simulations of biomolecular fragmentation and diffraction...

86
ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2019 Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Science and Technology 1815 Simulations of Biomolecular Fragmentation and Diffraction with Ultrafast X-ray Lasers CHRISTOFER ÖSTLIN ISSN 1651-6214 ISBN 978-91-513-0669-8 urn:nbn:se:uu:diva-382441

Upload: others

Post on 13-Jan-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

ACTAUNIVERSITATIS

UPSALIENSISUPPSALA

2019

Digital Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Science and Technology 1815

Simulations of BiomolecularFragmentation and Diffraction withUltrafast X-ray Lasers

CHRISTOFER ÖSTLIN

ISSN 1651-6214ISBN 978-91-513-0669-8urn:nbn:se:uu:diva-382441

Page 2: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Dissertation presented at Uppsala University to be publicly examined in Häggsalen,Ångströmlaboratoriet, Lägerhyddsvägen 1, Uppsala, Friday, 14 June 2019 at 10:15 for thedegree of Doctor of Philosophy. The examination will be conducted in English. Facultyexaminer: Professor Helmut Grubmüller (Theoretical and Computational Biophysics, MaxPlanck Institute for Biophysical Chemistry, Göttingen, Germany).

AbstractÖstlin, C. 2019. Simulations of Biomolecular Fragmentation and Diffraction with UltrafastX-ray Lasers. Digital Comprehensive Summaries of Uppsala Dissertations from theFaculty of Science and Technology 1815. 84 pp. Uppsala: Acta Universitatis Upsaliensis.ISBN 978-91-513-0669-8.

Studies of biomolecules have recently seen substantial developments. New X-ray lasers allowfor high-resolution imaging of protein crystals too small for conventional X-ray crystallography.Even structures of single particles have been determined at lower resolutions with these newsources. The secret lies in the ultrashort high-intensity pulses, which allow for diffraction andretrieval of structural information before the sample gets fragmented. However, the attainableresolution is still limited, in particular when imaging non-crystalline samples, making furtheradvancements highly desired. In this thesis, some of the resolution-limiting obstacles facingsingle particle imaging (SPI) of proteins are studied in silico.

As the X-ray pulse interacts with injected single molecules, their spatial orientation isgenerally unknown. Recovering the orientation is essential to the structure determinationprocess, and currently nontrivial. Molecular dynamics simulations show that the Coulombexplosion due to intense X-ray ionization could provide information pertaining to the originalorientation. Used in conjunction with current methods, this would lead to an enhanced three-dimensional reconstruction of the protein.

Radiation damage and sample heterogeneity constitute considerable sources of noise in SPI.Pulse durations are presently not brief enough to circumvent damage, causing the sample todeteriorate during imaging, and the accuracy of the averaged diffraction pattern is impairedby structural variations. The extent of these effects were studied by molecular dynamics.Our findings suggest that radiation damage in terms of ionization and atomic displacementpromotes a gating mechanism, benefiting imaging with longer pulses. Because of this, sampleheterogeneity poses a greater challenge and efforts should be made to minimize its impact.

X-ray lasers generate pulses with a stochastic temporal distribution of photons, affectingthe achievable resolution on a pulse-to-pulse basis. Plasma simulations were performed toinvestigate how these fluctuations influence the damage dynamics and the diffraction signal.The results reveal that structural information is particularly well-preserved if the temporaldistribution is skewed such that most photons are concentrated at the beginning.

While many obstacles remain, the prospect of atomic-resolution SPI is drawing ever closer.This thesis is but one of the stepping stones necessary to get us there. Once we do, thepossibilities are limitless.

Keywords: X-ray free-electron laser, X-ray imaging, Single particle imaging, Computersimulation, Radiation damage, Molecular dynamics, Diffraction theory, Coulomb explosion,Sample heterogeneity, Diffractive noise, XFEL, SPI

Christofer Östlin, Department of Physics and Astronomy, Molecular and Condensed MatterPhysics, Box 516, Uppsala University, SE-751 20 Uppsala, Sweden.

© Christofer Östlin 2019

ISSN 1651-6214ISBN 978-91-513-0669-8urn:nbn:se:uu:diva-382441 (http://urn.kb.se/resolve?urn=urn:nbn:se:uu:diva-382441)

Page 3: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

In loving memory of my Dad.

Page 4: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by
Page 5: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

List of papers

This thesis is based on the following papers, which are referred to in the textby their Roman numerals.

I Reproducibility of Single Protein Explosions Induced by X-ray

Lasers

C. Östlin, N. Tîmneanu, H. O. Jönsson, T. Ekeberg, A. V. Martin andC. Caleman.Physical Chemistry Chemical Physics, 20, 12381–12389 (2018).

II Is Radiation Damage the Limiting Factor in Single Particle

Imaging with X-ray Free-Electron Lasers?

C. Östlin, N. Tîmneanu, C. Caleman and A. V. Martin.Submitted to Structural Dynamics.

III Sample Heterogeneity in Single Particle Imaging using X-ray

Lasers

C. Östlin, T. Mandl, M. Brodmerkel, E. G. Marklund, A. V. Martin, N.Tîmneanu, C. Caleman.Manuscript.

IV Simulations of Radiation Damage as a Function of the Temporal

Pulse Profile in Femtosecond X-ray Protein Crystallography

H. O. Jönsson, N. Tîmneanu, C. Östlin, H. A. Scott and C. Caleman.Journal of Synchrotron Radiation, 22, 256–266 (2015).

V FreeDam – A Webtool for Free-Electron Laser-Induced Damage in

Femtosecond X-ray Crystallography

H. O. Jönsson†, C. Östlin†, H. A. Scott, H. N. Chapman, S. J. Aplin, N.Tîmneanu, C. Caleman.High Energy Density Physics, 26, 93–98 (2018).

Reprints were made with permission from the publishers.

†These authors contributed equally to this work.

Page 6: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by
Page 7: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

List of additional papers

VI Auger Electron and Photoabsorption Spectra of Glycine in the

Vicinity of the Oxygen K-edge Measured with an X-FEL

A. Sanchez-Gonzalez, T. R. Barillot, R. J. Squibb, P. Kolorenc, M.Agåker, V. Averbukh, M. J. Bearpark, C. Bostedt, J. D. Bozek, S.Bruce, S. Carron Montero, R. N. Coffee, B. Cooper, J. P. Cryan, M.Dong, J. H. D. Eland, L. Fang, H. Fukuzawa, M. Guehr, M. Ilchen, A.S. Johnsson, C. Liekhus-S, A. Marinelli, T. Maxwell, K. Motomura, M.Mucke, A. Natan, T. Osipov, C. Östlin, M. Pernpointner, V. S. Petrovic,M. A. Robb, C. Såthe, E. R. Simpson, J. G. Underwood, M. Vacher, D.J. Walke, T. J. A. Wolf, V. Zhaunerchyk, J-E. Rubensson, N. Berrah, P.H. Bucksbaum, K. Ueda, R. Feifel, L. J. Frasinski, J. P. Marangos.Journal of Physics B: Atomic, Molecular and Optical Physics, 48,234004 (2015).

VII Ultrafast Dynamics of Water Exposed to XFEL Pulses

C. Caleman, H. O. Jönsson, C. Östlin, N. Tîmneanu.Submitted to Proceedings of SPIE.

Page 8: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by
Page 9: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Contents

1 Introduction 13

2 Background 17

2.1 Biomolecular imaging . . . . . . . . . . . . . . . . . . . . . . 17Protein structure . . . . . . . . . . . . . . . . . . . . . . . . 17X-ray crystallography . . . . . . . . . . . . . . . . . . . . . 19

2.2 Radiation damage . . . . . . . . . . . . . . . . . . . . . . . . 23Direct photon ionization . . . . . . . . . . . . . . . . . . . . 25Auger decay and electron collision . . . . . . . . . . . . . . 26Atomic displacement . . . . . . . . . . . . . . . . . . . . . 27Long term damage processes . . . . . . . . . . . . . . . . . 27

2.3 Diffraction before destruction . . . . . . . . . . . . . . . . . . 28X-ray free-electron lasers . . . . . . . . . . . . . . . . . . . 29Serial femtosecond crystallography . . . . . . . . . . . . . . 31

2.4 Single particle imaging . . . . . . . . . . . . . . . . . . . . . 31The orientation problem . . . . . . . . . . . . . . . . . . . . 34Coulomb explosion . . . . . . . . . . . . . . . . . . . . . . 34Sample heterogeneity . . . . . . . . . . . . . . . . . . . . . 36Pulse profile . . . . . . . . . . . . . . . . . . . . . . . . . . 37

3 Methods 38

3.1 Molecular dynamics . . . . . . . . . . . . . . . . . . . . . . . 38Force fields . . . . . . . . . . . . . . . . . . . . . . . . . . 40Water models . . . . . . . . . . . . . . . . . . . . . . . . . 42GROMACS and XMD . . . . . . . . . . . . . . . . . . . . 43Ion mapping . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.2 Calculating diffraction patterns . . . . . . . . . . . . . . . . . 45Radial profile and SNR . . . . . . . . . . . . . . . . . . . . 47Pearson correlation . . . . . . . . . . . . . . . . . . . . . . 49

3.3 Non-LTE plasma simulations . . . . . . . . . . . . . . . . . . 51

4 Results and conclusions 53

4.1 Reproducibility of Coulomb explosions (Paper I) . . . . . . . . 534.2 Radiation damage in SPI (Paper II) . . . . . . . . . . . . . . . 574.3 Sample heterogeneity in SPI (Paper III) . . . . . . . . . . . . . 614.4 Pulse profile effects in SFX (Paper IV) . . . . . . . . . . . . . 644.5 The FreeDam database (Paper V) . . . . . . . . . . . . . . . . 66

Page 10: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

5 Outlook 67

6 Acknowledgements 69

7 Sammanfattning på svenska 71

8 References 74

Included papers

Paper I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85Paper II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95Paper III . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106Paper IV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113Paper V . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

Author contributionsThe papers presented in this thesis are the result of collaborative work betweena number of people and should be recognized as such. However, to clarify mypersonal contributions to the publications, I provide the following list:

Paper I–III: Main author and contributor – coordinated the project, per-formed simulations, designed and carried out analyses, led discussions.

Paper IV: Participated in discussions, proofreading and writing.

Paper V: Shared main author – designed and implemented the onlineGUI, participated in discussions.

Page 11: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Abbreviations

AGIPD Adaptive gain integrating pixel detectorAMBER Assisted model building with energy refinement force fieldAMO Beamline for atomic, molecular and optical scienceCCD Charge coupled deviceCHARMM Chemistry at Harvard macromolecular mechanics force fieldCryo-EM Cryogenic electron microscopyCSPAD Cornell-SLAC pixel-array detectorCXI Beamline for coherent X-ray imagingCXIDB Coherent X-ray imaging data bankEMC Expectation maximization compressionFLASH Free-electron laser in HamburgFRC Fourier ring correlationFWHM Full-width at half maximumGDVN Gas-dynamic virtual nozzleGROMACS Groningen machine for chemical simulationsHHG High-harmonic generationHIO Hybrid input-outputLAMMPS Large-scale atomic/molecular massively parallel simulatorLCLS Linac coherent light sourceMCP Microchannel plateNAMD Nanoscale molecular dynamicsNIST National Institute of Standards and TechnologyNon-LTE Non-local thermodynamic equilibriumNMR Nuclear magnetic resonanceOPLS Optimized potentials for liquid simulations force fieldPAL-XFEL Pohang accelerator laboratory X-ray free-electron laserPDB Protein data bankRMSD/RMSF Root-mean-square deviation/fluctuationSACLA Spring-8 Angstrom compact free-electron laserSASE Self-amplified spontaneous emissionSFX Serial femtosecond crystallographySHINE Shanghai high repetition rate XFEL and extreme light facilitySNR Signal-to-noise ratioSPI Single particle imagingSwissFEL Swiss free-electron laserTIP3P/TIP4P Transferable intermolecular potential with 3/4 pointsVMI Velocity map imagingXFEL X-ray free-electron laser

Page 12: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by
Page 13: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

1. Introduction

Observing is a fundamental cornerstone of science. By simply looking, wecan obtain information crucial to understanding the world around us. In someinstances this is a fairly easy task; a pair of healthy eyes have little difficultiesdistinguishing the prominent features of a plant or an animal. However, thismacroscopic view is limited and seldom enough to fully explain the underlyingmechanisms of the observed system. Seeing the green leaves on a maple treemight tell you that the organism is capable of photosynthesizing, but leavesyou unable to explain how photosynthesis happens. Watching the musculatureof a gazelle provides insight into its remarkable running ability, but how itsmuscles actually contract remains a mystery. There is clearly a need to gobeyond the imposed limits of human eyesight in order to accurately discernthe inner workings of nature.

To enter the microscopic world we have to rely on more advanced imag-ing techniques. These have their origins in the field of microscopy, whichoriginally emerged with the advent of modern optical microscopes in the 17thcentury. Since then, various novel approaches have been developed, and theoriginal method refined, to enhance the attainable resolution of the imagedsample. Following the discovery of X-rays, the last century saw the inventionof several X-ray based imaging techniques that offer remarkably high magnifi-cation powers. The perhaps most well-known is that of X-ray crystallography,which is capable of imaging the individual atoms in a sample (referred to asatomic resolution).

Conventional X-ray crystallography relies on coherent diffraction of X-raysfrom a crystallized sample. Diffraction patterns are collected continuouslywhile the sample is being rotated, essentially producing images of the samplefrom all angles. The ordered and repeating structure of a crystal enhancesthe signal and makes it more resilient to the harmful radiation. Both of theseeffects scale with the number of unit cells, the smallest repeating element ofthe crystal. This means that there is an inherent minimal size requirementfor imaging to be viable. Unfortunately, acquiring crystals of appropriate sizeis not always possible. For instance, a group of highly-promising drug targetsknown as membrane proteins are notoriously difficult to crystallize. Therefore,in order to determine their structure, an alternative method must be employed.

In their pioneering work from 2000, Neutze et al. [1] predicted that usinghigh-intensity pulses on the femtosecond time scale produced by X-ray free-electron lasers (XFELs) [2] would enable imaging of non-crystalline samples.In this approach each X-ray pulse is intense enough (on the order of 1012 pho-tons) to counteract the loss in signal compared to a crystal, but at the same

13

Page 14: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

time brief enough to outrun damage processes. Following the exposure, thesample is heavily ionized and quickly disintegrates in a violent Coulomb ex-plosion, but only after diffraction data has been collected. The concept cameto be known as "diffraction before destruction" and was supported by furthertheoretical studies [3, 4].

In 2006, diffraction before destruction was confirmed experimentally withsoft X-rays (32 nm wavelength) at the FLASH free-electron laser in Ham-burg, Germany [5]. Three years later, the first XFEL matching the proposedrequirements to reach atomic resolution, the Linac Coherent Light Source(LCLS) [6], became available. Within the first two years of operation, imag-ing of biological samples unfit for conventional X-ray crystallography wereattempted. Chapman et al. managed to assemble diffraction patterns into athree-dimensional volume from nanometer-sized photosystem I crystals, andshowed that it was comparable to known data obtained conventionally [7].Seibert et al. collected diffraction patterns from single virus particles and re-constructed two-dimensional images to 32 nm resolution [8]. The latter waslater expanded to three dimensions by Ekeberg et al., albeit at a lower resolu-tion of 125 nm [9]. Other non-crystalline, single particles were also imagedusing the same principle, such as RNAi microsponges [10], organelles [11]and entire cells [12]. Each study utilized the same basic methodology; as thesamples were destroyed upon diffracting they were continuously replaced ina serial fashion. This allowed for the accumulation of a large set of diffrac-tion patterns from different orientations, which could then be combined toultimately retrieve the three-dimensional structure.

The approach has since become increasingly standardized. When appliedto crystalline samples it has been named serial femtosecond X-ray crystal-lography (SFX) and has been used to determine several biological structures[13], up to resolutions as high as 1.20 Å [14]. However, in the case of non-crystalline single particles – referred to as single particle imaging (SPI) – thesuccess has been less pronounced. Resolutions of reconstructed images havebeen limited to tens of nanometers [15, 16], which is far from the desiredatomic resolution of ~1.5 Å.

In this thesis we aim to address some of the obstacles holding SPI back.Chapter 2 provides the necessary background to understand SPI and how itcompares to conventional X-ray crystallography. It also explains the physicalprocesses involved and how they are related to diffraction. Chapter 3 describesthe computational and mathematical methods employed in the presented stud-ies. Lastly, Chapter 4 summarizes the results of those studies, where the fol-lowing challenges in SPI were treated.

The orientation problem (Paper I)

Typically, the single particles are aerosolized with a gas-dynamic virtual noz-zle (GDVN) [17] and delivered to the XFEL beam via an aerodynamic lensstack [18] or a convergent nozzle [19]. As a result, the spatial orientation of

14

Page 15: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

the sample at the time of X-ray interaction is random and unknown. Findingthe orientation is nontrivial but crucial to the successful three-dimensional as-sembly of diffraction patterns, and one of the major challenges of SPI. Thesolution so far has been to orient the diffraction patterns algorithmically [20].Another potential approach is to control the orientation pre-pulse using elec-tric fields [21]. In Chapter 4.1, we investigate and discuss a third option: thepossibility of retrieving orientation data by mapping the ion trajectories of theCoulomb explosion.

Radiation damage (Paper II, V)

Despite the ultrashort pulse durations offered by the XFELs in operation to-day, radiation induced damage cannot be circumvented entirely. Ionizationleads to diffuse scattering, while the following structural decay modifies thesample. Such effects bring about noise in the measured diffraction pattern.However, some studies suggest that these processes may also promote a gat-ing effect such that the effective pulse duration becomes artificially shortened,for nanocrystals [22] as well as for single particles [23]. In Chapter 4.2, weexplore if a similar effect is observed when considering the Coulomb explo-sion and how the damage noise compares to other sources of noise. Moreover,Chapter 4.5 presents a database containing radiation damage data for varioussamples.

Sample heterogeneity (Paper III)

In SPI, each diffraction pattern is collected from a new sample. Ideally theyare all identical, but this is unlikely to be true in most cases. Especially bi-ological molecules, which we are concerned with here, are known to exhibitthermal variability, and some are even associated with ensembles of possiblefunctional conformations. The structural heterogeneity leads to a loss of co-herence as diffraction patterns are averaged to improve the signal quality andhas profound negative impact on the achievable resolution [24, 25]. Chapter4.3 explores how the signal loss due to sample heterogeneity relates to tem-perature and sample hydration.

Pulse profile (Paper IV)

Current XFEL-facilities generate pulses through self-amplified spontaneousemission (SASE), a process that is stochastic in nature. Consequently, thetime-distribution of photons is generally unknown and may fluctuate greatlybetween pulses. The temporal profile of the X-ray pulse affects the evolutionof damage and therefore has a direct impact on the imaging quality. However,the exact relationship between pulse profile and imaging is not well known.In Chapter 4.4 we delve deeper into this issue, which becomes increasinglyrelevant as techniques for single-shot pulse characterization [26,27] and shap-ing [28] are being developed.

15

Page 16: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

The aim of this work is to provide some results, ideas and insights in thepursuit of atomic resolution X-ray imaging of single particles. And with itsvast potential applicability in the fields of materials science and structural bi-ology, it is a worthwhile pursuit. The challenges are extensive, but despitethese hardships we are convinced that, once realized, SPI will have a tremen-dous impact. Just like X-ray crystallography before it, SPI promises to classifyas one of the most important scientific breakthroughs of its time.

16

Page 17: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

2. Background

In 1901, German physicist Wilhelm Conrad Röntgen was awarded the veryfirst Nobel prize in physics for his discovery of X-rays [29]. It was a break-through that would provide the stepping stone for countless advancements invarious scientific fields. While the medical imaging applications were im-mediately apparent, the properties of the novel radiation would also open upa new world of possibilities. The following years, methods utilizing X-rayswere developed and enabled unprecedented studies of matter at the submolec-ular level. For the first time ever, scientists could reveal atomic-level detailsabout solids, liquids, gases and even specific molecules. A multitude of addi-tional Nobel prizes were issued during the 20th century in connection to thesemethods and findings attributed to them (see for instance [30], [31] or [32]),and their use remain highly relevant even today.

2.1 Biomolecular imagingWhile X-ray based methods can be applied to virtually any sample, in-depthinvestigation of substances pertaining to life are of particular interest. Labeledas biomolecules, this broad category encompasses all molecules found in or-ganisms that are involved in life-essential processes or functions. It is com-monly divided into four subgroups of macromolecules: lipids, carbohydrates,nucleic acids and proteins. Some examples are the phospholipids that con-stitute the protective membrane around eukaryotic cells, the polysaccharidesthat store energy and fuel cellular activity, and the DNA that carry genetic in-formation. But the arguably most intriguing of biomolecules are the proteins.Despite consisting of a mere 20 different building blocks known as aminoacids, proteins display an impressive range of varying biological functions.Inside living organisms they have a crucial role in many important cellular pro-cesses, such as ion transport [33], immune response [34], cell signaling [35],mitosis [36], gene expression [37], and more. Clearly, understanding proteinsis vital to understanding life.

2.1.1 Protein structureWhat gives rise to this immense diversity? Much like how a spoon is moresuitable for eating soup than a fork is – despite both being made from the

17

Page 18: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

same material – the answer lies in their design. It is the practically endlessnumber of unique conformations that is the secret behind the many differentfunctions of proteins. Their structures are divided into four distinct levels,visualized in Figure 2.1, ranging from quaternary to primary. Typically, aprotein consist of a single chain of covalently bonded amino acids (a polypep-tide). Some proteins, however, form aggregates of multiple polypeptide chainsthat act in unison as a single entity. Each polypeptide chain in such complexis referred to as a subunit, and the global three-dimensional conformation iscalled the quaternary structure. This comprises the highest structural level.For an isolated subunit, or a single-chained protein, the global conformationis instead described by the tertiary structure. It is formed by a combinationof non-specific hydrophobic interactions and specific interactions such as hy-drogen bonds, disulfide bonds and salt bridges within the chain. This is thelevel that is typically implied when considering protein folding and the mostprevalent in regard to biological function.

