throbber
HHS Public Access
`Author manuscript
`Expert Opin Drug Deliv. Author manuscript; available in PMC 2016 November 21.
`Published in final edited form as:
`Expert Opin Drug Deliv. 2009 April ; 6(4): 405–420. doi:10.1517/17425240902824808.
`
`Page 1
`
`Prodrug approaches to improving the oral absorption of antiviral
`nucleotide analogues
`
`Larryn W Peterson1 and Charles E McKenna†,2
`1PhD candidate, University of Southern California, Department of Chemistry, Los Angeles, CA
`90089-0744, USA
`2Professor of Chemistry, University of Southern California, Department of Chemistry, Los
`Angeles, CA 90089-0744, USA
`
`Abstract
`Nucleotide analogues have been well accepted as therapeutic agents active against a number of
`viruses. However, their use as antiviral agents is limited by the need for phosphorylation by
`endogenous enzymes, and if the analogue is orally administered, by low bioavailability due to the
`presence of an ionizable diacid group. To circumvent these limitations, a number of prodrug
`approaches have been proposed. The ideal prodrug achieves delivery of a parent drug by
`attachment of a non-toxic moiety that is stable during transport and delivery, but is readily cleaved
`to release the parent drug once at the target. Here, a brief overview of several promising prodrug
`strategies currently under development is given.
`
`Keywords
`antiviral agents; oral bioavailability; prodrugs; pronucleotides
`
`1. Introduction
`Of the approximately 40 antiviral drugs formally licensed for use, half are nucleoside or
`nucleotide analogues [1]. Nucleoside drugs per se must usually be phosphorylated to the 5(cid:5321)-
`mono-, 5(cid:5321)-di-, and finally, 5(cid:5321)-triphosphate by intracellular or viral kinases [2] in order to
`inhibit their therapeutic targets. This requirement limits efficacy, as phosphorylation to the
`monophosphate by endogenous kinases is slow and typically is the rate-limiting step in
`human cells [3,4].
`
`The administration of a nucleoside drug as its monophosphate (NMP) is a well-known
`approach to overcoming this obstacle [3,5]. However, this entails a penalty in the form of
`decreased membrane permeability. Nucleotide analogues contain an ionizable –O-P(O)
`(OH)2 group that exists chiefly as a dianion at physiological pH, resulting in low oral
`bioavailability [5]. In addition, if a NMP succeeds in crossing the intestinal membrane, it
`
`†Author for correspondence: Tel: +1 213 740 7007; Fax: +1 213 740 0930; mckenna@usc.edu.
`Declaration of interest
`LP and CM are co-inventors on a patent related to a portion of the work discussed in this review.
`
`Gilead 2014
`I-MAK v. Gilead
`IPR2018-00125
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`

