• No results found

University of Groningen Discovery of Inhibitors by Combinatorial-Chemistry Approaches van der Vlag, Ramon

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Discovery of Inhibitors by Combinatorial-Chemistry Approaches van der Vlag, Ramon"

Copied!
27
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

Discovery of Inhibitors by Combinatorial-Chemistry Approaches

van der Vlag, Ramon

DOI:

10.33612/diss.146091529

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from

it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date:

2020

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

van der Vlag, R. (2020). Discovery of Inhibitors by Combinatorial-Chemistry Approaches. University of

Groningen. https://doi.org/10.33612/diss.146091529

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

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.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

Chapter 4

Optimized Inhibitors of MDM2 via an Attempted

Protein-Templated Reductive Amination

Innovative and efficient hit-identification techniques are required to accelerate drug

discovery. Protein-templated fragment ligations represent a promising strategy in

early drug discovery, enabling the target to assemble and select its binders from a

pool of building blocks. Development of new protein-templated reactions to access

a larger structural diversity and expansion of the variety of targets to demonstrate

the scope of the technique are of prime interest for medicinal chemists. Herein, we

present our attempts to use a protein-templated reductive amination to target

protein-protein interactions (PPIs), a challenging class of drug targets. We address

a flexible pocket, which is difficult to achieve by structure-based drug design. After

careful analysis we did not find one of the possible products in the kinetic

target-guided synthesis (KTGS) approach, however subsequent synthesis and biochemical

evaluation of each library member demonstrated that all the obtained molecules

inhibit MDM2. The most potent library member (K

i

= 0.095 μM) identified is almost

as active as Nutlin-3, a potent inhibitor of the p53–MDM2 PPI.

This chapter is adapted from the original publication:

R. van der Vlag*, M.Y. Unver*, T. Felicetti*, A. Twarda-Clapa, F. Kassim, C. Ermis, C.G. Neochoritis, B. Musielak, B. Labuzek, A. Dömling, T.A. Holak, A.K.H. Hirsch, Optimized Inhibitors of MDM2 via an Attempted Protein-Templated Reductive Amination,

ChemMedChem, 2020, 15, 370 * Shared first author

(3)

4.1 Introduction

Discovery of fast and efficient techniques to identify bioactive compounds constitutes an important part in today’s drug discovery. A way to accelerate hit-identification is the use of reversible reactions (dynamic combinatorial chemistry, DCC) or irreversible reactions (kinetic target-guided synthesis, KTGS), which are both categories of templated fragment

ligations.1 In these templated fragment ligations, the target selects its own inhibitors by

assembling the corresponding binders from a library of complementary building blocks or by binding and amplifying them from a library of compounds formed in a covalent bond-forming reaction. In KTGS, the biological target accelerates the irreversible reaction

between complementary building blocks upon binding,2,3 whereas in DCC a reversible

reaction between building blocks affords a dynamic combinatorial library (DCL) from which

the biological target selects and amplifies the best binder(s).4,5 Both techniques hold the

potential to accelerate drug discovery and are still relatively underexplored, especially in terms of target scope and availability of biocompatible reactions.

KTGS is a promising hit-identification strategy but only a few reactions with a limited

number of targets have been reported so far.1,2,6–19 The most widely used reaction is the

Huisgen 1,3-dipolar cycloaddition of azides and alkynes and most work in KTGS focuses on acetylcholine esterase (AChE) from various species, although different reactions and targets

have been explored.2,3

Protein-protein interactions (PPIs) are involved in many biological functions, such as intercellular communication and apoptosis. Targeting PPIs using small molecules is considered challenging given the flatness of the interface, a lack of small-molecule starting points for the future design and the difficulties in distinguishing real from artefactual

binding.20

P53 is a tumor-suppressor protein that is activated by cellular stress or damage and leads to cell-cycle arrest, apoptosis and DNA repair. MDM2 is the negative regulator of the

p53 protein and its overexpression leads to loss of p53 function.21,22 MDM2 has a

well-defined and deep pocket, unusual for PPIs, accommodating a hotspot triad consisting of

Trp23, Leu26 and Phe19 from p53.21,23,24 Therefore, design of high-affinity ligands to inhibit

MDM2 should focus on these hotspot amino acids of p53 (three-point pharmacophore

model). As a result, several small-molecule inhibitors have been reported.23–26 Some time

ago, we found the Leu26 pocket to be a flexible pocket, which is enlarged upon ligand

binding (four-point pharmacophore model).27 Very recently, we expanded this work with a

thorough SAR analysis.28 The flexibility of the pocket makes it very difficult to target using

structure-based drug design (SBDD) or other computational techniques such as virtual screening. Therefore, KTGS holds the potential to explore this flexible binding pocket by letting the protein select which combination of building blocks ideally fits.

Use of protein-templated reactions to interrupt PPIs has only been shown for the

Bcl-XL/BAK interaction11,17 and the 14-3-3 protein.7 These targets feature deep cavities in the

binding pocket, which makes them suitable for KTGS just like MDM2. The reactions used for

these two targets are the sulfo-click reaction between thio acids and sulfonyl azides and SN2

thiol ring opening of epoxides, respectively.

The reversible reaction between an aldehyde and an amine to afford imines has been

widely applied in DCC to target various biological targets (Figure 1).29–39 It is generally

(4)

building blocks in adjacent pockets, followed by imine formation and subsequent reduction, resulting in an irreversible amine bond. However, since in most cases the reduction agent is present in the mixture during most of the reaction time, it is unknown if this is actually true and difficult, if not impossible, to prove. Additionally, we think that protein-templated reductive amination is an example of KTGS instead of genuine DCC, in which the formed bond remains non-covalent.

Figure 1. Protein-templated reductive amination.

Since the formed imines are unstable in aqueous media, in situ reduction is performed in most cases. After analysis, frequently only the most amplified hits are synthesized and tested for activity. This approach, however, represents several pitfalls, such as: 1) it is assumed that the imines have a similar binding mode as the amines; and 2) the potency of the compound might change upon reduction. Previously, large differences in potency

between imine and amine have been reported.31,39 Furthermore, we noticed that in most

reports the potencies of the starting materials are not reported, although many of the starting materials are used as anchors.

Herein, we present our efforts to use protein-templated reductive amination in the KTGS context for the identification of inhibitors of the MDM2–p53 PPI. Furthermore, it is the first application of KTGS to address a flexible binding pocket (Leu26 pocket). Since we were unable to observe any of the reductive amination products, we evaluated all members of the library.

4.2 Results and Discussion

After the discovery of the extended Leu26 pocket in MDM2,27 we set out to explore this

flexible pocket. Having selected MDM2 as a target, we designed our potential inhibitor scaffold starting from the X-ray crystal structure of inhibitor 1 in complex with MDM2

(Figure 2, PDB ID: 4MDN).27

We designed and optimized a new inhibitor by using the molecular modeling program

SeeSAR.40 Inhibitor 1 occupies the three subpockets of MDM2 that are named according to

the corresponding p53 residues: the 6-chloroindole-2-hydroxamic acid moiety is hosted by the Trp23 pocket, the tert-butyl group occupies the Phe19 pocket and the large 4-clorobenzyl phenyl ether was found to fill the Leu26 and an induced subpocket. In order to occupy this extended subpocket in an optimal way by using the reductive amination reaction, we converted the 4-chlorobenzyl ether moiety into an amine, which can be assembled from the corresponding aldehyde 3 and amine 4 followed by in situ reduction (Scheme 1).

(5)

Figure 2. (Top) X-ray crystal structure of MDM2 in complex with docked inhibitor 1 (pink), superimposed with

designed inhibitor 2 (green) (PDB ID: 4MDN). Oxygen, nitrogen and chlorine atoms are shown in red, blue and light green, respectively. (Bottom) Design of inhibitor 2 as a product of reductive amination, based on the previously reported inhibitor 1.

Following the design of the initial inhibitor, we generated a combinatorial library by using aldehyde 3 as a core scaffold and eight different amines (4–11) to explore and fill the extended Leu26 pocket in the best manner and give rise to a focused SAR (Scheme 1).

Scheme 1. Selection of building blocks for the protein-templated reductive amination, which affords eight possible

(6)

The use of KTGS to fill flexible pockets is unprecedented. To explore this binding pocket and possibly open-up the Leu26 pocket even further, we used a benzyl- and naphthyl-amine in the library in addition to the aromatic amines to extend the length of the linker between the core scaffold and the amines. A potential extension reaching deeper into the pocket would be an important finding for future drug development.

Having selected the building blocks from commercially available amines, we synthesized the core scaffold (3) as shown in Scheme 2. Literature protocols afforded

compounds 12 and 13, in three and one step(s), respectively.41,42 A four-component Ugi

reaction of aldehyde 12, amine 13, formic acid (14) and tert-butyl-isocyanide (15) furnished the corresponding Ugi product 16 in 63% yield. Oxidation of the alcohol and subsequent hydrolysis of the ester led to the final aldehyde 3 in 86% yield over two steps.