Along the polypeptide chain we may find local structural elements (suchas spiraling alpha helices or pleated beta sheets) that define the secondarystructure. They are formed spontaneously during polymerization through anetwork of hydrogen bonds between amino acid residues, and aid the foldinginto the tertiary structure from the initial random coil. These are, in turn,highly correlated to the sequence of amino acids; the protein primary structure.

Figure 2.1. The four levels of protein structure. (a) Quaternary structure of proteincomplex with four subunits. Two ribonuclease I molecules (green), each interactingwith a human ribonuclease inhibitor (grey). (b) Tertiary structure of ribonucleaseI. The global shape shows both alpha helices (green) and beta sheets (purple). (c)

Secondary structure element. An alpha helix formed by 14 amino acid residues withsequence (from top to bottom) Met-Lys-Glu-Ser-Arg-Ala-Lys-Lys-Phe-Gln-Arg-Gln-His-Met. (d) Primary structure segment. The two amino acids methionine (top) andhistidine (bottom) sharing a covalent peptide bond (arrow). Colors indicate differentelements: grey, C; red, O; blue, N and yellow, S.

Because of their intimate connection to life, proteins have become espe-cially interesting to life sciences and drug discovery, and determining thestructure is an integral part of that process. This has led to the emergence of thefield of structural biology, and to the establishment of the Protein Data Bank(PDB) [38] where biomolecular structures are stored. Finding the structure

18

Page 19: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

is not an easy task, however. A hypothesis known as Anfinsen’s dogma [39]states that in the native physiological environment, the primary structure willuniquely determine the full three-dimensional configuration of the protein. Inspite of this, it is as of yet impossible to deduce the tertiary structure with suffi-cient accuracy based on the peptide sequence alone. Therefore, other methodsneed to be employed in order to establish the detailed structure of the moleculeand consequently understand its underlying mechanisms.

2.1.2 X-ray crystallographyThere are currently three predominant techniques in use for high-resolutionstructure determination: nuclear magnetic resonance (NMR) spectroscopy [40],cryogenic electron microscopy (cryo-EM) [41] and X-ray crystallography [42].NMR spectroscopy employs magnetic fields to exploit the quantum mechani-cal properties of the nuclei within the molecule in order to retrieve data aboutinteratomic distances, movement and bonding. The information can then becompiled to form a three-dimensional model of the sample. This method holdsthe advantage of being able to study proteins in solution, which can be sim-ilar or identical to its native environment. On the other hand, because NMRspectra become increasingly hard to interpret with molecular size, the tech-nique is generally limited to smaller proteins. Advancements have been madeto counter this drawback, but the upper limit of single domain monomericproteins currently remains at ~80 kDa [43], which roughly corresponds to amedium-sized protein with 700 amino acid residues.

An alternative to NMR spectroscopy that has gained considerable popular-ity in recent years is cryo-EM. By scattering electrons from frozen proteinsamples and focusing them onto a detector using an electromagnetic lens,structural data can be obtained. With recent advancements in detector tech-nology and reconstruction algorithms, cryo-EM has managed to reach near-atomic resolution imaging of important single proteins such as ribosomes,membrane proteins and hemoglobin [44–46]. As opposed to NMR spec-troscopy, this method is not limited by large samples. Instead, the biggestchallenge currently facing developers of cryo-EM is to increase the achievableresolution when the samples are particularly small. Another drawback is theact of freezing, which renders direct studies of protein dynamics impossible.

The third technique for probing the three-dimensional structure of proteinsis X-ray crystallography, which so far is the most widely used. As of May2019, about 90 % of the 140,000 protein structures deposited into the PDBhave been solved by X-ray crystallography and the number keeps growingby around 10,000 annually. Clearly, the importance of this method cannot beoverstated, and the mechanisms behind it (described below) constitute the veryfoundation for the studies presented in this thesis.

19

Page 20: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

In conventional X-ray crystallography, a three-dimensional crystalline sam-ple (see example in Figure 2.2a) is illuminated by a beam of monochromaticand coherent X-rays. The incoming photons interact with the electrons withinthe crystal and are reemitted without loss of energy – a process known as elas-tic scattering. While other processes may take place (see Section 2.2), elasticscattering constitute the basis for crystallography. The scattered photons canbe viewed as spherical waves with well-defined spatial origins correspondingto atomic positions, which are repeating due of the periodic structure of thecrystal. Waves originating from repeating elements separated by distance d,often referred to as the lattice spacing, will interfere constructively in specificdirections given by Bragg’s law [47]

2d sinθ = nλ , (2.1)

where θ is the incident angle of radiation relative to the crystal planes, n is apositive integer and λ the wavelength of the X-rays. The result is a diffrac-tion pattern with sharp, high-intensity spots called Bragg peaks (or sometimesreflections) that can be measured. Knowing the position of these peaks andtheir intensities it becomes possible to deduce the electron density, and ulti-mately the structure, behind the scattering material. Figure 2.2b below depictsa schematic visualization of Bragg’s law.

Figure 2.2. Protein crystal and Bragg’s law. (a) Crystalline sample of a malic en-zyme (image courtesy of NASA). Zoomed in is a conceptual illustration of a crystallattice at the atomic scale, showing the periodic arrangement of atoms. (b) Elasticscattering from two parallel crystallographic planes separated by distance d. The co-herent incoming X-rays (purple arrows) of wavelength λ interact with the atoms ineach plane at an incident angle θ and change direction without loss of energy. Theextra distance traversed by the lower beam is 2d sinθ , so the outgoing waves will re-main in phase and add constructively if this distance is equal to an integer multiple ofλ , as stated by Bragg’s law.

Similarly to a multiple slit diffraction setup, the contribution to construc-tive interference will grow with the amount of equivalent scatterers. Hence,by increasing the number of crystal lattice planes the intensity of Bragg peakswill increase while their size narrows. In the case of three-dimensional pro-tein crystals, the intensity scales with the number of unit cells squared. Since

20

Page 21: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

intensities drop with scattering angle [48], the enhanced intensity of a largercrystal allows for diffraction up to higher angles θ , which translates to higherresolution imaging. This follows from the fact that the 2θ change in directionbetween incoming and outgoing photons correspond to the magnitude of themomentum transfer |q| = q = 2π/d between the two. The inverse relation-ship implies that shorter interplanar distances require more momentum to betransferred. Inserting this into Equation 2.1 for the lowest order n = 1 gives

q =4πλ

sinθ , (2.2)

meaning that the momentum transfer ranges from qmin = 0 to qmax = 4π/λwith maximum at θmax = 90◦. In real space representation, these limits mapto d = ∞ and d = λ/2, respectively, so the highest resolution informationtheoretically obtainable through Bragg diffraction are on the length scales ofhalf the incoming wavelength. This sets an upper constraint on the chosenwavelength to 2− 3 Å, provided that atomic resolution is desired. However,the resolution is more restricted in practice and it is common to use photonenergies on the order of 10 keV (λ = 1.24 Å).

As a consequence of the relations between scattering angle θ , momentumtransfer q and spatial coherence length d, the measured diffraction pattern is areciprocal space representation of the crystal’s periodicity. If a photon detectoris placed centrally downstream of the sample (measuring forward scattering,see Figure 2.3a) in relation to the incoming X-ray beam, Bragg peaks encodingfor shorter length scales will appear further from the detector center.

The diffraction pattern from a stationary crystal subjected to an incomingX-ray beam of constant incident angle will only give two-dimensional infor-mation about the electron density. This can be compared to how a regular pho-tograph only shows a two-dimensional projection of three-dimensional space.In order to determine the three-dimensional structure of the sample, diffrac-tion patterns from different angles must be collected. This is usually doneby rotating the crystal during exposure while simultaneously measuring theBragg peaks. Once a full dataset has been recorded, the individual patternsare assembled into a three-dimensional volume in reciprocal space, which ismathematically linked to the electron density by the (inverse) Fourier trans-form.

It is worth noting that the structure retrieval process, as described here, isimmensely simplified. For instance, scattering from a sample is described byits structure factor

F(q) = |F(q)|eiφ(q), (2.3)

a mathematical complex function that includes both the amplitude |F(q)| andphase φ(q) of the scattered wave in reciprocal space. F(q) is the three-dimensional Fourier transform of the real space electron density, so in order todetermine the structure of the sample we first need to find this function. How-

21

Page 22: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

ever, when measuring diffraction patterns we are limited to detecting intensi-ties, which are proportional to the square modulus of the scattered amplitude,|F(q)|2. Information about the phase, on the other hand, is lost and needs tobe recovered before we can accurately reconstruct the electron density. This isknown as the phase problem and constitutes a considerable challenge in struc-ture determination. One way of solving the phase problem is by estimatingan initial set of phases, for example using direct methods [49], and iterativelyrefining the calculated structure factor until it corresponds well to the observeddata.

Figure 2.3. Protein X-ray crystallography workflow. (a) An incoming beam ofmonochromatic and coherent X-rays (represented by the wave vector k) is directed ata crystallized protein sample. The outgoing waves (wave vector k′) produce a diffrac-tion pattern that is sampled on the Ewald sphere with a photon detector. Positionsof the Bragg peaks are related to the momentum transfer q = k′ −k, the magnitudeof which is inversely proportional to spatial coherence in the sample. Hence, higherresolution structural information is encoded in spots closer to the detector edge (asindicated by the black arrow). The crystal is rotated during exposure to allow forthe collection of diffraction patterns from different angles. (b) The collected patternsare combined to form a three-dimensional image of the reciprocal space. Applyingthe inverse three-dimensional Fourier transform allows for the reconstruction of theelectron density, from which a structural model can be fitted.

Another simplification is the assembly of the three-dimensional diffractionvolume as depicted in Figure 2.3b. In X-ray crystallography, we are probingthe reciprocal lattice that represents the crystal. Elastic scattering results inBragg peaks at lattice points coinciding with the surface of a sphere centeredat the sample origin with a radius equal to the length of the wave vectors k andk′. This is known as the Ewald sphere [50]. A diffraction pattern will thereforerepresent the two-dimensional slice in reciprocal space defined by the sphere.But while the Ewald sphere has a curvature, our detectors are flat (as shown inthe figure), meaning that the recorded image is actually a gnomonic projectionof the spherical slice. The approximation becomes less accurate with smallersphere radius, i.e. shorter wavelengths, and at higher scattering angles, andmay need to be corrected for when assembling the three-dimensional volume.

22

Page 23: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

The aforementioned aspects of phase retrieval and Ewald sampling in X-raycrystallography are important and involve challenges that are non-trivial. Re-gardless, they are less relevant to the scope of this thesis, where the presentedsimplifications will suffice.

The success of X-ray crystallography is highly dependent on the qualityand, as previously mentioned, the size of the sample crystal. Holton andFrankel [51] estimated that a spherical lysozyme crystal on the order of mi-crometers in diameter is needed to obtain a complete dataset for structure de-termination up to 2 Å resolution. An even larger crystal is required when theunit cell is bigger than that of lysozyme. The size-dependence is troublesomesince a significant number of proteins are difficult or even impossible to crys-tallize, a fact that holds especially true for the subgroup of membrane proteins.It has been estimated that 20−30 % of all proteins in organisms are embeddedin the cell membrane [52], which makes them promising candidates as drugtargets. In fact, approximately half of all known targets for therapeutic drugsbelong to this subgroup [53]. But due to their limited ability to crystallize –often yielding nanocrystals at best – membrane protein structures are severelyunderrepresented in the PDB. As of 2019, less than 4 % of all proteins in thedatabase belong to this group.

The advantages of a large crystal is twofold. Apart from the repetitive ele-ments contributing to a stronger Bragg signal, a bigger crystal is also less sen-sitive to the radiation damage induced by the X-rays. The damage processesgive rise to noise in the diffraction pattern, but as the effects are stochasticallyspread over many copies of the protein, the noise is averaged and drownedout by the signal. For smaller crystals, however, damage quickly becomesproblematic. With the combined issues of weaker signal and non-negligiblenoise, the diffraction conditions become unfavorable and structure determina-tion compromised. In the next section we will discuss the relevant physicalmechanisms behind X-ray–matter interaction and how they relate to radiationdamage.

2.2 Radiation damageA sample, such as a protein crystal, will deteriorate over time when exposed toX-rays. This is a result of the radiation damage induced by the photons as theyinteract with the matter. In X-ray crystallography, these interactions happenprimarily with electrons through three distinct mechanisms1. The three pro-cesses are shown visually in Figure 2.4. We have previously discussed elasticscattering, the process that enables imaging by giving rise to the diffractionpatterns containing structural information. When elastically scattered, a pho-

1For a full description of processes omitted here the reader is referred to Hau-Riege, S. P.,"High-Intensity X-Rays – Interaction with Matter: Processes in Plasmas, Clusters, Moleculesand Solids", John Wiley & Sons, 2012 (ref. [54]).

23

Page 24: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

ton simply changes direction by transferring momentum to the electron, with-out losing energy. Since the sample is left unaltered, this process is not asource of radiation damage.

Figure 2.4. X-ray interaction with matter. A simple schematic showing the threepredominant mechanisms by which X-ray photons interact with electrons in matter.Photons are visualized as purple arrows with energies indicated by their oscillatingfrequencies, black horizontal lines show electron energy levels and the circles de-pict electrons (black) and holes (white). Elastic scattering (left): the photon retainsits energy but undergoes a change in direction of propagation through a transfer ofmomentum. Inelastic scattering (center): the energy of the incoming photon is par-tially transferred to the atom, which becomes excited or ionized. The leftover energyis emitted as a secondary photon. Note that this process is predominant for looselybound or quasi-free electrons. Photoelectric effect (right): the photon is absorbedcompletely by an electron. The acquired energy is enough to eject the electron fromits bound state, leaving a hole behind.

However, a photon may also inelastically scatter. In such event, the incom-ing photon transfers part of its energy to an electron and the leftover energyis incoherently emitted as a secondary photon. In this case the photon stillundergoes a change in direction, but the system becomes excited as a productof the energy transfer. If the transfer is sufficient enough to overcome the elec-tron binding energy, the involved atom might even become ionized. Due to theincoherence of the outgoing waves, inelastic scattering does not contribute tothe constructive interference forming Bragg peaks but instead generates noisein the diffraction pattern.

Lastly, the photon may not scatter at all, and instead be completely absorbedby an electron causing it to eject from the atom. This ionizing process isknown as the photoelectric effect and was first explained by Einstein in 1905[55]. The probabilities of the three processes taking place, described by theircross sections, depend on the energy of the photon and the type of elementit interacts with. For carbon in the context of X-ray crystallography wherephoton energies on the order of ~10 keV are used, the photoelectric effect isthe dominant process. At this energy, the cross section is over one order ofmagnitude higher for photoabsorption than it is for either type of scattering(see Figure 2.5).

24

Page 25: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Figure 2.5. Photon interaction cross sections for carbon. The photoelectric effect(dotted line) is dominating for photon energies around 10 keV relevant to X-ray crys-tallography with a cross section more than one order of magnitude higher than theelastic (solid line) and inelastic (dashed line) scattering counterparts. Data retrievedfrom the photon cross section database XCOM by the National Institute of Standardsand Technology (NIST) [56].

Ionization is the predominant antagonist of X-ray imaging. As we shall seebelow, the release of electrons leads to a host of effects that affect both signalstrength and noise levels.

2.2.1 Direct photon ionizationWhile inelastic scattering contributes to ionization of the sample, it is the pho-toelectric effect that represents the main driver of damage. Being the dominantprocess at the relevant energies, electrons are stripped from the constituentatoms of the crystal on time scales of ~100 attoseconds [57]. This is problem-atic since the free electrons alter the electron density, while simultaneouslycontributing to noise in the diffraction pattern. The change in electron density,and consequently in the structure factor that relates the diffraction pattern tothe structure, blurs the Bragg peaks and causes detrimental artefacts. On top ofthat, the released electrons give rise to diffuse scattering, which further masksthe signal and inhibits the obtainable resolution.

An electron ejected by the photoelectric effect, also known as a photoelec-tron, will have a kinetic energy equal to the difference between the incomingphoton energy and the ionization energy. Due to the energy conservation, thephotoelectrons generated by 10 keV photons in a protein crystal consistingmostly of light elements will be highly energetic. In fact, this is true evenfor the most tightly bound core electrons of the 1s inner shell of all relevant

25

Page 26: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

elements. For instance, a 1s electron of carbon has a binding energy of ap-proximately 290 eV. A photoabsorption of a 10 keV photon would thereforeyield a 9.7 keV photoelectron. Unless the electron escapes, all of this energywill be deposited into the sample, inducing further damage (see below). Leftbehind is an ion with an inner-shell vacancy referred to as a core hole. Both ofthese factors will promote additional secondary ionizations through the Augereffect and electron collisions.

2.2.2 Auger decay and electron collisionAn ion with a core hole generated by the photoelectric effect is in an excitedstate. Within a few femtoseconds of being created, the ion will relax by fillingthe vacancy with an electron from a higher-energy orbital. In doing so the en-ergy of the system is decreased, and the energy gain is compensated for in oneof two ways. Either a secondary photon is emitted corresponding to the energydifference, or another electron is released from binding (see Figure 2.6a). Theformer process (fluorescence) is more common for heavy elements and gener-ates incoherent noise in the measurement similarly to inelastic scattering. Thelatter, which is favored in light atoms, further ionizes the sample and is knownas Auger decay. Despite being independently reported by Austrian-Swedishphysicist Lise Meitner a year prior [58], it is named after French physicistPierre Victor Auger who discovered it in 1923 [59]. Electrons released thisway are referred to as Auger electrons.

So far we have discussed ionizations induced by photons, but electrons canproduce similar effects. If an electron with enough energy collides with anatom, it may liberate another electron by transfer of energy (see Figure 2.6b).All free electrons generated by the photoelectric effect and Auger decay willtherefore propagate further in the sample, releasing additional electrons alongthe way. With each collision the energy is distributed between the two outgo-ing electrons, so the cascade is eventually terminated as the electrons becomethermalized. This happens on a 1−10 fs time scale [60]. Cascades initiated byphotoelectrons are particularly damaging since they carry a lot of energy, andit has been estimated that a single photoelectron may result in up to ~500 moreionization events [61]. However, this number can be significantly decreased ifthe sample is small enough and the electrons escape. Simulations have shownthat the electron cloud generated from a photoelectron cascade in a biomolec-ular crystal have a spatial spread on the order of ~500 nm, so for samples ofsmaller dimensions many of the electrons will initially leave the sample in-stead of causing further ionizations [62]. Although it should be noted thatdamage generally cannot be fully avoided this way, as electrons eventuallybecome electrostatically trapped by the resulting net positive charge [63].

26

Page 27: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Figure 2.6. Secondary atomic processes. Simple Bohr model of an atom with twodistinct energy levels illustrating secondary mechanisms contributing to damage andnoise. (a) A core hole is filled by an electron from a higher energy level, resulting inemission of a photon (fluorescence) or an electron (Auger decay). (b) A free electroncollides with the atom and liberates one of its electrons by transferral of energy.

2.2.3 Atomic displacementThe formation of ions is the next side effect that is potentially detrimental todiffractive imaging. When electrons are removed, the positively charged ionswill exhibit repulsive Coulomb forces on each other. This triggers movementof the nuclei that cause the molecular structure to gradually degrade. As thecrystal periodicity is disrupted the signal deteriorates and is eventually lost,starting at high scattering angles. The time scales at which atomic displace-ment manifests is in the 10− 100 fs range and the effects can be marginallydelayed by cryogenic cooling [64]. While displacement happens at a slowerrate than ionization and electron cascades, it is still relevant in the context ofX-ray crystallography. A large crystal is less sensitive since the damage isspread over the unit cells, allowing for doses up to 30 MGy if the sample iscooled [65], so the diffraction pattern will not be significantly impacted. Butin cases where the crystal size is limited the loss of spatial coherence quicklybecomes substantial, especially if the incoming photon flux is high.

2.2.4 Long term damage processesRadiation damage also extends to longer time scales. Heavy ionization ofthe sample will lead to the formation of radicals that diffuse within the crys-tal. Due to their unpaired valence electron, radicals are highly reactive andwill therefore interact with the surrounding molecules and further degrade thesample. Free electrons may also recombine with ions and thereby alter thestructure factor. Lastly, the deposited energy leads to an inevitable increase intemperature. Proteins are often sensitive to heating and are likely to undergodenaturation, a process that disrupt their structure. All these effects first be-come relevant in the regime of picoseconds or longer [61]; considerably laterthan previously discussed processes.

27

Page 28: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

In summary: radiation damage in a crystal illuminated by X-rays is the prod-uct of primarily two processes, inelastic scattering and the photoelectric effect.The result is a combination of incoherently scattered photons (from inelasticscattering and fluorescence) that reach the detector and mask the desired elas-tic scattering signal, as well as the release of energetic electrons. Secondaryionizations follow through Auger decay and electron collisions, generatingeven more free electrons. These contribute to noise in the form of diffuse X-ray scattering and simultaneously suppress the signal as the electron density ischanged. Energetic electrons propagate though the sample, and the positivelycharged ions left behind start to repel as a consequence of the Coulomb forces.The structural degradation leads to further loss of coherence causing the signallevels to drop.

Given long enough exposure time, damage will manifest to the extent wherenoise overtakes signal, at which point structural data collection no longeris possible. Luckily, the resilience of large and cryocooled crystals effec-tively keeps noise levels down during the time needed to acquire a full three-dimensional dataset. This is not true in the case of small crystals, especiallysince the beam intensity needs to be increased to counteract the weaker contri-butions to constructive interference from elastic scattering. For this reason,when attempting to image more radiation sensitive samples, managing thedamage effects is paramount. One such approach is to use ultrashort X-raypulses, which we will discuss next.

2.3 Diffraction before destructionSince radiation damage in a crystal evolves over time, the effects can be lim-ited by reducing the time of X-ray exposure. A usable diffraction pattern canstill be obtained, provided that the photon flux is increased accordingly tocompensate for the briefer interaction window. Ideally, we want such X-raypulse to have a duration that is in the same regime as the timescales of thedamage processes, or shorter. While damage still commences in this setup,its effects does not extensively manifest until after the pulse has passed anda diffraction pattern has been recorded. However, this concept – known asdiffraction before destruction [66] – relies on the existence of a light sourcecapable of producing such ultrashort, ultraintense pulses.