`

`Peterson and McKenna
`
`Page 2
`
`then becomes a potential substrate for phosphohydrolases (phosphatases and 5(cid:5321)-
`nucleotidases), which remove the phosphate group [6]. The use of a nucleoside phosphonate
`–CH2-P(O)(OH)2 circumvents dephosphorylation, but decreased transport remains an
`obstacle.
`
`Formulation strategies [7–11] to overcome these limitations are beyond the scope of this
`short review, which has as its focus an alternative approach: prodrug modification of
`nucleotide drugs. Promoieties can be attached at a number of positions on an NMP or
`nucleotide analogue [12,13]. However, the introduction of promoieties at the phosphorus (–
`[O,CH2]-P(O)(X)(Y) where X,Y = OR, OR(cid:5321), NHR(cid:5322)) directly addresses the problem of
`blocking P-OH ionization in vivo. The attachment of a well-designed promoiety increases
`delivery of the drug to its target, provided that its biochemical and physical properties –
`including lipophilicity, site-specificity and chemical stability – are conducive to this end
`[5,13,14].
`
`A prodrug must be stable under delivery conditions [3,5], but it must be capable of
`conversion to its active parent drug in vivo [5], at a rate consistent with pharmacological
`efficacy. The prodrug and metabolized promoiety/promoieties should have low acute and
`chronic toxicity [5]. Control of these and other crucial properties, such as aqueous solubility
`and lipophilicity, remains a key challenge in the development of an effective prodrug.
`
`Esterification with pivaloyloxymethyl (POM), p-acyloxybenzyl (PAOB), or isopropyloxy–
`carbonyloxymethyl (POC) groups has been reviewed extensively [3,5,6,13,15,16] and will
`not be addressed here. Also, of recent interest, but omitted from this discussion is the
`approach of Hostetler et al. to improve the oral bioavailability of certain antiviral
`phosphonate drugs by esterification with an ether lipid ester that mimics the natural lipid
`lysophosphatidylcholine, thus potentially delivering the prodrug within the cell intact
`[4,17,18]. Our review will examine the prodrug approaches represented by the structures in
`Figure 1.
`
`2. Phosphoramidate ‘ProTide’ approach
`McGuigan has introduced prodrugs (‘ProTides’) based on an amino acid ester promoiety,
`attached to the drug (as a aryl monophosphate or phosphonate) via a P-N bond, applying this
`approach to: 4(cid:5321)-azidouridine [19], 4(cid:5321)-azidoadenosine [20], 2(cid:5321),3(cid:5321)-dideoxyuridine (ddU) [21],
`carbocyclic L-d4A (L-Cd4A) [22], stauvidine (d4T) [23], 9-[2-(phosphonomethoxy)
`ethyl]adenine (PMEA) [24], 3(cid:5321)-azidothymidine (AZT) [25], abacavir (ABC) [26] and
`tenofovir (PMPA, 9-[(R)-2-(phosphonomethoxy)propyl]adenine) [27].
`
`The original approach involved preparation of simple alkyloxy phosphoramidates (Figure
`2A(1)), but has evolved into aryloxy phosphoramide pronucleotides with distinct structure–
`activity relationship detail [28]. Phosphorodiamidates (Figure 2A(2)) were also prepared, but
`no biological benefit versus the phosphoramidates was observed and synthetic yields were
`lower [28]. Interestingly, analogues linked through an oxygen resulted in a significant
`decrease in antiviral activity [29], possibly because the nucleoside monophosphate was not
`released from the diester intermediate [28]. Diaryl pronucleotides (Figure 2A(3)) were not
`
`Expert Opin Drug Deliv. Author manuscript; available in PMC 2016 November 21.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`

