscriptsubscript𝑉𝑅𝑖\displaystyle p_{n}=\sum_{i=0}^{6}q_{i}^{\left(n\right)}\left(V_{R}\right)^{i}.italic_p start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( italic_V start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT . (2)

The dominant contribution to the fit comes from q1(0)superscriptsubscript𝑞10q_{1}^{\left(0\right)}italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 0 ) end_POSTSUPERSCRIPT, which indicates the behavior of the MBES is close to an ideal time-of-flight spectrometer where KE1/(tt0)2proportional-to𝐾𝐸1superscript𝑡subscript𝑡02KE\propto\nicefrac{{1}}{{\left(t-t_{0}\right)^{2}}}italic_K italic_E ∝ / start_ARG 1 end_ARG start_ARG ( italic_t - italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG, and that the mapping from time-of-flight to kinetic energy over retardation is largely independent of the value of the retardation voltage. The energy calibration agrees extremely well with our measurements, validating the accuracy of our numerical calculations for modeling the MBES.

3.3 Collection Efficiency

Refer to caption
Figure 3: Yield of nitrogen KLL𝐾𝐿𝐿KLLitalic_K italic_L italic_L Auger-Meitner electrons from N2O measured on the inner and outer anodes, as a function of applied solenoid current.

The collection efficiency of a MBES strongly depends on the kinetic energy of the electrons, the strength of the magnetic fields, and the retardation voltage applied to the electrostatic lens. Fig. 3 shows how the magnetic field strength of the solenoid influences the electron counts measured on the inner and outer anodes. This measurement was performed using nitrogen KLL𝐾𝐿𝐿KLLitalic_K italic_L italic_L Auger-Meitner and valence-shell photoemission from nitrous oxide (N2O) at a photon energy of 510510510510 eV. A retardation potential of 300300300300 V was applied to the electrostatic lens. The magnetic field strength was controlled by scanning the current applied to the solenoid coil. As the magnetic field strength of the solenoid (applied current) is increased, we observe an approximately linear increase in the detected electron counts on the outer anode (red curve in Fig. 3), which reaches a maximum at 1similar-toabsent1{\sim}1∼ 1 A. At this field strength, the cyclotron radius of the highest energy electrons emitted perpendicular to the spectrometer axis is small enough that almost all of the emitted electrons are directed onto the detector. As the current is further increased the electron trajectories are squeezed closer to the center of the flight tube and more electrons impinge on the inner anode, resulting in a corresponding decrease of the electron count rate on the outer anode.

Refer to caption
Figure 4: (A) Simulated values of the detection efficiency as a function of retardation voltage and kinetic energy over retardation. (B) Comparison of simulated detection efficiencies from panel (A) with experimentally measured yield of similar-to{\sim}100 eV photoelectrons measured at different retardation voltages (black points). (C) Simulated electron transmission from panel (A) for lower retardation voltages and electron kinetic energies.

We also studied the effect of the voltage applied to the retardation lens on the MBES collection efficiency. SIMION simulations show that the collection efficiency is almost independent of electron kinetic energy when no retardation field is applied to the electrostatic lens. Once a retarding field is applied, the collection efficiency vs. kinetic energy over retardation (i.e. the final kinetic energy of the electron through the flight tube) has a weak dependence on the applied retardation voltage. This behavior is illustrated in Fig. 4 (A).

We validate the performance of our SIMION model using the measured yield of 100similar-toabsent100{\sim}100∼ 100 eV nitrogen K𝐾Kitalic_K-shell photoelectrons ionized from N2O by 510510510510 eV x-rays. The black points in Fig. 4 (B) show the electron yield as a function of kinetic energy over retardation. This data was collected by scanning the applied retardation voltage while keeping the x-ray photon energy fixed at 510510510510 eV. The colored lines show the simulated detection efficiency for different retardation voltages selected from panel (A). In both experiment and simulation, we observe a steep decrease in collection efficiency as the retardation voltage approaches the initial kinetic energy of the electron. Panel (C) shows the region of panel (A) corresponding to retardation voltages below 50505050 V. At lower retardation voltages and kinetic energies, the detection efficiency depends more strongly on the retardation voltage.

3.4 Electron-electron covariance

Refer to caption
Figure 5: (A) Schematic of Auger-Meitner decay following nitrogen K𝐾Kitalic_K-shell photoionization of N2O by similar-to{\sim}790 eV x-rays. (B) Region of covariance map (Eqn. 3) of the electron kinetic spectrum following 790 eV ionization of N2O. (C) Partial covariance map (Eqn. 4) demonstrating suppression of spurious correlations in panel (B). This reveals correlation between photoelectrons from ionization of the nitrogen K𝐾Kitalic_K-shell (similar-to{\sim}380-400eV) and Auger-Meitner emission (similar-to{\sim}340-360eV), highlighted with the blue rectangle. The inset shows the average spectral profile of the incoming x-ray pulse, which produces the bimodal structure of the nitrogen K𝐾Kitalic_K-shell photoelectron spectrum.

To demonstrate an advantage of a high collection efficiency electron spectrometer such as our MBES, we performed an electron/electron covariance measurement on an N2O target ionized by 790790790790 eV x-rays. Ionization of the nitrogen K𝐾Kitalic_K-shell produced photoelectrons with a kinetic energy of 390similar-toabsent390{\sim}390∼ 390 eV. The core-ionized molecule decayed via Auger-Meitner emission, leading to the emission of a second electron with a kinetic energy of 350similar-toabsent350{\sim}350∼ 350 eV. A schematic of this decay process is shown in Fig. 5(A). We measure the kinetic energy spectrum of the resultant electrons with our MBES, using a retardation voltage of 250250250250 V applied to the electrostatic lens. For this measurement, we made use of the electrons detected on the outer anode to improve the effective collection efficiency of the MBES. We calculated the autocovariance of the electron kinetic energy spectrum across 40,000similar-toabsent40000{\sim}40,000∼ 40 , 000 XFEL shots [23]:

Cov[Y,X]=YXYX,Cov𝑌𝑋delimited-⟨⟩𝑌𝑋delimited-⟨⟩𝑌delimited-⟨⟩𝑋\mbox{Cov}[\vec{Y},\vec{X}]=\left\langle\vec{Y}\cdot\vec{X}\right\rangle-\left% \langle\vec{Y}\right\rangle\left\langle\vec{X}\right\rangle,Cov [ over→ start_ARG italic_Y end_ARG , over→ start_ARG italic_X end_ARG ] = ⟨ over→ start_ARG italic_Y end_ARG ⋅ over→ start_ARG italic_X end_ARG ⟩ - ⟨ over→ start_ARG italic_Y end_ARG ⟩ ⟨ over→ start_ARG italic_X end_ARG ⟩ , (3)

where X𝑋\vec{X}over→ start_ARG italic_X end_ARG is the electron kinetic energy spectrum, Y=XT𝑌superscript𝑋T\vec{Y}=\vec{X}^{\mbox{T}}over→ start_ARG italic_Y end_ARG = over→ start_ARG italic_X end_ARG start_POSTSUPERSCRIPT T end_POSTSUPERSCRIPT, and operator delimited-⟨⟩\langle\cdot\rangle⟨ ⋅ ⟩ indicates an average across the measured laser shots. Fig. 5 (B) shows a small region of the resultant covariance map. This map is dominated by strong covariance islands between electrons of kinetic energy 330400similar-toabsent330400{\sim}330-400∼ 330 - 400 and 265similar-toabsent265{\sim}265∼ 265 eV, which lie directly below the horizontal black line at 280280280280 eV. Electrons with kinetic energy similar-to{\sim}380–400 eV correspond to direct ionization from the N1s1𝑠1s1 italic_s shell, while electrons with energy similar-to{\sim}340–360 eV are produced by nitrogen Auger-Meitner emission. The electrons at 265similar-toabsent265{\sim}265∼ 265 eV are produced by ionization from the oxygen K𝐾Kitalic_K-shell.