Conventional X-ray crystallography experiments are usually carried outwith a synchrotron light source. At a synchrotron facility, electrons are ac-celerated in bunches to relativistic speeds around a closed-loop trajectory in astorage ring using magnetic fields. The field is synchronized to accommodatefor the increasing kinetic energy of the electrons as they accelerate, hence thename synchrotron. Once the electrons have reached the desired speed, theyare guided through an insertion device where the X-ray radiation is produced.This device (a wiggler or an undulator) consist of periodically arranged dipole

28

Page 29: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

magnets of alternating polarity that force the electron beam into a sine-like os-cillatory motion along the path of travel. The repeating transverse accelerationis an effect of the Lorentz force and results in emittance of the desired photons.While the product is an intense beam that does have pulsing properties due tothe bunched electrons, the pulses are neither intense enough nor brief enoughto enable damage-free imaging of nanocrystals.

To accomplish this feat, the pulses used need to be shorter than 100 fs (1fs = 10−15 s) while simultaneously delivering 1012 coherent photons of appro-priate energy. Producing such extreme light is not trivial, however. Shorterpulses can be attained at a synchrotron by using a slicing source to extractsmall portions of the electron bunches, but fewer electrons also equals fewerphotons. With the already limited photon count of approximately 106 pho-tons per synchrotron pulse [67], the intensity provided by only a fraction ofthe electrons would be many orders of magnitude below what is needed. Thisapproach is clearly not feasible for our purposes.

Other potential alternatives are high-harmonic generation (HHG) or usinglaser induced plasmas. In HHG, a pulsed laster is used to excite a gas result-ing in emittance of higher harmonics of the incoming beam. The generatedpulses are both spatially and temporally coherent and can be extremely short,reaching into the sub-femtosecond domain [68]. Unfortunately, the energiesand intensities provided both fall short when it comes to diffractive imaging.The energy aspect can be mitigated by instead inducing plasma formation ofa liquid or a metal with a powerful femtosecond laser. But while the burst-wise emission of photons from such plasma may reach energies well into thehard X-ray regime, the pulse lengths are undesirable. Additionally, the issueof intensity remains with fluxes being too low within a given bandwidth.

The only light source offering pulses meeting the proposed criteria cur-rently available are X-ray free-electron lasers (XFELs). Relying on principlesrelated to synchrotrons, they are able to generate femtosecond pulses of un-precedented intensities.

2.3.1 X-ray free-electron lasersAs opposed to the storage ring in a traditional synchrotron, free-electron lasersinstead utilize linear acceleration of electrons to relativistic speeds. Along thepath is an undulator where synchrotron radiation is emitted (see Section 2.3and Figure 2.7 below). As more photons are created, the electromagnetic fieldgrows stronger until the electric component of the field eventually starts to in-teract with the charged electrons. As a result, some of the electrons slow downwhile others to speed up, essentially compressing them spatially into distinctclusters. The phenomenon is known as microbunching, and neatly separatesthe clusters exactly one wavelength apart. In this configuration, monochro-matic X-rays emitted from the microbunches as they continue through the un-

29

Page 30: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

dulator will be in-phase. Photons will therefore interfere constructively andgive rise to a non-linear increase in beam power, enabling XFELs to generatelaser-like pulses with peak brilliances many orders of magnitude above thoseof synchrotrons.

Figure 2.7. Self-amplified spontaneous emission. The accelerated bunch of ran-domly distributed electrons enters the undulator, a structure of periodically arrangedmagnets with alternating polarity. Upon the influence of the magnetic field, the elec-trons start to oscillate and emit synchrotron radiation. The produced photons interactwith with the electrons, giving rise to microbunching, which in turn enhances the co-herence and intensity of the beam. (Image courtesy of European XFEL.)

A light source utilizing the described process, self-amplified spontaneousemission (SASE), was proposed already in the early 1970’s [2]. But it was notuntil nearly 40 years later that the first XFEL capable of producing hard X-raysat ~10 keV came online. Construction of the Linac Coherent Light Source(LCLS) at SLAC National Accelerator Laboratory in Stanford, California wasfinalized in 2009 [6] and contributed to hundreds of scientific publicationswithin the first five years of operation [69]. Numerous other facilities withsimilar capabilities have been built since – SACLA in Japan [70], PAL-XFELin South Korea [71], SwissFEL in Switzerland [72] and the European XFEL inGermany [73] – while others are underway, for instance the upgraded LCLS-IIin the US [74] and SHINE in China [75].

While the sub-100 femtosecond pulses from XFELs indeed are short enoughto outrun many of the damage processes and intense enough to produce aproper diffraction pattern from weakly scattering samples, one key issue re-mains. After interacting with the pulse the sample is invariably destroyed,meaning that the collected data only allows for structural reconstruction intwo dimensions. To extend this to three dimensions the crystal need to be re-placed continuously, which is a central concept behind the method of serialfemtosecond crystallography (SFX).

30

Page 31: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

2.3.2 Serial femtosecond crystallographyIn SFX, many copies of the small sample crystal are delivered in a steadystream to the X-ray focus. This creates a narrow interaction point of bothbeams. As a crystal cross this interaction point, it is hit by an XFEL pulseand diffracts onto the detector located downstream in the X-ray direction. Thesample quickly disintegrates from the exposure while it continues its trajectoryand is soon replaced by a new, undamaged copy. The recorded pattern is saved,and the process is repeated with the next crystal. Each crystal is randomlyoriented in space, so collecting a large number of patterns allows for a fullthree-dimensional dataset to be obtained.

This novel method of protein structure determination was first demonstratedexperimentally by Chapman et al. in 2011 [7]. In their setup, 0.2−4 μm pho-tosystem I crystals were introduced to the XFEL interaction point by a liquidwater microjet and the patterns were measured with two p-n junction charge-coupled devices (pnCCDs). Using 70 fs pulses of 1012 photons at 1.8 keV, theymanaged to collect enough patterns to assemble a three-dimensional diffrac-tion volume and calculate an electron density map. The map correspondedwell to the density determined by conventional X-ray crystallography up to aresolution of 8.5 Å – the theoretical maximum of their setup. Clearly, SFXwas a force to be reckoned with, as was further validated by the results itcontributed to the following years [76].

SFX offers the possibility of imaging crystals too small to provide enoughsignal at synchrotrons. But is there a limit to how small they can get? Theo-retically, no, since even single molecules have an electron density that scattersX-rays. However, due to the lack of repeating units, such samples are signif-icantly more sensitive to radiation damage and produce only a fraction of thesignal. Despite this, the prospect of single particle imaging (SPI) with X-rayshas long been viewed as the holy grail in structure determination, and with theadvent of XFELs we are closer than ever to making this dream a reality.

2.4 Single particle imagingSince many proteins cannot be crystallized, the potential for SPI is virtuallylimitless. Being able to establish the structure of any biomolecule would havea profound impact on structural biology and medical sciences. The techniquehas been demonstrated to produce damage-free renditions of particles up toresolutions of tens of nanometers [8, 77], but a range of challenges need to beovercome to achieve atomic resolution [15].

The idea is identical to SFX, where particles are continuously introduced tothe XFEL beam in vacuum and allowed to diffract. However, the low scatter-ing signal generated by individual molecules makes the technique extremelynoise sensitive. This puts constraints on how the sample is delivered, sincefor instance a liquid jet easily can produce enough extraneous noise to drown

31

Page 32: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

out the signal. Alternative sample delivery schemes have been developed forboth SPI and SFX, such as aerosolization by a gas-dynamic virtual nozzle(GDVN) [17] or electrospray injection methods [78], that aims to reduce theamount of residual water around the samples. While noise levels are drasti-cally decreased, the obvious downsides to such implementations is the changefrom a potentially native environment of the biomolecule, and what structuralrepercussions the change might entail. Although, at least for some proteins,the effects are seemingly minor [79, 80].

Figure 2.8. Single particle imaging. Aerosolized identical particles are continu-ously delivered to the XFEL beam in random and unknown orientations. As a samplereaches the interaction point it is intercepted by a femtosecond pulse, giving rise toa continuous diffraction pattern. The exposed sample is heavily ionized and rapidlydisintegrates in a violent Coulomb explosion as it exits the beam focus. It is then re-placed by a new, undamaged copy of the same molecule and the process is repeated.Once enough patterns has been collected, they may be combined to determine thethree-dimensional structure of the sample.

There is also a distinct difference between diffraction patterns from crys-talline and single particle samples. A crystal produces a discrete pattern withhigh-intensity Bragg peaks due to its periodicity, but such periodicity is miss-ing in a single particle. Instead, the resulting pattern is continuous with in-tensity variations arising from interference of coherent waves from withinthe same molecule. These features, called speckles, fortunately encode forthe same structural information the Bragg peaks do in the crystalline case,so structure retrieval is still possible. However, with a considerably lowersignal-to-noise ratio, this task becomes substantially more difficult and usu-ally requires many patterns to be averaged before they can be used in structuraldetermination.

32

Page 33: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

In the following sections, we will introduce the challenges currently facingSPI and, to some extent, SFX, which are the focus of this thesis. But to il-lustrate the complexity of these novel techniques, and to make clear that thelist is in no way exhaustive, we first dedicate a few paragraphs to examples ofother areas where improvements are necessary.

In both methods, there is currently no practical way of synchronizing thearrival of sample molecules with the XFEL pulses. Whether a given particlein the constant stream is intercepted by a pulse is simply a matter of probabil-ities, commonly called the hit rate. As a consequence, only a fraction of theparticles yield useable diffraction patterns. An increased hit rate would makeSPI more time efficient and reduce sample consumption, which can be limit-ing for samples that are expensive or otherwise hard to acquire in the neededquantities. A slower flow of particles or using higher sample concentrationsare two options, but both run the risk of clogging the injection system, and thelatter might have a negative impact on the total amount of sample used. Theincreased number of unwanted multiple hits and a greater risk of sample freez-ing are other side effects that would need to be taken into consideration [81].

Instrumental development constitute another avenue for ameliorating struc-tural studies at XFELs. Upgrades to the X-ray source to allow for even shorterpulses and higher intensities would better circumvent radiation damage andyield higher signal levels. An increased repetition rate of pulses would also bebeneficial, as hit rates would increase and data collection could be carried outfaster. The latter is already underway, with the newly inaugurated EuropeanXFEL offering a pulse repetition rate of 27,000 Hz – 225 times higher than theLCLS – and the upcoming LCLS-II promising to boost this number to one mil-lion shot per second. To take full advantage of these higher rates, the detectorsthat record the diffraction patterns need to be developed in tandem. The twotypes of detectors suitable for diffraction experiments used at the LCLS arep-n charge coupled devices (pnCCDs) and Cornell-SLAC pixel-array detec-tors (CSPADs). These are found at the atomic, molecular and optical science(AMO) beamline [82] and the coherent X-ray imaging (CXI) beamline [83]respectively, and both offer a readout rate matching the 120 Hz of the pulses.By comparison, the single particles, clusters, and biomolecules and serial fem-tosecond crystallography (SPB/SFX) beamline at the European XFEL housesthe so-called adaptive gain integrating pixel detector (AGIPD) [84]. While itvastly outperforms the LCLS detectors in terms of frame rate at 4,5 MHz, itrelies on storing a maximum of 352 images to an analog memory that are readout between pulse trains. So in a sense, is unable to utilize the full repetitionrate of the X-ray source. Clearly, new detector technologies to accommodatefor the improved capabilities of both current and future XFELs are highly de-sired.

Lastly, algorithm development plays a huge role in the success of SPI. Ro-bust software is an absolute necessity in order to successfully identify andsort suitable diffraction patterns, correctly assemble them in three dimen-

33

Page 34: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

sions, solve the phase problem, and ultimately retrieve the molecular struc-ture. While such algorithms have been developed and applied to single parti-cle diffraction data, such as the expectation maximization compression (EMC)algorithm for orientation recovery [20] and the hybrid input-output (HIO) al-gorithm for phase retrieval [85], further advancements are needed. A deepersummary of the current status of available computational resources for SPI(and of the method in general) can be found in a recent review article by Sunet al. [86].

We reiterate that all these aspects are of vital importance to both SPI andSFX, and should not be neglected. However, in this work we limit ourselves tothe challenges of orientation recovery, noise contributions from the Coulombexplosion and heterogeneous samples, as well as the temporal profile of theXFEL pulse.

2.4.1 The orientation problemThe currently most viable means of introducing single particles to the beamrelies on aerosolization of the particles, producing small droplets containingindividual molecules. Once in vaccum, much of the noise-causing solvent isideally evaporated without forcing structural modifications. A key drawbackto this approach, however, is the lack of control over the spatial orientation ofthe sample. Recovering the orientation from the diffraction pattern a posteriorican be done with a number of different algorithms [20,87–90], but the processis oftentimes troublesome and can deem patterns that are too noisy unusable.This may be rectified by regulating the orientation through the use of externalelectric fields [21], but this has yet to be verified experimentally.

In Paper I, we explore the possibility of orientation determination from theXFEL-initiated Coulomb explosion (see Figure 2.9). By mapping the ion tra-jectories predicted by molecular dynamics (MD) simulations of a realistic SPIsetup, we investigate the reproducibility of the explosions of four different pro-teins. This presents a new potential approach to the orientation problem whereion map data could be combined with the measured diffraction patterns to aidthe orientation recovery process. Depending on the level of reproducibility,such joint concept could significantly improve the quality of the assembledthree-dimensional volume, reduce time needed for data collection, and ulti-mately yield higher resolution structures.

2.4.2 Coulomb explosionIn the context of short-pulse imaging, long term damage processes (see Sec-tion 2.2.4) become largely irrelevant and can usually be disregarded. In somecases, this may extend to electron collision events as well. Simulations haveshown that for sample sizes < 500 nm diameter, many of the photoelectrons

34

Page 35: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Figure 2.9. XFEL-induced Coulomb explosion. Molecular dynamics simulation ofa single ubiquitin protein (white, H; grey, C; blue, N; red, O; yellow, S) exposed toan XFEL pulse. The X-ray interaction strips the sample of electrons, resulting in anexplosion driven by Coulomb forces. Most of the damage happens after the pulsepeak intensity at t = 0 fs, showcasing the diffraction before destruction concept. Thesimulated pulse (yellow arrow) consisted of 1012 photons at 8 keV and had a Gaussiantemporal profile with a full-width at half maximum (FWHM) duration of 50 fs. Thephotons were uniformly distributed in space over a 100 nm diameter focal spot.

freed by an XFEL pulse are able to escape the sample entirely [62]. This isbeneficial since it limits the extent of damage by secondary ionizations; oncean electron has left it can no longer collide with atoms within the sample. Onthe other hand, a deficit of electrons invariably leads to a net positive charge.For this reason, the smaller samples relevant to SPI are bound to quickly dete-riorate in a Coulomb explosion, as illustrated in Figure 2.9.

The Coulomb explosion mechanism is fundamentally different from the dy-namics of atomic displacement in a larger sample. By contrast, in a biggersample volume, released photoelectrons become trapped and distribute theirenergies through collisions and heating. The accumulation of free electronsinside the material leads to a form of layering effect where the outer shellpeels off due to positive charges, similarly to the previous case, and the innercore undergoes a hydrodynamic expansion from the increased electron pres-sure [4,63]. While the largest known human protein, titin, has a length of 1 μmand a width of 3−4 μm, most biomolecules are considerably smaller than the500 nm cited above. Hydrodynamic expansion effects therefore become in-consequential in SPI for most proteins, but may need to be taken into accountin the SFX analog. This size-dependence of the displacement is important toconsider when modeling the two scenarios.

The onset of atomic displacement from a Coulomb explosion happens ontime scales comparable to currently available pulse durations (10− 100 fs).Consequently, it is a source of noise that is expected to manifest in the diffrac-

35

Page 36: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

tion pattern. However, the nature and extent of this noise is not well under-stood, and it remains an open question whether the atomic motion necessi-tates even briefer pulses. In fact, for nanocrystals this is not the case. Bartyet al. showed that the disruption of the crystalline order provides a gatingeffect such that the diffraction signal is terminated prematurely, making thepulse appear shorter [91]. A later simulation study supported the result andconcluded that pulse duration is not a limiting factor in SFX, given sufficientX-ray flux [22]. Moreover, a similar gating effect has been suggested for sin-gle molecules based on a model taking spatially correlated damage processes,but not the Coulomb explosion, into account [23].

In Paper II, we investigate the noise caused by the Coulomb explosion of asingle particle. Diffraction patterns are calculated and analyzed to determineboth the severity of the noise and whether a gating effect is observed. Theprospect of SPI with longer pulses is intriguing since sub-10 fs pulses are yetunavailable at the intensities and photon energies required.

2.4.3 Sample heterogeneityOne of the principal challenges facing SPI is the low signal yield. Because abiomolecule is weakly scattering, a single diffraction pattern from each molec-ular orientation does not provide sufficient signal for structure determinationto high resolution. To mitigate this deficiency known as shot noise, many pat-terns from the same orientation must be averaged. Provided that all recordedpatterns come from identical samples, the averaging is well-motivated and willhave no impact on the resolution limit. But especially large biomolecules areknown to be dynamic and undergo constant structural adjustments. Some eventake on several distinct configurations. This gives rise to the issue of sampleheterogeneity.

If individual imaged molecules exhibit a structural variation, averaging oftheir patterns will result in a lowering of the speckle contrast. The consequentloss in resolution is worse than the root-mean-square deviation (RMSD) of theatomic positions [24]. With current aerosol injection schemes used in SPI, thissituation is highly probable. Suggestions to counteract the problems of sam-ple heterogeneity, while simultaneously repressing radiation damage, includesencapsulation by a hydration layer [79, 92] and cooling of the sample.

In Paper III, we study how the noise from sample heterogeneity is relatedto temperature and level of hydration. A lower temperature and the presenceof a protective water layer is expected to cause less structural variability of aprotein in vacuum. The impact of these two factors on the diffraction patternis analyzed, with the goal of understanding their importance and identifyingthe optimal conditions for SPI. Such information is likely to be valuable to thefurther development of sample delivery systems and techniques.

36

Page 37: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

2.4.4 Pulse profileAs a side effect of the SASE process, XFEL-pulses have a stochastic distri-bution of photons in time [93]. Since the X-rays both promote the onset ofdamage and generate the diffraction signal – two time-dependent processes –it is fair to suspect that the temporal profile will affect the quality of the mea-surement. With emerging techniques to both shape [28] and characterize [94]XFEL pulse profiles, understanding this relationship is vital.

In Paper IV, we evaluate the effects of the temporal structure of the pulsein a SFX experiment. Simulations are carried out to predict the evolution ofradiation damage within a protein nanocrystal for various pulse profiles. Theoutput is then analyzed to determine the rate of decay of the Bragg signal,and thereby identify ideal pulse profiles for imaging. Our findings can be ap-plied to at least two different routes towards improving diffractive imaging atXFELs. Firstly, they provide guidelines for pulse shaping before sample inter-action. Secondly, if a suboptimal pulse is used, a combined understanding thedamage evolution and characterization of the pulse may allow for correctionsof the diffraction pattern to be made. This way, the signal quality can be im-proved despite an imperfect temporal distribution of photons. While the studycenters around crystalline samples, we expect the results to extend to singleparticles.

Lastly, in Paper V, we present a database of radiation damage simulationsavailable to the public. It contains calculations based on a non-local thermo-dynamic equilibrium radiation transfer plasma code for a variety of materialsexposed to an XFEL pulse. Specific pulse parameters such as photon energy,fluence and duration are included, and the returned plots show the time evo-lution of average ionization levels, electron and ion temperatures, and atomicdisplacement during exposure. While not explicitly applicable to SPI, the sim-ulations have value to the SFX community as well as to others utilizing XFELsin their research.

37

Page 38: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

3. Methods

In this section we describe the tools used to collect and analyze data in thepresented papers. All studies are solely based on simulations employing threeseparate theoretical frameworks. Molecular dynamics was utilized in papersI–III to describe spatial atomic movement over time. These simulations werecombined with diffraction pattern calculations in papers II and III based onscattering theory. Lastly, a plasma radiation transfer code was used in papersIV and V to estimate the evolution of radiation damage.

3.1 Molecular dynamicsA common protocol used to study the atomic motion of systems such as pro-teins is molecular dynamics (MD). Based on classical mechanics, it relies onsolving Newton’s equations of motion in an iterative manner. While it ne-glects quantum effects, MD is a good approximative numerical approach thatallows for simulations of systems too complex and time spans too long for theproblem to be solved analytically.

The general workflow for MD is illustrated in Figure 3.1. The initial state(time t = 0) is given by the starting positions ri and velocities vi of all atomsi = 1,2, . . . ,N in the system. In the case of proteins, the positions are usuallygiven by a structure file obtained from crystallographic data, and the velocitiesare generated from a Maxwell-Boltzmann distribution defined by the specifiedtemperature and randomly assigned to the atoms. The first iteration of thesimulation commences by calculating the net forces Fi acting on the individualatoms based on an interaction potential V ,

Fi =−∂V∂ri

. (3.1)

The potential is a function of atomic positions outlined by user-determinedinteraction parameters known as a force field (see Section 3.1.1 below) thatis unchanged throughout the duration of the simulation. Once the forces areknown, the atomic accelerations ai are determined by Newton’s second law.With mi being the mass of atom i, its acceleration is

ai =Fi

mi(3.2)

38

Page 39: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

1. Specify positions ri and velocities vi at t = 0

2. Calculate forces Fi = −∂V/∂ri

3. Determine accelerations ai = Fi/mi

4. Update positions and velocities based on ai

5. Increase t by Δt

Iteratesteps 2–6

6. Return full trajectory of the system

t < Tt = T

Figure 3.1. The global MD algorithm. The simulation starts at time t = 0 with asystem of known atomic positions ri, usually given by a structure file. Velocities viare determined from a Maxwell-Boltzmann distribution and an interaction potential Vis specified by a force field. Newton’s equation of motions are solved to determinethe forces Fi acting within the system, and subsequently the accelerations ai of eachatom. Applying the accelerations for a small time window Δt yields new positionsand velocities that describe the time-evolved system. Lastly, the time parameter isprogressed by Δt, finishing the first iteration of the algorithm. Using the new coor-dinates and velocities, the process is repeated until the desired simulation length T isachieved.

from which new atomic positions ri and velocities vi can be computed bynumerical integration of the differential equations

ai =dvi

dtand vi =

dri

dt. (3.3)

Various integration schemes exist for this purpose, with the one employedin all MD simulations presented here being known as leapfrog integration.The main idea behind this method is to determine positions and velocities inan alternating manner, using the recursive equations

r(t+Δt) = r(t) +v(t+ 12 Δt)Δt (3.4)

v(t+ 12 Δt) = v(t− 1

2 Δt) +a(t)Δt, (3.5)

where atomic indices have been dropped for clarity and exponents have beenadded to denote timesteps. Since positions are calculated at times t = nΔt,while velocities are evaluated at the midway points t = 1

2 nΔt along the sametimeline (with n∈Z

+), the two are continuously overtaking each other through-out the simulation, similarly to a game of leapfrog.