`

`Peterson and McKenna
`
`Page 3
`
`active in kinase-deficient cells [30], due to poor intracellular delivery of the NMP or
`possibly chemical instability of the diaryl masking groups. Overall, aryloxy
`phosphoramidates (Figure 2A(4)) appear to hold the most promise for delivery of the
`therapeutic agent.
`
`Aryloxy phosphoramidates were designed to release the NMP intracellularly via both
`chemical and enzymatic mechanisms (Figure 2B). The first step in the activation process is
`cleavage of the amino acid ester by a carboxyesterase [28] to afford 6 (Figure 2B), although
`a thorough investigation by Venkatachalam and coworkers found that activation by a lipase
`or protease is also possible and that these enzymes have different specificities for the
`substituent on the aryl group, the amino acid and the stereochemistry at the phosphorus [31].
`Subsequent nucleophilic attack at the phosphorus by the carboxyl group releases the aryloxy
`group, forming a transient cyclic diester, which is hydrolyzed to form the amino acyl
`metabolite (AAM, Figure 2B(7)) [28]. In the final step, the amino acid moiety is cleaved by
`a phosphoramidase to release the nucleoside monophosphate (Figure 2B(8)) and an amino
`acid [28].
`
`McGuigan and coworkers have thoroughly studied the aryloxy phosphoramidates of d4T and
`have been able to gain extensive structure–activity relationship insight. In general, the
`methyl, ethyl and benzyl esters lead to potent activity, while bulkier esters (t-butyl and
`isopropyl) are significantly less active than the methyl ester [32], most likely due to the
`increased stability to enzymatic hydrolysis [20]. A quantitative structure–activity
`relationship (QSAR) study on the variation of amino acid esters further described the most
`potent esters as those with considerable lipophilicity slightly removed from the ester bond
`[33]. In a separate report, it was found that the conversion of AAM to NMP was inhibited
`when benzyl alcohol was released [34]. This inhibition was not observed when ethanol or
`methanol was released [34].
`
`Although there are some exceptions depending on the nature of the drug used, the most
`successful pronucleotides contain L-alanine as the amino acid [34,35]. Exchange of L-
`alanine for either glycine or L-leucine reduced the antiviral activity 70- and 13-fold,
`respectively [36]. When L-valine was used, the antiviral activity was reduced 147-fold, and
`furthermore, 100% of the intact prodrug was recovered when it was exposed to pig liver
`carboxyesterases [34]. When D-alanine was substituted for L-alanine, the potency was
`decreased 35-fold [34]. The exact reason for the preference for alanine remains unknown.
`When the achiral amino acid analogue (cid:867),(cid:867)-dimethylglycine was prepared, the antiviral
`activity was only reduced three-fold [36], which illustrates the fact that natural amino acids
`are not essential for activity. However, when such amino acids were used, a preference for
`(cid:867)-amino acids was observed [34]. The (cid:868)-amino acid phosphoramidates showed efficient
`ester cleavage, but no phenyl loss was detected, and the AAM was not observed [34]. This
`suggests a possible entropy barrier that increases with chain length.
`
`Studies to determine the optimal aryloxy group were also performed [37]. The greatest
`activity was achieved when the aryl group had a p-Cl substituent, and generally for aryl
`groups that function as mildly electron-withdrawing, lipophilic substituents [37]. The
`potential for toxicity of the released phenol was not discussed. A naphthyl group was also
`
`Expert Opin Drug Deliv. Author manuscript; available in PMC 2016 November 21.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`

`

`Peterson and McKenna
`
`Page 4
`
`reported to be an effective aryl moiety for delivering anticancer agents [38], and this activity
`is most likely transferable to antiviral agents.
`
`To obtain successful intracellular delivery of NMP, the pronucleotides need to be resistant to
`hydrolysis during the absorption and distribution process. The chemical stability of the
`pronucleotides was studied, and all exhibited satisfying stability over the range of pHs
`studied (2.0 – 7.4) [39]. The aryloxy phosphoramidates were significantly less resistant to
`decomposition in plasma or cell extracts, indicative of the need for enzymatic activation
`[39]. After overnight exposure to pig liver carboxyesterases, CEM cell extract, human serum
`and mouse serum, AAM was formed from a majority of the L-amino acid-containing
`pronucleotides [34]. When carbocyclic adenosine phosphoramidates were evaluated in
`intestinal and liver S9 homogenates, some of the most antivirally potent analogues exhibited
`complete decomposition over 1 h in intestinal homogenate [22], but isopropyl and t-butyl
`esters on the amino acid increased intestinal stability [22]. Similarly, D-alanine and glycine
`exhibited the highest intestinal stability [22], which highlights the complexity in obtaining a
`structure–activity relationship. Although the usage of this aryloxy phosphoramidate prodrug
`approach with nucleotide analogues containing a phosphonate may be more difficult due to
`decreased chemical stability [40], an example of successful application to tenofovir has been
`described [27].
`
`The pharmacokinetics and oral bioavailability of aryloxy phosphoramidates, specifically
`abacavir phosphoramidates, were examined [35]. When the abacavir methyl alaninyl–
`phosphoramidate was administered intravenously, the pronucleotide was rapidly cleared
`from the plasma with a half-life of 7 min [35]. Similar results were observed following oral
`administration [35]. However, the major metabolite observed was the AAM [35]. Total
`exposure to the pronucleotide and its active metabolites was reported to approach that
`estimated for a similar dose of the parent drug, abacavir, resulting in an overall
`bioavailability of 50% [35]. The epithelial permeability of a series of d4T aryloxy
`phosphoramidates was evaluated in Caco-2 and MDCK monolayers [41]. The
`pronucleotides exhibited relatively low permeability, which may be partially explained by
`their susceptibility to first-pass metabolism in the intestinal epithelial cells and by being
`substrates of P-gp [41]. In general, this work exemplifies the difficulty in delivering the
`NMP to the target while avoiding significant metabolism during absorption and distribution.
`To obtain optimal antiviral activity of each pronucleotide, the fine tuning of each element
`(amino acid, ester, and aryl moiety) is required.
`
`3. Monoester prodrugs
`Amino acid phosphoramidate monoesters designed to release the NMP after a single
`activation by an endogenous phosphoramidase have been described by Wagner, who has
`applied this approach to AZT [42,43] and ddA [44], as well as anticancer drugs [45].
`
`After the delivery of AZT monophosphate by a glycoslyated carrier attached through lysine
`was reported [46], Wagner and coworkers proposed that NMP could be efficiently delivered
`by non-polar amino acid phosphoramidate monoesters and that the aryl group was not
`necessary. Furthermore, these phosphoramidate monoesters were stable in cell culture and
`
`Expert Opin Drug Deliv. Author manuscript; available in PMC 2016 November 21.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`

