60
ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2021 Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Science and Technology 1999 Fragment-based drug discovery Novel methods and strategies for identifying and evolving fragment leads EDWARD A. FITZGERALD ISSN 1651-6214 ISBN 978-91-513-1106-7 urn:nbn:se:uu:diva-429950

Fragment-based drug discovery - DiVA portal

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Fragment-based drug discovery - DiVA portal

ACTAUNIVERSITATIS

UPSALIENSISUPPSALA

2021

Digital Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Science and Technology 1999

Fragment-based drug discovery

Novel methods and strategies for identifying andevolving fragment leads

EDWARD A. FITZGERALD

ISSN 1651-6214ISBN 978-91-513-1106-7urn:nbn:se:uu:diva-429950

Page 2: Fragment-based drug discovery - DiVA portal

Dissertation presented at Uppsala University to be publicly examined in A1:107a, BMC(Biomedicinsk Centrum), Husargatan 3, Uppsala, Wednesday, 24 February 2021 at 09:15 forthe degree of Doctor of Philosophy. The examination will be conducted in English. Facultyexaminer: Professor Hanna-Kirsti Schrøder Leiros (UiT The Arctic University of Norway).

AbstractFitzGerald, E. A. 2021. Fragment-based drug discovery. Novel methods and strategies foridentifying and evolving fragment leads. Digital Comprehensive Summaries of UppsalaDissertations from the Faculty of Science and Technology 1999. 59 pp. Uppsala: ActaUniversitatis Upsaliensis. ISBN 978-91-513-1106-7.

The need for new drugs became ever more apparent in the year 2020 when the world was facedwith a viral pandemic. How drugs are discovered and their relevance to society became partof daily discussions in workplaces and homes throughout the world. Consequently, efficientstrategies for preclinical drug discovery are clearly needed.

The aim of this thesis has been to contribute to the drug discovery process by developing novelmethods for fragment-based drug discovery (FBDD), a rapidly developing approach wheresuccess relies on access to sensitive and informative analytical methods as well as chemicalcompounds with suitable properties. This process is fundamentally dependent on the interplaybetween scientists and engineers across biology, chemistry and physics.

This project is characterized by the development and implementation of novel biophysicalmethods over a series of studies, which are subdivided into: 1. Development of biosensor assaysand approaches for challenging targets, 2. Discovery of fragments targeting dynamic proteinsusing biosensors, and 3. Reconstruction of ligands using fragment-based strategies.

A selection of diverse targets was used as challenging prototypes for the target agnosticmethodologies described herein. The targets in focus were: acetylcholine-binding protein(AChBP), a soluble homologue of ligand gated ion channels, and two complex multi-domainepigenetic enzymes lysine specific demethylase 1 (LSD1) and SET and MYND domain-containing protein 3 (SMYD3). Expression, purification, engineering of protein variants,and biochemical characterization were required before robust screening strategies could beestablished.

Three types of biosensors, based on different time-resolved and very sensitive detectionprinciples (SPR, SHG, GCI), were used to identify and characterize the kinetics of theinteractions of novel fragments for the proteins. For SPR, a variety of multiplexed assayswere designed for the screening of fragments against difficult targets. Notably, it led tothe identification of an allosteric ligand and site in SMYD3, which was subsequentlycharacterized kinetically and structurally using X-ray crystallography, and further evolved usingcomputational approaches.

An innovative SHG assay for the specific detection of ligands inducing conformationalchanges was developed and used for fragment screening against AChBP. It revealed thatfragments with a potential to serve as functional regulators of ligand gated ion channels canbe identified using this technique. The combined application of the novel biophysical andcomputational approaches enabled the identification of useful starting points for drug discoveryprojects.

Keywords: Biochemistry, Drug Discovery, Biophysics, Fragment-based drug discovery,Epigenetics, Biosensors, Surface Plasmon Resonance, Interaction Analysis, Second-Harmonics

Edward A. FitzGerald, Department of Chemistry - BMC, Biochemistry, Box 576, UppsalaUniversity, SE-75123 Uppsala, Sweden.

© Edward A. FitzGerald 2021

ISSN 1651-6214ISBN 978-91-513-1106-7urn:nbn:se:uu:diva-429950 (http://urn.kb.se/resolve?urn=urn:nbn:se:uu:diva-429950)

Page 3: Fragment-based drug discovery - DiVA portal

List of Papers

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

I FitzGerald, E.A., Vagrys, D., Opassi, G., Klein, H.F., Hamilton, D.J., Boronat, P., Cederfelt, D., Talibov, V.O., Abramsson, M.,Moberg, A., Lindgren, M.T., Holmgren, C., Dobritzsch, D., Da-vis, B., O’Brien, P., Wijtmans, M., van Muijlwijk-Koezen, J.E.,Hubbard, R.E., de Esch, I.J.P., Danielson, U.H. (2020) Multi-plexed experimental strategies for fragment library screening us-ing SPR biosensors. bioRxiv, doi: 10.1101/2020.12.23.424167.

II FitzGerald, E.A., Butko, M., Boronat, P., Cederfelt, D., Abramsson, M., Ludviksdottir, H., van Muijlwijk-Koezen, J.E., de Esch, I.J.P., Dobritzsch, D., Young, T., Danielson, U.H. (2020) Discovery of fragments targeting dynamic proteins using second-harmonic generation. Submitted

III Talibov, V.O., Fabini, E., FitzGerald, E.A., Tedesco, D., Cederfelt, D., Talu, M.J., Rachman, M.M., Mihalic, F., Manoni, E., Naldi, M., Sanese, P., Forte, G., Signorile, M.L., Barril, X., Simone, C., Bartolini, M., Dobritzsch, D., Del Rio, A., Dan-ielson, U.H. (2021) Discovery of allosteric ligand binding site in SMYD3 lysine methyltransferase. ChemBioChem, Accepted Au-thor Manuscript. doi: 10.1002/cbic.202000736.

IV FitzGerald, E.A., Rachman, M.M., Cederfelt, D., Zhang, H., Barril, X., Dobritzsch, D., Koehler, K., Danielson, U.H. (2021) Evolution of an allosteric ligand of the epigenetic modulator SMYD3 - Retrosynthesis using an integrated biophysical and computational approach. Manuscript

Reprints were made with permission from the respective publishers.

Page 4: Fragment-based drug discovery - DiVA portal

Contribution report

Paper I Planned the study. Produced and purified AChBP and LSD1. Designed and performed kinetic experiments, designed and performed thermal shift assays, screened initial crystallog-raphy conditions. Analysed data from these experiments. Co-wrote the manuscript.

Paper II Planned the study. Produced and purified AChBP, designed single cysteine mutants and conducted site directed mutagen-esis. Developed and performed SHG experiments, analysed the data. Developed and performed grating-coupled interfer-ometry assay, analysed data. Wrote the manuscript.

Paper III Participated in planning of the study, performed computa-tional experiments and analysed SMYD3 surface features, contributed to writing the manuscript.

Paper IV Planned the study. Produced and purified SMYD3. Partici-pated in the design of compounds, acquired compounds. De-veloped grating-coupled interferometry assay, performed ex-periments and analysed data. Set crystals. Wrote the manu-script.

Page 5: Fragment-based drug discovery - DiVA portal

Contents

Introduction ................................................................................................. 9Drug discovery ....................................................................................... 9Early stage & preclinical drug discovery ............................................. 10

Phases of preclinical drug discovery ............................................... 11High-throughput screening .............................................................. 11Fragment-based drug discovery ....................................................... 12An interdisciplinary and highly collaborative effort ....................... 15

Molecular recognition in drug discovery .................................................. 16Functional interactions beyond binding ............................................... 19

Protein-protein interactions ............................................................. 19Analysis of biomolecular interactions using label-free biosensor-based methods ................................................................................................ 20

Surface Plasmon Resonance (SPR) ................................................. 21Grating Coupled Interferometry (GCI) ............................................ 21

Analysing kinetic data with label-free biosensor-based methods ........ 23Interaction mechanisms ........................................................................ 24

1:1 (Langmuir) binding model ........................................................ 24Heterogenous ligand model ............................................................. 24Steady-state affinity ......................................................................... 25

Analysis of biomolecular interactions using labelled biosensor-based methods ................................................................................................ 25

Second-Harmonic Generation (SHG) .............................................. 26Orthogonal validation and structural insights ...................................... 26

Protein crystallography .................................................................... 27Computational methods ................................................................... 27

Aim ........................................................................................................... 29

Present investigation ................................................................................. 30Target proteins ...................................................................................... 31Compound libraries .............................................................................. 32Discovery of fragment hits (Papers I & II) .......................................... 33

Improving SPR biosensor-based screening strategies for complex and difficult targets .......................................................................... 33Extending hit identification beyond detection of binding -Detection of ligand-induced conformational changes ...................................... 38

Page 6: Fragment-based drug discovery - DiVA portal

Conclusions: Fragment library screening, SPR vs. SHG ................ 42Screening for allosteric ligands and confirming their binding modes (Paper III) ............................................................................................. 42Evolution of ligands – a combined biophysical and in silico approach (Paper IV) ............................................................................................. 46

Conclusions ............................................................................................... 49

Future at a glance ...................................................................................... 50

Sammanfattning ........................................................................................ 51

Acknowledgements ................................................................................... 53

References ................................................................................................. 56

Page 7: Fragment-based drug discovery - DiVA portal

Abbreviations

AC Aplysia californica AChBP Acetyl choline binding protein ADME Absorption, distribution, metabolism, excretion COREST Corepressor for REST DSF Differential scanning fluorimetry FBDD Fragment-based drug discovery FRAGNET Fragment networkGCI Grating coupled interferometry HTS High-throughput screeningHTVS High-throughput virtual screening KMT Lysine methyl transferase LGIC Ligand gated ion channel LS Lymnaea stagnalis LSD1 Lysine-specific demethylase 1 LSD2 Lysine-specific demethylase 2 MSCA Marie Skłodowska-Curie actions MST Microscale thermophoresis PDB Protein data bank PMI Principle moments of inertia PPI Protein-protein interactionPROTAC Proteolysis targeting chimera RI Refractive indexRU Response unitsSAM S-Adenosyl methionineSAR Structure–activity relationshipSBDD Structure based drug discoverySH- active Second harmonic activeSHG Second harmonic generationSMYD3 SET and MYND domain-containing protein 3SPR Surface plasmon resonancetcCA Carbonic anhydrases from Trypanosoma cruziTSA Thermal shift assayWHO World health organizationXRC Protein x-ray crystallography

Page 8: Fragment-based drug discovery - DiVA portal
Page 9: Fragment-based drug discovery - DiVA portal

9

Introduction

Drug discovery At the time of writing, the world is facing one of its most devastating pandem-ics in recorded history. The novel corona virus, SARS-CoV-2, has had a dev-asting impact on human lives, causing millions of deaths, with severe conse-quences to the global economy.1 It has never been more apt or timely to dis-cuss the importance and fundamentals of drug discovery. How are such pro-jects initiated, what is required, and how can science in the year 2021 help accelerate our discovery? Although the definition of a drug may have changed over a millennium, the concept is far from new. Taking an elixir, a potion, or a medicine has existed in one form or another for thousands of years. Evidence for this is scattered throughout history books. Whether it is the extraction of caffeine from the humble tea leaf in China ~2700 BCE or animals consuming natural products for their analgesic effects, e.g. the coca leaf being chewed by gorillas to suppress hunger and relieve pain. Fast forward to 1855 and we see how through advancement of scientific tech-nologies and scientific intervention the German chemist Fredrich Gaedcke isolated the cocaine alkaloid from the coca leaf dubbing it “erythroxyline”.2 The organic synthesis and structural elucidation came later in 1898 when Ri-card Willstätter performed the first recorded synthesis.3 It was from this point it became widely used as an anaesthetic4 or prescribed for many over the coun-ter indications including toothache drops for children.5 It was in the late 19th and early 20th centuries we see the development of the modern pharmaceutical industry on what was largely off the back of the petrochemical dye industry and fine chemical industries in Switzerland and North America.6 There are many examples of going from natural product to drug, or examples of seren-dipitous discoveries. However, the real advances came with structure based or structure guided drug discovery in what is now known as early-stage drug discovery (ESDD) or preclinical drug discovery.

