Quest for Novel Chemical Entities through Incorporation of Silicon in

Oct 17, 2017 - Jeffrey Aubé in the Eshelman School of Pharmacy, University of North Carolina at Chapel Hill, U.S.. Biography. D. Srinivasa Reddy rece...
10 downloads 30 Views 966KB Size
Subscriber access provided by University of Florida | Smathers Libraries

Perspective

Quest for Novel Chemical Entities through Incorporation of Silicon in Drug Scaffolds Remya Ramesh, and D. Srinivasa Reddy J. Med. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jmedchem.7b00718 • Publication Date (Web): 17 Oct 2017 Downloaded from http://pubs.acs.org on October 17, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Medicinal Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Quest for Novel Chemical Entities through Incorporation of Silicon in Drug Scaffolds Remya Ramesh, a,b and D. Srinivasa Reddy a,b * a CSIR-National Chemical Laboratory, Dr. Homi Bhabha Road, Pune, 411008, India b Academy of Scientific and Innovative Research (AcSIR), New Delhi, 110 025, India E-mail: [email protected]

Abstract In order to optimize a lead molecule for further development, bioisosteric replacements are generally adopted as one of the strategies. Silicon appears to be the right choice as a carbon isostere because of the similarity in chemical properties. Silicon can be strategically introduced in a molecule to modulate its drug like properties, providing medicinal chemists with an unconventional strategy for replacing a carbon atom. Silicon can also be introduced to replace other heteroatoms, and can act as a surrogate of functional groups such as olefin and amide as well. The present perspective focuses on the opportunities that silicon incorporation offers in drug discovery, with an emphasis on case studies where introduction of silicon has created a benefit over its carbon analog. We have tried to highlight all the recent developments in the field and briefly discuss the challenges associated with them.

1. About silicon Silicon is an element (atomic no. 14) that belongs to the third row of the periodic table, positioned just below carbon. It is the second most abundant element in

ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the earth’s crust, the first being oxygen. Although, both carbon and silicon exhibit a valency of four and form tetrahedral compounds, there are certain differences in their properties. Unlike silicon, carbon is the element of life displaying a very rich and unique chemistry, with a branch of organic chemistry entirely dedicated to carbon compounds. The ability of carbon to form diverse compounds is due to its unique property of self-linking (catenation). Silicon does not form stable π bonds and Si-Si σ bond (230 kJmol-1) is weaker than the Si-O bond (368 kJmol-1).1,2 Owing to the higher stability of Si-O bond, silicon occurs in nature as silicates and silica.

2. Silicon as bioisostere of carbon Bioisosteres are groups with similar physicochemical properties, and display similar biological activity in a broad sense. Bioisosteric replacements are commonly practised in medicinal chemistry, to modulate the properties of a molecule and for creating intellectual property (IP) space.

3-5

Due to the similarity of silicon to carbon,

over the years there has been growing interest in biological evaluation of organosilanes over the years. Silicon can be considered a bioisostere of carbon and offers an innovative avenue in drug discovery.6-13 The substitution of carbon with silicon in biologically active compounds is also known as the “silicon switch”. The effect of silicon substitution has also been explored in agrochemicals14,15 and odorants as well. The relationship between structure and odor with respect to silicon substitution is well studied.16-18 Tacke

19-25

and Wannagat26 are considered to

be pioneers in this field.

2 ACS Paragon Plus Environment

Page 2 of 71

Page 3 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

3. Silicon in medicinal chemistry The similarity between carbon and silicon makes silicon an ideal choice as a carbon bioisostere in drug discovery. Unlike elements such as tin, silicon does not have any inherent toxicity.7,27 The introduction of silicon also provides freedom to operate as most patents do not cover silicon as part of their claims. Some of the differences between carbon and silicon can be efficiently utilized in drug design and lead optimization. Such differences include:  Larger carbon-silicon bond length: The bond length of C-C is 1.54 Å and that of C-Si is 1.87 Å i.e. C-Si bond is almost 20% longer than a C-C bond.1 This can lead to changes in shape and conformation of the molecule and can also alter the way in which the molecule interacts with a receptor.  Increase in lipophilicity: Silicon analogs are more lipophilic than their carbon counterparts.28 Therefore substitution of carbon by silicon can lead to improved cell penetration and can ultimately result in an improvement in potency. The increase in lipophilicity can also be beneficial in the design of drugs targeting bacteria or the central nervous system.29 In some cases, it can also be detrimental, due to liabilities such as poor solubility and metabolic clearance. Therefore, one needs to balance the concept, depending on the drug target and its location in the body.  Difference in bonding preferences: Silicon prefers higher coordination numbers when compared to carbon.1,2 Silanediol does not undergo dehydration to form silanone whereas geminal diols are not stable in the case of carbon. Thus, silicon offers a novel chemical space and provides access to compounds

3 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

for which corresponding carbon analogs are not available. This has been exploited in designing transition state mimics. Silicon can also extend its coordination number to six, which is useful in the formation of coordination complexes.  Higher electropositivity of silicon: Silicon is more electropositive than carbon which leads to a difference in bond polarity. According to the Pauling scale, silicon has an electronegativity of 1.74, whereas carbon has a value of 2.50. This can cause subtle differences in the behaviour of silanols and carbinols.8 The utility of organosilanes in drug discovery has been reviewed in the past, as well.6-13 In 2013, a very useful perspective on medicinal applications of organosilicon compounds by Franz and Wilson was published in this journal and attracted the attention of many research groups.8 We are writing this perspective considering the importance and interest in the medicinal chemistry fraternity. A recent review by Hashimoto13 and book chapters by Tacke,12a and Sieburth12b are also worth mentioning here. Selected examples from literature are discussed in this perspective, highlighting the potential of silicon in medicinal chemistry. The focus is on case studies where the introduction of silicon has created a benefit over its analog.

4. Modification of metabolic pathways Any drug, after reaching the body, undergoes enzyme-catalyzed biochemical modification called metabolism.30 During this process, administered drugs get converted to more polar compounds and are excreted from the body. Due to the difference in bonding properties, silicon analogs can have a different metabolic fate compared to the parent carbon compound. The change in chemical reactivity could be

4 ACS Paragon Plus Environment

Page 4 of 71

Page 5 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

utilized in drug design to alter problematic metabolic pathways. This was nicely illustrated by Tacke and co-workers in the case of antipsychotic drug haloperidol (1). 31-34

Haloperidol causes a neurotoxic side effect that is attributed to its metabolite,

pyridinium compound 3 (figure 1). Compound 3 is structurally similar to 1-methyl-4phenylpyridinium (MMP+) that kills dopamine producing neurons of the brain and can induce Parkinson disease.35-38

Figure 1. Metabolism of haloperidol The corresponding silicon analog (4) follows a different metabolic pathway devoid of the dehydrated compound 5 and the toxic pyridinium metabolite.32,39 Ndealkylation and hydroxylative ring opening were found to be the major pathways leading to metabolites 6 − 8 (figure 2). The inherent stability of the Si-O bond and the instability of the Si=C makes the molecule to adopt this alternate pathway.

5 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. Silahaloperidol and major metabolites Similar carbon/silicon switching was also done in another related compound, trifluperidol (9) and analogous results were obtained (figure 3).32,40 In the case of silicon analog (10), hydroxylation of the piperidine ring leading to opening of the ring was found to be the major metabolic pathway. Reports indicate that there is significant difference in conformational behavior of the silicon analog. Due to increased C-Si bond length, silapiperidine chair form is more flattened compared to the piperidine chair conformer.31

Figure 3. Trifluperidol and its silicon analog Tacke and co-workers incorporated silicon in loperamide (11), a commonly prescribed anti-diarrheal drug.41 The structure of loperamide is similar to haloperidol, although they act on completely different targets. Compounds 11 and 12 showed

6 ACS Paragon Plus Environment

Page 6 of 71

Page 7 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

similar potency for the µ1 opioid receptor and had similar physicochemical properties.

Figure 4. Metabolism of loperamide and its silicon analog The metabolic pathways of both compounds (11 and 12) were found to have marked differences (figure 4). The major metabolites 13, 14 and 15 of loperamide were stable in circulation. Similar metabolites 16 and 17 were observed in the case of the silicon analog, but the analog of 15 was not detected. Instead, another metabolite 18 resulting from the α-oxidation and ring opening was found. The silanediol 18 showed high clearance and was rapidly eliminated from the circulation thus reducing the burden of metabolites.41 These reports show that silicon incorporation in drugs can lead to alternate metabolic pathways which can be advantageous.

7 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

5. Novel chemical space provided by silicon 5.1. Silanols as transition state mimics Proteases are enzymes that hydrolyze the amide bond of peptides and are involved in various physiological reactions. Inhibition of specific proteases has been found to be useful in controlling several diseases and hence proteases have become an interesting pharmaceutical target.42 Proteases stabilize the tetrahedral diol formed by the addition of water and this diol then collapses to individual peptides. Any compound that can mimic the enzyme-substrate interactions present in the tetrahedral intermediate, and does not undergo hydrolysis, could inhibit the enzyme (see structures 19 − 22).43

Figure 5. Design of protease inhibitors mimicking the transition state In fact, Sieburth pioneered the use of stable silanediol of the general formula 22 to mimic the gem diol transition state involved in protease hydrolysis.43-51 Silicon favors sp3 hybridization over sp2 and hence geminal silanediols are stable and do not undergo elimination to form the silanone (Si=O). Silanediol has gained attraction in medicinal chemistry as a bioisostere of carbonyl hydrate and some of the inhibitors

8 ACS Paragon Plus Environment

Page 8 of 71

Page 9 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

known in the literature are shown in figure 6 (compounds 23 – 26).43,46,47,50 The increased hydrogen bonding ability of silanols is an advantage which allows the silanediols to interact efficiently with the active site of enzyme. Simple silanediols undergo polymerization, but sterically hindered diols are found to be stable. Silanediol 26 is inferior to the carbinol 27, which had potency similar to the marketed drug indinavir (28). This decreased activity may be due to the larger size of silicon.49

Figure 6. Silanediols as protease inhibitors Olsen and co-workers synthesized analogs of vorinostat (29), a histone deacetylase (HDAC) inhibitor and studied their biological activities.52 Most of the HDAC’s require Zn2+ as a cofactor and are called Zn2+ dependent histone deacetylases. Silanediol 31 was synthesized owing to its ability to coordinate zinc with affinity similar to that of hydroxamic acids. The activities of all the compounds were tested against 11 HDAC isoforms (activity against two isoforms shown in figure

9 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

7). The trifluoromethyl compound 30 and silanediol 31 showed moderate activities and may be considered as promising leads.

Figure 7. Silanediol analog of vorinostat The biological activity of silanediols prompted organic chemists to develop efficient synthetic routes to access them.53 Studies were also carried out to understand the mechanism of inhibition as this knowledge is essential for taking the compounds forward.54 Since protease inhibitors represent an important class of therapeutic agents, silanediols have great potential and further research in this area can lead to more potent and selective inhibitors. Pietschnig and co-workers extended the concept of silanediols to silanetriols, in order to mimic the transition state of ester hydrolysis.55 Silanetriols with bulky substituents (32−34) were synthesized so as to prevent self-polymerization and were evaluated for their inhibition of the enzyme acetylcholinesterase (AChE). Compound 33 showed the highest inhibition rate and the IC50 was found to be 121 µM. This compound showed reversible inhibition of the enzyme which is a prerequisite for therapeutic agents. Although it is debatable, many researchers believe that irreversible inhibitors can lead to damage of the enzyme.56 Hence, reversible silanetriol inhibitors may find use in developing future therapeutic agents.

10 ACS Paragon Plus Environment

Page 10 of 71

Page 11 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Figure 8. Silanetriols as AChE inhibitors 5.2. Silicon containing complexes Silicon can exhibit coordination numbers higher than 4 due to the presence of empty 3d orbitals. The silicon complex Pc-4 (36), built on the framework of phthalocyanine (35), went to clinical trials for use in photodynamic therapy (PDT).5759

In PDT, the photosensitizer, when excited by light of particular wavelength,

interacts with molecular oxygen and generates reactive oxygen species.60-63 This reactive oxygen species (ROS) then kills the cells in the illuminated area by inducing apoptosis. The photosensitizer is cleared from healthy tissues more rapidly, which leads to their accumulation in tumour cells.62 The photophysical properties of phthalocyanines can be modified by varying the ring substituents, axial ligands, and the central metal atom. The silicon phthalocyanine 36 has the right balance between lipophilicity and hydrophilicity. After cellular uptake, Pc-4 accumulates in the mitochondria

and

causes

photocytotoxicity.