`

`Peterson and McKenna
`
`Page 5
`
`rat and human plasma [42]. A series of these compounds were synthesized containing an
`amino acid (tryptophan methyl ester [Figure 2C(9)] or phenylalanine methyl ester [Figure
`2C(11)]) connected via a P-N bond to the NMP with the other P-OH left as a free acid or
`esterified to a simple alkyl group [47]. The tryptophan monoester (Figure 2C(9)) exhibited
`the best antiviral activity, with an eight-fold increase over AZT with no cytotoxicity
`observed at the levels tested [47]. Further studies have been done to investigate the activation
`pathway of these pronucleotides and optimize their structures.
`
`The effect of changing the amino acid was studied in peripheral blood mononuclear cells
`(PBMC) [42]. The best antiviral activity was obtained with the L-alanine methyl ester [42],
`consistent with McGuigan’s ProTides. Furthermore, enhanced activity was observed with
`the L-tryptophan derivative (Figure 2C(9)) compared with the L-phenylalanine (Figure
`2C(11)), L-valine and L-leucine derivatives [42]. When evaluated in CEM cells, the L-
`alanine and L-phenylalanine derivatives exhibited antiviral activity comparable to AZT [42].
`This suggests that a simple structure–activity relationship does not exist. In order to avoid
`the polar carboxylate formed after interaction of the pronucleotides with carboxyesterases,
`the amino acid methyl ester was substituted by a methyl amide [42]. The authors reported
`that this exchange had little effect on the antiviral activity of the tryptophan derivatives,
`while the phenylalanine methyl amide derivatives exhibited increased potency [42].
`However, the methyl amide derivatives exhibited greater in vitro and in vivo stability [48].
`The antiviral activity did not exhibit a strong dependence on the amino acid stereochemistry
`[42], but the inclusion of the D-isomer versus the L-isomer led to decreased volumes of
`distribution [48]. Overall, the L-tryptophan methyl amide derivative (Figure 2C(10)) was
`selected for further studies.
`
`To better understand the differences in potency, Wagner and coworkers investigated the
`ability of the pronucleotide to deliver NMP intracellularly [42]. The antiviral activity is
`strongly related to the intracellular levels of nucleoside triphosphate. In both PBMCs and
`CEM cells, AZT was able to produce higher levels of AZT triphosphate than the
`pronucleotides [42]. However, when evaluated in CEM cells, the intracellular levels of the
`tryptophan methyl ester (Figure 2C(9)) and phenylalanine methyl ester (Figure 2C(11))
`pronucleotides did not plateau [49]. Therefore, the differences in potency may be derived
`from the ability of a phosphoramidase to cleave the P-N bond and release the NMP.
`
`The oral bioavailability, disposition and stability of the AZT phosphoramidate monoesters
`were evaluated in rats [50]. The phosphoramidate monoesters were stable in tissue
`homogenates, intestinal contents, and rat and human plasma [48,50]. The tryptophan methyl
`amide derivative (Figure 2C(10)) exhibited the best pharmacokinetic parameters. However,
`in simulated gastric fluids at pH 2.0, the pronucleotide exhibited a significantly reduced
`half-life of 5 h, but greater stability as the pH increased [50]. These results are consistent
`with greater chemical hydrolysis of P-N bonds at lower pH [51]. The pronucleotide was not
`detected in plasma or urine, which was confirmed in an in situ single pass perfusion study
`where little or no absorption of the pronucleotide in the 120 min perfusion period was
`detected [50]. AZT was observed in plasma and urine samples accounting for 29.5% of the
`dose, while 54.3% of the dose was recovered 4 h post-dosing (intravenously) as intact
`pronucleotide in the bile [50]. These results offer some possible explanations for the zero
`
`Expert Opin Drug Deliv. Author manuscript; available in PMC 2016 November 21.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`