In XFEL measurements, the shot-to-shot fluctuations of the spectral profile and pulse energy of the incoming x-ray pulse can produce spurious covariance islands which do not reflect the correlated emission of two photoproducts from the same ionization pathway. To remove these uninteresting correlations, here we calculate the multivariate partial covariance:

pCov[Y,X;I]=Cov[Y,X]Cov[Y,I](Cov[IT,I])1Cov[I,X],pCov𝑌𝑋𝐼Cov𝑌𝑋Cov𝑌𝐼superscriptCovsuperscript𝐼𝑇𝐼1Cov𝐼𝑋\mbox{pCov}[\vec{Y},\vec{X};\vec{I}]=\mbox{Cov}[\vec{Y},\vec{X}]-\mbox{Cov}[% \vec{Y},\vec{I}]\left(\mbox{Cov}[\vec{I}^{T},\vec{I}]\right)^{-1}\mbox{Cov}[% \vec{I},\vec{X}],pCov [ over→ start_ARG italic_Y end_ARG , over→ start_ARG italic_X end_ARG ; over→ start_ARG italic_I end_ARG ] = Cov [ over→ start_ARG italic_Y end_ARG , over→ start_ARG italic_X end_ARG ] - Cov [ over→ start_ARG italic_Y end_ARG , over→ start_ARG italic_I end_ARG ] ( Cov [ over→ start_ARG italic_I end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , over→ start_ARG italic_I end_ARG ] ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT Cov [ over→ start_ARG italic_I end_ARG , over→ start_ARG italic_X end_ARG ] , (4)

where I𝐼\vec{I}over→ start_ARG italic_I end_ARG is the single-shot x-ray spectrum and (Cov[IT,I])1superscriptCovsuperscript𝐼𝑇𝐼1\left(\mbox{Cov}[\vec{I}^{T},\vec{I}]\right)^{-1}( Cov [ over→ start_ARG italic_I end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , over→ start_ARG italic_I end_ARG ] ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT is the inverse of the autocorrelation of the spectrum. This partial covariance map is shown in Fig. 5 (B). In this map we observe that the spurious correlation between the electrons at 330400similar-toabsent330400{\sim}330-400∼ 330 - 400 and 265similar-toabsent265{\sim}265∼ 265 eV has been suppressed. We also observe a strong positive correlation away from the diagonal of the map, indicating that the emission of electrons with 390similar-toabsent390{\sim}390∼ 390 eV of kinetic energy is correlated with the emission of the 360360360360 eV Auger-Meitner electrons as illustrated in panel (A). This is a direct experimental observation of the photoelectron/Auger-Meitner electron emission process following K𝐾Kitalic_K-shell ionization of N2O. The kinetic energy structure of the Auger-Meitner emission results from the population of different final states in the N2O dication. We note that the nitrogen photoelectron features have a dual peak structure. This is the result of the average spectral intensity of the incoming x-ray spectrum, which has a double-peak structure as shown in an inset in each panel.

3.5 Spatial Sensitivity of MBES

Refer to caption
Figure 6: (A) Schematic illustrating the coordinate definitions. x: time-of-flight axis of the magnetic bottle and x-ray polarization direction, y: axis of the diffusive gas jet; z: direction of x-ray propagation.(B) Electron yield from N2O ionized by 516516516516 eV x-rays as a function of magnet alignment according to coordinate definitions in panel (A). (C) Simulated collection efficiency of the MBES for a point source of electrons displaced from the interaction point by the yz𝑦𝑧yzitalic_y italic_z plane.

In contrast to open area spectrometers (such as traditional VMI [34] or COLTRIMS [35]) time-of-flight spectrometers collect electrons from a localized region. In a magnetic bottle spectrometer the volume of the collection region is small, due to the strong spatial localization of the magnetic field. This spatial selectivity offers the opportunity to perform photoionization measurements which are highly selective to different positions along the focus of the ionizing radiation.

The collection efficiency of the MBES is sensitive to the alignment between the permanent magnet, the flight tube and the electron source. We can optimize the spectrometer alignment using the three-axis manipulator on which the magnet is mounted. We plot the dependence of the electron yield on the magnet position in Fig. 6 (B), which is sensitive to the spatial selectivity of the MBES. This measurement was performed using electrons with 100similar-toabsent100{\sim}100∼ 100 eV kinetic energy produced by nitrogen K𝐾Kitalic_K-shell ionization of N2O by 510similar-toabsent510{\sim}510∼ 510 eV x-rays. No retardation potential was applied to the lens. We observe a nearly equal dependence on collection efficiency for magnet motion along the axes perpendicular to the flight tube (y𝑦yitalic_y and z𝑧zitalic_z-axes), with a collection length of 1similar-toabsent1{\sim}1∼ 1 mm. The range of motion we are able to scan is limited by spatial constraints on the position of the magnet. The reduced sensitivity along the axis of the spectrometer (x𝑥xitalic_x-axis) is expected. Moving the magnet in this direction can affect the collection efficiency of electrons depending on their initial direction of emission [15]. The measurement shown in Fig. 6 is sensitive to the spatial selectivity of the MBES, but does not allow for its direct characterization. This is because we were unable to change the alignment between the flight tube and the electron source. Moreover, at the beam parameters available for the experiment the electron signal was produced by linear x-ray interactions. As a result, our effective electron source was highly extended along the direction of beam propagation. Nonetheless, we are able to investigate the spatial selectivity of the MBES in simulation. For similar parameters as the measurement shown in Fig. 6(B) (103 eV KE electrons, 0 V retardation), we simulate the spatial selectivity of the MBES by displacing a point source of electrons, emitted in a uniform distribution across the unit sphere, in a MBES where the magnet and flight tube are fully aligned. The direction of the source point displacement is perpendicular to the flight tube axis, which corresponds to any direction in the yz𝑦𝑧yzitalic_y italic_z plane illustrated in panel (A). The simulated dependence of detection efficiency on the source point position is shown in panel (C). We observe that in a well-aligned MBES, it is possible to strongly limit the collected electrons to a source with a spatial extent around 100 μ𝜇\muitalic_μm.

3.6 Energy Resolution

Refer to caption
Figure 7: Photoelectron kinetic energy spectrum of neon recorded at a photon energy of 1.35 keV. The three primary peaks of the Ne KLL𝐾𝐿𝐿KLLitalic_K italic_L italic_L Auger–Meitner emission spectrum become increasingly better resolved as the voltage applied to the electrostatic retardation lens is increased. The three dashed lines correspond to previously measured values of these features [33].

To investigate the energy resolution of the MBES and characterize the effect of the retardation lens, we performed a systematic scan of the retardation voltage. A higher retardation voltage is expected to produce higher energy resolution, because the energy resolution of an MBES δE/E𝛿𝐸𝐸\nicefrac{{\delta E}}{{E}}/ start_ARG italic_δ italic_E end_ARG start_ARG italic_E end_ARG is approximately linearly proportional to ΔT/TΔ𝑇𝑇\nicefrac{{\Delta T}}{{T}}/ start_ARG roman_Δ italic_T end_ARG start_ARG italic_T end_ARG [13].

Fig. 7 shows the measured photoemission kinetic energy spectrum of neon gas ionized by 1.35similar-toabsent1.35{\sim}1.35∼ 1.35 keV x-rays. We record this spectrum at six different retardation voltages, as indicated on the figure. The neon KLL𝐾𝐿𝐿KLLitalic_K italic_L italic_L Auger-Meitner emission has three dominant features separated by similar-to{\sim}30 eV, corresponding to different final states of the neon dication [33]. The expected energetic position of these three features is indicated by the vertical dotted lines. We observe that increasing the retardation voltage improves the energy resolution of the spectrometer: at higher retardation voltages the three primary KLL𝐾𝐿𝐿KLLitalic_K italic_L italic_L emission lines appear more clearly separated.

To quantify the resolving power of the MBES, we measured the widths of the photoionization features produced by K𝐾Kitalic_K-shell ionization of N2O as a function of retardation voltage. This measurement was performed using x-rays with a central photon energy of 516similar-toabsent516{\sim}516∼ 516 eV. The target molecule was chosen because the nitrogen K𝐾Kitalic_K-shell photoemission spectrum of N2O is well-studied and consists of two features corresponding to ionization of the central and terminal nitrogen sites at 412.5412.5412.5412.5 eV and 408.5408.5408.5408.5 eV, respectively [36].

The SASE FEL has a broad (similar-to{\sim}5 eV) spectral bandwidth and inherent shot-to-shot fluctuations in the spectral profile, which can significantly degrade spectroscopic resolution. To mitigate the effect of the SASE bandwidth and spectral jitter, and isolate the resolution of the MBES, we employed spectral domain ghost imaging [37, 28], as described in the appendix.