Scheme 2. Synthesis of aldehyde building block 3. Conditions and reagents: i) three-step synthesis41 ii) one-step

synthesis42 iii) MeOH, rt, 6 d, 63%; iv) a) DMP, CH2Cl2, rt, 3 h, b) 1 M LiOH, H2O-EtOH (1:1), rt, 18 h, 86%.

Biochemical evaluation of aldehyde building block 3 showed promising potency

against MDM2 (Ki = 0.55 ± 0.05 μM). We envisioned that additional interactions in the

flexible pocket would enhance compound potency. Therefore, we set up two experiments in parallel using the synthesized aldehyde building block and eight different commercially available amines (4–11), a protein-templated reaction and a blank reaction at pH = 7.4

(100 mM phosphate buffer, 10% DMSO).

In both reactions, we applied standard building block and reducing agent

concentrations of 100 and 200 µM, respectively. One of the stringent requirements for KTGS

is a substantial difference in reaction rate between the blank and the protein-templated reaction. As the imine formation between an aldehyde and an amine is fast, we used dilute

conditions to prevent product formation in the reference reaction for a certain time.43 As a

result, less protein is required, an important consideration especially for precious proteins. By using a reducing agent in the reaction mixture from the beginning, formed imines would be “frozen” by reduction, leading to detectable amounts of products. Therefore, the two

reactions were started by mixing all amines 4–11 (100 µM each), aldehyde 3 (100 µM) and

NaCNBH3 (200 µM). To the protein-templated reaction, we added MDM2 (100 µM). Then,

both reactions were carefully monitored using UPLC-MS analysis (wavelengths: 254 and 305 nm) for two days. After two days, signs of denaturation were observed in all MDM2 containing samples. We did not detect any trace of one of the reductive amination products, while a large amount of the starting materials were observed. Additionally, experiments in which MDM2 was replaced by bovine serum albumin (BSA), experiments without reducing

agent and experiments in which the potent MDM2 inhibitor Nutlin-3 (100 µM, Ki = 36 nM44)

was added, showed identical results (Figure 4). In order to estimate how much of the product(s) should be formed for detection, we assessed the detection limit of the UPLC-MS. The UV-visible spectrum of 3 showed a clear absorption band around 305 nm (Figure S10).

(7)

This wavelength has the benefit over 254 nm that there is a lot less background signal. Under identical circumstances as the protein-templated reaction, the detection limit of

aldehyde 3 was estimated to be <2.6 μM and <0.88 μM for 254 and 305 nm, respectively

(Figures S11–13). In case only one reductive amination product is formed and taking into account the protein denaturation step with acetonitrile, this corresponds to as low as 1.8% conversion at 305 nm. Even if all eight amines reacted in a similar fashion, generating all possible products, this would require only 14% conversion of 3. We believe that in order to claim a protein-templated reaction, we should be far above these numbers.

In order to determine the potency of each product we synthesized all possible products (2 and 17–23, Table 1), starting from the core scaffold 3 by using a reductive amination protocol (Scheme 3).

Figure 4. Close-up of the UPLC traces at 305 nm in the area of which all products are expected (after two days

reaction time). From top to bottom: blank (black), reaction without protein (red), MDM2 reaction (green), MDM2 reaction without reducing agent (blue), MDM2 reaction in presence of (±)-Nutlin-3 (yellow) and reaction in presence of BSA instead of MDM2 (purple). * Reduced aldehyde of 3. Note: amines 4, 5, 6, and 11 have limited or no absorption at 305 nm. The retention times of all starting materials and products are shown in Table S1.

Scheme 3. Synthesis of the inhibitors 2 and 17–23. (a) Conditions and reagents: amines 4–11, pyrrolidine, 4Å

molecular sieves, Na(CH3CO2)3BH, dry CH2Cl2, rt, 18 h.

10

Red. 3*

7 8 9 3

(8)

We evaluated the inhibitory potency of compounds 2 and 17–23 by using the

fluorescence polarization assay as previously described.45 MDMX is another p53 binding

protein and shows significant homology with MDM2. In order to investigate if the compounds selectively inhibit one of the proteins, we tested all compounds against MDM2 and MDMX (Table 1).

As can be seen from Table 1, the majority of the compounds (2 and 17–21, Ki = 0.40–

0.76 μM) show an activity against MDM2 in the same range as aldehyde 3 (Ki = 0.55 ±

0.05 μM) and lead compound 1 (Ki = 0.60 μM).27 Amine 23 (Ki = 3.18 ± 0.18 μM) is almost

six-fold less potent than aldehyde 3. Interestingly, 18 (Ki = 0.63 ± 0.07 μM) is three times more

potent than its regioisomer in which the CH2–NH is reversed (Ki = 1.9 μM).28 Surprisingly,

compound 22 shows a Ki value of 0.095 ± 0.010 μM, which makes it by far the most potent

inhibitor of this series. In fact, the compound is almost as potent against MDM2 as Nutlin-3

(Ki = 0.036 ± 0.009 μM44).

The inhibitory activities of the synthesized reductive amination products against

MDMX are rather comparable (Ki = 3.73–7.08 μM). Except for compound 17 which has a

slightly higher Ki of 12.2 ± 2.1 μM and 23, which is not active against MDMX under the

applied conditions. Aldehyde 3 (Ki = 11.4 ± 1.6 μM) also has a reduced inhibitory activity

against MDMX in comparison to MDM2. Most compounds have a preference for MDM2 by a factor 7–16. With a selectivity factor of 21, aldehyde 3 is slightly more selective for MDM2 than most of the amines. A remarkable exception is compound 22, which is over four times more selective for MDM2 than inhibitors 2 and 18–21 and two times more selective than

aldehyde 3. Although the activity of 22 against MDM2 is close to the Ki value of Nutlin-3, it

is around six times less selective.

There are several possible explanations as to why we did not find any of the reductive amination products in the protein-templated experiments. Very recently, Van der Veken and co-workers reported their efforts towards the application of the

(9)

Table 1. Biochemical evaluation of all reductive amination products. Experiments were performed in duplicate and

values are reported as average ± standard deviation. Note: at physiological pH there is a difference in charge expected between the aniline-type products and benzylamine-type compounds 17 and 23.

Inhibitor Ki (μM) MDM2 Ki (μM) MDMX Selectivity (MDMX/MDM2) 3 Aldehyde 0.55 ± 0.05 11.4 ± 1.6 21 ± 3.5 2 0.40 ± 0.05 4.18 ± 0.27 10 ± 1.5 17 0.76 ± 0.08 12.2 ± 2.1 16 ± 3.2 18 0.63 ± 0.07 4.56 ± 0.49 7.2 ± 1.1 19 0.75 ± 0.08 7.08 ± 0.52 9.4 ± 1.2 20 0.47 ± 0.04 3.73 ± 0.30 7.9 ± 0.9 21 0.49 ± 0.04 4.61 ± 0.39 9.4 ± 1.1 22 0.095 ± 0.010 3.95 ± 0.46 42 ± 6.5 23 3.18 ± 0.18 >15 ≥5 Nutlin-344 0.036 ± 0.009 9.38 ± 0.35 261 ± 66

The imine formation under near-physiological conditions showed to be most challenging. Furthermore, although their final products were also very potent against the protein target, the target itself interfered with the attempted reaction, which could also be the case in our study. Additionally, it could be that the reducing agent does not reach the imine, which is therefore hydrolyzed upon analysis. Last, there are clear differences in rigidity and hydrogen-bond donor/acceptor profiles between the imines and the amines. It can well be that the aldehyde and reductive amination products show activity against MDM2, but that the imines cannot be formed or do not fit in the active site. We performed crystallization studies to confirm the binding mode of the new inhibitors, but due to solubility problems we could not obtain any crystals.

To confirm the inhibitory activity of the compounds further, the uniformly 15N-labelled

MDM2 was titrated with an increasing concentration of compound 17, and 1H-15N HSQC

spectra were recorded after each new portion of the inhibitor was added. Instead of the most potent compound, 22, compound 17 was used due to its better solubility. The method is based on monitoring changes in NMR chemical shifts in protein amide backbone

(10)

resonances upon its interaction with a small molecule.47–50 In the course of titration, shifts

or disappearances of cross-peaks assigned to the amino acids of MDM2 affected by binding of 17 were observed, which corroborates the binding event (Figure 5).

Figure 5. (Left) 1H-15N HSQC NMR spectra: blue, reference spectrum of apo-MDM2; red, 1:5 (MDM2: inhibitor 17,

overtitrated). (Right) Chemical shift perturbations plotted onto the structure of the complex of MDM2 (gray) with p53 (green) (PDB ID: 1YCR51); residues which shift or disappear upon titration are labeled on the MDM2 surface

(red).