`

`Peterson and McKenna
`
`Page 6
`
`oral bioavailability of the pronucleotide. Not only will the pronucleotide exist as a charged
`species at physiological pH, but pre-systemic hydrolysis of the P-N bond would lead to a
`dianionic monophosphate. This fact, combined with its molecular weight, makes biliary
`excretion difficult to avoid [50].
`
`The phosphoramidate monoester pronucleotide approach was applied to AZT to create a
`pronucleotide with increased antiviral activity and decreased cytotoxicity. Unfortunately,
`decomposition of the phosphoramidate monoester in simulated gastric fluid was observed,
`and the pronucleotide exhibited little or no bioavailability when evaluated in rats.
`
`4. SATE pronucleotides
`S-Acyl-2-thioethyl (SATE) protecting groups for nucleotide drugs have been introduced for
`the delivery of a number of NMP including d4T [52], PMEA [53], elvucitabine ((cid:868)-L-FD4C)
`[54], acyclovir [55,56], AZT [57–59] and cytarabine (Ara-C) [60].
`
`The SATE approach utilizes both enzymatic and chemical mechanisms to activate the
`pronucleotide and release the NMP (Figure 3A) [61]. Removal of the S-acyl-2-thioethyl
`protecting group and release of the monophosphate is initiated by esterase-mediated
`hydrolysis of the acyl group, which produces a reactive thiol group [61]. Nucleophilic attack
`on the (cid:867)-carbon results in a reactive 2-mercaptoethyl ester that decomposes spontaneously
`to release the diester (Figure 3A(14)) and ethylene sulfide (episulfide) [61]. If the second
`ester is also a SATE group, the process is then repeated, resulting in release of the NMP
`[61]. It has been proposed that if SATE mixed pronucleotides are used, the aryl group in
`SATE phosphotriesters is cleaved by a type 1 phosphodiesterase and a phosphoramidase
`cleaves the protecting group bound via a P-N bond [61]. However, the cleavage of the aryl
`ester or phosphoramidate could be concomitant with cleavage of the SATE group [61].
`
`Although the release of the NMP is essential, a further point to consider is promoiety
`toxicity. In the activation process, carboxylic acids and episulfide were released [61]. The
`body can metabolize the carboxylic acids, and no cytotoxicity was reported for the episulfide
`[61]. When the SATE promoiety and its metabolites were evaluated in various cell lines, no
`additional cytotoxicity was observed [61]. In vivo toxicity studies in cynomolgus monkeys
`showed neither clinical symptoms nor behavior problems indicative of toxicity [61].
`
`In all cases, application of the bis(SATE) phosphotriester (Figure 3B(15)) approach has led
`to increased in vitro antiviral activity compared to the parent nucleoside and efficient
`delivery of the NMP. Bis(MeSATE) d4T monophosphate was 10- to 17-fold more potent
`than d4T in wild-type CEM cells and PBMCs, and furthermore, antiviral activity was
`retained in thymidine kinase-deficient cells [52]. A slight increase in cytotoxicity was also
`observed for the pronucleotides [52]. Administration of the bis(tBuSATE) Ara-C
`phosphotriester resulted in significant antiviral activity in a resistant cell line due to lack of
`the appropriate kinase [60].
`
`Stability and transport across a Caco-2 monolayer was used to determine the optimal R1
`group (Figure 3B) [53]. Transport of the pronucleotides across Caco-2 monolayers resulted
`in intact pronucleotide in the basolateral compartment only when the bis(tBuSATE)
`
`Expert Opin Drug Deliv. Author manuscript; available in PMC 2016 November 21.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`