Refer to caption
Figure 8: (A) Dispersion plot of photoelectron kinetic energy vs.𝑣𝑠vs.italic_v italic_s . photon energy for the N(1s) photoionization of N2O, obtained using spectral domain ghost imaging. (B) Binding energy spectrum for N(1s) photoionization of N2O obtained from dispersion plot in panel A. The measured peak width decreases with increasing retardation voltage. Black line shows previous high-resolution measurements made with a synchrotron source from [36]. The change in relative intensities at larger retardation voltages is a result of the MBES transmission function, see Fig. 4. (C) Widths corresponding to gaussian fits of the terminal photoline centered at 408.5 eV binding energy. The dashed black line at 0.70 eV corresponds to the measured width from reference [36].

An example of the two-dimensional map of photoelectron kinetic energy vs. incoming photon energy is presented in Fig. 8 (A). We extract the kinetic energy spectrum from the two-dimensional map by averaging across all photon energies after accounting for the energy dispersion of the photoelectron. The width of the measured photoelectron feature for different retardation values is shown in Fig.8 (B) and can be seen to decrease with increasing retardation. The MBES detection efficiency at very low values of kinetic energy over retardation is highly suppressed, as shown in Fig. 4. This effect results in a significant decrease in yield for the lower energy photoelectrons at a retardation of 90.5 V. We quantify the resolution by fitting a gaussian curve to the photoline of the terminal nitrogen site at 107 eV. The dependence of this width on retardation is shown in Fig. 8 (C). The results show that when there is no retardation voltage applied, it is still possible to discern the 4 eV separation between the central and terminal lines thanks to the long flight tube. Nonetheless by increasing the retardation voltage, the kinetic energy resolution can be significantly increased.

4 Energy-Dependent Modulation in Electron Transmission Efficiency

Refer to caption
Figure 9: (A) SIMION simulation of trajectories of similar-to{\sim}300 eV electrons with different initial emission directions through the MBES. The retardation voltage is 0 V and the solenoid current has been reduced from the standard operation parameters to amplify the radius of the cyclotron oscillation. (B) Zoom-in of the trajectories at the end of the flight tube, close to the electron detector. (C) Dependence of electron trajectories through the flight tube on the angle of direction of initial emission, relative to the flight tube axis. The dotted black line shows the central axis of flight tube.

Finally, we discuss an artifact that is common to magnetic bottle electron spectrometers, namely a strongly energy-dependent modulation of the electron transmission function which is often amplified when a retarding field is applied to the electrostatic lens. This artifact is a result of the betatron oscillations of the electrons as they travel through the magnetic-field of the flight tube and propagate toward the detector [38]. The cyclotron radius of these oscillations,

rC=2meKEe0|B|sinθf,subscript𝑟𝐶2subscript𝑚𝑒𝐾𝐸subscript𝑒0𝐵subscript𝜃𝑓r_{C}=\frac{\sqrt{2m_{e}KE}}{e_{0}|\vec{B}|}\sin{\theta_{f}},italic_r start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = divide start_ARG square-root start_ARG 2 italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_K italic_E end_ARG end_ARG start_ARG italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | over→ start_ARG italic_B end_ARG | end_ARG roman_sin italic_θ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT , (5)

depends on the electron kinetic energy KE𝐾𝐸KEitalic_K italic_E, the strength of the magnetic field B𝐵\vec{B}over→ start_ARG italic_B end_ARG, and the angle of the electron velocity relative to the flight tube axis, θfsubscript𝜃𝑓\theta_{f}italic_θ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT. The effect of this cyclotron motion is to induce a periodic modulation in the transverse distribution of electron trajectories as they move along the flight tube. The spatial period of the modulation, i.e. the distance between points where the electron spread is maximal, depends on the kinetic energy of the electron [38]:

lo(KE)=2π2meKEe0|B|cosθf.subscript𝑙𝑜𝐾𝐸2𝜋2subscript𝑚𝑒𝐾𝐸subscript𝑒0𝐵subscript𝜃𝑓l_{o}(KE)=2\pi\frac{\sqrt{2m_{e}KE}}{e_{0}|\vec{B}|}\cos{\theta_{f}}.italic_l start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT ( italic_K italic_E ) = 2 italic_π divide start_ARG square-root start_ARG 2 italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_K italic_E end_ARG end_ARG start_ARG italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | over→ start_ARG italic_B end_ARG | end_ARG roman_cos italic_θ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT . (6)

The size of the electron spatial distribution at the detector is therefore a function of the velocity of the electrons. If the flight tube is an integer multiple of losubscript𝑙𝑜l_{o}italic_l start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT, then the spatial extent of the electrons is quite small at the detector plane. In the other limiting case, if the flight tube length is a half-integer multiple of losubscript𝑙𝑜l_{o}italic_l start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT and the spatial extend of the electron distribution is maximal at the detector plane. The energy-dependent oscillations have consequences for the operation of a MBES: electrons arriving at the detector plane with radial displacement larger than the detector can miss the detector and fail to be detected. Thus, the electron detection efficiency of the MBES is periodically modulated with electron time-of-flight. In our spectrometer, there may be an additional decrease in transmission efficiency due to electrons impinging on the plates of the electrostatic retardation lens. This additional effect should be particularly pronounced at high retardation voltages, due to fringe fields from the electrostatic lens that act perpendicular to the time-of-flight axis.

This oscillatory behavior of the electron trajectories is illustrated in the simulations shown in Fig. 9. We plot ten trajectories of electrons emitted in arbitrary directions uniformly sampled on a sphere, with similar-to{\sim}300 eV initial kinetic energy. The dependence of the cyclotron radius on the direction of the electron velocity described in Eq. 5 manifests as the dependence of the trajectory on the initial emission angle, as shown in panel (C). This effect is often colloquially referred to as ‘sausaging’, because the three-dimensional shape of the electron trajectories through the flight tube is reminiscent of the shape of a link of sausages.

This oscillatory collection efficiency is observed in experiment in the measured electron time-of-flight spectra of x-ray (520520520520 eV) ionization of N2O molecules, shown in Fig. 10 (A). The strong periodic modulation of electron detection efficiency with the electron time-of-flight is apparent in the spectrum recorded on the inner anode of the detector. According to Eqn.��6, the modulation frequency should increase with increasing magnetic field, and thus increasing solenoid current. This effect can be seen in Fig. 10 (B) and (C), where the Fourier transform of the electron time-of-flight spectra recorded on the outer ((B)) and inner ((C)) sections of the detector is shown as a function of the current applied to the solenoid coil. There is a dominant Fourier component, indicated by the white dashed-line, whose frequency increases linearly with solenoid current in accordance with the well-known dependence of cyclotron frequency on magnetic field strength. Below 1.1similar-toabsent1.1{\sim}1.1∼ 1.1 A, the modulation is particularly pronounced on the outer section of the detector (outer anode), while above this value the effect dominates the inner anode. This is a result of the geometry of the two detectors: as we increase the solenoid current, the electron trajectories flatten to the central axis of the flight tube and the modulation amplitude grows stronger on the inner anode and less prominent on the outer.

Refer to caption
Figure 10: (A) Time-of-flight (ToF) spectrum for photoelectrons produced by ionization of N2O by 520520520520 eV x-rays, measured on the inner anode at a solenoid current of 2 A, showing a strong periodic modulation in detection efficiency. (B) Fourier transform of the ToF spectra as a function of applied solenoid current for the outer anode. (C) Fourier transform of the ToF spectra as a function of applied solenoid current for the inner anode. For frequency values greater than 0.01 GHz, the color scales in panels (B) and (C) have been multiplied by 50 and 25, respectively.

One solution to the periodic modulation in electron detection efficiency is to increase the solenoid current to reduce the cyclotron radius such that all electrons fall on the detector. However, increasing the solenoid current to the required value may be technically challenging, and also results in a decrease of electron energy resolution [13]. Instead we developed an analytical method to remove this artifact in data post-processing [39]. We demonstrate this routine using the electron spectrum recorded by our MBES following the ionization of gas-phase para-aminophenol at similar-to{\sim}252 eV. The time-of-flight spectrum was recorded with a retardation of 190190190190 V and is shown in Fig. 11 (A) before (dashed red) and after (dashed blue) removal of this artifact.