4.3 Conclusions

In conclusion, herein, we reported our efforts towards the first example of a reductive amination in KTGS applied to a PPI target. By using our target MDM2 in situ, we screened a library of compounds in one-pot, which did not reveal a clear hit over a course of two days. After synthesis of all library members, we found a very potent and rather selective inhibitor

of MDM2 (Ki = 0.095 ± 0.010 μM). Although, the KTGS method can find application in the

early stages of drug discovery, one should be careful in the use of imine-based chemistry. Particularly interesting scenarios for the use of KTGS in drug discovery are flexible protein pockets that are difficult to target by structure-based approaches.

(11)

4.4 Experimental Section

4.4.1 Protein purification procedures

4.4.1.1 MDM2

Fragments of human MDM2 (residues 1–118 in pET46) were expressed in the E. coli BL21-CodonPlus (DE3) RIL strain. Cells were cultured in a total volume of 5 L of LB or minimal medium at 37 °C, and induced with 1 mM IPTG at OD600nm of 0.8. Protein was expressed for 5 h at 37 °C. Cells were collected

by centrifugation, re-suspended in 120 mL PBS with protease inhibitor cocktail and lysed by sonication. Inclusion bodies that were collected by centrifugation, were washed twice with 120 mL PBS containing 0.05% Triton-X100 and once with 120 mL PBS and centrifuged after each wash. Purified inclusion bodies were solubilized in 20 mL of 6 M guanidine hydrochloride in 100 mM Tris-HCl, pH 8.0, containing 1 mM EDTA and 10 mM β-mercaptoethanol. The protein was dialyzed against 1 L of 4 M

guanidine hydrochloride, pH 3.5 supplemented with 10 mM β-mercaptoethanol. Following, the protein was refolded by dropwise addition into 1 L of 10 mM Tris-HCl, pH 7.0, containing 1 mM EDTA

and 10 mM β-mercaptoethanol and slow mixing overnight at 4 °C. Ammonium sulphate was added to a final concentration of 1.5 M, mixed for 2 h and centrifuged. The refolded protein was recovered on Butyl Sepharose 4 Fast Flow (GE Healthcare) previously equilibrated with refolding buffer containing 1.5 M (NH4)2SO4. MDM2 was eluted using 100 mM Tris-HCl buffer (pH 7.2), containing 5 mM

β-mercaptoethanol. Fractions containing the protein were pooled, concentrated to <10 mL and further purified by gel filtration on an S75 16/600 column (GE Healthcare) in 50 mM phosphate buffer pH 7.4 containing 150 mM NaCl and 5 mM DTT (FP/NMR buffer).

4.4.1.2 MDMX

Fragments of human MDMX (residues 1–134 in pET46) were expressed in E. coli BL21-CodonPlus (DE3) RIL strain. Cells were cultured in a total volume of 5 L of LB or minimal medium at 37 °C and induced with 0.5 mM IPTG at OD600nm of 0.6. Protein was expressed for 12 h at 20 °C. Cells were collected by

centrifugation, re-suspended in 120 mL 50 mM NaH2PO4 pH 8.0 with 300 mM NaCl, 10 mM imidazole

and protease inhibitor cocktail and lysed by sonication. The protein was purified under native conditions. The lysate was cleared by centrifugation and loaded on Ni-chelating sepharose (GE Healthcare) previously equilibrated with lysis buffer (50 mM NaH2PO4 pH 8.0 with 300 mM NaCl and 10 mM imidazole). MDMX was eluted with 50 mM NaH2PO4 pH 8.0 containing 300 mM NaCl and

300 mM imidazole. Fractions containing the protein were pooled, concentrated and further purified

by gel filtration on S75 16/600 column (GE Healthcare) in the FP/NMR buffer.

4.4.2 Fluorescence polarization assay

A fluorescence polarization (FP) assay was used to monitor interactions between MDM2 and MDMX proteins and their inhibitors.45

For each assay, fresh protein stocks of MDM2 (1–118) and MDMX (1–134) were thawed, and the protein concentrations were determined using Bradford method. The assay buffer contained 50 mM

NaCl, 10 mM Tris pH 8.0, 1 mM EDTA and 5% DMSO. Corning black 96-well NBS assay plates were used. The binding affinity of P2 peptide (sequence: LTFEHYWAQLTS, labelled with carboxyfluorescein) towards MDM2 and MDMX was first determined. For this purpose, 10 nM of the fluorescent P2 peptide was contacted with serial dilutions of tested protein (range from 750 to 0.012 nM for MDM2 and from 3750 to 0.10 nM for MDMX) in a final volume of 100 μL and fluorescence polarization was determined. Kd was determined by fitting the curve described to

(12)

FP = FPmin(FP𝑚𝑎𝑥− FP𝑚𝑖𝑛) ∙ c 𝐾d+ c

where FP is the determined value of fluorescence polarization, FPmin is the fluorescence polarization

for ligand only, FPmax is the fluorescence polarization at protein concentration saturating the ligand

and c is the protein concentration.

Competition binding assay was performed using 10 nM fluorescent P2 peptide and optimal protein concentration for the measurement calculated based on determined Kd according to Huang, 2003 (f0

= 0.8).52 Tested compounds were dissolved in DMSO at 50 µM. Serial dilutions (50 μM to 0.05 μM) were

prepared in DMSO.

All the experiments were prepared in duplicate, and plates were read 15 min after mixing of all assay components. Fluorescence polarization was determined using Tecan InfinitePro F200 plate reader with the 485 nm excitation and 535 nm emission filters. Fluorescence polarization values were expressed in millipolarization units (mP).

4.4.3

1

H-

15

N HSQC NMR experiment

Uniform 15N isotope labelling was obtained by expression of the protein in M9 minimal medium

containing 15NH4Cl as the sole nitrogen source. The final step of purification of MDM2 for NMR

consisted of gel filtration into the NMR buffer (50 mM phosphate buffer pH 7.4 containing 150 mM

NaCl, 5 mM DTT). 10% (v/v) of D2O was added to the samples to provide lock signal. Stock solutions of

inhibitors of MDM2/MDMX used for titration were prepared in DMSO-d6. The samples were prepared

by adding 50 mM ligand stock solution to the protein solution containing the 15N-labeled MDM2

fragment at a concentration of 0.3 mM. A 2D 1H-15N correlated heteronuclear single quantum

coherence (HSQC) NMR spectrum was recorded at 2–3 different ligand/protein ratios. All the spectra were recorded at 300 K using a Bruker Avance 600 MHz spectrometer. 1H-15N heteronuclear

correlations were obtained using the fast HSQC pulse sequence.53 Assignment of the amide backbone

resonances of MDM2 was obtained according to Stoll et al., 2001.54 The spectra were processed with

the TopSpin 3.2 software.

4.4.4 Modeling and docking

The inhibitor scaffold was modeled starting from the MDM2 X-ray crystal structure (PDB ID: 4MDN).27

Modeling and docking was performed using SeeSAR version 9.2.40 The distances between atoms in

and around the substituted formamide moiety in the X-ray crystal structure are so short, that the BiosolveIT programs show the structure as two connected cyclopropane rings (the authors also incorporated a remark at the beginning of the PDB file). Upon docking of this configuration, many poses were obtained, which are close to the original X-ray crystal structure, indicating the reproducibility by docking. Having restored the original 3D configuration manually, the molecule was docked by generating several poses. Whereas the indole part of the molecule is fully retained in position, the phenoxy moiety shifts a little bit towards the Leu26 pocket. However, the para-chlorine substituted phenyl more or less retains its original position. Subsequently, the pose of inhibitor 1, which resembles most closely the X-ray crystal structure was changed to 2, after which the structure was re-docked by generating several poses. The newly generated poses for 2 showed a high similarity to the poses of 1, which indicates that based on the docking no large shift or difference in binding mode is predicted. Based on these results it is expected that 2 can inhibit MDM2 in a similar way as compound 1, and that our KTGS strategy could potentially work to target the Leu26 pocket. A representative example of the docked poses of 1 and 2 is shown in Figure 2.

(13)

4.4.5 Attempted protein-templated reductive amination

To a 2 mL Eppendorf was sequentially added the eight building blocks 4—11 (0.5 μL each, 100 mM in DMSO), aldehyde 3 (0.5 μL, 100 mM in DMSO), DMSO (45.5 μL), phosphate buffer (235 μL, 100 mM, pH 7.4) and MDM2 (214 µL, 0.213 mM in phosphate buffer 100 mM, pH 7.4). After 10 minutes,

NaCNBH3 (1.0 μL, 100 mM in CH3CN) was added. The reaction mixture was allowed to rotate at room

temperature with 10 rpm using a Falc F205 rotary shaker. After one hour and after 1, 2 and 7 days, a 50 μL aliquot of reaction mixture was added to 50 μL of CH3CN, resulting in protein denaturation. The

precipitate was removed by centrifugation and filtration, whereupon the mixture was analyzed by UPLC-MS.