39

Page 40: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

With each iteration, the updated values are saved and the time variable t isadvanced by Δt. If t matches or exceeds the full simulation length T , the sim-ulation is terminated and the full time evolution of atomic coordinates (calledthe trajectory) is returned. If not, the next iteration of the algorithm is initiatedutilizing the new positions and velocities.

3.1.1 Force fieldsTo determine the net forces that drive the dynamical behavior of the system ina MD simulation, a force field must be specified. There are numerous forcefields available and selecting a suitable one simply depends on the nature of thesimulation. For biomolecules, some commonly used force fields are AMBER(assisted model building with energy refinement) [95], CHARMM (chemistryat Harvard macromolecular mechanics) [96, 97] and OPLS (optimized poten-tials for liquid simulations) [98–100]. In the MD-based studies presented here(papers I–III), the latter is exclusively used.

A force field defines the interatomic potentials of the system based on equa-tions and parameters derived from experiments. Typically, energy contribu-tions from bonded and nonbonded interactions are treated separately. Thetotal potential energy is therefore given by

V =Vbonded +Vnonbonded, (3.6)

although it may sometimes contain additional terms to include other contribu-tions such as restraints or external forces.

The bonded interactions specifically model covalent bonds and are gener-ally decomposed into three potential energy terms,

Vbonded =Vbonds +Vangles +Vtorsions, (3.7)

which are functions of bond lengths, angles and dihedral angles, respectively.The former two are commonly represented by quadratic functions, for exam-ple derived from Hooke’s law, which gives a good estimation of the chemicalproperties as well as a high computational efficiency. However, such model forVbond does not accommodate bond breaking since these functions grow indef-initely as the bond stretches (or compresses) from its equilibrium distance. Insituations where breaking is necessary, such as in papers I and II, we insteademploy a Morse potential [101]

Vbonds = ∑bonds

Dr(1− e−βr(r−r0))2, (3.8)

where Dr is the depth and βr is the steepness of the potential well, r is thebond distance and r0 is the equilibrium distance (where the potential is mini-mized). The remaining bonded terms used in aforementioned papers are unal-

40

Page 41: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

tered from the OPLS-model, where they are defined as

Vangles = ∑angles

kθ (θ −θ0)2, (3.9)

where kθ is the angular spring constant of the bond, θ is the bond angle andθ0 is the equilibrium angle, and

Vtorsions = ∑torsions

{4

∑n=1

Vn

2[1+ cos(n ·φ)]

}(3.10)

where Vn are Fourier coefficients and φ is the dihedral angle.Nonbonded interactions encompass long-range electrostatic forces and van

der Waals forces. In OPLS, they are evaluated for all intermolecular atoms andintramolecular atoms separated by at least three bonds. Similarly to above, thenonbonded contribution to the total potential can be additively written as

Vnonbonded =Velectrostatic +Vvan der Waals. (3.11)

The electrostatic term is computed using Coulomb’s law, summed over atompairs i j in the system. Denoting the Euclidian distance between two atoms byri j, and their respective charges by qi and q j, it is implemented in OPLS as

Velectrostatic = ∑i

∑j<i

14πε0

qiq j

ri j, (3.12)

where ε0 is the electric constant. Analogously, the van der Waals interactionsare modeled using the Lennard–Jones potential

Vvan der Waals = ∑i

∑j<i

4εi j

[(σi j

ri j

)12

−(σi j

ri j

)6]. (3.13)

Here, ε is the well depth and σ the Lennard–Jones radius, i.e. the (finite)distance at which the potential is zero. Both parameters are associated with thestandard geometric combining rules εi j =

√εiiε j j and σi j =√σiiσ j j. While

the long-distance van der Waals attraction is approximated by the (σ/r)6 termin expression 3.13, the (σ/r)12 term is included to account for Pauli repulsionat short interatomic distances. A 50 % scaling of Vnonbonded is applied to thespecial case of 1-4 interactions, where i and j are the distal atoms of a dihedralangle. This is to allow for the same set of parameters to be used for both inter-and intramolecular interactions. It also bears mentioning that OPLS does notinclude an explicit description of hydrogen bonds, since they already are wellapproximated by Coulomb and Lennard–Jones terms [102].

Taken together, the full interaction potential V is calculated from the pa-rameters Dr, βr, r0, kθ , θ0, Vn, q, ε and σ . To accurately reproduce experi-mental results, these parameters must be fitted not only to specific elements,

41

Page 42: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

but also to their chemical environment. There exists several parameteriza-tions for OPLS, but one of the most widely used is the all-atom (OPLS-AA)model. OPLS-AA lists the parameters for all atoms in different environmentsexplicitly. The parameters have been continuously refined over the years, mostrecently by Robertson, Tirado-Rives and Jorgensen [103], to better predict thebehavior of peptides and proteins. In papers I and II we apply the versionknown as OPLS-AA/L, as designed by Kaminski et al. [100].

3.1.2 Water modelsA standard way of enhancing the performance of the simulation is to treatcommonly occurring molecules separately, in particular water. By assigningan independent force field to the water molecules, their properties can be ac-curately represented while remaining simple enough to lower the computa-tional demands. Such force field is called a water model. The simplest wa-ter models are rigid, meaning that covalent bond lengths and angles are keptfixed. Nonbonded interactions, on the other hand, are usually implementedwith Coulomb and Lennard–Jones potentials as described above. While morecomplex models that include for instance flexible bonds and polarization ef-fects exist, the rigid models provide an accurate enough description for therelevant studies presented here.

Rigid water models may be further categorized by the number of interac-tion sites they employ. In this work, depending on the simulation type, we haveused both a three-site and a four-site model (see Figure 3.2). With three sites,there is a direct correspondence to the atoms of the water molecule. The sitesare distributed to match the water geometry and assigned with point chargesto mimic the dipole nature of water. A well-known rigid three-site model isTIP3P (transferable intermolecular potential with three points) [104]. In it, theO-site has a negative charge of −0.843 and is chemically bonded to two H-sites with charge +0.417 each. Both OH bonds have a constant length of 0.96Å and the rigid HOH angle is 104.5◦. Lennard–Jones parameters are only as-signed to the O-site in the original version of TIP3P, meaning that hydrogensare treated as points with no extension in space. However, a modified versionoptimized for the CHARMM force field includes additional Lennard–Jonesparameters for hydrogen atoms. Such model is more suitable for the presentedsimulations of single molecule Coulomb explosions, and was therefore cho-sen.

A related four-site model is TIP4P. The general geometry is identical toTIP3P, but it has an added fourth site located 0.15 Å from the O-site along thebisector of the HOH angle, towards the two hydrogens. The model also hasa slightly different distribution of charges; the negative charge is changed to−1.04 and placed at the fourth site rather than the oxygen, and the positivecharges at the H-sites are adjusted to +0.52 each. These alterations allow

42

Page 43: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Figure 3.2. Geometries of the TIP3P and TIP4P rigid water models. [left] InTIP3P, the oxygen site (red) has a charge of −0.843 and is covalently bonded to twohydrogen sites (grey) at an angle of 104.5◦. The bond length is 0.96 Å and the re-spective charges at the H-sites are +0.417. [right] In TIP4P, the overall geometry interms of bond lengths and angles is identical to TIP3P. However, the charge distribu-tion is modified by inclusion of a fourth site (purple) 0.15 Å away from the oxygenalong the bisector of the HOH angle toward the two H-sites is included. It hosts anegative charge of −1.04, while the O-site is neutral. The charges at the H-sites arealso changed to +0.52.

for a more favorable model of the electrostatic distribution over the moleculeand generally yield results more consistent with measured data. TIP4P wasimplemented in the vacuum simulations carried out in papers I–III.

3.1.3 GROMACS and XMDSeveral software packages have been developed to perform MD simulations.Some of the most frequently used are the large-scale atomic/molecular mas-sively parallel simulator (LAMMPS) [105], nanoscale molecular dynamics(NAMD) [106] and Groningen machine for chemical simulations (GROMACS)[107]. Each program suite comes with varying functionalities and optimiza-tions, but ultimately produce the same results if the specified simulation pa-rameters are identical. All MD simulations presented in this work was carriedout using GROMACS, for mainly two reasons. Firstly, the general experi-ence with MD within the research group was most profound with GROMACS,which circumvented the need to adopt the technicalities of a different softwarefrom scratch. Secondly, an earlier version of GROMACS include the exten-sion program XMD that provide code crucial to the Coulomb explosion studies(see below).

Historically, the first publicly available version of GROMACS was releasedin 1991. At the time it was developed at University of Groningen, Netherlands,but has since 2001 been primarily managed jointly between Uppsala Univer-sity and the Royal Institute of Technology, Sweden. Over the years, it hasbecome an immensely popular choice for conducting MD studies likely dueto its impressive speed, often outperforming similar programs by a factor of

43

Page 44: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

two or more. Additionally, GROMACS offers support for a large collection offorce fields and water models.

The specific version chosen for the presented studies was GROMACS 3.3.3,originally released in 2008. It hosts XMD, an extension that enables the sim-ulation of ionizing effects triggered by X-rays, that was removed from laterreleases. With XMD, the user can define the parameters of an X-ray pulseinteracting with the system of interest, making it well-suited for SPI-relatedstudies. These parameters include pulse duration, intensity, photon energyand size of the focal spot. XMD implements the predominant ionizing pro-cesses by dividing atom bound electrons into two levels, valence and core.It allows for the direct removal of a core electron (photoelectric effect) or avalence electron (inelastic scattering) and if a core hole is present it may befilled by one valence electron while releasing another (Auger decay). Ioniza-tion events are distributed stochastically over the atoms in the system based onthe pulse parameters and cross sections taken from the NIST XCOM database.Throughout the simulation, the electron configurations of all atoms are moni-tored and their respective cross sections updated. The resulting positive ioniccharges are included in the dynamics calculations and the free electrons aretreated as a homogeneously distributed background charge. As such, XMDdoes not consider secondary ionizations by electron collision, which is as-sumed to be a fair approximation in the context of SPI (see Section 2.2).

3.1.4 Ion mappingAnalysis of the ion distribution from the simulated Coulomb explosions inPaper I was carried out as follows. The positions and velocities of all atomsof a chosen species (in particular carbon and sulfur) in the sample moleculewere extracted from the final frame of each simulation. Using these values,the atoms were projected onto a unit sphere centered at the origin defined bythe initial center of mass of the biomolecule. A grid of points were distributedacross the sphere by equal splitting of the polar and azimuthal angle intervals,and spherical caps of a set radius were centered around each point. The resultwas a network of circular detection areas covering the sphere and allowed forthe ion distribution to be binned into pixels around the interaction point. Suchsetup mimics an idealized setup of commercially available microchannel plate(MCP) detectors covering the full spatial angle, with an increased resolutionby finer sampling due to the detector overlap.

The binning enables easy plotting of the ion distribution by equirectangularprojection, similarly to a world map, which helps in visually interpreting thedata. Moreover, it samples the two-variable ion distribution function so that itcan be described as a Fourier sum of spherical harmonics. The correspondingFourier coefficients defines a power spectrum, which combined with a condi-tion derived from the Nyquist–Shannon sampling theorem gives insight into

44

Page 45: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

the angular resolution of the data. In other words, the accuracy of orientationdetermination by ion mapping can be evaluated.

3.2 Calculating diffraction patternsPapers II and III investigate how various noise sources impact the single parti-cle diffraction pattern quality. To do so, structures from MD simulations wereextracted and fed into a program referred to as DiffSim developed by Mar-tin et al. [23]. DiffSim calculates the diffraction pattern based on the inputstructure (the elements and their positions) and user-specified experimentalparameters. These parameters include both intensity, energy and focus of theX-ray pulse, as well as placement, pixel count and size of the detector. Thecalculated patterns are instantaneous and noiseless, and therefore represent anidealized case. Instantaneous means that there are no dynamics involved; thiscan be viewed as a simultaneous arrival of all photons, or equivalently, thatthe structure remains static during exposure. Noiseless refers to the absenceof shot noise. Recall from Section 2.4.3 that shot noise arises from the fact thata biomolecule is weakly scattering. As a consequence, a single measured pat-tern will not provide sufficient signal, which necessitates averaging of manypatterns. Patterns determined by DiffSim correspond to the convergence ofsuch averaging where the number of included patterns approaches infinity, allfrom identical samples in the same orientation. However, the program does al-low for generation of noisy patterns by estimating shot noise through Poissonsampling of the individual pixels in the noiseless pattern.

Diffraction patterns are calculated by standardized methods originating fromscattering theory. X-rays are weakly scattering, meaning that the wave propa-gating through the sample does not significantly differ from the incident wave.Assuming that they are equal is known as the Born approximation and pro-vides a description good enough for our needs. Starting from a single atom j,the elastic X-ray scattering amplitude is then given by the Fourier transformof its electron density ρ(r), commonly called the atomic scattering factor

f j(q) =∫ρ(r)e−ir·qdr. (3.14)

Here, r is a position vector in real space from the center of mass of the atomand q = k′ − k is the momentum transfer between incoming and outgoingphotons with wave vectors k and k′, respectively. Assuming that the electrondensity is spherically symmetric, which is commonly done in crystallography,we can write it as ρ(r) = ρ(r), where r = |r|. Defining the magnitude ofthe scattering vector similarly, q = |q|, and Θ as the angle between the twowave vectors, the scalar product in the exponential becomes r ·q = rqcosΘ.Introducing these relations into Equation 3.14 in spherical coordinates, the

45

Page 46: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

expression simplifies to

f j(q) =∫ ∞

0

∫ π

0

∫ 2π

0ρ(r)e−irqcosΘr2 sinΘdφdΘdr

=4πq

∫ ∞

0rρ(r)sin(qr)dr.

(3.15)

Clearly, with a spherically symmetric electron density, the scattering factor isalso spherically symmetric and solely determined by the magnitude q of themomentum transfer vector. It is common to denote the scattering angle byθ = Θ

2 such that the relationship between q, the wavelength λ of the incident(and elastically scattered) light, and the scattering angle is

q = 4πsinθλ

. (3.16)

DiffSim determines these element-specific and angle-dependent scattering fac-tors at the relevant photon energies by applying a correction term to tabulatedforward scattering values, as described by Henke, Gullikson and Davis [108].It also employs Slater orbitals [109] to calculate the corresponding ionic scat-tering factors whenever ions are present in the system.

The scattering factor describes the amplitude of the scattered plane wave,which therefore can be written as

ψ j(q) = f j(q)e−iq·R j , (3.17)

where R j is the position vector of the atom. Knowing the individual f j(q)for all j = 1,2, . . . ,N atoms of a system, such as a molecule comprised ofindependent scatterers, we can extend the theory by writing the resulting waveas a vector sum of the individual waves:

ψ(q) =N

∑j=1

f j(q)e−iq·R j . (3.18)

However, we cannot directly measure the total scattering amplitude in exper-iments, and phase information is lost entirely. The detector instead capturesthe intensity, which is proportional to the probability

|ψ(q)|2 =(

N

∑j=1

f j(q)e−iq·R j

)(N

∑k=1

fk(q)eiq·Rk

)

=N

∑j=1

N

∑k=1

f j(q) fk(q)eiq·(Rk−R j)

=N

∑j=1

f 2j (q)+2

N

∑j=1

j−1

∑k=1

f j(q) fk(q)cos[q · (R j −Rk)].

(3.19)

46

Page 47: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

The last step follows from dividing the double sum into one set of terms wherej = k and another where j �= k. In the latter, the imaginary parts vanishes whenterms of interchanged values for j and k are added, hence the cosine replacingthe exponential. This further highlights the fact that phases are lost.

By implementing the theoretical derivation above, DiffSim is able to outputa pixelated diffraction pattern in the form of a matrix. The input structure filescontain both the atomic species, allowing for calculation of the proper atomicscattering factors for the chosen photon energy, and the atomic positions R j.Based on the given detector geometry, the software determines relevant scat-tering vectors q pointing to the center of each pixel along with correspondingsolid angles dΩ. The final intensity measurement is then computed as

I(q) = r2e P(q)dΩI0

[N

∑j=1

A j(q)+2N

∑j=1

j−1

∑k=1

B jk(q)

], (3.20)

where re is the classical electron radius, P(q) is a polarization factor and I0is the intensity of the incident X-ray pulse. A j(q) and B jk(q) are defined asthe respective summation terms in Equation 3.19. An example output of suchcalculation can be seen in Figure 3.3.

Figure 3.3. Single particle diffraction pattern and corresponding radial profile.

[left] Logarithmic heat map visualizing a noiseless diffraction pattern from a singleubiquitin molecule calculated by DiffSim, with colors indicating photon intensities.A virtual detector with 1516× 1516 pixels with a size of 110× 110 μm each wasplaced 50 mm behind the sample in the calculation. The X-ray pulse consisted of1012 photons at 8 keV (λ = 1.55 Å) distributed over a circular spot with a diameterof 100 nm, fully engulfing the sample molecule. [right] The corresponding radiallyintegrated profile of the pattern. A dashed line has been added at q = 0.5 nm−1,representing the black circle enclosing the central speckle in the diffraction pattern.

3.2.1 Radial profile and SNRIn a diffraction pattern such as the one above, the momentum transfer vectorq can be thought of as a vector pointing outwards with origin at the center

47

Page 48: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

of the pattern. We can therefore write it in polar coordinates as q = (q,ϕ),with q being the magnitude and ϕ the polar angle from a reference axis. Themagnitude corresponds inversely to distances in real space, explaining whythe circle of radius q in the pattern is usually called a resolution shell. Anassumption we make in papers II and III is that the spatial coherence of theproteins under study do not inherently have a strong directional bias. It isunlikely that the electron density of large, folded biomolecules would exhibita repeating structure that is significantly overrepresented in certain directions.By comparison, the following would not be a viable approach for a saturatedlong-chained fatty acid due to its linear shape. Based on this assumption, thediffraction pattern I(q) can be condensed into a one-variable radial profile byaveraging the intensities over the full angle ϕ:

I(q) =1

∫ 2π

0I(q)dϕ. (3.21)

The radial profile is, in other words, a function describing the intensity ex-pectation value over all resolution shells covered by the detector (see Figure3.3). One of its uses is for estimation of the q-dependence of both shot noiseand signal strength in I(q). Because the former can be modeled as a Poissonprocess, the mean shot noise level is simply proportional to

√I(q). As for the

signal, we define

σI(q) =

√(1

∫ 2π

0I2(q)dϕ

)− I2(q), (3.22)

which is the standard deviation of the intensity in each resolution shell. Inter-ference between outgoing waves give rise to the speckles seen in the patternthat hold the structural information, and σI(q) is a measure of the contrastof those speckles. A higher value suggest larger intensity fluctuations, mean-ing that speckles at the given resolution are more discernible from noise andtherefore that the signal strength is higher.

In the papers, the integration scheme of Equation 3.21 is applied to an av-eraged pattern μ(q) and a variance pattern σ2(q), calculated from a numberof individual diffraction patterns. The averaging mimics an SPI setup wherepatterns must be summed to enhance the signal-to-noise ratio (SNR), and thetwo quantities outlined above can be inferred from μ(q) and its radial pro-file μ(q). However, another source of noise emerges if the summed patternsdeviate from each other, for instance due to structural heterogeneity or dam-age dynamics. Such variational noise is captured well by the variance patternσ2(q), so we quantize it as the square root of its radial profile:

σ(q) =

√1

∫ 2π

0σ2(q)dϕ. (3.23)

48

Page 49: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

The established measures can eventually be combined to compute the ex-pected SNR from the data. Since shot noise and variational noise are treatedseparately, their contributions to the SNR can be evaluated independently.These ratios can be written

SNRshot(q) =σI(q)√μ(q)

and SNRvar(q) =σI(q)σ(q)

(3.24)

respectively, and show the impact of each noise source over the relevant reso-lution span. Since the two generally cannot be distinguished in experiments,we also treat the noise collectively to form a total ratio, SNRtot(q). It providesa more complete picture of the SNR we can expect and is calculated as

SNRtot(q) =σI(q)√

μ(q)+σ2(q), (3.25)

where the denominator is the combined standard deviation of the two indepen-dent noise variables given by the square root of the sum of their variances.

3.2.2 Pearson correlationA simple way of comparing the similarity between two calculated diffractionpatterns is by Pearson correlation. It is a common statistical tool that pro-vides a numerical measure of the linear correlation between two datasets. Thereturned value, sometimes referred to as Pearson’s r, may take on any valuebetween −1 and 1, with the two extremes indicating total negative or positivecorrelation, respectively. A value of r = 0 is the result of two uncorrelatedsets.

In the presented papers the Pearson correlation has two implementations.In the first, the correlation is applied to two diffraction patterns in a pixel-wisefashion, allowing for the complete datasets to be correlated. In this case, asingle r-value reflects the similarities between the two patterns. In the second,the intensity data is divided into subsets corresponding to different resolutionshells, which are correlated independently. This yields a range of r-values thatgives insight into the q-dependence of the correlation coefficient.