To remove the artifact, we treat the measured electron spectrum  D(t)𝐷𝑡D(t)italic_D ( italic_t ) as a product of the true electron spectrum S(t)𝑆𝑡S(t)italic_S ( italic_t ) and the transmission function that encodes the periodic features, which we approximate as Ts(t)=1acos(ωst)subscript𝑇𝑠𝑡1𝑎subscript𝜔𝑠𝑡T_{s}(t)=1-a\cos(\omega_{s}t)italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) = 1 - italic_a roman_cos ( italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_t ), where ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the period of oscillation in the transmission function. The measured data can then be written as

D(t)=S(t)Ts(t)=S(t)(1acos(ωst)).𝐷𝑡𝑆𝑡subscript𝑇𝑠𝑡𝑆𝑡1𝑎subscript𝜔𝑠𝑡D(t)=S(t)T_{s}(t)=S(t)(1-a\cos(\omega_{s}t)).italic_D ( italic_t ) = italic_S ( italic_t ) italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) = italic_S ( italic_t ) ( 1 - italic_a roman_cos ( italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_t ) ) . (7)

Filtering out the modulation frequency (ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT) from this transmission function with a Fourier filter will result in data loss because the Fourier transform of D(t)𝐷𝑡D(t)italic_D ( italic_t ) is given by the convolution rather than the sum,

D~(ω)=(S~T~s)(ω)2S~(ω)a(S~(ωωs)+S~(ω+ωs)).~𝐷𝜔~𝑆subscript~𝑇𝑠𝜔proportional-to2~𝑆𝜔𝑎~𝑆𝜔subscript𝜔𝑠~𝑆𝜔subscript𝜔𝑠\tilde{D}(\omega)=(\tilde{S}\ast\tilde{T}_{s})(\omega)\propto 2\tilde{S}(% \omega)-a\left(\tilde{S}(\omega-\omega_{s})+\tilde{S}(\omega+\omega_{s})\right).over~ start_ARG italic_D end_ARG ( italic_ω ) = ( over~ start_ARG italic_S end_ARG ∗ over~ start_ARG italic_T end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ( italic_ω ) ∝ 2 over~ start_ARG italic_S end_ARG ( italic_ω ) - italic_a ( over~ start_ARG italic_S end_ARG ( italic_ω - italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) + over~ start_ARG italic_S end_ARG ( italic_ω + italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ) . (8)

Attempts to filter out the shifted copies of S~~𝑆\tilde{S}over~ start_ARG italic_S end_ARG are likely to modify the extracted spectrum.

To avoid this problem, we first take the logarithm of the spectrum, which separates the ground truth spectrum and transmission function into a sum, which can be more easily separated by Fourier decomposition:

ln(D(t))=ln(S(t)Ts(t))=ln(S(t))+ln(1acos(ωt)).𝐷𝑡𝑆𝑡subscript𝑇𝑠𝑡𝑆𝑡1𝑎𝜔𝑡\ln(D(t))=\ln(S(t)T_{s}(t))=\ln(S(t))+\ln(1-a\cos(\omega t)).roman_ln ( italic_D ( italic_t ) ) = roman_ln ( italic_S ( italic_t ) italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_t ) ) = roman_ln ( italic_S ( italic_t ) ) + roman_ln ( 1 - italic_a roman_cos ( italic_ω italic_t ) ) . (9)

The final term can be written as a Fourier series as,

(ln(D))(ω)=𝐷𝜔absent\displaystyle\mathcal{F}(\ln(D))(\omega)=caligraphic_F ( roman_ln ( italic_D ) ) ( italic_ω ) = (ln(S))(ω)ln(1+a24)δ(ω)𝑆𝜔1superscript𝑎24𝛿𝜔\displaystyle\mathcal{F}(\ln(S))(\omega)-\ln\left(1+\frac{a^{2}}{4}\right)% \delta(\omega)caligraphic_F ( roman_ln ( italic_S ) ) ( italic_ω ) - roman_ln ( 1 + divide start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG ) italic_δ ( italic_ω )
+a(δ(ωωs)+δ(ω+ωs))𝑎𝛿𝜔subscript𝜔𝑠𝛿𝜔subscript𝜔𝑠\displaystyle+a(\delta(\omega-\omega_{s})+\delta(\omega+\omega_{s}))+ italic_a ( italic_δ ( italic_ω - italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) + italic_δ ( italic_ω + italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) )
+a24(δ(ω2ωs)+δ(ω+2ωs))+𝒪(a3).superscript𝑎24𝛿𝜔2subscript𝜔𝑠𝛿𝜔2subscript𝜔𝑠𝒪superscript𝑎3\displaystyle+\frac{a^{2}}{4}(\delta(\omega-2\omega_{s})+\delta(\omega+2\omega% _{s}))+\mathcal{O}(a^{3}).+ divide start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG ( italic_δ ( italic_ω - 2 italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) + italic_δ ( italic_ω + 2 italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ) + caligraphic_O ( italic_a start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) . (10)

The broadening at the harmonic  0.045rad/ns0.045𝑟𝑎𝑑𝑛𝑠0.045\nicefrac{{rad}}{{ns}}0.045 / start_ARG italic_r italic_a italic_d end_ARG start_ARG italic_n italic_s end_ARG due to the convolution in Eq. 8 is replaced by a narrowed peak after applying the logarithm, as shown in Fig. 11 (B). To recover the signal S(t)𝑆𝑡S(t)italic_S ( italic_t ), we filter out the harmonics of ωssubscript𝜔𝑠\omega_{s}italic_ω start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT from the log-scaled spectra using Butterworth filters to generate the frequency-domain filtered data (ln(D))(ω)𝐷𝜔\mathcal{F}(\ln(D))(\omega)caligraphic_F ( roman_ln ( italic_D ) ) ( italic_ω ). To recover the filtered data in real space Df(t)subscript𝐷𝑓𝑡D_{f}(t)italic_D start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ( italic_t ), we apply an inverse Fourier transform and then exponentiate. This method of examining the Fourier transform of the logarithm has been used to study situations where the Fourier transform contains periodic peaks, because it maps the convolution of two functions onto their sums [40].

Refer to caption
Figure 11: Demonstration of the filtering algorithm to remove structure caused by betatron motion of electrons in the magnetic bottle for the x-ray valence ionization spectrum of para-aminophenol. (A) Time-of-flight representation with the raw spectrum showing the artefact before (dashed red) and after (blue) applying the filtering algorithm. (B) FFT power spectrum at different points in the filtering algorithm. Red lines show the spectrum of the raw data before (tick dotted red line) and after (thin dotted red line) taking the logarithm, i.e. D(ω)𝐷𝜔D(\omega)italic_D ( italic_ω ) and (ln(D))(ω)𝐷𝜔\mathcal{F}(\ln(D))(\omega)caligraphic_F ( roman_ln ( italic_D ) ) ( italic_ω ). Blue lines show the spectrum of the log-filtered data before (thick blue line) and after (thin blue line) exponentiating, i.e. F(lnDf(ω))𝐹subscript𝐷𝑓𝜔F(\ln{D_{f}(\omega)})italic_F ( roman_ln italic_D start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ( italic_ω ) ) and D~f(ω)subscript~𝐷𝑓𝜔\tilde{D}_{f}(\omega)over~ start_ARG italic_D end_ARG start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ( italic_ω ). (C) Data in the electron kinetic energy representation before (blue dashed line) and after (black line) the filtering algorithm.

We also implemented the log-space Fourier filter algorithm on a two-dimensional measurement of the resonant oxygen Auger-Meitner emission from para-aminophenol. The X-ray photon energy was scanned over the oxygen K𝐾Kitalic_K-edge region from similar-to{\sim}505 to similar-to{\sim}550 eV, and the high-energy electron emission spectrum was recorded with 400400400400 V applied to the electrostatic lens. The resonant Auger-Meitner map (incoming x-ray photon energy vs. outgoing photoelectron energy) is shown in Fig. 12. The photoelectrons produced by x-ray ionization of the valence shell show a clear linear dispersion, and the resonant Auger-Meitner emission after promotion of an electron from the oxygen 1s1𝑠1s1 italic_s to the 2pπ2𝑝superscript𝜋2p\pi^{\ast}2 italic_p italic_π start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT orbital is clear at similar-to{\sim}532 eV. Above the oxygen 1s1𝑠1s1 italic_s ionization potential, the electron emission converges to the normal KLL𝐾𝐿𝐿KLLitalic_K italic_L italic_L Auger-Meitner emission.