Recently,

another

modified

phthalocyanine Pc-227 (37) was reported which upon photolysis gives 36 (figure 9).64

11 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 9. Silicon phthalocyanines Meggers and co-workers reported the biological activity of hydrolytically stable silicon containing octahedral polypyridyl complexes 38 − 41 (figure 10).65 Studies showed that the complexes 40 and 41 intercalate between the DNA base pairs. DNA intercalators have a variety of therapeutic applications mainly in the area of cancer therapy.66 A follow-up of this work, led to the synthesis of several hexacoordinate silicon complexes and NMR spectroscopy studies proved that they bind to G-quadruplex DNA.67 Because of non-toxicity and natural abundance, use of silicon is advantageous over the commonly used transition metals in designing complexes with pharmaceutical applications.

12 ACS Paragon Plus Environment

Page 12 of 71

Page 13 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Figure 10. Hypervalent silicon complexes as DNA intercalators Silatranes (42; Y= O) are compounds having pentavalent silicon with a transannular dative bond between N and Si.68,

69

Some of the compounds from this

class are reported to have interesting biological activities (antibacterial, antifungal, anticancer, antiviral, anti-inflammatory etc.).70-74 Kurup and co-workers reported that 1-ethoxysilatrane (43) inhibits cholesterol biosynthesis and rats treated with the compound showed a decrease in serum cholesterol.73,74 OC2H5 O

R Y Y

Si N

Y

Y = O; Silatrane Y = N; Azasilatrane Y = C; Carbasilatrane Y = S; Thiasilatrane

42

O

Si N 43

Figure 11. Silatranes

13 ACS Paragon Plus Environment

O

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

6. Increase in lipophilicity by introduction of silicon 6.1. Silicon incorporation to increase brain penetration The endothelial cells of the brain capillaries have tight junctions which is called blood brain barrier (BBB). This barrier selectively allows the passage of certain molecules to the brain and prevents others, making drug delivery to the brain a tough task.75,76 It is known that small lipophilic molecules diffuse through this barrier. The incorporation of silicon into a molecule increases its lipophilicity, which can increase the BBB penetration of a drug. Reddy and co-workers designed and synthesized silicon analogs of oxazolidinone antibiotics and evaluated their biological activity (figure 12).77 Linezolid is the first marketed oxazolidinone drug and is used as an antibiotic. The thiomorpholine Sutezolid is in clinical trials for treating tuberculosis. Several compounds were synthesized with silicon replacing the O/S atom of the ring. The brain pharmacokinetics of the most active compounds (44 − 46) revealed that silicon incorporation caused significant increase in brain to plasma ratio. Among them, the compound 46 had 29-fold higher brain/plasma ratio with respect to Linezolid. This increase in CNS exposures is due to the increase in lipophilicity of the compounds which is also evident from the cLogP values.

14 ACS Paragon Plus Environment

Page 14 of 71

Page 15 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Figure 12. Silicon incorporation in linezolid 6.2. Silicon analogs of campothecin The increase in lipophilicity due to silicon incorporation was utilized in the synthesis of stable campothecin analogs. Campothecin (47) is an anticancer natural product that inhibits the enzyme DNA topoisomerase I. The δ-lactone moiety of campothecin is susceptible to hydrolysis in vivo. The lactone form (47), which can distribute into the cell, is considered to be the active form. The carboxylate of 47 binds strongly to human serum albumin which causes an increase in the rate of lactone hydrolysis so as to maintain the equilibrium.78

15 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 71

O

N

N

OH

N O HO

O OH

N

O

H HO

O

O

campothecin (carboxylate form)

campothecin (47)

Si Si

Si HO N

O

N O

karenitecin (48a)

O

N

O

N

N

N HO

HO

O

O HO

O

HO O

DB-67 (48b)

O

DB-91 (48c)

Figure 13. Silicon analogs of campothecin An increase in lipophilicity leads to increased partitioning into red blood cells resulting in less hydrolysis of the lactone. Several silicon analogs were synthesized and two of them, karenitecin (48a) and DB-67(48b) were evaluated in human clinical trials for cancer.78-85 Another closely related compound, DB-91 (48c) penetrates the blood brain barrier and hence was tested for cancers of the central nervous system.85

7. Drug design utilizing the electropositivity of silicon 7.1. N+/ Si exchange Zifrosilone/MDL 73,745 (50) is a silicon compound that went to clinical trials for treating Alzheimer’s disease.86-88 It was proposed that Alzheimer’s patients have a deficiency of the neurotransmitter acetylcholine and so medications that help to increase the concentration of acetylcholine would be beneficial.89 The enzyme acetylcholineesterase (AChE) catalyzes the hydrolysis of acetylcholine and hence inhibitors of this enzyme were tested for Alzheimer’s disease. Trifluoromethyl

16 ACS Paragon Plus Environment

Page 17 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

ketones such as 49 were reported to inhibit AChE.90,91 An ammonium/silicon switching in 49 resulted in lipophilic compound 50, which also showed AChE inhibition (figure 14). The greater electropositivity of silicon creates an ionic charge that interacts with the cation binding site of the target enzyme. This drug design strategy (to create a positive center in a neutral molecule) can be used in medicinal chemistry as an alternate option.

Figure 14. N+/ Si exchange in zifrosilone Tacke and co-workers applied the concept of N+/Si exchange in the muscarinic M2 receptor modulators (figure 15).92,93 W84 (51) inhibits the dissociation of [3H]NMS, an orthosteric radiolabeled ligand of the receptor. The silicon analog 52 was also found to be an allosteric modulator of the receptor, but it exhibited positive allosteric modulation (enhances binding of [3H]NMS to receptor). Here silicon switching has transformed the molecule from allosteric inhibitor to allosteric activator. However, it was found that the neutral compound (2-fold exchange) was inactive.

17 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 71

Figure 15. Silicon analog of W84 7.2. Hydrogen bonding capability of silanols The hydrogen bonding ability and acidity of silanols is higher compared to corresponding carbinols, due to π bonding between p orbitals of oxygen and empty 3d orbitals of silicon.94 This property can be useful in pharmacophores where hydrogen bonding interactions are important. Baney and co-workers showed that silicon incorporation increased the antibacterial activity in a series of hydroxyl compounds 53 – 56 (figure 16).95

Figure 16. Alcohols and their corresponding silanols; X= C/Si Compound

Minimum Lethal Concentration (MLC in %, g/g) E. coli

S. aureus

P. aeruginosa

E. faecalis

53a

13.54

10.61

9.79

13.33

53b

2.36

2.48

2.36

3.15

54a

5.09

4.17

3.96

5.67

54b

1.04

0.80

0.87

1.14

55a

5.23

4.37

3.67

5.63

55b

1.23

1.04

1.00

1.32

56a

0.96

0.78

0.71

0.93

56b

0.27

0.26

0.35

0.42

Table 1. Antibacterial activity of compounds 53 − 56

18 ACS Paragon Plus Environment

Page 19 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

This increase in activity (see table 1) is probably due to increase in hydrogen bonding ability as well as lipophilicity. The authors also proved that the hydrogen bonding ability of silanols is almost twice of that of the corresponding alcohols with the help of IR spectroscopy. Another example where an increase in potency of silanols was reported is discussed in a later section (figure 22). Fujii and co-workers investigated the biological activity of a silanol analog (58) of known nuclear receptor modulator 57.96 The corresponding carbon analog 59 was synthesized for comparison purpose. The perfluoroalcohol groups have higher acidity than the corresponding hydrocarbon alcohols due to the presence of trifluoromethyl groups. The octanol-water partition coefficient of all the three compounds 57 − 59 was determined experimentally. It was found that the silanol 58 had an intermediate lipophilicity between that of 57 and 59. Compound 57 showed significant activity towards multiple nuclear receptors LXRα and β, FXR, PXR and RORγ, whereas alcohol 59 was not active. Silanol 58 exhibited activity towards PXR and RORγ comparable to that of the parent compound 57, whereas it was inactive towards LXRα and β. From molecular docking studies, it was ascertained that the hydroxyl group is not involved in any hydrogen bonding interaction with RORγ receptor. In the case of LXRα, LXRβ and PXR, the hydroxyl group interacts with His407 of the receptor protein. This indicates that both hydrogen bonding ability and lipophilicity has a major role to play in the activity of these compounds. From the biological activity results, it was suggested that the silanol group could be considered as an isostere of perfluoroalcohol group.

19 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 17. Silanol as an isostere of perfluoroalcohol group

8. Taking advantage of the increased bond length of silicon 8.1. Silicon as an isostere of olefinic double bond Combretastatin A-4, a natural product, displays potent anti-cancer activity by inhibiting tubulin polymerization. The compound has a cis-olefin that tends to isomerize in solution and hence there have been attempts to replace this olefin with any stable isostere.97 Nakamura et al. assumed that because of the larger C-Si bond length, a silicon linker can be used as a bioisostere of olefin (figure 18). Computational calculations showed that the distance between the benzene rings is almost the same in combretastatin A-4 (3 Å) and the silicon analog 60 (3.03 Å). In carbon analog 61, this distance was found to be 2.47 Å. Docking studies suggested that the newly designed silicon compound 60 and combretastatin A-4 occupied the colchine-binding site of tubulin. Several silicon analogs were synthesized and evaluated using in vitro assays for anti-cancer activity (human breast cancer cell line). The natural product showed an IC50 of 0.004 µM and the newly designed silicon analog (60) also showed activity (IC50 0.043 µM), but lesser than the natural product. One of the silicon analogs (62) showed an IC50 of 0.007 µM similar to the natural product, whereas the corresponding carbon analog 63 (IC50 0.075µM) was less active. The stability of the compounds in solution was also analyzed which showed that

20 ACS Paragon Plus Environment

Page 20 of 71

Page 21 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

compounds 60, 62 and 63 were stable. It was observed that the concentration of combretastatin decreased to nearly 85% under similar conditions.98

Figure 18. Silicon analogs of combretastatin A-4 The same concept was extended to aliphatic cis olefins and found to be successful.99 The cis-olefin present in PPAR-α agonist oleoylethanolamide (64) was substituted with a silyl linker and the biological activity of analogs 65 – 68 were evaluated. The diethyl derivative 66 showed moderate activity, but was less than that of compound 64. The substituents on silicon and the chain length influenced the activity of the compounds. The fatty acid chain occupies the hydrophobic pocket of the receptor and docking studies indicate that silyl group improves this hydrophobic interaction.

Figure 19. Silicon linker as olefin isostere

21 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

8.2. Silicon as amide surrogate Fujii and co-workers replaced the amide functionality in retinoic acid receptor inverse agonists belonging to the phenanthridinone class (69 and 74) with a silyl group and studied their biological activity.100,101 Docking studies showed that the amide is not involved in hydrogen bonding with the receptor and the binding site is hydrophobic.102 The analogs 70 − 73 showed similar IC50 for the inhibition of RORγ as that of parent compound 69. The activities of 71 (IC50 = 4.2 µM) and 73 (IC50 = 4.4µM) were slightly better than that of compound 69 (IC50 = 4.7µM). However, the inhibition of 75 (IC50 = 18 µM) was found to be 2-fold less compared to the carbon compound 74 (IC50 = 37 µM). Thus, in these examples taking advantage of silicon’s increased bond length to replace a olefin/ amide led to an increase in activity.

Figure 20. Silicon analogs of phenanthridinone derivatives

9. Silicon incorporation to improve the potency Lipophilic silicon analogs display better cell penetration compared to their carbon analogs and can lead to increase in potency in certain drugs. Other properties such as polarization of the bond, increased bond length, conformational changes etc. can also cause an increase in potency.