Mathematically, the first implementation is carried out as follows. Let A(q)and B(q) be two diffraction patterns of equal size to be correlated, with q

representing the different pixels. Next, let 〈A(q)〉 and 〈B(q)〉 denote the meanintensity over all pixels in each pattern. We first normalize the patterns bysubtraction of their respective means from all pixel values to obtain

A(q) = A(q)−〈A(q)〉 (3.26)

andB(q) = B(q)−〈B(q)〉. (3.27)

49

Page 50: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

The normalized datasets have an arithmetic mean equal to zero, securing thecorrelation measure to be scale invariant. This is important since the totalintensity recorded in each pattern may differ. Pearson’s r is then calculated as

r =∑q A(q)B(q)√

∑q A2(q)√

∑q B2(q). (3.28)

It is worth mentioning that calculation of the correlation coefficient only is car-ried out after a circular section in the center of both patterns has been removed.The reason is because this area largely overlaps with the incident (unscattered)X-ray beam and will therefore report intensities many orders of magnitudehigher than other sectors of the detector. Correlating patterns without first ap-plying this mask would consequently return values close to 1 and not reflectthe true correlation of the structural features. This also makes the theoreticalapproach more realistic as experimental detectors are equipped with a centralhole or beamstop to protect the instrument from the unscattered beam, settinga limit to the minimum scattering angle at which data can be collected.

Evaluating r as a function of q is done similarly by correlating each reso-lution shell separately. Because the patterns are pixelated, the shells are dis-cretized by dividing q into steps equal to the pixel size. As a result, everypixel coinciding with a horizontal ray from the detector center will belong todifferent resolution shells. To perform the correlations, we rewrite the pixelpositions in polar coordinates, q = (q,ϕ), and normalize the subsets individu-ally:

A(q,ϕ) = A(q,ϕ)−A(q), (3.29)

B(q,ϕ) = B(q,ϕ)−B(q). (3.30)

Here, A(q) and B(q) are the radial profiles as calculated in Equation 3.21.The correlation value at different resolutions q is then computed as before,where the sums span the full polar angle ϕ to include all pixels at the relevantdistance,

r(q) =∑ϕ A(q,ϕ)B(q,ϕ)√

∑ϕ A2(q,ϕ)√

∑ϕ B2(q,ϕ). (3.31)

This measure, sometimes referred to as the Fourier ring correlation (FRC)function, gives a more detailed picture of how the similarities between twodiffraction patterns vary as we transverse the resolution shells. Hence, it canbe helpful when investigating how pattern deviations limits the obtainable res-olution.

50

Page 51: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

3.3 Non-LTE plasma simulationsThe last two papers focus on the intricacies of SFX, in contrast to SPI, and em-ploy a different theoretical framework to evaluate effects of radiation damage.In these studies, we chose to adopt a non-local thermodynamic equilibrium(non-LTE) plasma code known as Cretin [110]. Such approach is necessarywhen investigating samples that are considerably larger than single molecules.Even in a crystal of a few hundred nanometers, electrons released by the X-rayinteraction will no longer escape the confines of the sample to the same extentas in a lone molecule. Consequentially, more energy is deposited into the sam-ple by secondary electron collision cascades – an effect that is disregarded inthe MD simulations discussed previously. Moreover, with the increased rate ofionization the sample rapidly undergoes a transformation into a plasma, whichgives rise to a host of additional properties that need to be taken into account.

A plasma is a highly ionized state of matter, where chemical bonds arebroken and unbound electrons and positively charged ions remain. In thisstate, various processes that impact the diffraction measurement become in-creasingly relevant. Apart from further ionization from colliding electrons,recombinations where electrons are recaptured into vacant atomic orbitals ac-companied by emission of photons may take place. The movement of freeelectrons also generate magnetic fields, which alters the kinetics of the system.These fields, along with electric fields arising from charges, can furthermorecause emittance of Bremsstrahlung upon interaction with energetic electrons,and thereby contribute to both noise and additional ionization.

While these effects theoretically could be included in a MD description ofthe system, it is not really viable in practice. Plasmas induced by an XFEL willnot be in a thermodynamic equilibrium, even locally, during the short timescales considered. The temperature of ions and electrons can vary greatlythroughout the system, and change over time due to radiative transfer. Thisimpacts the cross sections of the different processes, making them difficult toaccurately predict at an atom-by-atom level with MD. On top of that, sincethe studied systems are large in terms of number of atoms, the computationalcost would be tremendous. Instead, we use Cretin that allows for a collectivetreatment of the atoms.

Cretin is a multi-dimensional non-LTE plasma simulator specifically de-signed to model accretion disks around stars, but has proven to also be well-suited for laser-induced plasmas. It is based on a previously published codeby the same author [111], and has been successful in reproducing experimen-tal results from XFELs [91, 112]. As opposed to MD, Cretin does not takethe atomic structure of the sample into consideration; a sample is instead gen-eralized by its elemental composition and density. The initial geometry isconstructed as a number of equally sized adjacent continuum zones definedby the sample, each with identical pressure, temperature and electronic popu-lations. During the simulation, transferral of heat and radiation across shared

51

Page 52: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

zone boundaries is allowed, but mass transfer is not. This ensures that the el-emental composition remains intact. However, the density may still change asCretin allows for independent hydrodynamic expansion (or compression) ofthe zones.

How the simulation evolves depends on the user-defined X-ray pulse pa-rameters. Once set, Cretin tracks the time evolution of atomic kinetics, ionand electron temperatures, radiative transfer and X-ray absorption throughoutthe simulation. Atomic kinetics entail the electron configurations and transi-tion rates, and are evaluated by solving rate equations. Most ionization andexcitation processes, as well as their inverses, are included, and their rates arecalculated from tabulated data, equilibrated electron and ion temperatures andthe specifications of the XFEL pulse. Solving the atomic kinetics allow forprediction of the time evolution of the system, which is returned in terms ofelectronic states, ion temperatures and ion–ion collision frequencies.

Generally, utilizing a plasma approach allows for calculations of ultrafastenergy deposition and sample response during XFEL exposure, which is valu-able for data analysis. For instance, the output from such simulations wereused in Paper IV to study the change in Bragg peak intensities throughout thepulse. Plasma code may also provide beneficial predictions related to experi-mental design. For this reason we carried out Cretin simulations for a varietyof different materials, both organic and inorganic. The obtained data were an-alyzed and collected in a freely available database, as outlined in Paper V,with the intention of assisting the construction of future XFEL experiments.

52

Page 53: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

4. Results and conclusions

The following subsections present the main results from the individual studies.We also discuss the general conclusions that can be drawn from them. Notethat the texts merely serve as condensed summaries of the respective papers.Enclosed in this thesis are each publication, which the reader may consult fora more detailed view of the analyses.

4.1 Reproducibility of Coulomb explosions (Paper I)The first paper investigates the reproducibility of the Coulomb explosion inSPI. MD simulations predicting the explosion dynamics of four single proteinsamples, each with three different levels of hydration, were performed. Theaverage spherical ion distribution for different atomic species was then calcu-lated from 150 separate explosion events, where initial samples were hetero-geneous but in the same spatial orientation. Pulse parameters were represen-tative of current XFEL sources and unchanged between simulations. Figure4.1 displays the obtained distribution maps of carbon ions in the form of heatmaps, with an additional column representing a fully irreproducible explosionbehavior for comparison.

Definite signs of directional bias of the carbons can be seen in the figure.Most of the maps contain clear hotspots of ion densities significantly higher(and lower) than the random case. Hydration seems to enhance this effect, asthe hotspots grow increasingly distinct with thickening of the water shell. Thissuggests that the surface-bound water molecules provide a shielding mecha-nism that limits the accessible trajectories of the ions. Further analysis of thelysozyme sample (bottom row in Figure 4.1) revealed that the fraction of themap where ion densities measure five standard deviations outside of the meanof the random case increase from 2 % to 13 % when the hydration shell ischanged from 0 Å to 6 Å.

The shielding effect has bearing on the angular resolution of the ion dis-tribution. When projecting the 6 Å lysozyme carbon data onto spherical har-monics, we found from the power spectrum that orientation recovery shouldbe viable up to a resolution of 40◦ with respect to an equatorial line. While ahigher-resolved orientation is needed for accurate merging of diffraction pat-terns in SPI, the improvement from the 120◦ analog of waterless lysozyme isstill remarkable. As one might expect, the middle level of hydration yieldedan angular resolution between the extreme cases with an estimated value of72◦.

53

Page 54: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Figure 4.1. Carbon ion maps. The distribution of carbon ions averaged over 150 sim-ulated Coulomb explosions. Rows represent different proteins and the three leftmostcolumns the various hydration layers. The fourth column serves as a control, wherethe relevant number of ions have been uniformly spherically distributed and averaged150 times. Areas of higher or lower concentration of ions compared to the randomcase can clearly be distinguished for most of the samples. The color scale indicate theaverage number of ions, ranging from red (high) to blue (low).

Figure 4.2. Solvent oxygen ion maps. Theaverage distribution of oxygens found inthe hydration shell shows an inverse re-lationship with the corresponding carbonmaps, strengthening the notion of a shield-ing mechanism.

To further investigate the shield-ing by water, ion maps were plot-ted for solvent oxygen only, omit-ting oxygen atoms bound to the pro-tein. These can be viewed in Fig-ure 4.2 and show a striking comple-mentarity to the carbon maps. Ap-parently, carbons tend to migrate indirections where water oxygens arenot, and vice versa. This is in ac-cordance with the suggested shield-ing mechanism, which likely relatesto the structural arrangement of sol-vent molecules around the sample. Ifso, we would expect these maps toremain fairly stationary between re-peated measurements, provided thatthe level of hydration is similar, assimulations suggest that the hydro-

54

Page 55: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

gen bonding network of the solvent is highly reproducible [80]. By extension,a comparable robustness of the carbon maps should follow. The cited studyalso concluded that dehydration negatively affects the stability of the protein,indicating that the encompassing water provides a scaffold that restricts struc-tural alterations. This is certainly beneficial for SPI, where sample hetero-geneity is a key issue, and combined with the increased angular resolution ofthe carbon ion map a larger hydration shell may seem preferable. However,excess water lead to higher noise levels in the diffraction pattern and both ap-plicability and experimental implementation of measuring ion maps remainsunclear.

A similar analysis was conducted for the distribution of sulfur ions, whereall samples but lysozyme were excluded due to their limited sulfur content. In-terestingly, these heavier elements show exceptional explosion reproducibilitywith little angular spread, as visualized by the sharp hotspots in Figure 4.3.Close to the entire sphere records a statistically significant signal outside a 5σinterval around the random case mean for all samples, although the fractiondecreases slightly as the solvent layer is expanded (from ~100 % for 0 Å to96 % for 6 Å). This is in direct opposition to the behavior of carbon, possiblyindicating that the same shielding effect does not apply to the heavier sulfuratoms. Carbon maps also displayed large changes with hydration, which is notthe case for sulfur. Hotspots instead remain in similar positions regardless ofwater content, again suggesting a negligible interference from the solvent.

Figure 4.3. Sulfur ion maps of lysozyme. Clear hotspots indicate high levels ofreproducibility of sulfur trajectories, consequently yielding good angular resolutions.The distributions remain largely unchanged over all hydration shells, dismissing thesignificant impact by water seen in the corresponding carbon maps.

Both sulfur maps of hydrated lysozyme display an angular resolution of atleast 12.4◦, which is the upper limit of our analysis. For the dry protein, theanalogous value was estimated to 14.4◦. Clearly, sulfurs provide consider-ably more reproducible explosion patterns and should be superior to carbonfor orientation determination. Again the solvent seem to improve the resolu-tion, something that likely is a consequence of the increased structural stabilitydiscussed previously.

In our analyses, we treat atoms of the same element as indistinguishableand measure their distribution by binning. However, to gain a deeper under-

55

Page 56: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

standing of the explosion dynamics we also examined the exact trajectories ofeach individual sulfur atom. Figure 4.4 (bottom pane) below shows the sulfurangular trajectories extracted from all 150 explosions of waterless lysozymewith the different sulfurs indicated. While all sulfur ions exhibit reproducibleexplosion patterns, forming concentrated clusters around the sphere, some ofthem aggregate pairwise. These dual clusters also tend to have a smaller an-gular spread than those formed by single sulfurs. From the initial structureof the molecule (see Figure 4.4, top pane) it becomes apparent that the pairedclusters arise from atoms involved in disulfide bonds, suggesting that such in-teraction contributes to a higher explosion reproducibility. Disulfide bonds areexclusively formed between the sulfurs of cysteine residues in proteins – theonly other amino acid containing sulfur, methionine, cannot form these bonds.For this reason, explosion mapping of sulfur ions may be particularly suitablefor orientation of cysteine-rich samples.

Figure 4.4. Connecting the sulfur ion trajectories to the structure of lysozyme.

[top] A representation of the tertiary structure of lysozyme with the individual sulfurshighlighted and distinguished (numbered yellow spheres). Colored sticks outline thetwo sulfur-containing amino acids, cysteine (magenta) and methionine (orange). [bot-

tom] Combined results of the 150 MD simulations of dehydrated exploding lysozyme.Narrow clusters are formed by disulfide bonded sulfurs of cysteine (purple and yel-low), while the lone sulfurs of methionine (green and blue) exhibit a wider spread.

56

Page 57: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

4.2 Radiation damage in SPI (Paper II)Collecting high-quality single-particle diffraction patterns is a cumbersomeendeavor. Not only is the signal weak due to the small elastic X-ray inter-action cross sections, but it is also masked by noise stemming from varioussources. In the second paper, we estimated the severity of noise caused bythe Coulomb explosion and compared it to the effects of sample heterogene-ity and shot noise. By calculating and averaging noiseless diffraction patternsfrom every frame of the simulated lysozyme explosion events in one orien-tation, we obtained a pattern representative of what we ideally can expect tocompile from many shots. Since this pattern contained both the interferencefrom the Coulomb explosion and structural variability between samples, wecorrelated it to the averaged pattern from initial structures to isolate and mea-sure the impact of damage. Six different sets of pulse parameters were used;each combined from two intensities (1012 and 1013 photons) and three pulsedurations (5, 25 and 50 fs FWHM). Figure 4.5 shows the results from the pixel-wise Pearson correlation, with central pixels omitted, from the lower intensitypulses.

Figure 4.5. Pearson correlation of diffraction patterns. Vertical lines show the cor-relation values (given above each line) between the average time-integrated patternsand corresponding average initial structure patterns for the different pulse durations.Heterogeneity effects are captured by both patterns, and hence diminishes by the cor-relation procedure, while explosion dynamics only affect the former. Deviations fromperfect correlation scores of 1 are consequently due to the Coulomb explosion. In-cluded are also density plots from 50,000 pairwise correlations of noiseless patternsfrom heterogeneous structures (orange) and Poisson sampled patterns from a singlestructure (grey), with numbers showing the correlation value with highest probability.The density plots were smoothened using a standard Gaussian kernel and illustrateexpected correlation values in the presence of sample heterogeneity and shot noise,respectively. The simulated pulse had a total of 1012 photons at 8 keV, uniformly dis-tributed in space over a focal spot of 100 nm in diameter. X-ray pulse durations aregiven in FWHM of the Gaussian temporal profile.

57

Page 58: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

For reference, we performed two separate sets of 50,000 correlations each.In the first, noiseless patterns from randomly chosen heterogeneous structuresof lysozyme were correlated. In the second, the pattern of a single structurewas repeatedly Poisson sampled to emulate shot noise before correlating thenoisy patterns. This yielded two distributions of correlation values (visualizedas orange and grey density plots in Figure 4.5), allowing for comparisons be-tween noise contributions from damage, heterogeneity and shot-to-shot fluc-tuations.

A significant difference in correlation coefficients can be seen in the lower-intensity pulse data presented here. For the shortest pulse the r-value is closeto 1, indicating that almost no changes can be seen in the pattern due to theCoulomb explosion. As the pulse grows longer, damage has time to manifestto a greater extent and the correlation value subsequently decreases. Omittedhere is the data from the higher-intensity pulse of 1013 photons, which showsa similar trend but with lower values overall. This difference is to be expectedas more photons invariably lead to higher ionization levels and more damage.However, in both cases all calculated coefficients stay above the peak value ofthe distribution from heterogeneous samples. It is a remarkable result sinceit suggests that sample variation cause more noise when averaging patterns ofthe same orientation than the Coulomb explosion does, even at longer pulsedurations of 50 fs. The comparably low peak value for shot noise is unsur-prising since very few photons are scattered in these patterns, highlighting thereason averaging is necessary in the first place. The distribution is slightlyshifted toward the right with a peak at r = 0.045 when the photon count isincreased tenfold, but still remain far from overlapping with the other datapoints.

We investigated the resolution-dependence of r by correlating resolutionshells individually. As can be seen in Figure 4.6, the values tend to decrease aswe approach higher resolutions, with the decrease becoming steeper as longerpulses are employed. However, the loss of correlation is not as severe as forthe heterogeneous dataset, which undergoes a significantly sharper drop – inparticular around the 2.5 nm−1 mark. As a result, the exploding patterns re-main above the values of the heterogeneous patterns over the entire resolutionspan covered by the detector, excluding the central speckle that was omittedfrom the pixel-wise analysis. This relationship generally holds for the higherintensity pulses as well, except for in the low-resolution region up to q = 1.8nm−1 where the two longer pulses approximately match the heterogeneouscase. At higher resolutions, the difference becomes tangible for all pulses.Apparently, the fine structure information at high scattering angles remainslargely intact compared to when averaging heterogeneous patterns, even whenusing longer pulses, despite the sample experiencing damage in the form of aCoulomb explosion during the imaging procedure.

Similar effects have been observed in simulations of small crystals [22, 91]and in single particles with an alternative damage model [23], where they have

58

Page 59: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Figure 4.6. Pearson correlation of individual resolution shells. Plots showing howthe correlation coefficient varies with q. Black, purple and green lines are the corre-lations between the average time-integrated diffraction pattern of exploding lysozymeand the average initial structure pattern for three different pulse durations. The remain-ing lines represent the mean of the same analysis carried out with noiseless diffractionpatterns from heterogeneous lysozyme (orange) and noisy patterns of a single struc-ture (grey), with shaded areas indicating one standard deviation above and below eachcalculated value. The pulse parameters used here were identical to those specified inFigure 4.5.

been attributed to pulse gating. The theory is that damage lead to deteriorationof both signal and noise throughout the pulse, resulting in saturation of thesignal-to-noise ratio and diffraction patterns akin to those arising from shorterpulses. To determine whether a similar effect can be seen in our model, wetruncated the temporal sum of diffraction patterns at each time step and calcu-lated the accumulation of SNRtot(q) as described in Section 3.2.1. The time-evolution at q = 6.67 nm−1 (corresponding to 1.5 Å resolution) for all pulsesare shown in Figure 4.7.

A few interesting points can be made from the figure. Firstly, the highestSNRtot at pulse termination is achieved with the shortest pulse at the high-est intensity. Considering that a high intensity facilitates the scattering signalwhile a short duration allows for circumvention of damage, this is not surpris-ing. However, as the pulse duration is increased, damage eventually becomesinevitable and the higher photon count less favorable. In other words, with alonger exposure time, additional photons seemingly do more harm than good.While a short, high-intensity pulse is ideal it is not easily attainable at the nec-essary energies, so the tradeoff between the two parameters should be consid-ered. Secondly, changes to SNRtot diminish during the latter half of the pulses;even for the longest pulse durations it converges to a nonzero value. This im-plies that the structural signal at 1.5 Å resolution is not entirely drowned outby noise in the integrated pattern, and should be recoverable given enoughpatterns for averaging. Following the evolution of any of the plots, we ini-tially see a rapid increase in SNRtot while the sample is still largely intact.

59

Page 60: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Figure 4.7. Signal-to-noise ratio accumulation. SNRtot at q = 6.67 nm−1 calculatedfrom diffraction patterns integrated up to each time point and plotted for the durationof the pulses. A gating effect driven by radiation damage cause the longer-pulse linesto plateau and converge to a nonzero value before the pulses are terminated. Numbersindicate the concluding values. All pulses simulated had a Gaussian temporal profilewith maximum at t = 0 and FWHM durations as shown. The photon energy was fixedat 8 keV and the focal spot size 100 nm in diameter.

As damage starts to manifest, noise contributions eventually catch up with thesignal and the line reaches its peak. At this point the structure factor is becom-ing compromised and noisy scattering overtakes signal, causing a followingSNRtot decline. But somewhat surprisingly, the decline is not entirely detri-mental since it is soon abated and the plot plateaus. For the longest pulses, thishappens when a comparatively large fraction of photons still remain to interactwith the molecule. Such artificial shortening of the pulse is the hallmark of thegating effect mentioned previously, and we can conclude it to be apparent inour model – in agreement with the ionization and ion diffusion approach byMartin et al. [23] – as well.

Lastly, numerical calculations of shot noise and damage noise from time-integrated patterns showed that the former is significantly more prominent thanthe latter. Throughout the full resolution span, disregarding the central speckleregion, shot noise measured at least one order of magnitude higher than corre-sponding damage noise values. Taken together, these results suggest that dam-age might not be as disadvantageous as previously thought. Other sources ofnoise studied here (sample heterogeneity and shot-to-shot fluctuations) seemto have a more profound negative impact on the structure signal, while damageactually invokes a beneficial gating effect. Further investigations are required,in particular on the repercussions of heterogeneous samples where the litera-ture currently is sparse, but the conclusions from Paper II nonetheless supportthe possibility of successful SPI with longer pulses.

60

Page 61: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

4.3 Sample heterogeneity in SPI (Paper III)Expanding upon the preliminary findings regarding sample heterogeneity, thethird paper explores how temperature and sample hydration relates to hetero-geneity and the subsequent impact on image reconstruction in SPI. MD sim-ulations were carried out to represent the structural variability of two proteinsamples (ubiquitin and lysozyme) in three states of hydration (dehydrated, 3 Åwater shell and 6 Å water shell, see example structures in Figure 4.8) at eightdifferent temperatures in increments of 50 K ranging from 0 to 400 K, withthe exclusion of 50 K. Atomic displacement within the proteins was evaluatedby measuring the root-mean-square fluctuations (RMSF) of the constituentatoms throughout the simulations. It is related to the more common RMSD,but instead of measuring the collective deviation of all atoms between twostructures, the RMSF describes the displacement of a single atom from its av-erage position over all simulation frames. Figure 4.8 shows the distributionsof RMSF-values seen in lysozyme at four selected temperatures.

Figure 4.8. Positional fluctuations of atoms. To the left are surface representationsof lysozyme with the different levels of hydration used in the simulations, where greyrepresents the protein and blue indicates water. Density plots to the right show thedistribution of RMSF values of all atoms in the corresponding lysozyme samples.Colors here represent different temperatures. Atomic displacement increases withtemperature, but less so whenever surface-bound water is present. Note that solventmolecules are excluded from the distributions.