In the left panel, the periodic modulation is clearly seen and is most apparent in the resonant and normal Auger-Meitner emission at higher photon energies. We note that this modulation does not disperse with photon energy because it is a function of electron time-of-flight only. Removal of this periodic modulation has been performed by an enhanced application of the algorithm described above. To eliminate the possibility of distorting the two-dimensional map, a single transmission function T(t)𝑇𝑡T(t)italic_T ( italic_t ) was developed and then applied uniformly to each photon energy bin as

Df(t,hν)=D(t,hν)/T(t)subscript𝐷𝑓𝑡𝜈𝐷𝑡𝜈𝑇𝑡D_{f}(t,h\nu)=\nicefrac{{D(t,h\nu)}}{{T(t)}}italic_D start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ( italic_t , italic_h italic_ν ) = / start_ARG italic_D ( italic_t , italic_h italic_ν ) end_ARG start_ARG italic_T ( italic_t ) end_ARG (11)

where Df���(t,hν)subscript𝐷𝑓𝑡𝜈D_{f}(t,h\nu)italic_D start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ( italic_t , italic_h italic_ν ) denotes the two-dimensional filtered data. To compute the transmission function, the raw data is summed over photon energy and the 1D curve is filtered. By dividing the unfiltered and filtered data, a robust transmission function can be developed as

T(t)=Dall(t)/Dall,f(t)𝑇𝑡subscript𝐷𝑎𝑙𝑙𝑡subscript𝐷𝑎𝑙𝑙𝑓𝑡T(t)=\nicefrac{{D_{all}(t)}}{{D_{all,f}(t)}}italic_T ( italic_t ) = / start_ARG italic_D start_POSTSUBSCRIPT italic_a italic_l italic_l end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_D start_POSTSUBSCRIPT italic_a italic_l italic_l , italic_f end_POSTSUBSCRIPT ( italic_t ) end_ARG (12)

This transmission function is more robust compared to treating each photon energy bin separately, because the filter is nonlinear. By increasing the statistics when calculating the transmission function, we can more effectively filter the periodic transmission function.

Refer to caption
Figure 12: Resonant oxygen Auger-Meitner spectra of para-aminophenol before (A) and after (B) the filtering algorithm to remove the periodic modulation in detection efficiency. The x𝑥xitalic_x-axes shows the electron kinetic energy spectrum measured with a retardation of similar-to{\sim}400 eV and the y𝑦yitalic_y-axes show the x-ray central photon energy. Between x-ray photon energies of 530–535 eV we observe resonant Auger-Meitner emission following resonant oxygen 1s \rightarrow valence excitation. At photon energies above 535 eV, the spectrum converges to regular oxygen KLL𝐾𝐿𝐿KLLitalic_K italic_L italic_L Auger-Meitner emission. We also observe dispersive lines due to x-ray valence ionization.

5 Conclusion

In this paper, we have presented the design and performance of a MBES for x-ray photoelectron spectroscopy. Our spectrometer is installed at the TMO endstation of the Linac Coherent Light Source and is available to users of the facility. It is equipped with an einzel-stack retardation lenses and a segmented anode, which enables high kinetic-energy resolution (δE/E1%similar-to𝛿𝐸𝐸percent1\delta E/E\sim 1\%italic_δ italic_E / italic_E ∼ 1 %) measurements of high energy (several hundred eV) photoelectrons. The segmented anode also enables the simultaneous collection of data in two different modes: high collection efficiency, using electrons detected on both anodes, and high resolution, using only electrons detected on the inner anode. We achieve high energy x-ray photoelectron spectroscopy measurements when coupled to a noisy x-ray free-electron laser by employing spectral domain ghost imaging, and we achieve covariance measurements of correlated photoelectron/Auger-Meitner electron emission from N2O. We present a robust analysis procedure to correct for the well-known periodic modulation in detection efficiency due to cyclotron motion of the electron in the MBES flight tube. Upgrades to the instrument will include the integration of an ion time-of-flight spectrometer for concurrent ion/electron spectroscopy. Our spectrometer can currently be used for a variety of user experiments, and will be capable of operating in the MHz regime with the the upgraded high-repetition-rate LCLS-II source.

Acknowledgements

We are grateful to Jan Metje, Markus Gühr and Richard Squibb for useful discussions on the design of the MBES and to John Pennachio and Denise Welch for excellent technical support. KDB and DR acknowledges the support of the Department of Energy, Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy Sciences, Office of Science, grant No. DE-FG02-86ER13491. KB acknowledges the support of the US Department of Energy, Office of Science, Office of Workforce Development for Teachers and Scientists, Office of Science Graduate Student Research (SCGSR) program. The SCGSR program is administered by the Oak Ridge Institute for Science and Education (ORISE) for the DOE. ORISE is managed by ORAU under contract number DE-SC0014664 DR acknowledges the hospitality and support of SLAC during his sabbatical. NB acknowledges the support of the Department of Energy, Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy Sciences, Office of Science, grant No. DE-SC0012376

Author Declarations

Conflict of Interest

The authors have no conflicts to disclose.

Data Availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

6 Appendix

6.1 Spectral Domain Ghost Imaging

Spectral domain ghost imaging exploits the correlation between electron and x-ray photon spectra on a single-shot basis to extract the spectral response of the sample under investigation [27, 28, 41, 29]. In the single photon regime, the correlation of the photoelectron spectra with the photon flux of the incident x-ray pulse can be represented as a system of linear equations Ax=b𝐴𝑥𝑏Ax=bitalic_A italic_x = italic_b. A𝐴Aitalic_A and b𝑏bitalic_b are matrices whose columns index the spectral bins of the photon and electron energy spectra respectively, and whose rows correspond to different shots. The unknown partial ionization cross section is encoded in the x𝑥xitalic_x matrix. Due to the inherent measurement noise in A𝐴Aitalic_A and b𝑏bitalic_b, determining x𝑥xitalic_x via𝑣𝑖𝑎viaitalic_v italic_i italic_a least-square optimization is typically a numerically unstable operation. Therefore, we solve for x𝑥xitalic_x by minimizing the following cost function, which includes regularization parameters which enforce known qualities of the solution x𝑥xitalic_x:

Axb22+λ1x1+λ2Lx22+Ind+(x)superscriptsubscriptnorm𝐴𝑥𝑏22subscript𝜆1subscriptnorm𝑥1subscript𝜆2superscriptsubscriptnorm𝐿𝑥22𝐼𝑛subscript𝑑𝑥||Ax-b||_{2}^{2}+\lambda_{1}||x||_{1}+\lambda_{2}||Lx||_{2}^{2}+Ind_{+}(x)| | italic_A italic_x - italic_b | | start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | | italic_x | | start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | | italic_L italic_x | | start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_I italic_n italic_d start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( italic_x ) (13)

Here, the λ𝜆\lambdaitalic_λ are hyperparameters that weigh the different regularization terms. For the measurements presented in this work, the first term imposes sparsity defined by the L1subscript𝐿1L_{1}italic_L start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT-norm λ1subscript𝜆1\lambda_{1}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and the second term imposes smoothness λ2subscript𝜆2\lambda_{2}italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, defined by the laplacian operator. The last term imposes non-negativity in x𝑥xitalic_x. We employed a quadratic programming optimization [42, 43] to minimize equation 13 and obtain the optimal value of x𝑥xitalic_x. We choose the values of the hyperparameters by selecting appropriate values along the L-hypersurface [44, 45]. For all measurements, the spectral response x𝑥xitalic_x was robust to the values of the hyperparameters over several orders of magnitude.