22 ACS Paragon Plus Environment

Page 22 of 71

Page 23 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Kagechika and co-workers substituted silicon in place of carbon in 4substituted phenols and evaluated their estrogenic activity.103 Out of all tested compounds (76 − 83), the triethyl silicon derivative 81 showed the highest potency. Silicon substitution caused an increase in activity for all the compounds in the series (figure 21). Docking studies showed that the trialkyl group at p position lies in a hydrophobic pocket of the estrogen receptor, suggesting that the improvement in activity may be due to the increase in lipophilicity caused by silicon substitution.

Figure 21. Estrogenic activity of phenols Tacke studied the effect of C/Si bioisosterism in muscarinic antagonists (84 – 89, figure 22).104 As mentioned, muscarinic acetylcholine receptors are activated by the neurotransmitter acetylcholine and antagonists of this receptor are useful in treating Parkinson’s disease.

23 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 71

Figure 22. Silicon switch in muscarinic antagonists (X = C or Si) Silanols showed increased binding affinities for all the four receptor subtypes (M1 − M4) compared to the corresponding carbinols. The highest improvement in potency was found for the pair 88 (table 2). Compound

Si/C affinity ratio

pair

M1

M2

M3

M4

84

1.6

2.0

2.0

1.6

85

1.3

2.0

-

1.3

86

10.0

5.0

7.9

16

87

2.5

1.6

3.2

3.2

88

40

50

32

40

89

6.3

13

10

6.3

Table 2. Binding affinity ratios of silicon to carbon analogs

24 ACS Paragon Plus Environment

Page 25 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

DeGrado and co-workers reported organosilyl amine inhibitors of M2 protein of influenza A virus.105 It is known that lipophilicity can improve the potency of M2 inhibitors because the protein has a hydrophobic binding pocket. The spirane amines 90 − 93 were synthesized and the activities against A/M2-V27A were evaluated (figure 23). Silicon analogs 91 and 93 showed an increase in potency compared to their carbon counterparts 90 and 92. The authors suggest that this is probably due to the lipophilicity as well as larger ring size which provide a better binding of the molecule with the receptor protein.

Figure 23. Antiviral activity of silaspirane amines Tacke and co-workers studied the effect of carbon to silicon switch in the retinoid agonist SR11237 (94).106 Since silicon substitution causes an enlargement in ring size, the indane derivatives 95 and 98 were also made for the sake of comparison. Due to higher C-Si bond length, the ring size of 94 and 98 are expected to be same. The potencies of 94 and 97 were found to be similar with the silicon analog 97 exhibiting a slightly higher activity. However, in the cases of indanes the sila analog 98 was found to be almost 10-fold more potent compared to its corresponding carbon compound 95. Similar enhancement in activity (almost 10-fold) was also observed in the case of another retinoid agonist, tamibarotene (99 vs 96) and AM580 (4-[(5,6,7,8Tetrahydro-5,5,8,8-tetramethyl-2-naphthalenyl)carboxamido]benzoic acid).107

25 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 24. Silicon switching in retinoids Lee and co-workers investigated the anti-allergic activity of silicon substituted oxadiazoles. The silicon containing compound 100 was found to be more potent than the corresponding carbon derivative 101 using RBL-2H3 in vitro model. Compound 100 suppressed the mast cell degranulation in antigen sensitized cells in a dose dependant manner. The compound also effectively reduced anaphylaxis in an in vivo mouse model with an activity comparable to that of anti-histamine drug diphenylhydramine.108

Figure 25. 1,3,4-oxadiazoles with anti-allergic activity Smith and co-workers reported the synthesis and biological activity studies of sila analogs of quinoline/ferroquine derivatives and their metal complexes.109,110 Further extension of this work led to the identification of thiosemicarbazone complexes (103, 104) with antiplasmodial activity.111 IC50 values against the chloroquine sensitive P. falciparum strain NF54 are shown in figure 26. Silicon analog 102b was found to be highly potent compared to the carbon compound 102a.

26 ACS Paragon Plus Environment

Page 26 of 71

Page 27 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

However, in the case of metal complexes, the activities of carbon and silicon analogs (103a vs 103b; 104a vs 104b) were found to be comparable.

Figure 26. Antiplasmodial activity of thiosemicarbazone complexes Reddy and co-workers studied the effect of silicon substitution in morpholine antifungals. Fenpropidin (105) and fenpropimorph (106) are marketed fungicides whereas amorolfine (107) is applied topically for nail infections. Several compounds were synthesized and the sila analog 108 came out as the lead of the series with better MFC (minimum fungicidal concentration) than the parent compounds (105 − 107) against various human pathogenic fungi (table 3). The silicon analog 109 also showed good activity.112

Figure 27. Silicon switching in morpholine antifungals

27 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Compound

Page 28 of 71

Minimum fungicidal concentration (MFC in µg/mL) C. albicans

C. neoformans

C. glabrata

C. tropicalis

A. niger

105

0.5

32.0

16.0

> 256

> 256

106

4.0

64.0

128.0

> 256

> 256

107

8.0

64.0

32.0

> 256

> 256

108

0.5

2.0

2.0

64.0

256

109

8.0

4.0

8.0

64.0

64.0

Table 3. Antifungal activity of compounds 105 – 109 Reddy group repurposed the rimonabant scaffold for tuberculosis and studied the effect of silicon substitution on their potency. Rimonabant is an inverse agonist of the cannabinoid receptor CB1 and was marketed for treating obesity. A systematic study of rimonabant analogs (110 − 113) revealed that silicon substitution led to a dramatic improvement in anti-TB potency of this series (110 vs 112, 113). Two compounds 110 (MIC 0.031 µg/mL) and 111 (MIC 0.39 µg/mL) were identified as leads. This increase in potency may be due to the higher cell wall penetration of lipophilic silicon compounds.113

28 ACS Paragon Plus Environment

Page 29 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Si O N N

N NH

SAR optimization

NH

N N

Cl

Cl

N

Si

N

N N Cl

Cl

Cl Rimonabant CB1 receptor blocker for treating obesity MIC: 25 g/mL (Mtb, H37Rv)

Cl 110 MIC: 0.031 g/mL (Mtb, H37Rv) % plasma stability (human): 100 % metabolic stability in HLM: 20.3

N

Cl

N

N

Si

N Cl

N

Cl 111 MIC: 0.39 g/mL (Mtb, H37Rv) % plasma stability (human): 100 % metabolic stability in HLM: 93.1

N

N

N

N

Cl

N

Cl

Cl

Cl

110 MIC: 0.031 g/mL

112 MIC: 25 g/mL

113 MIC: 1.56 g/mL

Figure 28. Rimonabant analogs towards tuberculosis Padron and co-workers studied the effect of TBS group (a protecting group commonly used in organic chemistry) on the cytotoxic activity of substituted tetrahydropyrans.114 The cLogP of the TBS protected alcohols were found to be higher compared to free alcohols as well as other protecting groups. The growth inhibitory activities of the compounds were checked in two cancerous cell lines, HL60 and MCF7 (figure 29). The derivatives having TBS group at 3rd position showed anti-cancer activity (115 vs 114/ 116/ 117; 119 vs 118; 121 vs 120). This observed activity by the introduction of TBS group may be due to better cell penetration of the lipophilic TBS compounds.

29 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 29. Cytotoxic effects of tetrahydropyran derivatives (NA = not active) Along similar lines, Trindade and co-workers studied the role of TBS group in the antitumor activity of a series of 5-hydroxymethyl furfural derivatives (122 − 125).115 Compounds 124 and 125 were the most potent from this series (figure 30). Wang and co-workers reported that a 5’silyl group is essential for the antiviral activity of triazolyl thymidine analogs (see compounds 126 and 127).116 The improved activity of sila ethers compared to their hydrophilic derivatives was reported by other groups as well.117-119 These results indicate that the TBS group can be employed to increase cellular uptake of compounds. However, TBS groups are expected to be labile under gastric (acidic pH) conditions and the decrease in aqueous solubility is a concern.

30 ACS Paragon Plus Environment

Page 30 of 71

Page 31 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Figure 30. TBS groups to enhance potency Since silyl ethers can be cleaved to hydroxyl groups in vivo, they can be used in designing prodrugs.28,120 Peterson and co-workers reported the anti-proliferative activity of ureidoadenosine derivatives such as compound 128.121,122 It was known that hydroxyl derivative 129 binds to the ATP binding site of the kinase BMPR1b. Dysregulation of BMP (bone morphogenetic protein) signalling is associated with various pathological conditions such as cancer. Hence, it was suggested that the activity of silyl derivatives may be due to inhibition of BMPR1b by 129 and compound 128 may be a prodrug. Hydroxyl derivative 129 did not show any antiproliferative activity indicating that lipophilic TBS group is necessary to enhance cell

31 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

penetration. Cleavage of the silyl group in cytoplasm gives the active compound 129, which in turn inhibits the BMP signalling cascade.

Figure 31. Silyl derivative 128 is a prodrug

10. Silicon substitution to modulate pharmacological selectivity In certain compounds, silicon substitution leads to a change in receptor selectivity. In the case of compound 52 (discussed in section 7.1), silyl substitution transformed the molecule from allosteric inhibitor to allosteric activator. Hashimoto and co-workers reported a change in receptor selectivity by silicon substitution in compounds 130 − 132.123 It was previously reported that some of the analogs of the known vitamin D receptor (VDR) agonist 130 (X = C) showed slight androgen receptor (AR) antagonistic activity.124,125 The silicon analogs showed decreased VDR agonistic activity compared to their parent compounds. However, they displayed better androgen receptor antagonistic activity compared to carbon compounds (figure 32).

32 ACS Paragon Plus Environment

Page 32 of 71

Page 33 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Figure 32. Change in receptor selectivity by silicon switching Tacke and co-workers synthesized silicon analogs of venlafaxine (133) which is a marketed antidepressant (in racemic form) belonging to serotonin-noradrenaline reuptake inhibitor class.126 Enhanced acidity of silanols 134 and 135 can be beneficial leading to an increased hydrogen bonding interaction with the receptor. However, it was found that silanols 134 and 135 displayed a different pharmacological profile. Electron density studies suggested that the difference in conformation and electron distribution of the silicon analog is responsible for this change of behaviour.127

Figure 33. Silicon analogs of venlafaxine Silicon analogs showed decreased serotonin reuptake inhibition (table 4). Interestingly sila analog 134 showed higher selectivity for the noradrenaline reuptake transporter. The pure enantiomers were prepared by diastereomeric resolution. The authors further postulated that selective noradrenaline reuptake inhibitors maybe

33 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 71

useful in treating emesis (vomiting). Accordingly, (R)-134 was evaluated in animal models and showed anti-emetic activity at 5 mg/kg. 128 Compound

IC50 (µM) Serotonin

Noradrenaline

Dopamine

rac- 133

0.020

0.149

4.430

(R)-133

0.030

0.061

19.600

(S)-133

0.006

0.754

6.670

rac- 134

1.063

0.109

2.630

(R)- 134

3.168

0.251

5.270

(S)- 134

0.791

4.715

36.350

rac- 135

0.904

0.275

0.707

Table 4. Monoamine reuptake transporter inhibition of venlafaxine and analogs Compounds 136a - 139a were known to be highly selective σ1 ligands. The silicon substituted analogs 136b - 139b showed higher affinity for the σ1-receptor compared to the carbon compounds. At the same time, the affinity of the sila compounds for the σ2-receptor was found to be decreased. Hence, in this case silicon incorporation lead to an increase in receptor selectivity (σ1 over σ2). The compound 136b had a 221-fold selectivity for σ1 over σ2 receptor. Compound 139b had a 63-fold selectivity for σ1 receptor which is 21-fold higher compared to its carbon analog 139a.129

34 ACS Paragon Plus Environment

Page 35 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Figure 34. Affinities of compounds 136 -139 for the σ-receptor Recently, Tacke group reported the synthesis of several silicon incorporated 4,5,6,7-tetrahydrobenzo[d]thiazole derivatives of lead compound 140 (figure 35).130 They evaluated the in vitro biological activity using GPR81 and GPR109A, two novel and potential targets for the treatment of diabetes. Although silicon incorporation reduced potency against GPR81, it slightly improved the potency with GPR109A. This suggests that silicon analogs (141 and 142) can become dual activators by targeting both GPR81 and GPR109A simultaneously (table 5). This dual nature may find additional benefits in further pharmacological studies.