Presence of surrounding water suppresses the atomic motion, adding to thestructural stability of the protein [80], which is clearly seen here. As a conse-quence, hydrated samples will exhibit less heterogeneity compared to their drycounterparts, which is beneficial for SPI. However, excess water will simul-taneously contribute significantly to noisy scattering, especially since oxygenis a stronger scatterer than carbon. In order to examine the interplay betweenthese opposing effects, we studied the diffractive capabilities of the samples.

61

Page 62: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Instantaneous and noiseless diffraction patterns were calculated and averagedusing DiffSim from the heterogeneous MD structures, aligned in the samespatial orientation. The resulting patterns of lysozyme are shown in Figure 4.9where a significantly higher speckle contrast can be observed for the hydratedsamples, in particular with the 3 Å water coverage. Speckles become increas-ingly blurred when patterns from samples with a greater structural varianceare averaged, explaining why more contrast is lost at higher temperatures. Inthe case of fully dehydrated samples, such variance corresponds to the hetero-geneity inherent to the protein. Whenever water is present, this still holds true,but may also be an effect of structural changes of the solvent itself aroundthe molecule. If the water molecules are more loosely bound and not well-localized, the variation in their positions will cause additional noise. So whilethe enhanced structural stability of both proteins is comparable between thetwo levels of hydration, it is likely the greater mobility of water molecules inthe 6 Å-samples that hinders signal strength in comparison to the 3 Å-samples.

Figure 4.9. Averaged diffraction patterns of lysozyme. Each logarithmic heatmapcorresponds to the average of 500 noiseless diffraction patterns from heterogeneouslysozyme, generated from 1012 photons at 8 keV uniformly distributed over a circularspot with a diameter of 100 nm. Rows indicate hydration levels and columns the sam-ple temperature. The color scale shows measured intensities as number of photons.

The parameters used in the calculations allowed for diffraction up to 1.4Å resolution. However, with enough noise the attainable resolution will be

62

Page 63: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

lower. We employed the method of Fourier ring correlation (FRC) with afixed-value cutoff – see for instance [113] – to our datasets to assess the rela-tionship between sample heterogeneity and resolution limit. The FRC functionof q was calculated by randomly dividing the set of noiseless patterns into twoequally-sized subsets and correlating the resolution shells of their respectiveaverages. The resolution corresponding to the smallest scattering magnitudeat which FRC(q) = 0.5 was defined as the resolution limit. Obtained reso-lution limits for the different samples and temperatures are plotted in Figure4.10. At temperatures below 200 K, no change is observed and the experi-mental setup remains the bottleneck in terms of resolution when heterogeneityis the only source of noise. But as we go toward higher temperatures, atomicdisplacement increases (Figure 4.8), speckle contrast is lost (Figure 4.9), andour structural data yield a lower-resolution image at best (Figure 4.10). In par-ticular, both dehydrated samples suffer a significant setback with resolutionsrestricted to 2.5−3 Å at 400 K. At room temperature, which is more feasiblefor experimental conditions, lysozyme shows a 20 % gain in resolution uponaddition of a 3 Å water layer. For ubiquitin, the impact is even greater with aresolution-increase from 1.8 to 1.4 Å.

Figure 4.10. Highest attainable resolution in reconstructed 2D image. Resolutionlimits for the various samples and temperatures based on the averaging of 250 indi-vidual diffraction patterns. Note that sample heterogeneity is the only noise sourceconsidered.

Clearly, the stabilizing effects of a surrounding hydration shell offer benefitsto SPI, despite the extraneous noise generated by the water. Optimally theshell should be small, as samples with 6 Å water does not perform as well.We would expect this development to continue as even more water is added,eventually being limited to even lower resolutions than the dry sample, asscattering from water starts to dominate the diffraction pattern.

63

Page 64: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

4.4 Pulse profile effects in SFX (Paper IV)As a result of the SASE process, the time-distribution of photons in an XFELpulse is highly stochastic. How these fluctuations affect the imaging processis not well understood, and therefore the central question in Paper IV. Fourdifferent temporal pulse profiles with photon distributions as shown in Figure4.11 (left pane) were studied by plasma simulations in Cretin designed to rep-resent a SFX experiment. While the variation in flux differed, total fluences,photon energies and durations were unchanged between pulses. Total accumu-lated Bragg peak intensities from a 700 nm crystalline photosystem I sample(Figure 4.11, right pane) was found to attain intensities optimally surpassingnoise whenever the pulse used is front-loaded.

Figure 4.11. Temporal profiles and Bragg intensity accumulation. Simulated pulseshapes and the resulting accretion of Bragg signal from carbon atoms in a 700 nmphotosystem I crystal. [left] The time-distribution of photons for the various pulses, allwith a photon energy of 6 keV and an intensity of 2 ·106 J cm−2. [right] Accumulationof intensity in Bragg peaks as a function of time at momentum transfer q = 0.62 Å−1.Color coding is consistent between both panes. (Figure reproduced from Paper IV.)

If the number of photons arriving early is maximized, they will encounteran undamaged sample and therefore contribute more to the Bragg signal. Con-versely, an initially lower flux will produce less signal and simultaneously on-set ionizing processes that change the structure factor and impair the desiredscattering of impending photons. With the following progression of damage,the pulse will be gated as atoms are displaced and spatial coherence in the crys-tal is lost. The loss in constructive interference of scattered waves terminatesthe accumulation of Bragg peaks, and the integrated signal becomes saturated.Photons interacting with the sample after this point are essentially wasted interms of Bragg diffraction and may contribute to noise. A front-loaded pulse istherefore more suitable for bypassing the effects of both ionization and atomicdisplacement, even if the pulse duration is long (100 fs), as the bulk of photonswill interact with the pristine crystal.

Figure 4.12 shows how ionization and displacement independently affectthe scattering signal from carbon throughout the pulse at momentum trans-

64

Page 65: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

fer q = 0.62 Å−1, given the four different temporal profiles. The functionsshown are normalized such that 1 corresponds to the signal obtained fromthe undamaged crystal and 0 indicates a total signal termination. Atomic dis-placement (left pane) is less affected by pulse shape, where the decay remainsfairly consistent between the four. The small difference that can be observedis expected, since the front-loaded pulse deposits more energy into the sampleduring the first 20 fs of exposure compared to the other pulses. This guaran-tees a rapid plasma formation and a quicker increase in ion temperature andion–ion collision rate, which drive the atomic displacement. By comparison,for a pulse with a single peak in flux between 40 and 60 fs the effects will notbe as immediate, and signal loss from displacement will be delayed.

Figure 4.12. Bragg signal deterioration. Graphs showing the simulated loss indiffractive capability of carbons during a 100 fs pulse of varying temporal profilein a 700 nm photosystem I crystal. Values range between 0 and 1 and correspond tothe fraction of total Bragg intensity obtained at q = 0.62 Å−1 in comparison to that ofa perfect crystal. [left] Decay due to the displacement of atoms and loss of structuralcoherence. [right] Decay due to ionization and consequent altering of the structurefactor. All pulses had the same photon energy (6 keV) and intensity (2 ·106 J cm−2).Plot colors represent different temporal profiles as depicted in the inlet.

Decay due to ionization (right pane in Figure 4.12) is considerably moresensitive to variations in photon flux. A front-loaded pulse seems to be at adisadvantage since the initial heavy X-ray exposure cause a rapid decay indiffractive potential. However, these photons encounter a largely undisturbedcrystal and therefore ensure a substantial contribution to Bragg scattering. Sowhile the diffraction is terminated earlier with this pulse shape, the bulk in-tensity of the pulse has already intercepted the sample once this happens. Itclearly demonstrates the power of the diffraction before destruction principle.In the case of a delayed intensity peak, the situation is reversed. The low ini-tial flux generates little signal while inducing radiation damage that, despiteevolving at a slower rate, undermines the signal yield from the higher-fluxpeak.

The main conclusion of this paper is that the temporal profile indeed hasa significant impact on nanocrystalline imaging. From Figure 4.11 it is evi-dent that the total signal obtained post-pulse varies greatly between profiles.

65

Page 66: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Furthermore, the relation to different scattering angles and photon energies isnot linear. Understanding how these parameters are related, combined withnovel techniques for experimental pulse profiling and/or shaping, will allowfor better predictions of damage dynamics in terms of ionization and atomicdisplacement. In turn, this information can be used to properly rescale col-lected diffraction data to aid the structural determination from nanocrystals.The presented effects should extend to the single particle case where the sameprocesses apply, and likely prove to be particularly relevant as these samplesare especially prone to damage.

4.5 The FreeDam database (Paper V)Results presented in the previous section were based on non-LTE simulationscarried out with the Cretin software. Cretin is described in detail in Section3.3 and monitors the evolution of radiation damage in XFEL-induced plas-mas. Since any material composition can be specified, Cretin is not limitedto only biomolecules and can therefore be aidful in XFEL-based fields otherthan protein imaging. The output encompass average ionization, ion and elec-tron temperatures and atomic displacement as functions of time. FreeDam(free-electron laser damage simulation database), outlined in Paper V, is afree-to-use online resource hosting such data.

Simulation results in FreeDam can both be viewed visually and downloadedas text files for post-processing. An array of samples for a range of XFELpulse parameters have been added to FreeDam thus far, and the selection isplanned to be expanded upon in the future. Most relevant to biology are thecurrently available data files for water, lysozyme and photosystem I, while ex-amples of other materials are graphite, silicon and boron carbide. The usermay define the X-ray pulse in terms of duration, photon energy and total flu-ence and immediately retrieve the simulated data.

As of now, FreeDam does not support a direct pipeline to Cretin for on-demand simulations. Although such functionality is under consideration andwill likely be added in a later update, for now simulations has to be manuallyuploaded by the hosts. Scripts have been developed to perform this task, andwe encourage the scientific community to submit requests of materials andpulse specifications they would like to see added to FreeDam.

The purpose behind FreeDam is to make simulations of radiation damagereadily available to everyone. Access to the theoretical calculations is bene-ficial both for planning and designing XFEL experiments, as well as aidingthe preprocessing of data obtained experimentally. For instance, the evolutionof ionization and temperature in water during an XFEL-pulse presented inFreeDam could be highly relevant to both SFX and SPI, as samples typicallyare delivered to the interaction point in jets or droplets of water. The databasecan be found at https://freedam.desy.de/.

66

Page 67: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

5. Outlook

Structural determination of noncrystalline samples with XFELs at atomic res-olution remains an elusive pursuit. Since the introduction of the diffraction be-fore destruction approach back in 2000, considerable progress has been madeto overcome the many obstacles inherent to SPI. Improved sample deliveryschemes, upgraded laser sources and enhanced algorithms are but a few ofthe recent advances that have brought us closer to this ultimate goal. But keyissues remain and we still have a ways to go.

This thesis aims to address some of the challenges holding SPI back, inparticular those related to radiation damage, given the current capabilities ofultrafast X-ray sources. Granted, some of them could likely be overcome byachieving hard (~10 keV) X-ray pulses in the sub-5 fs duration regime with-out loss of intensity, an undertaking that should therefore remain a priority.However, a viable setup to consistently generate such extreme pulses have yetto be demonstrated, despite several decades of awareness of their potential.Alternative solutions rely not on the prevention of radiation damage, but onunderstanding, or even utilizing, its effects.

All the results presented here are based on theoretical simulations and needto be validated experimentally. This is in itself a challenge, as the neces-sary instrumentation in some cases is lacking. For instance, measuring thefull ion distribution map from a Coulomb explosion would require a sphericaldetector accepting the full 4π solid angle while resolving the molecular orien-tation. Conventional MCPs mounted around the interaction point would onlypartially cover the sphere due to space limitations. While this may be mit-igated by velocity map imaging (VMI), a technique where charged particlesare projected onto a detector using an electrostatic lens [114], the orientationproblem remains unsolved. Similarly, investigating the impact of the temporalpulse profile would require accurate diagnostic tools and pulse-shaping proto-cols. The latter is the main bottleneck as current methods relies on truncatingthe pulse intensity, which at this point is detrimental to SFX and, in particular,SPI. But regardless, these experimental limitations should in no way hinderstudies seeking to complement the in silico observations. Measuring partialion distribution maps or using lower-intensity pulses of well-defined temporalprofile may still provide valuable insight.

When findings are based on theoretical models, their validity directly relieson the accuracy of those models. It is consequently of vital importance thatthe models are constantly evaluated and refined. Adaptation to better reflectexperiments, as discussed in the previous paragraph, is one way. Another is

67

Page 68: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

to adjust the approximations made to emulate the true underlying physics to agreater degree. In the context of MD simulations this could mean the inclusionof quantum effects, which is becoming increasingly feasible with the rapidadvancement of computational power.

To a large extent, papers I–IV each focus on a singular issue pertaining toSPI and SFX, while neglecting others. Isolating a specific part of a greaterquestion certainly allows for a deeper investigation of that particular aspect,but runs the risk of jeopardizing the complete picture overview. In the worstcase, the results may even no longer remain true to the initial problem. For thisreason, it is important that future studies both (i) treat the issues covered heresimultaneously, and (ii) combine them with other known aspects of SPI andSFX. One example could be to examine sample heterogeneity concurrentlywith varying pulse profiles, another to include the resolution-limiting effectsof orientation recovery in addition to the other noise sources considered inPaper II. Ultimately, to enable continuous development of these promisingtechniques, the aforementioned complete picture needs to be established.

Lastly, data accessibility is essential. To maximize the likelihood of reach-ing atomic resolution with SPI, and to further improve SFX, all relevant find-ings need to be available to researchers worldwide. A good example is thecoherent X-ray imaging data bank (CXIDB) [115] where results from diffrac-tive measurements at XFELs are stored. The same holds true for simulationresults. Ideally, developed software should be open source and numerical datashould be stored in databases to be expanded upon. FreeDam of Paper V isbut one attempt to establish a representative of the latter. The more informa-tion we can collect, the more sophisticated our future pursuit, and subsequentprogress, will be.

Personally, I believe SPI to be well within reach and its capabilities to bevirtually limitless. In particular, the prospect of imaging proteins that so farhave eluded structural studies is indeed exciting. I have little doubt that oncewe get there, this novel technique will revolutionize structural biology andmedical sciences in an unprecedented manner. I would even go as far as tosay that it will become one of the most important scientific discoveries of thecentury.

And we are getting ever closer.

68

Page 69: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

6. Acknowledgements

My most profound gratitude is extended toward all the people that made thisundertaking possible. I deeply thank my supervisors Calle and Nic for theirpatience, guidance and support. Your help has truly been imperative and I feelprivileged to call you my friends. Now, and for many years to come.

Knowledge is never single-handedly held, so I would like to acknowledgeeveryone who has shared me theirs during my time in the division. To myeminent fellow PhD students – those who fearlessly left their countries in thepursuit of science, Rameez, Clara, Geethu, Eva and Deepak, and those whoextended their stay in native Sweden, Fredrik, Sebastian and Ludde – youreally are amazing. To the researchers and postdocs who through their kindopenness have contributed to countless interesting conversations and a veryfriendly environment. Fredrik, Robert, Isaak, Konstantin and Ronny, youmade this journey unforgettable. To the seniors who gave valuable feedback,your insight has unquestionably been crucial. In particular, thank you Olle forhelping out as external reviewer of my licentiate thesis and for your positiveencouragement within our group. To the rest of that same group, including theamazing people who has since left – may your future endeavors prove to berewarding, Olof, Davide, Rebecka and Nicklas. And to the colleagues whotruly should be mentioned, but regrettably slip my mind at the moment.

Financial, technical and instructional support are indispensable to any science-oriented institution. Without the generous economical aid from the Swedishresearch infrastructures SSF, STINT and the Carl Trygger foundation; theensued Swedish–German joint collaboration cluster RÅC; the German ad-vanced laser science center CFEL, the Helmholtz association and the long-established Volkswagen stiftung; as well as the Australian national linkageresearch program ARC centres of excellence, this work would not have seenthe light of day. Instrumental was also UPPMAX, the Uppsala universityhigh-performance computing resource, provided by SNIC. Finally, my mostearnest appreciation goes out to the numerous collaborators who assisted me.Obviously, Tomas, Andrew, Thomas, Howard, Henry and Steve have theirnames on the papers already, but they still deserve another mention. And toexpand further on this list: thank you Erik, Gösta, Alex and Harry.

69

Page 70: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Though particularly, I am forever indebted to those whose invaluable supporthas extended beyond the workplace. To Benedikt and Max at BMC, Justine

and Daniel in Melbourne, Salah and Max in Hamburg; thank you. I will alsonever forget Micke, Emil, Charlie, Julia, Hillevi and the numerous otherkindhearted Swedes in Australia. I thank my family, despite their consistentyawning whenever physics is brought up... Thank you Jessika, Ida and Tove,outstanding friends and life advisors, and Ebba, my personal sea urchin andunderstanding partner in crime. I love you all.

Most of all, to my late father Tommy.Always missed, never forgotten.

"Man ska aldrig ge upp."

70

Page 71: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

7. Sammanfattning på svenska

Proteiner är en grupp av molekyler som ofta kallas för livets byggstenar. An-ledningen är att dessa molekyler är intimt sammankopplade till de cellulärafunktioner som utförs i en levande organism. Immunförsvar, celldelning ochenergiproduktion är bara några exempel på processer där proteiner har en ny-ckelroll. Liv skulle helt enkelt inte kunna existera utan dem. Trots att de fyllerså otroligt många och varierade funktioner består alla proteiner av samma20 grundenheter. Dessa kallas för aminosyror. Varje protein består av enlång kedja av aminosyror som kan variera i längd, och det är den specifikasekvensen av aminosyror som skiljer mellan olika proteiner. Detta ger upphovtill en enorm diversitet, fastän antalet unika aminosyror är tämligen begränsat.

Varje protein har en tredimensionell struktur som uppstår när aminosyraked-jan veckas och binder till sig själv. Strukturen, som brukar kallas proteinetstertiärstruktur, är karaktäristisk för varje protein och av central betydelse fördess funktion. Så för att fullständigt förstå hur ett visst protein fungerar måstevi veta hur det ser ut – information som sedan kan användas för utveckling avexempelvis nya läkemedel.

Även om proteiner relativt sett är stora molekyler är de fortfarande alldelesför små för att urskiljas i ett vanligt ljusmikroskop. Därför krävs andra metoderför att bestämma tertiärstrukturen. Ett flertal tekniker finns tillgängliga, menden hittills vanligaste och mest framgångsrika heter röntgenkristallografi. Somnamnet antyder används röntgenljus för att "fotografera" proteinet, som förstmåste kristalliserats. En proteinkristall består av många identiska kopior avproteinet som är placerade på ett ordnat vis i förhållande till varandra. Närröntgenljuset träffar kristallen så sprids det och ger upphov till ett så kallatdiffraktionsmönster. Mönstret är direkt kopplat till proteinets struktur ochrepresenterar alltså en tvådimensionell bild av molekylen (likt ett fotografi).Genom att rotera kristallen samtidigt som den bestrålas kan många diffrak-tionsmönster samlas in från olika vinklar, vilket gör det möjligt att återskapahela proteinet i tre dimensioner.

Den största nackdelen med röntgenkristollografi är behovet av en kristall.Många proteiner är mycket svåra att kristallisera; i vissa fall är det endastmöjligt att framställa mycket små kristaller, i andra fall går det inte alls. Ochju mindre kristallen är, desto otydligare blir diffraktionsmönstret. Teoretisktsett skulle man kunna göra mönstret från en liten kristall tydligare genom attöka röntgenljusets intensitet, men eftersom strålningen är joniserande lederdet samtidigt till att provet snabbt förstörs. Så är hoppet helt ute för att avbildadessa mer problematiska proteiner?

71

Page 72: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Inte nödvändigtvis. Under de senaste årtiondena har en ny typ av rönt-genkällor utvecklats. Frielektronlasrar, som de kallas, har kapaciteten att gener-era mycket korta pulser av intensiv röntgenstrålning. Pulsens intensitet är såhög att den kompenserar för provets ringa storlek, samtidigt som den är så ko-rtvarig att provet helt enkelt inte hinner gå sönder innan ett diffraktionsmönsterhar uppmätts. Visserligen fås endast ett tvådimensionellt mönster från provetinnan den förstörs, men genom att kontinuerligt ersätta det med ett nytt kandiffraktionsmönster från olika vinklar erhållas. Resultatet är en fullständig up-psättning mönster för att återskapa strukturen i 3D, precis som när kristallenroteras i konventionell röntgenkristallografi.

Tillvägagångssättet har kommit att döpas till serial femtosecond crystallog-raphy (SFX) och har visat sig fungera väl för proteinkristaller som tidigare harvarit för små. Namnet kommer av tidsskalan för laserpulserna som typiskt settär under 100 femtosekunder. De är alltså otroligt korta – en femtosekund ären miljondels miljarddels sekund och förhåller sig till sekund som en sekundförhåller sig till 32 miljoner år. Förhoppningen är att SFX ska kunna utveck-las för att hantera mindre och mindre kristaller, för att eventuellt eliminerabehovet av kristallisering helt. Provet skulle då bestå av en enstaka protein-molekyl och metoden kallas därmed för single particle imaging (SPI).

Båda dessa tekniker är under pågående utveckling och behöver förbättrasbetydligt innan de helt kan ersätta den vanliga röntgenkristallografin. Fram-förallt SPI står inför många utmaningar som måste övervinnas. Den här avhan-dlingen presenterar totalt fem artiklar med forskningsresultat baserade på simu-leringsstudier, och syftar till att bemöta följande av dessa utmaningar:

- Molekylens orientering vid avbildningstillfället.- Brus i diffraktionsmönstret från strålskada och heterogena strukturer.- Pulsens tidsberoende.

Artikel I undersöker spridningen av joner när ett enstaka protein exploderarfrån att ha exponerats för en ultrakort röntgenpuls. Vid SPI injiceras provetsom ska avbildas i en slumpmässig orientering, vilket gör det svårt att avgörafrån vilken vinkel diffraktionsbilden är tagen. Vi fann att explosionen i höggrad är reproducerbar och därför kan ge information om just orienteringen närlaserpulsen träffar. Detektion av jonfragmenten skulle därför kunna under-lätta avsevärt vid hopsättningen av diffraktionsbilder and därmed effektiviseraavbildningsprocessen.