References

  • \bibcommenthead
  • Stolow et al. [2004] Stolow, A., Bragg, A.E., Neumark, D.M.: Femtosecond Time-Resolved Photoelectron Spectroscopy. Chemical Reviews 104(4), 1719–1758 (2004) https://doi.org/10.1021/cr020683w . Accessed 2022-05-02
  • Suzuki [2006] Suzuki, T.: Femtosecond Time-Resolved Photoelectron Imaging. Annual Review of Physical Chemistry 57(1), 555–592 (2006) https://doi.org/10.1146/annurev.physchem.57.032905.104601 . Accessed 2023-10-16
  • Neumark [2001] Neumark, D.M.: Time-resolved photoelectron spectroscopy of molecules and clusters. Annual review of physical chemistry 52(1), 255–277 (2001). Accessed 2014-02-26
  • Schuurman and Blanchet [2022] Schuurman, M.S., Blanchet, V.: Time-resolved photoelectron spectroscopy: the continuing evolution of a mature technique. Physical Chemistry Chemical Physics, 10–1039105885 (2022) https://doi.org/10.1039/D1CP05885A . Accessed 2022-08-22
  • Wernet et al. [2017] Wernet, P., Leitner, T., Josefsson, I., Mazza, T., Miedema, P.S., Schröder, H., Beye, M., Kunnus, K., Schreck, S., Radcliffe, P., Düsterer, S., Meyer, M., Odelius, M., Föhlisch, A.: Communication: Direct evidence for sequential dissociation of gas-phase Fe(CO)5 via a singlet pathway upon excitation at 266 nm. The Journal of Chemical Physics 146(21), 211103 (2017) https://doi.org/%****␣main.bbl␣Line␣125␣****10.1063/1.4984774 . Accessed 2023-10-09
  • Brauße et al. [2018] Brauße, F., Goldsztejn, G., Amini, K., Boll, R., Bari, S., Bomme, C., Brouard, M., Burt, M., De Miranda, B.C., Düsterer, S., Erk, B., Géléoc, M., Geneaux, R., Gentleman, A.S., Guillemin, R., Ismail, I., Johnsson, P., Journel, L., Kierspel, T., Köckert, H., Küpper, J., Lablanquie, P., Lahl, J., Lee, J.W.L., Mackenzie, S.R., Maclot, S., Manschwetus, B., Mereshchenko, A.S., Mullins, T., Olshin, P.K., Palaudoux, J., Patchkovskii, S., Penent, F., Piancastelli, M.N., Rompotis, D., Ruchon, T., Rudenko, A., Savelyev, E., Schirmel, N., Techert, S., Travnikova, O., Trippel, S., Underwood, J.G., Vallance, C., Wiese, J., Simon, M., Holland, D.M.P., Marchenko, T., Rouzée, A., Rolles, D.: Time-resolved inner-shell photoelectron spectroscopy: From a bound molecule to an isolated atom. Physical Review A 97(4), 043429 (2018) https://doi.org/10.1103/PhysRevA.97.043429 . Accessed 2023-10-09
  • Mayer et al. [2022] Mayer, D., Lever, F., Picconi, D., Metje, J., Alisauskas, S., Calegari, F., Düsterer, S., Ehlert, C., Feifel, R., Niebuhr, M., Manschwetus, B., Kuhlmann, M., Mazza, T., Robinson, M.S., Squibb, R.J., Trabattoni, A., Wallner, M., Saalfrank, P., Wolf, T.J.A., Gühr, M.: Following excited-state chemical shifts in molecular ultrafast x-ray photoelectron spectroscopy. Nature Communications 13(1), 198 (2022) https://doi.org/10.1038/s41467-021-27908-y . Number: 1 Publisher: Nature Publishing Group. Accessed 2022-12-13
  • Gabalski et al. [2023] Gabalski, I., Allum, F., Seidu, I., Britton, M., Brenner, G., Bromberger, H., Brouard, M., Bucksbaum, P.H., Burt, M., Cryan, J.P., Driver, T., Ekanayake, N., Erk, B., Garg, D., Gougoula, E., Heathcote, D., Hockett, P., Holland, D.M.P., Howard, A.J., Kumar, S., Lee, J.W.L., Li, S., McManus, J., Mikosch, J., Milesevic, D., Minns, R.S., Neville, S., Atia-Tul-Noor, Papadopoulou, C.C., Passow, C., Razmus, W.O., Röder, A., Rouzée, A., Simao, A., Unwin, J., Vallance, C., Walmsley, T., Wang, J., Rolles, D., Stolow, A., Schuurman, M.S., Forbes, R.: Time-Resolved X-ray Photoelectron Spectroscopy: Ultrafast Dynamics in CS 22{}_{\textrm{2}}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT Probed at the S 2p Edge. The Journal of Physical Chemistry Letters 14(31), 7126–7133 (2023) https://doi.org/10.1021/acs.jpclett.3c01447 . Accessed 2023-10-09
  • Al-Haddad et al. [2022] Al-Haddad, A., Oberli, S., González-Vázquez, J., Bucher, M., Doumy, G., Ho, P., Krzywinski, J., Lane, T.J., Lutman, A., Marinelli, A., Maxwell, T.J., Moeller, S., Pratt, S.T., Ray, D., Shepard, R., Southworth, S.H., Vázquez-Mayagoitia, A., Walter, P., Young, L., Picón, A., Bostedt, C.: Observation of site-selective chemical bond changes via ultrafast chemical shifts. Nature Communications 13(1), 7170 (2022) https://doi.org/10.1038/s41467-022-34670-2 . Number: 1 Publisher: Nature Publishing Group. Accessed 2023-10-13
  • McFarland et al. [2014] McFarland, B.K., Farrell, J.P., Miyabe, S., Tarantelli, F., Aguilar, A., Berrah, N., Bostedt, C., Bozek, J.D., Bucksbaum, P.H., Castagna, J.C., Coffee, R.N., Cryan, J.P., Fang, L., Feifel, R., Gaffney, K.J., Glownia, J.M., Martinez, T.J., Mucke, M., Murphy, B., Natan, A., Osipov, T., Petrović, V.S., Schorb, S., Schultz, T., Spector, L.S., Swiggers, M., Tenney, I., Wang, S., White, J.L., White, W., Gühr, M.: Ultrafast X-ray Auger probing of photoexcited molecular dynamics. Nature Communications 5(1), 4235 (2014) https://doi.org/10.1038/ncomms5235 . Accessed 2023-10-16
  • Wolf et al. [2017] Wolf, T.J.A., Myhre, R.H., Cryan, J.P., Coriani, S., Squibb, R.J., Battistoni, A., Berrah, N., Bostedt, C., Bucksbaum, P., Coslovich, G., Feifel, R., Gaffney, K.J., Grilj, J., Martinez, T.J., Miyabe, S., Moeller, S.P., Mucke, M., Natan, A., Obaid, R., Osipov, T., Plekan, O., Wang, S., Koch, H., Gühr, M.: Probing ultrafast pipi*/npi* internal conversion in organic chromophores via K-edge resonant absorption. Nature Communications 8(1), 29 (2017) https://doi.org/10.1038/s41467-017-00069-7 . Number: 1 Publisher: Nature Publishing Group. Accessed 2023-03-07
  • Barillot et al. [2021] Barillot, T., Alexander, O., Cooper, B., Driver, T., Garratt, D., Li, S., Al Haddad, A., Sanchez-Gonzalez, A., Agåker, M., Arrell, C., Bearpark, M. ., Berrah, N., Bostedt, C., Bozek, J., Brahms, C., Bucksbaum, P. ., Clark, A., Doumy, G., Feifel, R., Frasinski, L. ., Jarosch, S., Johnson, A. ., Kjellsson, L., Kolorenč, P., Kumagai, Y., Larsen, E. ., Matia-Hernando, P., Robb, M., Rubensson, J.-E., Ruberti, M., Sathe, C., Squibb, R. ., Tan, A., Tisch, J. . ., Vacher, M., Walke, D. ., Wolf, T. . ., Wood, D., Zhaunerchyk, V., Walter, P., Osipov, T., Marinelli, A., Maxwell, T. ., Coffee, R., Lutman, A. ., Averbukh, V., Ueda, K., Cryan, J. ., Marangos, J. .: Correlation-Driven Transient Hole Dynamics Resolved in Space and Time in the Isopropanol Molecule. Physical Review X 11(3), 031048 (2021) https://doi.org/10.1103/PhysRevX.11.031048 . Accessed 2022-08-19
  • Kruit and Read [1983] Kruit, P., Read, F.H.: Magnetic field paralleliser for 2pi electron-spectrometer and electron-image magnifier. Journal of Physics E: Scientific Instruments 16(4), 313–324 (1983) https://doi.org/10.1088/0022-3735/16/4/016 . Accessed 2022-02-08