Figure 35. Towards dual activators of GPR81 and GPR109A Compound 140

logD (pH 7.4) 4.2

EC50 (GPR81) 0.24 µM

35 ACS Paragon Plus Environment

EC50 (GPR109A) 2.1 µM

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 71

141

4.8

1.0 µM

1.2 µM

142

4.6

1.2 µM

1.3 µM

Table 5. In vitro biological activities of compounds 140 - 142

11. Silicon incorporation to improve the pharmacological properties Silicon incorporation can be used to enhance the drug like parameters of a lead molecule. Introduction of silicon in haloperidol (figure 2) resulted in an altered metabolic pathway that was devoid of the toxic pyridinium metabolite. Miller and coworkers reported biological activities of a sila analog of doramapimod (144) in which the t-butyl group was replaced with a trimethylsilyl moiety. Doramapimod (BIR-796, 143) is a p38 MAP kinase inhibitor, tested in humans for treating inflammatory disorders. Sila analog 144 showed slightly inferior in vitro activity against p38 MAP kinase compared to the parent compound 143. However, the human microsomal stability of the silicon derivative was higher than doramapimod. Both the compounds were tested for their in vivo efficacy in LPS induced TNF-α mice models and compound 144 with silicon incorporation showed better profile.131

Figure 36. Silicon substitution in p38 MAP kinase inhibitor

12. Silicon-containing amino acids 36 ACS Paragon Plus Environment

Page 37 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Unnatural amino acids are used to probe the bioactive conformation of peptides and find applications in peptidomimetics.132,133 Amino acids with silicon incorporation have several attractive features that can be useful in drug discovery.134,135 But, only a few silicon amino acids are known and with limited structural variations.136-150 Cavelier group has made significant contributions in this area.136-149 Some of the known silicon containing amino acids are shown in figure 37 (145 – 148). Incorporation of a lipophilic silicon moiety can improve the cell penetration as well as stability of the peptide towards biodegradation. An elaborative review on silicon amino acids appeared recently in literature from the Cavelier group.135

Figure 37. Some of the known sila amino acids The synthesis of silaproline 147 (Sip) was first reported by Cavelier group.144 The cLogP value of Fmoc-silaproline was experimentally found to be 14 times greater than Fmoc-proline, which shows that it is more lipophilic than proline. The authors incorporated this proline surrogate in place of proline in the neurotensin analog. Neurotensin is a neuropeptide containing 13 amino acids and has analgesic and hypothermic effects. The natural neurotensin undergoes proteolytic decomposition near to the proline residue and hence is injected along with enzyme inhibitors. The neurotensin analog with silaproline was active even in the absence of protease inhibitors which indicates that silicon incorporation improved the stability of the peptide in vivo.145 In another report, the replacement of proline with silaproline on the 37 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

hydrophobic face of a cell penetrating peptide led to a 20-fold enhancement in cellular uptake.146 Bristol-Myers Squibb has claimed silaproline containing compounds that modulate the activity of inhibitors of apoptosis (IAP). IAPs are proteins that regulate apoptosis and are reported to be overexpressed in cancer tissues.151 New analogs were tested for their inhibition of XIAP and compound 149 showed the highest potency.152

Figure 38. IAP modulator containing silaproline Silaproline was also incorporated in one of Merck’s lead compound MK-4882 (150), a hepatitis C virus (HCV) inhibitor. The Pro-Val dipeptide subunit is essential for the activity of compound 150. Several analogs were made by varying the central core and by manipulating the proline unit. Compounds 151 and 152 containing silicon and fluorine appeared to be promising candidates for further development. These compounds displayed good half-life (7.6 h and 6.7 h respectively; rat IV) and plasma exposure.153,154 Interestingly, silicon substitution led to an improvement in half-life of the compounds when compared to their carbon counterparts.

38 ACS Paragon Plus Environment

Page 38 of 71

Page 39 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Figure 39. HCV inhibitor containing silaproline Although not exactly relevant in this section, a recent report on antiviral compounds with silicon incorporation is worth mentioning. Wang’s group demonstrated the organosilicon compounds could be utilized towards antiviral agents, which targeting multidrug-resistant influenza A viruses. The antiviral activity and selectivity index were superior when compared to its counterpart made of carbon.155 The proline unit of captopril (153), an ACE inhibitor was replaced with silaproline and its biological activity was evaluated. However, silacaptopril (154) displayed a lower inhibition of ACE. This decrease in activity may be due to an increase in steric hindrance.141

39 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 40. Captopril and its silicon analog The replacement of leucine with TMS-alanine (145, figure 37) in the neurotensin analog led to an increased affinity for the neurotensin receptor. The peptide containing TMS-alanine showed potent analgesic effects in animal models of pain.140 Tacke and co-workers studied the effect of silicon incorporation in decapeptide cetrorelix, a GnRH (gonadotrophin-releasing hormone) antagonist. The TMS-alanine containing peptide showed significant activity in vivo with increased duration of action.156 β-amino acids with silicon incorporation have also been reported in literature. Jennifer and co-workers published the synthesis of silicon containing cyclic β-amino acids of the general formula 155.157 Skrydstrup and co-workers synthesized β-amino acids 156 and 157 and incorporated them in antimicrobial peptide alamethicin. One of the sila peptides showed 20-fold increase in calcein release compared to alamethicin.158

Figure 41. Silyl β-amino acids

40 ACS Paragon Plus Environment

Page 40 of 71

Page 41 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Marutake et al. synthesized and evaluated the silicon analogs of pregabalin (figure 42).159, 160 Pregabalin is an analog of gamma-amino butyric acid (GABA), a neurotransmitter present in the central nervous system. It is used to treat epilepsy and neuropathic pain. The anti-allodynic activity of the compounds 158 and 159 were checked in spinal nerve ligation models. The activity of 159 was found to be comparable to that of pregabalin. Interestingly, both the silicon analogs 158 and 159 did not show any anti-convulsant activity of pregabalin. Pregabalin causes dizziness and had hypalgesic effects that was not observed in rats dosed with 158 and 159. These results suggest that silicon analogs 158 and 159 can be considered as pregabalin analogs with reduced CNS side effects.159

Figure 42. Silicon containing GABA analogs Further optimization of the series led to the development of compounds 160 and 161.160 In rotarod experiments, rats administered with pregabalin lost their balance showing CNS side effects. The silagaba compounds were found to be devoid of any CNS side effects in rotarod experiments. PK experiments showed that the distribution of pregabalin into brain is more when compared to the silicon analogs that may be a reason for its side effects. The silagaba compounds showed weak binding to α2-δ protein which is presumed to be the target of pregabalin. This indicates that sila compounds may be binding to some other proteins causing similar allodynic activity as pregabalin.

41 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Along these lines, Reddy and co-workers synthesized silicon incorporated amino acids (alpha, beta and gamma; 162 − 164) with unusual 5,5-trans fusion starting from allylsilanes (figure 43). The ease of formation of this skeleton was explained by the larger C-Si bond length as compared to a C-C bond. The synthesis of lipophilic and constrained silicon analog of GABA (164) is interesting because conformationally rigid GABA analogs are known for the treatment of several CNS disorders. The sila-GABA derivative is similar to the marketed anti-epileptic drug gabapentin and another clinical candidate atagabalin.161 It is interesting to note that 5,5-trans fusion of diquinane resembles the stereochemical orientation of methyl groups of atagabalin.

Figure 43. Silicon amino acids with 5,5-trans fusion

13. Organosilanes that entered clinical trials Although there are no marketed drugs containing silicon, some of the compounds have entered human clinical trials (figure 44). To the best of our knowledge, nine compounds have been tested in humans; the results of which indicate

42 ACS Paragon Plus Environment

Page 42 of 71

Page 43 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

that silicon incorporation does not cause any toxicity concerns.7, 162 In fact, silabolin, a prodrug of the steroid nandrolone was available in Russia and used by athletes.28, 163

Figure 44. Organosilanes that entered human clinical trials

14. Challenges and way forward

43 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The benefits of introducing silicon in drug like scaffolds were discussed in detail in this perspective. However, this area is associated with certain challenges and drawbacks. One of the main challenges is to design efficient synthetic routes to access the target molecule containing silicon. The availability and cost of starting materials is a concern and the synthesis may involve uncommon reagents. Due to the inherent instability of certain bonds such as Si-H and Si-X, less structural variations are possible in organosilanes. Pharmaceutical companies may find this field less attractive because working on a scaffold not present in any marketed drug is riskier. Although, increased lipophilicity is advantageous for improving potency, it can hamper the druggable properties of the molecule such as aqueous solubility. Hence, the right balance between lipophilicity and hydrophilicity should be maintained while striving for more potent molecules. Although silicon offers potential avenues to medical chemists, the area remains underexplored. Advancement in synthetic methodologies which provide easy access to organosilicon compounds can attract more research groups to work on this interesting field.166-168 Recently, Arnold and co-workers reported the synthesis of organosilanes using heme proteins under physiological conditions.169 Prior to these findings, the synthesis of organosilanes using living organisms were not known. Although we have seen many stories with the concept of “silicon incorporation”, none of these has resulted in a marketed drug. All that is needed is one silicon-containing drug to make it to the market, to fuel interests in siliconincorporation, as seen with the “deuterium trick”. Once this is done, then more people from industry and academia can take up this concept seriously and use it in their lead optimization projects. The coming years will hopefully see a few candidates with

44 ACS Paragon Plus Environment

Page 44 of 71

Page 45 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

silicon incorporation in clinical evaluation and possibly the identification of further opportunities (may be serendipitous!) with this interesting class of compounds.

Author Information *Corresponding author: Dr. D. Srinivasa Reddy; E mail ID: [email protected]

Notes The authors declare no competing financial interest.

Acknowledgements We thank our colleagues Vasudevan, Rahul, Seetharam, Gorakh, and Akshay for their contributions

to

silicon-based

medicinal

chemistry

from

the

group

and

discussions/suggestions while preparing this perspective. Thanks to Aubrie Harland, School of Pharmacy, University of North Carolina for proofreading the article. We are also thankful to the reviewers for their useful suggestions in revising the manuscript. We acknowledge funding from CSIR, New Delhi through XII Five Year Plan projects (ORIGIN: CSC0108; NICE-P: CSC0109; NCL-IGIB joint program: BSC0124) and BIRAC (Biotechnology Industry Research Assistance Council), DBT, New Delhi for the support through CRS Scheme (BT/CRS0046/CRS-02/12).

Abbreviations [3H]NMS, [3H]N-methylscopolamine; LXR, liver X receptor; FXR, farnesoid X receptor; ROR, retinoid-related orphan receptors; HL60, Human promyelocytic leukemia cell line; BMPR1b, bone morphogenetic protein receptor type-1B;

45 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biographies Remya Ramesh completed her M.Sc. in Applied chemistry from Cochin University of Science and Technology, India, in 2011. She began her research career in CSIRNational Chemical Laboratory, Pune, India under the guidance of Dr. D. Srinivasa Reddy. Her doctoral research was mainly focused on the synthesis of sex pheromones and biologically active compounds containing silicon. Currently, she is working as a postdoctoral research associate with Prof. Jeffrey Aubé in the Eshelman School of Pharmacy, University of North Carolina at Chapel Hill, USA. D. Srinivasa Reddy received his Ph.D. in 2000 from University of Hyderabad (Prof. Goverdhan Mehta) followed by postdoctoral work at the University of Chicago (Prof. Sergey Kozmin) and the University of Kansas (Prof. Jeffrey Aubé). During the subsequent 7 years in drug discovery pharma industry, he successfully led a couple of drug discovery programs, among which one of the molecules discovered by his team is currently in Phase-II clinical trials. His current research interests are total synthesis of biologically active natural products and medicinal chemistry with a special emphasis on silicon incorporation. He was recognized with many awards including the Shanti Swarup Bhatnagar (SSB) prize in chemical sciences, a prestigious recognition in India. Recently, he was appointed as an editor of Bioorganic & Medicinal Chemistry Letters.

References (1)

Clayden, J.; Greeves, N.; Warren, S. Organic Chemistry, Oxford University Press, Oxford, second edition, 2012, 668-677.