Control reactions

In addition to the reductive amination reaction, the following control reactions were performed: • Blank reaction (negative control) using buffer instead of MDM2

• BSA reaction using BSA (449 μL, 0.111 mM in phosphate buffer 100 mM, pH 7.4) instead of MDM2

• MDM2 reaction in the presence of (±)-Nutlin-3 (2.0 μL, 25 mM in DMSO, solid was obtained from Sanbio B.V., NL)

• MDM2 reaction with the addition of 1 μL CH3CN instead of NaCNBH3 solution

Formation of the individual reductive amination products

Each product of the library was formed by reacting each amine (4–11) with aldehyde 3 in the presence of reducing agent. The amine (2.5 μL, 100 mM in DMSO) and aldehyde 3 (2.5 μL, 100 mM in DMSO) were added to a mixture of MilliQ (450 μL) and DMSO (40 μL). After 10 minutes, NaCNBH3 (5 μL,

100 mM in CH3CN) was added, and the reaction mixtures were followed by UPLC-MS over time (1, 2,

4 and 7 days).

UPLC-MS method

All reactions were repeated twice and analyzed using a Thermo Fischer Scientific Vanquish with LCQ Fleet detector (of which the detection limit was estimated). All analyses were performed using a reversed-phase UPLC column (ACQUITY BEH C8 Column, 130 Å, 1.7 µm, 2.1 mm x 150 mm) on which 1 μL sample was injected. The eluents, acetonitrile and water contained 0.1% of formic acid. The library components were eluted with a gradient from (water): 95% → 10% over 4 min, then at 10% for 11 min, after which the column was re-equilibrated to 95%. The MS (m/z 100-700, ESI+) and the absorbance spectra at 254 and 305 nm were carefully analyzed and compared.

(14)

4.4.6 UV-vis absorption spectra of 3, 22 and 24

Compounds 3, 22 and 24 (pre-cursor of 12) were dissolved in MilliQ/ACN 50:50 containing 0.1% formic acid to obtain 0.1 mM solutions. After centrifugation, the solutions were added to a 96-well plate and UV-visible spectra were recorded from 230 – 700 nm using a BioTek Synergy H1 hybrid plate reader to obtain absorption maxima of 0.6 – 1. The blank absorption was subtracted and the absorption spectra were normalized at 305 nm.

4.4.7 Synthesis

General Experimental Details

Starting materials and reagents were purchased from Aldrich or TCI Europe. Yields refer to analytically pure compounds and were not optimized. All solvents were reagent-grade and if necessary, SPS-grade. Column chromatography was performed on silica gel (Silicycle® SiliaSepTM 40–63 μM 60 Å). TLC

was performed with silica gel 60/Kieselguhr F254. Solvents used for the column chromatography were pentane, ethyl acetate, dichloromethane and methanol. 1H and 13C spectra were recorded at 25 °C on

a Varian AMX400 spectrometer (400 MHz for 1H and 101 MHz for 13C, respectively) or a Bruker

Ascend™ 600 MHz (600 MHz for 1H and 151 MHz for 13C, respectively). Chemical shifts (δ) are reported

in ppm relative to the residual solvent peak. Splitting patterns are indicated as (s) singlet, (d) doublet, (t) triplet, (q) quartet, (m) multiplet, (br) broad. The coupling constants (J) are given in Hz. High-resolution mass spectra were recorded with an Orbitrap XL (Thermo Fisher Scientific; Esi pos. mode). UPLC-MS analysis was performed using a Thermo Fischer Scientific Vanquish with LCQ Fleet detector (purity analysis of 20–22) or a Waters Acquity Ultra Performance LC with Acquity TQD detector (purity analysis of the remaining final products). Melting points were measured on a Stuart®SMP11 50 W melting point apparatus.

Final products give a mixture of rotamers in NMR (CD3OD at 25oC). Major peaks were

identified and are reported in 1H-NMR spectra. For the 13C-NMR spectra, all peaks corresponding to

the complete mixture of rotamers are reported. All compounds are ≥95% purity as shown by UPLC-MS analysis (254 nm).

General procedure for compounds 2 and 17–23.

To a solution of compound 3 (1.00 eq), amine (1.05 eq), and pyrrolidine (0.10 eq) in dry CH2Cl2, 4 Å

molecular sieves (100 mg per mmol 3) and Na(CH3CO2)3BH (2.0 eq) were added. The reaction mixture

was stirred at room temperature for 18 h under a nitrogen atmosphere. Then, the mixture was filtered over Celite®, and the filtrate was evaporated under vacuum. The crude product was purified by flash chromatography column, eluting with ammonia-infused (1/1 (v/v) ammonia/CH2Cl2 mixed and CH2Cl2

(15)

Ethyl 3-(2-(tert-butylamino)-1-{formyl[4-(hydroxymethyl)benzyl]amino}-2-oxoethyl)-6-chloro-1H-indole-2-carboxylate (16)

In a 50-mL round-bottomed flask charged with MeOH (10 mL), aldehyde

12 (700 mg, 2.78 mmol) was added to amine 13 (381 mg, 2.78 mmol) and

the yellow reaction mixture was stirred for one hour at room temperature. Then, formic acid (110 μL, 2.78 mmol) and tert-butyl isocyanide (312 μL, 2.78 mmol) were added and the reaction mixture was stirred at room temperature for six days. Then, the red reaction mixture was evaporated under vacuum, and the orange crude was purified by flash column chromatography eluting with CH2Cl2/MeOH (99:1 → 95:5). Compound 16

was obtained as a pale yellow solid (871 mg, 1.74 mmol, 63% yield). M.p. = 113–115 °C; HRMS (ESI+) for C26H31ClN3O5 [M + H]+: 500.195 (calc.), 500.195 (found). 1H-NMR (400 MHz, CD3OD, major rotamer,

53%) δ 8.38 (s, 1H), 7.79 (d, 1H, J = 8.8 Hz), 7.39 (d, 1H, J = 1.8 Hz), 7.12 (dd, 1H, J = 8.8 and 1.9 Hz), 7.00 (d, 2H, J = 7.8 Hz), 6.76 (d, 2H, J = 7.9 Hz), 6.19 (s, 1H), 5.11 (d, 1H, J = 15.2 Hz), 4.46 (s, 2H), 4.38 (q, 2H, J = 7.0 Hz), 4.32 – 4.17 (m, 1H), 1.41 (t, 3H, J = 7.1 Hz), 1.23 (s, 9H); 13C-NMR (101 MHz, CD3OD) (mixture of rotamers) δ 171.2 (C), 170.9 (C), 166.6 (CH), 166.0 (CH), 162.2 (C), 162.1 (C), 141.4 (C), 141.3 (C), 138.0 (C), 137.9 (C), 137.8 (C), 137.4 (C), 132.1 (C), 131.8 (C), 129.5 (C), 128.5 (C), 128.4 (2xCH), 127.3 (2xCH), 127.2 (2xCH), 126.9 (C), 126.7 (2xCH), 126.1 (C), 123.5 (CH), 123.3 (CH), 122.6 (CH), 122.4 (CH), 115.9 (C), 114.0 (C), 113.3 (CH), 113.2 (CH), 64.8 (CH2), 64.7 (CH2), 62.2 (CH2), 62.1 (CH2), 58.3 (CH), 54.2 (CH), 52.54 (C), 52.52 (C), 51.2 (CH2), 48.0 (CH2), 28.9 (CH3), 28.8 (CH3), 14.8 (3xCH3), 14.7 (3xCH3). 3-{2-(tert-Butylamino)-1-[formyl(4-formylbenzyl)amino]-2-oxoethyl}-6-chloro-1H-indole-2-carboxylic acid (3)

To a solution of 16 (400 mg, 0.801 mmol) in dry CH2Cl2 (16 mL), Dess-Martin

periodinane (350 mg, 0.825 mmol, 1.03 eq.) was added portion-wise. The reaction was stirred at room temperature for 3 h under a N2 atmosphere.

Then, the mixture was quenched with a solution of Na2S2O3 (10%, 6 mL),

followed by a saturated aq. solution of NaHCO3 (10 mL) and extracted with

CH2Cl2. The combined organic layer was washed with a saturated aq.

solution of NaHCO3 and a saturated aq. solution of NaCl. Then the organic

layer was dried over MgSO4, filtered and evaporated to dryness to give the

oxidized intermediate as pale yellow solid that was quickly dissolved in EtOH (8 mL). The reaction mixture was cooled down using an ice-water bath and an aq. solution of LiOH (1.0 M, 0.19 g, 8.00 mmol, 8 mL) was added dropwise. Afterwards, the ice bath was removed, and the reaction mixture was stirred at room temperature for 18 h. The reaction mixture was again cooled down using an ice-water bath, and the mixture was slowly acidified with an aq. HCl (1 M) solution to pH 6. Filtration

of the precipitate afforded compound 3 as an off-white solid (324 mg, 0.689 mmol, 86% yield). M.p. = 211–213 ° C; HRMS (ESI+) for C24H25ClN3O5 [M + H]+: 470.148 (calc.), 470.148 (found). 1H-NMR (600

MHz, DMSO-d6, major rotamer, 75%) δ 13.47 (br. s, 1H, -OH), 11.81 (br. s, 1H, indole-NH), 9.86 (s, 1H),