Page 10: Fragment-based drug discovery - DiVA portal

10

Early stage & preclinical drug discovery Structure based drug discovery (SBDD), often referred to as rational drug de-sign, is a method where prior knowledge of the chemical structure of a known drug, or the structure of a target of interest is exploited to rationally design or guide the process. An example of this, upon the discovery of the antibiotic penicillin, scientists rapidly produced subsequent variations of the parent mol-ecule by producing semi synthetic β-lactam compounds with various chemical modifications decorated around the β-lactam ring.7 This led to the discovery of more potent, and/or, drugs which were less prone to antibiotic resistance. As technologies advanced and scientific methodology improved, focus was placed on understanding the binding pocket of the 3D structure of a target of interest. This became more and more viable as the number of structures in the protein-data bank (PDB) became increasingly available.8 Due to advances in recombinant protein production and methods of structural elucidation i.e. X-ray crystallography (XRC) and nuclear magnetic resonance (NMR), experi-mentalists were able to visualise and understand the binding mode of com-pounds interacting with a target protein. This opened up the possibility of us-ing computational methods to visualise, model and improve future iterations of compounds by complementing the shape and charge of the binding pocket. This process in the drug discovery pipeline (Figure 1) is often described col-lectively as “preclinical discovery” and incorporates target validation, hit identification, and lead generation. Once a hit has been optimised, it then moves to the clinical phases of the pipeline.

Figure 1. Schematic of the drug discovery pipeline. Both the cost and the risk of failure increase as the process moves from one stage to the next.

Page 11: Fragment-based drug discovery - DiVA portal

11

Phases of preclinical drug discovery Target validation, screening, hit generation and lead optimisation are inter-twined. During these steps an iterative movement from one step to the previ-ous can often happen. A critical point in progressing hits to the next step in the pipeline is the orthogonal validation of the compounds using a method based on a different assay principle. This process can be seen as a decision workflow and is summarised in Figure 2.

High-throughput screening In the early nineties, big pharma invested heavily in enormous infrastructure for high-throughput screening (HTS). This included building libraries of hun-dreds of thousands or even millions of compounds, and establishing advanced robotics to handle the required throughput. A typical HTS library is character-ized by drug- or lead-like compounds which follow Lipinski’s rule of 59 and have a molecular weight of <500 Da. In theory, these compounds should have the potential to display a high affinity interaction, i.e. have KD-values in the low micromolar to nanomolar range for various targets of interest. In most cases, a very sensitive biochemical assay is established to rapidly screen these very large libraries. If compounds meet the criteria of the assay, e.g. inhibit the target, they are considered as hits. After validation of the hits, thus exclud-ing compounds appearing to be hits due to assay artifacts (false positives), the compounds are structurally optimised to improve their physicochemical and biological properties. Figure 3 illustrates a typical HTS approach .

Figure 3. Illustration representing a typical HTS approach. A target of interest is shown in green, with HTS library defined with interconnected shapes. Although com-pounds fit the hypothetical binding pocket there is little room for optimisation.

Figure 2. Decision workflow in going from a validated target to a preclinical lead or clinical candidate. If the 3D structure of the protein is known, compound libraries can be screened virtually in silico before being validated by biophysical methods.

Page 12: Fragment-based drug discovery - DiVA portal

12

However robust these assays are, they often fall short for many reasons: 1. Although libraries can contain millions of compounds, they often lack the complexity required for novel targets, and probe a fraction of available chem-ical space. 2. These compounds are usually remnants from previous projects and can lack the required specificity for novel target classes. 3. Due to the chemical composition and being lead-like or drug-like they can limit flexibil-ity in further optimisation. Despite hits being readily identified, oftentimes they are characterized by large hydrophobic moieties leading to “greasy” leads with unfavourable pharmacokinetic profiles. Mike Hann of GSK suggests that these compounds suffer from “molecular obesity”10 i.e. large bulky com-pounds, which due to their hydrophobicity, present a risk in future candidate nomination, leaving little room for optimisation. Consequently, the molecules resulting from this type of optimisation often have problems with ADME pro-files such as solubility and oral bioavailability.

Fragment-based drug discovery Fragment-based drug discovery (FBDD) is now an accepted and validated mainstay in the drug discovery process.11 Its core principles and theory was built on the higher sensitivity and throughput of biophysical assays, starting out with NMR, followed by X-ray crystallography and SPR, methods that evolved and had the required throughput for fragment libraries covering a suf-ficiently large chemical space. It was catalysed by a landmark paper in 1996 where researchers at Abbott describe the identification of high affinity ligands using SAR by NMR.12 In the early 2000’s FBDD went from being a niche technique to a legitimate discipline which was rapidly adopted by both indus-try and academia.13 It was seen as viable means for discovering novel chemical entities for a drug discovery or a chemical biology project.

This traction was due to advances in experimental design and due to the de-velopment of increasing sensitivity in biophysical instrumentation.14 Rather than relying on massive chemical libraries, the focus is placed on reducing the complexity of the libraries to fragments of compounds. This reductionist ap-proach to HTS suggests, in theory, that a much broader chemical space can be more efficiently probed by using structurally diverse compounds with lower molecular weight than one would conventionally find in a HTS or drug-like compound library.15

For FBDD, libraries typically follow a “rule of three” concept, in which the molecular weight of a fragment is <300, the cLogP is ≤3, the number of hy-drogen bond donors is ≤3 and the number of hydrogen bond acceptors is ≤3.16 A typical fragment library is between 500 and 2500 fragments. Due to the smaller size one can explore a far greater chemical space by exploring strate-gies of fragment linking/merging or growing (Figure 4). This also provides

Page 13: Fragment-based drug discovery - DiVA portal

13

flexibility when later optimising for lead generation, including at the candidate nomination stages, when different physicochemical properties are adjusted to predict favourable ADME properties for the drug candidate.

As mentioned earlier, the drug discovery process is often a long, risky and costly business. Despite this fact, we are already seeing numerous examples of novel therapeutics enter late stage clinical trials and more importantly, to date, four FDA-approved drugs have been discovered by fragment based ap-proaches: vemurafenib17, erdafitinib18, venetoclax19 and pexidartinib20 (Fig-ure 5).

Figure 4. Illustration of a typical FBDD screen against a target of interest (green), with a fragment library (colored shapes). A fragment can bind and 1. be grown, guided by structural and medicinal chemistry routes, or 2. If two fragments are found close to each other they may be linked. In both instances the compounds are optimised to favor ADME and bioavailability.

Page 14: Fragment-based drug discovery - DiVA portal

14

Figure 5. Current drugs developed by a fragment-based approach and the initial frag-ment hits from which they were derived.

Page 15: Fragment-based drug discovery - DiVA portal

15

An interdisciplinary and highly collaborative effort FBDD relies on the tight interplay between many disciplines. For a project to succeed, no one scientist or a single academic can realistically expect to lead a campaign from target identification through to a marketed drug. In the pre-clinical stages the assays and expertise required in going from hit to lead are vast. With this knowledge, and in order to tackle what may feel to be a gar-gantuan task, industry and academia have seen the strength in partnership. As a result of the highly specialized skillset required in FBDD, a pan European project was established under the Marie Skłodowska-Curie actions (MSCA) to train early career researchers in FBDD. It was dubbed FRAGNET, for frag-ment network, and promotes interdisciplinary research by training researchers in all aspects related to FBDD.21 The researchers are placed at either industrial or academic host institutions and specialize in one of the four pillars defined by the FRAGNET consortium (Figure 6): 1. Libraries 2. Screening 3. Design, and 4. Optimisation. For example, a researcher involved in the development of biophysical assays will rely on libraries developed by the consortia, simi-larly those involved in design will rely on collaboration and be guided by part-ners developing the screening assays and subsequent lead optimisation. This interplay will become apparent throughout the course of this thesis.

Figure 6. Overview of the FRAGNET strategy, encompassing the four pillars of FBDD conducted in the network; 1. Libraries 2. Screen 3. Design and 4. Optimisa-tion. Adapted from 22.

Page 16: Fragment-based drug discovery - DiVA portal

16

Molecular recognition in drug discovery

To understand what a hit is, and how it is identified, it is important to under-stand and indeed define what a molecular interaction is, the recognition event, and what it means in a drug discovery context. The term molecular recognition can be broadly defined as a specific interaction between two or more biomol-ecules. These interactions may be comprised of one or more of the following: hydrogen bonding, hydrophobic forces, van der Waals and π-π interactions, resonant interaction effects, or metal coordination. These interactions can be further simplified as requiring a degree of complementarity to be formed i.e. negative charge being attracted by positive charge. When thinking of molecular recognition in a drug discovery context, one is often presented with the concept of a lock and key mechanism23 where the key (inhibitor or drug), being complementary to a specific lock (enzyme), fits the lock which will then lead to a preferred or desired outcome. However, this oversimplification treats the enzyme as a rigid object and fails to account for the dynamic nature and flexibility of the system. In fact, the interaction often involves a myriad of conformational changes required for ligand binding. When considering the design of inhibitors or drugs it can be useful to consider the mechanism and kinetics of the interaction (Figure 7).

Figure 7. Simple representation of molecular recognition in the reversible interaction between a protein (e.g. an enzyme) and a ligand (e.g. an inhibitor) forming a pro-tein:ligand complex by interacting at a specific site (e.g. an active site).

In Figure 7, enzyme inhibition is used as an example to illustrate how a mo-lecular entity fits into a specific site on a protein, with favourable shape com-plementarity. If the interaction forces are also complementary, this can result

Page 17: Fragment-based drug discovery - DiVA portal

17

in a high affinity interaction with favourable kinetics for the formation of a stable complex. Here the inhibitor is a hypothetical drug that could lead to a therapeutic effect. However, a drug can be many things, it can inhibit, activate or modulate a target. The target can be any biomolecule which can be modu-lated or manipulated to induce a therapeutic outcome. Targets can be proteins, protein complexes, enzymes, RNA, DNA, antibodies and so on. Designing a drug or “ligand” can be achieved as described above. During the hit identification stages, particularly with fragments which are defined by weak transient interactions, being able to rank hits is important. This can be thought of in terms of thermodynamics or kinetics, hits can be ranked by the fragment’s affinity for the target and be expressed in the following equation and defined as the equilibrium dissociation constant, KD.

Equation 1

Where [P] and [L] are the free concentrations of target and ligand, [PL] the concentration of the binary complex, koff the dissociation rate constant, and kon

the association rate constant. The interaction can also be expressed in terms of the free energy ( G) of the interaction, which can be separated into two thermodynamic parameters, en-thalpy ( H) and entropy ( S), as defined by (Equation 2).

Equation 2 The energy of the interaction is associated to the equilibrium dissociation con-stant, KD as defined in Equation 3

Equation 3 and the van’t Hoff equation (Equation 4),

Equation 4

Page 18: Fragment-based drug discovery - DiVA portal

18

where T is the absolute temperature in Kelvin (K) and R is the gas constant. From Equations 3 and 4, it can be inferred that similar values of ∆G may have differences in ∆H and ∆S. Furthermore, a correlation to the thermodynamic data with an affinity is also true. The same is true for kinetic rates, since with a system at equilibrium the KD values can be obtained from the kinetic rates (Equation 1). Similarly, from the energy diagram below, ∆Gon and ∆Goff relate to kon and koff since they are inverse relationships. It also shows why ∆G and KD can be con-stant even though the other parameters may vary.