`

`Peterson and McKenna
`
`Page 7
`
`phosphotriester was evaluated, whereas no intact pronucleotide or metabolites were
`observed for the bis(MeSATE) phosphotriester [53]. Although very low amounts of the
`intact pronucleotide were detected, the greatest uptake in the Caco-2 monolayer was
`observed for the bis(tBuSATE) phosphotriester, which could be due to the increased
`enzymatic stability or lipophilicity of the pivaloyl group [53]. The bis(tBuSATE)
`phosphotriester was extensively metabolized to the monoester during transport in vivo,
`resulting in no intact pronucleotide observed in the plasma [62]. Some bis(tBuSATE)
`phosphotriesters exhibit poor aqueous solubility and had to be administered with 5%
`dimethyl sulfoxide (DMSO), which affected the esterase activity and significantly increased
`the half-lifes of the pronucleotides [62].
`
`In an attempt to obtain more favorable stability and bioavailability, a SATE group bearing a
`functionalized acyl moiety was investigated (Figure 3B(16)) [57,59,62]. The addition of one
`hydroxyl group led to antiviral activity comparable to the nucleoside and increased activity
`compared to the bis(tBuSATE) phosphotriester [62]. The addition of a hydroxyl group to the
`bis(tBuSATE) promoiety decreased the hydrolysis rate and increased the half-life and
`aqueous solubility of the pronucleotide [62]. Furthermore, intact pronucleotide was observed
`when the bis(hydroxyl-tBuSATE) phosphotriester was evaluated in a Caco-2 monolayer
`[62].
`
`Several variations of the SATE pronucleotide approach have been described. Initial attempts
`at the intracellular delivery of NMP with SATE esters involved the use of two SATE groups
`[52,55,60,62]. Although these bis(SATE) pronucleotides (Figure 3B(15 and 16)) exhibited
`increased antiviral activity and decreased cytotoxicity, the removal of the second SATE
`group proved difficult and proceeded much more slowly due to the negative charge at the
`phosphate [61]. Therefore, Gosselin and Imbach evaluated two different kinds of SATE
`mixed esters: aryl SATE phosphotriesters (Figure 3B(17)) [64,65] and SATE
`phosphoramidate diesters (Figure 3B(18)) [64,67].
`
`A phenyl (tBuSATE) mixed phosphotriester was synthesized, but the NMP was not released
`from the prodrug when studied in cell extract [63]. Several derivatives of L-tyrosine SATE
`phosphotriesters (Figure 3B(17)) were investigated for stability and antiviral activity [64,65].
`Since the presence of the free carboxylic acid on the tyrosine residue resulted in decreased
`antiviral activity, most likely due to decreased membrane permeability, the moiety was
`modified to contain polar, but not anionic, functionalities [66]. The pronucleotide with the
`shortest half-life exhibited the best antiviral activity [64]. The resulting mixed SATE
`phosphotriesters exhibited enhanced antiviral activity in CEM kinase-deficient cells,
`illustrating the successful delivery of the NMP intracellularly [64].
`
`Another solution to the slow activation of the bis(SATE) phosphotriesters was the SATE
`phosphoramidate diesters (Figure 3B(18)) modeled after the phosphoramidates of McGuigan
`and Wagner. Initial studies with various alkylamines illustrated that the rate limiting
`hydrolysis of the phosphoramidate was dependent on the basicity and bulk of the amine
`[65,67]. However, when the pKa of the amine was appropriate (approx. 5 – 11.2), the
`phosphoramidate diesters effectively delivered NMP intracellularly [67]. Phosphoramidate
`diesters of AZT were as potent as AZT in wild-type CEM cells and retained significant
`
`Expert Opin Drug Deliv. Author manuscript; available in PMC 2016 November 21.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`