  • Tsuboi et al. [1988] Tsuboi, T., Xu, E.Y., Bae, Y.K., Gillen, K.T.: Magnetic bottle electron spectrometer using permanent magnets. Review of Scientific Instruments 59(8), 1357–1362 (1988) https://doi.org/10.1063/1.1139722 . Accessed 2023-10-16
  • Cheshnovsky et al. [1987] Cheshnovsky, O., Yang, S.H., Pettiette, C.L., Craycraft, M.J., Smalley, R.E.: Magnetic time-of-flight photoelectron spectrometer for mass-selected negative cluster ions. Review of Scientific Instruments 58(11), 2131–2137 (1987) https://doi.org/10.1063/1.1139475 . Accessed 2023-08-31
  • Radcliffe et al. [2007] Radcliffe, P., Düsterer, S., Azima, A., Li, W.B., Plönjes, E., Redlin, H., Feldhaus, J., Nicolosi, P., Poletto, L., Dardis, J., Gutierrez, J.P., Hough, P., Kavanagh, K.D., Kennedy, E.T., Luna, H., Yeates, P., Costello, J.T., Delyseries, A., Lewis, C.L.S., Glijer, D., Cubaynes, D., Meyer, M.: An experiment for two-color photoionization using high intensity extreme-UV free electron and near-IR laser pulses. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 583(2-3), 516–525 (2007) https://doi.org/10.1016/j.nima.2007.09.014 . Accessed 2023-10-16
  • Mucke et al. [2012] Mucke, M., Förstel, M., Lischke, T., Arion, T., Bradshaw, A.M., Hergenhahn, U.: Performance of a short “magnetic bottle” electron spectrometer. Review of Scientific Instruments 83(6), 063106 (2012) https://doi.org/10.1063/1.4729256 . Accessed 2023-10-16
  • Hikosaka et al. [2014] Hikosaka, Y., Sawa, M., Soejima, K., Shigemasa, E.: A high-resolution magnetic bottle electron spectrometer and its application to a photoelectron–Auger electron coincidence measurement of the L2,3VV Auger decay in CS2. Journal of Electron Spectroscopy and Related Phenomena 192, 69–74 (2014) https://doi.org/10.1016/j.elspec.2014.01.017 . Accessed 2023-10-13
  • Eland et al. [2003] Eland, J.H.D., Vieuxmaire, O., Kinugawa, T., Lablanquie, P., Hall, R.I., Penent, F.: Complete Two-Electron Spectra in Double Photoionization: The Rare Gases Ar, Kr, and Xe. Physical Review Letters 90(5), 053003 (2003) https://doi.org/10.1103/PhysRevLett.90.053003 . Accessed 2020-10-10
  • Frasinski et al. [2013] Frasinski, L.J., Zhaunerchyk, V., Mucke, M., Squibb, R.J., Siano, M., Eland, J.H.D., Linusson, P., Meulen, P., Salén, P., Thomas, R.D., Larsson, M., Foucar, L., Ullrich, J., Motomura, K., Mondal, S., Ueda, K., Osipov, T., Fang, L., Murphy, B.F., Berrah, N., Bostedt, C., Bozek, J.D., Schorb, S., Messerschmidt, M., Glownia, J.M., Cryan, J.P., Coffee, R.N., Takahashi, O., Wada, S., Piancastelli, M.N., Richter, R., Prince, K.C., Feifel, R.: Dynamics of Hollow Atom Formation in Intense X-Ray Pulses Probed by Partial Covariance Mapping. Physical Review Letters 111(7), 073002 (2013) https://doi.org/10.1103/PhysRevLett.111.073002 . Accessed 2023-03-07
  • Zhaunerchyk et al. [2015] Zhaunerchyk, V., Kamińska, M., Mucke, M., Squibb, R.J., Eland, J.H.D., Piancastelli, M.N., Frasinski, L.J., Grilj, J., Koch, M., McFarland, B.K., Sistrunk, E., Gühr, M., Coffee, R.N., Bostedt, C., Bozek, J.D., Salén, P., Meulen, P.V.D., Linusson, P., Thomas, R.D., Larsson, M., Foucar, L., Ullrich, J., Motomura, K., Mondal, S., Ueda, K., Richter, R., Prince, K.C., Takahashi, O., Osipov, T., Fang, L., Murphy, B.F., Berrah, N., Feifel, R.: Disentangling formation of multiple-core holes in aminophenol molecules exposed to bright X-FEL radiation. Journal of Physics B: Atomic, Molecular and Optical Physics 48(24), 244003 (2015) https://doi.org/10.1088/0953-4075/48/24/244003 . Accessed 2023-10-16
  • Kornilov et al. [2013] Kornilov, O., Eckstein, M., Rosenblatt, M., Schulz, C.P., Motomura, K., Rouzée, A., Klei, J., Foucar, L., Siano, M., Lübcke, A., Schapper, F., Johnsson, P., Holland, D.M.P., Schlathölter, T., Marchenko, T., Düsterer, S., Ueda, K., Vrakking, M.J.J., Frasinski, L.J.: Coulomb explosion of diatomic molecules in intense XUV fields mapped by partial covariance. Journal of Physics B: Atomic, Molecular and Optical Physics 46(16), 164028 (2013) https://doi.org/10.1088/0953-4075/46/16/164028 . Accessed 2020-03-13
  • Frasinski [2016] Frasinski, L.J.: Covariance mapping techniques. Journal of Physics B: Atomic, Molecular and Optical Physics 49(15), 152004 (2016) https://doi.org/10.1088/0953-4075/49/15/152004 . Accessed 2023-10-13
  • Squibb et al. [2018] Squibb, R.J., Sapunar, M., Ponzi, A., Richter, R., Kivimäki, A., Plekan, O., Finetti, P., Sisourat, N., Zhaunerchyk, V., Marchenko, T., Journel, L., Guillemin, R., Cucini, R., Coreno, M., Grazioli, C., Di Fraia, M., Callegari, C., Prince, K.C., Decleva, P., Simon, M., Eland, J.H.D., Došlić, N., Feifel, R., Piancastelli, M.N.: Acetylacetone photodynamics at a seeded free-electron laser. Nature Communications 9(1), 63 (2018) https://doi.org/10.1038/s41467-017-02478-0 . Number: 1 Publisher: Nature Publishing Group. Accessed 2020-10-10
  • Pathak et al. [2020] Pathak, S., Ibele, L.M., Boll, R., Callegari, C., Demidovich, A., Erk, B., Feifel, R., Forbes, R., Di Fraia, M., Giannessi, L., Hansen, C.S., Holland, D.M.P., Ingle, R.A., Mason, R., Plekan, O., Prince, K.C., Rouzée, A., Squibb, R.J., Tross, J., Ashfold, M.N.R., Curchod, B.F.E., Rolles, D.: Tracking the ultraviolet-induced photochemistry of thiophenone during and after ultrafast ring opening. Nature Chemistry 12(9), 795–800 (2020) https://doi.org/10.1038/s41557-020-0507-3 . Number: 9 Publisher: Nature Publishing Group. Accessed 2021-03-05
  • Walter et al. [2022] Walter, P., Osipov, T., Lin, M.-F., Cryan, J., Driver, T., Kamalov, A., Marinelli, A., Robinson, J., Seaberg, M.H., Wolf, T.J.A., Aldrich, J., Brown, N., Champenois, E.G., Cheng, X., Cocco, D., Conder, A., Curiel, I., Egger, A., Glownia, J.M., Heimann, P., Holmes, M., Johnson, T., Lee, L., Li, X., Moeller, S., Morton, D.S., Ng, M.L., Ninh, K., O’Neal, J.T., Obaid, R., Pai, A., Schlotter, W., Shepard, J., Shivaram, N., Stefan, P., Van, X., Wang, A.L., Wang, H., Yin, J., Yunus, S., Fritz, D., James, J., Castagna, J.-C.: The time-resolved atomic, molecular and optical science instrument at the Linac Coherent Light Source. Journal of Synchrotron Radiation 29(Pt 4), 957–968 (2022) https://doi.org/10.1107/S1600577522004283 . Accessed 2023-03-08