Figure 8. Energy diagram of a ligand (L) interacting with a protein (P) in a simple one step mechanism (from Figure 7), highlighting the relationship between binding energy states and kinetics.

Page 19: Fragment-based drug discovery - DiVA portal

19

Functional interactions beyond binding

Protein-protein interactions As described above there are multiple ways to consider drugging a target in search of a therapeutic effect. It can be useful to contemplate other modes of binding other than a molecular event at the active site of an enzyme. Proteins can be involved in multiple cellular processes and can play various roles in a biological context such as signalling, macromolecular structure, messaging, catalysis, and many more. During these processes, proteins can have complex protein-protein interac-tions, so much so it has become a scientific discipline onto itself i.e. the study of the interactome. With clever databases it has become easier to identify and visualise these.24 In the previous example of an enzyme inhibitor complex, we often think of a protein-small molecule interaction. However, the inhibition or activation of a molecular process may occur as a result of a protein-protein interaction (PPI) leading to the formation of a protein-protein complex (Fig-ure 9). Here the concept of complementarity becomes evident, where there needs to be a favourable interaction at the interface of each protein binding partner. This can be seen as a series of hot spots which collectively form a high affinity complex where the degree of complementarity can be compared to a Velcro fastener on clothing sticking. This has become of value to identify and perturb for a therapeutic outcome.25 Can we use fragments to inhibit or perturb such protein-protein interactions?

Figure 9. Illustration of a simple protein-protein interaction, highlighting comple-mentary surfaces with “hot spots” forming the interaction.

Page 20: Fragment-based drug discovery - DiVA portal

20

Analysis of biomolecular interactions using label-free biosensor-based methods Two label-free biosensor-based approaches are used in this thesis. Surface plasmon resonance (SPR) and Grating coupled interferometry (GCI). Both have an output in the form of an interaction kinetic curve or “sensorgram”, where the interaction kinetics can be quantified through time resolved exper-iments. They differ in their detection principle, but have similar methods of surface preparation, experimental design, and data analysis. Throughput and sensitivity are defined by instrument configurations.

Figure 10. Overview of biosensors and time-resolved experimental output. (a) SPR biosensor with Kretschmann configuration. (b) A modified Mach-Zehnder biosensing interferometer used in GCI. (c) Output from SPR and GCI biosensors, in the form of an interaction kinetic curve (sensorgram), comprised of 1. pre-injection baseline measurement, 2. association phase, and 3. dissociation phase.

Page 21: Fragment-based drug discovery - DiVA portal

21

Surface Plasmon Resonance (SPR) The SPR instrumentation which is most commonly used is based on a Kretschmann configuration (Figure 10a), and was made popular by the launch of Biacore in 1991.26 The sensor is comprised of a gold sensor surface, a microfluidic flow cell, a prism, and a CCD detector. The detection principle is such that the interaction between biomolecules is detected by surface plas-mon resonance. This phenomena occurs at the surface of a derivatised thin gold layer, usually ~5 µm26 between two layers which have a different refrac-tive index. In biosensors these are represented by a glass layer and a sample solution layer. The gold surface is derivatised with a target of interest which has been immobilised via a flexible hydrogel consisting of a carboxylated dex-tran layer. Due to surface plasmon resonance, the intensity of the reflected light has a minimum at an angle (θr) which depends on the refractive index of the medium close to the gold layer. By monitoring this angle continuously, it is possible to detect changes taking place at the surface in real time, e.g. inter-actions with immobilised proteins. Signals are dominated by changes in mass at the surface, but also changes in the refractive index resulting from differ-ences in DMSO concentrations between analyte samples and running buffers, or conformational changes can give rise to detectable signals, requiring careful pipetting, extensive controls and data analysis procedures.

Grating Coupled Interferometry (GCI) Grating coupled interferometry is also a refractive index (RI) based, mass de-pendent biosensing technique. Detectors and sensor setups can differ in de-sign. The WAVE™ developed by Creoptix AG utilizes a modified Mach-Zehnder interferometer. In this configuration a waveguide is combined with optical interferometry to detect changes in (RI). The waveguide is composed of Ta2O5 which has a high RI contrast, permitting the detection of small changes to the RI.27 Said waveguide is covered by SiO2, apart from the 5 mm sensing region. The sensing region is functionalised with a surface matrix, which can be modified to attach biomolecules of interest. The sensor is de-fined by three different gratings on the waveguide; two incoupling gratings, the 1st grating is found passing through the sensing region, and the 2nd after the sensing region, functioning as a reference (Figure 10b, 1st 2nd) and a 3rd after the sensing surface (Figure 10b 3rd). The 1st and 2nd gratings are used to guide the two incoming light rays i.e. the measuring light and reference light (Figure 10b) in the waveguide, creating the arms of the interferometer.27,28 The 5 mm sensing region is probed by an evanescent field from the electro-magnetic wave that is propagating through the waveguide. As in SPR, the ev-anescent wave senses changes to the RI caused by a mass change of analytes interacting with functionalised biomolecules on the sensing surface.27–29 The

Page 22: Fragment-based drug discovery - DiVA portal

22

2nd, acting as reference grating, produces a phase shifted wave in the wave-guide upon the waves combining. This final shifted wave is outcoupled in the 3rd grating which directs the light to a detector which converts the interference pattern to a sensorgram. Both SPR and GCI visualise interactions in a signal to time plot i.e. a kinetic curve or sensorgram (Figure 10c). In the case of SPR, the change of the SPR angle (Resonance Units, RU) is plotted against time, while for GCI, the re-sponse of the sensing surface (Response, pg/mm2) is plotted against time. The interaction kinetic curves consist of three distinct parts; 1. pre-injection base-line recorded with only running buffer flowing over the surface, 2. association phase recorded upon injection of analyte into the running buffer, and 3. disso-ciation phase where the analyte is washed from the sensor surface. These curves can be used to identify the interaction mechanism and determination of kinetic rate constants by fitting different kinetic models to the data.

Page 23: Fragment-based drug discovery - DiVA portal

23

Analysing kinetic data with label-free biosensor-based methods The time resolved data obtained from biosensor-based analysis of simple in-teractions between a protein and a small molecule (illustrated in Figure 7) al-lows the establishment of the interaction mechanism and the determination of the associated kinetic parameters.30 Experimental SPR data and the basic mathematical analysis typically used and the graphic visualisations of the pro-cedure are shown in Figure 11. The principle of global fitting of a suitable interaction model to the data using non-linear regression analysis is shown in Figure 11a, while Figure 11b illustrates a corresponding steady state analy-sis, based on report points taken from the sensorgrams.

Figure 11. An example of sensorgrams generated for a series of different concentra-tions of lobeline interacting with immobilised AChBP. (a) Kinetic rate constants (kon and koff) and the corresponding equilibrium constant (KD) are obtained by globally fitting an equation representing a simple 1:1 interaction model (Scheme 1, Equation 6, black line) to the complete dataset. (b) The “dose response” plot is based on report points taken at as an average of the signal during a short time period at steady state, here marked at the end of the injection in (a). The vertical line represents the ligand concentration equal to the KD-value.

For the simplest interaction typically studied, i.e. a reversible 1:1 interaction, the mechanism can mathematically be described per Equation 5:

Equation 5

where R is the response in RU, dR/dt is the change of response over time, Rmax is the maximum response in RU and C the concentration of analyte being passed over the sensing surface. However, interactions are often more complex and the model needs to be es-tablished and expressed in suitable mathematical terms for data analysis. For

Page 24: Fragment-based drug discovery - DiVA portal

24

fragment interactions with their targets, a handful of models can be useful and are indeed important for the selection of screening hits and characterization of the features of the interaction. The interaction models used in this thesis are summarised below.

Interaction mechanisms

1:1 (Langmuir) binding model As is illustrated in Figure 7, the simplest interaction model is a reversible 1:1 1-step interaction between two biomolecules. Its name is derived from thesimilarity of this model to the Langmuir binding adsorption of gas to a sur-face.31

Scheme 1

P is the immobilised protein i.e. a target protein, L is an injected analyte ligand i.e. a small molecule or fragment, kon and koff are the rate constants

This interaction model is described by the following differential equation:

Equation 6

Where the equilibrium dissociation constant is:

Equation 7

Heterogenous ligand model A heterogenous ligand model assumes several independent interactions and that the analyte is capable of interacting with multiple forms of the protein, but with different kinetics. This can be due to multiple protein populations on the surface, either arising dynamically as a result of structural flexibility and conformational changes, or be a consequence of interactions with structural variants representing different forms of the protein. They can also be experi-mental artifacts, resulting from a poor experimental set-up such as a subopti-mal immobilisation procedure. As the simplest heterogeneous binding model assumes two individual interactions, it is described as two independent 1:1 interaction events. The kinetic parameters and the equilibrium dissociation constants are calculated separately for each binding reaction.

Page 25: Fragment-based drug discovery - DiVA portal

25

Scheme 2

Steady-state affinity Oftentimes it is not possible to resolve the kinetic steps of an interaction due to mechanistic complexities, or to quantify the transient and weak interactions, commonly observed fragments. In such cases, it is useful to perform a steady state analysis, using report points taken at equilibrium (Figure 11). Assuming a Langmuir model, it is possible to estimate an equilibrium dissociation con-stant (KD) from the relationship of the response level (R) and the ligand con-centration [L], using the following equation.

I Equation 8

Where Rmax is the theoretical response at steady-state and RI is the refractive index contribution caused by a bulk effect in the sample.

Analysis of biomolecular interactions using labelled biosensor-based methods As previously described, SPR and GCI biosensors are considered label free assays. The target protein requires no pre-experiment sample preparation and the immobilisation strategies have become rather routine. The detection prin-ciple as outlined earlier is reliant on a mass change at the sensor surface, which then translates to a digital signal permitting extrapolation of kinetic rate con-stants. However, it fails to account for structural changes or a functional out-come upon the protein:ligand target formation. Although it has been alluded to that conformational changes can be detected using SPR,32 the analysis and interpretation of complex sensorgrams can be elusive. Observing conforma-tional changes in real-time is possible using nuclear magnetic resonance NMR33 but is limited in throughput and is highly dependent on being able to obtain protein in a suitable form and quantity.

Page 26: Fragment-based drug discovery - DiVA portal

26

Second-Harmonic Generation (SHG) In this thesis, second harmonic generation (SHG) was explored as a novel technology of potential advantage for FBDD. SHG is a labelled biosensor and is used to detect direct conformational changes in real time. The sensing sur-face is comprised of a lipid bilayer which is functionalised with a protein of interest that has been labelled with a dye probe. The detection method is based on an optical readout whereby biomolecules of interest are made second-har-monic active (SH-active) through the incorporation of SH-active dye probes. Conformational changes are detected spectroscopically using SHG, a non-lin-ear process where two photons from an incident laser are converted into a single photon of twice the energy,34 the efficiency of which is highly depend-ent on the angular orientation of the SH-active probes with respect to the sur-face normal where the biomolecules of interest are tethered Figure 12. When the SH active probe is conjugated to the biomolecule of interest and is subse-quently tethered to the surface of a well, any ligand-induced conformational change, which results in a net dye movement, will be detected with a change in the SHG signal.35 The detection principle is highlighted in Figure 12.

Figure 12. Principle for SHG biosensor assays. Affinity-tagged biomolecules conju-gated with an SH-active dye (blue) and tethered onto a lipid bilayer (orange) through either His-tag:Ni/NTA or biotinylated Avi-tag:avidin interactions. Incoming light at 800 nm (red arrow) is directed at the dye, which transforms two photons of this light into one photon of light with twice the energy (400 nm), the second-harmonic light (blue light).