`

`Peterson and McKenna
`
`Page 8
`
`antiviral activity in kinase-deficient cells [58]. Interestingly, the isopropylamino derivative
`exhibited the greatest potency, which illustrates the flexibility in the amine moiety and
`demonstrates that an (cid:867)-amino acid is not a structural requirement [58].
`
`Périgaud and coworkers have recently investigated the use of aryl SATE phosphotriesters
`[57,59]. Building on the ability of (hydroxyl-tBu)SATE to increase the solubility and
`stability of the prodrug, functionalization of the acyl moiety with polar groups was applied
`to the phenyl SATE mixed phosphotriester analogues. Additionally, an analogue with a
`valine replacing the acyl group was studied, but the compound was unstable in cell extract
`and did not maintain antiviral activity in a thymidine kinase (TK) deficient cell line [57,59].
`Introduction of one hydroxyl group on the acyl moiety resulted in greater stability in cell
`extracts compared to the tBuSATE analogue [57]. The derivative containing two hydroxyl
`groups showed decreased enzymatic stability compared to the (hydroxyl-tBu)SATE
`analogue, but it resulted in a significant loss in activity in a TK-deficient cell line [57]. The
`monohydroxylated prodrug showed anti-HIV activity in the micromolar range comparable to
`the tBuSATE analogue, and the solubility of the pronucleotide was greatly increased [57],
`which demonstrates the necessity of obtaining optimized pharmacokinetic properties to
`achieve effective prodrugs.
`
`Application of the SATE pronucleotide approach to numerous nucleoside monophosphates
`and nucleotide analogues has resulted in increased antiviral activity in vitro. However, when
`evaluated for transport across Caco-2 monolayers and for bioavailability, no intact prodrug
`was observed illustrating premature hydrolysis. As pointed out by Gosselin, Imbach, and
`Périgaud, these facts illustrate the necessary balance that needs to be achieved between
`lipophilicity, solubility and enzymatic stability [61].
`
`5. CycloSal prodrugs
`Meier et al. have shown that salicyl alcohol is an effective bifunctional masking unit for
`nucleotides, that is cleaved by a pH-dependent mechanism to deliver the active drug [68].
`They have illustrated the utility of this approach using various nucleoside and nucleotide
`analogues including acyclovir [69], 9-[2-(phosphonomethoxy)ethyl]adenine (PMEA,
`adefovir) [70], 2(cid:5321),3(cid:5321)-dideoxyadenosine (ddA) [71], 2(cid:5321),3(cid:5321)-dideoxy-2(cid:5321),3(cid:5321)-didehydroadenosine
`(d4A) [71], 5-[(E)-2-bromovinyl]-2(cid:5321)-deoxyuridine (BVdU or brivudin) [72,73], 2(cid:5321),3(cid:5321)-
`dideoxy-2(cid:5321),3(cid:5321)-didehydrothymidine (d4T) [74–79] and carbocyclic 3(cid:5321)-azidothymidine
`analogues [80].
`
`Their original goal in creating these pronucleotides was to find a masking unit that would
`deliver the nucleotide analogue exclusively by a chemical mechanism. Initial attempts using
`bis(alkyl), bis(phenyl), or bis(benzyl) nucleotide triesters proved unsuccessful at releasing
`the NMP by a purely chemical mechanism [6]. The charge formed once one ester was
`cleaved led to a stable compound resistant to further chemical hydrolysis [16]. However,
`Meier and coworkers found that they could successfully mask the phosphate with phenyl and
`benzyl esters of salicyl alcohol, while the nucleoside was attached by esterification via the
`5(cid:5321)-hydroxyl group (Figure 4A(19)) [81]. These esters are distinct enough to achieve
`differentiated chemical hydrolysis independent of any enzymatic activity [82]. This principle
`
`Expert Opin Drug Deliv. Author manuscript; available in PMC 2016 November 21.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`