46 ACS Paragon Plus Environment

Page 46 of 71

Page 47 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(2)

Bertrand, G. Chemistry. The modest undressing of a silicon center. Science 2004, 305, 783-785.

(3)

Meanwell, N. A. Synopsis of some recent tactical application of bioisosteres in drug design. J. Med. Chem. 2011, 54, 2529−2591.

(4)

Patani, G. A.; LaVoie, E. J. Bioisosterism: a rational approach in drug design. Chem. Rev. 1996, 96, 3147-3176.

(5)

Lima, L. M.; Barreiro, E. J. Bioisosterism: a useful strategy for molecular modification and drug design. Curr. Med. Chem. 2005, 12, 23-49.

(6)

Gately, S.; West, R. Novel therapeutics with enhanced biological activity generated by the strategic introduction of silicon isosteres into known drug Scaffolds. Drug Dev. Res. 2007, 68, 156−163.

(7)

Bains, W.; Tacke, R. Silicon chemistry as a novel source of chemical diversity in drug design. Curr. Opin. Drug Discovery Dev. 2003, 6, 526−543.

(8)

Franz, A. K.; Wilson, S. O. Organosilicon molecules with medicinal applications. J. Med. Chem. 2013, 56, 388−405.

(9)

Mills, J. S.; Showell, G. A. Exploitation of silicon medicinal chemistry in drug discovery. Expert Opin. Investig. Drugs 2004, 13, 1149-1157.

(10) Mills, J. S.; Showell, G. A. Chemistry challenges in lead optimization: silicon isosteres in drug discovery. Drug Discovery Today 2003, 8, 551-556. (11) Pooni, P. K.; Showell, G. A. Silicon switches of marketed drugs. Mini Rev. Med. Chem. 2006, 6, 1169-1177. (12) (a) Tacke, R.; Do¨rrich, S. Drug design based on the carbon/ silicon switch strategy. Top. Med. Chem. 2016, 17, 29–60. (b) Sieburth, S. M. Silicon mimics of unstable carbon. Top. Med. Chem. 2016, 17, 61–86. Both the

47 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

chapters were published in Atypical elements in drug design, Springer, San Fransisco, edited by Jacob Schwarz. (13) Fujii, S.; Hashimoto, Y. Progress in the medicinal chemistry of silicon: C/Si exchange and beyond. Future Med. Chem. 2017, 9, 485-505. (14) Koyanagi, T.; Haga, T. Bioisosterism in agrochemicals. ACS Symp. Ser. 1995, 584, 15–24. (15) Sieburth, S. McN. Isosteric replacement of carbon with silicon in the design of safer chemicals. ACS Symp. Ser. 1996, 640, 74–83. (16) Doszczak, L.; Gasperi, T.; Saint-Dizier, A.; Loreto, M. A.; Enders, D. Silylating reagents: a powerful tool for the construction of isosteric analogs of highly branched odorants. Chem Biodivers. 2004, 12, 1921-1935. (17) Liu, J.; Zhang, Q.; Li, P.; Qu, Z.; Sun, S.; Ma, Y.; Su, D.; Zong, Y.; Zhang, J. Six-membered silacycle odorants: synthesis and olfactory characterization of Si analogues of artemone, β-dynascone, and herbac. Eur. J. Inorg. Chem. 2014, 3435–3440. (18) Liu, J.; Zou, Y.; Fan, W.; Mao, J.; Chai, G.; Li, P.; Qu, Z.; Zong, Y.; Zhang, J.; Kraft, P. Synthesis and olfactory properties of silicon-containing analogs of rosamusk, romandolide, and applelide: insights into the structural parameters of linear alicyclic musks. Eur. J. Org. Chem. 2016, 976–982 (19) Doszczak, L.; Kraft, P.; Weber, H. –P.; Bertermann, R.; Triller, A.; Hatt, H.; Tacke, R. Prediction of perception: probing the hOR17-4 olfactory receptor model with silicon analogues of bourgeonal and lilial. Angew. Chem. Int. Ed. 2007, 46, 3367 –3371

48 ACS Paragon Plus Environment

Page 48 of 71

Page 49 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(20) Büttner, M. W.; Metz, S.; Kraft, P.; Tacke, R. Silicon-based noncyclic woodyambery odorants:  synthesis and olfactory characterization of 4,4,6,6tetramethylheptan-2-one and its sila-analogues. Organometallics 2007, 26, 3925–3929. (21) Büttner, M. W.; Burschka, C.; Junold, K.; Kraft, P.; Tacke, R. Disilaokoumal:

a

silicon

analogue

of

the

ambergris

odorant

okoumal.

ChemBioChem 2007, 8, 1447–1454. (22) Metz, S.; Nätscher, J. B.; Burschka, C.;Götz, K.; Kaupp, M.; Kraft, P.; Tacke, R. Disila-phantolide and derivatives: synthesis and olfactory characterization of

silicon-containing

derivatives

of

the

musk

odorant

phantolide.

Organometallics 2009, 28, 4700–4712. (23) Sunderkötter, A.; Lorenzen, S.; Tacke, R.; Kraft, P. Novel silicon-based patchouli odorants of the trialkyl(1-hydroxy-1-methylethyl)silane type: design, synthesis, and olfactory properties. Chem. Eur. J. 2010, 16, 7404– 7421. (24) Geyer, M.; Bauer, J.; Burschka, C.; Kraft, P.; Tacke, R. Synthesis and olfactory characterization of novel silicon-containing acyclic dienone musk odorants. Eur. J. Inorg. Chem. 2011, 2769–2776. (25) Förster, B.; Bertemann, R.; Kraft, P.; Tacke, R. Sila-rhubafuran and derivatives: synthesis and olfactory characterization of novel silicon containing odorants. Organometallics 2014, 33, 338-346. (26) Wannagat, U.; Damrath, V.; Huch, V.; Veith, M.; Harder, U. Sila-Riechstoffe und

riechstoffisostere

XII*.

Geruchsvergleiche

49 ACS Paragon Plus Environment

homologer

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 50 of 71

organoelementverbindungender vierten Hauptgruppe (C, Si, Ge, Sn). J. Organomet. Chem. 1993, 443, 153–165. (27) Tacke, R.; Zilch, H. Sila-substitution — a useful strategy for drug design? Endeavour 1986, 10, 191–197. (28) Millership, J. S.; Shanks, M. L. Prodrugs utilizing a reversible silyl linkage. Int. J. Pharm. 1986, 28, 19. (29) Fujii, S. Expanding the chemical space of hydrophobic pharmacophores: the role of hydrophobic substructures in the development of novel transcription modulators. Med. Chem. Commun. 2016, 7, 1082-1092. (30) Lin, J. H.; Lu, A. Y. H. Role of pharmacokinetics and metabolism in drug discovery and development. Pharmacol. Rev. 1997, 49, 403-449. (31) Tacke, R.; Nguyen, B.; Burschka, C.; Lippert, W. P.; Hamacher, A.; Urban, C.; Kassack, M. U. Sila-trifluperidol, a silicon analogue of the dopamine (D2) receptor

antagonist

trifluperidol:

synthesis

and

pharmacological

characterization. Organometallics 2010, 29, 1652–1660. (32) Johansson, T.; Weidolf, L.; Popp, F.; Tacke, R.; Jurva, U. In vitro metabolism of haloperidol and sila-haloperidol: new metabolic pathways resulting from carbon/silicon exchange. Drug Metab. Dispos. 2010, 38, 73-83. (33) Tacke, R.; Popp, F.; Mueller, B.; Theis, B.; Burschka, C.; Hamacher, A.; Kassack, M. U.; Schepmann, D.; Wuensch, B.; Jurva, U.; Wellner, E. Silahaloperidol, a silicon analogue of the dopamine (D2) receptor antagonist haloperidol: synthesis, pharmacological properties, and metabolic fate. ChemMedChem 2008, 3, 152−164.

50 ACS Paragon Plus Environment

Page 51 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(34) Tacke, R.; Heinrich, T.; Bertermann, R.; Burschka, C.; Hamacher, A.; Kassack, M. U. Sila-haloperidol:  a silicon analogue of the dopamine (D2) receptor antagonist haloperidol. Organometallics 2004, 23, 4468−4477. (35) Dauer, W.; Przedborski, S. Parkinson’s disease: mechanisms and models. Neuron 2003, 39, 889-909. (36) Subramanyam, B.; Rollema, H.; Woolf, T.; Castagnoli, N. Jr. Identification of a potentially neurotoxic pyridinium metabolite of haloperidol in rats. Biochem Biophys Res Commun. 1990, 166, 238-244. (37) Subramanyam, B.; Pond, S. M.; Eyles, D. W.; Whiteford, H. A.; Fouda, H. G.; Castagnoli, N. Jr. Identification of potentially neurotoxic pyridinium metabolite in the urine of schizophrenic patients treated with haloperidol. Biochem Biophys Res Commun. 1991, 181, 573-578. (38) Subramanyam, B.; Woolf, T.; Castagnoli, N. Jr. Studies on the in vitro conversion of haloperidol to a potentially neurotoxic pyridinium metabolite. Chem. Res. Toxicol. 1991, 4, 123-128. (39) Svennebring, A. The role of intramolecular self-destruction of reactive metabolic intermediates in determining toxicity. J. Appl. Toxicol. 2016, 36, 483-500. (40) Tacke, R.; Nguyen, B.; Burschka, C.; Lippert, W. P.; Hamacher, A.; Urban, C.; Kassack, M. U. Sila-Trifluperidol, a silicon analogue of the dopamine (D2) receptor

antagonist

trifluperidol:

synthesis

and

pharmacological

characterization. Organometallics 2010, 29, 1652–1660. (41) Geyer, M.; Wellner, E.; Jurva, U.; Saloman, S.; Armstrong, D.; Tacke, R. Can silicon make an excellent drug even better? An in vitro and in vivo head-to-

51 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

head comparison between loperamide and its silicon analogue sila-loperamide. ChemMedChem. 2015, 10, 911-924. (42) Turk, B. Targeting proteases: successes, failures and future prospects. Nat. Rev. Drug Discovery 2006, 5, 785-799. (43) Chen, C. A.; Sieburth, S. McN.; Glekas, A.; Hewitt, G. W.; Trainor, G. L.; Erickson-Viitanen, S.; Garber, S. S.; Cordova, B.; Jeffry,S.; Klabe, R. M. Drug design with a new transition state analog of the hydrated carbonyl: silicon-based inhibitors of the HIV protease. Chem. Biol. 2001, 8, 1161-1166. (44) Sieburth, S. McN.; Nittoli, T.; Mutahi, A. M.; Guo, L. Silanediols: a new class of potent protease inhibitors. Angew. Chem. Int. Ed. 1998, 37, 812-814. (45) wa Mutahi, M.; Nittoli, T.; Guo, L.; Sieburth, S. McN. Silicon-based metalloprotease inhibitors: synthesis and evaluation of silanol and silanediol peptide analogues as inhibitors of angiotensin-converting enzyme. J. Am. Chem. Soc. 2002, 124, 7363-7375. (46) Kim, J.; Glekas, K.; Sieburth, S. McN. Silanediol-based inhibitor of thermolysin. Bioorg. Med. Chem. Lett. 2002, 12, 3625–3627. (47) Kim, J.; Sieburth, S. McN. Silanediol peptidomimetics. Evaluation of four diastereomeric ACE inhibitors. Bioorg. Med. Chem. Lett. 2004, 14, 2853– 2856. (48) Juers, D. H.; Kim, J.; Matthews, B. W.; Sieburth, S. McN. Structural analysis of silanediols as transition-state-analogue inhibitors of the benchmark metalloprotease thermolysin. Biochemistry 2005, 44, 16524−16528. (49) Sieburth, S. McN.; Chen, C.-A. Silanediol protease inhibitors: from conception to validation. Eur. J. Org. Chem. 2006, 311−322.