Orthogonal validation and structural insights Although not trivial, identifying fragment hits using the aforementioned meth-odology is possible. The ability to orthogonally validate identified hits is an extremely important aspect of the discovery process. The gold standard for orthogonal validation is usually a method which gives structural insights into the interaction between protein and fragment. Normally, this would result in a 3D structure which can be visualised and, in some cases, used for computa-tional modelling. The go-to method is often X-ray crystallography (XRC) but as other fields advance and technologies improve, we see more and more

Page 27: Fragment-based drug discovery - DiVA portal

27

structural information being deposited into the protein-data bank (PDB)8 from other methods including nuclear magnetic resonance (NMR) and cryogenic electron microscopy (CryoEM). Oftentimes the inability to generate structural information can be one of biggest stumbling blocks in the preclinical phases. However strategies are being improved for advancing hits in the absence of structural information.36 In the studies presented herein, XRC and in silico methods play a critical role.

Protein crystallography Protein X-ray crystallography (XRC) is a method which provides direct struc-tural information on a proteins 3D structure. It can be used to elucidate the structural details of a protein:ligand complex and thus prove highly fruitful when advancing hits to leads.37 Based on the theory of three-dimensional scat-tering of X-ray light from an ordered crystal lattice, the intensities of the scat-tered light vary depending on the atomic structural arrangement of the mole-cules in the lattice. This diffraction pattern is unique for each molecule and can be used to deduce the protein structure using Braggs law.38 The diffraction pattern is combined with phase information (e.g. from previously known ho-mologous or identical protein structures) and is then inverse Fourier trans-formed to map the electron density of the molecule.39 Applied to FBDD, fragment cocktails can be co-crystallized, or soaked into an already formed protein crystal.40 If successful the formed complex can be evaluated at atomic scale by studying the diffraction patterns. If the structure can be resolved it may elucidate a mode of action and provide additional in-formation aiding the process of going from “hit-to-lead”. This can be particu-larly useful for fragments given their weak interactions with target proteins, as crystallography conditions can permit high concentrations of fragment so-lutions. However, there can be some drawbacks to this method, including neg-ative effects of high fragment concentrations on crystal stability and packing patterns, and the time-consuming trial and error process of getting a protein construct in a suitable form to produce crystals. XRC plays a vital role in Pa-pers I-IV described herein.

Computational methods If a 3D structure of the protein is available, it opens up enormous potential for in silico studies. Not only does it enable understanding of the target in greater detail, but it also provides the opportunity to model many aspects of the sys-tem. In this thesis, a series of computational methods were used to work around limitations of a purely experimental approach. The methods used in-clude solvent mapping,41,42 molecular docking,43 and a technique known as dynamic undocking or DUcK.44 Through collaboration we further expanded

Page 28: Fragment-based drug discovery - DiVA portal

28

the repertoire of tools available to us, by using a method for fragment growing and ligand optimization, which is highlighted in Paper IV. The methods used will be discussed briefly in terms of what information they can provide and how they have guided the studies described herein. The in silico approach followed a multifaceted procedure illustrated in Figure 13, where mixed solvent mapping using pyMDmix41,42 and pocket prediction using Fpocket45 were applied to localise binding pockets of interest. Once identified, compounds were docked using rDock43, either from smaller fo-cused libraries or using high-throughput virtual screening (HTVS), to rank, score and predict binding modes. To further scrutinize the predicted interac-tions and to reduce the number of compounds to a more manageable size, dy-namic undocking44 was performed. This method uses steered molecular dy-namics to examine the strength of a key hydrogen bond by extracting or pull-ing the ligand out of the binding pocket. The extraction profile can be studied, allowing further ranking of previously identified ligands. Finally, compounds identified from in silico studies are validated using a series of biophysical as-says e.g., XRC, SPR, and GCI.

Figure 13. Overview of the computational methods explored. (1) Solvent mapping highlights identified binding pockets of interest (pink, blue, and green). (2) Docking identifies compounds of interest and predicts binding mode. (3) Dynamic undocking probes key interactions to further rank putative ligands. (4) Biophysical methods val-idate the in silico predictions.

Page 29: Fragment-based drug discovery - DiVA portal

29

Aim

The aim of this PhD thesis was to develop and implement novel biophysical methods for state-of-the-art fragment-based drug discovery. The targets cho-sen were agnostic of disease or target class but were selected as translatable model systems for novel and efficient strategies. This has involved the devel-opment and optimisation of biophysical assays for screening fragment librar-ies and subsequent in-depth characterization of interactions involved in going from a hit-to-lead. This was set into the following sub-goals:

• Identify targets of interest, explore site directed mutagenesis and ex-press different gene constructs in prokaryotic and eukaryotic systems.

• Perform biochemical and biophysical characterization of produced target proteins.

• Develop novel SPR biosensor-based FBDD fragment screening cam-paigns incorporating a series of challenging targets, characterized by underrepresented target classes or binding modalities to generate pu-tative hits for a hit-to-lead campaign.

• Develop SHG biosensor based FBDD fragment screening campaigns to understand the impact of conformational change in an FBDD con-text.

• Characterize allosteric ligands and pockets, and their potential drug-gability, in particular for PPIs.

• Develop strategies in going from hit-to-lead in an efficient manner exploring chemical space using computational approaches.

In brief, this thesis aimed to build on, improve, and develop biophysical meth-odologies to identify and characterize fragments interacting with their targets. It goes beyond basic hit identification by exploring functional outcomes in the protein:ligand complex formation, and further develops strategies in going from hit-to-lead.

Page 30: Fragment-based drug discovery - DiVA portal

30

Present investigation

Biophysical methods have become critical tools for the early stages of pre-clinical drug discovery. At a very basic level, they can be used to identify the existence of a molecular interaction, understand its features and quantify the affinity and kinetics of said interaction. From a drug discovery context, this knowledge can then be applied to define structural starting points for the de-velopment of a future drug. In this thesis, the focus was placed on how to implement and develop novel biophysical methods for FBDD, especially for complex and dynamic targets, or for ligands with specific modes-of-action. In Paper I, the fundamentals of how to use SPR for FBDD are described, including what is required for se-lecting and producing a suitable target protein and how one ought to consider the construction of a fragment library. The paper illustrates the implementa-tion of time resolved SPR biosensor-based strategies to identify and select fragment hits towards a panel of target proteins, and later use SPR and various orthogonal methods to understand and validate the interactions. In Paper II the focus is shifted from identifying fragment hits based on their affinities for the target, to their functional effects as a consequence of complex formation with the target. Specifically, the study uses SHG to identify fragments inducing conformational changes in AChBP, a structurally dynamic target protein. Pa-per III describes an experimental strategy to identify allosteric ligands through competitive SPR screening. A small library of compounds was experimentally screened against SMYD3 and in silico methods were employed to further our knowledge of the target and to validate the experimental observations. Having identified an allosteric ligand in Paper III, Paper IV builds on the knowledge acquired in Papers I-III, and flips the paradigm so that biophysical methods take on the role to support computational lead discovery approaches. The studies herein represent a series of projects which deal with many com-plex protein targets, ranging from models of LGICS (Paper I & II) through to epigenetic target proteins (Paper I, III, & IV). These studies are unified by using fragments as starting points for lead discovery, not only for understand-ing the interactions, but also how we can use various tools at our disposal to improve and accelerate the drug discovery process. It is therefore crucial to consider how target proteins or compound libraries are obtained. Targets needed to be expressed and purified with a high purity and quality, while a

Page 31: Fragment-based drug discovery - DiVA portal

31

fragment library had to be curated. Practical considerations for experimental screening strategies using biosensor-based approaches had to be considered, including: immobilisation methods, assay buffers, protein stability, control ex-periments, and data analysis. It was essential to establish a set of robust meth-ods to confirm series of fragment hits using the biosensor-based approaches. A series of orthogonal methods was chosen to validate fragment hits, these included X-ray crystallography (XRC), Microscale thermophoresis (MST), and Differential scanning fluorimetry (DSF).

Target proteins Often dismissed as a trivial task, the expression, purification and molecular manipulation of proteins should not be dismissed. The studies described herein are presented with a focus on the experimental outcome in terms of an assay being developed and or hits being identified or advanced in one form or another, a success story if you will. However, a considerable effort was re-quired to get to a stage where the target protein could be obtained in a suitably pure and active form. Many other projects were initiated but were abandoned due to difficulties in producing the proteins in required amounts or suitable forms, these included LSD1n,46 a neuronal specific splicing variant of canon-ical LSD1,47 Lysine Specific Demethylase 2 (LSD2),48 and a carbonic anhy-drase from Trypanosoma cruzi.49 The successfully acquired proteins described herein and in further detail in the papers, were expressed and purified according to previously published proce-dures when possible,50–53 although they often required optimisation. Both pro-karyotic and eukaryotic expression systems were used. The purity, stability and functionality of target proteins were assessed using a series of biochemical and biophysical techniques. Proceeding with a protein which is unsuitable be-cause of stability, activity or purity issues is a futile endeavour and can lead to suboptimal or misleading experimental outcomes.

Page 32: Fragment-based drug discovery - DiVA portal

32

Compound libraries A total of three structurally diverse compound libraries were used in the stud-ies described herein. A library containing 61 drug-like compounds and pep-tides was put together specifically for a screen against SMYD3 in Paper III. To increase the chemical space of the compounds to be screened against a number of other targets, two other fragment libraries were used instead. The smaller and simpler fragment library of the two contained only 90 compounds (FL90). It was repurposed from a crystallographic screening library54 and was screened against 3 Trypanosoma species (Paper I). It was seen as a managea-ble size, affordable to many labs regardless of infrastructure for screening. A larger library with 1056 compounds (FL1056) was collated from multiple sources specifically for the work described in this thesis, thus enabling the exploration of a much larger chemical space. It encompassed a set of frag-ments from the Chemical Biology Consortium Sweden (CBCS) and com-pounds synthesized within the FRAGNET consortium. The latter included compounds which contained 3D modalities55, an often debated theory56 in FBDD but an area of interest in FRAGNET. FL1056 was used in Papers I & II. An analysis of the structural diversity of the FL90 and FL1056 libraries was conducted using a plot of the principal moments of inertia (PMI), Figure 14.

Figure 14 PMI plots showing the shape diversity in FL90 (a) and FL1056 (b) high-lighting the shape diversity in the larger FL1056.

In Paper IV, much larger chemical libraries were explored using computa-tional methods, where in principle millions of compounds are screened virtu-ally and a handful of the most interesting compounds are followed up experi-mentally.

Page 33: Fragment-based drug discovery - DiVA portal

33

Discovery of fragment hits (Papers I & II) How do you find fragment starting points for lead discovery? What screening methods can be used to do this? What do we need to initiate such a campaign? In two separate studies, based on different screening technologies, we lay the foundations for what is required, and what considerations should be accounted for when embarking on a FBDD project. Practical issues, control experiments, data analysis, and orthogonal validation are explored.

Improving SPR biosensor-based screening strategies for complex and difficult targets In Paper I we demonstrated how to implement the latest and most sensitive SPR instrumentation to initiate FBDD campaigns for a selection of very chal-lenging targets. A series of assays were designed to be capable of detecting the weak and transient interactions expected in FBDD, going beyond conven-tional strategies and emphasising the special experimental measures required when analysing fragments. These are difficult to detect due to their low mo-lecular weight and typically have very rapid and low affinity interactions. A series of 9 targets from challenging target classes were screened against. These included intrinsically disordered proteins (IDPs), ligand gated ion channels (LGICs), metalloenzymes, and epigenetic complexes. These targets were con-sidered applicable to a multiplexed strategy by encompassing a thorough ex-perimental design in which many interactions could be probed, and assays could be tailored to identify fragments with various binding modes, including allosteric sites and PPI interfaces. To establish these assays, many parameters had to be considered. Proteins were immobilised on the sensor chip surface using a series of methods, ex-ploring coupling techniques and their effects on binding i.e. amine coupling, thiol coupling, streptavidin capture, and cross-linked His-capture. An immo-bilisation strategy was established for each target, taking into consideration target presentation to the surface of the sensor chip and exploiting an affinity tag when available Figure 15.