Artikel II och III utforskar problematiken med brus i diffraktionsmönstrenfrån enskilda proteiner. Även om de röntgenpulser som finns tillgängliga idagär väldigt korta så är exponeringstiden fortfarande för lång för att helt un-vika laserns skadande inverkan. Att provets struktur förändras något underavbildningen kommer således att reflekteras i diffraktionsmönstret i form avbrus. Detta kan jämföras med att ta ett forografi av något i rörelse med enlång slutartid – bilden blir suddig. Brus uppstår också när diffraktionsbilderna

72

Page 73: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

kombineras om de olika kopiorna av proteinet inte har varit helt identiska,vilket typiskt sett är fallet. Artiklarna fokuserar på båda dessa källor till brusmed syfte att avgöra hur stor inverkan de har för strukturbestämningen. Dethuvudsakliga resultatet var att strukturell varians bland proverna tycks utgöraett större problem än den nuvarande pulslängden för SPI, eftersom strålskadanvirtuellt förkortar exponeringstiden genom en så kallad "gating"-effekt.

Artikel IV granskar hur pulsens fördelning av ljuspartiklar över tid påverkaravbildningen. Frielektronlaserpulser har typiskt en slumpmässig tidsfödel-ning av fotoner, vilket misstänks ha en betydande inverkan på det resulterandediffraktionsmönstret. Våra simuleringar visade att skadeutvecklingen i provetvarierar kraftigt med pulsens tidsprofil, även om det totala antalet fotoner ochpulslängden är oförändrade. Det leder vidare till kvalitetsskillnader i diffrak-tionsmönstret. Optimala förhållanden erhålls om majoriteten av fotoner ärfördelade tidigt i pulsen, då provet fortfarande är intakt.

Artikel V presenterar en webbdatabas med simuleringsresultat för diversematerial. Simuleringarna beskriver strålskadeutvecklingen när materialet ut-sätts för en röntgenpuls med given intensitet, energi och varaktighet, och för-modas vara av stor nytta till forskare välden över. Tillgänglig data är i nulägetaningen begränsat, men planeras att utökas i framtiden.

Resultaten utgör små men väsentliga aspekter för utvecklingen av SPI och,till viss del, SFX. Även om många utmaningar kvarstår så har vi redan kommitlångt i jakten efter kristallfri avbildning med röntgenljus. Och jakten fortsätter,för få potentiella metoder utlovar en så omfattande belöning som SPI. Väl därkommer den sannolikt att revolutionera hela den medicinska vetenskapen, ochtvivellöst utgöra en av de absolut mest betydelsfulla genombrotten i moderntid.

73

Page 74: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

8. References

[1] R. Neutze, R. Wouts, D. van der Spoel, E. Weckert, and J. Hajdu, “Potentialfor biomolecular imaging with femtosecond X-ray pulses,” Nature 406,752–757, Aug. 2000.

[2] J. M. J. Madey, “Stimulated Emission of Bremsstrahlung in a PeriodicMagnetic Field,” Journal of Applied Physics 42, 1906–1913, Apr. 1971.

[3] Z. Jurek, G. Faigel, and M. Tegze, “Dynamics in a Cluster under the Influenceof Intense Femtosecond Hard X-Ray Pulses,” The European Physical JournalD - Atomic, Molecular and Optical Physics 29, 217–229, May 2004.

[4] M. Bergh, N. Tîmneanu, and D. van der Spoel, “Model for the Dynamics of aWater Cluster in an X-Ray Free Electron Laser Beam,” Physical Review E 70,051904, Nov. 2004.

[5] H. N. Chapman, A. Barty, M. J. Bogan, S. Boutet, M. Frank, S. P. Hau-Riege,S. Marchesini, B. W. Woods, S. Bajt, W. H. Benner, R. A. London, E. Plönjes,M. Kuhlmann, R. Treusch, S. Düsterer, T. Tschentscher, J. R. Schneider,E. Spiller, T. Möller, C. Bostedt, M. Hoener, D. A. Shapiro, K. O. Hodgson,D. van der Spoel, F. Burmeister, M. Bergh, C. Caleman, G. Huldt, M. M.Seibert, F. R. N. C. Maia, R. W. Lee, A. Szöke, N. Timneanu, and J. Hajdu,“Femtosecond Diffractive Imaging with a Soft-X-Ray Free-Electron Laser,”Nature Physics 2, 839–843, Dec. 2006.

[6] P. Emma, R. Akre, J. Arthur, R. Bionta, C. Bostedt, J. Bozek, A. Brachmann,P. Bucksbaum, R. Coffee, F.-J. Decker, Y. Ding, D. Dowell, S. Edstrom,A. Fisher, J. Frisch, S. Gilevich, J. Hastings, G. Hays, P. Hering, Z. Huang,R. Iverson, H. Loos, M. Messerschmidt, A. Miahnahri, S. Moeller, H.-D.Nuhn, G. Pile, D. Ratner, J. Rzepiela, D. Schultz, T. Smith, P. Stefan,H. Tompkins, J. Turner, J. Welch, W. White, J. Wu, G. Yocky, and J. Galayda,“First Lasing and Operation of an Ångstrom-Wavelength Free-Electron Laser,”Nature Photonics 4, 641–647, Sept. 2010.

[7] H. N. Chapman, P. Fromme, A. Barty, T. A. White, R. A. Kirian, A. Aquila,M. S. Hunter, J. Schulz, D. P. DePonte, U. Weierstall, R. B. Doak, F. R. N. C.Maia, A. V. Martin, I. Schlichting, L. Lomb, N. Coppola, R. L. Shoeman,S. W. Epp, R. Hartmann, D. Rolles, A. Rudenko, L. Foucar, N. Kimmel,G. Weidenspointner, P. Holl, M. Liang, M. Barthelmess, C. Caleman,S. Boutet, M. J. Bogan, J. Krzywinski, C. Bostedt, S. Bajt, L. Gumprecht,B. Rudek, B. Erk, C. Schmidt, A. Hömke, C. Reich, D. Pietschner, L. Strüder,G. Hauser, H. Gorke, J. Ullrich, S. Herrmann, G. Schaller, F. Schopper,H. Soltau, K.-U. Kühnel, M. Messerschmidt, J. D. Bozek, S. P. Hau-Riege,M. Frank, C. Y. Hampton, R. G. Sierra, D. Starodub, G. J. Williams, J. Hajdu,N. Timneanu, M. M. Seibert, J. Andreasson, A. Rocker, O. Jönsson,M. Svenda, S. Stern, K. Nass, R. Andritschke, C.-D. Schröter, F. Krasniqi,M. Bott, K. E. Schmidt, X. Wang, I. Grotjohann, J. M. Holton, T. R. M.

74

Page 75: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Barends, R. Neutze, S. Marchesini, R. Fromme, S. Schorb, D. Rupp,M. Adolph, T. Gorkhover, I. Andersson, H. Hirsemann, G. Potdevin,H. Graafsma, B. Nilsson, and J. C. H. Spence, “Femtosecond X-ray proteinnanocrystallography,” Nature 470, 73–77, Feb. 2011.

[8] M. M. Seibert, T. Ekeberg, F. R. N. C. Maia, M. Svenda, J. Andreasson,O. Jönsson, D. Odic, B. Iwan, A. Rocker, D. Westphal, M. Hantke, D. P.DePonte, A. Barty, J. Schulz, L. Gumprecht, N. Coppola, A. Aquila, M. Liang,T. A. White, A. Martin, C. Caleman, S. Stern, C. Abergel, V. Seltzer, J.-M.Claverie, C. Bostedt, J. D. Bozek, S. Boutet, A. A. Miahnahri,M. Messerschmidt, J. Krzywinski, G. Williams, K. O. Hodgson, M. J. Bogan,C. Y. Hampton, R. G. Sierra, D. Starodub, I. Andersson, S. Bajt,M. Barthelmess, J. C. H. Spence, P. Fromme, U. Weierstall, R. Kirian,M. Hunter, R. B. Doak, S. Marchesini, S. P. Hau-Riege, M. Frank, R. L.Shoeman, L. Lomb, S. W. Epp, R. Hartmann, D. Rolles, A. Rudenko,C. Schmidt, L. Foucar, N. Kimmel, P. Holl, B. Rudek, B. Erk, A. Hömke,C. Reich, D. Pietschner, G. Weidenspointner, L. Strüder, G. Hauser, H. Gorke,J. Ullrich, I. Schlichting, S. Herrmann, G. Schaller, F. Schopper, H. Soltau,K.-U. Kühnel, R. Andritschke, C.-D. Schröter, F. Krasniqi, M. Bott, S. Schorb,D. Rupp, M. Adolph, T. Gorkhover, H. Hirsemann, G. Potdevin, H. Graafsma,B. Nilsson, H. N. Chapman, and J. Hajdu, “Single Mimivirus ParticlesIntercepted and Imaged with an X-Ray Laser,” Nature 470, 78–81, Feb. 2011.

[9] T. Ekeberg, M. Svenda, C. Abergel, F. R. N. C. Maia, V. Seltzer, J.-M.Claverie, M. Hantke, O. Jönsson, C. Nettelblad, G. van der Schot, M. Liang,D. P. DePonte, A. Barty, M. M. Seibert, B. Iwan, I. Andersson, N. D. Loh,A. V. Martin, H. Chapman, C. Bostedt, J. D. Bozek, K. R. Ferguson,J. Krzywinski, S. W. Epp, D. Rolles, A. Rudenko, R. Hartmann, N. Kimmel,and J. Hajdu, “Three-Dimensional Reconstruction of the Giant MimivirusParticle with an X-Ray Free-Electron Laser,” Physical Review Letters 114,098102, Mar. 2015.

[10] M. Gallagher-Jones, Y. Bessho, S. Kim, J. Park, S. Kim, D. Nam, C. Kim,Y. Kim, D. Y. Noh, O. Miyashita, F. Tama, Y. Joti, T. Kameshima, T. Hatsui,K. Tono, Y. Kohmura, M. Yabashi, S. S. Hasnain, T. Ishikawa, and C. Song,“Macromolecular structures probed by combining single-shot free-electronlaser diffraction with synchrotron coherent X-ray imaging,” NatureCommunications 5, 3798, Dec. 2014.

[11] M. F. Hantke, D. Hasse, F. R. N. C. Maia, T. Ekeberg, K. John, M. Svenda,N. D. Loh, A. V. Martin, N. Timneanu, D. S. D. Larsson, G. van der Schot,G. H. Carlsson, M. Ingelman, J. Andreasson, D. Westphal, M. Liang,F. Stellato, D. P. DePonte, R. Hartmann, N. Kimmel, R. A. Kirian, M. M.Seibert, K. Mühlig, S. Schorb, K. Ferguson, C. Bostedt, S. Carron, J. D.Bozek, D. Rolles, A. Rudenko, S. Epp, H. N. Chapman, A. Barty, J. Hajdu,and I. Andersson, “High-throughput imaging of heterogeneous cell organelleswith an X-ray laser,” Nature Photonics 8, 943–949, Dec. 2014.

[12] G. van der Schot, M. Svenda, F. R. N. C. Maia, M. Hantke, D. P. DePonte,M. M. Seibert, A. Aquila, J. Schulz, R. Kirian, M. Liang, F. Stellato, B. Iwan,J. Andreasson, N. Timneanu, D. Westphal, F. N. Almeida, D. Odic, D. Hasse,G. H. Carlsson, D. S. D. Larsson, A. Barty, A. V. Martin, S. Schorb,

75

Page 76: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

C. Bostedt, J. D. Bozek, D. Rolles, A. Rudenko, S. Epp, L. Foucar, B. Rudek,R. Hartmann, N. Kimmel, P. Holl, L. Englert, N.-T. Duane Loh, H. N.Chapman, I. Andersson, J. Hajdu, and T. Ekeberg, “Imaging single cells in abeam of live cyanobacteria with an X-ray laser,” Nature Communications 6,5704, Dec. 2015.

[13] L. C. Johansson, B. Stauch, A. Ishchenko, and V. Cherezov, “A Bright Futurefor Serial Femtosecond Crystallography with XFELs,” Trends in BiochemicalSciences 42, 749–762, Sept. 2017.

[14] T. Masuda, M. Suzuki, S. Inoue, C. Song, T. Nakane, E. Nango, R. Tanaka,K. Tono, Y. Joti, T. Kameshima, T. Hatsui, M. Yabashi, B. Mikami, O. Nureki,K. Numata, S. Iwata, and M. Sugahara, “Atomic resolution structure of serineprotease proteinase K at ambient temperature,” Scientific Reports 7, 45604,Dec. 2017.

[15] A. Aquila, A. Barty, C. Bostedt, S. Boutet, G. Carini, D. dePonte, P. Drell,S. Doniach, K. H. Downing, T. Earnest, H. Elmlund, V. Elser, M. Gühr,J. Hajdu, J. Hastings, S. P. Hau-Riege, Z. Huang, E. E. Lattman, F. R. N. C.Maia, S. Marchesini, A. Ourmazd, C. Pellegrini, R. Santra, I. Schlichting,C. Schroer, J. C. H. Spence, I. A. Vartanyants, S. Wakatsuki, W. I. Weis, andG. J. Williams, “The Linac Coherent Light Source Single Particle ImagingRoad Map,” Structural Dynamics 2, 041701, July 2015.

[16] R. P. Kurta, J. J. Donatelli, C. H. Yoon, P. Berntsen, J. Bielecki, B. J. Daurer,H. DeMirci, P. Fromme, M. F. Hantke, F. R. N. C. Maia, A. Munke,C. Nettelblad, K. Pande, H. K. N. Reddy, J. A. Sellberg, R. G. Sierra,M. Svenda, G. van der Schot, I. A. Vartanyants, G. J. Williams, P. L. Xavier,A. Aquila, P. H. Zwart, and A. P. Mancuso, “Correlations in Scattered X-RayLaser Pulses Reveal Nanoscale Structural Features of Viruses,” PhysicalReview Letters 119, Oct. 2017.

[17] D. P. DePonte, U. Weierstall, K. Schmidt, J. Warner, D. Starodub, J. C. H.Spence, and R. B. Doak, “Gas Dynamic Virtual Nozzle for Generation ofMicroscopic Droplet Streams,” Journal of Physics D: Applied Physics 41,195505, Oct. 2008.

[18] P. Liu, P. J. Ziemann, D. B. Kittelson, and P. H. McMurry, “Generating ParticleBeams of Controlled Dimensions and Divergence: I. Theory of ParticleMotion in Aerodynamic Lenses and Nozzle Expansions,” Aerosol Science andTechnology 22, 293–313, Jan. 1995.

[19] R. A. Kirian, S. Awel, N. Eckerskorn, H. Fleckenstein, M. Wiedorn,L. Adriano, S. Bajt, M. Barthelmess, R. Bean, K. R. Beyerlein, L. M. G.Chavas, M. Domaracky, M. Heymann, D. A. Horke, J. Knoska, M. Metz,A. Morgan, D. Oberthuer, N. Roth, T. Sato, P. L. Xavier, O. Yefanov, A. V.Rode, J. Küpper, and H. N. Chapman, “Simple Convergent-Nozzle AerosolInjector for Single-Particle Diffractive Imaging with X-Ray Free-ElectronLasers,” Structural Dynamics 2, 041717, July 2015.

[20] N.-T. D. Loh and V. Elser, “Reconstruction algorithm for single-particlediffraction imaging experiments,” Physical Review E 80, 026705, Aug. 2009.

[21] E. G. Marklund, T. Ekeberg, M. Moog, J. L. Benesch, and C. Carleman,“Controlling Protein Orientation in Vacuum Using Electric Fields,” TheJournal of Physical Chemistry Letters Sept. 2017.

76

Page 77: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

[22] C. Caleman, N. Tîmneanu, A. V. Martin, H. O. Jönsson, A. Aquila, A. Barty,H. A. Scott, T. A. White, and H. N. Chapman, “Ultrafast self-gating Braggdiffraction of exploding nanocrystals in an X-ray laser,” Optics Express 23,1213, Jan. 2015.

[23] A. V. Martin, J. K. Corso, C. Caleman, N. Timneanu, and H. M. Quiney,“Single-molecule imaging with longer X-ray laser pulses,” IUCrJ 2, 661–674,Nov. 2015.

[24] F. R. N. C. Maia, T. Ekeberg, N. Tîmneanu, D. van der Spoel, and J. Hajdu,“Structural variability and the incoherent addition of scattered intensities insingle-particle diffraction,” Physical Review E 80, Sept. 2009.

[25] A. V. Martin and H. M. Quiney, “Coherence loss by sample dynamics andheterogeneity in x-ray laser diffraction,” Journal of Physics B: Atomic,Molecular and Optical Physics 49, 244001, Dec. 2016.

[26] I. Grguraš, A. R. Maier, C. Behrens, T. Mazza, T. J. Kelly, P. Radcliffe,S. Düsterer, A. K. Kazansky, N. M. Kabachnik, T. Tschentscher, J. T. Costello,M. Meyer, M. C. Hoffmann, H. Schlarb, and A. L. Cavalieri, “Ultrafast X-raypulse characterization at free-electron lasers,” Nature Photonics 6, 852–857,Dec. 2012.

[27] W. Helml, A. R. Maier, W. Schweinberger, I. Grguraš, P. Radcliffe, G. Doumy,C. Roedig, J. Gagnon, M. Messerschmidt, S. Schorb, C. Bostedt, F. Grüner,L. F. DiMauro, D. Cubaynes, J. D. Bozek, T. Tschentscher, J. T. Costello,M. Meyer, R. Coffee, S. Düsterer, A. L. Cavalieri, and R. Kienberger,“Measuring the temporal structure of few-femtosecond free-electron laserX-ray pulses directly in the time domain,” Nature Photonics 8, 950–957, Dec.2014.

[28] A. Marinelli, R. Coffee, S. Vetter, P. Hering, G. N. West, S. Gilevich, A. A.Lutman, S. Li, T. Maxwell, J. Galayda, A. Fry, and Z. Huang, “OpticalShaping of X-Ray Free-Electron Lasers,” Physical Review Letters 116,254801, June 2016.

[29] W. C. Rontgen, “On a New Kind of Rays,” Science 3, 227–231, Feb. 1896.[30] J. D. Watson and F. H. C. Crick, “Molecular Structure of Nucleic Acids: A

Structure for Deoxyribose Nucleic Acid,” Nature 171, 737–738, Apr. 1953.[31] D. C. Hodgkin, J. Pickworth, J. H. Robertson, K. N. Trueblood, R. J. Prosen,

and J. G. White, “Structure of Vitamin B12 : The Crystal Structure of theHexacarboxylic Acid derived from B12 and the Molecular Structure of theVitamin,” Nature 176, 325–328, Aug. 1955.

[32] B. Schechter, “Nobel Prize in Chemistry to Hauptman and Karle,” PhysicsToday 38, 20–21, Dec. 1985.

[33] E. Gouaux, “Principles of Selective Ion Transport in Channels and Pumps,”Science 310, 1461–1465, Dec. 2005.

[34] A. G. Pockley, “Heat Shock Proteins as Regulators of the Immune Response,”The Lancet 362, 469–476, Aug. 2003.

[35] W. Cho and R. V. Stahelin, “Membrane-Protein Interactions in Cell Signalingand Membrane Trafficking,” Annual Review of Biophysics and BiomolecularStructure 34, 119–151, June 2005.

[36] T. J. Mitchison and E. D. Salmon, “Mitosis: A History of Division,” NatureCell Biology 3, E17–E21, Jan. 2001.

77

Page 78: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

[37] R. D. Kornberg, “The molecular basis of eukaryotic transcription,”Proceedings of the National Academy of Sciences 104, 12955–12961, Aug.2007.

[38] H. M. Berman, John Westbrook, Zukang Feng, Gary Gilliland, T. N. Bhat,Helge Weissig, Ilya N. Shindyalov, and Philip E. Bourne, “The Protein DataBank,” Nucleic Acids Research 28, 235–242, Jan. 2000.

[39] C. B. Anfinsen, “Principles that Govern the Folding of Protein Chains,”Science 181, 223–230, July 1973.

[40] K. Wuthrich, “Protein Structure Determination in Solution by NuclearMagnetic Resonance Spectroscopy,” Science 243, 45–50, Jan. 1989.

[41] Y. Cheng, N. Grigorieff, P. A. Penczek, and T. Walz, “A Primer toSingle-Particle Cryo-Electron Microscopy,” Cell 161, 438–449, Apr. 2015.

[42] N. Jones, “Crystallography: Atomic secrets,” Nature 505, 602–603, Jan. 2014.[43] D. P. Frueh, A. C. Goodrich, S. H. Mishra, and S. R. Nichols, “NMR methods

for structural studies of large monomeric and multimeric proteins,” CurrentOpinion in Structural Biology 23, 734–739, Oct. 2013.

[44] X.-c. Bai, I. S. Fernandez, G. McMullan, and S. H. Scheres, “Ribosomestructures to near-atomic resolution from thirty thousand cryo-EM particles,”eLife 2, e00461, Feb. 2013.

[45] M. Liao, E. Cao, D. Julius, and Y. Cheng, “Structure of the TRPV1 ion channeldetermined by electron cryo-microscopy,” Nature 504, 107–112, Dec. 2013.

[46] M. Khoshouei, R. Danev, J. M. Plitzko, and W. Baumeister, “Revisiting theStructure of Hemoglobin and Myoglobin with Cryo-Electron Microscopy,”Journal of Molecular Biology 429, 2611–2618, Aug. 2017.

[47] W. H. Bragg and W. L. Bragg, “The Reflection of X-rays by Crystals,”Proceedings of the Royal Society A: Mathematical, Physical and EngineeringSciences 88, 428–438, July 1913.

[48] M. Bergh, G. Huldt, N. Tîmneanu, F. R. N. C. Maia, and J. Hajdu, “Feasibilityof imaging living cells at subnanometer resolutions by ultrafast X-raydiffraction,” Quarterly Reviews of Biophysics 41, 181, Nov. 2008.

[49] C. Giacovazzo, “Direct Methods and Powder Data: State of the Art andPerspectives,” Acta Crystallographica Section A Foundations ofCrystallography 52, 331–339, May 1996.