  • Driver et al. [2020] Driver, T., Li, S., G. Champenois, E., Duris, J., Ratner, D., J. Lane, T., Rosenberger, P., Al-Haddad, A., Averbukh, V., Barnard, T., Berrah, N., Bostedt, C., H. Bucksbaum, P., Coffee, R., F. DiMauro, L., Fang, L., Garratt, D., Gatton, A., Guo, Z., Hartmann, G., Haxton, D., Helml, W., Huang, Z., LaForge, A., Kamalov, A., F. Kling, M., Knurr, J., Lin, M.-F., A. Lutman, A., P. MacArthur, J., P. Marangos, J., Nantel, M., Natan, A., Obaid, R., T. O’Neal, J., H. Shivaram, N., Schori, A., Walter, P., Wang, A.L., A. Wolf, T.J., Marinelli, A., P. Cryan, J.: Attosecond transient absorption spooktroscopy: a ghost imaging approach to ultrafast absorption spectroscopy. Physical Chemistry Chemical Physics 22(5), 2704–2712 (2020) https://doi.org/10.1039/C9CP03951A . Publisher: Royal Society of Chemistry. Accessed 2024-01-10
  • Li et al. [2021] Li, S., Driver, T., Al Haddad, A., Champenois, E.G., Agåker, M., Alexander, O., Barillot, T., Bostedt, C., Garratt, D., Kjellsson, L., Lutman, A.A., Rubensson, J.-E., Sathe, C., Marinelli, A., Marangos, J.P., Cryan, J.P.: Two-dimensional correlation analysis for x-ray photoelectron spectroscopy. Journal of Physics B: Atomic, Molecular and Optical Physics 54(14), 144005 (2021) https://doi.org/10.1088/1361-6455/abcdf1 . Accessed 2021-08-27
  • Wang et al. [2023] Wang, J., Driver, T., Allum, F., Papadopoulou, C.C., Passow, C., Brenner, G., Li, S., Düsterer, S., Noor, A.T., Kumar, S., Bucksbaum, P.H., Erk, B., Forbes, R., Cryan, J.P.: Photon energy-resolved velocity map imaging from spectral domain ghost imaging. New Journal of Physics 25(3), 033017 (2023) https://doi.org/10.1088/1367-2630/acc201 . Publisher: IOP Publishing. Accessed 2023-10-27
  • Osipov et al. [2018] Osipov, T., Bostedt, C., Castagna, J.-C., Ferguson, K.R., Bucher, M., Montero, S.C., Swiggers, M.L., Obaid, R., Rolles, D., Rudenko, A., Bozek, J.D., Berrah, N.: The LAMP instrument at the Linac Coherent Light Source free-electron laser. Review of Scientific Instruments 89(3), 035112 (2018) https://doi.org/10.1063/1.5017727 . Accessed 2020-02-26
  • Seaberg et al. [2022] Seaberg, M., Lee, L., Morton, D., Cheng, X., Cryan, J., Curiel, G.I., Dix, B., Driver, T., Fox, K., Hardin, C., Kamalov, A., Li, K., Li, X., Lin, M.-F., Liu, Y., Montagne, T., Obaid, R., Sakdinawat, A., Stefan, P., Whitney, R., Wolf, T., Zhang, L., Fritz, D., Walter, P., Cocco, D., Ng, M.L.: The X-ray Focusing System at the Time-Resolved AMO Instrument. Synchrotron Radiation News, 1–9 (2022) https://doi.org/10.1080/08940886.2022.2066416 . Accessed 2023-10-16
  • Larsen et al. [2023] Larsen, K.A., Borne, K., Obaid, R., Kamalov, A., Liu, Y., Cheng, X., James, J., Driver, T., Li, K., Liu, Y., Sakdinawat, A., David, C., Wolf, T.J.A., Cryan, J.P., Walter, P., Lin, M.-F.: Compact single-shot soft X-ray photon spectrometer for free-electron laser diagnostics. Optics Express 31(22), 35822–35834 (2023) https://doi.org/10.1364/OE.502105 . Publisher: Optica Publishing Group. Accessed 2024-01-21
  • Krause et al. [1970] Krause, M.O., Stevie, F.A., Lewis, L.J., Carlson, T.A., Moddeman, W.E.: Multiple excitation of neon by photon and electron impact. Physics Letters A 31(2), 81–82 (1970) https://doi.org/10.1016/0375-9601(70)91039-X . Accessed 2024-03-19
  • Eppink and Parker [1997] Eppink, A.T.J.B., Parker, D.H.: Velocity map imaging of ions and electrons using electrostatic lenses: Application in photoelectron and photofragment ion imaging of molecular oxygen. Review of Scientific Instruments 68(9), 3477–3484 (1997) https://doi.org/10.1063/1.1148310 . Accessed 2023-10-13
  • Dörner et al. [2000] Dörner, R., Mergel, V., Jagutzki, O., Spielberger, L., Ullrich, J., Moshammer, R., Schmidt-Böcking, H.: Cold target recoil ion momentum spectroscopy: a ‘momentum microscope’to view atomic collision dynamics. Physics Reports 330(2-3), 95–192 (2000)
  • Griffiths et al. [1991] Griffiths, W.J., Correia, N., Keane, M.P., Brito, A.N., Svensson, S., Karlsson, L.: Doubly ionized states of N2O studied by photon-induced Auger electron and double charge transfer spectroscopies. Journal of Physics B: Atomic, Molecular and Optical Physics 24(19), 4187 (1991). Accessed 2013-11-02
  • Li et al. [2022] Li, S., Driver, T., Rosenberger, P., Champenois, E.G., Duris, J., Al-Haddad, A., Averbukh, V., Barnard, J.C.T., Berrah, N., Bostedt, C., Bucksbaum, P.H., Coffee, R., DiMauro, L.F., Fang, L., Garratt, D., Gatton, A., Guo, Z., Hartmann, G., Haxton, D., Helml, W., Huang, Z., LaForge, A.C., Kamalov, A., Knurr, J., Lin, M.-F., Lutman, A.A., MacArthur, J.P., Marangos, J.P., Nantel, M., Natan, A., Obaid, R., O’Neal, J.T., Shivaram, N.H., Schori, A., Walter, P., Wang, A.L., Wolf, T.J.A., Zhang, Z., Kling, M.F., Marinelli, A., Cryan, J.P.: Attosecond Coherent Electron Motion in Auger-Meitner Decay. Science 375(6578), 285–290 (2022) https://doi.org/10.1126/science.abj2096 . arXiv: 2105.08854. Accessed 2022-04-11
  • Barba et al. [2020] Barba, ., Bučar, K., Krušič, ., Žitnik, M.: Magnetic bottle electron spectrometer driven by electron pulses. Review of Scientific Instruments 91(7), 073108 (2020) https://doi.org/10.1063/5.0012523 . Accessed 2023-08-29
  • O’Neal [2022] O’Neal, J.T.: Tracking Charge Migration with Attosecond X-ray Pulses from a Free-Electron Laser. Doctoral, Stanford, Stanford, CA (December 2022)
  • Childers et al. [1977] Childers, D.G., Skinner, D.P., Kemerait, R.C.: The cepstrum: A guide to processing. Proceedings of the IEEE 65(10), 1428–1443 (1977) https://doi.org/10.1109/PROC.1977.10747 . Conference Name: Proceedings of the IEEE
  • Li et al. [2021] Li, S., Driver, T., Alexander, O., Cooper, B., Garratt, D., Marinelli, A., Cryan, J.P., Marangos, J.P.: Time-resolved pump–probe spectroscopy with spectral domain ghost imaging. Faraday Discussions, 10–1039000122 (2021) https://doi.org/10.1039/D0FD00122H . Accessed 2021-03-12
  • [42] Wang, J.: FDGI: A package for spectral-domain ghost imaging (https://pypi.org/project/FDGI/). https://github.com/congzlwag/spook Accessed 2023-10-16
  • Stellato et al. [2020] Stellato, B., Banjac, G., Goulart, P., Bemporad, A., Boyd, S.: OSQP: an operator splitting solver for quadratic programs. Mathematical Programming Computation 12(4), 637–672 (2020) https://doi.org/10.1007/s12532-020-00179-2
  • Belge et al. [1998] Belge, M., Kilmer, M.E., Miller, E.L.: Simultaneous multiple regularization parameter selection by means of the L-hypersurface with applications to linear inverse problems posed in the wavelet transform domain. In: Mohammad-Djafari, A. (ed.) San Diego, CA, pp. 328–336 (1998). https://doi.org/10.1117/12.323812 . http://proceedings.spiedigitallibrary.org/proceeding.aspx?articleid=958902 Accessed 2023-08-20
  • Belge et al. [2002] Belge, M., Kilmer, M.E., Miller, E.L.: Efficient determination of multiple regularization parameters in a generalized L-curve framework. Inverse Problems 18(4), 1161 (2002) https://doi.org/10.1088/0266-5611/18/4/314 . Accessed 2023-06-12