`

`Peterson and McKenna
`
`Page 9
`
`was validated by comparing the half-lifes of the triesters in phosphate buffer at pH 7.3 and
`cell extracts with fetal calf serum, which showed similar half-lifes for the triesters in both
`media [83].
`
`The successful cycloSal pronucleotide releases the nucleotide (Figure 4B(23)) and the
`salicyl alcohol (Figure 4B(24)) intracellularly [81]. The nucleotide is released through a
`cascade, which is initiated by cleavage of the phenyl ester via nucleophilic attack on the
`phosphorus atom by hydroxide to form the diester (Figure 4B(22)) (step a1). The ring, now
`activated by the strong electron-donating hydroxyl group, allows for cleavage of the benzyl
`ester to yield the nucleotide analogue (Figure 4B(23)) and salicyl alcohol (Figure 4B(24))
`via a SN1-type reaction at the benzyl position (step a2). It is also possible that hydrolysis of
`the benzyl ester can occur to first produce a charged intermediate (Figure 4B(25)) (step b1),
`which then reacts with water to yield 26 (Figure 2B) (step b2). Further hydrolysis of 26
`(Figure 2B) does not occur, thus preventing the release of the nucleotide analogue. However,
`in hydrolysis studies, the major products were the NMP and salicyl alcohol. When evaluated
`for cytotoxicity in mice, salicyl alcohol showed no toxicity [82].
`
`As the cycloSal pronucleotides were designed to release the active drug via a chemical
`cascade mechanism, the stability and hydrolysis pathways of these pronucleotides can be
`fine-tuned by varying the substituents on the aromatic ring. Acceptor substituents in the 5-
`or 6-position decrease stability, while donor substituents at the 3- or 5-position increase the
`stability of the triesters (Figure 4A(19)) [82]. Bulky substituents (tert-butyl groups) at the 3-
`and/or 5-position increase the amount of the phenyl phosphate diester (Figure 4B(25))
`observed. When substitution was made at the benzyl position, the half-life decreased
`drastically compared to the unsubstituted analogue and the major product was the diester
`(Figure 4B(25)) in hydrolysis studies [82]. However, the addition of a donor substituent at
`the 6-position caused the major hydrolysis product to be the desired diester (Figure 4B(22))
`[82].
`
`Although there is a benefit – lack of dependence on enzyme expression differences in
`tissues, individuals and species – to the use of a pronucleotide activated by a chemical
`mechanism, the possibility of extracellular release of the active drug or efflux of the
`pronucleotide, due to the establishment of a concentration equilibrium across the cell
`membrane, cannot be ignored. To remedy these potential problems, Meier et al. designed a
`way to trap the pronucleotide inside the cell [77,79,84]. In theory, the attachment of a moiety
`to the aromatic ring that can be enzymatically activated to release a more polar group will
`prevent penetration of the cellular membrane by the compound and trap the pronucleotide
`[75]. The initial attempts included the use of an esterase to release an alcohol or carboxylic
`acid [85]. The released alcohol group was not polar enough to prevent efflux of the
`pronucleotide, and when a two, three, or four carbon linker was used to attach a methyl or
`benzyl ester to the aromatic ring, the compounds did not show good esterase affinity [78].
`The feasibility of acetoxymethyl (AM) and pivaloyloxymethyl (POM) esters as enzyme-
`cleavable triggers was demonstrated by their significantly decreased stability in cell extracts
`versus plasma, illustrating selective activation of these compounds intracellularly [73].
`Attempts at intracellular trapping of the pronucleotide with an amino acid ester trigger
`moiety resulted in a large differential between the buffer and cell extract stability and
`
`Expert Opin Drug Deliv. Author manuscript; available in PMC 2016 November 21.
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`Author Manuscript
`
`

`

`Peterson and McKenna
`
`Page 10
`
`sustained antiviral activity in kinase-deficient cells [84]. Although a strongly polar group
`was required to trap the pronucleotide in the cell, Meier and coworkers found

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket