<<

VU Research Portal

Trypanosoma brucei B1 as a drug target for Human African Trypanosomiasis Jansen, C.J.W.

2015

document version Publisher's PDF, also known as Version of record

Link to publication in VU Research Portal

citation for published version (APA) Jansen, C. J. W. (2015). Trypanosoma brucei phosphodiesterase B1 as a drug target for Human African Trypanosomiasis.

General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal ?

Take down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

E-mail address: [email protected]

Download date: 05. Oct. 2021

An introduction into and their potential role as drug targets for neglected diseases

Chapter 1

4

CHAPTER 1 1.1 Human African Trypanosomiasis

Human African Trypanomiasis (HAT), also known as African sleeping sickness, is a deadly infectious disease caused by the kinetoplastid Trypanosoma brucei (Figure 1A). The spread of HAT is restricted to sub-Saharan Africa by the prevalence of the disease vector, several species of tsetse fly (genus Glossina, Figure 1B).1, 2 As shown in Figure 1C, the tsetse flies are found across 38 central African countries, where they are able to cause localized epidemics following infection by feeding on the blood of infected humans, livestock or wild animals.3 Following eradication efforts and given that not all tsetse flies become infected or are able to become infected, 24 countries have reported recent cases of HAT, leaving an estimated 70 million people at risk of infection.2, 4, 5 While the number of annually reported cases has remained under 10,000 since 2009, the disease has had three major epidemics in the last century.1, 6, 7 Furthermore, efforts to control the spread of the disease are hampered by the remoteness of outbreaks and conflicts in several endemic regions, leading to incomplete reporting of new HAT cases and an increased risk of another major epidemic.8

Figure 1: A) A stylized representation of Trypanosoma brucei infected blood based on scanning and transmission electron micrographs.9 B) A tsetse fly drawing blood with its

5 proboscis inserted in human skin.10 C) A map showing the spread of the distribution of the tsetse fly across Africa overlaid with the cases of HAT reported over the period 2000- 2009.11 The images in A, B and C are reproduced without adaptation from the source referenced.

Two different forms of HAT exist and these result from infection with either T.b. gambiense or T.b. rhodesiense. The two strains have different epidemiologies, with T.b. gambiense endemic to central and western Africa accounting for 97% of infections and T.b. rhodesiense endemic to the eastern Africa.5 Disease progression and therapy are also affected by the strain. Infection by T.b. gambiense progresses from the haemolymphatic phase (stage 1) to the meningoencephalitis phase (stage 2) after 2-4 years. In a few cases patients have cleared the parasites during stage 1 without treatment, however once a patient passes to stage 2 the disease is invariably fatal without treatment.12 Infection by T.b. rhodesiense is more aggressive with progression from stage 1 to stage 2 usually occurring within 8 weeks and no cases of parasitological clearance without treatment have been reported.13

The treatment options for HAT have improved in recent years. However significant issues remain.14 For HAT caused by T.b. gambiense the first-line treatment for stage 1 is pentamidine and the second-line treatment is suramin, for stage 2 the first-line treatment is NECT (nifurtimox and eflornithine) and the second-line treatment is melarsoprol (Figure 2). For HAT caused by T.b. rhodesiense the first-line treatment for stage 1 is suramin and the second line treatment is pentamadine, for stage 2 the first-line treatment is melarsoprol and there is no second-line treatment.15 Each of these treatments are regarded as essential medicines by the WHO and they are now available free of charge to endemic countries. The introduction of eflornithine and later NECT to replace melarsoprol as the treatment for stage 2 HAT caused by T.b. gambiense, has improved treatment outcomes significantly, with deaths following treatment dropping from about 6% to 1%.16, 17 The situation for stage 2 HAT caused by T.b. rhodesiense has not seen such an improvement as eflornithine and NECT are not effective against this parasite strain. Given the similarities between the two subspecies, this is an indication that resistance to drug treatments may be able to arise in HAT caused by T.b. gambiense. Resistance to treatments is an ongoing threat for parasitic

6

diseases and in the case of HAT, eradication of the disease is unlikely given the reservoir CHAPTER 1 of trypanosomes found in livestock and wild life.18 This necessitates the discovery of novel drugs to prepare for the eventuality of resistance to current treatments. Given the areas to which HAT is endemic, a further complication in treating HAT is that each of these treatments requires either intravenous or intramuscular administration. Therefore, a key feature sought in novel treatments is that they be orally available.

Figure 2: The chemical structures of compounds that form part of the WHO recommended essential medicines for the treatment of HAT, suramin, pentamidine, merlasoprol, and the two active components of NECT treatment, nifurtimox and eflornithine. Two compounds currently in clinical trials for the treatment of HAT, SCYX-7158 and fexinidazole.

7

There are several promising drug candidates undergoing clinical trials for the treatment of HAT. A promising compound is SCYX-7158, a benzoxaborole currently in phase 1 clinical trials and showing the potency and pharmacokinetics consistent with a single dose oral treatment for stage 2 HAT (Figure 2).19 This compound was discovered and developed using phenotypic screening and the mode of action remains unknown. To find out if the mode of action might be related to phosphodiesterase (PDE) inhibition, SCYX-7158 was synthesized in our lab and found to inhibit human PDE4B with an IC50 of 80 µM and TbrPDEB1 with an IC50 of >100 µM, ruling out PDE inhibition as a mode of action (unpublished). Another compound in development for the treatment of HAT is fexinidazole which is currently undergoing phase II/III clinical trials.20 This too would be an orally available drug effective against both strains of trypanosomes causing HAT. While such advances are needed, even with a new drug reaching the market the need for novel treatments would remain, since resistance can be expected to arise. A further benefit to the discovery of multiple novel drugs is that the process of resistance can be slowed significantly by the use of multiple drugs with orthogonal modes of action. For these reasons, developing new drugs with novel targets for the treatment of HAT remains a high priority for institutes tackling neglected diseases.21 The drug target that was explored in the work presented here is phosphodiesterases, specifically TbrPDEB1.

1.2 Phosphodiesterases

Phosphodiesterases (PDEs) play a key role in signaling cascades that involve the second messenger molecules cyclic (cAMP) and cyclic guanosine monophosphate (cGMP). Through the hydrolysis of cAMP and cGMP to AMP and GMP respectively, PDEs regulate the concentrations of the cyclic nucleotides and thereby signal transmission. Cyclic nucleotide signaling in parasites shows potential as a drug target due to the important role PDEs play in the life cycles of many parasites.22-24 The Trypanosoma brucei cyclic nucleotide signaling pathway begins with the activation of membrane bound adenylyl cyclases (ACs).25 The ACs are bound to the cellular surface with specific distribution densities determined by localization (for example in flagellar ACs) and life cycle stage.26 The ACs are activated by an extracellular receptor domain that passes a signal

8

24, 27 via a transmembrane domain to an intracellular catalytic domain. The activated catalytic CHAPTER 1 domain converts ATP to cAMP passing the signal into cell and amplifying it though the creation of many second messenger molecules. The ACs act as sensors, but also play a role in inhibiting the innate immune response of the host.28 Downstream effectors of cAMP signaling are still poorly understood, however cAMP dependent protein kinase As (PKAs) have been identified.29 The concentration of cAMP determines signaling to downstream effectors and signaling is regulated by the activity of PDEs that hydrolyze the cAMP into AMP, reducing the cAMP concentration. The cyclic nucleotide signaling pathway has been found to play important roles in motility and cytokinesis of Trypanosoma brucei.30, 31.

There are 5 trypanosomal PDE genes, A, B1, B2, C and D and the same classification of genes is seen in leishmanial parasites. Plasmodial parasites also encode 5 PDE genes, however they are classed, αA, αB, β, γ and δ. Not all of the parasite PDEs are drug targets. In Trypanosoma brucei the inhibition of cytokinesis is observed when the activity of both TbrPDEB1 and TbrPDEB2 has been inhibited through either RNAi or pharmacological means.32-34 Together these form a target for the control of parasite proliferation, however due to their structural similarity and the correlation in inhibitor activity between them, drug discovery efforts have focused on TbrPDEB1.32, 33, 35, 36 Since TbrPDEA in not essential and TbrPDEC is inactive, they are not considered to be drug targets, whereas it is not yet known whether TbrPDED is essential.24, 37

In humans the signaling cascade usually begins when an extracellular stimulus is detected by a transmembrane receptor that upon activation stimulates the activity of adenylyl cyclase (AC) or guanylyl cyclase (GC). While active, AC continuously catalyzes the conversion of ATP into cAMP and GC continuously catalyzes the conversion of GTP into cGMP. This results in high concentrations of cAMP or cGMP and a strong amplification of the signal. The cyclic nucleotides then pass the signal on, usually to protein kinase A (PKA) for cAMP or protein kinase G (PKG) for cGMP. In turn these protein kinases activate proteins further down the signaling cascade. The specific roles of cAMP and cGMP are dependent on, amongst others, the signaling cascade, the cell type and the location of the cyclic nucleotides in the cell. The concentration of the cyclic nucleotides is reduced by PDEs and there are 21 genes that encode human PDEs. These are divided into 11 gene families that encode

9 structurally related PDE subtypes. This diversity allows PDEs to regulate a diverse range of signaling outcomes.38-40

Each PDE contains a catalytic domain, a lengthy N-terminus that may contain one or more structured domains and an unstructured C-terminus (Figure 3).41 Multiple isoforms (resulting from splice variants) of the PDE genes result in differences in the PDE sequence, or truncation of the PDE sequence when compared to the canonical sequence. The truncations typically occur at the carboxy-terminal unstructured region or at the amino- terminal region where one or more structured domain may be absent. Differences in the sequence of the catalytic domain are not seen and isoforms without a catalytic domain will be inactive. However PDEs resulting from alternative spicing may show altered catalytic activity due to the influence of the amino-terminal and carboxy-terminal regions on the substrate access to the catalytic site. Sequence differences in, or truncation of, amino- terminal domains may also affect the localization and activation mechanism of a PDE and generating inhibitors selective for specific isoforms is of interest for drug discovery.42

The structured regions found in the N-termini of PDEs have diverse roles and the role of the same domain family may vary between PDE families. The GAF domains have been shown to modulate PDE activity, play a role in PDE dimerization and are able to bind the cyclic nucleotides cAMP and cGMP.43 Differences in the GAF-A and GAF-B domains of PDE2, PDE5, PDE6, PDE10, PDE11 and parasite PDEB families influence the specific substrate binding and dimerization roles of the GAF domains in each case.44, 45 The UCR domains regulate PDE4 activity and crystallography has shown direct contact between UCR and inhibitors bound to the active site of PDE4D.46 The activation of PDE1 by Ca2+- calmodulin through interaction with the CaM binding domain regulates PDE1 activity.47 The PDE amino-terminal domains influence PDE localization, oligomerization and activity, alternative splicing of the N-terminus further diversifies the influence of PDEs on signaling pathways.48

10

CHAPTER 1

Figure 3: The structures and domains of the 11 human PDE families are shown along with the 3 parasite PDEs. The amino-terminal domains may regulate the activity of the conserved catalytic domain, play a role in the localization of PDEs or the interaction with protein partners. The amino-terminal domains include the CaM-binding domain (CaM), GAF domains, transmembrane domain (TM domain), targeting domain (TD), upstream conserved regions (UCRs), signal regulatory domain (REC), PAS domain, Pat7 nuclear localization, FYVE-type domains and coiled coil regions. The names of PDE subtypes are given in blue in cases where crystal structures of the PDE have been published and the specific domains that have been crystalized are shown in blue.

The catalytic activity of PDEs in the hydrolysis of cAMP and cGMP is dependent on the PDE family and to a lesser extent the PDE subtype. An overview of the selectivity of PDEs for the hydrolysis of cAMP, cGMP or both cAMP and cGMP is provided in Figure 4. The parasite PDEs of greatest interest as drug targets are shown along with the human PDEs. Of the trypanosomal PDEs, TbrPDEA and both TbrPDEB1 and TbrPDEB2 hydrolyse cAMP, while TbrPDEC is enzymatically inactive and the enzymatic activity of TbrPDED is

11 unknown.24, 37 Although rates of catalysis may vary across PDE subtypes, the substrate selectivity remains consistent within PDE families.

Figure 4: The substrate selectivity of PDEs, PDEs listed on the left selectively hydrolyze cGMP to GMP, those in the middle display a dual selectivity and hydrolyze both cGMP to GMP and cAMP to AMP and those on the right selectively hydrolyze cAMP to AMP.

The catalytic domains of phosphodiesterases contain a conserved substrate binding site where the hydrolysis of cyclic nucleotides is catalyzed. The binding of the cyclic nucleotide substrate is stabilized by the combination of a narrow hydrophobic region, dubbed the hydrophobic clamp, and multiple hydrogen bonds to a conserved glutamine residue. The phosphate group of the substrate is positioned in close proximity to the metal binding region of the pocket. Two metal ions are bound to phosphodiesterases, one of which is a Zn2+ ion and the second of which is most often Mg2+, but may be Mn2+ or another metal ion.49, 50 The

12

metal ions coordinate the phosphate group of the substrate and an activated water molecule CHAPTER 1 attacks the phosphate breaking a phosphate ester bond. The binding and hydrolysis of cAMP and AMP in PDE4 is shown in Figure 5, which reveals a conserved binding mode for the substrate and product.51 The hydrolysis of cGMP to GMP by other PDEs follows a similar mechanism.

Figure 5: A-B) The binding mode of cAMP (cyan, PDBcode: 2PW3) and AMP (magenta, PDBcode: 1ROR) bound to PDE4 (gray, PDBcode: 1ROR) highlighting the key residues in substrate stabilization and metal binding. The key residues, which interact with the substrate, are a conserved glutamine residue (red), two hydrophobic residues which form a clamp (yellow), a phenylalanine and an isoleucine (which may be a leucine or valine in other PDEs), and a phenylalanine (which may be a tyrosine in other PDEs). The metal ions zinc and magnesium are shown along with a key residue in metal binding, a conserved aspartate (blue). C) The hydrolysis of cAMP to AMP with the bond broken indicated in red and the added water in blue.

Through the analysis of PDE crystal structures published in the Protein Databank (PDB), a set of 57 residues were identified as the PDE pocket residues. A novel nomenclature for PDE pocket residues was devised, which combines the amino acid code, isoform residue number, pocket region and pocket residue number (Figure 6A). The residues were divided

13 into 10 pocket regions, Q, Q1, Q2, HC, HC1, HC2, S, MB, MB1 and MB2 (Figure 6B). The Q region refers to the conserved glutamine, QQ.50, which is flanked by the Q1 and Q2 regions. The HC region refers to the hydrophobic clamp made up of a conserved phenylalanine at FHC.52 and a hydrophobic residue at (I/L/V)HC.32 which may be isoleucine, leucine, or valine. The HC region is flanked by the HC1, HC2 and S regions, S region refers to a region exposed to bulk water that may contain multiple water filled subpockets. The MB region refers to the metal binding region in which two metal ions are coordinated by several conserved residues. The Zn2+ ion is coordinated to HMB.03, HMB.04, DMB.05, DMB.22 and two water molecules and the second metal ion is coordinated with DMB.05 and 5 water molecules.

An estimate of importance of the PDE pocket residues to ligand binding can be derived from the percentage of ligands, which show interactions with the residues in crystal structures. Using interaction fingerprints (IFPs) to identify interactions, these percentages are shown as color coding in Figure 6C (a complete analysis of this data is provided in Chapter 2). From this figure it is clear that of the 57 PDE pocket residues, 16 pocket residues do not form interactions with ligands (17 including Q2.44) and a further 12 form interactions with 3 ligands or less. A significant number of interactions are formed between ligands and 28 residues, and of these 7 residues form interactions with over 75% of ligands and can be described as key ligand binding residues, (I/L/V)MB1.17, (I/L/M)HC1.23, (I/L/V)HC.32, (F/Y)S.35, (F/G/L/M/V)S.40, QQ.50 and FHC.52. The generic pharmacophore for PDE inhibitors is determined by these interactions, chiefly the interactions with the hydrophobic clamp residues, (I/V/L)HC.32 and FHC.52, and (F/Y)S.35 that together form a hydrophobic cleft invariably occupied by aromatic or fully conjugated ring systems, and the adjacent QQ.50 with which over 90% of PDE inhibitors form hydrogen bonds.

Another way to assess the role of particular residues is to perform mutagenesis studies. As an enzyme there are multiple outcomes that can be tested, the catalytic activity of the mutated PDE, inhibitor binding to the mutated PDE, or the inhibitory activity of PDE inhibitors against mutated PDEs. An overview of the mutagenesis studies and outcomes for PDEs is provided in Figure 6D. From this overview it is clear that the identities of key residues involved in ligand binding are important to PDE function and inhibition, as might

14

be expected. There is a correlation for these residues with the residues found to form most CHAPTER 1 interactions with ligands (Figure 6C), however in the metal binding region mutating residues has a significant impact on PDE activity, while these residues do not play a significant role in ligand binding.

15

Figure 6: A) A nomenclature is presented that combines the standard amino acid reference containing the single letter amino acid code (red) and isoform specific residue number (purple) with the PDE pocket residue region name (blue) and the PDE pocket residue

16

number (green). When referencing PDE pocket residues of a subtype the isoform number CHAPTER 1 may be omitted (YHC1.01) and when referencing PDE pocket residues across the families the amino acid code and isoform number may be omitted (HC1.01). B) The 57 PDE binding site residues, defined, colored, and labeled according to the novel PDE pocket regions and nomenclature presented in Chapter 2. The names of the residues are made up of the amino acid code, regions of the PDE pocket (Q, Q1, Q2, HC, HC1, HC2, MB, MB1, MB2 and S) and the relative position of the pocket residue in the PDE sequence (the isoform residue numbers are not shown). C) The pocket residues are colored according to percentage of PDE crystal structures in which the residues are found to form interactions with ligands, using IFP analysis to determine interactions (Chapter 2). D) An overview of the mutational studies performed on PDEs. The size of the spheres denotes the number of mutational studies performed on a particular residue. The color denotes the average change in the reported parameter of PDE function or inhibition.

1.3 Drugs targeting PDEs

PDE activity plays an important role in a wide variety of signaling pathways, making PDEs attractive therapeutic targets. The value of modulating PDE activity is evident when considering the 23 PDE inhibitors that are approved for use as drugs (Figure 7). This is an impressive number of drugs for a single proteins class, although it is noted that some of these compounds bind not only PDEs, but also other targets (consider for example the multipharmacological profile of felodipine, levosimendan, , , and the ). Several drugs which act through PDE inhibition were not known to be PDE inhibitors during their development (,52 ,53 ,54 [1848, Georg Merck],55 and enoximone56). is a derivative of papaverine developed as No-Spa in Hungary in 1961 by Chinoin. However the rational of its design is not reported.

17

Figure 7: The drug molecules approved for therapeutic use which are known to inhibit PDE activity.

The first drug developed with PDE inhibition known to be a probable mode of action was the PDE3 inhibitor , a derivative of amrinone (Figure 8A). The identification of milrinone as a clinical candidate followed phenotypic screening using inotropic activity to identify hits.57 The discovery of followed from the earlier discovery of the selective PDE4 inhibitor (Figure 8B).58 An extensive structure-activity relationship study around rolipram was published prior to the discovery of roflumilast, that included a PDE4 inhibitor with a 1500 fold improvement in potency over rolipram.59 Roflumilast became the first drug designed as a PDE4 inhibitor to reach the market in 2011, following the failures of rolipram, and piclamilast in clinical trials due to lack of efficacy or failure to show sufficient safety.60

The difficulty in developing PDE4 inhibitors with acceptable safety profiles was made worse by side effects associated with PDE4 inhibition itself. These include nausea, emesis, abdominal pain and diarrhea.61-63 These have been factors in the failure of rolipram, cilomilast and piclamilast and nausea and diarrhea appear on the label of Daxas© (roflumilast).60 The severity and frequency of side-effects are dependent on several factors; the PDE4 subtype (with PDE4D associated with emesis), the specific PDE4 isoforms or

18

localized populations of PDE (e.g. in the brain) inhibited and the distribution of the inhibitor CHAPTER 1 in the body.64 Additionally PDE4 inhibitors may act as dual inhibitors of PDE3 raising the potential of cardiac safety concerns.65 The importance of selectivity in the development of PDE inhibitors is highlighted by the lengthy path to develop the first drug targeted at PDE4 inhibition.

N N H2N

O N O N H H Amrinone Milrinone PDE3 IC50 15 M PDE3 IC50 1.5 M A

N N Cl Cl O HN Cl Cl HN O HN O

O O O F O O O F Rolipram Piclamilast Roflumilast PDE4 IC50 1.5 M PDE4 IC50 1nM PDE4 IC50 0.8 nM B

N O N H NH N O N O N O N N NO HO N NH N NH NH N NO NO O O HO NH2 P N N S O O O O cGMP pyrazolopyrimidone PDE5 IC50 2.0 M PDE5 IC50 0.3 M PDE5 IC50 3.5 nM C non-selective

O O HN O HN HN NN N N N N O O O HN O NO O OO -CCE GR30040X Hydantoin Lead 2a PDE5 IC 8nM PDE5 IC 5nM D PDE5 IC50 0.8 M PDE5 IC50 0.3 M 50 50 Figure 8: The discovery of the drugs, milrinone (A), roflumilast (B), sildenafil (C), and tadalafil (D), that target PDEs.

Sildenafil stands out as the first rationally designed PDE inhibitor to reach the market and a remarkably successful drug under the name Viagra (Figure 8C). The program was initiated by Pfizer to discover a treatment for hypertension and other cardiovascular indications in

19

1985, when little was known about the role of cGMP or the target PDE5.66 The group studied the electronic distributions of cGMP and zaprinast (known to bind PDE5 weakly) to determine the dipole moments and the conformations of the ligands when crystalized. This enabled the suggestion of alternative hetrocycles and decorations to the scaffold that maintained the ligand’s ability to form core interactions. This resulted in an intermediate pyrazolopyrimidone with a 10 fold higher inhibition of PDE5 than zaprinast. From here the focus was on mimicking the phosphate group of cGMP to further stabilize binding, achieved with the addition of a sulfonamide group. Despite the vector of the sildenafil sulfonamide differing greatly from that of the cGMP phosphate in later crystal structures, sildenafil proved 1000 fold more potent than zaprinast and much more selective for PDE5. The clinical trials gave surprising results that led to sildenafil being labeled as a treatment for erectile dysfunction (ED) instead of as a treatment for cardiovascular disease, though the label was later extended to include pulmonary arterial hypertension (PAH).

The success of sildenafil on the market, lead to several me-too drugs, , , tadalafil and most recently . The discovery routes of vardenafil, udenafil (developed in South Korea) and avanafil (developed in Japan) are not published. The discovery of tadalafil progressed from ethyl β-carboline-3-carboxylate (β-CCE), which was found to inhibit PDE5 (Figure 8D).67 In a program of medicinal chemistry exploring the chemical space around β-CCE the intermediate GR30040X was discovered to improve the inhibition of PDE5. This was optimized further to the hydantoin lead 2a resulting in a 50 fold improvement in PDE5 inhibition and again to form tadalafil with a minor improvement in PDE5 inhibition.68 The exact progression has been questioned in court by Vanderbilt University. In their claim, which was rejected by the court, they state that a compound (8- (4-hydroxy phenylthio)-IBMX) in a research proposal they provided to the Glaxo France laboratory was used to progress the discovery of tadalafil.69

20

CHAPTER 1 10000 1000 inhibition 100 10 PDE5

1 Fold sildenafil vardenafil udenafil tadalafil avanafil PDE1 PDE6 PDE11

Figure 9: The fold inhibitions of PDE1, PDE6 and PDE11 relative to the inhibition of PDE5 for drugs that target PDE5.70, 71

All drugs targeting PDEs developed so far were conceived before the first PDE crystal structures were published. This precludes the use of structure-based drug design methods to support the development of compounds selective for one particular PDE family. As has been discussed in relation to PDE4, selectivity is a key aspect of drugs targeting PDEs as the unintentional inhibition of off target PDEs can lead to significant side effects. This is perhaps most visible in the various drugs targeting PDE5 which show some correlation between PDE selectivity and side effects (Figure 9). Inhibition of PDE1 is associated with cardiac events and flushing, side-effects most common to sildenafil, though these effects may also result from PDE5 inhibition.72 Inhibition of PDE6 (found in the retina) can lead to visual disturbances, a prominent side effect of sildenafil and udenafil. Inhibition of PDE11 is associated with myalgia and back pain, side effects most prominent for tadalafil.72 Developing PDE inhibitors into drugs today means being aware of the selectivity profiles of compounds across a panel of PDEs in order to identify the desired selectivity and structural biology plays a key role in that process.

1.4 PDE Crystal structure analysis

The 168 PDE crystal structures available in the PDB are a valuable resource for the discovery of novel PDE inhibitors. An extensive analysis of the PDE crystal structures is

21 presented in Chapter 2, in which IFPs were used to assess the interactions formed between ligands, metals, water molecules and the PDE binding pocket in each of the PDE crystal structures. Such analyses provide medicinal chemists with an overview of complex structural data in a practical format allowing them to quickly compare multiple PDE crystal structures.

The effective visualization of the information from a PDE-ligand complex can help the interpretation and comparison of multiple PDE crystal structures. Retaining information about the complex when viewing just the ligand is a way to compare many crystal structures in an efficient manner, as shown in Figure 10. Here the colors of the pocket regions close to atoms are projected onto the atoms in a figure of the ligand structure. In this way a single figure of the ligand can be used to convey information about the binding mode of the ligand in the binding site and the degree to which each atom of the ligand interacts with regions of the pocket. The color coded ligand figures are generated by applying a distance dependent translucence in relation to each of the regions for each atom of the ligand and then overlaying and averaging the resulting figures.

Figure 10: Visualizing occupation of pocket regions in single ligand figures using the color coding of the PDE pocket. A) The pocket color coding used to color ligand atoms. B) The two enantiomers of rolipram bound to PDE4B (PDB code 1XN0). C) The conformation of piclamilast as bound to PDE4B (PDB code 1XM4). D) The conformation of roflumilast as bound to PDE4B (PDB code 1XMU).

The comparison of the shapes and pharmacophore points of PDE binding pockets can also provide information on the similarity between PDEs. A new method was devised for this task. The Pymol plugin, Castp was used to generate binding sites from each published PDE crystal structure using a standardized algorithm (Figure 11A).73 The binding sites were then overlaid using the software application ROCS to score the similarity between PDE binding

22

74 pockets. ROCS was developed to perform 3D pharmacophore searches on small CHAPTER 1 molecules, applying it to binding sites was a novel use. The package uses a combination of shape overlap and pharmacophore overlap to optimize and rank the overlap of structures and generates a set of coordinates for the best fit. Averaging the scores across the PDE subtypes provided the input for the table shown in Figure 11B. This can be used to identify the most likely off targets when targeting a particular PDE subtype, or to identify similar PDEs when selecting compounds to test on a novel PDE subtype for which few inhibitors are known. The same process can be applied across protein classes to identify potential off- target binding sites and similar binding sites for which ligands may prove valuable starting points and against which drugs may potentially be repurposed. Looking for example at the TbrPDEB1 and lmjPDEB1 parasite PDEs, the most similar binding pockets are indicated to be PDE4B and PDE4D, a finding supported by sequence similarity, analysis of the binding sites and the activities of TbrPDEB1 inhibitors across the complete set of human PDEs.35, 75

23

Figure 11: A) The process followed to compare the binding sites of PDE crystal structures. In the first step the binding site is extracted by Castp and in the second the binding sites are compared by ROCs. B)A ROCS analysis of the binding sites extracted from PDE crystal structures showing the highest sum of shape and color (pharmacophore similarity) between crystal structures of any two PDE subtypes. The cells are colored as a heat map with green indicating high similarity and red indicating lack of similarity.

1.5 PDEs as targets for the treatment of neglected tropical diseases

The catalytic domains of both human and parasite PDEs share a high degree of structural similarity (Figure 12A). This similarity enriches the chances of identifying novel parasite PDE inhibitors when screening collections of human PDE inhibitors and has led to the discovery of a number of early hit molecules in parasite PDE drug discovery efforts.32, 36, 76, 77 However, compounds are needed that are selective for parasite PDEs, so once an inhibitor of a parasite PDE is found it must be optimized to achieve selectivity against the human PDE with the greatest similarity. In the case of TbrPDEB1 and LmjPDEB1 it is the human PDE4 that shows the greatest similarity structurally and pharmacologically. Achieving selectivity requires ligands bound to parasite PDEs to form interactions with residues that differ from the residues at the same position the human PDE or that occupy space that is not accessible in the human PDE. Subtle factors such as protein flexibility and water network disruption also play a role, but are hard to target in ligand design. In the case of TbrPDEB1 and human PDE4, the key interactions between PDE4 inhibitors bound to PDE4 and TbrPDEB1 are conserved. In order to develop TbrPDEB1 selective compounds, they need to interact with an unconserved region of the pocket. The parasite specific P- pocket is a region that differs between parasite and human PDEs and has been the focus for developing selective parasite PDE inhibitors (Figure 12B-E).24, 35, 77 The P-pocket is very similar in TbrPDEB1 and LmjPDEB1, in TcrPDEC the position and size differ and in PDE4, like other human PDEs the P-pocket is not present.

24

CHAPTER 1

25

Figure 12: A) A superposition of crystal structures of TbrPDEB1 (blue, 4I1535), LmjPDEB1 (green, 2R8Q77) with IBMX bound (blue), TcrPDEC (purple, 3V9437) with WYQ bound (pink) and PDE4B (yellow, 1XM478) with piclamilast bound (green). Close ups of the binding pocket are shown in B (TbrPDB1), C (LmjPDEB1), D (TcrPDEC) and E (PDE4B) with the P-pocket indicated in parasite PDEs.

HAT is being targeted through inhibition of TbrPDEB1 and TbrPDEB2 simultaneously. This has been shown to result in a halt of cell proliferation, lysis of the cells and in vivo clearance of the parasites from infected mice in a series of validation studies.32, 34, 75 The sequence identity when TbrPDEB1 and TbrPDEB2 are aligned is 75% and no residues on the surface of the binding site differ (Figure 13) coupled with the correlation in the inhibition of these enzymes, this allows efforts to be focused on just TbrPDEB1. 32, 33, 35, 36

Figure 13: A superposition a TbrPDEB1 crystal structure (blue, 4I15) and a homology model of TbrPDEB2 (red) showing the residues that differ as sticks.

26

39

The exact roles of cAMP in the Trypanosoma brucei life cycle is still uncertain. It plays a CHAPTER 1 role in differentiation of trypanosomes through the activation of protein kinase A (PKA).79 The concentration of cAMP also impacts cytokinesis during cell division. When TbrPDEB1 and TbrPDEB2 are inhibited or inactivated through RNAi, the internal cAMP concentration increases dramatically.34 Cells appear phenotypically affected as cytokinesis begins and multiple nuclei and kinetoplasts appear, a process that eventually leads to cell lysis.32-34 Targeting PDEs may also be applicable to treat other neglected tropical diseases.

Leishmaniasis is a disease found across the intertropical and temperate regions of the world where it is caused by infection by Leishmania species (Leishmania major, Leishmania infantum, Leishmania donovani, Leishmania mexicana or Leishmania braziliensis) that are spread by the phlebotomine sandfly. As in the case of trypanosomes, it is the PDEB family which is has been most studied as a drug target. However in the case of Leishmanial PDEB1 and PDEB2 the targets have not been validated and it remains uncertain whether they are essential.80 A particular complication in assessing the sensitivity of Leishmania species to PDE inhibition is that the parasites can be present in either promastigote or amastigote forms and the amastigotes are found within human cells. Assays have relied on visual inspection of human cells for the content of amastigotes, or high throughput yet poorly predictive promastigote screening, though the first biochemical assays have recently been developed within the T4-302 project.81, 82

Chagas disease is found in South America where it is spread by the faeces of triatomine bugs infected with Trypanosoma cruzi. In the case of Trypanosoma cruzi inhibition of PDEC has been validated as a means of controlling the parasite.23 The crystal structures of Trypanosoma cruzi PDEC show a binding site ,which diverges significantly from those of the PDEB families in Trypanosoma brucei or Leishmania species.

Malaria is no longer considered a neglected tropical disease and significant and ongoing efforts are being made to control the spread of malaria.83 Nevertheless, continued drug discovery efforts are prudent given the rate at which resistance to antimalarial drugs has grown in the past.84 Malaria is caused by infection with one of the Plasmodium species (Plasmodium falciparum, Plasmodium vivax, Plasmodium knowlesi, amongst others)

27 following a bite by an infected female Anopheles mosquito. Although Plasmodium PDEs are yet to be validated as drug targets for the control of malaria, there are indications that PDE inhibitors could function as antiplasmodials.22, 85

Figure 14: A phylogenetic tree showing the evolutionary relationships between known human and parasite PDEs and a series of putative parasite PDEs.

Looking beyond these four diseases there are other neglected tropical diseases caused by parasites where research on PDEs is still in its infancy. In order to identify PDE sequences in these parasites a series of BLAST searches were performed starting from each human PDE canonical sequence and each verified parasite PDE. The results were filtered to identify sequences derived from the following genera, Trypanosoma, Leishmania, Plasmodium, Schistosoma, Brugia, Wuchereria, Loa, Onchocerca and Wolbachia. The

28

remaining sequences were further filtered on the presence of residues known to play key CHAPTER 1 roles in the function of PDEs. The remaining sequences were realigned and the phylogenic tree shown in Figure 14 was generated. Two genera were not present in the results, Onchocerca and Wolbachia. The putative PDEs of Schistosoma, Brugia, Wuchereria and Loa were found to show greater similarity to human PDEs than to those of Trypanosoma, Leishmania and Plasmodium. This suggests that screening human PDE inhibitors against Schistosoma, Brugia, Wuchereria and Loa could prove to be an excellent starting point to validate these PDEs as potential drug targets.

1.6 Research Aim

The aim of the research presented in this thesis was to discover novel selective inhibitors of TbrPDEB1 with the potential for development into drugs to treat HAT. In support of that aim, improving the structural understanding of PDEs was set as a priority in order to allow rational drug design and support future PDE drug discovery efforts.

1.7 PDE Drug Discovery

During the research presented in this thesis a diverse range of drug-discovery methods were deployed with the aim of discovering parasite PDE inhibitors. Several key drug-discovery resources were available at the start of the project: the publicly available crystal structures of many human PDEs and the parasite PDE LmjPDEB1, a proprietary lead compound series that potently inhibits parasite PDEB subtypes, a screening assay for parasite and human PDE inhibition and a phenotypic parasite screening assay. Over the course of the project crystal structures of TbrPDEB1 became available opening the door to further drug- discovery methods.

In Chapter 3 TbrPDEB1 inhibitors were found using fragment-based drug design (FBDD). In FBDD, molecules that satisfy the definition of a fragment, e.g. molecules with 20 or fewer heavy atoms, are identified as starting points for further optimization. These initial fragments can be identified by fragment screening using computational screening,

29 pharmacological screening, or, as in this case, from literature. Fragments tend to show weak activity. However, when corrected by the number of atoms, the ligand efficiency (LE) of fragments can be high. Maintaining a high LE while growing fragments is a key challenge in FBDD. Structural knowledge of the ligand binding site and the binding mode of the fragment provide the means to rationally grow the fragment while trying to keep the LE high. In the case of targeting TbrPDEB1, a homology model was created of TbrPDB1 using the LmjPDEB1 crystal structure as a reference. This homology model was used to predict the binding modes of the fragments and intermediates during the fragment growing process.

In Chapter 4 virtual screening was used to identify novel TbrPDEB1 inhibitors using the first TbrPDEB1 crystal structure to be published. Virtual screening is the process of identifying active ligands from a database of molecular structures. The identification can proceed in a ligand-based fashion, through the comparison of molecules in the database to a reference molecule, or a pharmacophore model. A pharmacophore model abstracts several positions in space using the conformations of one or more reference molecules. A sphere at each isolated position describes one or more molecular features and rules regarding which spheres and features must be matched by conformations of ligands in the database to be considered hits. If a protein structure is available, structure based virtual screening may be used to identify hits from a database. In this case binding modes of the molecules in the database are predicted using docking software and one or more scoring functions are used to identify hits. In the virtual screening method presented in Chapter 4, a database of commercially available molecules was filtered and then docked into an unliganded TbrPDEB1 crystal structure. The interactions between each docked pose and the protein were calculated in the form of interaction fingerprints (IFPs) and a score was generated by comparing the IFPs to the IFP of a reference compound. Ranking combined the docking scores and IFP similarity scores with selected references, to identify hit compounds.

In Chapter 5 ligands were discovered using structure guided design with multiple liganded TbrPDEB1 crystal structures and scaffold merging. In structure guided design, crystal structures of intermediate compounds guide the discovery of target compounds, often in an iterative process. In scaffold merging, substructures from multiple ligands are brought together to form a new molecule. Once crystal structures showing the binding modes of

30

potent but non-selective TbrPDEB1 inhibitors described in Chapter 4 were available, it CHAPTER 1 became clear these were not targeting a parasite specific sub-pocket, dubbed the P-pocket. Through the analysis of all published PDE crystal structures, scaffolds were identified to merge with those of the potent non-selective TbrPDEB1 inhibitors, with the aim of rigidifying the ligands and forcing occupation of P-pocket. The resulting inhibitors proved to be both potent and selective inhibitors of TbrPDEB1.

1.8 Project T4-302

The research presented in this thesis has taken place within the context of a consortium financed mainly by Top Institute Pharma (TI Pharma). The aims of the consortium were, to validate the use of parasite-specific PDE inhibitors as therapeutic agents in Leishmania and Trypanosoma infections; develop suitable medicinal chemistry leads targeting PDEs as drug candidates in these and other neglected tropical diseases (NTDs); and to provide a new platform for TI Pharma to exploit rational mechanism-based approaches to NTD discovery. Resources to achieve these aims were provided by each of the consortium members.

31

Figure 15: An overview of the consortium members and their contributions to project T4- 302. The aim of this project was to identify phosphodiesterase inhibitors with the potential to be developed into drugs, like Viagra © shown at the center.

The consortium contained seven members, TI Phama, University of Bern, Nycomed, Mercachem, VU University Amsterdam, IOTA Pharmaceuticals, DNDi and the Royal Tropical Institute (Figure 15). TI Pharma coordinated the public-private partnership in which private contributions were matched by public contributions. The University of Bern is where phosphodiesterases (PDEs) were first found to be essential for parasite proliferation in Trypanosoma brucei and Leishmania major and they provided phenotypic screening against the parasites. Nycomed (now part of Takeda), provided access to a series of Trypanosoma brucei phosphodiesterase B1 (TbrPDEB1) inhibitors discovered through

32

high throughput screening, along with pharmacological screening against parasite and CHAPTER 1 human PDEs. Mercachem provided novel PDE inhibitor synthesis contributions and support from experienced medicinal chemists. VU University Amsterdam provided novel PDE inhibitor computational design and synthesis contributions and novel pharmacology technique research. IOTA provided access to a fragment library for phenotypic and enzymatic screening and performed biochemical and biophysical screening. DNDi provided phenotypic screening and basic toxicology screening of project compounds. The Royal Tropical Institute in Amsterdam developed novel screening methods to identify parasite infections and provided phenotypic screening against parasites.

1. Simarro, P.; Cecchi, G.; Paone, M.; Franco, J.; Diarra, A.; Ruiz, J.; Fevre, E.; Courtin, F.; Mattioli, R.; Jannin, J., The Atlas of human African trypanosomiasis: a contribution to global mapping of neglected tropical diseases. International Journal of Health Geographics 2010, 9 (1), 57. 2. Simarro, P. P.; Cecchi, G.; Franco, J. R.; Paone, M.; Diarra, A.; Ruiz‐Postigo, J. A.; Fevre, E. M.; Mattioli, R. C.; Jannin, J. G., Estimating and mapping the population at risk of sleeping sickness. PLoS Negl Trop Dis 2012, 6 (10), e1859. 3. Spinage, C., The Tsetse Fly I: Africa’s Bane and Benefice. In African Ecology, Springer Berlin Heidelberg: 2012; pp 821‐866. 4. Simarro, P. P.; Jannin, J.; Cattand, P., Eliminating Human African Trypanosomiasis: Where Do We Stand and What Comes Next? PLoS Med 2008, 5 (2), e55. 5. Weiss, B. L.; Wang, J.; Maltz, M. A.; Wu, Y.; Aksoy, S., Trypanosome Infection Establishment in the Tsetse Fly Gut Is Influenced by Microbiome‐ Regulated Host Immune Barriers. PLoS Pathog 2013, 9 (4), e1003318. 6. Steverding, D., The history of African trypanosomiasis. Parasites & Vectors 2008, 1 (1), 3. 7. Simarro, P. P.; Franco, J. R.; Diarra, A.; Postigo, J. A. R.; Jannin, J. G., Diversity of human African trypanosomiasis epidemiological settings requires fine‐ tuning control strategies to facilitate disease elimination. Research and Reports in Tropical Medicine 2013, 4, 1‐6. 8. Chappuis, F.; Lima, M. A.; Flevaud, L.; Ritmeijer, K., Human African trypanosomiasis in areas without surveillance. Emerg. Infect. Dis. 2010, 16 (2), 354‐356. 9. Wheeler, R., Silent but deadly. In Blender, Cell Picture Show ‐ Parasites, 2013. 10. Robert, P., Sleeping Sickness is Back. Corbis: 2000.

33

11. Wertheim, H. F. L.; Hornby, P.; Woodall, J. P., Atlas of Human Infectious Diseases. Wiley‐Blackwell: 2012. 12. Jamonneau, V.; Ilboudo, H.; Kaboré, J.; Kaba, D.; Koffi, M.; Solano, P.; Garcia, A.; Courtin, D.; Laveissière, C.; Lingue, K.; Büscher, P.; Bucheton, B., Untreated Human Infections by Trypanosoma brucei gambiense Are Not 100% Fatal. PLoS Negl Trop Dis 2012, 6 (6), e1691. 13. MacLean, L.; Reiber, H.; Kennedy, P. G. E.; Sternberg, J. M., Stage Progression and Neurological Symptoms in Trypanosoma brucei rhodesiense Sleeping Sickness: Role of the CNS Inflammatory Response. PLoS Negl Trop Dis 2012, 6 (10), e1857. 14. Steverding, D., The development of drugs for treatment of sleeping sickness: a historical review. Parasites & Vectors 2010, 3 (1), 15. 15. Kennedy, P. G. E., Clinical features, diagnosis, and treatment of human African trypanosomiasis (sleeping sickness). The Lancet Neurology 2013, 12 (2), 186‐194. 16. Alirol, E.; Schrumpf, D.; Amici Heradi, J.; Riedel, A.; de Patoul, C.; Quere, M.; Chappuis, F., Nifurtimox‐Eflornithine Combination Therapy for Second‐Stage Gambiense Human African Trypanosomiasis: Médecins Sans Frontières Experience in the Democratic Republic of the Congo. Clin. Infect. Dis. 2013, 56 (2), 195‐203. 17. Schmid, C.; Richer, M.; Bilenge, C. M. M.; Josenando, T.; Chappuis, F.; Manthelot, C. R.; Nangouma, A.; Doua, F.; Asumu, P. N.; Simarro, P. P.; Burri, C., Effectiveness of a 10‐Day Melarsoprol Schedule for the Treatment of Late‐Stage Human African Trypanosomiasis: Confirmation from a Multinational Study (Impamel II). J. Infect. Dis. 2005, 191 (11), 1922‐1931. 18. Van den Bossche, P.; Ky‐Zerbo, A.; Brandt, J.; Marcotty, T.; Geerts, S.; De Deken, R., Transmissibility of Trypanosoma brucei during its development in cattle. Trop. Med. Int. Health 2005, 10 (9), 833‐839. 19. Wring, S.; Gaukel, E.; Jacobs, R.; Chanda, S.; Gualano, V.; Evène, E.; Donazzolo, Y.; Latreille, M.; Don, R.; Mowbray, C.; Tarral, A. In SCYX‐7158 (AN5568): CNS Exposure Predicted from First‐in‐Human Clinical Studies Indicates a Single Oral Dose Treatment is Possible for Sterile Cures of Stage 2 Human African Trypanosomiasis, ASTMH 62nd Annual Meeting Washington D.C., Washington D.C., 2013. 20. Mäser, P.; Wittlin, S.; Rottmann, M.; Wenzler, T.; Kaiser, M.; Brun, R., Antiparasitic agents: new drugs on the horizon. Curr. Opin. Pharmacol. 2012, 12 (5), 562‐566. 21. R&D Portfolio. In Patient Needs‐Driven Collaborative R&D Model for Neglected Diseases, DNDi Drugs for Neglected Diseases initiative: 2014.

34

22. Howard, B.; Thompson, P.; Manallack, D., Active site similarity between CHAPTER 1 human and Plasmodium falciparum phosphodiesterases: considerations for antimalarial drug design. J. Comput. Aided Mol. Des. 2011, 25 (8), 753‐762. 23. King‐Keller, S.; Li, M.; Smith, A.; Zheng, S.; Kaur, G.; Yang, X.; Wang, B.; Docampo, R., Chemical Validation of Phosphodiesterase C as a Chemotherapeutic Target in Trypanosoma cruzi, the Etiological Agent of Chagas' Disease. Antimicrob. Agents Chemother. 2010, 54 (9), 3738‐3745. 24. Shakur, Y.; de Koning, H. P.; Ke, H.; Kambayashi, J.; Seebeck, T., Therapeutic potential of phosphodiesterase inhibitors in parasitic diseases. Handb Exp Pharmacol. 2011, (204), 487‐510. 25. Leung, K.; Manna, P.; Boehm, C.; Maishman, L.; Field, M., Cell Biology for Immune Evasion: Organizing Antigenic Variation, Surfaces, Trafficking, and Cellular Structures in Trypanosoma brucei. In Trypanosomes and Trypanosomiasis, Magez, S.; Radwanska, M., Eds. Springer Vienna: 2014; pp 1‐39. 26. Saada, E. A.; Kabututu, Z. P.; Lopez, M.; Shimogawa, M. M.; Langousis, G.; Oberholzer, M.; Riestra, A.; Jonsson, Z.; Wohlschlegel, J. A.; Hill, K. L., Insect stage‐ specific receptor adenylate cyclases are localized to distinct subdomains of the Trypanosoma brucei flagellar membrane. Eukaryot. Cell 2014. 27. Seebeck, T.; Schaub, R.; Johner, A., cAMP signalling in the kinetoplastid protozoa. Curr Mol Med. 2004, 4 (6), 585‐99. 28. Salmon, D.; Vanwalleghem, G.; Morias, Y.; Denoeud, J.; Krumbholz, C.; Lhommé, F.; Bachmaier, S.; Kador, M.; Gossmann, J.; Dias, F. B. S.; De Muylder, G.; Uzureau, P.; Magez, S.; Moser, M.; De Baetselier, P.; Van Den Abbeele, J.; Beschin, A.; Boshart, M.; Pays, E., Adenylate Cyclases of Trypanosoma brucei Inhibit the Innate Immune Response of the Host. Science 2012, 337 (6093), 463‐466. 29. Jones, N. G.; Thomas, E. B.; Brown, E.; Dickens, N. J.; Hammarton, T. C.; Mottram, J. C., Regulators of Trypanosoma brucei Cell Cycle Progression and Differentiation Identified Using a Kinome‐Wide RNAi Screen. PLoS Pathog 2014, 10 (1), e1003886. 30. Langousis, G.; Hill, K. L., Motility and more: the flagellum of Trypanosoma brucei. Nat Rev Micro 2014, 12 (7), 505‐518. 31. Salmon, D.; Bachmaier, S.; Krumbholz, C.; Kador, M.; Gossmann, J. A.; Uzureau, P.; Pays, E.; Boshart, M., Cytokinesis of Trypanosoma brucei bloodstream forms depends on expression of adenylyl cyclases of the ESAG4 or ESAG4‐like subfamily. Mol. Microbiol. 2012, 84 (2), 225‐242. 32. de Koning, H. P.; Gould, M. K.; Sterk, G. J.; Tenor, H.; Kunz, S.; Luginbuehl, E.; Seebeck, T., Pharmacological Validation of Trypanosoma brucei Phosphodiesterases as Novel Drug Targets. J. Infect. Dis. 2012, 206 (2), 229‐237. 33. Orrling, K. M.; Jansen, C.; Vu, X. L.; Balmer, V.; Bregy, P.; Shanmugham, A.; England, P.; Bailey, D.; Cos, P.; Maes, L.; Adams, E.; van den Bogaart, E.; Chatelain, E.; Ioset, J.‐R.; van de Stolpe, A.; Zorg, S.; Veerman, J.; Seebeck, T.; Sterk, G. J.; de

35

Esch, I. J. P.; Leurs, R., Catechol Pyrazolinones as Trypanocidals: Fragment‐Based Design, Synthesis, and Pharmacological Evaluation of Nanomolar Inhibitors of Trypanosomal Phosphodiesterase B1. J. Med. Chem. 2012, 55 (20), 8745‐8756. 34. Oberholzer, M.; Marti, G.; Baresic, M.; Kunz, S.; Hemphill, A.; Seebeck, T., The Trypanosoma brucei cAMP phosphodiesterases TbrPDEB1 and TbrPDEB2: flagellar enzymes that are essential for parasite virulence. FASEB J. 2007, 21 (3), 720‐731. 35. Jansen, C.; Wang, H.; Kooistra, A. J.; de Graaf, C.; Orrling, K. M.; Tenor, H.; Seebeck, T.; Bailey, D.; de Esch, I. J. P.; Ke, H.; Leurs, R., Discovery of Novel Trypanosoma brucei Phosphodiesterase B1 Inhibitors by Virtual Screening against the Unliganded TbrPDEB1 Crystal Structure. J. Med. Chem. 2013, 56 (5), 2087‐ 2096. 36. Bland, N. D.; Wang, C.; Tallman, C.; Gustafson, A. E.; Wang, Z.; Ashton, T. D.; Ochiana, S. O.; McAllister, G.; Cotter, K.; Fang, A. P.; Gechijian, L.; Garceau, N.; Gangurde, R.; Ortenberg, R.; Ondrechen, M. J.; Campbell, R. K.; Pollastri, M. P., Pharmacological Validation of Trypanosoma brucei Phosphodiesterases B1 and B2 as Druggable Targets for African Sleeping Sickness. J. Med. Chem. 2011, 54 (23), 8188‐8194. 37. Wang, H.; Kunz, S.; Chen, G.; Seebeck, T.; Wan, Y.; Robinson, H.; Martinelli, S.; Ke, H., Biological and Structural Characterization of Trypanosoma cruzi Phosphodiesterase C and Implications for Design of Parasite Selective Inhibitors. J. Biol. Chem. 2012, 287 (15), 11788‐11797. 38. McDonough, K. A.; Rodriguez, A., The myriad roles of cyclic AMP in microbial pathogens: from signal to sword. Nat Rev Micro 2012, 10 (1), 27‐38. 39. Gould, M. K.; de Koning, H. P., Cyclic‐nucleotide signalling in protozoa. FEMS Microbiol. Rev. 2011, 35 (3), 515‐541. 40. Keravis, T.; Lugnier, C., Cyclic nucleotide phosphodiesterase (PDE) isozymes as targets of the intracellular signalling network: benefits of PDE inhibitors in various diseases and perspectives for future therapeutic developments. Br. J. Pharmacol. 2012, 165 (5), 1288‐1305. 41. Francis, S.; Houslay, M.; Conti, M., Phosphodiesterase Inhibitors: Factors That Influence Potency, Selectivity, and Action. In Phosphodiesterases as Drug Targets, Francis, S. H.; Conti, M.; Houslay, M. D., Eds. Springer Berlin Heidelberg: 2011; Vol. 204, pp 47‐84. 42. Houslay, M. D.; Adams, D. R., Putting the lid on phosphodiesterase 4. Nat Biotech 2010, 28 (1), 38‐40. 43. Schultz, J., Structural and Biochemical Aspects of Tandem GAF Domains. In cGMP: Generators, Effectors and Therapeutic Implications, Schmidt, H. H. W.; Hofmann, F.; Stasch, J.‐P., Eds. Springer Berlin Heidelberg: 2009; Vol. 191, pp 93‐ 109.

36

44. Heikaus, C. C.; Pandit, J.; Klevit, R. E., Cyclic Nucleotide Binding GAF CHAPTER 1 Domains from Phosphodiesterases: Structural and Mechanistic Insights. Structure 2009, 17 (12), 1551‐1557. 45. Kunz, S.; Luginbuehl, E.; Seebeck, T., Gene Conversion Transfers the GAF‐ A Domain of Phosphodiesterase TbrPDEB1 to One Allele of TbrPDEB2 of Trypanosoma brucei. PLoS Negl Trop Dis 2009, 3 (6), e455. 46. Burgin, A. B.; Magnusson, O. T.; Singh, J.; Witte, P.; Staker, B. L.; Bjornsson, J. M.; Thorsteinsdottir, M.; Hrafnsdottir, S.; Hagen, T.; Kiselyov, A. S.; Stewart, L. J.; Gurney, M. E., Design of phosphodiesterase 4D (PDE4D) allosteric modulators for enhancing cognition with improved safety. Nat Biotech 2010, 28 (1), 63‐70. 47. Goraya, T. A.; Cooper, D. M. F., Ca2+‐calmodulin‐dependent phosphodiesterase (PDE1): Current perspectives. Cell. Signal. 2005, 17 (7), 789‐ 797. 48. Omori, K.; Kotera, J., Overview of PDEs and Their Regulation. Circ. Res. 2007, 100 (3), 309‐327. 49. Huai, Q.; Colicelli, J.; Ke, H., The Crystal Structure of AMP‐Bound PDE4 Suggests a Mechanism for Phosphodiesterase Catalysis†,‡. Biochemistry (Mosc). 2003, 42 (45), 13220‐13226. 50. Liu, S.; Mansour, M. N.; Dillman, K. S.; Perez, J. R.; Danley, D. E.; Aeed, P. A.; Simons, S. P.; LeMotte, P. K.; Menniti, F. S., Structural basis for the catalytic mechanism of human phosphodiesterase 9. Proc. Natl. Acad. Sci. U.S.A. 2008, 105 (36), 13309‐13314. 51. Xu, R. X.; Rocque, W. J.; Lambert, M. H.; Vanderwall, D. E.; Luther, M. A.; Nolte, R. T., Crystal Structures of the Catalytic Domain of Phosphodiesterase 4B Complexed with AMP, 8‐Br‐AMP, and Rolipram. J. Mol. Biol. 2004, 337 (2), 355‐ 365. 52. Benotti, J. R.; Grossman, W.; Braunwald, E.; Davolos, D. D.; Alousi, A. A., Hemodynamic Assessment of Amrinone. N. Engl. J. Med. 1978, 299 (25), 1373‐ 1377. 53. Fleming, J. S.; Buyniski, J. P., A potent new inhibitor of aggregation and experimental thrombosis, anagrelide (BL‐4162A). Thromb. Res. 1979, 15 (3– 4), 373‐388. 54. Shintani, S.; Watanabe, K.; Kawamura, K.; Mori, T.; Tani, T.; Toba, Y.; Sasabe, H.; Nakagiri, N.; Hongoh, O.; Fujita, S.; et al., General pharmacological properties of cilostazol, a new antithrombotic drug. Part II: Effect on the peripheral organs. Arzneimittelforschung. 1985, 35 (7A), 1163‐72. 55. Hopkins, H. S., Protoplasmic effects of papaverine, histamine and other drugs, in relation to the theory of smooth muscle contraction. American Journal of Physiology ‐‐ Legacy Content 1922, 61 (3), 551‐561.

37

56. Schnettler, R. A.; Dage, R. C.; Grisar, J. M., 4‐Aroyl‐1,3‐dihydro‐2H‐ imidazol‐2‐ones, a new class of cardiotonic agents. J. Med. Chem. 1982, 25 (12), 1477‐1481. 57. Pastelin, G.; Mendez, R.; Kabela, E.; Farah, A., The search for a digitalis substitute II milrinone (Win 47203) its action on the heart‐lung preparation of the dog. Life Sci. 1983, 33 (18), 1787‐1796. 58. Hatzelmann, A.; Schudt, C., Anti‐Inflammatory and Immunomodulatory Potential of the Novel PDE4 Inhibitor Roflumilast in Vitro. J. Pharmacol. Exp. Ther. 2001, 297 (1), 267‐279. 59. Ashton, M. J.; Cook, D. C.; Fenton, G.; Karlsson, J.‐A.; Palfreyman, M. N.; Raeburn, D.; Ratcliffe, A. J.; Souness, J. E.; Thurairatnam, S.; Vicker, N., Selective Type IV Phosphodiesterase Inhibitors as Antiasthmatic Agents. The Syntheses and Biological Activities of 3‐(Cyclopentyloxy)‐4‐methoxybenzamides and Analogs. J. Med. Chem. 1994, 37 (11), 1696‐1703. 60. Brown, W. M., Treating COPD with PDE 4 inhibitors. International Journal of Chronic Obstructive Pulmonary Disease 2007, 2 (4), 517 ‐ 533. 61. Heaslip, R. J.; Evans, D. Y., Emetic, central nervous system and pulmonary activities of rolipram in the dog. Eur. J. Pharmacol. 1995, 286 (3), 281‐290. 62. Giembycz, M. A., An update and appraisal of the cilomilast Phase III clinical development programme for chronic obstructive pulmonary disease. Br. J. Clin. Pharmacol. 2006, 62 (2), 138‐152. 63. Lipworth, B. J., Phosphodiesterase‐4 inhibitors for and chronic obstructive pulmonary disease. The Lancet 365 (9454), 167‐175. 64. Hirose, R.; Manabe, H.; Nonaka, H.; Yanagawa, K.; Akuta, K.; Sato, S.; Ohshima, E.; Ichimura, M., Correlation between emetic effect of phosphodiesterase 4 inhibitors and their occupation of the high‐affinity rolipram binding site in Suncus murinus brain. Eur. J. Pharmacol. 2007, 573 (1–3), 93‐99. 65. Movsesian, M.; Wever‐Pinzon, O.; Vandeput, F., PDE3 inhibition in dilated cardiomyopathy. Curr. Opin. Pharmacol. 2011, 11 (6), 707‐713. 66. Campbell, S. F., Science, art and drug discovery: a personal perspective. Clin. Sci. 2000, 99, 255‐260. 67. Daugan, A.; Grondin, P.; Ruault, C.; Le Monnier de Gouville, A.‐C.; Coste, H.; Kirilovsky, J.; Hyafil, F.; Labaudinière, R., The Discovery of Tadalafil: A Novel and Highly Selective PDE5 Inhibitor. 1: 5,6,11,11a‐Tetrahydro‐1H‐ imidazo[1‘,5‘:1,6]pyrido[3,4‐b]indole‐1,3(2H)‐dione Analogues. J. Med. Chem. 2003, 46 (21), 4525‐4532. 68. Daugan, A.; Grondin, P.; Ruault, C.; Le Monnier de Gouville, A.‐C.; Coste, H.; Linget, J. M.; Kirilovsky, J.; Hyafil, F.; Labaudinière, R., The Discovery of Tadalafil: A Novel and Highly Selective PDE5 Inhibitor. 2: 2,3,6,7,12,12a‐ hexahydropyrazino[1‘,2‘:1,6]pyrido[3,4‐b]indole‐1,4‐dione Analogues. J. Med. Chem. 2003, 46 (21), 4533‐4542.

38

69. Civil Action No: 05‐506‐SLR. 2008; pp 1‐29. CHAPTER 1 70. Kouvelas, D.; Goulas, A.; Papazisis, G.; Sardeli, C.; Pourzitaki, C., PDE5 inhibitors: in vitro and in vivo pharmacological profile. Curr Pharm Des. 2009, 15 (30), 3464‐75. 71. Wang, R.; Burnett, A. L.; Heller, W. H.; Omori, K.; Kotera, J.; Kikkawa, K.; Yee, S.; Day, W. W.; DiDonato, K.; Peterson, C. A., Selectivity of Avanafil, a PDE5 Inhibitor for the Treatment of Erectile Dysfunction: Implications for Clinical Safety and Improved Tolerability. The Journal of Sexual Medicine 2012, 9 (8), 2122‐2129. 72. Smith, W. B.; McCaslin, I. R.; Gokce, A.; Mandava, S. H.; Trost, L.; Hellstrom, W. J., PDE5 inhibitors: considerations for preference and long‐term adherence. Int. J. Clin. Pract. 2013, 67 (8), 768‐780. 73. Dundas, J.; Ouyang, Z.; Tseng, J.; Binkowski, A.; Turpaz, Y.; Liang, J., CASTp: computed atlas of surface topography of proteins with structural and topographical mapping of functionally annotated residues. Nucleic Acids Res. 2006, 34 (suppl 2), W116‐W118. 74. ROCS, version 3.0.0; Openeye Scientific Software: Santa Fe, NM. 75. Orrling, K. M.; Jansen, C.; Vu, X. L.; Balmer, V.; Bregy, P.; Shanmugham, A.; England, P.; Cos, P.; Maes, L.; Adams, E.; van de Bogaart, E.; Chatelain, E.; Ioset, J.‐ R.; Stolpe, A.; Zorg, S.; Veerman, J.; Seebeck, T.; Sterk, G. J.; de Esch, I. J. P.; Leurs, R., Catechol Pyrazolinones as Trypanocidals: Fragment‐Based Design, Synthesis and Pharmacological Evaluation of Nanomolar Inhibitors of Trypanosomal Phosphodiesterase B1. J. Med. Chem. 2012. 76. Kunz, S.; Kloeckner, T.; Essen, L.‐O.; Seebeck, T.; Boshart, M., TbPDE1, a novel class I phosphodiesterase of Trypanosoma brucei. Eur. J. Biochem. 2004, 271 (3), 637‐647. 77. Wang, H.; Yan, Z.; Geng, J.; Kunz, S.; Seebeck, T.; Ke, H., Crystal structure of the Leishmania major phosphodiesterase LmjPDEB1 and insight into the design of the parasite‐selective inhibitors. Mol. Microbiol. 2007, 66 (4), 1029‐1038. 78. Card, G. L.; England, B. P.; Suzuki, Y.; Fong, D.; Powell, B.; Lee, B.; Luu, C.; Tabrizizad, M.; Gillette, S.; Ibrahim, P. N.; Artis, D. R.; Bollag, G.; Milburn, M. V.; Kim, S.‐H.; Schlessinger, J.; Zhang, K. Y. J., Structural Basis for the Activity of Drugs that Inhibit Phosphodiesterases. Structure 2004, 12 (12), 2233‐2247. 79. Seebeck, T.; Gong, K.; Kunz, S.; Schaub, R.; Shalaby, T.; Zoraghi, R., cAMP signalling in Trypanosoma brucei. Int. J. Parasitol. 2001, 31 (5–6), 491‐498. 80. Johner, A.; Kunz, S.; Linder, M.; Shakur, Y.; Seebeck, T., Cyclic nucleotide specific phosphodiesterases of Leishmania major. BMC Microbiol. 2006, 6 (1), 1‐ 13. 81. van den Bogaart, E.; Schoone, G. J.; England, P.; Faber, D.; Orrling, K. M.; Dujardin, J.‐C.; Sundar, S.; Schallig, H. D. F. H.; Adams, E. R., Simple Colorimetric Trypanothione Reductase‐Based Assay for High‐Throughput Screening of Drugs

39 against Leishmania Intracellular Amastigotes. Antimicrob. Agents Chemother. 2014, 58 (1), 527‐535. 82. van den Bogaart, E.; Schoone, G. J.; Adams, E. R.; Schallig, H. D. F. H., Duplex quantitative Reverse‐Transcriptase PCR for simultaneous assessment of drug activity against Leishmania intracellular amastigotes and their host cells. International Journal for Parasitology: Drugs and Drug Resistance 2014, 4 (1), 14‐ 19. 83. Wells, T. N. C.; Alonso, P. L.; Gutteridge, W. E., New medicines to improve control and contribute to the eradication of malaria. Nat Rev Drug Discov 2009, 8 (11), 879‐891. 84. Wongsrichanalai, C.; Pickard, A. L.; Wernsdorfer, W. H.; Meshnick, S. R., Epidemiology of drug‐resistant malaria. The Lancet Infectious Diseases 2002, 2 (4), 209‐218. 85. Beghyn, T. B.; Charton, J.; Leroux, F.; Laconde, G.; Bourin, A.; Cos, P.; Maes, L.; Deprez, B., Drug to Genome to Drug: Discovery of New Antiplasmodial Compounds. J. Med. Chem. 2011, 54 (9), 3222‐3240.

40