Page 34: Fragment-based drug discovery - DiVA portal

34

Figure 15. Visualisation of targets immobilised to sensor surfaces in random orienta-tion via covalent amine coupling (a) and in fixed orientation via non-covalent SA-biotin conjugation (b).

In order to develop a sensing surface which can detect the weak transient in-teractions of fragments, immobilisation levels of the targets were guided by a theoretical maximum response from a fragment which was calculated using the following equation:

Equation 9

Where Rmax is the maximal response, MWanalyte is the average molecular weight in Dalton of a fragment in the fragment library, RUligand is the response of immobilised ligand and MWligand is the molecular weight of the target. When optimising assays, it is critical to establish if the immobilised protein is functional on the surface. To assess sensor surface sensitivity and stability, tool compounds (where available) were therefore used at saturating concen-trations. This revealed if the sensor surface degraded over time and permitted the calculation of Rmax for ligands interacting with sufficient affinity for satu-ration to be estimated. In the absence of controls, sensor surfaces were directly compared with each other in the presence and absence of cofactors or protein binding partners. Figure 16 shows the screening strategy set up. FL90 was screened against farnesyl pyrophosphate synthase (FPPS) from three Trypanosoma species, while FL1056 was screened against 6 targets (acetylcholine-binding protein (AChBP), lysine specific demethylase 1 (LSD1) with and without the protein cofactor COREST, tauK18, and protein tyrosine phosphatase 1B (PTP1B)). The screening cascades for FL90 and FL1056 differed, mainly by the inclu-sion of a pre-screen to clean up the newly collated FL1056 library and remove promiscuous binders. For FL90 two different SPR biosensor instruments were used, differing in experimental design and throughput. For the FL90 library, only the high-throughput instrument was used.

Page 35: Fragment-based drug discovery - DiVA portal

35

Figure 16. Screening cascade and experimental strategy for the multiplexed SPR ap-proach: An average of 1% of compounds fell out in the pre-screen. A pre-screen was deemed unnecessary for FL90 as the library was characterized by a high solubility and was well validated during crystallographic studies. Binding level screens identi-fied between 10-15 % hit rates, with these fragments being brought to hit confirma-tion, testing concentration response in an affinity screen. For FL90, 4-6 compounds were considered final hits and brought to orthogonal validation. With FL1056 final hits were 11-19 depending on the target, these too were brought to validation. Tar-gets: Farnesyl pyrophosphate synthase (FPPS), acetylcholine-binding protein (AChBP), lysine specific demethylase 1 (LSD1) with and without the protein cofactor COREST, tauK18, and protein tyrosine phosphatase 1B (PTP1B).

Hit selection criteria and fragment ranking differed in the two screening cas-cades. One of the most important steps was to confirm that the hits had kinetic profiles that represented specific interactions with the targets. Figure 17 illus-trates typical sensorgram for fragments, of which only some are considered to be associated with useful fragments.

Page 36: Fragment-based drug discovery - DiVA portal

36

Figure 17. Kinetic profiles of fragments, ranging from a preferred rapid 1:1 “square pulse” (green), rapid interactions >1:1 (red), slow dissociation (yellow) and slow secondary interactions (blue).

The data from the FL90 screen could, due to its relatively small size, be man-ually inspected after the primary screen using a single concentration of each fragment (binding level screen). Fragments which displayed signals between 30-70% of their theoretical Rmax were brought forward to the secondary screenusing a series of fragment concentrations (affinity screen). In the case ofFL1056, 10 % of hits displaying favourable kinetic profiles characterized byfast kinetics and typical 1:1 interactions with rapid kon and koff i.e. a squarepulse, were carried forward to an affinity screen (Figure 17).

Identified hits were confirmed during affinity level screening for both librar-ies. They were ranked by affinity (KDapp), ligand efficiency (LE) or via binding efficiency (BE) (Figure 18). The best hits typically had micromolar to nano-molar affinities. Those showing a clear dose dependency and with sensor-grams representing mechanistically well-defined interactions, were brought through to orthogonal validation using microscale thermophoresis (MST), X-ray crystallography, NMR, or thermal shift assays (TSA), depending on the target and the availability of orthogonal assays.

Figure 18 summarises the outcome of the screening experiments, illustrating overlaps between hits for different targets (Figure 18a) and the characteristics of selected hits (Figure 18b).

Page 37: Fragment-based drug discovery - DiVA portal

37

Figure 18. (a) Schematic of overlapping hits using FL90 (left) and FL1056 (right). (b) table containing BE, LE and KD for orthogonally validated hits. To conclude, using the strategies established in Paper I, fragment hits can be readily identified against a plethora of complex targets, often considered to be unsuitable for an SPR-biosensor-based approach. This was achieved through innovative assay design, immobilisation strategies, and data analysis. The out-come of the screening campaign was not hindered by the fragment library used for screening. In both instances, the methods described, routinely produced validated hits which can be used to feed the pipeline in a preclinical lead dis-covery setting. Hit rates differed depending on the target which was screened and ranged from ~0.5-1%, with apparent KD values in the mid nanomolar to micromolar range. Having an orthogonal assay established, preferably one which can elucidate structural information of the interaction will be key to advancing hits to leads. This study was used to identify fragment hits, charac-terized by a high affinity for the target which also show preferred kinetic pro-files for a kinetically “ideal” fragment. SPR instrumentation permitted the de-

Page 38: Fragment-based drug discovery - DiVA portal

38

tection of weak interactions but lacked information on a functional or struc-tural outcome. It provides a yes or no result with kinetic information but relies heavily on orthogonal validation for structural insight.

Extending hit identification beyond detection of binding -Detection of ligand-induced conformational changes In Paper II a new unique approach was exploited for fragment library screen-ing against targets where ligand induced conformational changes can be ex-pected. It involved using a target protein labelled with a second harmonic ac-tive (SH-active) dye to establish an assay to identify fragment hits which in-duce a conformational change upon protein:ligand complex formation [PL*] per scheme 3, i.e. SHG was used to identify ligands altering the structure [PL*], while in Paper I ligands can only be identified that bound during [PL] formation. The detection principle and an overview of SHG technology is pre-sented earlier in the thesis.

Scheme 3 Where P = protein, L= ligand, and PL*= conformational change or altered structure after protein:ligand complex formation. This project represents an entirely exploratory approach for FBDD, as it was unknown if SHG could detect fragment interactions seeing that they are weak and transient, and may not be able to induce sufficiently large conformational changes. To build this assay we chose AChBP, a model system for the ligand-gated ion channels (LGIC’s) which represents a target class where conforma-tional changes are necessary for function57 and for which the discovery of both agonists and antagonists would be of relevance for therapeutic development. AChBP is a water-soluble homologue of the extracellular domain of the lig-and-binding domain of nicotinic receptors (nAChR) and other LGICs (Figure 19). It is derived from the great pond snail Lymnaea stagnalis and has been established as a model system for studies of fundamental mechanisms of lig-and-binding, gating and ion transport in these ion channels.58

Page 39: Fragment-based drug discovery - DiVA portal

39

Figure 19. Structures of (a) a full pentameric LGIC complete with extracellular, transmembrane, and intracellular domains (PDB ID 2BG9), (b) Ls-AChBP, corre-sponding to the extracellular domain of the homopentameric LGIC. The structure is characterised by the highly conserved C-loop orthosteric site (inset) (PDB ID 1UW6).

An assay was established on wildtype AChBP to test the viability of using SHG to observe conformational changes in the protein. It showed that the tar-get could be labelled and tethered to a lipid bilayer, whilst maintaining its functionality. Furthermore, it was possible to detect the ligand induced con-formational changes with know (partial) agonists and antagonists. Once the assay had been established, we chose to challenge the system with a fragment screen. Our experiment asked three major questions: 1. Can frag-ments induce a conformational change, 2. Does this assay have the sensitivity to detect these changes, and 3. Can we design an assay that would allow us to probe dynamic regions of the protein. Questions 1 and 2 can be addressed by screening the wildtype protein but question 3 would require additional thought with a more complicated experimental design. To address question 3, a series of single cysteine mutants were engineered towards conformationally dy-namic regions of the protein through site directed mutagenesis. These mutants were expressed using the baculovirus expression system in Sf9 insect cells, and purified as previously described.53 An overview of these mutants is pre-sented in Figure 20.

Page 40: Fragment-based drug discovery - DiVA portal

40

Figure 20. AChBP shown in grey surfaces (right), and monomer of AChBP shown in cartoon with cysteine mutants highlighted and labelled (left).

Each construct went through a series of optimisations to get the required SH active probes or dye:protein ratio required for the assay. Constructs C4 and C6 failed during the conjugations stages and were not pursued for screening purposes. However, C1, C2, C3, and C5 were successfully conjugated with labelled probes and gave expected responses to tool agonists and antagonists.

These constructs were then individually screened with the same fragment li-brary from Paper I i.e. FL1056 using the cascade outlined (Figure 21a). A hit rate between 10-22% was observed depending on the construct, with many hits overlapping between constructs. A subset of hits which hit across wildtype, C3, C5 were selected for orthogonal validation by XRC and GCI.

Page 41: Fragment-based drug discovery - DiVA portal

41

Figure 21. SHG biosensor-based screening of a fragment library against AChBP. (a) Screening cascade for each of the screened constructs. (b) Hits identified on each corresponding construct and overlap between constructs, and (c) Hits and hit rate expressed as a percentage of total library.

It can be concluded from Paper II that screening of the same fragment library as with SPR, i.e. FL1056, we could use SHG to identify a set of bone fide hits across a series of constructs of AChBP. In this study we showed that fragments are indeed capable of inducing conformational changes and furthermore, that the SHG assay developed had the sensitivity to detect these changes. Different regions of a protein can thus be probed through the introduction of site-spe-cific probes. Although not entirely necessary for a successful outcome, they permit a greater level of detail about the location of binding sites for hits.

Page 42: Fragment-based drug discovery - DiVA portal

42

Conclusions: Fragment library screening, SPR vs. SHG Ideally, when embarking on a FBDD screening campaign, one would prefer to have a technique which is optimal in terms of quantitative and qualitative data readouts. Having a technique which is all “pros”, with little or no “cons” to identify putative hits is probably more of a pipedream than a reality in the immediate future. When deciding which technique to use, there are many pa-rameters to be considered, ranging from the level of information attained, throughput, assay development, and time required. Using SPR for label free interaction analysis is considered by many the gold standard, its adoption by industry speaks for its robustness and reliability, it requires small quantities of target protein and assays can be developed in a relatively short time. SHG is still in its infancy and requires a considerable investment in assay develop-ment, but offers unrivalled data for observing conformational change in real time. Both techniques reliably identify hits which could be validated using orthogonal methods. Interestingly, applying the most stringent hit calling thresholds in both assays resulted in no overlap of hits between the two techniques. However, if we go back and reanalyse the hit calling thresholds from the SPR screen, it clearly shows SHG hits are being omitted. In such a scenario, it could be postulated that perhaps we are excluding functionally relevant fragments in hit calling when using affinity as the key selection criteria. Highlighting further, the pre-viously unmet need for biophysical techniques to correctly characterize real time complex conformational changes which are induced upon ligand binding.

Screening for allosteric ligands and confirming their binding modes (Paper III) Sometimes fragment hits are found interacting elsewhere, other than the ac-tive-site, and this can be used to our advantage. It is possible to design screen-ing strategies to probe and indeed sniff out these allosteric sites. Are there additional experimental strategies or designs to help guide this discovery pro-cess? Can computational methods verify, guide or indeed validate the experi-mental process? In this study combining biophysical and computational meth-ods help us get around limitations of a purely experimental approach. In Papers I-II, fragments were readily identified using a series of different bi-ophysical techniques and experimental assay design. In paper III we returned to SPR to screen an epigenetic target, SET and MYND domain-containing protein 3 (SMYD3) a lysine methyl transferase (KMT), with a complex role in cellular regulation. The validation of SMYD3 as a viable target remains elusive, it is suggested to be of relevance in certain cancer pathologies, but

Page 43: Fragment-based drug discovery - DiVA portal

43

concrete evidence is lacking.59 SMYD3 is characterized by two known bind-ing sites: 1. A co-factor site S-adenosyl methionine (SAM), and 2. A substrate site connected by a methylation tunnel to 1, which accommodates the methyl-ation on the substrate, shown in Figure 22.

Figure 22. SMYD3 highlighting the SAM co-factor site, SAM shown in blue sticks, and EPZ031686 bound in the substrate site. Zinc atoms shown in green PDB ID 5CCM.

There have been two inhibitors of SMYD3 previously described,60 both bind-ing to the substrate binding site, EPZ031686 and EPZ030456. However, we chose to develop an assay which aimed to identify allosteric sites which may help uncover some of the complexities of this elusive methyltransferase, or perhaps provide a validated hit for further exploration in the hit-lead process. In this experimental strategy a competitive biosensor-based screen was estab-lished, with a total of three different protein surfaces representing APO en-zyme, active site blocked enzyme, and chemically denatured enzyme.

A racemic compound, diperodon with a molecular weight of 384 Da. was identified as a hit using the experimental strategy outlined in Figure 23.

Page 44: Fragment-based drug discovery - DiVA portal

44

The hit was further validated using XRC, where the structural information of the interaction was elucidated at the opposite face of the protein with respect to the active site. In parallel to the first XRC studies, in silico methods were explored to predict any potential binding sites other than the previously stud-ied active site and cofactor site. The identification of potential binding pockets was predicted using a combination of solvent mapping and pocket detection using pyMDmix61 and fPocket45. A series of sites were predicted and scored, most notably the diperodon site was correctly predicted and divided into two sub pockets ranked 4th and 8th out a potential 15 pockets which were identified as highlighted in Figure 24.

Figure 23. Screening of a compound library against SMYD3. (a) An overview the biosensor layout, (b) data for the focused library of 40 compounds which were screened highlighting identified hit, and (c) chemical structure of the screening hit.

Page 45: Fragment-based drug discovery - DiVA portal

45

Figure 24. Identification of potential allosteric sites by in silico pocket detection. (a-d) Potentially druggable cavities identified by fPocket calculations. The surfaces show contours of residues lining the pockets, (b) showing the active site, (c) Diperodon binding site (split into two pockets: P4 and P8) and other pockets, and (d) the SAM site.

Table 1. Overview of predicted pockets and scores

This analysis indicated that the diperodon binding pocket is indeed druggable and can be exploited later to develop a series of other ligands which may have completely different chemical scaffolds with optimised physicochemical pa-rameters. This pocket was further scrutinised to investigate a potential PPI with HSP90 using biosensor-based analysis, which was subsequently compli-mented with cellular assays using colon cancer cell lines. Although the SMYD3 HSP90 interaction could be confirmed, with a relatively weak affin-ity estimated at 13 µM, no direct competition between HSP90 and diperodon for their interaction with SMYD3 was observed. Perhaps analogues of dipero-don with a higher affinity may be capable of direct competition.

Page 46: Fragment-based drug discovery - DiVA portal

46

Evolution of ligands – a combined biophysical and in silico approach (Paper IV) Having identified hits, what happens next? How do we progress these ligands to lead like compounds? Are there additional methods to aid the discovery process, other than embarking on a traditional SAR or medicinal chemistry routes? An allosteric ligand with a relatively weak affinity (43 μM) was discovered in paper III, (Figure. 25). We here asked how we could use a combination of in silico and biophysical methods to improve the interaction of the compound with its target while ensuring that it had suitable physicochemical properties. Optimisation of affinity and kinetic parameters are essential for progressing a hit to a lead, and could at the same time permit the discovery of novel chemical probes. In the case of SMYD3, the allosteric site is thought to be involved in a PPI interaction with the C-terminal domain of HSP90,62 but its in vivo func-tion has not yet been established. Here we chose a bidirectional approach to explore the chemical space available to us for optimisation of diperodon. In the first instance we chose to explore a SAR by catalogue approach, and in the second a computational compound growing strategy. For both strategies, the compounds of interest were acquired and tested experimentally for binding to SMYD3 using a GCI biosensor assay developed specifically for the purpose.

Figure 25. (a) crystal structure of SMYD3 with diperodon bound in the allosteric site interacting with GLU 189 and LYS378. (b) bonding network found in PDB ID 7BJ1. The first approach resulted in the selection of 21 compounds for experimental evaluation, of which 2 were observed to bind with a higher affinity for SMYD3 than diperodon. An overview of the validated compounds is shown in Table 2.

Page 47: Fragment-based drug discovery - DiVA portal

47

In the second approach diperodon was stripped back to its most basic fragment features and a fragment evolution platform was applied to try and make sig-nificant gains in affinity. At each step the fragment was evolved 2 heavy atoms at a time over a total of 7 iterations. The criteria were such that the compounds should be available to order in the commercially available space. A total of 11 compounds were acquired and tested, with 3 compounds showing a simple 1:1 reversible interaction with favourable affinities above the control (Table 2). Table 2. Successfully evolved allosteric SMYD3 ligands with improved affinity over controls, using two computational approaches.

Concluding thoughts on Paper IV: in an academic setting, a bottleneck in the development of lead like compounds depends greatly on the expertise of the group. In FBDD, a tight collaborative approach is required in going from hit-to-lead. It requires relentless synergy across multiple disciplines. In our case the bottleneck was rigorous medicinal chemistry and iterative synthesis of an-alogues, which can take a team of FTE weeks or months to get through. Our approach was to try and “hack” this tried and tested method by combining novel computational strategies complemented with near neighbour analogues, all within the commercially available space. A key criterion was that com-pounds should be available within seven days of our decision to proceed or not. This project is continuing to evolve and has not yet reached a final state where optimised ligands have been identified. It requires iterative rounds guided by crystallography. Although the biophysical assays confirmed binding and in some cases with a higher affinity than our starting compound, we did not ob-serve the ten-fold or hundred-fold gains in affinity that could be expected in going from hit to candidate nomination.63 If structural elucidation had proved positive it could easily be seen how this would guide the design of subsequent compounds.

Page 48: Fragment-based drug discovery - DiVA portal

48

Another reality is that the pocket we have identified and characterized may not be suitable for a fragment growing strategy. Furthermore it may be that we have hit the goldilocks zone with diperodon, meaning that we have hit the sweet spot in ligand optimisation. Strategies moving forward may include structural elucidation via NMR or labelled assays as described in Paper II. If affinity permitted, this computational compound design may enable the devel-opment of a proteolysis targeting chimera (PROTAC) i.e. a heterobifunctional ligand inducing target degradation via hijacking of the ubiquitin-proteasome degradation machinery.

Page 49: Fragment-based drug discovery - DiVA portal

49

Conclusions

The overall goal of this thesis was to develop biophysical methods and to ap-ply them to fragment-based drug discovery, beyond the scope of what is cur-rently considered to be state-of-the-art. It has been motivated by the fact that investments in R&D by big pharma has increased dramatically in the last dec-ades, and the realisation that the number of approved drugs per year has never been as low. The previous brute force approaches of discovering drugs i.e. screening millions of compounds against a target is proving less and less use-ful. The need for rational and creative experimental design, combined with state-of-the-art biophysical and computational methods can aid the rapid de-velopment of therapeutics into the future. By understanding and evaluating the molecular interactions of ligands with their target protein, affords an un-precedented opportunity to play a part in the development of the next genera-tion of medicines. It is expected that the major findings of this thesis will play a role in that fu-ture, and my specific contributions can be summarised as follows.

• Developed new strategies for the identification of fragments interact-ing with several complex and dynamic drug targets by multiplexing SPR biosensor assays, and the application of hit validation methods.

• Established SHG biosensor assays for detection of fragment induced conformational changes and showed proof-of-principle for fragment library screening against a conformationally dynamic target.

• Established strategies for structural analysis of fragment hits using a combination of experimental and computational methods, and identi-fication of fragment evolution paths for AChBP, LSD1 and, SMYD3.

• Evolution of hits against SMYD3 using a multifaceted approach where biophysical and computational methods go hand in hand in ac-celerating the discovery process.

These findings have played an important role in the FRAGNET consortium, and it is expected they will form the basis of future drug discovery projects.

Page 50: Fragment-based drug discovery - DiVA portal

50

Future at a glance

This thesis was driven by a personal curiosity to further understand drug dis-covery and more specifically, interactions at a molecular level. In my previous work, I was involved in developing 3D human tissue-tissue cell culture mod-els of various organ types. These models were applied to innovative microflu-idics and were later used for drug development.64–68 However, in my opinion, there is a disconnect between what is happening at a macromolecular level i.e., at a cellular level, and at the molecular level in terms of drug discovery. If we are to be successful in the development of to-morrow’s drugs there needs to be a stronger interplay between the biophysical methods and cellular methods. Putting observed interactions from biophysical experiments into a cellar context with real time detection, may have profound consequences on the attrition rates in the drug discovery process. In this thesis fragment compounds are discovered interacting with epigenetic enzymes, whether they will have a functional outcome is yet to be elucidated. The beauty of targeting epigenetic enzymes based on overexpression in disease pathologies means one can either knock out their catalytic activity or, if com-pounds are interacting outside the active site, one hopes they can interfere with a PPI causing a loss of function. In the future, one could envisage taking a more holistic approach to drug dis-covery, by combining knowledge from both disciplines. In such a future the development of PROTAC molecules applied to cancer models using organ-on-chip technology could unify the discovery process between microfluidic models and remove years from the drug discovery timelines.

Page 51: Fragment-based drug discovery - DiVA portal

51

Sammanfattning

Utveckling av nya läkemedel är ett brett forskningsområde som har en stor betydelse för hur vi kan behandla många vanliga sjukdomar idag och nya som vi ännu inte känner. År 2020 dominerades av ett tidigare okänt virus, SARS CoV-2, som orsakar sjukdomen COVID-19. Vi har alla bevittnat hur effek-terna av viruset slagit hårt över hela världen. COVID-19 har blivit den pan-demi som många fruktat skulle drabba oss, men ingen visste vilket virus eller annan organism som skulle komma eller när. Under 2020 har kunskapen om viruset och dess egenskaper ökat mycket snabbt. Olika strategier för att för-hindra dess spridning och hur sjukdomen skulle behandlas har testats och framtagningen av ett vaccin har varit en enastående vetenskaplig framgång. Men det saknas fortfarande specifika läkemedel som är effektiva för behand-ling av patienter som infekteras och riskerar att utveckla svår COVID-19. I denna avhandling har målet varit att ta fram nya metoder för att utveckla nya läkemedel. Forskningen utgår ifrån en strategi benämnd ”fragment-baserad läkemedelsutveckling” (eng. Fragment-Based Drug Discovery, FBDD). Stra-tegin introducerades i mitten av 1990-talet och har blivit mycket populär ef-tersom den gör att forskare snabbt kan komma igång och visa vägen, även med relativt begränsade resurser. Det finns flera viktiga orsaker till att strategin fick ett snabbt genombrott, men till stor del handlar det om att nya känsliga analytiska tekniker utvecklats och blivit mer allmänt tillgängliga, enklare att använda och effektivare i analys av många prover. Att välja rätt teknik och anpassa den till projektet är dock kritiskt för att lyckas med FBDD. Utveckling och användning av biosensorer för fragment-baserad läkemedels-utveckling har varit ett av de viktigare teknikutvecklingsområdena för strate-gin sedan starten. Avhandlingen bidrar konkret till denna forskning. Biosen-sorer används för att identifiera små och relativt enkla molekyler (fragment) som kan binda till det protein som valts som läkemedlets måltavla (eng. ”tar-get”). Om målproteinet har en central roll i sjukdomen som ska behandlas kan bindningen av den lilla molekylen till dess yta göra att proteinet förlorar sin funktion och därmed nås en önskad klinisk effekt. De fragment som identifi-erats fungerar som tidiga läkemedelsprototyper (eng. ”lead compound”). Ef-tersom läkemedel vanligen är större och mer komplexa än dessa små läkeme-delsprototyper, sker en optimering av prototyperna genom att metodiskt för-ändra dess kemiska struktur och/eller lägga till viktiga kemiska grupper så de

Page 52: Fragment-based drug discovery - DiVA portal

52

får önskade egenskaper. För att lyckas i läkemedelsprojekt behövs en god till-gång till målproteinet i en form som är lämpligt för de metoder man väljer för sin strategi. Samtidigt behövs det stora bibliotek med kemiska föreningar som kan screenas och kemiska metoder för att ta fram nya substanser med optime-rade egenskaper. De projekt som beskrivs i avhandlingen har tagit FBDD-forskningen vidare genom att dels utveckla nya metoder som bättre utnyttjar potentialen hos redan väletablerade biosensorer, och dels utforskat potentialen hos nya typer av bi-osensorer. Biosensorer har blivit viktiga inom fragment-baserad läkemedels-utveckling eftersom de i realtid och med hög känslighet kan mäta hur en mo-lekyl binder till ett protein som fästs på en yta. Tekniken används tidigt i pro-cessen när man letar prototyper och även genomgående under optimeringen av prototyperna till färdiga läkemedelsmolekyler. Forskningen som presente-ras i avhandlingen har även omfattat dator-baserade studier för att bättre förstå vilka prototyper som är mest intressanta och hur de kan optimeras. Metoderna har kombinerats för att lösa utmaningar i projekt där nuvarande tekniker inte fungerar tillräckligt bra. Orsaken till svårigheterna att använda biosensorer och andra instrument för fragment-baserad läkemedelsutveckling är ofta att proteinet som är intressant som mål för läkemedlet inte är helt enkelt att hantera i en form som gör att den analytiska tekniken fungerar. Syftet med forskningen som presenteras i av-handlingen har varit att utveckla nya biosensor-baserade metoder som funge-rar på proteiner som ofta undvikits som mål för läkemedel av praktiska skäl, även om de skulle vara lämpade rent biologiskt. De många målproteiner som ingått i studierna har varit involverade i sjukdo-mar innefattande sömnsjuka (tbFPPS, tcFPPS, hFPPS), olika hjärnfunktioner (AChBP) och demenssjukdomar (tau), samt cancer (LSD1, SMYD3). Alla målproteiner har bidragit till metodutvecklingen och med de nya metoderna har nya läkemedelsprototyper och principer för hur de kan optimeras till läke-medel identifierats. Metoderna förväntas kunna användas för att utveckla mor-gondagens mediciner, t.ex. mot sjukdomar som COVID-19.

Page 53: Fragment-based drug discovery - DiVA portal

53

Acknowledgements

There is an old African adage “it takes a village to raise a child”. What it takes to turn a bright eyed, bushy tailed student into a PhD scholar is something far greater, less tangible, and involves a great deal of dedicated mentors, col-leagues, collaborators, friends and most importantly family. Helena, your mentorship over the last four years has been transformative. Your patience, guidance and dedication to teaching is inspiring. Thank you for taking a chance on an Irish chancer, and for making me feel so welcome in your group. Your support throughout the entirety of the PhD has been un-relenting. Not only have you taught me the joys of biophysics and how to do creative science in academia, but you also taught me to teach, to understand and to listen. More importantly, you have bestowed upon me the great art of Swedish pragmatism, a life skill I thank you for. To my colleagues at Beactica, Wei, Rajiv, Mattis, Thomas, and Ulf, thank you for all the insightful discussions, meetings, coffee breaks and after work get togethers. Johan, for your patience, knowledge and guidance during my first fragment screens at Beactica, I thank you. Konrad my co-supervisor, thanks for all your help with chemical structures and identifying compounds, I still think we have a nanomolar PROTAC in us, just got to get that crystal-lography to behave. A sincere thank you to Per for your support, inspiring discussions and for passing on the wise knowledge that you acquired that sometimes “no data is good data”. It would be impossible to make it this far without the top-notch crew, past and present, we have in the Danielson Lab. I want to take a moment to sincerely thank students that worked on projects with me, Louise, Michael, Hildur, Mia Z, and Mia A your input, help, and creativity in progressing some of our trickiest projects cannot be underestimated. We don’t do easy in the Danielson lab but we do have fun! Mia A, I am so pleased you are continuing with your SCIENCEEE and wish you all the best with your PhD. Vladimir thank you for putting up with my discussions on the weather, and for your expertise with SMYD3. Giulia it was a blast, you taught me that no matter what everything will be Bene. Running that big study at Cytiva was a tough feat but we got there, thank you. Daniela the mother of crystals, thank you for putting up with me as an office mate and for not killing me for every time I had some wild

Page 54: Fragment-based drug discovery - DiVA portal

54

new idea to try with fragments and crystallography. Mostafa thanks for being a great colleague and always having kind words. Gun, whenever I had a prob-lem you were there. Thank you for all of our chats, the advice, and for teaching me some of your tricks of the trade. I will never forget your reaction when I stood proudly in front of a hideous set-up I rigged for a batch purification, needless to say we optimised it. Nadine, wish we had more time together, it has been a joy. Thank you for all the pep talks during my writing periods. Team SciLife shout out to you Helena N and Eldar for listening to my wan-dering thoughts about experiments and for your advice with all things Biacore. Eldar thanks for being my lunch buddy and for never bringing a lunchbox the first few years. Team USUN, Sandra, Kyhati and Emil, thank you for all the laughs and for the advice during group meetings. To my mentors, teachers and colleagues at Biokemi, Micke, thank you for shar-ing your knowledge and enthusiasm of biochemistry. Gunnar, for your inspir-ing discussions and for instilling in me your passion for biochemistry, I thank you. Doreen, a sincere thank you for your continued support with all of crystal-lography. A special mention to Eva and Francoise, thank you for putting up with my unannounced visits and for your support and help over the years. To the rest of my budding biochemists on the third floor, and that hard-to-find location where Joana, Maxim and Emiliano sit…thank you for keeping me so-cial. To the ever-expanding Ivarsson group, and my co-supervisor Ylva, thank you for making the corridors a happy place to be. Johanna thank you for really listening at times when it was most needed, you don’t need me to wish you luck with the rest of your PhD…luck is for chancers, you are a winner. Keep killing it! Caroline your life advice was always welcomed, thank you for all the chats. Your dedication in the lab and work ethic is inspiring. You got this, keep smashing it. Ali javviiissst you made it, thanks for all the ridiculous conversations we shared at ridiculous hours of the night at the BMC. Leandro thanks for your help with my troublesome spreadsheets, always helpful, always smiling, thank you. Susanne I feel like I done the highest concentration of laughing with you, thank you for laughing at my terrible jokes. Maxim I am still waiting for that ski trip, failing that, let’s just go to Germany and hang with Brunhilda’s dad? Best of luck with everything mate. Dilip thank you for being a great office mate. Gina thanks for all the chats and good luck with everything. To the ESRs of Fragnet, the Fraglets ! What a journey, thank you for the “good times, arrright”. To the JNTα crew, we laughed, and then we laughed some more. To my right honourable gentleman Darius, thank you for being you. Sebastian a true friend, thank you for all the memories in San Diego and in Basel. To Moira for all the hard work on our papers, thank you. Pierre thanks for shedding light on the crystals, see what I done there…it was great having you in Uppsala and getting to spend time with you at the home source

Page 55: Fragment-based drug discovery - DiVA portal

55

at the VU. Thank you to the PIs and especially to Iwan and Jaqueline for making the Fragnet journey extra memorable. To my Californian friends at Biodesy, thank you for all the great memories working in South San Francisco. Shout out to the lunch crew Bruce, Ben, Gabriel and David. Serra, Tracey and Ariel thank you for supporting the project and for hosting me. An enormous thank you to Margaret for all those crazy long days we shared, filled with butterfingers and espresso. Your dedi-cation and excel skills are mind blowing. Thank you for your patience and for teaching me the joys of a Bravo. There is an international team of friends who have had a profound impact dur-ing my PhD. Claire, thank you for being my office mate when I first arrived to this new land, for sharing a semla when cloning sucked and most im-portantly for your continued presence over the last few years. Taking me under your wing and our road trip in Cali. Lasting memories I will cherish. Valerie, thank you for always checking in on me, for your advice and for the phone calls when they were really needed. To Derek and Kerri, my Boston parents thank you for taking care of me when I am on the other side of the Atlantic. Ms. Kerri your pep talks, peeps and care packages have all helped me to make it this far. Derek, I reckon the five guys played a part too. To the Swedes in my life, Anders and Mathilde thank you for making me feel so welcome and for all the memorable evenings at Bergagatan. Linus thank you for keeping me social this summer. Anna and Max, for keeping me motivated, fed (you rival my grandmother) and inspired during the last years. Anna thank you for entertaining my silly ideas “NeuroPrecis” and for your guidance. Erik, you’re a scholar, a gent and a good judge of bad horses. Thank you for all the co fe fe breaks over the last few years, I hope the committee of longfellows can continue to meet post PhD. Vanessa the biggest legend in existence, thank you for you for your omnipresence and for being you. Kom-pis aka Josefine chairperson of the committee of Nycklegris, your support and friendship over the last twelve months has been invaluable. Thank you for teaching me how to be Swedish, how to midsummer, and how to cook!! To my family and their unrelenting support in my pursuit of “just one more piece of paper”. To Mum, Dad and Jenna, the best is yet to come. To my grandparents, Maura and Edward Snr, Granny I promise I am done. The phone calls to The Boatport on Sundays got me through the toughest times. I would not be where I am without you. Lian my Yoko, you’ve supported me in more ways than you can imagine. I cannot wait to spend the rest of my life with you.

Frodegatan 5G, January 7, 2021

Page 56: Fragment-based drug discovery - DiVA portal

56

References

1. Organization, W. H. & others. Coronavirus disease 2019 (COVID-19): situation report, 72. (2020).

2. Gaedcke, F. Ueber das Erythroxylin, dargestellt aus den Blättern des in Süd-Amerika kultivierten Strauches Erythroxylon Coca Lam. (1854).

3. Willstätter, R. & Hollander, C. Synthese der Ecgoninsäure. Justus Liebigs Ann. Chem. 326, 79–90 (1903).

4. Lanska, D. J. JL Corning and vagal nerve stimulation for seizures in the 1880s. Neurology 58, 452–459 (2002).

5. Small, D. Masters of Healing: Cocaine and the Ideal of the Victorian Medical Man. J. Vic. Cult. 21, 3–20 (2016).

6. Chandler, A. D. Shaping the industrial century: The remarkable story of the evolution of the modern chemical and pharmaceutical industries. 46, (Harvard University Press, 2009).