52 ACS Paragon Plus Environment

Page 52 of 71

Page 53 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(50) Singh, S.; Sieburth, S. McN. Serine protease inhibition by a silanediol peptidomimetic. Org. Lett. 2012, 14, 4422−4425. (51) Kim, J. K.; Sieburth, S. McN. Synthesis and properties of a sterically unencumbered δ-silanediol amino acid. J. Org. Chem. 2012, 77, 2901–2906. (52) Madsen, A. S.; Kristensen, H. M. E.; Lanz, G.; Olsen, C. A. The effect of various zinc binding groups on inhibition of histone deacetylases 1–11. ChemMedChem 2014, 9, 614-626. (53) Organ, M. G.; Buon, C.; Decicco, C. P.; Combs, A. P. A concise synthesis of silanediol-based transition-state isostere inhibitors of proteases. Org. Lett. 2002, 4, 2683-2685. (54) Madsen, J. L. H.; Andersen, T. L.; Santamaria, S.; Nagase, H.; Enghild, J. J.; Skrydstrup, T. Synthesis and evaluation of silanediols as highly selective uncompetitive inhibitors of human neutrophil elastase. J. Med. Chem. 2012, 55, 7900−7908. (55) Blunder, M.; Hurkes, N.; Spirk, S.; List, M.; Pietschnig, R. Silanetriols as in vitro inhibitors for AChE. Bioorg. Med. Chem. Lett. 2011, 21, 363–365. (56) González-Bello, C. Designing irreversible inhibitors- worth the effort? ChemMedChem 2016, 11, 22-30. (57) Oleinick, N. L.; Antunez, A. R.; Clay, M. E.; Rihter, B. D.; Kenney, M. E. New phthalocyanine photosensitizers for photodynamic therapy. Photochem. Photobiol. 1993, 57, 242–247. (58) Zaidi, S. I. A.; Agarwal, R.; Eichler, G.; Rihter, D.; Kenney, M. E.; Mukhtar, H. Photodynamic effects of new silicon phthalocyanines: in vitro studies

53 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

utilizing rat hepatic microsomes and human erythrocyte ghosts as model membrane sources. Photochem. Photobiol. 1993, 58, 204–210. (59) Baron, E. D.; Malbasa, C. L.; Santo-Domingo, D.; Fu, P.; Miller, J. D.; Hanneman, K. K.; Hsia, A. H.; Oleinick, N. L.; Colussi, V. C.; Cooper, K. D. Silicon phthalocyanine (Pc 4) photodynamic therapy is a safe modality for cutaneous neoplasms: results of a phase 1 clinical trial. Lasers Surg. Med. 2010, 42, 728–735. (60) Mfouo-Tynga, I.; Abrahamse, H. Cell death pathways and phthalocyanine as an efficient agent for photodynamic cancer therapy. Int. J. Mol. Sci. 2015, 16, 10228-10241. (61) Morris, R. L.; Varnes, M. E.; Kenney, M. E.; Li, Y. –S.; Azizuddin, K.; McEnery, M. W.; Oleinick, N. L. The peripheral benzodiazepine receptor in photodynamic therapy with the phthalocyanine photosensitizer Pc 4. Photochem. Photobiol. 2002, 75, 652–661. (62) Josefsen, L. B.; Boyle, R. W. Photodynamic therapy and the development of metal-based photosensitisers. Met. Based Drugs 2008, Article ID 276109, doi:10.1155/2008/276109 (63) Allen, C. M.; Sharman, W. M.; Van Lier, J. E. Current status of phthalocyanines in the photodynamic therapy of cancer. J. Porphyrins Phthalocyanines 2001, 5, 161–169. (64) Li, J.; Yang, Y.; Zhang, P.; Sounik, J. R.; Kenney, M. E. Synthesis, properties and drug potential of the photosensitive alkyl- and alkylsiloxy-ligated silicon phthalocyanine Pc 227. Photochem. Photobiol. Sci. 2014, 13, 1690-1698.

54 ACS Paragon Plus Environment

Page 54 of 71

Page 55 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(65) Xiang, Y.; Fu, C.; Breiding, T.; Sasmal, P. K.; Liu, H.; Shen, Q.; Harms, K.; Zhang, L.; Meggers, E. Hydrolytically stable octahedral silicon complexes as bioactive scaffolds: application to the design of DNA intercalators. Chem. Commun. 2012, 48, 7131–7133. (66) Wheate, N. J.; Brodie, C. R.; Collins, J. G.; Kemp, S.; Aldrich-Wright, J. R. DNA intercalators in cancer therapy: organic and inorganic drugs and their spectroscopic tools of analysis. Mini Rev. Med. Chem. 2007, 7, 627-648. (67) Henker, J.; Wirmer-Bartoschek, J.; Bendel, L. E.; Xiang, Y.; Fu, C.; Harms, K.; Schwalbe, H.; Meggers, E. Progress in the synthesis and bioactivity of hexacoordinate silicon (IV) complexes. Eur. J. Inorg. Chem. 2016, 5161– 5170. (68) Frye, C.

L.; Vogel, G.

E.; Hall, J. A.

Triptych-siloxazolidines:

Pentacoordinate bridgehead silanes resulting from transannular interaction of nitrogen and silicon. J. Am. Chem. Soc. 1961, 83, 996–997. (69) Puri, J. K.; Singh, R.; Chahal, V. K. Silatranes: a review on their synthesis, structure, reactivity and applications. Chem. Soc. Rev. 2011, 40, 1791–1840. (70) Voronkov, M. G. Biological activity of silatranes. Top Curr. Chem. 1979, 84, 77–135. (71) Black, C. A.; Ucci, J. W.; Vorpagel, J. S.; Mauck, M. C.; Fenlon, E. E. Stereoselective and improved syntheses and anticancer testing of 3′-Osilatranylthymidines. Bioorg. Med. Chem. Lett. 2002, 12, 3521–3523. (72) Grna, A.; Koo, P. H.; Hogan, J. Antitumor effect of novel silatranes on renal cell carcinoma in mice. Anticancer Res.1992, 12, 565–569.

55 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(73) Mehta, P. P.; Ramasarma, T.; Kurup, C. K. R. Hypocholesterolemic action of 1-ethoxysilatrane. Biochim. Biophys. Acta 1987, 920, 102-104. (74) Mehta, P. P.; Ramasarma, T.; Kurup, C. K. R. Investigations on the mechanism of the hypocholesterolemic action of 1-ethoxysilatrane. Mol. Cell Biochem. 1990, 97, 75-85. (75) Oller-Salvia, B.; Sánchez-Navarro, M.; Giralta, E.; Teixidó, M. Blood–brain barrier shuttle peptides: an emerging paradigm for brain delivery. Chem. Soc. Rev. 2016, 45, 4690-4707. (76) Gabathuler, R. Approaches to transport therapeutic drugs across the bloodbrain barrier to treat brain diseases. Neurobiol. Dis. 2010, 37, 48-57. (77) Seetharamsingh, B.; Ramesh, R.; Dange, S. S.; Khairnar, P. V.; Singhal, S.; Upadhyay, D.; Veeraraghavan, S.; Viswanadha, S.; Vakkalanka, S.; Reddy, D. S. Design, synthesis, and identification of silicon incorporated oxazolidinone antibiotics with improved brain exposure. ACS Med. Chem. Lett. 2015, 6, 1105−1110. (78) Bom, D.; Curran, D. P.; Kruszewski, S.; Zimmer, S. G.; Strode, J. T.; Kohlhagen, G.; Du, W.; Chavan, A. J.; Fraley, K. A.; Bingcang, A. L.; Latus, L. J.; Pommier, Y.; Burke, T. G. The novel silatecan 7-tert-butyldimethylsilyl10-hydroxycamptothecin displays high lipophilicity, improved human blood stability, and potent anticancer activity. J. Med. Chem. 2000, 43, 3970−3980. (79) Van Hattum, A. H.; Pinedo, H. M.; Schluper, H. M. M.; Hausheer, F. H.; Boven, E. New highly lipophilic camptothecin BNP1350 is an effective drug in experimental human cancer. Int. J. Cancer 2000, 88, 260–266.

56 ACS Paragon Plus Environment

Page 56 of 71

Page 57 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(80) Van Hattum, A. H.; Schluper, H. M.; Hausheer, F. H.; Pinedo, H. M.; Boven, E. Novel camptothecin derivative BNP1350 in experimental human ovarian cancer: determination of efficacy and possible mechanisms of resistance. Int. J. Cancer 2002, 100, 22–29. (81) Daud, A.; Valkov, N.; Centeno, B.; Derderian, J.; Sullivan, P.; Munster, P.; Urbas, P.; DeConti, R. C.; Berghorn, E.; Liu, Z.; Hausheer, F.; Sullivan, D. Phase II trial of Karenitecin in patients with malignant melanoma: clinical and translational study. Clin. Cancer Res. 2005, 11, 3009−3016. (82) Venditto, V. J.; Simanek, E. E. Cancer therapies utilizing the camptothecins: a review of the in vivo literature. Mol. Pharmaceutics 2010, 7, 307−349. (83) Arnold, S. M.; Rinehart, J. J.; Tsakalozou, E.; Eckardt, J. R.; Fields, S. Z.; Shelton, B. J.; DeSimone, P. A.; Kee, B.K.; Moscow, J. A.; Leggas, M. A. Phase I study of 7-t-butyldimethylsilyl-10-hydroxycamptothecin in adult patients with refractory or metastatic solid malignancies. Clin. Cancer Res. 2010, 16, 673–680. (84) Lopez-Barcons, L. A.; Zhang, J.; Siriwitayawan, G.; Burke, T. G.; PerezSoler, R. The novel highly lipophilic topoisomerase I inhibitor DB67 is effective in the treatment of liver metastases of murine CT-26 colon carcinoma. Neoplasia 2004, 6, 457–467. (85) Yeh, T. -K.; Li, C. -M.; Chen, C. -P.; Chuu, J. -J.; Huang, C. -L.; Wang, H. S.; Shen, C. -C.; Lee, T. -Y.; Chang, C. -Y.; Chang, C. -M.; Chao, Y. -S.; Lin, C. -T.; Chang, J. -Y.; Chen, C. -T. Antitumor activities and pharmacokinetics of silatecans DB-67 and DB-91. Pharmacol. Res. 2010, 61, 108–115.

57 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 58 of 71

(86) Hornsperger, J. M.; Collard, J. N.; Heydt, J. G.; Giacobinit, E.; Funes, S.; Dow, J.; Schirlin, D. Trimethylsilylated trifluoromethyl ketones, a novel class of acetylcholinesterase inhibitors: biochemical and pharmacological profile of MDL 73,745. Biochem. Soc. Trans. 1994, 22, 758-763. (87) Cutler, N. R.; Seifert, R. D.; Schleman, M. M.; Sramek, J. J.; Szylleyko, O. J.; Howard,

D.

R.;

Barchowsky,

Acetylcholinesterase

inhibition

A.; Wardle, by

zifrosilone:

T.

S.; Brass,

E.

pharmacokinetics

P. and

pharmacodynamics. Clin. Pharmacol. Ther. 1995, 58, 54-61. (88) Zhu, X. D.; Giacobini, E.; Hornsperger, J. –M. Effect of MDL 73,745 on acetylcholine and biogenic amine levels in rat cortex. Eur. J. Pharmacol. 1995, 276, 93–99. (89) Coyle, J. T.; Price, D. L.; Delong, M. R. Alzheimer's disease: a disorder of cortical cholinergic innervation. Science 1983, 219, 1184-1190. (90) Brodbeck, U.; Schweikert, K.; Gentinetta. R.; Rottenberg, M. Fluorinated aldehydes

and

ketones

acting

as

quasi-substrate

inhibitors

of

acetylcholinesterase. Biochim. Biophys. Acta 1979, 567, 357-369. (91) Nair, H. K.; Lee, K.; Quinn, D. M. m-(N,N,N-Trimethy1ammonio) trifluoroacetophenone: a femtomolar inhibitor of acetylcholinesterase. J. Am. Chem. Soc. 1993, 115, 9939-9941. (92) Daiss, J. O.; Duda-Johner, S.; Burschka, C.; Holzgrabe, U.; Mohr, K.; Tacke, R. N+/Si replacement as a tool for probing the pharmacophore of allosteric modulators of muscarinic M2 receptors: synthesis, allosteric potency, and positive cooperativity of silicon-based W84 derivatives. Organometallics 2002, 21, 803–811.

58 ACS Paragon Plus Environment

Page 59 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(93) Duda-Johner, S.; Daiß, J. O.; Mohr, K.; Tacke, R. Synthesis and pharmacological characterization of new silicon-based W84-type allosteric modulators for ligand binding to muscarinic M2 receptors. J. Organomet. Chem. 2003, 686, 75-83. (94) West, R.; Baney, R. H. Hydrogen bonding studies. II. The acidity and basicity of silanols compared to alcohols. J. Am. Chem. Soc. 1959, 81, 6145-6148. (95) Kim, Y. -m.; Farrah, S.; Baney, R. H. Structure–antimicrobial activity relationship for silanols, a new class of disinfectants, compared with alcohols and phenols. Int. J. Antimicrob. Agents 2007, 29, 217−222. (96) Toyama, H.; Sato, S.; Shirakawa, H.; Komai, M.; Hashimoto, Y.; Fujii, S. Altered activity profile of a tertiary silanol analog of multi-targeting nuclear receptor modulator T0901317. Bioorg. Med. Chem. Lett. 2016, 26, 18171820. (97) Tron, G. C.; Pirali, T.; Sorba, G.; Pagliai, F.; Busacca, S.; Genazzani, A. A. Medicinal chemistry of combretastatin A4:  present and future directions. J. Med. Chem. 2006, 49, 3033-3044. (98) Nakamura, M.; Kajita, D.; Matsumoto, Y.; Hashimoto, Y. Design and synthesis of silicon-containing tubulin polymerization inhibitors: replacement of the ethylene moiety of combretastatin A-4 with a silicon linker. Bioorg. Med. Chem. 2013, 21, 7381–7391. (99) Kajita, D.; Nakamura, M.; Matsumoto, Y.; Ishikawa, M.; Hashimoto, Y.; Fujii, S. Design and synthesis of silicon-containing fatty acid amide derivatives as novel peroxisome proliferator-activated receptor (PPAR) agonists. Bioorg. Med. Chem. Lett. 2015, 25, 3350–3354. 59 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 60 of 71

(100) Nishiyama, Y.; Nakamura, M.; Misawa, T.; Nakagomi, M.; Makishima, M.; Ishikawa,

M.;

Hashimoto,

Y.

Structure-activity

relationship-guided

development of retinoic acid receptor-related orphan receptor gamma (RORγ)-selective inverse agonists with a phenanthridin-6(5H)-one skeleton from a liver X receptor ligand. Bioorg. Med. Chem. 2014, 22, 2799-2808. (101) Toyama, H.; Nakamura, M.; Hashimoto, Y.; Fujii, S. Design and synthesis of novel ROR inverse agonists with a dibenzosilole scaffold as a hydrophobic core structure. Bioorg. Med. Chem. 2015, 23, 2982–2988. (102) Fauber, B. P.; de Leon Boenig, G.; Burton, B.; Eidenschenk, C.; Everett, C.; Gobbi, A.; Hymowitz, S. G.; Johnson, A. R.; Liimatta, M.; Lockey, P.; Norman, M.; Ouyang, W.; Rene, O.; Wong, H. Structure-based design of substituted hexafluoroisopropanol- arylsulfonamides as modulators of RORc. Bioorg. Med. Chem. Lett. 2013, 23, 6604-6609. (103) Fujii, S.; Miyajima, Y.; Masuno, H.; Kagechika, H. Increased hydrophobicity and estrogenic activity of simple phenols with silicon and germaniumcontaining substituents. J. Med. Chem. 2013, 56, 160−166. (104) Tacke, R.; Terunuma, D.; Tafel, A.; Muhleisen, M.; Forth, B.; Waelbroeck, M.; Gross, J.; Mutschler, E.; Friebe, T.; Lambrecht, G. Fluorine-containing derivatives of the muscarinic antagonists sila-pridinol and sila-difenidol: syntheses and antimuscarinic properties. J. Organomet. Chem. 1995, 501, 145-154. (105) Wang, J.; Ma, C.; Wu, Y.; Lamb, R. A.; Pinto, L. H.; DeGrado, F. Exploring organosilane amines as potent inhibitors and structural probes of influenza A virus M2 proton channel. J. Am. Chem. Soc. 2011, 133, 13844–13847.

60 ACS Paragon Plus Environment

Page 61 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(106) Lippert, W. P.; Burschka, C.; Gotz, K.; Kaupp, M.; Ivanova, D.; Gaudon, C.; Sato, Y.; Antony, P.; Rochel, N.; Moras, D.; Gronemeyer, H.; Tacke, R. Silicon analogues of the RXR-selective retinoid agonist SR11237 (BMS649): chemistry and biology. ChemMedChem 2009, 4, 1143–1152. (107) Tacke, R.; Muller, V.; Buttner, M. W.; Lippert, W. P.; Bertermann, R.; Daiss, J. O.; Khanwalkar, H.; Furst, A.; Gaudon, C.; Gronemeyer, H. Synthesis and pharmacological characterization of disila-AM80 (disila-tamibarotene) and disila-AM580, silicon analogues of the RARα-selective retinoid agonists AM80 (Tamibarotene) and AM580. ChemMedChem 2009, 4, 1797–1802. (108) Reddy, G. D.; Park, S. -J.; Cho, H. M.; Kim, T. -J.; Lee, M. E. Antiallergic activity profile in vitro RBL-2H3 and in vivo passive cutaneous anaphylaxis mouse model of new sila-substituted 1,3,4-oxadiazoles. J. Med. Chem. 2012, 55, 6438−6444. (109) Li, Y.; de Kock, C.; Smith, P. J.; Guzgay, H.; Hendricks, D. T.; Naran, K.; Mizrahi, V.; Warner, D. F.; Chibale, K.; Smith, G. S. Synthesis, characterization, and pharmacological evaluation of silicon-containing aminoquinoline organometallic complexes as antiplasmodial, antitumor, and antimycobacterial agents. Organometallics 2013, 32, 141–150. (110) Li, Y.; de Kock, C.; Smith, P. J.; Chibale, K.; Smith, G. S. Synthesis and evaluation of a carbosilane congener of ferroquine and its corresponding halfsandwich ruthenium and rhodium complexes for antiplasmodial and β‑ hematin inhibition activity. Organometallics 2014, 33, 4345–4348. (111) Adams, M.; de Kock, C.; Smith, P. J.; Land, K. M.; Liu, N.; Hopper, M.; Hsiao, A.; Burgoyne, A. R.; Stringer, T.; Meyer, M.; Wiesner, L.; Chibale, K.;

61 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Smith, G. S. Improved antiparasitic activity by incorporation of organosilane entities into half-sandwich ruthenium(II) and rhodium(III) thiosemicarbazone complexes. Dalton Trans. 2015, 44, 2456-2468. (112) Jachak, G. R.; Ramesh, R.; Sant, D. G.; Jorwekar, S. U.; Jadhav, M. R.; Tupe, S. G.; Deshpande, M.V.; Reddy, D. S. Silicon incorporated morpholine antifungals: design, synthesis, and biological evaluation. ACS Med. Chem. Lett. 2015, 6, 1111−1116. (113) Ramesh, R.; Shingare, R. D.; Kumar, V.; Anand, A.; Swetha, B.; Veeraraghavan, S.; Viswanadha, S.; Ummanni, R.; Gokhale, R.; Reddy, D. S. Repurposing of a known drug scaffold: identification of novel sila analogues of rimonabant as potent antitubercular agents. Eur. J. Med. Chem. 2016, 122, 723-730. (114) Donadel, O. J.; Martín, T.; Martín, V. S.; Villar, J.; Padron, J. M. The tertbutyl dimethyl silyl group as an enhancer of drug cytotoxicity against human tumor cells. Bioorg. Med. Chem. Lett. 2005, 15, 3536–3539. (115) António, J. P. M.; Frade, R. F. M.; Santos, F. M. F.; Coelho, J. A. S.; Afonso, C. A. M.; Gois, P. M. P.; Trindade, A. F. NHC catalysed direct addition of HMF to diazo compounds: synthesis of acyl hydrazones with antitumor activity. RSC Adv. 2014, 4, 29352-29356. (116) Vernekar, S. K. V.; Qiu, L.; Zhang, J.; Kankanala, J.; Li, H.; Geraghty, R. J.; Wang, Z. 5'-Silylated 3'-1,2,3-triazolyl thymidine analogues as inhibitors of west nile virus and dengue virus. J. Med. Chem. 2015, 58, 4016-4028.

62 ACS Paragon Plus Environment

Page 62 of 71

Page 63 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(117) Szilágyi, A.; Fenyvesi, F.; Majercsik, O.; Pelyvás, I. F.; Bácskay, I.; Fehér, P.; Váradi, J.; Vecsernyés, M.; Herczegh, P. Synthesis and cytotoxicity of leinamycin antibiotic analogues. J. Med. Chem. 2006, 49, 5626-5630. (118) León, L. G.; Miranda, P. O.; Martín, V. S.; Padrón, J. I.; Padrón, J. M. Antiproliferative

activity

of

4-chloro-5,6-dihydro-2H-pyrans.

part

2:

enhancement of drug cytotoxicity. Bioorg. Med. Chem. Lett. 2007, 17, 30873090. (119) Zablotskaya, A.; Segal, I.; Popelis, Y.; Grinberga, S.; Shestakova, I.; Nikolajeva, V.; Ezeb, D. Silyl modification of biologically active compounds. 13. synthesis, cytotoxicity and antibacterial action of N-methyl-N-(2triorganylsiloxyethyl)-1,2,3,4-tetrahydro(iso)quinolinium

iodides.

Appl.

Organometal. Chem. 2013, 27, 114–124. (120) Pungitore, C. R.; León, L. G.; García, C.; Martín, V. S.; Tonn, C. E.; Padrón, J. M. Novel antiproliferative analogs of the Taq DNA polymerase inhibitor catalpol. Bioorg. Med. Chem. Lett. 2007, 17, 1332–1335. (121) Shelton, J. R.; Cutler, C. E.; Browning, M. S.; Balzarini, J.; Peterson, M. A. Synthesis and SAR of 2’,3’-bis-O-substituted N6, 5’-bis-ureidoadenosine derivatives: implications for prodrug delivery and mechanism of action. Bioorg. Med. Chem. Lett. 2012, 22, 6067–6071. (122) Shelton, J. R.; Cutler, C. E.; Oliveira, M.; Balzarini, J.; Peterson, M. A. Synthesis,

SAR,

antiproliferative

and

preliminary

N6,5′-bis-ureido-

mechanistic

and

evaluation

of

novel

5′-carbamoyl-N6-ureidoadenosine

derivatives. Bioorg. Med. Chem. 2012, 20, 1008-1019.

63 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 64 of 71

(123) Nakamura, M.; Makishima, M.; Hashimoto, Y. Development of siliconcontaining bis-phenol derivatives as androgen receptor antagonists: selectivity switching by C/Si exchange. Bioorg. Med. Chem. 2013, 21, 1643–1651. (124) Hosoda, S.; Tanatani, A.; Wakabayashi, K.; Makishima, M.; Imai, K.; Miyachi, H.; Nagasawa, K.; Hashimoto, Y. Ligands with a 3,3diphenylpentane skeleton for nuclear vitamin D and androgen receptors: dual activities and metabolic activation. Bioorg. Med. Chem. 2006, 14, 5489-5502. (125) Maruyama, K.; Noguchi-Yachide, T.; Sugita, K.; Hashimoto, Y.; Ishikawa, M. Novel selective anti-androgens with a diphenylpentane skeleton. Bioorg. Med. Chem. Lett. 2010, 20, 6661-6666. (126) Daiss, J. O.; Burschka, C.; Mills, J. S.; Montana, J. G.; Showell, G. A.; Warneck, J. B. H.; Tacke, R. Sila-venlafaxine, a sila-analogue of the serotonin/noradrenaline reuptake inhibitor venlafaxine: synthesis, crystal structure analysis, and pharmacological characterization. Organometallics 2006, 25, 1188-1198. (127) Luger, P.; Dittrich, B.; Tacke, R. Invariom based electron density studies on the

C/Si

analogues

haloperidol/sila-haloperidol

and

venlafaxine/sila-

venlafaxine. Org. Biomol. Chem. 2015, 13, 9093-9106. (128) Showell, G. A.; Barnes, M. J.; Daiss, J. O.; Burschka, C.; Mills, J. S.; Montana, J. G.; Tacke, R. Warneck, J. B. H. (R)-Sila-venlafaxine: a selective noradrenaline reuptake inhibitor for the treatment of emesis. Bioorg. Med. Chem. Lett. 2006, 16, 2555–2558. (129) Tacke, R.; Bertermann, R.; Burschka, C.; Dörrich, S.; Fischer, M.; Müller, B.; Meyerhans, G.; Schepmann, D.; Wünsch, B.; Arnason, I.; Bjornsson, R. High-

64 ACS Paragon Plus Environment

Page 65 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

affinity,

selective

σ ligands

silaspiro[naphthalene-1,4’-piperidine]

of

the

type:

1,2,3,4-tetrahydro-1,4’-

syntheses,

structures,

and

pharmacological properties. ChemMedChem 2012, 7, 523 – 532. (130) Geyer, M.; Baus, J. A.; Fjellström, O.; Wellner, E.; Gustafsson, L.; Tacke, R. Synthesis and pharmacological properties of silicon containing GPR81 and GPR109A agonists. ChemMedChem 2015, 10, 2063-2070. (131) Barnes, M. J.; Conroy, R.; Miller, D. J.; Mills, J. S.; Montana, J. G.; Pooni, P. K.; Showell, G. A.; Walsh, L. M.; Warneck, J. B. H. Trimethylsilylpyrazoles as novel inhibitors of p38 MAP kinase: a new use of silicon bioisosteres in medicinal chemistry. Bioorg. Med. Chem. Lett. 2007, 17, 354–357. (132) Stevenazzi, A.; Marchini, M.; Sandrone, G.; Vergani, B.; Lattanzio, M. Amino acidic scaffolds bearing unnatural side chains: an old idea generates new and versatile tools for the life sciences. Bioorg. Med. Chem. Lett. 2014, 5349-5356. (133) Blaskovich, M. A. T. Unusual amino acids in medicinal chemistry. J. Med. Chem. 2016, 59, 10807−10836. (134) Mortensen, M.; Husmann, R.; Veri, E.; Bolm, C. Synthesis and applications of silicon-containing α-amino acids. Chem. Soc. Rev. 2009, 38, 1002−1010. (135) Rémond, E.; Martin, C.; Martinez, J.; Cavelier, F. Silicon-containing amino acids: synthetic aspects, conformational studies, and applications to bioactive peptides. Chem. Rev. 2016, 116, 11654–11684. (136) Cavelier, F.; Marchand, D.; Martinez, J.; Sagan, S. Biological activity of silylated amino acid containing substance P analogues. J. Peptide Res. 2004, 63, 290–296.

65 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 66 of 71

(137) Cavelier, F.; Marchand, D.; Martinez, J. α,α′-Disubstituted amino acids with silylated side chains as lipophilic building blocks for the synthesis of peptaibol analogues. Chem. Biodivers. 2008, 5, 1279-1287. (138) Fanelli, R.; Besserer-Offroy É.; René A.; Côté J.; Tétreault, P.; ColleretteTremblay, J.; Longpré, J. –M.; Leduc, R.; Martinez, J.; Sarret, P.; Cavelier, F. Synthesis

and

characterization

in

vitro

and

in

vivo

of

(L)‑

(Trimethylsilyl)alanine containing neurotensin analogues. J. Med. Chem. 2015, 58, 7785−7795. (139) Fanelli, R.; Salah, K. B. H.; Inguimbert, N.; Didierjean, C.; Martinez, J.; Cavelier, F. Access to α,α-disubstituted disilylated amino acids and their use in solid-phase peptide synthesis. Org. Lett. 2015, 17, 4498−4501. (140) Cavelier, F.; Marchand, D.; Mbassi, P.; Martinez, J.; Marraud, M. Conformational studies of proline-, thiaproline- and dimethylsilaprolinecontaining diketopiperazines. J. Pept. Sci. 2006, 12, 621–625. (141) Dalkas, G. A.; Marchand, D.; Galleyrand, J. –C.; Martinez, J.; Spyroulias, G. A.; Cordopatisa, P.; Cavelier, F. Study of a lipophilic captopril analogue binding to angiotensin I converting enzyme. J. Pept. Sci. 2010, 16, 91–97. (142) Martin, C.; Legrand, B.; Lebrun, A.; Berthomieu, D. Martinez, J.; Cavelier, F. Silaproline helical mimetics selectively form an all-trans PPII Helix. Chem. Eur. J. 2014, 20, 14240–14244. (143) Rene´,

A.;

Vanthuyne,

(Trimethylsilyl)alanine

N.;

synthesis

Martinez, exploiting

J.;

Cavelier,

(L)-

hydroxypinanone-induced

diastereoselective alkylation. Amino Acids 2013, 45, 301–307.

66 ACS Paragon Plus Environment

F.

Page 67 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

(144) Vivet, B.; Cavelier, F.; Martinez, J. Synthesis of silaproline, a new proline surrogate. Eur. J. Org. Chem. 2000, 807−811. (145) Cavelier, F.; Vivet, B.; Martinez, J.; Aubry, A.; Didierjean, C.; Vicherat, A.; Marraud, M. Influence of silaproline on peptide conformation and bioactivity. J. Am. Chem. Soc. 2002, 124, 2917−2923. (146) Pujals, S.; Fernandez-Carneado, J.; Kogan, M. J.; Martinez, J.; Cavelier, F.; Giralt, E. Replacement of a proline with silaproline causes a 20-fold increase in the cellular uptake of a pro-rich peptide. J. Am. Chem. Soc. 2006, 128, 8479−8483. (147) Marchand, D.; Martinez, J.; Cavelier, F. Straightforward synthesis of chiral silylated amino acids through hydrosilylation. Eur. J. Org. Chem. 2008, 3107– 3112. (148) Martin, C.; Vanthuyne, N.; Miramon, H.; Martinez, J.; Cavelier, F. Resolution of protected silaproline for a gram scale preparation. Amino Acids 2012, 43, 649-655. (149) Vivet, B.; Cavelier, F.; Martinez, J.; Didierjean, C.; Marraud, M.; Aubry, A. A silaproline-containing dipeptide. Acta Cryst. 2000, C56, 1452-1454. (150) Bartoccini, F.; Bartolucci, S.; Lucarini, S.; Piersanti, G. Synthesis of boronand silicon-containing amino acids through copper- catalysed conjugate additions to dehydroalanine derivatives. Eur. J. Org. Chem. 2015, 3352–3360. (151) Liston, P.; Fong, W. G.; Korneluk, R. G. The inhibitors of apoptosis: there is more to life than Bcl2. Oncogene 2003, 22, 8568–8580. (152) Kim, K. S. IAP anatgonists. World Patent Application WO-2014/011712, 2014.

67 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(153) Nair, A. G. Zeng, Q.; Selyutin, O.; Rosenblum, S. B.; Jiang, Y.; Yang, D. Y.; Keertikar, K.; Zhou, G.; Dwyer, M. P.; Kim, S. H.; Shankar, B.; Yu, W.; Tong, L.; Chen, L.; Mazzola, R.; Caldwell, J.; Tang, H.; Allard, M. L.; Buckle, R. N.; Gauuan, P. J.; Holst, C. L.; Martin, G. S.; Naicker, K. P.; Vellekoop, S.; Agrawal, S.; Liu, R.; Kong, R.; Ingravallo, P.; Xia, E.; Zhai, Y.; Nomeir, A.; Kozlowski, J. A. Discovery of silyl proline containing HCV NS5A inhibitors with pan-genotype activity: SAR development. Bioorg. Med. Chem. Lett. 2016, 26, 1475–1479. (154) Nair, A. G.; Keertikar, K. M.; Kim, S. H.; Kozlowski, J. A.; Rosenblum, S.; Selyutin, O. B.; Wong, M.; Yu, W.; Zeng, Q. Fused tricyclic silyl compounds and methods of use thereof for the treatment of viral diseases. World Patent Application WO-2011/112429, 2011. (155) Hu, Y.; Wang, Y.; Li, F.; Ma, C.; Wang, J. Design and expeditious synthesis of organosilanes as potent antivirals targeting multidrug-resistant influenza A viruses. Eur. J. Med. Chem. 2017, 135, 70-76. (156) Tacke, R.; Merget, M.; Bertermann, R.; Bernd, M.; Beckers, T.; Reissmann, T. Syntheses and properties of Silicon- and Germanium-containing α-amino acids and peptides: a study on C/Si/Ge bioisosterism. Organometallics 2000, 19, 3486-3497. (157) Matthews, J. L.; McArthur, D. R.; Muir, K. W. Synthesis of novel silasubstituted β-amino acids. Tetrahedron Lett.2002, 43, 5401−5404. (158) Madsen, J. L. H.; Hjørringgaard, C. U.; Vad, B. S.; Otzen, D.; Skrydstrup, T. Incorporation of β-silicon-β3-amino acids in the antimicrobial peptide

68 ACS Paragon Plus Environment

Page 68 of 71

Page 69 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

alamethicin provides a 20-fold increase in membrane permeabilization. Chem. Eur. J. 2016, 22, 8358 –8367. (159) Muratake, H.; Ito, A.; Toda, T.; Suzuki, H.; Fukasawa, H.; Tsuda, M; Inoue, K.;

Sugiyama,

K.;

Shudo,

K.

(R)-

and

(S)-4-Amino-3-

(trimethylsilyl)methylbutanoic acids ameliorate neuropathic pain without central nervous system-related side effects. Bioorg. Med. Chem. Lett. 2012, 22, 7602–7604. (160) Fukasawa, H.; Muratake, H.; Ito, A.; Suzuki, H.; Amano, Y.; Nagae, M.; Sugiyama, K.; Shudo, K. Silicon-containing GABA derivatives, silagaba compounds, as orally effective agents for treating neuropathic pain without central-nervous-system-related side effects. ACS Chem. Neurosci. 2014, 5, 525−532. (161) Ramesh, R.; Reddy, D. S. Zinc mediated allylations of chlorosilanes promoted by ultrasound: Synthesis of novel constrained sila amino acids. Org. Biomol. Chem. 2014, 12, 4093-4097. (162) Information from the internet via www.clinicaltrials.gov accessed on July 14, 2017. (163) Information from the website www.ergo-log.com/silabolin.html accessed on May 1, 2017. (164) Higginbotham, K. B.; Lozano, R.; Brown, T.; Patt, Y. Z.; Arima, T.; Abbruzzese, J. L.; Thomas, M. B. A phase I/II trial of TAC-101, an oral synthetic retinoid, in patients with advanced hepatocellular carcinoma. J. Cancer Res. Clin. Oncol. 2008, 134, 1325-1335.

69 ACS Paragon Plus Environment

Journal of Medicinal Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(165) Tse, F. L. S.; Jaffe, J. M.; Dain, J. G. Pharmacokinetics of compound 58–112, a potential skeletal muscle relaxant, in man. J. Clin. Pharmacol. 1984, 24, 47– 57. (166) Toutov, A. A.; Liu, W. –B.; Betz, K. N.; Stoltz, B. M.; Grubbs, R. H. Catalytic C–H bond silylation of aromatic heterocycles. Nat. Protoc. 2015, 10, 1897-1903. (167) New ref: Toutov, A. A.; Liu, W. –B.; Betz, K. N.; Fedorov, A.; Stoltz, B. M.; Grubbs, R. H. Silylation of C–H bonds in aromatic heterocycles by an earthabundant metal catalyst. Nature 2015, 518, 80-84. (168) Franz, A. K.; Dreyfuss, P. D.; Schreiber, S. L. Synthesis and cellular profiling of diverse organosilicon small molecules. J. Am. Chem. Soc. 2007, 129, 1020-1021. (169) Kan, S. B. J.; Lewis, R. D.; Chen, K.; Arnold, F. H. Directed evolution of cytochrome c for carbon–silicon bond formation: Bringing silicon to life. Science 2016, 354, 1048-1051.

70 ACS Paragon Plus Environment

Page 70 of 71

Page 71 of 71

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Medicinal Chemistry

Table of Contents graphic

71 ACS Paragon Plus Environment