8.29 (s, 1H), 7.76 (d, 1H, J = 8.8 Hz), 7.73 (s, 1H, amide-NH), 7.56 (d, J = 8.2 Hz, 2H), 7.32 (d, 1H, J = 2.0 Hz), 7.09 (dd, 1H, J = 8.8, 1.9 Hz), 7.01 (d, 2H, J = 7.9 Hz), 6.14 (s, 1H), 4.97 (d, 1H, J = 16.0 Hz), 4.36 (d, 1H, J = 16.1 Hz), 1.15 (s, 9H); 13C-NMR (151 MHz, DMSO-d 6) (mixture of rotamers) δ 192.6 (CH), 192.5 (CH), 168.88 (C), 168.82 (C), 163.80 (CH), 163.55 (CH), 162.39 (C), 161.82 (C), 145.78 (C), 144.89 (C), 135.8 (C), 135.7 (C), 134.53 (C), 134.52 (C), 128.9 (C), 128.72 (2xCH), 128.66 (2xC), 128.5 (C), 128.4 (2xCH), 127.3 (2xCH), 125.9 (2xCH), 125.5 (C), 124.7 (C), 122.3 (CH), 122.1 (CH), 120.7 (CH), 120.5 (CH),

(16)

113.9 (C), 112.6 (C), 111.9 (CH), 111.7 (CH), 56.2 (CH), 51.9 (CH), 50.6 (C), 50.5 (C), 49.2 (CH2), 46.6

(CH2), 28.5 (3xCH3), 28.3 (3xCH3).

3-(2-(tert-Butylamino)-1-(N-(4-(((4-chlorophenyl)amino)methyl)benzyl)formamido)-2-oxoethyl)-6-chloro-1H-indole-2-carboxylic acid (2)

Following the general procedure, starting from 4-chloroaniline (12.7 mg, 0.0966 mmol) as amine, compound 2 was obtained as a white solid (20 mg, 0.034 mmol, 34% yield). M.p. >119 °C (degradation); HRMS (ESI+) For C30H31Cl2N3O4 [M + H]+:

581.172 (calc.), 581.171 (found). 1H-NMR (400 MHz, CD3OD,

major rotamer, 75%) δ 8.36 (s, 1H), 7.71 (d, 1H, J = 8.8 Hz), 7.46 (br. s, 1H), 7.36 (d, 1H, J = 1.8 Hz), 7.12 (d, 2H, J = 7.9 Hz), 7.04 – 6.97 (m, 5H), 6.54 (d, 2H, J = 8.7 Hz), 6.42 (s, 1H), 5.03 (d, 1H,

J = 15.1 Hz), 4.36 (d, 1H, J = 15.1 Hz), 4.18 (s, 2H), 1.15 (s, 9H), note: of the exchangeable protons only

one broad singlet was observed; 13C-NMR (101 MHz, CD3OD) δ 171.2, 170.9, 166.7, 166.0, 148.87,

148.85, 139.8, 139.6, 137.7, 137.6, 137.5, 136.8, 131.6, 130.6, 129.7, 128.9, 128.7, 127.8, 127.7, 127.0, 126.2, 123.4, 123.2, 122.3, 122.2, 122.1, 115.1, 115.0, 113.2, 113.1, 58.3, 54.3, 52.5, 52.4, 51.2, 48.3, 48.2, 28.9, 28.8.

3-{2-(tert-Butylamino)-1-[formyl(4-{[(1-naphthylmethyl)amino]methyl}benzyl)amino]-2-oxoethyl}-6-chloro-1H-indole-2-carboxylic acid (17)

Following the general procedure, starting from 1-naphthylmethylamine (21 mg, 0.13 mmol) as amine, compound 17 was obtained as a white solid (20 mg, 0.033 mmol, 26% yield). M.p. 214–217 °C; HRMS (ESI+) for C35H36ClN4O4

[M + H]+: 611.242 (calc.), 611.243 (found). 1H-NMR (400 MHz,

DMSO-d6, major rotamer, 80%) δ 11.63 (br. s, 1H), 8.28 (s, 1H),

8.14 – 8.07 (m, 1H), 7.93 (d, 1H, J = 9.4 Hz), 7.86 (d, 1H, J = 8.1 Hz), 7.68 (d, 1H, J = 8.8 Hz), 7.60 (s, 1H), 7.58 – 7.43 (m, 4H), 7.35 (d, 1H, J = 1.8 Hz), 7.17 (d, 2H, J = 8.0 Hz), 7.08 – 6.99 (m, 2H), 6.96 (d, 2H, J = 8.0 Hz), 6.30 (s, 1H), 4.86 (d, 1H, J = 15.5 Hz), 4.36 (d, 1H, J = 15.5 Hz), 4.24 (s, 2H), 3.87 (s, 2H), 1.11 (s, 9H), note: the carboxylic acid proton was not observed; 13C-NMR (101 MHz, DMSO-d6) (mixture of rotamers) δ 169.2, 169.0,

163.8, 163.5, 163.4, 137.7, 137.0, 135.6, 135.3, 133.3, 131.40, 131.35, 128.4, 128.04, 127.95, 127.8, 127.7, 127.6, 127.2, 126.83, 126.77, 126.11, 126.05, 125.8, 125.7, 125.3, 125.0, 123.9, 123.9, 122.4, 121.8, 120.0, 112.2, 111.8, 56.6, 51.5, 50.41, 50.40, 48.5, 46.7, 28.5, 28.3.

3-(1-(N-(4-(((4-Bromophenyl)amino)methyl)benzyl)formamido)-2-(tert-butylamino)-2-oxoethyl)-6-chloro-1H-indole-2-carboxylic acid (18)

Following the general procedure, starting from 4-bromoaniline (17.2 mg, 0.100 mmol) as amine, compound 18 was obtained as a white solid (12 mg, 0.019 mmol, 20% yield). M.p. >129 °C (degradation); HRMS (ESI+) for C30H31BrClN4O4 [M + H]+:

625.121 (calc.), 625.121 (found). 1H-NMR (400 MHz, CD3OD,

major rotamer, 77%) δ 8.37 (s, 1H), 7.71 (d, 1H, J = 8.8 Hz), 7.35 (d, 1H, J = 1.9 Hz), 7.16 – 7.10 (m, 4H), 7.04 – 6.98 (m, 3H), 6.50 (d, 2H, J = 8.7 Hz), 6.46 (s, 1H), 5.01 (d, 1H, J = 15.0 Hz), 4.38 (d, 1H, J = 15.0 Hz), 4.18 (s, 2H), 1.15 (s, 9H), note: the carboxylic acid and amine protons were not observed; 13C-NMR (101 MHz, CD3OD) (mixture of rotamers) δ 171.7, 171.2, 168.0, 166.7, 166.1,

(17)

149.30, 149.26, 139.8, 139.4, 137.9, 136.93, 136.86, 136.7, 132.6, 130.5, 130.4, 129.2, 128.0, 128.0, 127.4, 127.1, 126.4, 123.3, 122.9, 121.8, 121.6, 115.6, 115.5, 112.9, 112.8, 112.3, 108.9, 108.8, 58.3, 54.7, 52.3, 52.2, 51.1, 48.4, 48.3, 48.1, 28.9, 28.8.

3-(2-(tert-Butylamino)-1-(N-(4-(((2,4-dimethylphenyl)amino)methyl)benzyl)formamido)-2-oxoethyl)-6-chloro-1H-indole-2-carboxylic acid (19)

Following the general procedure, starting from 2,4-dimethylaniline (12.1 mg, 0.100 mmol) as amine, compound

19 was obtained as a white solid (19 mg, 0.033 mmol, 34% yield).

M.p. >132 °C (degradation); HRMS (ESI+) for C32H36ClN4O4,

[M + H]+ 575.242 (calc.), 575.241 (found). 1H-NMR (400 MHz,

CD3OD, major rotamer, 75%) δ 8.37 (s, 1H), 7.71 (d, 1H, J = 8.8

Hz), 7.30 (d, 1H, J = 1.8 Hz), 7.13 (d, 2H, J = 8.0 Hz), 7.05 – 6.91 (m, 3H), 6.87 – 6.72 (m, 2H), 6.43 – 6.36 (m, 2H), 5.05 (d, 1H, J = 15.0 Hz), 4.34 (d, 1H, J = 15.1 Hz), 4.25 (s, 2H), 2.16 (s, 3H), 2.14 (s, 3H), 1.15 (s, 9H), note: the carboxylic acid and amine protons were not observed; 13C-NMR (101 MHz, CD

3OD) (mixture of rotamers) δ 171.2,

166.7, 166.1, 145.1, 140.4, 140.2, 137.6, 137.0, 136.6, 131.8, 131.7, 130.7, 129.1, 128.10, 128.07, 128.0, 127.8, 126.98, 126.95, 126.4, 123.7, 123.3, 122.9, 121.9, 121.8, 113.00, 112.98, 112.9, 112.0, 111.9, 58.3, 54.5, 52.4, 52.3, 51.2, 28.9, 28.8, 20.5, 17.8.

3-(2-(tert-Butylamino)-1-(N-(4-(((4-chloro-2 methylphenyl)amino)methyl)benzyl)formamido)-2-oxoethyl)-6-chloro-1H-indole-2 carboxylic acid (20)

Following the general procedure, starting from 4-chloro-2-methylaniline (15.7 mg, 0.111 mmol) as amine, compound 20 was obtained as a white solid (13.9 mg, 0.0233 mmol, 22% yield). M.p. >123 °C (degradation); HRMS (ESI+) for C31H33Cl2N4O4 [M + H]+: 595.187 (calc.), 595.187 (found). 1H-NMR (400 MHz, CD3OD, major rotamer, 70%) δ 8.36 (s, 1H),

7.73 (d, 1H, J = 8.8 Hz), 7.34 (d, 1H, J = 1.9 Hz), 7.10 (d, 2H,

J = 7.9 Hz), 7.03 (dd, J = 8.8, 1.9 Hz, 1H), 6.99 – 6.88 (m, 4H),

6.37 (d, 1H, J = 8.6 Hz), 6.33 (s, 1H), 5.07 (d, 1H, J = 15.1 Hz), 4.33 (d, 1H, J = 15.4 Hz), 4.27 (s, 2H), 2.15 (s, 3H), 1.15 (s, 9H), note: the carboxylic acid and amine protons were not observed; 13C-NMR (101

MHz, CD3OD) (mixture of rotamers) δ 171.3 (C), 171.0 (C), 166.7 (CH), 166.1 (CH), 146.4 (C), 146.3 (C),

140.0 (C), 139.8 (C), 137.7 (C), 137.4 (C), 137.2 (C), 136.7 (C), 131.3 (C), 131.1 (C), 130.4 (CH), 129.1 (CH), 127.8 (CH), 127.6 (CH), 127.3 (CH), 127.1 (C), 127.0 (CH), 126.3 (C), 125.4 (C), 125.3 (C), 123.4 (CH), 123.0 (CH), 122.1 (CH), 122.02 (CH), 121.96 (C), 121.9 (C), 114.0 (C), 113.1 (CH), 113.0 (CH), 112.3 (CH), 112.2 (CH), 58.3 (CH), 54.4 (CH), 52.5 (C), 52.3 (C), 51.2 (CH2), 48.3 (CH2), 48.2 (CH2), 48.1 (CH2),

(18)

3-(1-(N-(4-(((4-Bromo-2-methylphenyl)amino)methyl)benzyl)formamido)-2-(tert-butylamino)-2-oxoethyl)-6-chloro-1H-indole-2-carboxylic acid (21)

Following the general procedure, starting from 4-bromo-2-methylaniline (19 mg, 0.10 mmol) as amine, compound 21 was obtained as a white solid (21 mg, 0.033 mmol, 33% yield). M.p. >141 °C (degradation); HRMS (ESI+) for C31H33BrClN4O4

[M + H]+ 639.137 (calc.), 639.137 (found). 1H-NMR (400 MHz,

CD3OD, major rotamer, 75%) δ 8.37 (s, 1H), 7.71 (d, 1H, J = 8.8

Hz), 7.34 (d, 1H, J = 1.9 Hz), 7.17 – 6.91 (m, 7H), 6.41 (s, 1H), 6.33 (d, 1H, J = 8.6 Hz), 5.05 (d, 1H, J = 15.0 Hz), 4.35 (d, 1H,

J = 15.1 Hz), 4.27 (s, 2H), 2.15 (s, 3H), 1.14 (s, 9H) , note: the carboxylic acid and amine protons were

not observed; 13C-NMR (101 MHz, CD3OD) (mixture of rotamers) δ 171.5 (C), 171.1 (C), 167.1 (C), 166.7

(CH), 166.1 (CH), 146.8 (C), 146.7 (C), 139.9 (C), 139.7 (C), 137.7 (C), 137.2 (C), 137.0 (C), 136.6 (C), 133.1 (CH), 130.9 (C), 130.7 (C), 130.31 (CH), 130.30 (CH), 129.1 (CH), 127.8 (CH), 127.6 (CH), 127.2 (C), 127.1 (CH), 126.3 (C), 125.8 (C), 125.7 (C), 123.3 (CH), 122.9 (CH), 121.9 (CH), 121.8 (CH), 113.02 (C), 112.99 (CH), 112.9 (CH), 112.82 (CH), 112.75 (CH), 112.2 (C), 108.9 (C), 108.8 (C), 58.3 (CH), 54.6 (CH), 52.4 (C), 52.2 (C), 51.1 (CH2), 48.4 (CH2), 48.1 (CH2), 48.0 (CH2), 28.9 (3xCH3), 28.8 (3xCH3), 17.6 (2xCH3). 3-(1-(N-(4-(((4-Bromonaphthalen-1-yl)amino)methyl)benzyl)formamido)-2-(tert-butylamino)-2-oxoethyl)-6-chloro-1H-indole-2-carboxylic acid (22)

Following the general procedure, starting from 1-amino-4-chloronaphthalene as amine (40.1 mg, 0.181 mmol), compound 22 was obtained as a white solid (24.4 mg, 0.0361 mmol, 21% yield) after purification by prep-HPLC (Phenomenex Kinetex EVO C18 250x10 mm, H2O/Acetonitrile (0.10% formic

acid): 0.9/0.1 --> 0.1/0.9). M.p. >168 °C (degradation); HRMS (ESI+): C34H33BrClN4O4 [M + H]+, 675.137 (calc.), 675.136

(found). 1H-NMR (400 MHz, CD3OD, major rotamer, 65%) δ

8.35 (s, 1H), 8.13 – 8.08 (m, 2H), 7.73 (d, 1H, J = 8.8 Hz), 7.58 – 7.53 (m, 1H), 7.52 – 7.43 (m, 2H), 7.31 (d, 1H, J = 1.9 Hz), 7.14 (d, 2H, J = 7.4 Hz), 7.06 (dd, 1H, J = 5.2, 1.8 Hz), 6.91 (d, 2H, J = 7.6 Hz), 6.34 – 6.27 (m, 1H), 6.23 (s, 1H), 5.09 (d, 1H, J = 14.9 Hz), 4.40 (s, 2H), 4.36 – 4.23 (m, 1H), 1.13 (s, 9H), note: the carboxylic acid and amine protons were not observed;

13C-NMR (151 MHz, CD3OD) (mixture of rotamers) δ 171.2 (C), 170.9 (C), 166.7 (CH), 166.1 (CH), 145.34

(C), 145.30 (C), 139.6 (C), 139.4 (C), 137.7 (C), 137.61 (C), 137.57 (C), 136.8 (C), 133.44 (C), 133.41 (C), 131.63 (C), 131.57 (C), 131.55 (CH), 131.53 (CH), 129.0 (2xCH), 128.2 (2xCH), 128.00 (CH), 127.97 (CH), 127.8 (2xCH), 127.7 (CH), 127.1 (CH), 126.22 (C), 126.19 (CH), 126.1 (CH), 123.5 (CH), 123.2 (CH), 122.7 (CH), 122.6 (CH), 122.4 (CH), 122.2 (CH), 113.2 (CH), 113.1 (CH), 109.6 (C), 109.4 (C), 106.13 (CH), 106.10 (CH), 58.3 (CH), 54.4 (CH), 52.5 (C), 52.4 (C), 51.2 (CH2), 48.3 (CH2), 48.2 (CH2), 48.1 (CH2), 28.9 (3xCH3), 28.7 (3xCH3).

(19)

3-[1-[{4-[(Benzylamino)methyl]benzyl}(formyl)amino]-2-(tert-butylamino)-2-oxoethyl]-6-chloro-1H-indole-2-carboxylic acid (23)

Following the general procedure, starting from benzylamine (10.7 mg, 0.0999 mmol) as amine, compound 23 was obtained as a white solid (17 mg, 31% yield). M.p.241–243 °C; HRMS (ESI+) for C31H34ClN4O4 [M + H]+: 561.226 (calc.), 561.226 (found). 1H-NMR (400 MHz, DMSO-d6, major rotamer, 82%) δ 11.51 (br.

s, 1H), 8.27 (s, 1H), 7.65 (d, 1H, J = 8.7 Hz), 7.58 (br. s, 1H), 7.41 (d, 2H, J = 7.1 Hz), 7.38 – 7.21 (m, 4H), 7.13 (d, 2H, J = 8.0 Hz), 7.02 – 6.98 (m, 1H), 6.93 (d, 2H, J = 8.0 Hz), 6.35 (s, 1H), 4.82 (d, 1H, J = 15.6 Hz), 4.35 (d, 1H, J = 15.6 Hz), 3.85 (s, 2H), 3.80 (s, 2H), 1.13 (s, 9H), note: of the exchangeable protons only two broad singlets were observed; 13C-NMR (101 MHz, DMSO-d6) (mixture

of rotamers) δ 169.3, 163.78, 163.76, 137.3, 135.5, 135.1, 129.0, 128.8, 128.3, 128.8, 128.2, 127.6, 127.2, 127.1, 125.8, 125.1, 121.7, 119.8, 111.7, 56.7, 50.7, 50.5, 50.4, 46.6, 28.5, 28.3.

4.5 Acknowledgments

Funding was granted to A.K.H.H. by the Netherlands Organization for Scientific Research (NWO-CW, VIDI grant 723.014.008) and by the Helmholtz-Association’s Initiative and Networking Fund; to A.K.H.H. and A.D. by the COFUND ALERT (grant agreement No. 665250); to T.A.H. by Grant UMO-2014/12/W/NZ1/00457 from the National Science Centre, Poland; and to A.D. by the National Institute of Health (NIH) (2R01GM097082-05), the European Lead Factory (IMI) under grant agreement number 115489, the Qatar National Research Foundation (NPRP6-065-3-012). Moreover, funding was received through ITN “Accelerated early stage drug discovery” (AEGIS, grant agreement No 675555), Hartstichting (ESCAPE-HF, 2018B012) and KWF Kankerbestrijding grant (grant agreement No. 10504).

4.6 Contributions from co-authors

M.Y. Unver designed the project. C.G. Neochoritis provided helpful suggestions for the Ugi reaction. R. van der Vlag, M.Y. Unver and T. Felicetti performed KTGS experiments. T. Felicetti synthesized 2, 17–19 and 23. Enzyme-inhibition studies were carried out by A.

Twarda-Clapa and B. Labuzek. 1H-15N HSQC analysis was performed by B. Musielak. Part of

(20)

4.7 References

(1) Jaegle, M.; Wong, E. L.; Tauber, C.; Nawrotzky, E.; Arkona, C.; Rademann, J. Angew. Chemie Int. Ed.

2017, 56 (26), 7358–7378.

(2) Bosc, D.; Jakhlal, J.; Deprez, B.; Deprez-Poulain, R. Future Med. Chem. 2016, 8 (4), 381–404. (3) Unver, M. Y.; Gierse, R. M.; Ritchie, H.; Hirsch, A. K. H. J. Med. Chem. 2018, 61 (21), 9395–9409. (4) Mondal, M.; Hirsch, A. K. H. Chem. Soc. Rev. 2015, 44 (8), 2455–2488.

(5) Frei, P.; Hevey, R.; Ernst, B. Chem. – A Eur. J. 2019, 25 (1), 60–73. (6) Nguyen, R.; Huc, I. Angew. Chemie Int. Ed. 2001, 40 (9), 1774–1776.

(7) Maki, T.; Kawamura, A.; Kato, N.; Ohkanda, J. Mol. BioSyst. 2013, 9 (5), 940–943.

(8) Asaba, T.; Suzuki, T.; Ueda, R.; Tsumoto, H.; Nakagawa, H.; Miyata, N. J. Am. Chem. Soc. 2009, 131 (20), 6989–6996.

(9) Weber, L. Drug Discov. Today Technol. 2004, 1 (3), 261–267.

(10) Oueis, E.; Nachon, F.; Sabot, C.; Renard, P.-Y. Chem. Commun. 2014, 50 (16), 2043–2045. (11) Kulkarni, S. S.; Hu, X.; Doi, K.; Wang, H.-G.; Manetsch, R. ACS Chem. Biol. 2011, 6 (7), 724–732. (12) Greasley, S. E.; Marsilje, T. H.; Cai, H.; Baker, S.; Benkovic, S. J.; Boger, D. L.; Wilson, I. A. Biochemistry

2001, 40 (45), 13538–13547.

(13) Inglese, J.; Benkovic, S. J. Tetrahedron 1991, 47 (14–15), 2351–2364.

(14) Lewis, W. G.; Green, L. G.; Grynszpan, F.; Radić, Z.; Carlier, P. R.; Taylor, P.; Finn, M. G.; Sharpless, K. B. Angew. Chemie Int. Ed. 2002, 41 (6), 1053–1057.

(15) Manetsch, R.; Krasiński, A.; Radić, Z.; Raushel, J.; Taylor, P.; Sharpless, K. B.; Kolb, H. C. J. Am. Chem. Soc. 2004, 126 (40), 12809–12818.

(16) Krasiński, A.; Radić, Z.; Manetsch, R.; Raushel, J.; Taylor, P.; Sharpless, K. B.; Kolb, H. C. J. Am. Chem. Soc. 2005, 127 (18), 6686–6692.

(17) Hu, X.; Sun, J.; Wang, H.-G.; Manetsch, R. J. Am. Chem. Soc. 2008, 130 (42), 13820–13821.

(18) Gelin, M.; Poncet-Montange, G.; Assairi, L.; Morellato, L.; Huteau, V.; Dugué, L.; Dussurget, O.; Pochet, S.; Labesse, G. Structure 2012, 20 (6), 1107–1117.

(19) Chase, J. F. A.; Tubbs, P. K. Biochem. J. 1969, 111 (2), 225–235. (20) Dömling, A. Curr. Opin. Chem. Biol. 2008, 12 (3), 281–291.

(21) Vassilev, L. T.; Vu, B. T.; Craves, B.; Carvajal, D.; Podlaski, F.; Filipovic, Z.; Kong, N.; Kammlott, U.; Lukacs, C.; Klein, C.; Fotouhi, N.; Liu, E. A. Science 2004, 303 (5659), 844–848.

(22) Brown, C. J.; Lain, S.; Verma, C. S.; Fersht, A. R.; Lane, D. P. Nat. Rev. Cancer 2009, 9 (12), 862–873. (23) Popowicz, G. M.; Dömling, A.; Holak, T. A. Angew. Chemie Int. Ed. 2011, 50 (12), 2680–2688. (24) Zhao, Y.; Aguilar, A.; Bernard, D.; Wang, S. J. Med. Chem. 2015, 58 (3), 1038–1052. (25) Khoury, K.; Popowicz, G. M.; Holak, T. A.; Dömling, A. Medchemcomm 2011, 2 (4), 246–260. (26) Nguyen, D.; Liao, W.; Zeng, S. X.; Lu, H. Pharmacol. Ther. 2017, 178, 92–108.

(27) Bista, M.; Wolf, S.; Khoury, K.; Kowalska, K.; Huang, Y.; Wrona, E.; Arciniega, M.; Popowicz, G. M.; Holak, T. A.; Dömling, A. Structure 2013, 21 (12), 2143–2151.

(28) Neochoritis, C. G.; Atmaj, J.; Twarda-Clapa, A.; Surmiak, E.; Skalniak, L.; Köhler, L.-M.; Muszak, D.; Kurpiewska, K.; Kalinowska-Tłuścik, J.; Beck, B.; Holak, T. A.; Dömling, A. Eur. J. Med. Chem. 2019, 182, 111588.

(29) Huc, I.; Lehn, J.-M. Proc. Natl. Acad. Sci. 1997, 94 (6), 2106–2110.

(30) Hochgürtel, M.; Kroth, H.; Piecha, D.; Hofmann, M. W.; Nicolau, C.; Krause, S.; Schaaf, O.; Sonnenmoser, G.; Eliseev, A. V. Proc. Natl. Acad. Sci. 2002, 99 (6), 3382–3387.

(31) Valade, A.; Urban, D.; Beau, J.-M. J. Comb. Chem. 2007, 9 (1), 1–4.

(32) Hochgürtel, M.; Biesinger, R.; Kroth, H.; Piecha, D.; Hofmann, M. W.; Krause, S.; Schaaf, O.; Nicolau, C.; Eliseev, A. V. J. Med. Chem. 2003, 46 (3), 356–358.

(21)

(33) Schmidt, M. F.; Isidro-Llobet, A.; Lisurek, M.; El-Dahshan, A.; Tan, J.; Hilgenfeld, R.; Rademann, J. Angew. Chemie Int. Ed. 2008, 47 (17), 3275–3278.

(34) Zameo, S.; Vauzeilles, B.; Beau, J.-M. Angew. Chemie Int. Ed. 2005, 44 (6), 965–969. (35) Zameo, S.; Vauzeilles, B.; Beau, J.-M. European J. Org. Chem. 2006, 2006 (24), 5441–5444. (36) Nasr, G.; Petit, E.; Supuran, C. T.; Winum, J.-Y.; Barboiu, M. Bioorg. Med. Chem. Lett. 2009, 19 (21),

6014–6017.

(37) Nasr, G.; Petit, E.; Vullo, D.; Winum, J.-Y.; Supuran, C. T.; Barboiu, M. J. Med. Chem. 2009, 52 (15), 4853–4859.

(38) Fang, Z.; He, W.; Li, X.; Li, Z.; Chen, B.; Ouyang, P.; Guo, K. Bioorg. Med. Chem. Lett. 2013, 23 (18), 5174–5177.

(39) Valade, A.; Urban, D.; Beau, J.-M. ChemBioChem 2006, 7 (7), 1023–1027.

(40) SeeSAR version 9.2; BioSolveIT GmbH; Sankt Augustin (Germany). Sankt Augustin, Germany 2019. (41) Popowicz, G. M.; Czarna, A.; Wolf, S.; Wang, K.; Wang, W.; Dömling, A.; Holak, T. A. Cell Cycle 2010, 9

(6), 1104–1111.

(42) Zheng, J.; Li, Y.; Sun, Y.; Yang, Y.; Ding, Y.; Lin, Y.; Yang, W. ACS Appl. Mater. Interfaces 2015, 7 (13), 7241–7250.

(43) Godoy-Alcántar, C.; Yatsimirsky, A. K.; Lehn, J.-M. J. Phys. Org. Chem. 2005, 18 (10), 979–985. (44) Shangary, S.; Qin, D.; McEachern, D.; Liu, M.; Miller, R. S.; Qiu, S.; Nikolovska-Coleska, Z.; Ding, K.;

Wang, G.; Chen, J.; Bernard, D.; Zhang, J.; Lu, Y.; Gu, Q.; Shah, R. B.; Pienta, K. J.; Ling, X.; Kang, S.; Guo, M.; Sun, Y.; Yang, D.; Wang, S. Proc. Natl. Acad. Sci. 2008, 105 (10), 3933–3938.

(45) Czarna, A.; Popowicz, G. M.; Pecak, A.; Wolf, S.; Dubin, G.; Holak, T. A. Cell Cycle 2009, 8 (8), 1176– 1184.

(46) Gladysz, R.; Vrijdag, J.; Van Rompaey, D.; Lambeir, A.-M.; Augustyns, K.; De Winter, H.; Van der Veken, P. Chem. – A Eur. J. 2019, chem.201901917.

(47) Li, Y.; Kang, C. Molecules 2017, 22 (9), 1399–1419.

(48) Powers, R. Expert Opin. Drug Discov. 2009, 4 (10), 1077–1098. (49) Barile, E.; Pellecchia, M. Chem. Rev. 2014, 114 (9), 4749–4763.

(50) D’Silva, L.; Ozdowy, P.; Krajewski, M.; Rothweiler, U.; Singh, M.; Holak, T. A. J. Am. Chem. Soc. 2005, 127 (38), 13220–13226.

(51) Kussie, P. H.; Gorina, S.; Marechal, V.; Elenbaas, B.; Moreau, J.; Levine, A. J.; Pavletich, N. P. Science

1996, 274 (5289), 948–953.

(52) X. Huang. J. Biomol. Screen. 2003, 8 (1), 34–38.

(53) Mori, S.; Abeygunawardana, C.; Johnson, M. O.; Vanzijl, P. C. M. J. Magn. Reson., Ser. B 1995, 108 (1), 94–98.

(54) Stoll, R.; Renner, C.; Hansen, S.; Palme, S.; Klein, C.; Belling, A.; Zeslawski, W.; Kamionka, M.; Rehm, T.; Mühlhahn, P.; Schumacher, R.; Hesse, F.; Kaluza, B.; Voelter, W.; Engh, R. A.; Holak, T. A. Biochemistry

(22)

4.8 Supporting figures

MDM2 MDMX

Ki = 0.40 ± 0.05 µM Ki = 4.18 ± 0.27 µM

Figure S1. Inhibitory activity of compound 2 towards MDM2/MDMX.

MDM2 MDMX

Ki = 0.76 ± 0.08 µM Ki = 12.16 ± 2.15 µM

Figure S2. Inhibitory activity of compound 17 towards MDM2/MDMX.

MDM2 MDMX

Ki = 0.63 ± 0.07 µM Ki = 4.56 ± 0.49 µM

Figure S3. Inhibitory activity of compound 18 towards MDM2/MDMX. 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F rac tio n o f b o u n d r ep o rt er p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM)

(23)

MDM2 MDMX

Ki = 0.75 ± 0.08 µM Ki = 7.08 ± 0.52 µM

Figure S4. Inhibitory activity of compound 19 towards MDM2/MDMX.

MDM2 MDMX

Ki = 0.47 ± 0.04 µM Ki = 3.73 ± 0.30 µM

Figure S5. Inhibitory activity of compound 20 towards MDM2/MDMX.

MDM2 MDMX

Ki = 0.49 ± 0.04 µM Ki = 4.61 ± 0.39 µM

Figure S6. Inhibitory activity of compound 21 towards MDM2/MDMX. 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F rac tio n o f b o u n d r ep o rt er p e p tid e Inhibitor concentration (μM)

(24)

MDM2 MDMX

Ki = 0.095 ± 0.010 µM Ki = 3.95 ± 0.46 µM

Figure S7. Inhibitory activity of compound 22 towards MDM2/MDMX.

MDM2 MDMX

Ki = 3.18 ± 0.18 µM Not active

(64% of reporter peptide bound at 50 μM)

Figure S8. Inhibitory activity of compound 23 towards MDM2/MDMX.

MDM2 MDMX

Ki = 0.55 ± 0.05 µM Ki = 11.4 ± 1.6 µM

Figure S9. Inhibitory activity of compound 3 towards MDM2/MDMX. -0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d re p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F rac tio n o f b o u n d r ep o rt er p e p tid e Inhibitor concentration (μM) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 0.0001 0.001 0.01 0.1 1 10 100 F ra c tio n o f b o u n d r e p o rt e r p e p tid e Inhibitor concentration (μM)

(25)

Figure S10. UV-vis absorption spectrum of 3 (red), 22 (blue) and 24 (green) in MilliQ/ACN 50:50 (with 0.1% FA).

For comparison, data was normalized at 305 nm (λmax of 3).

Figure S11. Determining the detection limit of the UPLC-MS system of aldehyde 3 at 254 nm (close-up). From top

to bottom: 0.30, 0.10, 0.033, 0.011, 0.0037, 0.0012, 0.00041 mg/mL and blank. The peak of 3 at 0.0012 mg/mL (2.6 μM,purple) can be clearly distinguished from the blank, while the peak at 0.00041 mg/mL (cyan) cannot be reliably identified. RT = Retention Time, MA = Measured Area.

0 0.5 1 1.5 2 2.5 230 330 430 530 630 A bs orbance Wavelength (nm) 3 22 24 0 0.25 0.5 0.75 1 1.25 1.5 250 300 350 400 A bs orbance Wavelength (nm) 3 22 24

(26)

Figure S12. Determining the detection limit of the UPLC-MS system of aldehyde 3 at 305 nm (close-up). From top

to bottom: 0.30, 0.10, 0.033, 0.011, 0.0037, 0.0012, 0.00041 mg/mL and blank. The peak of 3 at 0.00041 mg/mL (0.88 μM,cyan) can be clearly distinguished from the blank. RT = Retention Time, MA = Measured Area.

Figure S13. Plot of the integrated area versus concentration of 3 in the UPLC-MS vial at wavelength 254 nm (blue)

and 305 nm (red). Linear fits are shown using dotted lines.Solving both equations for a value of three times the measured area of the blank gives an estimated detection limit of 1.56 μM and 1.07 μM for 254 and 305 nm, respectively. Based on this calculation, comparing the traces visually would result in a higher detection limit for 254 nm and a lower detection limit for 305 nm.

Table S1. Overview retention times of all compounds. Compound Ret. time

(min)

Compound Ret. time (min) A m ine s tar ti ng m at er ial s 4 5.54 R edu ct iv e a m ina ti o n pr o du ct s 2 8.72 5 6.50 18 8.79 6 5.45 19 8.60 7 6.16 17 7.44 8 7.12 20 8.92 9 7.46 21 8.99 10 8.18 22 9.21 11 1.87 23 7.28 Aldehyde 3 8.00 (±)-Nutlin-3 7.37 y = 2838x - 1088.8 R² = 1 y = 2621.1x - 1457.9 R² = 1 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 100 200 300 400 500 600 700 M e asur e d A re a (x 10 6) Concentration of 3 in vial (µM) 254 nm 305 nm

(27)

Referenties

GERELATEERDE DOCUMENTEN

Then I would like to thank specifically: Alwin, bedankt voor de vele dingen die je hebt georganiseerd, daarnaast was de EFMC in Ljubljana super!. Jonas, making

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.. Downloaded

The research presented in this thesis was performed in the research group of Drug Design of Groningen Research Institute of Pharmacy at the University of Groningen, the

Other useful techniques include capillary electrophoresis (CE), biolayer interferometry (BLI) and microscale thermophoresis (MST). In our lab,.. the fragment screening strategy

A second DSF- based screen (TSA buffer screen), similar in approach to that used in other laboratories, 17 was used to identify an dialysis buffer for the removal

tested the synthesis of tetrazole 2a in various solvents such as toluene, DMF, DMSO and acetonitrile at various reaction time (6 to 24h) and temperature from 55 to 160

Surprisingly, when we used 2M methanolic hydrochloric acid as solvent, the yield of 6-membered (6a) product significantly increased to 64% and minor 7-membered (7a) product (6%

The solvents were removed under vacuum and the crude product was purified by flash column chromatography to give pure product (8).. The resulting mixture was stirred