7. KONG, K.-F., Schneper, L. & Mathee, K. Beta-lactam antibiotics: from antibiosis to resistance and bacteriology. Apmis 118, 1–36 (2010).

8. Burley, S. K. et al. Protein Data Bank (PDB): the single global macromolecular structure archive. in Protein Crystallography 627–641 (Springer, 2017).

9. Lipinski, C. A., Lombardo, F., Dominy, B. W. & Feeney, P. J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 23, 3–25 (1997).

10. Hann, M. M. Molecular obesity, potency and other addictions in drug discovery. Medchemcomm 2, 349–355 (2011).

11. Lamoree, B. & Hubbard, R. E. Current perspectives in fragment-based lead discovery ( FBLD ). 0, 453–464 (2017).

12. Shuker, S. B., Hajduk, P. J., Meadows, R. P. & Fesik, S. W. Discovering High-Affinity Ligands for Proteins: SAR by NMR. Science (80-. ). 274, 1531–1534 (1996).

13. Baker, M. Fragment-based lead discovery grows up. Nat. Rev. Drug Discov. 12, 5–7 (2013).

14. Renaud, J.-P. et al. Biophysics in drug discovery: impact, challenges and opportunities. Nat. Rev. Drug Discov. 15, 679 (2016).

15. Keserű, G. M. et al. Design principles for fragment libraries: maximizing the value of learnings from pharma fragment-based drug discovery (FBDD) programs for use in academia. J. Med. Chem. 59, 8189–8206 (2016).

16. Congreve, M., Carr, R., Murray, C. & Jhoti, H. A’rule of three’for fragment-based lead discovery? Drug Discov. Today 19, 876–877 (2003).

Page 57: Fragment-based drug discovery - DiVA portal

57

17. Tsai, J. et al. Discovery of a selective inhibitor of oncogenic B-Raf kinase with potent antimelanoma activity. 105, 3041–3046 (2008).

18. Murray, C. W., Newell, D. R. & Angibaud, P. A successful collaboration between academia{,} biotech and pharma led to discovery of erdafitinib{,} a selective FGFR inhibitor recently approved by the FDA. Med. Chem. Commun. 10, 1509–1511 (2019).

19. Souers, A. J. et al. ABT-199, a potent and selective BCL-2 inhibitor, achieves antitumor activity while sparing platelets. Nat. Med. 19, 202–208 (2013).

20. Tap, W. D. et al. Structure-Guided Blockade of CSF1R Kinase in Tenosynovial Giant-Cell Tumor. N. Engl. J. Med. 373, 428–437 (2015).

21. Romasanta, A. K. S. et al. Career development in fragment-based drug discovery. Drug Discov. Today Technol. (2020). doi:https://doi.org/10.1016/j.ddtec.2020.10.001

22. FRAGNET. Available at: https://fragnet.eu/. (Accessed: 7th December 2020)

23. Koshland Jr., D. E. The Key–Lock Theory and the Induced Fit Theory. Angew. Chemie Int. Ed. English 33, 2375–2378 (1995).

24. Mosca, R., Céol, A. & Aloy, P. Interactome3D: adding structural details to protein networks. Nat. Methods 10, 47 (2013).

25. Ruffner, H., Bauer, A. & Bouwmeester, T. Human protein--protein interaction networks and the value for drug discovery. Drug Discov. Today 12, 709–716 (2007).

26. Johnsson, B., Löfås, S. & Lindquist, G. Immobilization of proteins to a carboxymethyldextran-modified gold surface for biospecific interaction analysis in surface plasmon resonance sensors. Anal. Biochem. 198, 268–277 (1991).

27. Patko, D., Cottier, K., Hamori, A. & Horvath, R. Single beam grating coupled interferometry: high resolution miniaturized label-free sensor for plate based parallel screening. Opt. Express 20, 23162–23173 (2012).

28. Kozma, P., Kehl, F., Ehrentreich-Förster, E., Stamm, C. & Bier, F. F. Integrated planar optical waveguide interferometer biosensors: A comparative review. Biosens. Bioelectron. 58, 287–307 (2014).

29. Kozma, P., Hámori, A., Kurunczi, S., Cottier, K. & Horvath, R. Grating coupled optical waveguide interferometer for label-free biosensing. Sensors Actuators B Chem. 155, 446–450 (2011).

30. Healthcare, G. E. BiacoreTM Assay Handbook. Gen. Electr. Co. Boston, MA, USA (2012).

31. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 40, 1361–1403 (1918).

32. Geitmann, M. et al. Interaction Kinetic and Structural Dynamic Analysis of Ligand Binding to Acetylcholine-Binding Protein. Biochemistry 49, 8143–8154 (2010).

33. Millett, F. & Raftery, M. A. An NMR method for characterizing conformation changes in proteins. Biochem. Biophys. Res. Commun. 47, 625–632 (1972).

Page 58: Fragment-based drug discovery - DiVA portal

58

34. Tran, R. J., Sly, K. L. & Conboy, J. C. Applications of Surface Second Harmonic Generation in Biological Sensing. Annu. Rev. Anal. Chem. (Palo Alto. Calif). 10, 387—414 (2017).

35. Moree, B. et al. Protein Conformational Changes Are Detected and Resolved Site Specifically by Second-Harmonic Generation. Biophys. J. 109, 806–815 (2015).

36. Erlanson, D. A., Davis, B. J. & Jahnke, W. Fragment-Based Drug Discovery: Advancing Fragments in the Absence of Crystal Structures. Cell Chem. Biol. 26, 9–15 (2019).

37. Blundell, T. L., Jhoti, H. & Abell, C. High-throughput crystallography for lead discovery in drug design. Nat. Rev. Drug Discov. 1, 45–54 (2002).

38. BRAGG, W. H. The Reflection of X-Rays by Crystals. Nature 91, 477 (1913).

39. Cowtan, K. Fitting molecular fragments into electron density. Acta Crystallogr. Sect. D Biol. Crystallogr. 64, 83–89 (2008).

40. Hartshorn, M. J., Murray, C. W., Cleasby, A., Frederickson, M. & Tickle, I. J. Fragment-Based Lead Discovery Using X-ray Crystallography. J Med Chem 48, 403–413 (2005).

41. Arcon, J. P. et al. Molecular dynamics in mixed solvents reveals protein--ligand interactions, improves docking, and allows accurate binding free energy predictions. J. Chem. Inf. Model. 57, 846–863 (2017).

42. Alvarez-Garcia, D. & Barril, X. Molecular simulations with solvent competition quantify water displaceability and provide accurate interaction maps of protein binding sites. J. Med. Chem. 57, 8530–8539 (2014).

43. Ruiz-Carmona, S. et al. rDock: a fast, versatile and open source program for docking ligands to proteins and nucleic acids. PLoS Comput Biol 10, e1003571 (2014).

44. Ruiz-Carmona, S. et al. Dynamic undocking and the quasi-bound state as tools for drug discovery. Nat. Chem. 9, 201–206 (2017).

45. Guilloux, V. Le, Schmidtke, P. & Tuffery, P. Fpocket : An open source platform for ligand pocket detection. 11, 1–11 (2009).

46. Laurent, B. et al. A specific LSD1/KDM1A isoform regulates neuronal differentiation through H3K9 demethylation. Mol. Cell 57, 957–970 (2015).

47. Shi, Y. et al. Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell 119, 941–953 (2004).

48. Fang, R. et al. Human LSD2/KDM1b/AOF1 regulates gene transcription by modulating intragenic H3K4me2 methylation. Mol. Cell 39, 222–233 (2010).

49. Nocentini, A., Cadoni, R., Dumy, P., Supuran, C. T. & Winum, J.-Y. Carbonic anhydrases from Trypanosoma cruzi and Leishmania donovani chagasi are inhibited by benzoxaboroles. J. Enzyme Inhib. Med. Chem. 33, 286–289 (2018).

50. Opassi, G. et al. Establishing Trypanosoma cruzi farnesyl pyrophosphate synthase as a viable target for biosensor driven fragment based lead discovery. Protein Sci. pro.3834 (2020). doi:10.1002/pro.3834

51. Yang, J. et al. Macrocyclic Peptides Uncover a Novel Binding Mode for Reversible Inhibitors of LSD1. ACS Omega 5, 3979–3995 (2020).

Page 59: Fragment-based drug discovery - DiVA portal

59

52. Vagrys, D., Davidson, J., Chen, I., Hubbard, R. E. & Davis, B. Exploring IDP--Ligand Interactions: Tau K18 as a Test Case. Int. J. Mol. Sci. 21, 5257 (2020).

53. Celie, P. H. N. et al. Nicotine and Carbamylcholine Binding to Nicotinic Acetylcholine Receptors as Studied in AChBP Crystal Structures. Neuron 41, 907–914 (2004).

54. Huschmann, F. U. et al. Structures of endothiapepsin-fragment complexes from crystallographic fragment screening using a novel, diverse and affordable 96-compound fragment library. Acta Crystallogr. Sect. FStructural Biol. Commun. 72, 346–355 (2016).

55. Downes, T. D. et al. Design and Synthesis of 56 Shape-Diverse 3D Fragments. Chemistry 26, 8969 (2020).

56. Erlanson, D. A., Fesik, S. W., Hubbard, R. E., Jahnke, W. & Jhoti, H. Twenty years on: the impact of fragments on drug discovery. Nat. Rev. Drug Discov. 15, 605 (2016).

57. Bocquet, N. et al. X-ray structure of a pentameric ligand-gated ion channel in an apparently open conformation. Nature 457, 111–114 (2009).

58. Nys, M., Kesters, D. & Ulens, C. Structural insights into Cys-loop receptor function and ligand recognition. Biochem. Pharmacol. 86, 1042–1053 (2013).

59. Fabini, E. et al. Unveiling the biochemistry of the epigenetic regulator SMYD3. Biochemistry 58, 3634–3645 (2019).

60. Mitchell, L. H. et al. Novel oxindole sulfonamides and sulfamides: EPZ031686, the first orally bioavailable small molecule SMYD3 inhibitor. ACS Med. Chem. Lett. 7, 134–138 (2016).

61. Alvarez-garcia, D. & Barril, X. Molecular Simulations with Solvent Competition Quantify Water Displaceability and Provide Accurate Interaction Maps of Protein Binding Sites. (2014). doi:10.1021/jm5010418

62. Brown, M. A. et al. C-terminal domain of SMYD3 serves as a unique HSP90-regulated motif in oncogenesis. Oncotarget 6, 4005 (2015).

63. Rees, D. C., Congreve, M., Murray, C. W. & Carr, R. Fragment-based lead discovery. Nat. Rev. Drug Discov. 3, 660–672 (2004).

64. Maoz, B. M. et al. A linked organ-on-chip model of the human neurovascular unit reveals the metabolic coupling of endothelial and neuronal cells. 36, (2018).

65. Herland, A. et al. Distinct contributions of astrocytes and pericytes to neuroinflammation identified in a 3D human blood-brain barrier on a chip. PLoS One 11, e0150360 (2016).

66. Park, T.-E. et al. Hypoxia-enhanced Blood-Brain Barrier Chip recapitulates human barrier function and shuttling of drugs and antibodies. Nat. Commun. 10, 2621 (2019).

67. Novak, R. et al. Robotic fluidic coupling and interrogation of multiple vascularized organ chips. Nat. Biomed. Eng. 4, 407–420 (2020).

68. Maoz, B. M. et al. Organs-on-Chips with combined multi-electrode array and transepithelial electrical resistance measurement capabilities. Lab Chip 17, 2294–2302 (2017).

Page 60: Fragment-based drug discovery - DiVA portal

Acta Universitatis UpsaliensisDigital Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Science and Technology 1999

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-429950

ACTAUNIVERSITATIS

UPSALIENSISUPPSALA

2021