[50] P. P. Ewald, “Introduction to the dynamical theory of X-ray diffraction,” ActaCrystallographica Section A: Crystal Physics, Diffraction, Theoretical andGeneral Crystallography 25, 103–108, Jan. 1969.

[51] J. M. Holton and K. A. Frankel, “The minimum crystal size needed for acomplete diffraction data set,” Acta Crystallographica Section D BiologicalCrystallography 66, 393–408, Apr. 2010.

[52] A. Krogh, B. Larsson, G. von Heijne, and E. L. Sonnhammer, “Predictingtransmembrane protein topology with a hidden markov model: Application tocomplete genomes11Edited by F. Cohen,” Journal of Molecular Biology 305,567–580, Jan. 2001.

[53] L. Tiefenauer and S. Demarche, “Challenges in the Development of FunctionalAssays of Membrane Proteins,” Materials 5, 2205–2242, Nov. 2012.

[54] S. P. Hau-Riege, High-Intensity X-Rays-Interaction with Matter: Processes inPlasmas, Clusters, Molecules and Solids. John Wiley & Sons, 2012.

78

Page 79: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

[55] A. Einstein, “Über einen die Erzeugung und Verwandlung des Lichtesbetreffenden heuristischen Gesichtspunkt,” Annalen der Physik 17, 132–148,1905.

[56] S. Seltzer, “XCOM-Photon Cross Sections Database, NIST StandardReference Database 8,” 1987.

[57] M. Ossiander, J. Riemensberger, S. Neppl, M. Mittermair, M. Schäffer,A. Duensing, M. S. Wagner, R. Heider, M. Wurzer, M. Gerl,M. Schnitzenbaumer, J. V. Barth, F. Libisch, C. Lemell, J. Burgdörfer,P. Feulner, and R. Kienberger, “Absolute timing of the photoelectric effect,”Nature 561, 374–377, Sept. 2018.

[58] O. Hardouin Duparc, “Pierre Auger – Lise Meitner: ComparativeContributions to the Auger Effect,” International Journal of MaterialsResearch 100, 1162–1166, Sept. 2009.

[59] P. Auger, “Secondary β -Rays Produced in a Gas by X-Rays,” Comptes RendusAcad. Sci. Paris 177, 169, 1923.

[60] S. P. Hau-Riege, “Nonequilibrium electron dynamics in materials driven byhigh-intensity x-ray pulses,” Physical Review E 87, May 2013.

[61] E. F. Garman, “Radiation Damage in Macromolecular Crystallography: WhatIs It and Why Should We Care?,” Acta Crystallographica Section D BiologicalCrystallography 66, 339–351, Apr. 2010.

[62] C. Caleman, G. Huldt, F. R. N. C. Maia, C. Ortiz, F. G. Parak, J. Hajdu, D. vander Spoel, H. N. Chapman, and N. Timneanu, “On the Feasibility ofNanocrystal Imaging Using Intense and Ultrashort X-ray Pulses,” ACS Nano 5,139–146, Jan. 2011.

[63] S. P. Hau-Riege, R. A. London, and A. Szoke, “Dynamics of BiologicalMolecules Irradiated by Short X-Ray Pulses,” Physical Review E 69, 051906,May 2004.

[64] A. Gonzalez, A. W. Thompson, and C. Nave, “Cryo-protection of ProteinCrystals in Intense X-ray Beams,” Review of Scientific Instruments 63,1177–1180, Jan. 1992.

[65] R. L. Owen, E. Rudino-Pinera, and E. F. Garman, “Experimentaldetermination of the radiation dose limit for cryocooled protein crystals,”Proceedings of the National Academy of Sciences 103, 4912–4917, Mar. 2006.

[66] H. N. Chapman, C. Caleman, and N. Timneanu, “Diffraction beforedestruction,” Philosophical Transactions of the Royal Society B: BiologicalSciences 369, 20130313–20130313, June 2014.

[67] C. Bressler and M. Chergui, “Ultrafast X-Ray Absorption Spectroscopy,”Chemical Reviews 104, 1781–1812, Apr. 2004.

[68] T. Popmintchev, M.-C. Chen, P. Arpin, M. M. Murnane, and H. C. Kapteyn,“The attosecond nonlinear optics of bright coherent X-ray generation,” NaturePhotonics 4, 822–832, Dec. 2010.

[69] C. Bostedt, S. Boutet, D. M. Fritz, Z. Huang, H. J. Lee, H. T. Lemke,A. Robert, W. F. Schlotter, J. J. Turner, and G. J. Williams, “Linac CoherentLight Source: The First Five Years,” Reviews of Modern Physics 88, 015007,Mar. 2016.

[70] T. Ishikawa, H. Aoyagi, T. Asaka, Y. Asano, N. Azumi, T. Bizen, H. Ego,K. Fukami, T. Fukui, Y. Furukawa, S. Goto, H. Hanaki, T. Hara, T. Hasegawa,

79

Page 80: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

T. Hatsui, A. Higashiya, T. Hirono, N. Hosoda, M. Ishii, T. Inagaki,Y. Inubushi, T. Itoga, Y. Joti, M. Kago, T. Kameshima, H. Kimura, Y. Kirihara,A. Kiyomichi, T. Kobayashi, C. Kondo, T. Kudo, H. Maesaka, X. M.Maréchal, T. Masuda, S. Matsubara, T. Matsumoto, T. Matsushita, S. Matsui,M. Nagasono, N. Nariyama, H. Ohashi, T. Ohata, T. Ohshima, S. Ono,Y. Otake, C. Saji, T. Sakurai, T. Sato, K. Sawada, T. Seike, K. Shirasawa,T. Sugimoto, S. Suzuki, S. Takahashi, H. Takebe, K. Takeshita, K. Tamasaku,H. Tanaka, R. Tanaka, T. Tanaka, T. Togashi, K. Togawa, A. Tokuhisa,H. Tomizawa, K. Tono, S. Wu, M. Yabashi, M. Yamaga, A. Yamashita,K. Yanagida, C. Zhang, T. Shintake, H. Kitamura, and N. Kumagai, “ACompact X-Ray Free-Electron Laser Emitting in the Sub-Ångström Region,”Nature Photonics 6, 540–544, June 2012.

[71] J. Choi, J. Huang, H. Kang, M. Kim, C. Yim, T. Lee, J. Oh, Y. Parc, J. Park,S. Park, et al., “Design of the PAL XFEL,” Journal of the Korean PhysicalSociety 50, 1372, May 2007.

[72] B. D. Patterson, P. Beaud, H. H. Braun, C. Dejoiea, G. Ingold, C. Milne,L. Patthey, B. Pedrini, J. Szlachentko, and R. Abela, “Science Opportunities atthe SwissFEL X-Ray Laser,” CHIMIA International Journal for Chemistry 68,73–78, Feb. 2014.

[73] E. A. Schneidmiller and M. V. Yurkov, “Photon beam properties at theEuropean XFEL,” DESY Technical Report, Sept. 2011.

[74] J. Galayda, “The Linac Coherent Light Source-II Project,” Proceedings of the5th International Particle Accelerator Conference, 935–937, June 2014.

[75] Z. Zhao, D. Wang, Z.-H. Yang, and L. Yin, “SCLF: An 8-GeV CW SCRFLinac-Based X-Ray FEL Facility in Shanghai,” Proceedings of the 38thInternational Free Electron Laser Conference, 182–184, Feb. 2018.

[76] I. Schlichting, “Serial femtosecond crystallography: The first five years,”IUCrJ 2, 246–255, Mar. 2015.

[77] M. J. Bogan, W. H. Benner, S. Boutet, U. Rohner, M. Frank, A. Barty, M. M.Seibert, F. Maia, S. Marchesini, S. Bajt, B. Woods, V. Riot, S. P. Hau-Riege,M. Svenda, E. Marklund, E. Spiller, J. Hajdu, and H. N. Chapman, “SingleParticle X-ray Diffractive Imaging,” Nano Letters 8, 310–316, Jan. 2008.

[78] J. Bielecki, M. F. Hantke, B. J. Daurer, H. K. N. Reddy, D. Hasse, D. S. D.Larsson, L. H. Gunn, M. Svenda, A. Munke, J. A. Sellberg, L. Flueckiger,A. Pietrini, C. Nettelblad, I. Lundholm, G. Carlsson, K. Okamoto,N. Timneanu, D. Westphal, O. Kulyk, A. Higashiura, G. van der Schot,D. Loh, T. E. Wysong, C. Bostedt, T. Gorkhover, B. Iwan, M. Seibert,T. Osipov, P. Walter, P. Hart, M. Bucher, A. Ulmer, D. Ray, G. Carini, K. R.Ferguson, I. Andersson, J. Andreasson, J. Hajdu, and F. R. Maia, “Electrospraysample injection for single-particle imaging with X-ray lasers,” bioRxiv,453456, Oct. 2018.

[79] A. Patriksson, E. Marklund, and D. van der Spoel, “Protein Structures underElectrospray Conditions,” Biochemistry 46, 933–945, Jan. 2007.

[80] E. G. Marklund, D. S. D. Larsson, D. van der Spoel, A. Patriksson, andC. Caleman, “Structural stability of electrosprayed proteins: Temperature andhydration effects,” Physical Chemistry Chemical Physics 11, 8069, 2009.

[81] W. Liu, A. Ishchenko, and V. Cherezov, “Preparation of microcrystals in

80

Page 81: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

lipidic cubic phase for serial femtosecond crystallography,” Nature Protocols9, 2123–2134, Sept. 2014.

[82] K. R. Ferguson, M. Bucher, J. D. Bozek, S. Carron, J.-C. Castagna, R. Coffee,G. I. Curiel, M. Holmes, J. Krzywinski, M. Messerschmidt, M. Minitti,A. Mitra, S. Moeller, P. Noonan, T. Osipov, S. Schorb, M. Swiggers,A. Wallace, J. Yin, and C. Bostedt, “The Atomic, Molecular and OpticalScience instrument at the Linac Coherent Light Source,” Journal ofSynchrotron Radiation 22, 492–497, May 2015.

[83] S. Boutet and G. J Williams, “The Coherent X-ray Imaging (CXI) instrumentat the Linac Coherent Light Source (LCLS),” New Journal of Physics 12,035024, Mar. 2010.

[84] B. Henrich, J. Becker, R. Dinapoli, P. Goettlicher, H. Graafsma, H. Hirsemann,R. Klanner, H. Krueger, R. Mazzocco, A. Mozzanica, H. Perrey, G. Potdevin,B. Schmitt, X. Shi, A. Srivastava, U. Trunk, and C. Youngman, “The adaptivegain integrating pixel detector AGIPD a detector for the European XFEL,”Nuclear Instruments and Methods in Physics Research Section A:Accelerators, Spectrometers, Detectors and Associated Equipment 633,S11–S14, May 2011.

[85] J. R. Fienup, “Phase retrieval algorithms: A comparison,” Applied Optics 21,2758, Aug. 1982.

[86] Z. Sun, J. Fan, H. Li, and H. Jiang, “Current Status of Single Particle Imagingwith X-ray Lasers,” Applied Sciences 8, 132, Jan. 2018.

[87] R. Fung, V. Shneerson, D. K. Saldin, and A. Ourmazd, “Structure from fleetingillumination of faint spinning objects in flight,” Nature Physics 5, 64–67, Jan.2009.

[88] D. Giannakis, P. Schwander, and A. Ourmazd, “The symmetries of imageformation by scattering I Theoretical framework,” Optics Express 20, 12799,June 2012.

[89] S. Kassemeyer, A. Jafarpour, L. Lomb, J. Steinbrener, A. V. Martin, andI. Schlichting, “Optimal mapping of x-ray laser diffraction patterns into threedimensions using routing algorithms,” Physical Review E 88, Oct. 2013.

[90] O. M. Yefanov and I. A. Vartanyants, “Orientation determination insingle-particle x-ray coherent diffraction imaging experiments,” Journal ofPhysics B: Atomic, Molecular and Optical Physics 46, 164013, Aug. 2013.

[91] A. Barty, C. Caleman, A. Aquila, N. Timneanu, L. Lomb, T. A. White,J. Andreasson, D. Arnlund, S. Bajt, T. R. M. Barends, M. Barthelmess, M. J.Bogan, C. Bostedt, J. D. Bozek, R. Coffee, N. Coppola, J. Davidsson, D. P.DePonte, R. B. Doak, T. Ekeberg, V. Elser, S. W. Epp, B. Erk, H. Fleckenstein,L. Foucar, P. Fromme, H. Graafsma, L. Gumprecht, J. Hajdu, C. Y. Hampton,R. Hartmann, A. Hartmann, G. Hauser, H. Hirsemann, P. Holl, M. S. Hunter,L. Johansson, S. Kassemeyer, N. Kimmel, R. A. Kirian, M. Liang, F. R. N. C.Maia, E. Malmerberg, S. Marchesini, A. V. Martin, K. Nass, R. Neutze,C. Reich, D. Rolles, B. Rudek, A. Rudenko, H. Scott, I. Schlichting, J. Schulz,M. M. Seibert, R. L. Shoeman, R. G. Sierra, H. Soltau, J. C. H. Spence,F. Stellato, S. Stern, L. Strüder, J. Ullrich, X. Wang, G. Weidenspointner,U. Weierstall, C. B. Wunderer, and H. N. Chapman, “Self-terminatingdiffraction gates femtosecond X-ray nanocrystallography measurements,”

81

Page 82: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Nature Photonics 6, 35–40, Jan. 2012.[92] S. P. Hau-Riege, R. A. London, H. N. Chapman, A. Szoke, and N. Timneanu,

“Encapsulation and Diffraction-Pattern-Correction Methods to Reduce theEffect of Damage in X-Ray Diffraction Imaging of Single BiologicalMolecules,” Physical Review Letters 98, 198302, May 2007.

[93] K. Tiedtke, A. Azima, N. von Bargen, L. Bittner, S. Bonfigt, S. Düsterer,B. Faatz, U. Frühling, M. Gensch, C. Gerth, N. Guerassimova, U. Hahn,T. Hans, M. Hesse, K. Honkavaar, U. Jastrow, P. Juranic, S. Kapitzki,B. Keitel, T. Kracht, M. Kuhlmann, W. B. Li, M. Martins, T. Núñez,E. Plönjes, H. Redlin, E. L. Saldin, E. A. Schneidmiller, J. R. Schneider,S. Schreiber, N. Stojanovic, F. Tavella, S. Toleikis, R. Treusch, H. Weigelt,M. Wellhöfer, H. Wabnitz, M. V. Yurkov, and J. Feldhaus, “The soft x-rayfree-electron laser FLASH at DESY: Beamlines, diagnostics and end-stations,”New Journal of Physics 11, 023029, Feb. 2009.

[94] C. Behrens, F.-J. Decker, Y. Ding, V. A. Dolgashev, J. Frisch, Z. Huang,P. Krejcik, H. Loos, A. Lutman, T. J. Maxwell, J. Turner, J. Wang, M.-H.Wang, J. Welch, and J. Wu, “Few-femtosecond time-resolved measurements ofX-ray free-electron lasers,” Nature Communications 5, 3762, Dec. 2014.

[95] Y. Duan, C. Wu, S. Chowdhury, M. C. Lee, G. Xiong, W. Zhang, R. Yang,P. Cieplak, R. Luo, T. Lee, J. Caldwell, J. Wang, and P. Kollman, “APoint-Charge Force Field for Molecular Mechanics Simulations of ProteinsBased on Condensed-Phase Quantum Mechanical Calculations,” Journal ofComputational Chemistry 24, 1999–2012, Dec. 2003.

[96] A. D. MacKerell, D. Bashford, M. Bellott, R. L. Dunbrack, J. D. Evanseck,M. J. Field, S. Fischer, J. Gao, H. Guo, S. Ha, D. Joseph-McCarthy,L. Kuchnir, K. Kuczera, F. T. K. Lau, C. Mattos, S. Michnick, T. Ngo, D. T.Nguyen, B. Prodhom, W. E. Reiher, B. Roux, M. Schlenkrich, J. C. Smith,R. Stote, J. Straub, M. Watanabe, J. Wiórkiewicz-Kuczera, D. Yin, andM. Karplus, “All-Atom Empirical Potential for Molecular Modeling andDynamics Studies of Proteins,” The Journal of Physical Chemistry B 102,3586–3616, Apr. 1998.

[97] A. D. Mackerell, M. Feig, and C. L. Brooks, “Extending the Treatment ofBackbone Energetics in Protein Force Fields: Limitations of Gas-PhaseQuantum Mechanics in Reproducing Protein Conformational Distributions inMolecular Dynamics Simulations,” Journal of Computational Chemistry 25,1400–1415, Aug. 2004.

[98] W. L. Jorgensen and J. Tirado-Rives, “The OPLS [optimized potentials forliquid simulations] potential functions for proteins, energy minimizations forcrystals of cyclic peptides and crambin,” Journal of the American ChemicalSociety 110, 1657–1666, Mar. 1988.

[99] W. L. Jorgensen, D. S. Maxwell, and J. Tirado-Rives, “Development andTesting of the OPLS All-Atom Force Field on Conformational Energetics andProperties of Organic Liquids,” Journal of the American Chemical Society118, 11225–11236, Jan. 1996.

[100] G. A. Kaminski, R. A. Friesner, J. Tirado-Rives, and W. L. Jorgensen,“Evaluation and Reparametrization of the OPLS-AA Force Field for Proteinsvia Comparison with Accurate Quantum Chemical Calculations on Peptides,”

82

Page 83: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

The Journal of Physical Chemistry B 105, 6474–6487, July 2001.[101] P. M. Morse, “Diatomic Molecules According to the Wave Mechanics. II.

Vibrational Levels,” Physical Review 34, 57–64, July 1929.[102] E. Neria, S. Fischer, and M. Karplus, “Simulation of activation free energies in

molecular systems,” The Journal of Chemical Physics 105, 1902–1921, Aug.1996.

[103] M. J. Robertson, J. Tirado-Rives, and W. L. Jorgensen, “Improved Peptide andProtein Torsional Energetics with the OPLS-AA Force Field,” Journal ofChemical Theory and Computation 11, 3499–3509, July 2015.

[104] W. L. Jorgensen, J. Chandrasekhar, J. D. Madura, R. W. Impey, and M. L.Klein, “Comparison of Simple Potential Functions for Simulating LiquidWater,” The Journal of Chemical Physics 79, 926–935, July 1983.

[105] S. Plimpton, “Fast Parallel Algorithms for Short-Range Molecular Dynamics,”Journal of Computational Physics 117, 1–19, Mar. 1995.

[106] J. C. Phillips, R. Braun, W. Wang, J. Gumbart, E. Tajkhorshid, E. Villa,C. Chipot, R. D. Skeel, L. Kalé, and K. Schulten, “Scalable MolecularDynamics with NAMD,” Journal of Computational Chemistry 26, 1781–1802,Dec. 2005.

[107] D. Van der Spoel, E. Lindahl, B. Hess, G. Groenhof, A. E. Mark, and H. J. C.Berendsen, “GROMACS: Fast, Flexible, and Free,” Journal of ComputationalChemistry 26, 1701–1718, Dec. 2005.

[108] B. Henke, E. Gullikson, and J. Davis, “X-Ray Interactions: Photoabsorption,Scattering, Transmission, and Reflection at E = 50-30,000 eV, Z = 1-92,”Atomic Data and Nuclear Data Tables 54, 181–342, July 1993.

[109] J. C. Slater, “Atomic Shielding Constants,” Physical Review 36, 57–64, July1930.

[110] H. A. Scott, “Cretin—a Radiative Transfer Capability for LaboratoryPlasmas,” Journal of Quantitative Spectroscopy and Radiative Transfer 71,689–701, Oct. 2001.

[111] H. A. Scott and R. W. Mayle, “GLF - A Simulation Code for X-Ray Lasers,”Applied Physics B Laser and Optics 58, 35–43, Jan. 1994.

[112] J. Andreasson, B. Iwan, A. Andrejczuk, E. Abreu, M. Bergh, C. Caleman, A. J.Nelson, S. Bajt, J. Chalupsky, H. N. Chapman, R. R. Fäustlin, V. Hajkova, P. A.Heimann, B. Hjörvarsson, L. Juha, D. Klinger, J. Krzywinski, B. Nagler, G. K.Pálsson, W. Singer, M. M. Seibert, R. Sobierajski, S. Toleikis, T. Tschentscher,S. M. Vinko, R. W. Lee, J. Hajdu, and N. Tîmneanu, “Saturated Ablation inMetal Hydrides and Acceleration of Protons and Deuterons to keV Energieswith a Soft-x-Ray Laser,” Physical Review E 83, Jan. 2011.

[113] B. von Ardenne, M. Mechelke, and H. Grubmüller, “Structure determinationfrom single molecule X-ray scattering with three photons per image,” NatureCommunications 9, Dec. 2018.

[114] A. T. J. B. Eppink and D. H. Parker, “Velocity map imaging of ions andelectrons using electrostatic lenses: Application in photoelectron andphotofragment ion imaging of molecular oxygen,” Review of ScientificInstruments 68, 3477–3484, Sept. 1997.

[115] F. R. N. C. Maia, “The Coherent X-ray Imaging Data Bank,” Nature Methods9, 854–855, Sept. 2012.

83

Page 84: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

This work is partly based on the licentiate thesisAdvances in Biomolecular Imaging with X-ray Free-Electron Lasers

by the same author, presented in 2017 at Uppsala University, Sweden.

84

Page 85: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by
Page 86: Simulations of Biomolecular Fragmentation and Diffraction ...uu.diva-portal.org/smash/get/diva2:1307570/FULLTEXT01.pdfhigh-intensity pulses on the femtosecond time scale produced by

Acta Universitatis UpsaliensisDigital Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Science and Technology 1815

Editor: The Dean of the Faculty of Science and Technology

A doctoral dissertation from the Faculty of Science andTechnology, Uppsala University, is usually a summary of anumber of papers. A few copies of the complete dissertationare kept at major Swedish research libraries, while thesummary alone is distributed internationally throughthe series Digital Comprehensive Summaries of UppsalaDissertations from the Faculty of Science and Technology.(Prior to January, 2005, the series was published under thetitle “Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Science and Technology”.)

Distribution: publications.uu.seurn:nbn:se:uu:diva-382441

ACTAUNIVERSITATIS

UPSALIENSISUPPSALA

2019